id
stringlengths 27
33
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 13
1.81M
|
---|---|---|---|
warning/0007/math-ph0007005.html
|
ar5iv
|
text
|
# Symmetric Hilbert spaces arising from species of structures
## 1 Introduction
Symmetric Hilbert spaces play a role in physics as the state spaces of many particle systems. The type of particle dictates the type of symmetrization: bosons require complete symmetrization and fermions complete antisymmetrization.
More general ways of of symmetrization, although apparently not realized in nature, have been studied for their own sake: parastatistics and interpolations by a parameter $`q[1,1]`$ between the above two cases .
All these constructions lead to quantum fields or generalized Brownian motions, each with their own generalized Gauss distributions . One particularly important case is $`q=0`$: the free Brownian motion, exhibiting the Wigner distribution. This case is related to free independence in the same way as the case $`q=1`$ of complete symmetrization is related to ordinary commutative independence.
Although there are results indicating that these two are the only notions of independence, more relaxed conditions such as the weak factorization property , or pyramidal independence are satisfied in a variety of examples.
In this paper we study combinatorial ways of symmetrization. Our starting point is the following observation. The category $`E`$ of finite sets has as its isomorphy classes the natural numbers $``$, and for each object $`U`$ in class $`n`$ there are $`n!`$ symmetries. This leads to the Fock space
$$l^2(,\frac{1}{n!})=:\mathrm{\Gamma }_E().$$
Taking for an annihilation operator $`a`$ the left shift on this space, we find that the field operator $`X:=a+a^{}`$ has distribution given by ()
$$\delta _{\mathrm{}},X^n\delta _{\mathrm{}}=\frac{1}{\sqrt{2\pi }}_{\mathrm{}}^{\mathrm{}}x^ne^{\frac{1}{2}x^2}𝑑x.$$
On the other hand, if we consider the category $`L`$ of finite sequences (or linear orderings on a set), we obtain
$$\mathrm{\Gamma }_L()=l^2()$$
and, since $`a^{}`$ is now the right shift ,
$$\delta _{\mathrm{}},X^n\delta _{\mathrm{}}=\frac{1}{\pi }_2^2x^n\sqrt{4x^2}𝑑x.$$
We conclude that the Gauss and Wigner distributions are produced by the concepts of ‘set’ and ‘sequence’. Our program in this paper is to generalize the Fock space construction to such combinatorial concepts as ‘tree’, ‘graph’ and ‘cycle’.
The proper framework for this undertaking turns out to be Joyal’s notion of a combinatorial species of structures . These are defined as functors from the category of finite sets with bijections to the category of finite sets with maps. Combinatorial species of structures can be viewed as coefficients of the Taylor expansion of analytic functors and lead to Joyal’s notion of a tensorial species, very close to the our $`\mathrm{\Gamma }_F(𝒦)`$. This circle of ideas is introduced in Sections 2 and 3.
A natural way to introduce an annihilation operator into this context is via the operation of removal of a point from a structure, which is called differentiation $`FF^{}`$ by Joyal. We thus arrive at operators
$$a(k):\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_F^{}(𝒦),a^{}(k):\mathrm{\Gamma }_F^{}(𝒦)\mathrm{\Gamma }_F(𝒦),(k𝒦)$$
However, these operators can only be added, in order to yield field operators, if the species $`F`$ and $`F^{}`$ are the same. This holds in two cases: the species $`E`$ of sets and the species $`E_\pm `$ of oriented sets, related to the Bose and Fermi symmetries. Natural as this may be, we cannot move any further if we do not modify the operation of removal of a point in some way.
Now in fact, already in the case of sequences it is required that the last point is removed. In the same way, we may require in the case of trees that only leaves may be picked off (points that leave the tree connected when removed). In the case of cycles we may require that the chain, coming from a broken, cycle must be connected up again. All of this leads to the study of suitable transformations between $`F`$ and $`F^{}`$, which are the subject of Section 4.
Our approach to symmetric Hilbert spaces and field operators provides a tool for creating new examples, and is particularly transparent due to the use of combinatorial objects which are easy to visualize. The two examples of “q-deformations” appearing in and are cast in the form of combinatorial Fock spaces for the species of ballots and the species of simple directed graphs respectively. In section 5 we point out the connection between this combinatorial approach and the one based on positive definite functions on pair partitions . We describe how the operations between species can be extended to the weights, illustrating this by examples.
## 2 Species of Structures
This section is a brief introduction to the combinatorial theory of species of structures , , insofar as is needed here.
We are concerned with the different kinds — or ‘species’ — of structures that can be imposed on a set $`U`$. The basic idea is that such a species is characterized by the way it transforms under permutations of the set $`U`$.
It will be convenient to consequently adhere to von Neumann’s construction of the natural numbers according to which $`0=\mathrm{}`$ and $`n+1=n\{n\}`$, so that the number $`n`$ coincides with the set $`\{0,1,2,\mathrm{},n1\}`$.
Definition. A species of structures is a rule $`F`$ which
(i) produces for each finite set $`U`$ a finite set $`F[U]`$,
(ii) produces for each bijection $`\sigma :UV`$ a function $`F[\sigma ]:F[U]F[V]`$.
The function $`F[\sigma ]`$ should have the following functorial properties:
(a) for all bijections $`\sigma :UV`$, $`\tau :VW`$, we have $`F[\tau \sigma ]=F[\tau ]F[\sigma ]`$,
(b) for the identity map $`\mathrm{Id}_U:UU`$, $`F[\mathrm{Id}_U]=\mathrm{Id}_{F[U]}`$.
The elements of $`F[U]`$ are called $`F`$-structures on $`U`$ and the function $`F[\sigma ]`$ describes the transport of $`F`$-structures along $`\sigma `$. Note that $`F[\sigma ]`$ is a bijection by the functorial property of $`F`$.
We denote by $`H_s`$ the stabilizer $`\{\sigma \mathrm{S}(U):F[\sigma ](s)=s\}`$ of the structure $`sF[U]`$.
Examples.
1. The species $`E`$ of sets is given by
$`E[U]`$ $`=`$ $`\{U\}.`$
$`E[\sigma ]`$ $`:`$ $`UV\text{if}\sigma :UV.`$
Thus the only $`E`$-structure over $`U`$ is the set $`U`$ itself. The stabilizer of this structure coincides with the whole permutation group $`H_s=\mathrm{S}(U)`$.
2. The species $`L`$ of linear orderings:
$$L[U]=\{f:|U|U;f\text{bijective}\}$$
where $`|U|=\{0,1,2,\mathrm{},|U|1\}`$ is the cardinality of $`U`$. The transport along the bijection $`\sigma :UV`$ is given by
$$L[\sigma ](f)=\sigma f.$$
The stabilizer of each linear ordering is trivial. The cardinality of the set of structures $`L[U]`$ is equal to that of the permutation group $`\mathrm{S}(U)`$.
3. The species $`𝒞`$ of cyclic permutations:
$`𝒞[U]`$ $`=`$ $`\{\pi \mathrm{S}_U|\pi ^k(u)u\text{for all }uU,k<|U|\};`$
$`𝒞[\sigma ]:\pi `$ $``$ $`\sigma \pi \sigma ^1.`$
Each structure $`\pi 𝒞[U]`$ has a nontrivial stabilizer $`H_\pi =\{\pi ^k|k<|U|\}`$, the number of structures is
$$|𝒞[U]|=\frac{|U|!}{|H_\sigma |}=(|U|1)!$$
Definition. A species of structures $`F`$ is called molecular if the permutation group acts transitively on its structures. A molecular species can be characterized by the conjugacy class of the stabilizer of any of its structures. Indeed for $`s,tF[U]`$ and $`s=F[\sigma ](t)`$ we have $`H_s=\sigma H_t\sigma ^1`$. By a well-known combinatorial lemma we have for each structure $`s`$:
$$|F[U]||H_s|=|U|!\mathrm{for}sF[U]$$
In general a species of structure may not be molecular, in which case it is a sum of species:
Definition. Let $`F,G`$ be species of structures. Then their sum $`F+G`$ is the species defined by the disjoint union
$$(F+G)[U]=F[U]G[U],$$
and the transport along the bijection $`\sigma :UV`$ is given by:
$$(F+G)[\sigma ](s)=\{\begin{array}{cc}F[\sigma ](s)\hfill & \text{if }sF[U]\hfill \\ G[\sigma ](s)\hfill & \text{if }sG[U]\hfill \end{array}$$
The canonical decomposition of a species $`F`$ is its decomposition as a sum $`F=F_0+F_1+F_2+\mathrm{}`$ where $`F_n`$ denotes the $`nth`$ level of $`F`$:
$$F_n[U]=\{\begin{array}{cc}F[U]\hfill & \text{if }|U|=n\hfill \\ \mathrm{}\hfill & \text{if }|U|n\hfill \end{array}$$
The simplest species having structures at only one level is the species of singletons $`X`$:
$$X[U]=\{\begin{array}{cc}\{U\}\hfill & \text{if }|U|=1\hfill \\ \mathrm{}\hfill & \text{otherwise}\hfill \end{array}$$
Besides addition, there is a number of other operations between species by which to construct new species out of simpler ones. Following a standard notation , we use the sum symbol to denote disjoint reunion.
Definition. Let $`F,G`$ be two species of structures. We define the product species $`FG`$ as:
$$(FG)[U]=\underset{(U_1,U_2)}{}F[U_1]\times G[U_2]$$
where the sum runs over all partitions of the set $`U`$ into disjoint parts $`U_1`$ and $`U_2`$. The transport along the bijection $`\sigma :UV`$ of the structure $`s=(f,g)(FG)[U]`$ is:
$$(FG)[\sigma ](s)=(F[\sigma _1](f),G[\sigma _2](g))$$
where $`fF[U_1]`$ , $`gG[U_2]`$ and $`\sigma _1`$, $`\sigma _2`$ are the restrictions of $`\sigma `$ to the sets $`U_1`$ and $`U_2`$ respectively.
The stabilizer of $`s=(f,g)`$ is $`H_{(f,g)}=H_fH_g\mathrm{S}(U_1)\mathrm{S}(U_2)\mathrm{S}(U_1+U_2)`$.
As an example let us consider the n-th power of the species $`X`$ of singletons:
$$X^n[U]=\{\begin{array}{cc}\{(u_1,..,u_n)|u_iU,u_iu_j\mathrm{for}ij\}\hfill & \text{if }|U|=n\hfill \\ \mathrm{}\hfill & \text{otherwise}\hfill \end{array}$$
It is clear that the species $`X^n`$ and $`L_n`$ are essentially the same. Indeed there exists a natural bijection between $`X^n[U]`$ and $`L_n[U]`$:
$$(u_1,..,u_n)(u:nU:iu_i).$$
Remark: In the language of category theory, a species of structures $`F`$ is a functor from the category $`𝔹`$ of finite sets with bijections to the category $`𝔼`$ of finite sets with functions.
Definition. A morphism from the species of structures $`F`$ to the species $`G`$ is a natural transformation of functors, that is a family of functions $`m_U:F[U]G[U]`$ such that:
$$G[\sigma ]m_U=m_VF[\sigma ]\text{for all}\sigma :UV.$$
An isomorphism is an invertible morphism.
Definition. The cartesian product $`F\times G`$ of two species of structures $`F`$ and $`G`$ is given by:
$`(F\times G)[U]`$ $`=`$ $`F[U]\times G[U]`$ (1)
$`(F\times G)[\sigma ](f,g)`$ $`=`$ $`(F[\sigma ](f),G[\sigma ](g))`$
The canonical decomposition of the cartesian product is:
$$F\times G=\underset{n=0}{\overset{\mathrm{}}{}}F_n\times G_n.$$
(2)
A structure $`(f,g)(F\times G)[U]`$ has the stabilizer $`H_{(f,g)}=H_fH_g\mathrm{S}(U)`$.
An operation which will play an important role later is the derivation.
Definition. The derivative $`F^{}`$ of a species $`F`$ is a species whose set of structures over a finite set $`U`$ is given by:
$$F^{}[U]=F[U\{U\}]$$
and $`F^{}[\sigma ](s)=F[\sigma ^+](s)`$ where $`\sigma ^+:U\{U\}V\{V\}`$ is the extension of $`\sigma :UV`$:
$$\sigma ^+(x)=\{\begin{array}{cc}\sigma (x)\hfill & \text{if }xU\text{,}\hfill \\ V\hfill & \text{if }x=U\text{.}\hfill \end{array}$$
Remark: The term $`\{U\}`$ in $`U\{U\}`$ is just any additional point, not belonging to $`U`$. In particular for $`U=n`$ we have $`U\{U\}=n+1`$. If no confusion arises, we may write $`U\{U\}`$ as $`U+\{\}`$. The transport along bijections is the one inherited from the species $`F`$ but it is restricted to those transformations that keep the point $``$ fixed. The stabilizer of a structure $`s`$ when considered as a $`F^{}`$structure is different from its stabilizer as a $`F`$structure:
$$sF^{}[U]=F[U+\{\}]H_s^F^{}=H_s^F\mathrm{S}(U).$$
As explained in the introduction, we wish to compare successive levels of a species, i.e. to compare $`F`$ with $`F^{}`$. In this direction there is a small
###### Lemma 2.1
There are only two species (up to multiplicity) which satisfy $`F=F^{}`$.
Proof. Clearly, the species $`F`$ must have the same number of structures at all levels. For $`sF[n]`$, the stabiliser $`H_s`$ satisfies $`|H_s|\frac{n!}{|F[n]|}`$ which for $`n`$ big enough, reduces the possibilities to either the whole symmetric group $`\mathrm{S}(n)`$ or the subgroup $`A(n)`$ of even permutations. In the first case we obtain the species $`E`$ of sets which has only one structure at each level, in the second we have the species $`E^\pm `$ of oriented sets with exactly two structures at each level
$$E^\pm [U]=\{U_+,U_{}\}$$
the stabiliser of each structure being $`H_{U_\pm }=A(U)`$.
Besides these two ideal cases, we are interested in species $`F`$ whose structure at successive levels “resemble” each other. That means that $`F_n[U]`$ and $`F_{n+1}[U+\{\}]`$ should contain structures that behave similarly under permutations of $`U`$. Suppose that we are given a morphism $`m`$ from a subspecies $`F_1`$ of $`F^{}`$ to $`F`$ ($`F^{}=F_1+F_2`$). Then the $`F`$-structures which belong to the image of this morphism are similar to their preimages in the sense that their stabilizers contain those of their preimages. The action of the morphism $`m`$ can be encoded in a weight on the species $`F\times F^{}`$.
Definition. A weighted species $`(F,\omega )`$ consists of
1. a species of structures $`F`$
2. a family of functions $`\omega _U:F[U]`$ called weights,
such that for a bijection $`\sigma :UV`$ one has $`\omega _VF[\sigma ]=\omega _U`$.
The weight $`\omega _m`$ associated to the morphism $`m:F_1F`$ is the indicator function of its graph:
$$\omega _{m,U}(f,g)=\{\begin{array}{cc}\delta _{f,m(g)}\hfill & \text{if }gF_1[U],fF[U]\text{,}\hfill \\ 0\hfill & \text{if }gF_1[U]\text{.}\hfill \end{array}$$
One of the most interesting operations between species is the composition.
Definition. Let $`F`$ and $`G`$ two species of structures such that $`G[\mathrm{}]=\mathrm{}`$. The composition $`FG`$ is a species whose structures on a set $`U`$ are made in the following way:
1. make a partition $`\pi `$ of the set $`U`$;
2. choose an $`F`$-structure over the set $`\pi `$: $`fF[\pi ]`$;
3. for each $`p\pi `$ choose a structure $`g_pG[p]`$. Then the triple $`(\pi ,f,(g_p)_{p\pi })`$ is a structure in $`FG[U]`$. The transport along $`\sigma :UV`$ is the natural one.
In brief, an $`FG`$ structure is an $`F`$-assembly of $`G`$-structures. As an example consider the following combinatorial equation.
$$𝒜=XE(𝒜)$$
(3)
This equation implicitly defines the species $`𝒜`$ of rooted trees. Here is an explicit definition:
$`𝒜[U]`$ $`=`$ $`\{f:UU|_{uU}:f^k(u)\text{ is eventually constant}\};`$
$`𝒜[\sigma ]`$ $`:`$ $`f\sigma f\sigma ^1.`$
The constant is the root of the tree. The preimage of the root consists of roots of subtrees. One can thus consider the tree $`f`$ as the pair $`(\text{root}(f),\{f_a|af^1(\text{root}(f))\})`$ with $`f_a𝒜[U_a]`$ the subtree of $`f`$ with root $`a`$:
$$U_a=\{uU|k\text{such that}f^k=a\},f_a=f_{U_a}$$
We finally note
$$U=\{\text{root}(f)\}\underset{af^1(\text{root}(f))}{}U_a$$
thus completing the bijection between $`𝒜[U]`$ and $`XE(𝒜)[U]`$.
## 3 Fock Spaces and Analytic Functors
In this section we will describe how one can associate to a species of structures an endofunctor of the category of Hilbert spaces with contractions. We call the images of this functor symmetric spaces associated to the species $`F`$ and as we shall see in the following sections, they are suitable for constructing algebras of creation and annihilation operators, by exploiting the symmetry properties of the species $`F`$.
Following Joyal we define a special class of endofunctors of the category of sets with maps.
Definition. Let $`F[]`$ be a species of structures. The analytic functor $`F()`$ is an endofunctor of the category Set of sets with maps, defined by:
$$F(J)=\stackrel{~}{\underset{U}{}}F[U]\times J^U$$
(4)
where $`J^U=\{c|c:UJ\}`$ and the symbol $`\stackrel{~}{}_U`$ means the set of equivalence classes under bijective transformations:
$$F[U]\times J^U(s,c)(F[\sigma ](s),c\sigma ^1)F[V]\times J^V$$
for $`\sigma :UV`$. We call the elements of $`J`$ “colors”. Thus, an element in $`F(J)`$ is an orbit of J-colored F-structures denoted by $`[s,c]`$. Alternatively
$$F(J)=\stackrel{~}{\underset{U}{}}F[U]\times J^U=\underset{n=0}{\overset{\mathrm{}}{}}F[n]\times J^n/\mathrm{S}(n).$$
Remark. This relation can be viewed as a Taylor expansion of the set $`F(J)`$, which explains the name “analytic functor” for $`F()`$ .
Parallel to the functor $`F()`$ we define another endofunctor, this time on the category Hilb of Hilbert spaces with contractions. For any Hilbert space $``$ and a finite set $`U`$ we denote by $`^U`$ the Hilbert space arising from the positive definite kernel on $`^U`$ given by
$$k(\underset{uU}{}\psi _u,\underset{u^{}U}{}\phi _u^{})=\underset{uU}{}\psi _u,\phi _u.$$
For every bijection $`\sigma :UV`$ there is a unitary transformation $`U(\sigma ):^U^V`$ obtained by linear extension of:
$$U(\sigma ):\underset{uU}{}\psi _u\underset{vV}{}\psi _{\sigma ^1(v)}.$$
Definition. Let $`F`$ be a species of structures. For each Hilbert space $`𝒦`$ we construct the symmetric Hilbert space
$$\mathrm{\Gamma }_F(𝒦):=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}_{\mathrm{symm}}^2(F[n]𝒦^n)$$
(5)
where the subscript “symm” denotes the invariance under the natural action of the symmetric group $`\mathrm{S}(n)`$:
$$\mathrm{\Psi }U(\sigma )\mathrm{\Psi }F[\sigma ^1].$$
The factor $`\frac{1}{n!}`$ refers to the inner product on $`\mathrm{}_{\mathrm{symm}}^2`$.
Remark. There is an equivalent way of writing $`\mathrm{\Gamma }_F(𝒦)`$:
$$\mathrm{\Gamma }_F(𝒦)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}^2(F[n])_{\mathrm{S}(n)}𝒦^n$$
(6)
where $`_{\mathrm{S}(n)}`$ means that we consider only the subspace of the tensor product whose vectors are invariant under the action of $`\mathrm{S}(n)`$.
Let us choose an orthonormal basis $`(e_j)_{jJ}`$ for the Hilbert space $`𝒦`$. Let $`(e_c)_{cJ^n}`$ be the basis of $`𝒦^n`$ given by $`e_c:=_{jn}e_{c(j)}`$, and
$`\gamma _{F,J}:F(J)`$ $``$ $`[0,\mathrm{})`$
$`\gamma _{F,J}([s,c])`$ $`=`$ $`|H_{(s,c)}|`$
where $`[s,c]`$ denotes the orbit of the colored structure $`(s,c)`$.
###### Lemma 3.1
There is a unitary equivalence between $`\mathrm{}^2(F(J),\gamma _{F,J})`$ and $`\mathrm{\Gamma }_F(𝒦)`$, given by
$$U\delta _{[s,c]}=\delta _s_{\mathrm{S}(n)}e_c:=\underset{\sigma \mathrm{S}(n)}{}\delta _{F[\sigma ](s)}e_{c\sigma ^1}$$
Proof. Considering $`𝒦^n`$ as $`\mathrm{}^2(J^n)`$ we may write
$`U\delta _{[s,c]}={\displaystyle \underset{\sigma \mathrm{S}(n)}{}}\delta _{F[\sigma ](s)}e_{c\sigma ^1}`$
$`={\displaystyle \underset{\sigma \mathrm{S}(n)}{}}\delta _{F[\sigma ](s),c\sigma ^1}=|H_{(s,c)}|\mathrm{𝟏}_{[s,c]}`$
It follows that
$`U\delta _{[s,c]}^2`$ $`=`$ $`{\displaystyle \frac{1}{n!}}|H_{(s,c)}|^2|[s,c]|=|H_{(s,c)}|`$
$`=`$ $`\delta _{[s,c]}^2.`$
Since the functions $`\mathrm{𝟏}_{[s,c]}`$ span the space $`\mathrm{}_{\mathrm{symm}}^2(F[n]\times J^n)`$, the operator $`U`$ is surjective and hence unitary.
Remark. For a constant coloring $`c`$ we have $`\delta _{[s,c]}^2=|H_s|`$, whereas for all colors different, $`\delta _{[s,c]}^2=1`$.
Certain operations with species of structures extend to analytic species and to the symmetric spaces: addition, multiplication and substitution.
1) Addition. As $`(F+G)(J)`$ is the disjoint union of $`F(J)`$ and $`G(J)`$, we have
$$\mathrm{\Gamma }_{F+G}(𝒦)=\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_G(𝒦).$$
2) Multiplication. Similarly, we have
$`(FG)(J)=\stackrel{~}{{\displaystyle \underset{U}{}}}({\displaystyle \underset{U_1+U_2=U}{}}F[U_1]\times G[U_2])\times J^U`$
$`=`$ $`\stackrel{~}{{\displaystyle \underset{U}{}}}{\displaystyle \underset{U_1+U_2=U}{}}F[U_1]\times G[U_2]\times J^{U_1+U_2}=\stackrel{~}{{\displaystyle \underset{U_1,U_2}{}}}F[U_1]\times G[U_2]\times J^{U_1+U_2}`$
$`=`$ $`F(J)\times G(J).`$
which suggests the following unitary transformation from $`\mathrm{\Gamma }_{FG}(𝒦)`$ to
$`\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_G(𝒦)`$:
$$T:\delta _{[(f,g),c]}\delta _{[f,c_1]}\delta _{[g,c_2]}$$
for $`fF[n]`$, $`gG[m]`$, $`cJ^{m+n}`$ and $`c_1,c_2`$ the restrictions of $`c`$ to $`n`$ respectively $`m`$. Indeed the map preserves orthogonality and is isometric:
$`\delta _{[(f,g),c]}^2`$ $`=`$ $`|H_{((f,g),c)}|=|H_{(f,c_1)}H_{(g,c_2)}|=|H_{(f,c_1)}||H_{(g,c_2)}|`$
$`=`$ $`\delta _{[f,c_1]}^2\delta _{[g,c_2]}^2`$
From now on we will consider $`\mathrm{\Gamma }_{FG}(𝒦)`$ and $`\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_G(𝒦)`$ as identical, without mentioning the unitary $`T`$.
3) Substitution. we start with the analytic functors:
$`(FG)(J)`$ $`=`$ $`\stackrel{~}{{\displaystyle \underset{U}{}}}(FG)[U]\times J^U`$
$`=\stackrel{~}{{\displaystyle \underset{U}{}}}(\stackrel{~}{{\displaystyle \underset{\pi }{}}}F[\pi ]\times G^\pi [U])\times J^U`$ $`=`$ $`\stackrel{~}{{\displaystyle \underset{\pi }{}}}(F[\pi ]\times \stackrel{~}{{\displaystyle \underset{U}{}}}G^\pi [U])\times J^U`$
$`=\stackrel{~}{{\displaystyle \underset{\pi }{}}}F[\pi ]\times G^\pi (J)`$ $`=`$ $`F(G(J))`$
where we have used $`G^\pi (J)=G(J)^\pi `$, which follows from the multiplication property. At the level of symmetric spaces we have the unitary transformation from $`\mathrm{\Gamma }_F(\mathrm{\Gamma }_G(𝒦))`$ to $`\mathrm{\Gamma }_{FG}(𝒦)`$:
$$T:\delta _{[f,C]}\delta _{[f,(g_a)_{a\pi },c]}$$
with the following relations for the structures appearing above: $`fF[\pi ]`$, $`C:\pi G(J)`$ such that $`C(a)=[g_a,c_a]`$, and $`c_a=c_a`$. Let us check the isometric property:
$`\delta _{[f,C]}^2`$ $`=`$ $`{\displaystyle \underset{a\pi }{}}\delta _{C(a)}^2|H_{(f,C)}|={\displaystyle \underset{a\pi }{}}\delta _{[g_a,c_a]}^2|H_{f,C}|`$
$`=`$ $`{\displaystyle \underset{a\pi }{}}|H_{(g_a,c_a)}||H_{f,C}|=\delta _{[f,(g_a)_{a\pi },c]}^2`$
Symmetric Fock space. The symmetric Hilbert space associated to the species of sets is the well known symmetric Fock space:
$$\mathrm{\Gamma }_E(𝒦)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}_{\mathrm{symm}}^2(E[n]𝒦^n)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}𝒦^{_sn}.$$
Full Fock space. For the linear orders we obtain the full Fock space:
$$\mathrm{\Gamma }_L(𝒦)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}_{\mathrm{symm}}^2(L[n]𝒦^n)=\underset{n=0}{\overset{\mathrm{}}{}}𝒦^n.$$
Antisymmetric Fock space. We recall from lemma 2.1 that the species $`E^\pm `$ of oriented sets has two structures at all levels
$$E^\pm [U]=\{U_+,U_{}\}$$
which are mapped into each other by odd permutations and have as stabiliser the group $`A(U)`$ of even permutation. The representation of $`\mathrm{S}(n)`$ on $`\mathrm{}^2(E^\pm [n])`$ contains two one-dimensional irreducible sub-representations, the symmetric and the antisymmetric representation. Accordingly the symmetric Hilbert space associated to $`E^\pm `$ is the direct sum of the symmetric and antisymmetric Fock spaces:
$$\mathrm{\Gamma }_{E^\pm }(𝒦)=\mathrm{\Gamma }_s(𝒦)\mathrm{\Gamma }_a(𝒦).$$
Remark. The set of species as defined in the previous section can be enlarged by defining the virtual species as equivalence classes of pair of species of structures under the equivalence relation:
$$(F_1,G_1)(F_2,G_2)F_1+G_2=F_2+G_1$$
One denotes the equivalence class of $`(F,G)`$ by $`FG`$. Thus we can say that the antisymmetric Fock space is associated to the virtual species $`E^\pm E`$.
## 4 Creation and Annihilation Operators
In this section we will describe a general framework for constructing \*-algebras of operators on symmetric Hilbert spaces by giving the action of the generators of these algebras, the creation and annihilation operators. In particular in the case of the species of sets $`E`$ and linear orderings $`L`$, we obtain the well known canonical commutation relations algebra (C.C.R.), respectively the algebra of creation/annihilation operators on the full Fock space.
The starting point is the observation that the operation of derivation of species of structures can be interpreted as removal of point $``$ from a structure. This makes it possible to define operators between the symmetric Hilbert spaces of a species of structure $`F`$ and its derivative $`F^{}`$.
We will consider now “colored” $`F`$-structures. Let $`J`$ be the set of “colors” and $`iJ`$. We have the map
$$a^{}(i):F^{}[U]\times J^UF[U+\{\}]\times J^{U+\{\}}$$
such that
$$a^{}(i):(s,c)(s,c_i^+)$$
where $`c_i^+:U+\{\}J`$ is given by:
$$c_i^+(u)=\{\begin{array}{cc}c(u)\hfill & \text{if }uU\hfill \\ i\hfill & \text{if }u=\hfill \end{array}$$
As we did in section 3, we pass to the set of orbits of $`J`$-colored $`F`$-structures. The map $`a^{}(i)`$ projects to a well defined map from $`F^{}(J)`$ to $`F(J)`$:
$`a^{}(i):F^{}[U]\times J^U/\mathrm{S}(U)`$ $``$ $`F[U+\{\}]\times J^{U+\{\}}/\mathrm{S}(U+\{\})`$
$`a^{}(i):[s,c]`$ $``$ $`[s,c_i^+]`$
But as $`F(J)`$ determines an orthogonal basis of the space $`\mathrm{\Gamma }_F(𝒦)`$ for $`(e_j)_{jJ}`$ orthogonal basis in $`𝒦`$, we can extend $`a^{}(i)`$ by linearity to an operator
$$a^{}(i):\mathrm{\Gamma }_F^{}(𝒦)\mathrm{\Gamma }_F(𝒦).$$
The adjoint of $`a^{}(i)`$ acts in the opposite direction:
$$a(i):\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_F^{}(𝒦).$$
The problem with this definition is that in general the species $`F`$ and $`F^{}`$ are distinct which means that one cannot take the “field operators” $`a^{}(i)+a(i)`$ and only certain products of creation and annihilation operators are well defined. In section 2 we pointed out that the “similarity” of the structures of the species $`F`$ and $`F^{}`$ can be encoded in a weight on the cartesian product $`F\times F^{}`$. Let $`\omega `$ be such a weight. Then $`\omega _U:F[U]\times F^{}[U]`$ such that for all $`sF[U],tF^{}[U]`$ and $`\sigma :UW`$ we have:
$$\omega (s,t)=\omega (F[\sigma ](s),F^{}[\sigma ](t))$$
We will use this to define creation and annihilation operators which act on the same space $`\mathrm{\Gamma }_F(𝒦)`$. In the sequel we will refer to the pair $`(\mathrm{\Gamma }_F(𝒦),\omega )`$ as combinatorial Fock space.
Definition. a) The annihilation operator (before symmetrization) associated to the species $`F`$ and weight $`\omega `$ is defined by:
$$\stackrel{~}{a}(h):\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)$$
$$(\stackrel{~}{a}(h)\phi )(f)=\underset{gF[n+1]}{}\omega (f,g)\mathrm{inp}_n(h,\phi (g))$$
where $`fF[n],h𝒦`$ and $`\mathrm{inp}_k(h,)`$ is the operator:
$$\mathrm{inp}_k(h,\psi _0..\psi _n)=h,\psi _k\psi _0..\psi _{k1}\psi _{k+1}..\psi _n$$
for $`k\{0,1,..,n\}`$.
b) The creation operator (before symmetrization) is:
$$\stackrel{~}{a}^{}(h):\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)$$
$$(\stackrel{~}{a}^{}(h)\phi )(f)=(n+1)\underset{gF[n]}{}\overline{\omega (g,f)}\mathrm{tens}_n(h,\phi (g))$$
where $`fF[n+1],h𝒦`$ and $`\mathrm{tens}_k(h,)`$ is the operator:
$$\mathrm{tens}_k(h,\psi _0..\psi _{n1})=\psi _0..\psi _{k1}h\psi _k..\psi _{n1}$$
for $`k\{0,1,..,n\}`$.
Remark. In order to avoid domain problems for $`\stackrel{~}{a},\stackrel{~}{a}^{}`$, we will restrict to weights which are bounded, $`|\omega (t,s)|C`$ for all $`t,s`$. Then
$$a(h)\psi _nn^{\frac{1}{2}}Ch\psi _na^{}(h)\psi _n(n+1)^{\frac{1}{2}}Ch\psi _n$$
for $`\psi _n\mathrm{}^2(F[n]𝒦^n)`$ thus $`a(h),a^{}(h)`$ have well defined extensions to the domain $`D(N^{\frac{1}{2}})`$ $`(N\psi _n=n\psi _n)`$. As this will not play a major role here, we will omit specifying the domain, usually the the vectors considered should belong to $`D(N^{\frac{1}{2}})`$.
We consider now the symmetrized creation and annihilation operators which act on the symmetric Hilbert space and which are the main object of our investigation.
###### Lemma 4.1
The unsymmetrized annihilation operator $`\stackrel{~}{a}(h)`$ restricts to a well defined operator $`a(h)`$ on the symmetric Hilbert space $`\mathrm{\Gamma }_F(𝒦)`$:
Proof. Let $`\phi \mathrm{\Gamma }_F(𝒦)`$. Then $`\phi (F[\sigma ](f))=U(\sigma )\phi (f)`$ for all $`\sigma \mathrm{S}(n),fF[n]`$ and
$`(a(h)\phi )(F[\sigma ](f))`$ $`=`$ $`{\displaystyle \underset{g}{}}\omega (F[\sigma ](f),g)\mathrm{inp}_n(h,\phi (g))`$
$`={\displaystyle \underset{g^{}}{}}\omega (f,g^{})\mathrm{inp}_n(h,\phi (F[\stackrel{~}{\sigma }]g^{}))`$ $`=`$ $`{\displaystyle \underset{g^{}}{}}\omega (f,g^{})U(\sigma )\mathrm{inp}_n(h,\phi (g^{}))`$
$`=`$ $`U(\sigma )(a(h)\phi )(f)`$
where $`\stackrel{~}{\sigma }:n+1n+1`$ is given by
$$\stackrel{~}{\sigma }(i)=\{\begin{array}{cc}\sigma (i)\hfill & \text{if }in\hfill \\ n\hfill & \text{if }i=n\hfill \end{array}$$
###### Lemma 4.2
The operator $`\stackrel{~}{a}^{}(h)`$ is the adjoint of $`\stackrel{~}{a}(h)`$ on the unsymmetrized space $`_{n=0}^{\mathrm{}}\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)`$.
Proof. Let $`\phi ,\psi `$ be two vectors in $`_{n=0}^{\mathrm{}}\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)`$ and $`\phi _n,\psi _n\frac{1}{n!}\mathrm{}^2(F[n]𝒦^n)`$ their components on level $`n`$.
Then we have:
$`\psi ,\stackrel{~}{a}^{}(h)\phi ={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{(n+1)!}}{\displaystyle \underset{g}{}}(n+1)\psi _{n+1}(g),(\stackrel{~}{a}^{}(h)\phi _n)(g)`$
$`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!}}{\displaystyle \underset{f,g}{}}\psi _{n+1}(g),\overline{\omega (f,g)}\phi _n(f)h`$
$`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!}}{\displaystyle \underset{f,g}{}}\omega (f,g)\mathrm{inp}_n(h,\psi _{n+1}(g)),\phi _n(f)`$
$`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!}}{\displaystyle \underset{f}{}}(\stackrel{~}{a}(h)\psi _{n+1})(f),\phi _n(f)=\stackrel{~}{a}(h)\psi ,\phi `$
We will restrict our attention to the action of the creation and annihilation operators on the symmetric Hilbert space $`\mathrm{\Gamma }_F(𝒦)`$. From Lemma 4.1 the annihilation operator $`a(h)`$ is well defined on $`\mathrm{\Gamma }_F(𝒦)`$. We call its adjoint on the symmetric Hilbert space, the symmetrized creation operator. If $`P`$ is the projection to $`\mathrm{\Gamma }_F(𝒦)`$ then the symmetrized creation operator is:
$$a^{}(h)\phi =P\stackrel{~}{a}^{}(h)\phi \text{ for }\phi \mathrm{\Gamma }_F(𝒦)$$
###### Lemma 4.3
Let $`fF[n+1]`$ and $`\tau _{n,k}\mathrm{S}(n+1)`$ the transposition of $`n`$ and $`k`$. Then for any $`\phi \mathrm{\Gamma }_F(𝒦)`$ the action of the symmetrized creation operator has the expression:
$$(a^{}(h)\phi )(f)=\underset{k=0}{\overset{n}{}}\underset{g}{}\overline{\omega (g,F[\tau _{n,k}](f))}U(\tau _{n,k})(\phi (g)h).$$
Proof. We have:
$`(P\stackrel{~}{a}^{}(h)\phi )(f)={\displaystyle \frac{1}{(n+1)!}}{\displaystyle \underset{\sigma \mathrm{S}(n+1)}{}}U(\sigma )(\stackrel{~}{a}^{}(h)\phi )(F[\sigma ^1]f)`$ (7)
$`=`$ $`{\displaystyle \frac{1}{n!}}{\displaystyle \underset{\sigma \mathrm{S}(n+1)}{}}{\displaystyle \underset{g}{}}\overline{\omega (g,F[\sigma ](f))}U(\sigma ^1)(\phi (g)h)`$
If $`\sigma \mathrm{S}(n+1)`$ and $`\sigma ^1(n)=k`$ then $`\rho =\tau _{n,k}\sigma ^1\mathrm{S}(n)`$. Thus the sum over all permutations can be split into a sum over $`kn+1`$ and one over $`\mathrm{S}(n)`$. Moreover, from the definitions of $`\mathrm{\Gamma }_F(𝒦)`$ and that of a weight we know that
$`U(\rho )\phi (g)`$ $`=`$ $`\phi (F[\rho ](g))`$
$`\omega (g,F[\rho ^1\tau _{n,k}](f))`$ $`=`$ $`\omega (F[\rho ](g),F[\tau _{n,k}](f))`$
which substituted into the sum (7) gives:
$`{\displaystyle \frac{1}{n!}}{\displaystyle \underset{k=0}{\overset{n}{}}}{\displaystyle \underset{\rho \mathrm{S}(n)}{}}{\displaystyle \underset{g}{}}\overline{\omega (F[\rho ](g),F[\tau _{n,k}](f))}U(\tau _{n,k})(\phi (F[\rho ](g))h)`$
$`=`$ $`{\displaystyle \underset{k=0}{\overset{n}{}}}{\displaystyle \underset{g^{}}{}}\overline{\omega (g^{},F[\tau _{n,k}](f))}U(\tau _{n,k})(\phi (g^{})h).`$
Sometimes algebras are defined by giving relations among generators as for example commutation relations. We will give next explicit formulas for the product of a creation and an annihilation operator.
###### Lemma 4.4
Let $`fF[n]`$ and $`\phi \mathrm{\Gamma }_F(𝒦)`$. Then
$$(a^{}(h_1)a(h_2)\phi )(f)=\underset{k=0}{\overset{n1}{}}\underset{f^{}}{}(\overline{\omega }\omega )_k(f,f^{})\text{tens}_k(h_1,\text{inp}_k(h_2,\phi (f^{})))$$
(8)
where we have made the notation
$$(\overline{\omega }\omega )_k(f,f^{})=\underset{g}{}\overline{\omega (g,F[\tau _{n1,k}](f))}\omega (g,F[\tau _{n1,k}](f^{}))$$
(9)
Proof. By applying successively the definitions of $`a^{}(h_1)`$ and $`a(h_2)`$ we have:
$`(a^{}(h_1)a(h_2)\phi )(f)={\displaystyle \underset{k=0}{\overset{n1}{}}}{\displaystyle \underset{g}{}}\overline{\omega (g,F[\tau _{n1,k}](f))}U(\tau _{n1,k})(a(h_2)\phi )(g)h_1`$
$`={\displaystyle \underset{k=0}{\overset{n1}{}}}{\displaystyle \underset{g,f^{}}{}}\overline{\omega (g,F[\tau _{n1,k}](f))}\omega (g,f^{})U(\tau _{n1,k})(\text{inp}_{n1}(h_2,\phi (f^{}))h_1)`$
$`={\displaystyle \underset{k=0}{\overset{n1}{}}}{\displaystyle \underset{g,f^{}}{}}\overline{\omega (g,F[\tau _{n1,k}](f))}\omega (g,F[\tau _{n1,k}]f^{})\text{tens}_k(h_1,\text{inp}_k(h_2,\phi (f^{})))`$
$`={\displaystyle \underset{k=0}{\overset{n1}{}}}{\displaystyle \underset{f^{}}{}}(\overline{\omega }\omega )_k(f,f^{})\text{tens}_k(h_1,\text{inp}_k(h_2,\phi (f^{}))).`$
###### Lemma 4.5
Let $`fF[n]`$ and $`\phi \mathrm{\Gamma }_F(𝒦)`$. Then:
$`(a(h_1)a^{}(h_2)\phi )(f)`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{n1}{}}}{\displaystyle \underset{f^{}}{}}(\omega \overline{\omega })_k(f,f^{})\mathrm{tens}_k(h_2,\mathrm{inp}_k(h_1,\phi (f^{})))`$
$`+`$ $`h_1,h_2{\displaystyle \underset{f^{}}{}}(\omega \overline{\omega })_n(f,f^{})\phi (f^{})`$
where we have made the notation
$$(\omega \overline{\omega })_k(f,f^{})=\underset{g}{}\omega (f,g)\overline{\omega (f^{},F[\tau _{n,k}](g))}$$
(10)
Proof. We use the definitions of $`a(h_1)`$ and $`a^{}(h_2)`$:
$`(a(h_1)a^{}(h_2)\phi )(f)={\displaystyle \underset{g}{}}\omega (f,g)\text{inp}_n(h_1,(a^{}(h_2)\phi )(g))`$
$`=`$ $`{\displaystyle \underset{g,f^{}}{}}{\displaystyle \underset{k=0}{\overset{n}{}}}\omega (f,g)\overline{\omega (f^{},F[\tau _{n,k}](g))}\text{inp}_n(h_1,U(\tau _{n,k})(\phi (f^{})h_2))`$
$`=`$ $`{\displaystyle \underset{k=0}{\overset{n1}{}}}{\displaystyle \underset{f^{}}{}}(\omega \overline{\omega })_k(f,f^{})\text{tens}_k(h_2,\text{inp}_k(h_1,\phi (f^{})))`$
$`+`$ $`h_1,h_2{\displaystyle \underset{f^{}}{}}(\omega \overline{\omega })_n(f,f^{})\phi (f^{})`$
### 4.1 Examples.
We will describe a few known operator algebras in the language developed so far and a new algebra based on the species $`𝒜`$ of rooted trees.
1) Sets: The combinatorial Fock space is $`(E,\omega _E)`$ with $`E[U]=\{U\}`$ and $`\omega (\{U\},\{U+\{\}\})=1`$. We use lemmas 4.4 and 4.5 to calculate the commutator of the creation and annihilation operator:
$`(a(h_1)a^{}(h_2)a^{}(h_2)a(h_1))\phi (f)=h_1,h_2(\omega \overline{\omega })_n(f,f^{})\phi (f^{})`$
$`+`$ $`{\displaystyle \underset{k=0}{\overset{n1}{}}}((\omega \overline{\omega })_k(f,f^{})(\overline{\omega }\omega )_k(f,f^{}))\text{tens}_k(h_2,\text{inp}_k(h_1,\phi (f^{}))`$
But $`(\omega \overline{\omega })_k(f,f^{})=(\overline{\omega }\omega )_k(f,f^{})=(\omega \overline{\omega })_n(f,f^{})=\delta _{f,f^{}}`$ for all $`kn`$ which implies the C.C.R.:
$$a(h_1)a^{}(h_2)a^{}(h_2)a(h_1)=h_1,h_2\mathrm{𝟏}$$
In particular it is clear that $`\mathrm{\Gamma }_E(𝒦)`$ is the symmetric Fock space over the Hilbert space $`𝒦`$.
2) Linear Orders: Let $`(L,\omega _L)`$ be the combinatorial Fock space with
$$L[U]=\{f:U\{0,1,..,|U|1\}\}$$
and
$$\omega _L(f,g)=\delta _{f,g_U}\text{for}fL[U],gL[U+\{\}]$$
where
$$\delta _{f,g_U}=\{\begin{array}{cc}1\hfill & \text{if }f(u)=g(u)\text{ for }uU\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}$$
From (10) we have:
$$(\omega \overline{\omega })_k(f,f^{})=\underset{g}{}\delta _{f,g_U}\delta _{f^{},L[\tau _{n,k}](g)_U}=\delta _{k,n}\delta _{f,f^{}}$$
Then by applying Lemma 4.5 we obtain
$$a(h_1)a^{}(h_2)=h_1,h_2\mathrm{𝟏}$$
which characterizes the algebra of creation and annihilation operators on the full Fock space .
3) Oriented Sets: We refer to the previous sections for the definition of the species $`E^\pm `$ of oriented sets. The weight $`\omega _{E^\pm }`$ is given by
$`\omega _{E^\pm }(U_+,U_+^{})`$ $`=`$ $`\omega _{E^\pm }(U_{},U_{}^{})=1,`$
$`\omega _{E^\pm }(U_+,U_{}^{})`$ $`=`$ $`\omega _{E^\pm }(U_{},U_+^{})=0`$ (11)
where $`U^{}=U+\{\}`$. With the help of the “switching” sign operator
$$𝐠\phi (\pm )=\phi ()$$
we obtain the $`𝐠`$-commutation relations:
$$a(h_1)a^{}(h_2)𝐠a^{}(h_2)a(h_1)=h_1,h_2\mathrm{𝟏}$$
(12)
As we saw in the previous section, the space $`\mathrm{\Gamma }_{E^\pm }(𝒦)`$ is isomorphic to the direct sum of the symmetric and antisymmetric Fock space over $`𝒦`$:
$$\mathrm{\Gamma }_{E^\pm }(𝒦)=\mathrm{\Gamma }_s(𝒦)\mathrm{\Gamma }_a(𝒦)$$
through the transformation:
$$\phi _s=\phi (+)+\phi ()\text{and}\phi _a=\phi (+)\phi ()$$
then the $`𝐠`$-commutation relations can be written equivalently as:
$$(a(h_1)a^{}(h_2)a^{}(h_2)a(h_1))\phi _s=h_1,h_2\phi _s$$
and
$$(a(h_1)a^{}(h_2)+a^{}(h_2)a(h_1))\phi _a=h_1,h_2\phi _a$$
4) Rooted Trees: We recall the definition of the species $`𝒜`$:
$$𝒜[U]=\{f:UU|f^k(u)=\text{root}(f)U\text{for}k|U|,uU\}$$
with the transport along $`\sigma `$: $`𝒜[\sigma ](f)=\sigma f\sigma ^1`$. We note that $`𝒜[\mathrm{}]=\mathrm{}`$. We consider a natural weight which can be described as follows: it takes value 1 on those pairs of trees for which the second is obtained by adding a leaf to the first one, and takes value 0 for the rest. Thus for $`t_1𝒜[U]`$ and $`t_2𝒜[U+\{\}]`$ the weight is:
$$\omega _𝒜(t_1,t_2)=\left\{\begin{array}{cc}1\hfill & \text{if }t_1(u)=t_2(u)\text{ for }uU\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}\right\}:=\delta _{t_1,t_2_U}$$
We will compute the commutator of $`a(h_2)`$ with $`a^{}(h_1)`$. For this we need to obtain the expressions of $`(\overline{\omega }\omega )_k(,)`$ and $`(\omega \overline{\omega })_k(,)`$. We start with
$`(\omega \overline{\omega })_n(f,f^{})`$ $`=`$ $`{\displaystyle \underset{g}{}}\omega _𝒜(f,g)\overline{\omega _𝒜(f^{},g)}={\displaystyle \underset{g}{}}\delta _{f,g_n}\delta _{f^{},g_n}`$ (13)
$`=`$ $`\delta _{f,f^{}}{\displaystyle \underset{g}{}}\delta _{f,g_n}=n\delta _{f,f^{}}`$
The factor $`n`$ appears because there are $`n`$ possible way of attaching a leaf to the tree $`f`$ each one giving a tree $`g`$ such that $`g_n=f`$. For $`k<n`$ we have
$$(\omega \overline{\omega })_k(t,t^{})=\underset{g}{}\omega _𝒜(t,g)\overline{\omega _𝒜(t^{},𝒜[\tau _{n,k}](g))}=\underset{g}{}\delta _{t,g_n}\delta _{t^{},𝒜[\tau _{n,k}](g)_n}.$$
At most one term in this sum is different from zero, for the tree $`g`$ satisfying:
$$\{\begin{array}{cc}g(i)=t(i)\hfill & \text{if }in\hfill \\ g(j)=t^{}(j)\hfill & \text{if }jn\{k\}\hfill \\ g(n)=t^{}(k)\hfill & \end{array}$$
(14)
On the other hand
$$(\overline{\omega }\omega )_k(t,t^{})=\underset{g^{}}{}\overline{\omega _𝒜(g^{},𝒜[\tau _{n1,k}](t))}\omega _𝒜(g^{},𝒜[\tau _{n1,k}](t^{}))$$
$$=\underset{g}{}\delta _{g^{},𝒜[\tau _{n1,k}](t)_{n1}}\delta _{g^{},𝒜[\tau _{n1,k}](t^{})_{n1}}=\delta _{𝒜[\tau _{n1,k}](t)_{n1},𝒜[\tau _{n1,k}](t^{})_{n1}}$$
$$=\{\begin{array}{cc}1\hfill & \text{if }t(i)=t^{}(i)\text{ for all }in,ik\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}$$
(15)
Finally from (14), (15) we conclude that $`(\overline{\omega }\omega )_k(t,t^{})=(\omega \overline{\omega })_k(t,t^{})`$ for $`k\{0,1,..n1\}`$.
Let us define the “vertex number” operator $`N`$ by
$$(N\phi )(t)=n\phi (t)$$
for $`t𝒜[n]`$. The usual commutation relations between $`N`$ and the creation operator hold:
$$[N,a^{}(h)]=a^{}(h)$$
By using Lemmas 4.4 and 4.5 we obtain the following:
###### Theorem 4.1
The following commutation relations hold on the combinatorial Fock space $`(𝒜,\omega _𝒜)`$:
$$a(h_1)a^{}(h_2)a^{}(h_2)a(h_1)=Nh_1,h_2$$
(16)
Remark: Notice that the vacuum of $`\mathrm{\Gamma }_𝒜(𝒦)`$ is an eigenvector of $`N`$ with eigenvalue 1.
5) Simple Directed Graphs: Let us define a species whose structures are directed graphs for which any pair of vertices is connected by at most one edge:
$$𝒟_s[U]=\{gU\times U|(u,v)g(v,u)g\}$$
(17)
where the transport along $`\sigma `$ is given by $`\sigma \times \sigma `$.
Let $`g_1𝒟_s[U]`$ and $`g_2𝒟_s[U+\{\}]`$. Then $`\omega (g_1,g_2)0`$ if and only if $`g_2`$ contains $`g_1`$ as a subset and all edges of $`g_2`$ connecting the vertex $``$ with vertices in $`U`$ are oriented from $``$ to $`U`$. We make the following convenient notation for the set of edges going out of a vertex $`a`$ of $`g𝒟_s[V]`$ :
$$v_a(g)=\{(a,v)|(a,v)g\}=\{a\}\times e_a(g)$$
The weight $`\omega ^{𝒟_s,q}`$ depends on the real parameter $`0q1`$ and is defined by:
$$\omega ^{𝒟_s,q}(g_1,g_2)=\delta _{g_2,g_1+v_{}(g_2)}(q^{|U||v_{}(g_2)|}(1q)^{|v_{}(g_2)|})^{\frac{1}{2}}$$
In the rest of this section we prove that $`(𝒟_s,\omega ^{𝒟_s,q})`$ is a realization of the q-commutation relations ().
###### Theorem 4.2
On $`(𝒟_s,\omega ^{𝒟_s,q})`$ we have:
$$a(h_1)a^{}(h_2)qa^{}(h_2)a(h_1)=h_1,h_2$$
Proof. We employ lemmas 4.5 and 4.4. First:
$`(\omega \overline{\omega })_n(f,f^{})={\displaystyle \underset{g}{}}\omega ^{𝒟_s,q}(f,g)\overline{\omega ^{𝒟_s,q,}(f^{},g)}`$
$`={\displaystyle \underset{g}{}}\delta _{g,f+v_n(g)}\delta _{g,f^{}+v_n(g)}q^{n|v_n(g)|}(1q)^{|v_n(g)|}`$
$`=\delta _{f,f^{}}{\displaystyle \underset{v_nn}{}}q^{n|v_n|}(1q)^{|v_n|}=\delta _{f,f^{}}`$ (18)
It remains to be proved that $`(\omega \overline{\omega })_k(f,f^{})=q(\overline{\omega }\omega )_k(f,f^{})`$ for
$`k\{0,1,..n1\}`$ and $`f,f^{}𝒟_s[n]`$.
In the sum
$$(\omega \overline{\omega })_k(f,f^{})=\underset{g}{}\omega ^{𝒟_s,q}(f,g)\overline{\omega ^{𝒟_s,q,}(f^{},𝒟_s[\tau _{n,k}](g))}$$
the only nonzero contribution comes from $`g𝒟_s[U+\{\}]`$ such that:
$$g=f+\{n\}\times e_n(g)=(fv_k(f))+v_k(f)+\{n\}\times e_n(g)$$
and
$`𝒟_s[\tau _{n,k}](g)`$ $`=`$ $`(fv_k(f))+\{n\}\times e_k(f)+\{k\}\times e_n(g)`$
$`=`$ $`f^{}+\{n\}\times e_k(f)`$
which together imply
$$(fv_k(f))+\{k\}\times e_n(g)=f^{}.$$
But this means that $`e_n(g)=e_k(f^{})`$ and $`fv_k(f)=f^{}v_k(f^{})`$. Then we have the expression
$$(\omega \overline{\omega })_k(f,f^{})=\delta _{fv_k(f),f^{}v_k(f^{})}q^{n\frac{|v_k(f)|+|v_k(f^{})|}{2}}(1q)^{\frac{|v_k(f)|+|v_k(f^{})|}{2}}$$
(19)
On the other hand in $`(\overline{\omega }\omega )_k(f,f^{})`$ we get only the contribution from those $`g`$ for which:
$$g^{}=𝒟_s[\tau _{n1,k}](f)_{n1}=𝒟_s[\tau _{n1,k}](f^{})_{n1}$$
Thus we obtain
$$(\overline{\omega }\omega )_k(f,f^{})=\delta _{fv_k(f),f^{}v_k(f^{})}q^{n1\frac{|v_k(f)|+|v_k(f^{})|}{2}}(1q)^{\frac{|v_k(f)|+|v_k(f^{})|}{2}}.$$
(20)
Finally from (19) and (20) we have the desired expression:
$$(\omega \overline{\omega })_k(f,f^{})=q(\overline{\omega }\omega )_k(f,f^{})$$
## 5 Fock States and Operations with Combinatorial Fock Spaces
The operations between species of structures described in Section 2 are helpful in understanding the action of creation and annihilation operators in terms of elementary ones. The guiding example is Green’s representation of the operators appearing in parastatistics, as sums of bosonic (fermionic) operators with the “wrong” commutation relations . Similar ideas appear in where the author considers macroscopic fields as linear combinations of basic bosonic fields with various commutation relations.
Thus, the first question we address in this section is the following: given two combinatorial Fock spaces $`(F,\omega _F)`$ and $`(G,\omega _G)`$, is there a natural natural weight associated to the species $`F+G`$, $`FG`$, $`F\times G`$, $`FG`$ ? The second question is related to the notion of positive definite functions on pair partitions. A general theory of such functions has been introduced in in connection with the so called generalized Brownian motion.
### 5.1 Fock States
We will start with the latter question by introducing the necessary definitions.
Definition. Let $`S`$ be a finite ordered set. We denote by $`𝒫_2(S)`$ is the set of pair partitions of $`S`$, that is $`𝒱𝒫_2(S)`$ if $`𝒱=\{V_1,..,V_r\}`$ where each $`V_i`$ is an ordered set containing two elements $`V_i=(k_i,l_i)`$ with $`k_i,l_iS`$, $`k_i<l_i`$ and $`\{V_1,..,V_r\}`$ is a partition of $`S`$ ($`V_iV_j=\mathrm{}`$ for $`ij`$ and $`_{i=1}^rV_i=S`$). The set of all pair partitions is
$$𝒫_2(\mathrm{})=\underset{r=1}{\overset{\mathrm{}}{}}𝒫_2(2r).$$
Let $`𝒦`$ be a Hilbert space. We denote by $`𝒞_𝒦`$ the -algebra obtained from the free algebra with generators $`c(f)`$ and $`c^{}(f)`$, ($`f𝒦`$) divided by the relations:
$$c^{}(\lambda f_1+\mu f_2)=\lambda c^{}(f_1)+\mu c^{}(f_2),\lambda ,\mu ,f_1,f_2𝒦,$$
and
$$c^{}(f)=(c(f))^{}.$$
We are interested in a particular type of positive functionals on $`𝒞_𝒦`$, called Fock states which have the following expression on monomials of creation and annihilation operators:
$$\rho _t(c^\mathrm{}_1(f_1)..c^\mathrm{}_n(f_n))=\{\begin{array}{cc}0\hfill & \text{if }n\text{ odd}\hfill \\ _{𝒱=\{V_1,..,V_{\frac{n}{2}}\}}\rho _t[V_1]..\rho _t[V_{\frac{n}{2}}]t(𝒱)\hfill & \text{if }n\text{ even}\hfill \end{array}$$
(21)
the sum running over all pair partitions $`𝒱`$ in $`𝒫_2(2r)`$, and the symbols $`\mathrm{}_i`$ standing for creation or annihilation. For $`V=(k,l)𝒱`$
$$\rho _t[V]=f_k,f_lQ(\mathrm{}_k,\mathrm{}_l)$$
with the 2 by 2 covariance matrix
$$Q=\left(\begin{array}{ccc}\rho (c_ic_i)& \rho (c_ic_i^{})& \\ \rho (c_i^{}c_i)& \rho (c_i^{}c_i^{})& \end{array}\right)=\left(\begin{array}{ccc}0& 1& \\ 0& 0& \end{array}\right).$$
where $`c_i=c(e_i)`$ and $`e_i`$ is an arbitrary normalized vector in $`𝒦`$.
Let us consider a real subspace $`𝒦_{}`$ of $`𝒦`$ such that $`𝒦=𝒦_{}i𝒦_{}`$. The sub-algebra of $`𝒞_𝒦`$ generated by the “field operators” $`\omega (f)=c(f)+c^{}(f)`$ with $`f𝒦_{}`$ is denoted by $`𝒜_𝒦`$. If the restriction of the functional $`\rho _t`$ to the algebra $`𝒜_𝒦𝒞_𝒦`$ is a state, then we call the function
$$t:𝒫_2(\mathrm{}).$$
positive definite . In particular if $`\rho _t`$ is a state on $`𝒞_𝒦`$ then $`t`$ is positive definite. The converse is not true in general. We will show first that the vacuum state of a symmetric Hilbert space is an example of Fock state.
###### Proposition 5.1
Let $`(F,\omega _F)`$ be a combinatorial Fock space with $`|F[\mathrm{}]|=1`$. Let $`\mathrm{\Omega }_F`$ be the vacuum vector of $`\mathrm{\Gamma }_F(𝒦)`$, $`𝒦`$ a Hilbert space. Then the functional $`\rho _F()=\mathrm{\Omega }_F,\mathrm{\Omega }_F`$ is a Fock state on $`𝒞_𝒦`$.
In order to prove this proposition we need to introduce one more tool.
Definition. Let $`𝒦`$ be a Hilbert space, $`(F,\omega _F)`$ a combinatorial Fock space and $`A(𝒦)`$ a bounded operator on $`𝒦`$. The second quantization of $`A`$ is defined by
$`d\mathrm{\Gamma }_F(A):\mathrm{\Gamma }_F(𝒦)`$ $``$ $`\mathrm{\Gamma }_F(𝒦)`$
$`(d\mathrm{\Gamma }_F(A)\phi )(g)`$ $`=`$ $`d\mathrm{\Gamma }(A)(\phi (g))`$
where the meaning of $`d\mathrm{\Gamma }(A)`$ on the right side is
$$d\mathrm{\Gamma }(A):𝒦^n𝒦^n$$
$$d\mathrm{\Gamma }(A):\psi _0..\psi _{n1}\underset{k=0}{\overset{n1}{}}\psi _0..A\psi _k..\psi _{n1}$$
Remark. The second quantization operator is a well defined operator on $`\mathrm{\Gamma }_F(𝒦)`$. Indeed let $`\phi \mathrm{\Gamma }_F(𝒦)`$ and $`\tau \mathrm{S}(n)`$, then
$`(d\mathrm{\Gamma }_F(A)\phi )(F[\tau ](g))`$ $`=`$ $`d\mathrm{\Gamma }(A)\phi (F[\tau ](g))`$
$`=d\mathrm{\Gamma }(A)U(\tau )\phi (g)`$ $`=`$ $`U(\tau )(d\mathrm{\Gamma }_F(A)\phi )(g)`$
and thus $`d\mathrm{\Gamma }_F(A)\phi \mathrm{\Gamma }_F(𝒦)`$. We have used the invariance of $`d\mathrm{\Gamma }(A)`$ under permutations:
$$d\mathrm{\Gamma }(A)U(\tau )=U(\tau )d\mathrm{\Gamma }(A).$$
###### Lemma 5.1
We have the following commutation relations:
$`[a(h),d\mathrm{\Gamma }_F(A)]`$ $`=`$ $`a(A^{}h)`$ (22)
$`[d\mathrm{\Gamma }_F(A),d\mathrm{\Gamma }_F(B)]`$ $`=`$ $`d\mathrm{\Gamma }_F([A,B])`$ (23)
Proof. Let $`\phi \mathrm{\Gamma }_F(𝒦)`$, $`fF[n],gF[n+1]`$. Then
$`(a(h)d\mathrm{\Gamma }(A)\phi )(f)`$ $`=`$ $`{\displaystyle \underset{g}{}}\omega (f,g)\mathrm{inp}_n(h,(d\mathrm{\Gamma }(A)\phi )(g))`$
$`=`$ $`{\displaystyle \underset{g}{}}\omega (f,g)\mathrm{inp}_n(h,{\displaystyle \underset{k=0}{\overset{n}{}}}\mathrm{𝟏}..A..\mathrm{𝟏}\phi (g))`$
$`=`$ $`{\displaystyle \underset{g}{}}\omega (f,g){\displaystyle \underset{k=0}{\overset{n1}{}}}\mathrm{𝟏}..A..\mathrm{𝟏}\mathrm{inp}_n(h,\phi (g))`$
$`+{\displaystyle \underset{g}{}}\omega (f,g)\mathrm{inp}_n(h,\mathrm{𝟏}..A\phi (g))`$
$`=`$ $`({\displaystyle \underset{k=0}{\overset{n1}{}}}\mathrm{𝟏}..A..\mathrm{𝟏}){\displaystyle \underset{g}{}}\omega (f,g)\mathrm{inp}_n(h,\phi (g))`$
$`+{\displaystyle \underset{g}{}}\omega (f,g)\mathrm{inp}_n(A^{}h,\phi (g))`$
$`=`$ $`d\mathrm{\Gamma }_F(A)a(h)+a(A^{}h)`$
which proves (22). The other commutator follows directly from the definition of the second quantization operator.
###### Lemma 5.2
Let $`\{e_j\}_{jJ}`$ be an orthonormal basis of $`𝒦`$ and denote $`a_j^{\mathrm{}}=a(e_j)^{\mathrm{}}`$. Then the following equation holds:
$$a_i\underset{k=1}{\overset{n}{}}a_{i_k}^\mathrm{}_k\mathrm{\Omega }=\underset{k=1}{\overset{n}{}}\delta _{i,i_k}\delta _{\mathrm{}_k,}a_{i_0}\underset{p=1}{\overset{k1}{}}a_{i_p}^\mathrm{}_pa_{i_0}^{}\underset{q=k+1}{\overset{n}{}}a_{i_q}^\mathrm{}_q\mathrm{\Omega }$$
(24)
where the colors $`(i_k)_{k=0,..,n}`$ satisfy the property $`i_ki_0`$ for all $`k=1,..n`$.
Proof. For simplicity we denote $`\mathrm{\Psi }=_{k=1}^na_{i_k}^\mathrm{}_k\mathrm{\Omega }`$. We notice that $`a_{i_0}\mathrm{\Psi }=0`$ due to the assumption that $`i_ki_0`$ for all $`k=1,..n`$. Then using (22) we get
$$a_i\mathrm{\Psi }=[a_{i_0},d\mathrm{\Gamma }(|i_0i|)]\mathrm{\Psi }=a_{i_0}d\mathrm{\Gamma }(|i_0i|)\mathrm{\Psi }$$
(25)
By successively applying the following commutator
$$[d\mathrm{\Gamma }(|i_0i|),a_{i_k}^\mathrm{}_k]=\delta _{i_k,i}\delta _{\mathrm{}_k,}a_{i_0}^{}$$
we obtain the sum in (24).
Proof of Proposition 5.1. It is clear that $`\rho _F`$ is a positive linear functional on $`𝒞_𝒦`$. We need to prove that it has the expression (21). From linearity of the creation operators and anti-linearity of the annihilation operators we conclude that it is sufficient to consider the vectors $`f_i`$ in (21) belonging to the chosen orthogonal basis. From
$$\rho _F(\underset{k=1}{\overset{n}{}}a_{i_k}^\mathrm{}_k)=\mathrm{\Omega },\underset{k=1}{\overset{n}{}}a_{i_k}^\mathrm{}_k\mathrm{\Omega }$$
and considering the fact that the creation operator increases the level by one while the annihilation operator decreases it by one, we deduce that nonzero expectations can appear only if $`n`$ is even and the number of creators in the monomial $`_{k=1}^na_{i_k}^\mathrm{}_k`$ is equal to that of annihilators. Furthermore $`a_{i_1}^\mathrm{}_1`$ must be an annihilator and $`a_{i_n}^\mathrm{}_n`$, a creator. We will thus consider that this is the case.
We put the monomial in the form $`a_{i_1}_{k=2}^na_{i_k}^\mathrm{}_k`$ and apply lemma 5.2. We obtain a sum over all pairs $`(a_{i_1},a_{i_k}^{})`$ of the same color ($`i_1=i_k`$) and replace $`i_1`$ by a new color $`i_0`$. We go now to the next annihilator in each term of the sum and repeat the procedure, the new color which we add this time being different from all the colors used previously. After $`\frac{n}{2}`$ steps we obtain a sum containing all possible pairings of annihilators and creators of the same color in $`_{k=1}^na_{i_k}^\mathrm{}_k`$:
$$\rho _F(\underset{k=1}{\overset{n}{}}a_{i_k}^\mathrm{}_k)=\underset{𝒱=\{V_1,..,V_{\frac{n}{2}}\}}{}\underset{p=1}{\overset{\frac{n}{2}}{}}\delta _{i_{k_p},i_{l_p}}Q(\mathrm{}_{k_p},\mathrm{}_{l_p})t(𝒱)$$
with $`V_p=(k_p,l_p)`$ and $`t(𝒱)`$ is given by
$$t(𝒱)=\rho _F(\underset{k=1}{\overset{n}{}}a_{j_k}^\mathrm{}_k)$$
such that $`j_{k_p}=j_{l_p^{}}`$ if and only if $`p=p^{}`$, for $`p,p^{}\{1,..,\frac{n}{2}\}`$, $`\mathrm{}_{k_p}`$ is annihilator and $`\mathrm{}_{l_p}`$ is creator.
Thus for each combinatorial Fock space $`(F,\omega _F)`$ (which has a vacuum), the vacuum state is described by a positive definite function $`t_F`$ on $`𝒫_2(\mathrm{})`$.
Remark. We observe that the result can be generalized to a larger range of states and monomials. Let us partition the index set $`J`$ of the orthonormal basis of the Hilbert space $`𝒦`$
$$J=J_1+J_2$$
and choose a state $`\rho _\mathrm{\Phi }()=\mathrm{\Phi },\mathrm{\Phi }`$ and monomials $`_{k=1}^na_{j_k}^\mathrm{}_k`$ such that $`j_kJ_1`$ and $`\mathrm{\Phi }\mathrm{\Gamma }_F(𝒦_2)\mathrm{\Gamma }_F(𝒦)`$ is a normalized vector where $`𝒦_2`$ is the subspace of $`𝒦`$ with the orthogonal basis $`\{e_j\}_{jJ_2}`$. Then it is easy to see that the argument used in the above proof still holds and $`\rho _\mathrm{\Phi }`$ is a Fock state for $`𝒞_{𝒦_1}`$. In general $`\rho _\mathrm{\Phi }`$ and $`\rho _F`$ do not coincide. When they do coincide we say that $`\rho _F`$ has the pyramidal independence property .
### 5.2 Operations with symmetric Hilbert Spaces
We pass now to the first question which we have posed in the beginning of this section. The various operations between species offer the opportunity of creating new symmetric Hilbert spaces which sometimes give rise to interesting interpolations between the two members. For the definitions of the operations we refer back to Section 2.
1) Sums. Let $`(F,\omega _F)`$ and $`(G,\omega _G)`$ be two combinatorial Fock spaces. From Section 3 we know that
$$\mathrm{\Gamma }_{F+G}=\mathrm{\Gamma }_F\mathrm{\Gamma }_G.$$
Note that the vacuum of $`\mathrm{\Gamma }_{F+G}`$ has dimension $`2`$ if $`F[\mathrm{}]\mathrm{}G[\mathrm{}]`$. The natural weight on $`F+G`$ is
$$\omega _{F+G}(t_1,t_2)=\omega _F(t_1,t_2)+\omega _G(t_1,t_2)$$
which gives rise to operators
$$a_{F+G}(h)=a_F(h)0+0a_G(h).$$
We consider a linear combination of the two vacua (for $`|F[\mathrm{}]|=|G[\mathrm{}]|=1)`$
$$\mathrm{\Omega }_\lambda =\sqrt{\lambda }\mathrm{\Omega }_F+\sqrt{1\lambda }\mathrm{\Omega }_G.$$
The corresponding state $`\rho _{F+G,\lambda }()=\mathrm{\Omega }_\lambda ,\mathrm{\Omega }_\lambda `$ interpolates linearly between $`\rho _F`$ and $`\rho _G`$
$$\rho _{F+G,\lambda }=\lambda \rho _F+(1\lambda )\rho _G$$
and the same is true for the positive definite functions
$$t_{F+G,\lambda }=\lambda t_F+(1\lambda )t_G.$$
(26)
1) Products. Let $`(F,\omega _F)`$ and $`(G,\omega _G)`$ be two combinatorial Fock spaces. We consider the product species $`FG`$. As we have proved in Section 3, there is the following isomorphism
$$\mathrm{\Gamma }_{FG}(𝒦)=\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_G(𝒦).$$
(27)
Again there is a natural weight for the species $`FG`$. For $`fF[U_1],gG[U_2],f^{}F^{}[U_1],g^{}G^{}[U_2]`$:
$`\omega _{FG,\lambda }((f,g),(f,g^{}))`$ $`=`$ $`\sqrt{\lambda }\omega _G(g,g^{})`$
$`\omega _{FG,\lambda }((f,g),(f^{},g))`$ $`=`$ $`\sqrt{1\lambda }\omega _F(f,f^{})`$
all other values of $`\omega _{FG,\lambda }`$ being 0.
From (27) and the expression of $`\omega _{FG}`$ we obtain
$$a_{FG}^{\mathrm{}}(h)=\sqrt{\lambda }a_F^{\mathrm{}}(h)\mathrm{𝟏}+\sqrt{1\lambda }\mathrm{𝟏}a_G^{\mathrm{}}(h)$$
If $`|F[\mathrm{}]|=|G[\mathrm{}]|=1`$ then the state $`\rho _{FG}()=\mathrm{\Omega }_F\mathrm{\Omega }_G,\mathrm{\Omega }_F\mathrm{\Omega }_G`$ generates the positive definite function:
$$t_{FG}(𝒱)=\underset{𝒱_1,𝒱_2}{}\lambda ^{|𝒱_1|}(1\lambda )^{|𝒱_2|}t_F(𝒱_1)t_G(𝒱_2)$$
where the sum runs over all partitions of $`𝒱`$ in two sets, $`𝒱_1`$ and $`𝒱_2`$.
Example: The Green representation of the (Fermi) parastatistics of order p is an example of application of the product of species. We consider the p-th power $`(E_\pm )^p`$ of the species of oriented sets $`E_\pm `$. Then the annihilation operators are
$$a(h)=\frac{1}{\sqrt{p}}\underset{k=1}{\overset{p}{}}a^{(k)}(h)$$
and the vacuum state is $`\rho ()=\mathrm{\Omega },\mathrm{\Omega }`$ where $`a^{(k)}`$ is the term corresponding to the k-th term in the product and
$$\mathrm{\Omega }=\mathrm{\Omega }_a^{(1)}..\mathrm{\Omega }_a^{(p)}$$
is the tensor product of the antisymmetric vacua of each of the species $`E_\pm ^{(k)}`$.
3) Cartesian Products. Let $`(F,\omega _F)`$ and $`(G,\omega _G)`$ be two combinatorial Fock spaces. We consider the cartesian product species $`F\times G`$. The corresponding weight has the expression:
$$\omega _{F\times G}((f,g),(f^{},g^{}))=\omega _F(f,f^{})\omega _G(g,g^{})$$
We note that $`\omega _{F\times G}`$ satisfies the invariance condition stated in the definition of the weight.
###### Proposition 5.2
Let $`(F,\omega _F)`$ and $`(G,\omega _G)`$ be two combinatorial Fock spaces both having a single structure on $`\mathrm{}`$. Then the positive definite function associated to the vacuum state of $`(F\times G,\omega _{F\times G})`$ satisfies:
$$t_{F\times G}(𝒱)=t_F(𝒱)t_G(𝒱)$$
(28)
for all $`𝒱𝒫_2(\mathrm{})`$.
Proof. We construct a linear operator $`T`$ from $`\mathrm{\Gamma }_{F\times G}(𝒦)`$ to $`\mathrm{\Gamma }_F(𝒦)\mathrm{\Gamma }_G(𝒦)`$ with the property that its restriction to a certain subspace $`\mathrm{\Gamma }_{F\times G}^{\mathrm{ext}}`$ of $`\mathrm{\Gamma }_{F\times G}(𝒦)`$, is an isometry. The subspace $`\mathrm{\Gamma }_{F\times G}^{\mathrm{ext}}`$ is spanned by vectors $`\delta _{[(f,g),c]}`$ of the orthogonal basis $`(F\times G)(J)`$ of $`\mathrm{\Gamma }_{F\times G}(𝒦)`$ which have all colors different from each other, i.e. $`c(i)c(j)`$ for $`ij`$. We refer to Section 3 for the definitions related to the orthogonal basis of $`\mathrm{\Gamma }_{F\times G}(𝒦)`$.
The action of $`T`$ on the basis vectors is:
$`T:\mathrm{\Gamma }_{F\times G}^{\mathrm{ext}}(𝒦)`$ $``$ $`\mathrm{\Gamma }_F^{\mathrm{ext}}(𝒦)\mathrm{\Gamma }_G^{\mathrm{ext}}(𝒦)`$
$`\delta _{[(f,g),c]}`$ $``$ $`\delta _{[f,c]}\delta _{[g,c]}`$
We check that the operator is well defined. Indeed the map
$`i:{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}(F\times G)[n]\times J^n`$ $``$ $`({\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}F[n]\times J^n)\times ({\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}G[n]\times J^n)`$
$`((f,g),c)`$ $``$ $`((f,c),(g,c))`$
commutes with the action of $`\mathrm{S}(n)`$ on the two sides, at each level and thus projects to a well defined map on the quotient:
$`i:(F\times G)(J)`$ $``$ $`F(J)\times G(J)`$
$`[(f,g),c]`$ $``$ $`([f,c],[g,c])`$
This means that $`T`$ is well defined. But as we have shown in Section 3, the vectors $`\delta _{[(f,g),c]}`$, $`\delta _{[f,c]}`$ and $`\delta _{[g,c]}`$ for which $`c(i)c(j)`$ if $`ij`$, have norm one which implies that $`T`$ is an isometry.
Let us now consider the vector
$$\phi _F^{(p)}=\underset{k=1}{\overset{p}{}}a_{F,i_k}^\mathrm{}_k\mathrm{\Omega }_F$$
the colors $`(i_k)_{k=1,..,n}`$ satisfying the condition that there are no three identical colors, and if there exists $`k_1<k_2`$ such that $`i_{k_1}=i_{k_2}`$, then $`a_{i_{k_1}}^{\mathrm{}_{k_1}}=a_{i_{k_1}}`$ and $`a_{i_{k_2}}^{\mathrm{}_{k_2}}=a_{i_{k_2}}^{}`$. It is clear that $`\phi _F^{(p)}\mathrm{\Gamma }_F^{\mathrm{ext}}(𝒦)`$. Analogously we define $`\phi _G^{(p)}`$ and $`\phi _{F\times G}^{(p)}`$. We want to prove by induction w.r.t. $`p`$ that the action of the isometry $`T`$ is such that
$$T:\phi _{F\times G}^{(p)}\phi _F^{(p)}\phi _G^{(p)}.$$
(29)
This implies in particular (28), when the monomial $`_{k=1}^pa_{i_k}^\mathrm{}_k`$ contains equal number of creators and annihilators pairing each other according to color, no two pairs having the same color.
For $`p=0`$ we have $`T(\mathrm{\Omega }_{F\times G})=\mathrm{\Omega }_F\mathrm{\Omega }_G`$. Suppose (29) holds for p. Then there are two possibilities for increasing the length of the monomial $`_{k=1}^pa_{i_k}^\mathrm{}_k`$ by one: either by adding on the first position a creation operator $`a_{i_0}^{}`$ such that the color $`i_0`$ does not appear in the rest of the monomial, or by adding an annihilation operator $`a_{i_0}`$ such that the term $`a_{i_0}^{}`$ appears once in the rest of the monomial. We treat the two cases separately.
1.) suppose that we have $`\phi _{F\times G}^{(p)}=_{k=1}^pa_{i_k}^\mathrm{}_k\mathrm{\Omega }_{F\times G}`$, $`i_0i_k`$ which has the decomposition
$$\phi _{F\times G}^{(p)}=\underset{[(f,g),c]}{}\phi ([(f,g),c])\delta _{[(f,g),c]}$$
with $`\phi ([(f,g),c])`$. Then
$$a_{F\times G,i_0}^{}\phi _{F\times G}^{(p)}=\underset{[(f,g),c],(f^{}g^{})}{}\phi ([(f,g),c])\omega _{F\times G}((f,g),(f^{},g^{}))\delta _{[(f^{},g^{}),c_{i_0}^+]}$$
which implies
$`T(a_{F\times G,i_0}^{}\phi _{F\times G}^{(p)})`$ $`=`$ $`{\displaystyle \underset{[(f,g),c],(f^{}g^{})}{}}\phi ([(f,g),c])a_{F,i_0}^{}\delta _{[f,c]}a_{G,i_0}^{}\delta _{[g,c]}`$
$`=a_{F,i_0}^{}a_{G,i_0}^{}T(\phi _{F\times G}^{(p)})`$ $`=`$ $`a_{F,i_0}^{}\phi _F^{(p)}a_{G,i_0}^{}\phi _G^{(p)}.`$
2.) suppose that we have $`\phi _{F\times G}^{(p)}=_{k=1}^pa_{i_k}^\mathrm{}_k\mathrm{\Omega }_{F\times G}`$ such that the term $`a_{i_0}^{}`$ appears exactly one time in the the monomial $`_{k=1}^pa_{i_k}^\mathrm{}_k`$. We use again the Fourier decomposition
$$\phi _{F\times G}^{(p)}=\underset{[(f,g),c]}{}\phi ([(f,g),c])\delta _{[(f,g),c]}$$
and identify in each orbit $`[(f,g),c](F\times G)(J)`$, a representant $`((f,g),c)(F\times G)[n]\times J^n`$ such that $`c(n1)=i_0`$. Then
$$a_{F\times G,i_0}\phi _{F\times G}^{(p)}=\underset{[(f,g),c],(f^{}g^{})}{}\phi ([(f,g),c])\omega _{F\times G}((f^{},g^{}),(f,g))\delta _{[(f^{},g^{}),c_{i_0}^{}]}$$
where $`c_{i_0}^{}`$ is the restriction of $`c`$ to the set $`n1`$. Finally
$`T(a_{F\times G,i_0}\phi _{F\times G}^{(p)})`$ $`=`$ $`{\displaystyle \underset{[(f,g),c],(f^{}g^{})}{}}\phi ([(f,g),c])\omega _{F\times G}((f^{},g^{}),(f,g))T(\delta _{[(f^{},g^{}),c_{i_0}^{}]})`$
$`=`$ $`{\displaystyle \underset{[(f,g),c],(f^{}g^{})}{}}\phi ([(f,g),c])\omega _F(f,f^{})\omega _G(g,g^{})\delta _{[f^{},c_{i_0}^{}]}\delta _{[g^{},c_{i_0}^{}]}`$
$`=`$ $`a_{F,i_0}\phi _F^{(p)}a_{G,i_0}\phi _G^{(p)}`$
which proves the induction hypothesis for $`p+1`$ and the proposition.
Application: Combining the result of the previous proposition and certain variations on the species of rooted trees, we investigate more general commutation relations of the type:
$$[a(h_1),a^{}(h_2)]=h_1,h_2f(N)$$
with $`f:𝐍𝐑`$ and $`N`$ the number operator characterized by
$$[N,a^{}(h)]=a^{}(h).$$
###### Theorem 5.1
Let $`P`$ be a real polynomial with positive coefficients. Then the commutation relations
$$[a(h_1),a^{}(h_2)]=h_1,h_2P(N)$$
are realizable on a symmetric Hilbert space.
We split the proof in a few lemmas.
###### Lemma 5.3
Let $`(F,\omega _F)`$ and $`(G,\omega _G)`$ be two symmetric Hilbert spaces for which the commutation relations hold
$`[a_F(h_1),a_F^{}(h_2)]`$ $`=`$ $`h_1,h_2a(N)`$
$`[a_G(h_1),a_G^{}(h_2)]`$ $`=`$ $`h_1,h_2b(N)`$
where $`a,b`$ are real functions. Then on $`(F\times G,\omega _{F\times G})`$ we have
$$[a_{F\times G}(h_1),a_{F\times G}^{}(h_2)]=h_1,h_2(ab)(N)$$
Proof. This is a direct application of Lemmas 4.4, 4.5 and the following equations:
$`(\omega \overline{\omega })_k((f,g),(f^{},g^{}))`$ $`=`$ $`(\omega \overline{\omega })_k(f,f^{})(\omega \overline{\omega })_k(g,g^{})`$
$`(\overline{\omega }\omega )_k((f,g),(f^{},g^{}))`$ $`=`$ $`(\overline{\omega }\omega )_k(f,f^{})(\overline{\omega }\omega )_k(g,g^{})`$
###### Lemma 5.4
Let $`𝒜`$ be the species of rooted trees. Let $`f𝒜[U]`$,
$`g𝒜[U+\{\}]`$ and
$$\stackrel{~}{\omega }_𝒜^c(f,g)=\omega _𝒜(f,g)+c^{\frac{1}{2}}\delta _{f_{},g}$$
a modification of the weight $`\omega _𝒜`$ defined in section 4, with $`c`$, a positive constant. The structure $`f_{}𝒜[U+\{\}]`$ is defined by:
$$f_{}(u)=\{\begin{array}{cc}f(u)\hfill & \text{if }u\text{root}(f)\hfill \\ \hfill & \text{if }u=\text{root}(f)\text{.}\hfill \end{array}$$
Then on $`(𝒜,\stackrel{~}{\omega }_𝒜^c)`$ we have:
$$[a(h_1),a^{}(h_2)]=h_1,h_2(N+c).$$
Proof. This is similar to the proof of Theorem 4.1, with an additional contribution to $`(\omega \overline{\omega })_n(f,g)`$ coming from the term $`c^{\frac{1}{2}}\delta _{f_{},g}`$ in $`\stackrel{~}{\omega }_𝒜^c`$.
###### Lemma 5.5
Let $`𝒜\times 𝒜`$ be the species of ordered pairs of rooted trees. We define the weight
$$\omega _{𝒜\times 𝒜}^c((f,g),(f^{},g^{}))=\omega _𝒜(f,g)\omega _𝒜(f^{},g^{})+c^{\frac{1}{2}}\delta _{f_{},f^{}}\delta _{g_{},g^{}}.$$
Then on $`(𝒜\times 𝒜,\omega _{𝒜\times 𝒜}^c)`$ we have
$$[a(h_1),a^{}(h_2)]=h_1,h_2(N^2+c).$$
Proof. Similar to the previous two lemmas.
Proof of Theorem 5.1. The polynomial $`P(x)`$ has a canonical expression as product of polynomials of the type $`x+c`$ and $`x^2+c`$ with $`c0`$. The theorem follows by applying the previous 3 lemmas.
Remark. The result can be extended to power series with positive coefficients and infinite radius of convergence. In particular for $`0q1`$
$$s(x)=q^x=\underset{k=0}{\overset{\mathrm{}}{}}\frac{1}{n!}(\mathrm{log}q)^nx^n$$
gives the commutation relations
$$[a_i,a_j^{}]=q^N\delta _{i,j}$$
which characterize the $`q`$-deformations , , up to a “rescaling” of the creation and annihilation operators with a function of $`N`$.
4) Compositions. Let $`(F,\omega _F)`$ and $`(G,\omega _G)`$ be two combinatorial Fock spaces. We recall that the composition of $`G`$ in $`F`$ is a species whose structures are $`F`$-assemblies of $`G`$-structures:
$$FG[U]=\underset{\pi }{}F[\pi ]\times \underset{p\pi }{}G[p].$$
We would like to define the annihilation and creation operators for the species $`FG`$ by making use of the available weights $`\omega _F`$ and $`\omega _G`$. Apart from the condition $`|G[\mathrm{}]|=0`$ we require $`|G[1]|=1`$. We consider an arbitrary structure $`(f,\pi ,(g_p)_{p\pi })FG[U]`$ where $`\pi `$ is a partition of the finite set $`U`$. Then we note that there are two essentially different possibilities to “add” a new point $``$, to the set $`U`$: one can enlarge the size of $`\pi `$ by creating a partition of $`U+\{\}`$ of the form $`\pi ^+=\pi +\{\{\}\}`$, or one can keep the size of $`\pi `$ constant by adding $``$ to one of the sets $`p\pi `$ and obtain the partition $`\pi _p^+`$. Between $`\pi `$ and $`\pi _p^+`$ there is the bijection
$$\alpha _p:p^{}\{\begin{array}{cc}p^{}\hfill & \text{if }p^{}p\hfill \\ p+\{\}\hfill & \text{if }p^{}=p\hfill \end{array}$$
We recognize that in the first case the weight $`\omega _F`$ should play a role, while in the second, the weight $`\omega _G`$. According to the properties of the species $`F`$, one can further distinguish among the subsets to which $``$ is added, by choosing (as we did for the creation and annihilation operators) a weight $`\omega _{F,ϵ}`$ on the cartesian product $`F\times ϵ`$ where $`ϵ`$ is the species of elements: $`ϵ[U]=U`$. Putting together the three data $`(\omega _F,\omega _G,\omega _{F,ϵ})`$, we define:
$`\omega _{FG}((f,\pi ,(g_p)_{p\pi }),(f^{},\pi ^{},(g_p^{}^{})_{p^{}\pi ^{}}))=\omega _F(f,f^{}){\displaystyle \underset{p\pi }{}}\delta _{g_p,g_p^{}}`$
$`+`$ $`{\displaystyle \underset{p\pi }{}}\delta _{f^{},F[\alpha _p](f)}\omega _{F,ϵ}(f,p)\omega _G(g_p,g_{p+\{\}}^{}).`$
where $`fF[\pi ],g_pG[p]`$, etc.
Remark.We find this definition rather natural and broad enough to cover some interesting examples. One can easily check that $`\omega _{FG}`$ satisfies the invariance property characterizing the weights.
Example: The species Bal of ordered partitions or Ballots is the composition of $`L`$ (the species of linear orderings ), with $`E_+`$ (the species of nonempty sets). A typical structure over a finite set $`U`$ looks like: $`s=(U_1,..,U_k)`$ with $`(U_p)_{p\{1,..,k\}}`$, a partition of $`U`$. The vacuum is the empty sequence $`\text{Bal}[\mathrm{}]=\{\mathrm{}\}`$. We use the weights $`\omega _E`$ and $`\omega _L`$ as defined in section 4. The action of the creation operator at the combinatorial level can be described as follows: we can add the point $``$ in the last subset of the sequence $`s=(U_1,..,U_k)`$ and obtain $`s_k^+=(U_1,..,U_k+\{\})`$, or we can create a new subset $`U_{k+1}=\{\}`$ and position it at the end of the sequence $`s`$, producing $`s^+=(U_1,..,U_{k+1})`$. We see that in this case the weight $`\omega _{L,ϵ}`$ is simply identifying the last element of the sequence: $`\omega _{L,ϵ}(s,U_k)=\delta _{U_k,U_p}`$. For the vacuum we set $`\omega _{\mathrm{Bal}}(\{\mathrm{}\},\{\})=1`$. We use $`0q1`$ as an interpolation parameter:
$$\omega _{\mathrm{Bal}}(s,s^{})=q^{\frac{1}{2}}\delta _{s_k^+,s^{}}+(1q)^{\frac{1}{2}}\delta _{s^+,s^{}}$$
(30)
Let us denote by $`t_{\mathrm{Bal}}`$ the positive definite function associated to the vacuum state of the combinatorial Fock space $`(\text{Bal},\omega _{\mathrm{Bal}})`$, as defined in subsection 5.1. Following we associate to any pair partition $`𝒱𝒫_2(\mathrm{})`$ a set $`B(𝒱)=\{𝒱_1,..,𝒱_k\}`$ such that $`𝒱=𝒱_1\mathrm{}𝒱_k`$ is the decomposition of $`𝒱`$ into connected sub-partitions or blocks.
###### Theorem 5.2
Let $`𝒱𝒫_2(\mathrm{})`$. Then
$$t_{\mathrm{Bal}}(𝒱)=q^{|𝒱||B(𝒱)|}$$
(31)
Proof. We split the task of proving (31) into two simpler ones: first we prove the strong multiplicativity property for $`t`$:
$$t(𝒱)=\underset{i=1}{\overset{k}{}}t(𝒱_i)\mathrm{if}B(𝒱)=\{𝒱_1,..,𝒱_k\}$$
and then for $`𝒱`$ consisting of a single block, $`t_{\mathrm{Bal}}(𝒱)=q^{|𝒱|1}`$. The proof of the strong multiplicativity is analogous to that of Proposition 5.2. We consider an orthogonal basis $`(e_j)_{jJ}`$ of the Hilbert space $`𝒦`$ and a partition $`J=J_1+J_2`$ of $`J`$ with the corresponding relation $`𝒦=𝒦_1𝒦_2`$. We define an isometry
$$S:\mathrm{\Gamma }_{\mathrm{Bal}}^{\mathrm{ext}}(𝒦_1)\mathrm{\Gamma }_{\mathrm{Bal}}^{\mathrm{ext}}(𝒦_2)\mathrm{\Gamma }_{\mathrm{Bal}}^{\mathrm{ext}}(𝒦_1𝒦_2)$$
and we will prove that it has a natural action on monomials of creation and annihilation operators:
$$S(\underset{k}{}a_{i_k}^\mathrm{}_k\mathrm{\Omega }\underset{p}{}a_{j_p}^\mathrm{}_p\mathrm{\Omega })=\underset{k}{}a_{i_k}^\mathrm{}_k\underset{p}{}a_{j_p}^\mathrm{}_p\mathrm{\Omega }.$$
(32)
We recall that the two monomials satisfy certain properties which are described in Proposition 5.2. The multiplicativity of $`t_{\mathrm{Bal}}`$ follows from equation (32) and the isometric property of $`S`$.
The action of $`S`$ on the orthogonal bases defined in Section 3 is:
$$\delta _{[s_1,c_1]}\delta _{[s_2,c_2]}\underset{s}{}q^{\frac{a(s)}{2}}(1q)^{\frac{b(s)}{2}}\delta _{[s,c]}$$
where, for arbitrary $`s_1=(U_1,..,U_k)`$ and $`s_2=(V_1,..,V_p)`$, the sum runs over all $`s=(V_1,..,V_{p1},V,U,U_2,..,U_k)`$ with $`V_pV,UU_1`$ and $`U+V=U_1+V_p`$. The coloring $`c`$ restricts to $`c_1`$ and $`c_2`$ on the sets $`_\alpha U_\alpha `$ respectively $`_\beta V_\beta `$. The coefficients appearing on the right side are $`a(s)=|V||V_p|`$ and $`b(s)=|U|`$. As $`\delta _{[s,c]}=1`$ and $`a(s)+b(s)=|U_1|`$, we obtain
$`S(\delta _{[s_1,c_1]}\delta _{[s_2,c_2]})^2`$ $`=`$ $`{\displaystyle \underset{j=0}{\overset{|U_1|}{}}}\left({\displaystyle \genfrac{}{}{0pt}{}{|U_1|}{j}}\right)q^k(1q)^{|U_1|k}`$
$`=1`$ $`=`$ $`\delta _{[s_1,c_1]}\delta _{[s_2,c_2]}^2,`$
which proves the isometry property. The equation (32) follows by induction w.r.t. $`k`$. For $`k=0`$ is is obvious that
$$S\left(\mathrm{\Omega }\underset{p}{}a_{j_p}^\mathrm{}_p\mathrm{\Omega }\right)=\underset{p}{}a_{j_p}^\mathrm{}_p\mathrm{\Omega }.$$
Then one can check on the basis vectors that
$$S(a_j^\mathrm{}_j\mathrm{𝟏})=a_j^\mathrm{}_jS$$
for $`jJ_1`$, which provides the tool for the incrementation of $`k`$.
We pass now to the expression of $`t_{\mathrm{Bal}}(𝒱)`$ for a one block partition $`𝒱`$. The basic observation is that the creation and annihilation operators have the following form, stemming from that of $`\omega _{\mathrm{Bal}}`$ (see (30) ):
$$a_i^\mathrm{}_i=q^{\frac{1}{2}}a_{E,i}^\mathrm{}_i+(1q)^{\frac{1}{2}}a_{L,i}^\mathrm{}_i$$
with the choice $`a_{E,i}^{}\mathrm{\Omega }=0`$. Let $`M_𝒱=_{l=1}^{2n}a_{i_l}^\mathrm{}_l`$ be a monomial associated to the pair partition $`𝒱𝒫_2(2n)`$. It is sufficient to prove that the only nonzero contribution to $`M_𝒱\mathrm{\Omega }`$ is brought by the the term $`a_{L,i_1}_{l=2}^{2n1}a_{E,i_l}a_{L,i_{2n}}^{}\mathrm{\Omega }=q^{n1}\mathrm{\Omega }`$. Indeed the action of $`a_{L,i_1}^{}`$ at the combinatorial level is to increase the number of subsets in a sequence by 1. Thus the terms which are nonzero must contain an equal number of creation and annihilation operators of type $`L`$. Let us consider such a term. Then there exist $`1l_1l_22n`$ such that on the positions $`l_1`$ and $`l_2`$ we have annihilation respectively creation operators of type $`L`$ and for $`l_1ll_2`$ we have type $`E`$ operators. We have identified a submonomial
$$m=a_{L,l_1}\underset{l=l_1+1}{\overset{l_21}{}}a_{E,i_l}^\mathrm{}_la_{L,l_2}^{}$$
which is nonzero only if it corresponds to a pair partition, that is if all creation and annihilation operators pair each other according to the color. But this is possible only when $`l_1=1`$ and $`l_2=2n`$ because $`𝒱`$ is a one-block pair partition.
4) Free Products. Inspired by the notion of freeness introduced by
Dan Voiculescu we make the following:
Definition. Let $`(F_\alpha )_{\alpha J}`$ be a finite set of species of structures with
$`F_\alpha [\mathrm{}]=\{\mathrm{}\}`$ for all $`\alpha J`$. The free product of $`(F_\alpha )_{\alpha J}`$ is the species defined by:
$`_{\alpha J}(F_\alpha )[U]`$ $`=`$ $`\{(\pi ,(s_1,..,s_p))|\pi =(U_1,..,U_p)\mathrm{Bal}[U],s_iF_{\alpha _i}[U_i],`$ (33)
$`\alpha _i\alpha _{i+1}\text{for}i=1,..p1\}`$
for $`U\mathrm{}`$ and $`_{\alpha J}F_\alpha [\mathrm{}]=\{\mathrm{}\}`$. The transport is induced from the species $`(F_\alpha )_{\alpha J}`$ and Bal. From the definition it is clear that we have the following combinatorial equation:
$$_{\alpha J}(F_\alpha )=1+\underset{p1}{}\underset{\alpha _1\alpha _2..\alpha _p}{}F_{\alpha _1+}F_{\alpha _2+}\mathrm{}F_{\alpha _p+}$$
and using the property $`\mathrm{\Gamma }_{FG}=\mathrm{\Gamma }_F\mathrm{\Gamma }_G`$ we obtain
$$\mathrm{\Gamma }_{_{\alpha J}(F_\alpha )}(𝒦)=_{\alpha J}(\mathrm{\Gamma }_{F_\alpha },\mathrm{\Omega }_\alpha )$$
where the last object is the Hilbert space free product .
The corresponding weight is similar to the one used for the species Bal. For $`f_iF_{\alpha _i}[U_i]`$ and $`f_i^{}F_{\alpha _i^{}}[V_i]`$, it has the expression:
$`\omega _{_{\alpha J(F_\alpha )}}((\pi ,(f_1,..,f_p)),(\pi ^{},(f_1^{},..,f_q^{})))=`$
$`\delta _{p,q}\delta _{\alpha _p,\alpha _p^{}}{\displaystyle \underset{i=1}{\overset{p1}{}}}\delta _{f_i,f_i^{}}\omega _{\alpha _p}(f_p.f_p^{})+\delta _{p+1,q}{\displaystyle \underset{i=1}{\overset{p}{}}}\delta _{f_i,f_i^{}}\omega _{\alpha _q^{}}(\{\mathrm{}\},f_q^{}).`$
Moreover the creation and annihilation operators can be written like
$$a_{_{\alpha J(F_\alpha )},i}^{\mathrm{}}=\underset{\alpha }{}a_{F_\alpha ,i}^{\mathrm{}}$$
with the relations ,
$$a_{F_\alpha ,i}a_{F_\beta ,j}^{}=0$$
for $`\alpha \beta `$.
|
warning/0007/nlin0007008.html
|
ar5iv
|
text
|
# Fractal Dimensions of the Hydrodynamic Modes of Diffusion
## 1. Introduction
The statistical mechanics of nonequilibrium processes has been the subject of renewed attention. Computer studies have shown that the classical motions of systems with large numbers of particles are typically chaotic . These observations have motivated the theoretical study of low-dimensional model systems where one can obtain quantitative information about nonequilibrium processes in terms of the chaotic properties of the microscopic dynamics of the systems.
In this context, several relationships have been established between transport properties such as diffusion or viscosity and characteristic quantities of chaos such as Lyapunov exponents, Kolmogorov-Sinai (KS) entropy per unit time, and fractal dimensions . These relationships have been obtained for two different types of dynamical systems: (a) deterministic systems with Gaussian thermostats and, (b) open Hamiltonian systems with absorbing boundaries.
In the deterministic systems with Gaussian thermostats , the forces ruling the motion of the particles are modified in order to mimic a heat pump removing the excess energy introduced by the external fields or other nonequilibrium constraints. The modified dynamical system is still time-reversal symmetric but volumes are no longer preserved in phase space so that trajectories, in general, converge toward a fractal attractor which has an invariant Sinai-Ruelle-Bowen (SRB) measure, absolutely continuous with respect to Lebesgue measure in unstable directions and fractal in stable directions. On average, the phase-space volumes contract at a rate which increases with the external fields and, thus, with the dissipation of heat. Since the overall contraction rate is equal to the sum of all the Lyapunov exponents, the dissipation turns out to be related to the characteristic quantities of chaos. In this way, formulae have been obtained for the viscosity of an externally sheared fluid and also for the diffusion of a Gaussian-thermostated Lorentz gas . In this last case, the formula relating the diffusion coefficient to the sum of the Lyapunov exponents was also rigorously proved . In the thermostated Lorentz gas, the kinetic energy is constant (instead of the total energy) and the phase space of the flow has thus the dimension three. The Lyapunov exponents of this system are $`\lambda ^+>0>\lambda ^{}`$ with $`|\lambda ^{}|>\lambda ^+`$ for a nonvanishing field force $`F`$. In the three-dimensional phase space, the fractal attractor $`𝒜_F`$ has a KS entropy equal to the positive Lyapunov exponent $`h_{\mathrm{KS}}=\lambda ^+`$ according to Pesin’s equality, and its information dimension is given by Young’s formula: $`D_\mathrm{I}=2+h_{\mathrm{KS}}/|\lambda ^{}|`$. In this expression for the fractal dimension, the 2 stands for the integer dimensions of the flow direction and of the unstable direction which both are equal to one for the attractor of a flow. However, the attractor is fractal in the stable direction which contributes its partial information dimension $`d_\mathrm{I}=h_{\mathrm{KS}}/|\lambda ^{}|`$ to the total information dimension of the attractor $`𝒜_F`$. The diffusion coefficient of the Gaussian-thermostated Lorentz gas at temperature $`T`$ has been alternatively expressed as
$`𝒟`$ $`=`$ $`\underset{F0}{lim}\left({\displaystyle \frac{k_\mathrm{B}T}{F}}\right)^2\left[\lambda ^{}(𝒜_F)\lambda ^+(𝒜_F)\right]`$ (1)
$`=`$ $`\underset{F0}{lim}\left({\displaystyle \frac{k_\mathrm{B}T}{F}}\right)^2\left[|\lambda ^{}(𝒜_F)|h_{\mathrm{KS}}(𝒜_F)\right]`$
$`=`$ $`\underset{F0}{lim}\left({\displaystyle \frac{k_\mathrm{B}T}{F}}\right)^2\lambda ^+(𝒜_F)c_\mathrm{I}(𝒜_F),`$
where the first form is a consequence of the average contraction of phase-space volumes due to the Gaussian thermostatting mechanism and the relation between the rate of contraction and the rate of entropy production in systems with Gaussian thermostats, the second form follows from the first one by Pesin’s equality, and the third one is obtained by Young formula for the partial information codimension $`c_\mathrm{I}=1d_\mathrm{I}=1h_{\mathrm{KS}}/|\lambda ^{}|`$ and from the fact that $`lim_{F0}|\lambda ^{}(𝒜_F)|=lim_{F0}\lambda ^+(𝒜_F)`$ . In the limit of arbitrarily small external field $`F`$, the information dimension of the fractal attractor approaches the phase-space dimension in a way controlled by the diffusion coefficient, which is the basis of this kind of formulae.
Very similar formulae have been obtained for open Hamiltonian systems with absorbing boundaries . Such boundaries are naturally introduced in systems of scattering type in which particles or trajectories undergo transient collisions before escaping out of a phase-space region of varying geometry. The absorbing boundaries have the effect of driving the system out of equilibrium without a modification of the forces which can remain Hamiltonian, i.e., time-reversal symmetric and also volume preserving. The escape of trajectories leads to the formation of a fractal repeller, described by a invariant Gibbs measure which is fractal in both the unstable and the stable directions. For such systems, the rate of escape is given as the difference between the sum of positive Lyapunov exponents and the KS entropy , calculated with respect to the invariant Gibbs measure on the repeller. This escape-rate formula was rigorously proved for Anosov maps with rectangular holes . For the conservative Lorentz gas with absorbing boundaries separated by a distance $`L`$, the Lyapunov exponents are $`\lambda ^+>0>\lambda ^{}=\lambda ^+`$ and the diffusion coefficient has been obtained by the following formulae where the quantities are evaluated for the fractal repeller $`_L`$:
$`𝒟`$ $`=`$ $`\underset{L\mathrm{}}{lim}\left({\displaystyle \frac{L}{\pi }}\right)^2\left[\lambda ^+(_L)h_{\mathrm{KS}}(_L)\right]`$ (2)
$`=`$ $`\underset{L\mathrm{}}{lim}\left({\displaystyle \frac{L}{\pi }}\right)^2\lambda ^+(_L)c_\mathrm{I}(_L)`$
$`=`$ $`\underset{L\mathrm{}}{lim}\left({\displaystyle \frac{L}{\pi }}\right)^2\lambda ^+(_L)c_\mathrm{H}(_L),`$
where $`c_\mathrm{I}=1h_{\mathrm{KS}}/\lambda ^+`$ denotes the partial information codimension in the stable or unstable directions, which are equivalent. This information codimension can be replaced by the partial Hausdorff codimension $`c_\mathrm{H}`$ of the fractal repeller in the limit $`L\mathrm{}`$ . According to Eq. (2), the fractal repeller fills the phase space in the limit where the absorbing boundaries are sent to infinity. The way the dimension of the fractal repeller approaches the phase-space dimension is controlled by the diffusion coefficient, and leads to the relationship between the transport coefficient and the codimension. The formula (2) has been generalized to all the transport coefficients .
The similarities between formulae (1) and (2) have been discussed for several years . In particular, a generalization of these formulae in the presence of both an external field and absorbing boundaries was obtained . However, each of the two formulae depends very much on the type of invariant set, an attractor or a repeller, and on the invariant measure, SRB or Gibbs, in terms of which they are derived. Thus, since the corresponding measures and fractal structures are so different in the two cases, it has not yet been possible to provide satisfying reasons why the formulae are so similar.
The purpose of the present paper is to introduce a unifying approach which transcends the aforementioned differences and in which a third and new formula can be derived which expresses the diffusion coefficient in terms of the positive Lyapunov exponent of the system and the Hausdorff dimension of a fractal curve that describes a generalized hydrodynamic mode of the diffusion process. This Hausdorff dimension is not, therefore, the dimension of an attractor or a repeller. The novelty of the present unifying approach is that we consider an abstract fractal curve directly associated with the hydrodynamic modes of relaxation toward a stationary state such as the equilibrium state. These hydrodynamic modes exist whether the system is thermostated or not, on the condition that its spatial extension is large compared with characteristic microscopic lengths, in order to sustain a process of transport by diffusion. We suppose that the time evolution of the concentration of tracer particles can be decomposed by a spatial Fourier transform into modes characterized by a wavenumber $`k`$ or, equivalently, a wavelength $`L=2\pi /k`$. Each mode describes the inhomogeneities of concentration having this specified spatial periodicity and the time evolution of a non-periodic concentration is obtained by superposition of all the modes. Each mode has an exponential relaxation rate. For a deterministic dynamical system, these hydrodynamic modes can be defined in analogy with the conditionally invariant measures by compensating the exponential decay of the mode amplitude with an appropriate renormalization at each time step. In the present context, the conditionally invariant measures defining the hydrodynamic modes are complex because of the Fourier transform. A theory of these hydrodynamic modes has been described elsewhere , where these conditionally invariant measures are obtained as the eigendistributions of a Perron-Frobenius operator of the dynamical system. The main result we have to keep in mind here is that these complex measures are singular with respect to the Lebesgue measure so that they do not have density functions. Instead, they have cumulative functions defined by the measure of a variable set in phase space. These cumulative functions are complex, continuous and nondifferentiable. When plotted in the complex plane, they depict fractal curves of von Koch’s type with Hausdorff dimensions between one and two. A recent work by the present authors has showed that the singular character of the hydrodynamic modes accounts for the entropy production of irreversible thermodynamics.
In the present paper, we shall show that the Hausdorff dimension of the fractal curves associated with the hydrodynamic modes is controlled by the diffusion coefficient, which leads to a new relationship between transport and chaos. We shall here prove our result for a whole class of multi-baker maps. This may appear restrictive but several important results were first proved for the multi-baker before being extended to more general chaotic systems. In particular, let us mention that the multi-baker models have been used for the study of the nonequilibrium steady states and the entropy production .
The multi-baker models have the advantage of being exactly solvable so that they are therefore appropriate for the intensive studies of different nonequilibrium properties. Many of their properties are known in detail .
The organization of the paper is as follows. In Sec. 2, we discuss general models of deterministic random walks, which include both conservative and so-called dissipative cases. In Sec. 3, the time evolution operator on the statistical ensembles of trajectories is introduced. The eigendistributions and their cumulative functions are derived. Two different types of stationary solutions to the time evolution operator are discussed in Sec. 4, one for conservative systems, the other for dissipative ones. The computation of the Hausdorff dimensions of the cumulative functions of the eigendistributions is done in Sec. 5. Conclusions are drawn in Sec. 6.
## 2. Deterministic Models of Diffusion
The simplest examples of reversible dynamical systems with diffusive properties are based upon simple stochastic processes. The most commonly studied is the discrete symmetric dyadic random walk where independent particles are allowed to move on the sites of a one-dimensional discrete lattice and the time evolution is determined by the condition that the probability is the same for a particle to move to the nearest neighboring site, on the right or left .
In order to study this process as one with a phase-space dynamics, we consider the set of all the possible infinite (both in the past and future) trajectories of such particles. It consists of sequences $`\{n_i\}_{i=\mathrm{}}^+\mathrm{}`$ of integers $`n_i`$ with the restriction that the difference between the integers at successive time steps is always equal to plus or minus one, i.e. $`i,`$
$$n_in_{i1}=\pm 1.$$
The stochastic time evolution is thus replaced by the shift operator on those sequences,
$$\mathrm{\Sigma }\left(\{n_i\}\right)=\{n_i^{}\},$$
with $`n_i^{}=n_{i+1}.`$
Alternatively, a convenient representation of phase space is to consider binary sequences of zeros and ones labeling respectively hops to the left and right. A trajectory of the random walker is then given by an integer coordinate representing, say, the position at time zero and an arbitrary binary sequence $`\{\omega _i\}_{i=\mathrm{}}^+\mathrm{},`$ $`\omega _i\{0,1\}i,`$ coding all its (past and future) displacements.
This phase space is a two-dimensional continuum, which is best represented by points on the unit square. The transposition from symbolic sequences to points of the unit square is given by the dyadic expansion of $`x`$ and $`y.`$ Let
$`x`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _i}{2^{i+1}}},`$ (3)
$`y`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _i}{2^i}},`$ (4)
i.e. the point $`(x,y)`$ on the unit square is coded by the bi-infinite sequence $`\{\omega _i\}_{i=\mathrm{}}^+\mathrm{},`$ with the positive indices (including zero) coding the $`x`$-component and the negative ones the $`y`$-component.
The reason for introducing this representation of trajectories is that the time evolution represented by the shift operator on binary sequences is isomorphic to the baker-map acting on points $`(x,y)`$ of the unit square,
$$(x,y)\{\begin{array}{cc}(2x,y/2),\hfill & 0x<1/2,\hfill \\ (2x1,(y+1)/2),\hfill & 1/2x<1.\hfill \end{array}$$
(5)
The proof that the shift on binary sequences and the baker map are isomorphic is straightforward : the baker map, Eq. (5), takes $`(x,y)`$ to $`(x^{},y^{})`$ with the dyadic expansion
$`x^{}`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _{i+1}}{2^{i+1}}},`$ (6)
$`y^{}`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _{i+1}}{2^i}}.`$ (7)
In other words, the couple $`(x^{},y^{})`$ is coded by the shifted sequence $`\{\omega _i^{}\}_{i=\mathrm{}}^+\mathrm{},`$ with $`\omega _i^{}=\omega _{i+1}.`$
In order to describe the one-dimensional dyadic symmetric random walk, we need to add the aforementioned integer $`n`$ which labels the site of the one-dimensional lattice where the particle is located at the current time. Accordingly, this random walk is isomorphic to the symmetric dyadic multi-baker map
$$B_2:(n,x,y)\{\begin{array}{cc}(n1,2x,y/2),\hfill & 0x<1/2,\hfill \\ (n+1,2x1,(y+1)/2),\hfill & 1/2x<1,\hfill \end{array}$$
(8)
acting in the phase space $`\times [0,1]^2.`$
Likewise, a one-dimensional non-symmetric $`r`$-adic random walk, where a particle located at site $`n`$ has probability $`p_j`$ to hop to the site $`n+j`$, with the integer $`j`$ taking values between $`(r1)/2`$ and $`(r1)/2`$ (assuming $`r`$ to be odd<sup>1</sup><sup>1</sup>1The case $`r`$ even is treated similarly, with the modification that the probability $`p_0`$ for a particle to remain at the same position is set to zero.) and $`p_{(r1)/2}+\mathrm{}+p_{(r1)/2}=1`$, can also be mapped on a deterministic multi-baker map.
Similarly to the dyadic process, we code the trajectories of such a process by sequences $`\{\omega _i\}_{i=\mathrm{}}^+\mathrm{},`$ with $`\omega _i\{0,\mathrm{},r1\}.`$ Since every symbol $`\omega _i`$ has respective weight $`p_{\omega _i(r1)/2},`$ the $`r`$-adic expansion of $`x`$ and $`y`$ takes a form slightly more complicated than Eqs. (3)-(4). We can write those expansions under the recursive form
$`x(\omega _0,\omega _1,\mathrm{})`$ $`=`$ $`{\displaystyle \underset{j<\omega _0}{}}p_{[j(r1)/2]}+p_{[\omega _0(r1)/2]}x(\omega _1,\omega _2,\mathrm{}),`$ (9)
$`y(\omega _1,\omega _2,\mathrm{})`$ $`=`$ $`{\displaystyle \underset{j<\omega _1}{}}p_{[(r1)/2j]}+p_{[(r1)/2\omega _1]}y(\omega _2,\omega _3,\mathrm{}).`$ (10)
The $`r`$-adic multi-baker map that mimics the random process is given by
$$B_r:(n,x,y)\{\begin{array}{cc}\hfill (n\frac{r1}{2},& \frac{x}{p_{(r1)/2}},p_{(r1)/2}y),\hfill \\ & 0x<p_{(r1)/2},\hfill \\ \hfill (n\frac{r3}{2},& \frac{xp_{(r1)/2}}{p_{(r3)/2}},p_{(r3)/2}y+p_{(r1)/2}),\hfill \\ & p_{(r1)/2}x<p_{(r1)/2}+p_{(r3)/2},\hfill \\ \hfill \mathrm{}& \\ \hfill (n+\frac{r1}{2},& \frac{x_{j=(r1)/2}^{(r3)/2}p_j}{p_{(r1)/2}},p_{(r1)/2}y+\underset{j=(r3)/2}{\overset{(r1)/2}{}}p_j),\hfill \\ & \underset{j=(r1)/2}{\overset{(r3)/2}{}}p_jx<1.\hfill \end{array}$$
(11)
The action of $`B_r`$ on points $`x`$ and $`y`$ with $`r`$-adic expansions given by Eqs. (9)-(10) is easily found to yield
$$\begin{array}{c}B_r[n,x(\omega _0,\omega _1,\mathrm{}),y(\omega _1,\omega _2,\mathrm{})]=\hfill \\ \hfill [n+\omega _0(r1)/2,x(\omega _1,\omega _2,\mathrm{}),y(\omega _0,\omega _1,\mathrm{})].\end{array}$$
(12)
Notice the reversed order of the $`p_i`$’s along the $`y`$-coordinate in Eq. (11). This ensures the time reversibility of the map : The involution $`T(x,y)=(1y,1x)`$ is a time-reversal operator or reversal symmetry for $`B_r`$ in the sense that
$$TB_rT=B_r^1.$$
(13)
The multi-baker map, Eq. (11), is chaotic with the mean positive Lyapunov exponent
$$\lambda ^+=\underset{j=(r1)/2}{\overset{(r1)/2}{}}p_j\mathrm{ln}\frac{1}{p_j}=h_{\mathrm{KS}}>\mathrm{\hspace{0.33em}0},$$
(14)
which is equal to its Kolmogorov-Sinai entropy, and the mean negative Lyapunov exponent
$$\lambda ^{}=\underset{j=(r1)/2}{\overset{(r1)/2}{}}p_j\mathrm{ln}p_j<\mathrm{\hspace{0.33em}0},$$
(15)
both evaluated under the forward dynamics.
The $`r`$-adic multi-baker maps (11) constitute simplified models of the Poincaré-Birkhoff mappings ruling the collision dynamics of a point-like particle elastically bouncing on hard disks fixed in the plane . In particular, the multi-baker models share many of the chaotic properties of these Lorentz-type billiards.
###### Example 1.
The dyadic biased random walk where a particle moves to the left with probability $`p_1q`$ and to the right with probability $`p_1=1q`$ is coded by binary sequences which are mapped onto the unit square by the expansion
$`x`$ $`=`$ $`\omega _0q+{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}\omega _iq{\displaystyle \underset{l=0}{\overset{i1}{}}}p_{(2\omega _l1)},`$ (16)
$`y`$ $`=`$ $`\omega _1(1q)+{\displaystyle \underset{i=2}{\overset{\mathrm{}}{}}}\omega _i(1q){\displaystyle \underset{l=1}{\overset{i+1}{}}}p_{(12\omega _l)}.`$ (17)
This transposition yields the so-called reversible dissipative multi-baker map
$$B_2^\mathrm{d}:(n,x,y)\{\begin{array}{cc}(n1,x/q,(1q)y),\hfill & 0x<q,\hfill \\ (n+1,(xq)/(1q),1q+qy),\hfill & qx<1.\hfill \end{array}$$
(18)
###### Example 2.
The triadic totally symmetric random walk, with $`p_1=p_0=p_11/3,`$ was discussed by the authors in . Similarly to Eqs. (3)-(4), infinite sequences of zeros, ones and twos are mapped onto the unit square by the expansion
$`x`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _i}{3^{i+1}}},`$ (19)
$`y`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\omega _i}{3^i}}.`$ (20)
The triadic multi-baker map is simply
$$B_3:(n,x,y)\{\begin{array}{cc}(n1,3x,y/3),\hfill & 0x<1/3,\hfill \\ (n,3x1,(y+1)/3),\hfill & 1/3x<2/3,\hfill \\ (n+1,3x2,(y+2)/3),\hfill & 2/3x<1.\hfill \end{array}$$
(21)
###### Example 3.
The biased triadic random walk, where a particle hops to the left with probability $`p_1s_L,`$ to the right with probability $`p_1s_R,`$ and remains at its position with probability $`p_0=s_N1s_Ls_R,`$ is one of the processes studied by Tél, Vollmer, and Breymann . The transposition of the infinite symbolic sequences to the points of the unit square is similar to Eqs. (16)-(17) and yields the reversible dissipative triadic multi-baker map
$$B_3^\mathrm{d}:(n,x,y)\{\begin{array}{cc}(n1,x/s_L,s_Ry),\hfill & 0x<s_L,\hfill \\ (n,(xs_L)/s_N,s_Ny+s_R),\hfill & s_Lx<s_L+s_N,\hfill \\ (n+1,(xs_Ls_N)/s_R,s_Ly+s_N+s_R),\hfill & s_L+s_Nx<1.\hfill \end{array}$$
(22)
The equivalence between the non-area-preserving map (22) and an area-preserving multi-baker map with an extra variable of energy was discussed in Ref. .
## 3. Time Evolution of Statistical Ensembles
In order to study the relaxation to stationarity of a large ensemble of random walkers, we restrict the one-dimensional lattice to a ring of $`L`$ sites and impose periodic boundary conditions, that is we identify site $`0`$ with site $`L.`$ The phase space is thus restricted to $`\times [0,1]^2`$ with $`=\{1,\mathrm{},L\}`$.
The time evolution of a density $`\rho _t(n,x,y)`$ is determined by the Perron-Frobenius operator : $`\rho _{t+1}(n,x,y)=𝒫\rho _t(n,x,y)=\rho _t\left[B_r^1(n,x,y)\right]/J\left[B_r^1(n,x,y)\right]`$, where we have introduced the Jacobian of the transformation, $`J\left[B_r^1(n,x,y)\right]=p_i/p_i`$ if $`_{j>i}p_jy<_{ji}p_j.`$ The average number of particles in a cylinder $`\mathrm{\Omega }`$ of the $`n^{\mathrm{th}}`$ cell is given by $`\mu _t(\mathrm{\Omega })=_\mathrm{\Omega }𝑑x𝑑y\rho _t(n,x,y)`$, with time evolution $`\mu _{t+1}(\mathrm{\Omega })=\mu _t\left[B_r^1(\mathrm{\Omega })\right]`$.
To display the fractal forms underlying the relaxation to stationarity, we consider the cumulative function defined as the measure of a cylinder $`\mathrm{\Omega }=[0,1]\times [0,y]`$ inside the $`n^{\mathrm{th}}`$ site, i. e.
$$g_t(n,y)=_0^1𝑑x_0^y𝑑y^{}\rho _t(n,x,y^{}).$$
(23)
We can assume, without loss of generality, that the distribution function $`\rho _t(n,x,y)`$ is uniform with respect to $`x`$, since the $`x`$-direction is uniformly expanding. Thus the distribution will become uniform along the $`x`$ direction, for long times, independently of the initial conditions. Thus we can write $`\mu _t(n,[0,x]\times [0,y])=xg_t(n,y)`$ and find that the time evolution of the cumulative function, $`g_t(n,y)`$, is given by
$$g_{t+1}(n,y)=\{\begin{array}{cc}\hfill p_{(r1)/2}g_t(n+\frac{r1}{2},& \frac{y}{p_{(r1)/2}}),\hfill \\ & 0y<p_{(r1)/2},\hfill \\ \hfill p_{(r1)/2}g_t(n+\frac{r1}{2},1)& \\ \hfill +p_{(r3)/2}g_t(n+\frac{r3}{2},& \frac{yp_{(r1)/2}}{p_{(r3)/2}}),\hfill \\ & p_{(r1)/2}y<p_{(r1)/2}+p_{(r3)/2},\hfill \\ \hfill \mathrm{}& \\ \hfill \underset{j=(r1)/2}{\overset{(r3)/2}{}}p_jg_t(nj,1)& \\ \hfill +p_{(r1)/2}g_t(n\frac{r1}{2},& \frac{y_{j=(r3)/2}^{(r1)/2}p_j}{p_{(r1)/2}}),\hfill \\ & \underset{j=(r3)/2}{\overset{(r1)/2}{}}p_jy<1.\hfill \end{array}$$
(24)
In order to identify the hydrodynamic modes and their eigenfunctions, we consider $`y=1`$ in Eq. (24). This yields the matrix equation
$$g_{t+1}(n,1)=\underset{j=(r1)/2}{\overset{(r1)/2}{}}p_jg_t(nj,1).$$
(25)
The eigenmodes of this equation have the form $`\psi _k(n)=\mathrm{exp}(ikn)`$ with eigenvalue
$$\chi _k=\underset{j=(r1)/2}{\overset{(r1)/2}{}}p_j\mathrm{exp}(ikj).$$
(26)
Moreover, the values of the wavenumber $`k`$ are restricted to $`k=2\pi m/L,`$ with $`m(modL),`$ by the periodic boundary conditions.
According to the above considerations, we may suppose that the cumulative function $`g_t(n,y)`$ of the deterministic multi-baker map may be expanded as :
$$g_t(n,y)=\underset{k}{}\chi _k^ta_k\psi _k(n)F_k(y),$$
(27)
where $`a_k`$ are coefficients set by the initial conditions and $`F_k(y)`$ is solution of the system
$$F_k(y)=\{\begin{array}{cc}\frac{p_{\left(r1\right)/2}}{\chi _k}\mathrm{exp}\left(ik\frac{r1}{2}\right)F_k\left(\frac{y}{p_{(r1)/2}}\right),\hfill & \\ 0y<p_{(r1)/2},\hfill & \\ \frac{p_{(r1)/2}}{\chi _k}\mathrm{exp}\left(ik\frac{r1}{2}\right)\hfill & \\ +\frac{p_{(r3)/2}}{\chi _k}\mathrm{exp}\left(ik\frac{r3}{2}\right)F_k\left(\frac{yp_{(r1)/2}}{p_{(r3)/2}}\right),\hfill & \\ p_{(r1)/2}y<p_{(r1)/2}+p_{(r3)/2},\hfill & \\ \mathrm{}\hfill & \\ & \\ _{j=(r1)/2}^{(r3)/2}\frac{p_j}{\chi _k}\mathrm{exp}(ikj)\hfill & \\ +\frac{p_{(r1)/2}}{\chi _k}\mathrm{exp}\left(ik\frac{r1}{2}\right)F_k\left(\frac{y_{j=(r3)/2}^{(r1)/2}p_j}{p_{(r1)/2}}\right),\hfill & \\ _{j=(r3)/2}^{(r1)/2}p_jy<1,\hfill & \end{array}$$
(28)
where we used the identity $`F_k(1)=1`$.
###### Remark 1.
As long as
$$p_j/\chi _k<1,j,$$
(29)
Eq. (28) is a contracting functional equation of de Rham type with a unique continuous solution corresponding to the cumulative function of the hydrodynamic eigendistribution of wavenumber $`k`$ of the Perron-Frobenius operator.
On large spatial scales, the transport process in the multi-baker chain can be modeled by the advection-diffusion equation
$$_tc=v_nc+𝒟_n^2c,$$
(30)
where $`c=c(n,t)=g_t(n,y=1)`$ is the particle concentration given by the cumulative function at $`y=1`$, $`v`$ is the average velocity of the particles drifting in a biased random walk, and $`𝒟`$ is the diffusion coefficient. The result (30) is a consequence of the decomposition (27) of the cumulative function in terms of the hydrodynamic modes $`\psi _k(n)=\mathrm{exp}(ikn)`$. Indeed, if we substitute the decomposition (27) into Eq. (30), we find that the hydrodynamic modes $`c(n,t)=\chi _k^t\psi _k(n)`$ are approximate solutions under the approximation that the terms in $`k^3`$, $`k^4`$,… are neglected, in which case we find that
$$\mathrm{ln}\chi _k=ivk𝒟k^2+O(k^3).$$
(31)
This approximation is justified for the advection-diffusion equation (30) because the large-scale limit is equivalent to the small wavenumber limit $`k0`$. In this way, an expansion of the eigenvalue (26) in powers of the wavenumber $`k`$ allows us to identify the drift velocity:
$$v\underset{j=(r1)/2}{\overset{(r1)/2}{}}jp_j,$$
(32)
and the diffusion coefficient:
$$𝒟\frac{1}{2}\underset{j=(r1)/2}{\overset{(r1)/2}{}}j^2p_j\frac{1}{2}\left(\underset{j=(r1)/2}{\overset{(r1)/2}{}}jp_j\right)^2.$$
(33)
in terms of the parameters of the mapping (11).
###### Remark 2.
For a symmetric random walk, with $`r`$ odd, and where $`p_j=p_jj,`$ the eigenvalue is real and Eq. (26) becomes :
$$\chi _k=p_0+2\underset{j=1}{\overset{(r1)/2}{}}p_j\mathrm{cos}(kj).$$
(34)
In this symmetric case, the drift velocity vanishes, $`v=0`$, and the random walk is purely diffusive with the diffusion coefficient
$$𝒟=\underset{j=1}{\overset{(r1)/2}{}}j^2p_j.$$
(35)
For totally symmetric $`r`$-adic random walks for which $`p_j=1/r`$ $`j`$, the eigenvalue is given by
$$\chi _k=\frac{\mathrm{sin}(kr/2)}{r\mathrm{sin}(k/2)},$$
(36)
and the diffusion coefficient is
$$𝒟=\frac{r^21}{24}.$$
(37)
For small $`k`$, we can expand $`F_k`$ in powers of $`k`$ and obtain the equivalent of a gradient expansion :
$$F_k(y)=F_0(y)+ikT(y)+O\left(k^2\right).$$
(38)
The first term of this equation is a real function $`F_0`$ and corresponds to the stationary state of eigenvalue 1. In the next section, we will differentiate between regular and singular stationary solutions. The second term in Eq. (38) is purely imaginary, i.e. $`T`$ is a real function. Note that
$$\frac{1}{\chi _k}=1+ikv+O\left(k^2\right),$$
(39)
where $`v`$ is the drift velocity (32) of the random walker due to a possible bias. Substituting Eq. (38) into Eq. (28) and, using Eq. (39), we find the general expression satisfied by $`T(y)`$ :
$$T(y)=\{\begin{array}{cc}p_{(r1)/2}\left[\left(\frac{r1}{2}+v\right)F_0\left(\frac{y}{p_{(r1)/2}}\right)+T\left(\frac{y}{p_{(r1)/2}}\right)\right],\hfill & \\ & 0y<p_{(r1)/2},\hfill \\ p_{(r1)/2}\left(\frac{r1}{2}+v\right)\hfill & \\ +p_{(r3)/2}[(\frac{r3}{2}+v)F_0\left(\frac{yp_{(r1)/2}}{p_{(r3)/2}}\right)\hfill & \\ +T\left(\frac{yp_{(r1)/2}}{p_{(r3)/2}}\right)],\hfill & \\ & p_{(r1)/2}y<p_{(r1)/2}+p_{(r3)/2},\hfill \\ \mathrm{}\hfill & \\ & \\ _{j=(r1)/2}^{(r3)/2}p_j\left(j+v\right)\hfill & \\ +p_{(r1)/2}[(\frac{r1}{2}+v)F_0\left(\frac{y_{j=(r3)/2}^{(r1)/2}p_j}{p_{(r1)/2}}\right)\hfill & \\ +T\left(\frac{y_{j=(r3)/2}^{(r1)/2}p_j}{p_{(r1)/2}}\right)],\hfill & \\ & _{j=(r3)/2}^{(r1)/2}p_jy<1.\hfill \end{array}$$
(40)
In Sec. 4, we will make the connection between this equation and the functional equation defining a function introduced by Takagi .
## 4. Stationary State
The stationary eigenstate of the Perron-Frobenius operator has eigenvalue 1, with a cumulative function $`F_0`$ given by Eq. (28) for a vanishing wavenumber $`k=0`$ :
$$F_0(y)=\{\begin{array}{cc}p_{(r1)/2}F_0\left(\frac{y}{p_{\left(r1\right)/2}}\right),\hfill & 0y<p_{(r1)/2},\hfill \\ p_{(r1)/2}+p_{(r3)/2}F_0\left(\frac{yp_{\left(r1\right)/2}}{p_{\left(r3\right)/2}}\right),\hfill & \\ & p_{(r1)/2}y<p_{(r1)/2}+p_{(r3)/2},\hfill \\ \text{etc.}\hfill & \end{array}$$
(41)
For a symmetric process, where $`p_j=p_jj,`$ the solution to this equation is simply $`F_0(y)=y,`$ corresponding to the uniform density of the equilibrium stationary state.
###### Example 4.
The dyadic symmetric random walk has the stationary state given by $`F_0(y)=y`$. Equation (40) thus becomes
$$T(y)=\{\begin{array}{cc}y+\frac{1}{2}T(2y),\hfill & 0y<\frac{1}{2},\hfill \\ 1y+\frac{1}{2}T(2y1),\hfill & \frac{1}{2}y<1.\hfill \end{array}$$
(42)
The solution to this functional equation is the Takagi function , which justifies the name generalized Takagi function for $`T`$ in Eq. (40). The Takagi function is known to be a nowhere differentiable function with Hausdorff dimension $`D_\mathrm{H}(T)=1`$. As will be seen later, this property is shared by the whole class of functions $`T`$ given by Eq. (40).
For a non-symmetric process, the contraction and expansion factors differ so that, in this case, $`F_0`$ is a Lebesgue singular function .
###### Example 5.
The cumulative function of the stationary state of the reversible dissipative multi-baker map, Eq. (18), of Example 1 satisfies
$$F_0(y)=\{\begin{array}{cc}qF_0\left(\frac{y}{1q}\right),\hfill & 0y<1q,\hfill \\ q+(1q)F_0\left(\frac{y1+q}{q}\right),\hfill & 1qy<1.\hfill \end{array}$$
(43)
We show in Fig. 1 this function evaluated for $`q=0.3,0.35,\mathrm{},0.7.`$ For this map, the generalized Takagi function, Eq. (40), is found to be a solution of
$$T(y)=\{\begin{array}{cc}2(1q)F_0(y)+qT\left(\frac{y}{1q}\right),\hfill & 0y<1q,\hfill \\ 2q[1F_0(y)]+(1q)T\left(\frac{y1+q}{q}\right),\hfill & 1qy<1.\hfill \end{array}$$
(44)
This function is depicted in Fig. 2 for $`q=0.5`$ to $`0.7.`$ For values of $`q`$ lower than $`0.5,`$ we note that $`T`$ has the symmetry $`T[q](y)=T[1q](1y).`$
## 5. Hausdorff Dimension of the Hydrodynamic modes
In this section, we investigate the fractal properties of the cumulative functions $`F_k`$ of the hydrodynamic modes for $`k0`$, as given by Eq. (28), and we compute their Hausdorff dimensions.
In order to compute the solutions of Eq. (28), we return to the symbolic dynamics and replace $`y`$ by the $`r`$-adic expansion, Eq. (10). We consider points $`y`$ with finite $`r`$-adic expansion $`y=y(\omega _1,\mathrm{},\omega _l)`$ and rewrite Eq. (28) in terms of the symbolic sequences :
$$\begin{array}{c}F_k[y(\omega _1,\mathrm{},\omega _l)]=\underset{j<\omega _1}{}\frac{p_{[j(r1)/2]}}{\chi _k}\mathrm{exp}\{ik[j(r1)/2]\}\hfill \\ \hfill +\frac{p_{[\omega _1(r1)/2]}}{\chi _k}\mathrm{exp}\{ik[\omega _1(r1)/2]\}F_k[y(\omega _2,\mathrm{},\omega _l)].\end{array}$$
(45)
Starting with $`F_k[\mathrm{}]=0`$, we can successively compute
$$F_k[y(\omega _1)],F_k[y(\omega _1,\omega _2)],F_k[y(\omega _1,\omega _2,\omega _3)],\text{etc}.$$
All these points are plotted in the complex plane where they form a fractal curve $`(\mathrm{Re}F_k,\mathrm{Im}F_k)`$ similar to von Koch’s curve, under certain conditions (see the following examples).
###### Example 6.
Consider the triadic totally symmetric random walk of Example 2. From Eq. (34), the eigenvalues are found to be $`\chi _k=(1/3)(1+2\mathrm{cos}k)`$. The condition, Eq. (29), that Eq. (28) be contracting becomes $`(1+2\mathrm{cos}k)>1`$, which is satisfied as long as $`\pi /2<k<\pi /2`$. In Fig. 3, we plot the imaginary versus real parts of $`F_k`$ for $`k=0.1`$ and $`0.5`$. The first of these two graphs for the smallest wavenumber $`k=0.1`$, Fig.3a, is dominated by the linear contributions to $`F_k`$, i.e. $`F_k(y)y+ikT(y)`$. As we argued in , the first order term in $`k`$, $`T`$, is the triadic equivalent of the Takagi function Eq. (42). In this case, $`T(y)`$ is solution of the functional equation
$$T(y)=\{\begin{array}{cc}y+\frac{1}{3}T(3y),\hfill & 0y<\frac{1}{3},\hfill \\ \frac{1}{3}+\frac{1}{3}T(3y1),\hfill & \frac{1}{3}y<\frac{2}{3},\hfill \\ 1y+\frac{1}{3}T(3y2),\hfill & \frac{2}{3}y<1.\hfill \end{array}$$
(46)
The deviations – observed in Fig.3b – of the curve $`(\mathrm{Re}F_k,\mathrm{Im}F_k)`$ with respect to the generalized Takagi function (46) are due to contributions which are second order in the wavenumber $`k`$. As we will show, these deviations are responsible for the fact that the Hausdorff dimension of this curve is not one for non-zero wavenumbers $`k`$. In fact, $`D_\mathrm{H}(F_k)>1`$.
###### Example 7.
Still considering the triadic totally symmetric random walk, we observe a rather spectacular fractal for the large wavenumber $`k=\pi /3`$ which corresponds to the eigenvalue $`\chi _{\pi /3}=2/3`$. This case is shown in Fig. 4. The points 0, 1/3, 2/3 and 1 are at the vertices of a half hexagon, which is responsible for the apparent hexagonal symmetry of this figure.
###### Example 8.
Another interesting case corresponds to the critical value $`k_c`$ of $`k`$ for which Eq. (28) looses its contracting property, i.e. $`\chi _{k_c}=p_j`$ in one of the conditions (29). This leads to interesting solutions of Eq. (28). For instance, in the triadic totally symmetric random walk, $`k_c=\pi /2`$ yields
$$F_{\pi /2}(y)=\{\begin{array}{cc}iF_{\pi /2}(3y),\hfill & 0y<1/3,\hfill \\ i+F_{\pi /2}(3y1),\hfill & 1/3y<2/3,\hfill \\ 1i+iF_{\pi /2}(3y2),\hfill & 2/3y<1.\hfill \end{array}$$
(47)
$`F_{\pi /2}(y)`$ is thus a non-continuous function of $`y`$. In the complex plane, the curve disintegrates into a disjoint set of points forming the square lattice $`^2`$. Figure 5 displays this set constructed with the $`3^{10}`$ points $`y(\omega _1,\mathrm{},\omega _{10})`$.
Similar functions can be observed for the 2-adic and 5-adic totally symmetric random walks for $`k_c=\pi /3`$. In the 2-adic case, the set is a regular triangular lattice; in the 5-adic case, a regular hexagonal lattice.
We turn to the computation of the Hausdorff dimension of $`F_k`$, i.e., of the curve $`(\mathrm{Re}F_k,\mathrm{Im}F_k)`$ drawn in the complex plane. The above examples show that the plot of $`(\mathrm{Re}F_k,\mathrm{Im}F_k)`$ forms a continuous curve in the complex plane as long as $`|k|<k_c`$. However, the curve densely fills plain domains of the complex plane already for lower values of the wavenumber, $`k_f<|k|<k_c`$. For these values, the Hausdorff dimension of the curve saturates at the dimension two of the complex plane so that $`D_\mathrm{H}(k)=2`$ for $`k_f|k|<k_c`$. At the limiting value $`|k|=k_f`$, the plot of $`(\mathrm{Re}F_k,\mathrm{Im}F_k)`$ forms a fractal curve similar to the Lévy dragon which has a Hausdorff dimension equal to two .
For still lower values of the wavenumber $`|k|<k_f`$, we consider the Hausdorff measure, $`\mathrm{\Gamma }_l^d(F_k),`$ of the cylinder sets specified by the sequences $`\{\omega _1,\mathrm{},\omega _l\}`$ :
$$\mathrm{\Gamma }_l^d(F_k)=\underset{\omega _1,\mathrm{},\omega _l}{}\left|\mathrm{\Delta }F_k[y(\omega _1,\mathrm{},\omega _l)]\right|^d,$$
(48)
where
$$\mathrm{\Delta }F_k[y(\omega _1,\mathrm{},\omega _l)]=F_k[y(\omega _1,\mathrm{},\omega _l+1)]F_k[y(\omega _1,\mathrm{},\omega _l)],$$
with the notation
$$\{\omega _1,\mathrm{},\omega _l+1\}=\{\begin{array}{cc}\{\omega _1,\mathrm{},\omega _{l1},\omega _l+1\},\hfill & \omega _l<r1,\hfill \\ \{\omega _1,\mathrm{},\omega _{l1}+1,0\},\hfill & \omega _l=r1,\hfill \end{array}$$
(49)
and we make the further convention that $`y(r1,\mathrm{},r1,r)=1`$. The quantities $`\mathrm{\Delta }F_k[y(\omega _1,\mathrm{},\omega _l)]`$ are the segments of the curve, $`(\mathrm{Re}F_k,\mathrm{Im}F_k)`$ at the $`l^{\mathrm{th}}`$ step of its construction. In the complex plane, the $`l^{\mathrm{th}}`$ approximant of the curve is covered by disks of diameter given by the absolute value of the segments $`\mathrm{\Delta }F_k[y(\omega _1,\mathrm{},\omega _l)]`$, appearing on the right hand side of Eq. (48). We recall that the Hausdorff dimension, $`D_\mathrm{H}(k),`$ of $`F_k`$ is the value of $`d`$ such that
$$\underset{l\mathrm{}}{lim}\mathrm{\Gamma }_l^d(F_k)=\{\begin{array}{cc}\mathrm{},\hfill & d<D_\mathrm{H}(k),\hfill \\ 0,\hfill & d>D_\mathrm{H}(k).\hfill \end{array}$$
(50)
With the help of Eq. (45), it is straightforward to check that
$`\left|\mathrm{\Delta }F_k[y(\omega _1,\mathrm{},\omega _l)]\right|`$ $`=`$ $`{\displaystyle \frac{p_{[\omega _1(r1)/2]}}{|\chi _k|}}\left|\mathrm{\Delta }F_k[y(\omega _2,\mathrm{},\omega _l)]\right|,`$ (51)
$`=`$ $`{\displaystyle \frac{1}{|\chi _k|^l}}{\displaystyle \underset{i=1}{\overset{l}{}}}p_{[\omega _i(r1)/2]}.`$
Hence, the Hausdorff measure
$$\mathrm{\Gamma }_l^d(F_k)=\frac{1}{|\chi _k|^{ld}}\underset{\omega _1,\mathrm{},\omega _l}{}\underset{i=1}{\overset{l}{}}p_{[\omega _i(r1)/2]}^{}{}_{}{}^{d}=\left(\frac{1}{|\chi _k|^d}\underset{j=(r1)/2}{\overset{(r1)/2}{}}p_{j}^{}{}_{}{}^{d}\right)^l.$$
(52)
According to Eq. (50), the Hausdorff dimension thus satisfies
$$|\chi _k|^{D_\mathrm{H}(k)}=p_{(r1)/2}^{D_\mathrm{H}(k)}+\mathrm{}+p_{(r1)/2}^{D_\mathrm{H}(k)}.$$
(53)
###### Example 9.
For the triadic totally symmetric random walk, the Hausdorff dimension of $`F_{\pi /3},`$ shown in Fig. 4, satisfies
$$\left(\frac{2}{3}\right)^{D_\mathrm{H}(\pi /3)}=3\left(\frac{1}{3}\right)^{D_\mathrm{H}(\pi /3)}.$$
Therefore $`D_\mathrm{H}(\pi /3)=\mathrm{ln}(3)/\mathrm{ln}(2)`$.
###### Remark 3.
It is not possible to solve Eq. (53) and find a general expression for $`D_\mathrm{H}(k)`$ for all processes. However for the case of totally symmetric $`r`$-adic random walks for which $`p_j=1/r,j`$, we can solve Eq. (53) and find
$$D_\mathrm{H}(k)=\{\begin{array}{cc}\frac{\mathrm{ln}\left(r\right)}{\mathrm{ln}\left(r\chi _k\right)},\hfill & |k|<k_f,\hfill \\ 2,\hfill & k_f|k|<k_c.\hfill \end{array}$$
(54)
$`k_f`$ is the wavenumber such that $`\chi _k=1/\sqrt{r}`$ where the curve starts to densely fill the complex plane. $`k_c`$ is the wavenumber such that $`\chi _k=1/r`$, above which the cumulative function $`F_k`$ no longer forms a continuous curve in the complex plane. For the symmetric dyadic case $`r=2`$, the Hausdorff dimension (54) has been previously derived .
###### Example 10.
A numerical computation of Eq. (54) for the totally symmetric triadic random walk is shown in Fig. 6. In this case, $`D_\mathrm{H}(k)=\mathrm{ln}(3)/\mathrm{ln}(1+2\mathrm{cos}k)`$ for $`|k|<k_f=1.1960\mathrm{}`$, while $`D_\mathrm{H}(k)=2`$ for $`k_f|k|<k_c=\pi /2`$.
For small wavenumbers, i.e. $`|k|k_f`$, we can expand the absolute value of the eigenvalue $`\chi _k`$, given by Eq. (26), in terms of the diffusion coefficient (33) of the general $`r`$-adic random walk :
$$|\chi _k|=1𝒟k^2+O(k^4).$$
(55)
Notice that the absolute value of the eigenvalue does not depend on the drift velocity $`v`$ in the linear and quadratic terms of its expansion in powers of the wavenumber $`k`$.
To establish a connection between the diffusion coefficient and the Hausdorff dimension of the hydrodynamic modes (for which $`|k|k_f`$), we let
$$D_\mathrm{H}(k)=1+ak^2+O(k^4),$$
(56)
and expand
$$\underset{j}{}p_j^{D_\mathrm{H}(k)}=1+ak^2\underset{j}{}p_j\mathrm{ln}p_j+O(k^4).$$
(57)
Substituting Eqs. (55) and (57) into Eq. (53), we find
$$a=\frac{𝒟}{{\displaystyle \underset{j}{}}p_j\mathrm{ln}p_j^1}=\frac{𝒟}{\lambda ^+},$$
(58)
where $`\lambda ^+`$ is the positive Lyapunov exponent (14) of our system.
We can thus state the following:
###### Theorem 1.
For an $`r`$-adic multi-baker map with probabilities $`\{p_{(r1)/2},\mathrm{},p_{(r1)/2}\}`$, the cumulative function of the hydrodynamic mode of wavenumber $`k`$, Eq. (28), forms in the complex plane a curve with the Hausdorff dimension
$$D_\mathrm{H}(k)=1+\frac{𝒟}{\lambda ^+}k^2+O\left(k^4\right),$$
(59)
where $`\lambda ^+`$ is the positive Lyapunov exponent (14) of this system and $`𝒟`$ is its diffusion coefficient as specified by Eq. (33).
As a corollary, we have the following:
###### Theorem 2.
Under the same conditions as in Theorem 1, the diffusion coefficient of the system is given in terms of the positive Lyapunov exponent $`\lambda ^+`$ and the Hausdorff dimension $`D_\mathrm{H}(k)`$ of the cumulative function of the hydrodynamic mode of wavenumber $`k`$ according to
$$𝒟=\lambda ^+\underset{k0}{lim}\frac{D_\mathrm{H}(k)1}{k^2}.$$
(60)
###### Remark 4.
The expression (60) bears a relation to the formula (2) giving the diffusion coefficient in terms of the Lyapunov exponent and the Hausdorff codimension of the fractal repeller $`_L`$ of orbits trapped in a 2-dimensional ordered Lorentz gas of size $`L`$ . The analogy is best seen by writing Eq. (60) for a hydrodynamic mode of the smallest possible nonvanishing wavenumber $`k=2\pi /L`$ in a periodic chain of length $`L`$, whereupon Eq.(60) takes the form
$$𝒟=\underset{L\mathrm{}}{lim}\left(\frac{L}{2\pi }\right)^2\lambda ^+\left[D_\mathrm{H}(2\pi /L)1\right],$$
(61)
where $`D_\mathrm{H}(2\pi /L)`$ is the dimension of the fractal curve $`F_k`$ with $`k=2\pi /L`$. Although $`ak^2`$ in Eq. (56) plays a role very similar to the codimension $`c_\mathrm{H}(_L)`$ of Eq. (2), it cannot be identified as such because it is the coefficient of a correction to $`D_\mathrm{H}`$ above 1, which stems from the definition of $`F_k`$ as a cumulative function.
## 6. Conclusions
In this paper, we have shown with the help of simple models of deterministic diffusion that, in the last stages of the relaxation toward the stationary state, the hydrodynamic modes have cumulative functions with fractal Hausdorff dimensions. For small enough wavenumbers, we were able to derive an expression for the Hausdorff dimensions in terms of the diffusion coefficient and positive Lyapunov exponent of the system. Accordingly, the diffusion coefficient can be computed in terms of the Hausdorff dimension and the positive Lyapunov exponent, which is the content of Theorem 2.
Our formula (60) establishes a new relationship between the diffusion coefficient and the characteristic quantities of chaos, which generalizes both the equation (1) from the thermostated-system approach and the equation (2) from the escape-rate formalism, as anticipated in the Introduction. On the one hand, Eq. (60) remains applicable whether the system is thermostated (i.e., in the non-area-preserving case) or not (i.e., in the area-preserving case). On the other hand, no absorbing boundary is required in the present approach where Eq. (60) was derived. Accordingly, the number of particles is kept constant in systems to which Eq. (60) applies. In this regard, our new formula (60) unifies the previous formulae (1) and (2). The novelty of the present approach is to consider the Hausdorff dimension of an abstract curve in the complex plane, instead of the Hausdorff (or information) dimension of either an attractor or a repeller. This abstract curve is the cumulative function of the hydrodynamic mode of diffusion, which is a general concept of nonequilibrium statistical mechanics.
In view of the generality of the considerations developed in the present paper, we conjecture that Eq. (59) for the Hausdorff dimension of the hydrodynamic modes extends to spatially extended Axiom-A systems with two degrees of freedom which sustain transport by diffusion, and more generally to such chaotic systems with two degrees of freedom.
Moreover, the similarity between Eq. (2) and Eq. (61) of the escape-rate formalism suggests a generalization of our formalism to other transport processes. As shown by Dorfman and Gaspard , the escape-rate formalism generalizes to other transport phenomena by setting up a first passage problem in the space of the Helfand moment . In much the same way, we can infer that the eigendistributions of the time evolution in the space of the Helfand moment have cumulative functions with fractal Hausdorff dimensions given by a gradient expansion analogous to Eq. (59), where the diffusion coefficient is replaced by the corresponding transport coefficient. These questions will be addressed in more detail elsewhere.
## acknowledgments
The authors wish to acknowledge a very helpful correspondence with Brian Hunt. T. G. thanks M. Courbage for discussions. P. G. thanks the National Fund for Scientific Research (FNRS Belgium) and the InterUniversity Attraction Pole Program of the Belgian Federal Office of Scientific, Technical and Cultural Affairs for financial support. JRD thanks the National Science Foundation for support under grant PHY 98-20824.
|
warning/0007/astro-ph0007317.html
|
ar5iv
|
text
|
# Recent results by the MOA group on gravitational microlensing
## 1 Observing strategy of MOA
The gravitational microlensing technique<sup>1,2</sup> is being used for a variety of purposes by a Japan/NZ group called MOA. A 60-cm wide-field telescope at the Mt John Observatory in New Zealand and a large format CCD camera that was supplied by the National Astronomical Observatory of Japan are used<sup>3</sup>. The small aperture and wide field of the telescope favour the study of microlensing events in which the magnification is high. Previous experience has shown that the peaks of these events include significant information on the locations<sup>1</sup><sup>1</sup>1Knowledge of the locations of lenses is useful for determining the fraction of dark matter composed of ’machos’. See, e.g., Sumi and Honma<sup>4</sup>. of lens-stars<sup>5</sup>, on the sizes of source-stars<sup>5</sup>, on the atmospheres of source-stars<sup>6</sup>, and on the planetary systems of lens-stars<sup>7,8</sup>. Approximately 20 square degrees in both the galactic bulge and the LMC are observed a few times per night, weather permitting. The bulge data are analysed on-line (see section 4 below) and alerts issued of high magnification events for follow-up by other groups. Also, some alerts issued by other groups are followed. For some events the baseline magnitude is below the limiting magnitude of MOA. Baseline data are required from other groups with larger telescopes, or from the Hubble Space Telescope, for these events.
## 2 Inverse ray shooting with the Kalaka parallel computer
Data for individual events are analysed using the inverse ray shooting technique of Wambsganss<sup>9</sup>. This includes the effects of multi-component lenses and limb-darkened sources of finite size, but it requires considerable computing power. The parallel computer ’Kalaka’ at the University of Auckland is used for this work <sup>2</sup><sup>2</sup>2http://www.scitec.auckland.ac.nz/peter/kalaka.html. This has 200 processors. The full parameter space for a binary lens and a finite-sized source star can be compared with a given dataset in a few hours. For a triple lens a few months is required.
## 3 Test analysis of OGLE-2000-BUL-12
An event with magnification $`50`$ that was reported recently by the OGLE group, OGLE-2000-BUL-12, has been analysed using the Kalaka computer. Data for the event were obtained from the early warning alert site of the OGLE group<sup>3</sup><sup>3</sup>3http://www.astrouw.edu.pl/ftp/ogle/ogle2/ews/ews.html. These data were derived from the ’fixed position’ version of DoPHOT<sup>10</sup>, and they do not allow for possible blending of the source star. The present analysis is therefore preliminary. The light curve for the event is shown above in Fig. 1 together with a best single-lens fit. It was found that better fits could be obtained with two classes of binary lenses, corresponding to a lens star with either a low-mass or a high-mass planet, respectively. These are described below.
### 3.1 Low-mass planet solution
Reasonable fits to the data were obtained with a planet at a projected separation from the lens star of $`(0.91.1)R_E`$ and mass fraction $`(14)\times 10^5`$, as shown in Fig. 2. For a lens mass $`0.3M_{}`$, and lens and source distances $``$ 7 and 9 kpc respectively, these solutions correspond to a planet with mass $`(14)M_E`$ at a projected separation of $`2`$ AU. A non-limb-darkened source star with radius $`1R_{}`$ was assumed in the fitting.
### 3.2 High-mass planet solution
Reasonable fits to the data were also obtained with a planet at a projected separation of $`(0.50.6)R_E`$ or $`(1.62.0)R_E`$ and mass fraction $`(36)\times 10^3`$, as shown in Fig.3. The two distinct solutions, one with the planet inside the Einstein radius and one with it outside, are as expected<sup>11</sup>. The solutions correspond to a planet with mass $`(12)M_J`$ at $`1`$ AU or $`(34)`$ AU.
### 3.3 Discussion
As noted above, the above analysis is only preliminary, because blending has not been allowed for. Additional solutions with a planet very close to the Einstein radius or with more than one planet may also be possible. Further data for the event, if available, should differentiate between the possible solutions. Further data may also become available in the future that would pinpoint the planetary parameters. In approximately one decade, the lens and source stars may diverge sufficiently to enable the next generation of space telescopes to resolve them separately. In this case, the mass of the lens star, and the distances of the lens and source stars, may be directly measurable. This would leave no parameters undetermined.
## 4 MOA alert system using difference imaging
On-line data analysis has recently been implemented by MOA. This uses difference imaging in a manner that is similar to that of Alard and Lupton<sup>12</sup>. This enables events that are magnified from beneath the limiting magnitude to be readily detected, and also it ameliorates the effects of crowding and blending that are necessarily present. Some details of the technique were described previously<sup>13</sup>. Full details will be presented in the future. Alerts of events in progress are being reported on the web<sup>4</sup><sup>4</sup>4http://www.phys.canterbury.ac.nz/physib/alert/alert.html.
## Acknowledgments
The authors thank the OGLE group for making their data available and for discussion, and Peter Dobcsanyi and Michael Harre for assistance. Financial support by the Marsden Fund of NZ and the conference organizers is gratefully acknowledged.
|
warning/0007/quant-ph0007076.html
|
ar5iv
|
text
|
# Schrödinger-Cat Entangled State Reconstruction in the Penning Trap
## I Introduction
A single electron trapped in a Penning trap is a unique quantum system in that it allows the measurement of fundamental physical constants with striking accuracy. Recently, the electron cyclotron degree of freedom has been cooled to its ground state, where the electron may stay for hours, and quantum jumps between adjacent Fock states have been observed . It is therefore evident that the determination of the genuine (possibly entangled) quantum state of the trapped electron is an important issue, with implications in the very foundations of physics, and in particular of quantum mechanics. After the pioneering work of Vogel and Risken , several methods have been proposed in order to reconstruct the quantum state of light and matter , which range from quantum tomography through quantum state endoscopy , to Wigner function determination from outcome probabilities . Also, different techniques have been proposed which allow to deal with entangled states.
In fact, entanglement has been defined as one of the most puzzling features of quantum mechanics, and it is at the heart of quantum information processing. Some fascinating examples of the possibilities offered by sharing quantum entanglement are quantum teleportation , quantum dense coding , entanglement swapping , quantum cryptography , and quantum computation . A striking achievement in this field has been the recent entanglement of four trapped ions .
In the present work we propose to reconstruct the full entangled state (combined cyclotron and spin state) of an electron in a Penning trap by using a modified version of quantum state tomography. Previous proposals need the a priori knowledge of the spin state and therefore are not able to deal with entangled states. Our method, on the contrary, has the ability of measuring the full (entangled) pure state of the two relevant degrees of freedom of the electron. In order to reach this scope, our method takes advantage of the magnetic bottle configuration to perform simultaneous measurements of the cyclotron excitation and of the $`z`$ component of the spin as a function of the phase of an applied driving electromagnetic field. The complete structure of the cyclotron-spin quantum state is then obtained with the help of a tomographic reconstruction from the measured data.
The present paper is organized as follows: In Sec. II we outline the basic model of an electron trapped in a Penning trap, while in Sec. III we describe the main idea of our reconstruction procedure. In Sec. IV and V we concentrate on the measurement of the spin and on the tomographic reconstruction of the cyclotron states, respectively. We present the results of our numerical simulations in Sec. VI, and conclude briefly in Sec. VII.
## II The basic model
Let us consider the motion of an electron trapped by the combination of a homogeneous magnetic field $`𝐁_0`$ along the positive $`z`$ axis and an electrostatic quadrupole potential in the $`xy`$ plane, which is known as a Penning trap . The corresponding Hamiltonian can be written as
$$\widehat{H}=\frac{1}{2m}[𝐩\frac{q}{c}𝐀]^2+\frac{qV_o}{2d^2}\left(\widehat{z}{}_{}{}^{2}\frac{\widehat{x}{}_{}{}^{2}+\widehat{y}^2}{2}\right),$$
(1)
where $`𝐀=(B_oy/2,B_ox/2,0)`$ is the vector potential, $`c`$ is the speed of light, $`d`$ characterizes the dimensions of the trap, $`V_o`$ is the electrostatic potential applied to its electrodes, and $`q`$ the electron charge.
The spatial part of the electronic wave function consists of three degrees of freedom, but neglecting the slow magnetron motion (whose characteristic frequency lies in the kHz region), here we only consider the axial and cyclotron motions, which are two harmonic oscillators radiating in the MHz and GHz regions, respectively. The spin dynamics results from the interaction between the magnetic moment of the electron and the magnetic field, so that the total quantum Hamiltonian is
$$\widehat{H}_{\mathrm{tot}}=\mathrm{}\omega _z(\widehat{a}_{}^{}{}_{z}{}^{}\widehat{a}_z+1/2)+\mathrm{}\omega _c(\widehat{a}_{}^{}{}_{c}{}^{}\widehat{a}_c+1/2)+\frac{\mathrm{}\omega _s}{2}\widehat{\sigma }_z.$$
(2)
In the previous expression we have introduced the lowering operator for the cyclotron motion
$$\widehat{a}_c=\frac{1}{2}\left[\frac{1}{2\beta \mathrm{}}(\widehat{p}_xi\widehat{p}_y)i\beta (\widehat{x}i\widehat{y})\right],$$
(3)
where $`\beta =m\omega _c/2\mathrm{}`$ and $`\omega _c=qB_o/mc`$ is the resonance frequency associated to the cyclotron oscillation. For the axial motion we have
$$\widehat{a}_z=\left(\frac{m\omega _z}{2\mathrm{}}\right)^{1/2}\widehat{z}+i\left(\frac{1}{2m\mathrm{}\omega _z}\right)^{1/2}\widehat{p}_z,$$
(4)
where $`\omega _{z}^{}{}_{}{}^{2}=qV_o/md^2`$. In the last term of Eq. (2), $`\widehat{\sigma }_z`$ is the Pauli spin matrix and $`\omega _s=(g/2)\omega _c`$.
The obtained Hamiltonian (2) is then made of three independent terms. Even though the only physical observable experimentally detectable is the axial momentum $`\widehat{p}_z`$, in the following both the cyclotron and spin states will be reconstructed. Considering the eigenstates of $`\widehat{\sigma }_z`$
$$|=\left(\begin{array}{c}1\\ 0\end{array}\right),|=\left(\begin{array}{c}0\\ 1\end{array}\right),$$
(5)
we can write the most general pure state of the trapped electron in the form
$$|\mathrm{\Psi }=c_1|\psi _1|+c_2|\psi _2|,$$
(6)
$`|\psi _1`$ and $`|\psi _2`$ being two unknown cyclotron states. The complex coefficients $`c_1`$ and $`c_2`$, satisfying the normalization condition $`|c_1|^2+|c_2|^2=1`$, are also to be determined.
The electronic state (6) possesses two very interesting features: first, if the cyclotron states $`|\psi _1`$ and $`|\psi _2`$ are macroscopically distinguishable, $`|\mathrm{\Psi }`$ is a typical example of Schrödinger-cat state . Second, the full state of the trapped electron is an entangled state between the spin and cyclotronic degrees of freedom (unless $`|\psi _1=|\psi _2`$). Introducing the total density operator $`\widehat{R}=|\mathrm{\Psi }\mathrm{\Psi }|`$ associated to the pure state $`|\mathrm{\Psi }`$, we can express the corresponding total density matrix $`R`$ in the basis of the eigenstates (5) of $`\widehat{\sigma }_z`$ in the form
$$R=\left(\begin{array}{cc}|c_1|^2|\psi _1\psi _1|& c_1c_2^{}|\psi _1\psi _2|\\ & \\ c_2c_1^{}|\psi _2\psi _1|& |c_2|^2|\psi _2\psi _2|\end{array}\right),$$
(7)
whose elements are operators. Its diagonal elements represent the possible cyclotron states, while the off-diagonal ones are the quantum coherences and contain information about the quantum interference effects due to the entanglement between the spin and cyclotron degrees of freedom.
It is also possible to give a phase-space description of the complete quantum state of the trapped electron by introducing the Wigner-function matrix whose elements are given by
$$W_{ij}(\alpha )=\widehat{\delta }_{ij}(\alpha \widehat{a})=\mathrm{Tr}[\widehat{R}\widehat{\delta }_{ij}(\alpha \widehat{a})],$$
(8)
where $`i,j=1,2`$ and $`\widehat{\delta }_{ij}(\alpha \widehat{a})`$ is an operator in the product Hilbert space $`=_{\mathrm{cyc}}_{\mathrm{spin}}`$ defined as
$$\widehat{\delta }_{ij}(\alpha \widehat{a})=|ij|\widehat{\delta }(\alpha \widehat{a}).$$
(9)
In the previous expression the operator-valued delta function $`\widehat{\delta }(\alpha \widehat{a})`$ is the Fourier transform of the displacement operator $`\widehat{D}(\alpha )=\mathrm{exp}(\alpha \widehat{a}^{}\alpha ^{}\widehat{a})`$.
## III Measurement Scheme and Reconstruction Procedure
The basic idea of our reconstruction procedure is very simple: Adding a particular inhomogeneous magnetic field—known as the “magnetic bottle” field —to that already present in the trap, it is possible to perform a simultaneous measurement of both the spin and the cyclotronic excitation numbers. Repeated measurements of this type allow us to recover the probability amplitudes associated to the two possible spin states and the cyclotron probability distribution $`P(n_c)`$ in the Fock basis. The reconstruction of the cyclotron density matrices $`\rho _{ii}=|\psi _i\psi _i|`$ ($`i=1,2`$) in the Fock basis is then possible by employing a technique similar to the Photon Number Tomography (PNT) which exploits a phase-sensitive reference field that displaces in the phase space the particular state one wants to reconstruct .
In close analogy with the procedure described in Refs. , the coupling between the different degrees of freedom in Eq. (2) is obtained modifying the vector potential with the addition of the magnetic bottle field so that $`𝐀`$ takes the form
$$𝐀=\frac{1}{2}[B_0\widehat{y}b\left(\widehat{y}\widehat{z}^2\frac{\widehat{y}^3}{3}\right),B_0\widehat{x}+b\left(\widehat{x}\widehat{z}^2\frac{\widehat{x}^3}{3}\right),0].$$
(10)
Such a vector potential gives rise to an interaction term in the total Hamiltonian,
$$\widehat{H}_{\mathrm{int}}=\mathrm{}\kappa \left[\left(\widehat{a}_c^{}\widehat{a}_c+\frac{1}{2}\right)+\frac{g}{2}\frac{\widehat{\sigma }_z}{2}\right]\widehat{z}^2,$$
(11)
where the coupling constant $`\kappa =qb/mc`$ is directly related to the strength $`b`$ of the magnetic bottle field.
Eq. (11) describes the the fact that the axial angular frequency is affected both by the number of cyclotron excitations $`\widehat{n}_c=\widehat{a}_c^{}\widehat{a}_c`$ and by the eigenvalue of $`\widehat{\sigma }_z`$. In terms of the lowering operator $`\widehat{a}_z`$ for the axial degree of freedom, the Hamiltonian that describes the interaction among the axial, cyclotron, and spin motions can be written as
$$\widehat{H}_{\mathrm{czs}}=\mathrm{}\omega _z\left(\widehat{a}_z^{}\widehat{a}_z+\frac{1}{2}\right)\frac{\widehat{\mathrm{\Omega }}_z^2}{\omega _z^2},$$
(12)
where the operator frequency $`\widehat{\mathrm{\Omega }}_z`$ is given by
$$\widehat{\mathrm{\Omega }}_z^2=\omega _z^2+\frac{\mathrm{}\kappa }{m}\left[\left(\widehat{a}_c^{}\widehat{a}_c+\frac{1}{2}\right)+\frac{g}{4}\widehat{\sigma }_z\right].$$
(13)
The operator $`\widehat{\mathrm{\Omega }}_z`$ is the modified axial frequency which can be experimentally measured after the application of the inhomogeneous magnetic bottle field. What is actually measured is an electric current (which is proportional to $`\widehat{p}_z`$) that gives the axial frequency shift $`\widehat{\mathrm{\Omega }}_z^2`$ . One immediately sees that the spectrum of $`\widehat{\mathrm{\Omega }}_z`$ is discrete: Since the electron $`g`$ factor is slightly (but measurably ) different from 2, $`\widehat{\mathrm{\Omega }}_z`$ assumes a different value for every pair of eigenvalues of $`\widehat{n}_c`$ and $`\widehat{\sigma }_z`$.
## IV Spin Measurements
If one can perform a large set of measurements of $`\widehat{\mathrm{\Omega }}_z`$ in such a way that before each measurement the state $`|\mathrm{\Psi }`$ is always prepared in the same way, it is possible to recover the probabilities $`P()`$ and $`P()`$ associated to the two possible eigenvalues of $`\widehat{\sigma }_z`$, namely $`|c_1|`$ and $`|c_2|`$. Recalling Eq. (6), we have
$`P()`$ $`=`$ $`\mathrm{Tr}_{\mathrm{cyc}}[|\mathrm{\Psi }\mathrm{\Psi }|]=|c_1|^2,`$ (15)
$`P()`$ $`=`$ $`\mathrm{Tr}_{\mathrm{cyc}}[|\mathrm{\Psi }\mathrm{\Psi }|]=|c_2|^2=1P().`$ (16)
However, this kind of measurement does not allow to retrieve the relative phase $`\theta `$ between the complex coefficients $`c_1`$ and $`c_2`$ in the superposition (6). We can then add a time-dependent magnetic field $`𝐛_0(t)`$ oscillating in the $`xy`$ plane perpendicular to the trap axis , i.e.
$$𝐛_0(t)=b_0\mathrm{cos}(\omega t)\widehat{𝐱}+b_0\mathrm{sin}(\omega t)\widehat{𝐲}.$$
(17)
The resulting interaction Hamiltonian in the interaction picture is
$`\widehat{H}_{\mathrm{rot}}^\mathrm{I}(t)`$ $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{i}{\mathrm{}}}\widehat{H}_0t\right]\widehat{H}_{\mathrm{rot}}(t)\mathrm{exp}\left[{\displaystyle \frac{i}{\mathrm{}}}\widehat{H}_0t\right]`$ (18)
$`=`$ $`{\displaystyle \frac{\mathrm{}}{2}}[(\omega _s\omega )\widehat{\sigma }_z+\omega _R\widehat{\sigma }_x],`$ (19)
where
$`\widehat{H}_0`$ $`=`$ $`{\displaystyle \frac{\mathrm{}\omega }{2}}\widehat{\sigma }_z,`$ (21)
$`\widehat{H}_{\mathrm{rot}}(t)`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{2}}[(\omega _s\omega )\widehat{\sigma }_z`$ (23)
$`+\omega _R(\widehat{\sigma }_x\mathrm{cos}(\omega t)+\widehat{\sigma }_y\mathrm{sin}(\omega t))].`$
In the above equations, $`\widehat{H}_{\mathrm{rot}}(t)`$ is the interaction Hamiltonian in a frame rotating at the driving frequency $`\omega `$, while $`\omega _R=gqb_0/2mc`$ is the Rabi frequency.
The evolution of the state (6) subjected to the Hamiltonian (19) in the resonant case $`\omega =\omega _s`$, yields the state
$`|\mathrm{\Psi }(\overline{t})`$ $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{i}{\mathrm{}}}\widehat{H}_{\mathrm{rot}}^\mathrm{I}(t)\overline{t}\right]|\mathrm{\Psi }`$ (24)
$`=`$ $`{\displaystyle \frac{\sqrt{2}}{2}}[(c_1|\psi _1ic_2|\psi _2)|`$ (26)
$`+(ic_1|\psi _1+c_2|\psi _2)|],`$
obtained applying the driving field (17) for a time $`\overline{t}=\pi /2\omega _R`$. We can now repeat the spin measurements just as we have described above in the case of the unknown initial state $`|\mathrm{\Psi }`$: Soon after $`𝐛_0(t)`$ is switched off, the magnetic bottle field is applied again and the spin measurement is performed. Repeating this procedure over and over again (with the same unknown initial state) for a large number of times, it is possible to recover the probabilities $`\overline{P}()`$ and $`\overline{P}()`$ associated to the two spin eigenvalues for the state $`|\mathrm{\Psi }(\overline{t})`$ of Eq. (26). Without loss of generality, we can assume $`c_1R`$, $`c_2=|c_2|e^{i\theta }`$, and $`\psi _1|\psi _2=re^{i\beta }`$, which yield
$`\overline{P}()`$ $`=`$ $`{\displaystyle \frac{1}{2}}[1+2r|c_1||c_2|\mathrm{sin}(\theta +\beta )]`$ (28)
$`\overline{P}()`$ $`=`$ $`{\displaystyle \frac{1}{2}}[12r|c_1||c_2|\mathrm{sin}(\theta +\beta )].`$ (29)
It is important to note that the probabilities $`\overline{P}()`$ and $`\overline{P}()`$ can be experimentally sampled and that the modulus $`r`$ and the phase $`\beta `$ of the scalar product $`\psi _1|\psi _2`$ can be both derived from the reconstruction of the cyclotron density matrices $`\rho _{11}`$ and $`\rho _{22}`$, as we shall explain in the next section. Thus we are able to find the relative phase $`\theta `$ by simply inverting one of the two Eqs. (IV), e.g.
$$\theta =\mathrm{arcsin}\left[\frac{2\overline{P}()1}{2r|c_1||c_2|}\right]\beta .$$
(30)
The resulting $`\pi `$ ambiguity in the $`\mathrm{arcsin}`$ function in the right hand side of Eq. (30) can be eliminated by choosing a second interaction time $`\overline{t}^{}`$ and repeating the procedure above.
## V Tomographic Reconstruction of the Cyclotron States
Let us consider again the state of Eq. (6): every time $`\widehat{\mathrm{\Omega }}_z`$ (and therefore $`\widehat{\sigma }_z`$) is measured, the total wave function is projected onto $`||n_c`$ or $`||n_c`$, where $`|n_c`$ is a cyclotron Fock state. We then propose a tomographic reconstruction technique in which the state to be measured is combined with a reference field whose complex amplitude is externally varied (as it is usually done in optical homodyne tomography ) in order to displace the unknown density operator in the phase space (a technique very close to the PNT scheme ). In particular, we shall sample the cyclotron density matrix in the Fock basis by varying only the phase $`\phi `$ of the reference field, leaving unaltered its modulus $`|\alpha |`$ .
Following Ref. , immediately before the measurement of $`\widehat{\mathrm{\Omega }}_z`$, we apply to the trap electrodes a driving field generated by the vector potential
$$𝐀=(2\frac{mc}{\beta |q|}Im[ϵe^{i\omega _ct}],2\frac{mc}{\beta |q|}Re[ϵe^{i\omega _ct}],0),$$
(31)
where $`ϵ`$ is the field amplitude, which gives rise to a Hamiltonian term of the form
$$\widehat{H}_{\mathrm{drive}}=i\mathrm{}(ϵe^{i\omega _ct}\widehat{a}_c^{}ϵ^{}e^{i\omega _ct}\widehat{a}_c).$$
(32)
The time evolution of the projected density operator $`\widehat{\rho }_{ii}`$ ($`i=1,2`$) according to the Hamiltonian (32) may then be written in the cyclotron interaction picture as
$`\widehat{\rho }_{ii}(ϵ,t)`$ $`=`$ $`\mathrm{exp}\left({\displaystyle \frac{i}{\mathrm{}}}\stackrel{~}{H}_{\mathrm{drive}}t\right)\widehat{\rho }_{ii}(0)\mathrm{exp}\left({\displaystyle \frac{i}{\mathrm{}}}\stackrel{~}{H}_{\mathrm{drive}}t\right)`$ (33)
$`=`$ $`\widehat{D}^{}(\alpha )\widehat{\rho }_{ii}(0)\widehat{D}(\alpha ),`$ (34)
where we have defined the complex parameter $`\alpha =ϵt`$ ($`t`$ being the interaction time) and $`\stackrel{~}{H}_{\mathrm{drive}}`$ is given by
$$\stackrel{~}{H}_{\mathrm{drive}}=i\mathrm{}(ϵ\widehat{a}_c^{}ϵ^{}\widehat{a}_c).$$
(35)
The right-hand side of Eq. (34) is then the desired displaced density operator, where the displacement parameter $`\alpha `$ is a function of both the strength $`ϵ`$ of the driving field and the interaction time $`t`$. Thus we can interpret the quantity
$`P^{(i)}(n_c,\alpha )`$ $`=`$ $`\mathrm{Tr}[\widehat{\rho }_{ii}(\alpha )|n_cn_c|]`$ (36)
$`=`$ $`n_c|\widehat{D}^{}(\alpha )\widehat{\rho }_{ii}(0)\widehat{D}(\alpha )|n_c`$ (37)
$`=`$ $`n_c,\alpha |\widehat{\rho }_{ii}(0)|n_c,\alpha ,`$ (38)
as the probability of finding the cyclotron state $`|\psi _i`$ with an excitation number $`n_c`$ after the application of the driving field of amplitude $`ϵ`$ for a time $`t`$. Fixing a particular value of $`\alpha `$, and measuring $`\widehat{n}_c`$, it is then possible to recover the probability distribution (38) performing many identical experiments.
Expanding the density operator $`\widehat{\rho }_{ii}`$ in the Fock basis, and defining $`N_c`$ as an appropriate estimate of the maximum number of cyclotronic excitations (cut-off), we have
$$P^{(i)}(n_c,\alpha )=\underset{k,m=0}{\overset{N_c}{}}n_c,\alpha |kk|\widehat{\rho }_{ii}|mm|n_c,\alpha .$$
(39)
The projection of the displaced number state $`|n_c,\alpha `$ onto the Fock state $`|m`$ can be obtained (generalizing the result derived in Ref. ) as
$`m|n,\alpha `$ $`=`$ $`m|\widehat{D}(\alpha )|n=e^{|\alpha |^2}\sqrt{{\displaystyle \frac{\nu !}{\mu !}}}|\alpha |^{\mu \nu }`$ (42)
$`\times \mathrm{exp}\left\{i(mn)\left[\phi \pi \theta (nm)\right]\right\}`$
$`\times L_\nu ^{\mu \nu }(|\alpha |^2),`$
where $`\theta (x)`$ is the Heaviside function, $`L_\nu ^\mu `$ is the associated Laguerre polynomial, and $`\mu =\mathrm{max}\{m,n\}`$, $`\nu =\mathrm{min}\{m,n\}`$. Inserting Eq. (42) into Eq. (39), we get
$`P^{(i)}(n,\alpha )`$ $`=`$ $`e^{|\alpha |^2}{\displaystyle \underset{k,m=0}{\overset{N_c}{}}}\sqrt{{\displaystyle \frac{\nu !\overline{\nu }!}{\mu !\overline{\mu }!}}}|\alpha |^{\mu +\overline{\mu }\nu \overline{\nu }}L_\nu ^{\mu \nu }(|\alpha |^2)`$ (45)
$`\times e^{i(mk)\phi \pi \left[(mn)\theta (nm)(kn)\theta (nk)\right]}`$
$`\times L_{\overline{\nu }}^{\overline{\mu }\overline{\nu }}(|\alpha |^2)k|\widehat{\rho }_{ii}|m,`$
where $`\overline{\mu }=\mathrm{max}\{k,n\}`$ and $`\overline{\nu }=\mathrm{min}\{k,n\}`$.
Let us now consider, for a given value of $`|\alpha |`$, $`P^{(i)}(n,\alpha )`$ as a function of $`\phi `$ and calculate the coefficients of the Fourier expansion
$$P^{(s,i)}(n,|\alpha |)=\frac{1}{2\pi }_0^{2\pi }𝑑\phi P^{(i)}(n,\alpha )e^{i\phi },$$
(46)
for $`s=0,1,2,\mathrm{}`$. Combining Eqs. (45) and (46), we get
$$P^{(s,i)}(n,|\alpha |)=\underset{m=0}{\overset{N_cs}{}}G_{n,m}^{(s)}(|\alpha |)m+s|\widehat{\rho }_{ii}|m,$$
(47)
where we have introduced the matrices
$`G_{n,m}^{(s)}(|\alpha |)`$ $`=`$ $`e^{i\pi [(s+mn)\theta (nsm)(mn)\theta (nm)]}`$ (48)
$`\times `$ $`e^{|\alpha |^2}\left({\displaystyle \frac{\nu !\overline{\nu }!}{\mu !\overline{\mu }!}}\right)^{\frac{1}{2}}|\alpha |^{\mu +\overline{\mu }\nu \overline{\nu }}L_\nu ^{\mu \nu }(|\alpha |^2)`$ (49)
$`\times `$ $`L_{\overline{\nu }}^{\overline{\mu }\overline{\nu }}(|\alpha |^2),`$ (50)
with $`\overline{\mu }=\mathrm{max}\{m+s,n\}`$ and $`\overline{\nu }=\mathrm{min}\{m+s,n\}`$.
We may now note that if the distribution $`P^{(i)}(n,\alpha )`$ is measured for $`n[0,N]`$ with $`NN_c`$, then Eq. (47) represents for each value of $`s`$ a system of $`N+1`$ linear equations between the $`N+1`$ measured quantities and the $`N_c+1s`$ unknown density matrix elements. Therefore, in order to obtain the latter, we only need to invert the system
$$m+s|\widehat{\rho }_{ii}|m=\underset{n=0}{\overset{N}{}}M_{m,n}^{(s)}(|\alpha |)P^{(s,i)}(n,|\alpha |),$$
(51)
where the matrices $`M`$ are given by $`M=(G^TG)^1G^T`$. It is possible to see that these matrices satisfy the relation
$$\underset{n=0}{\overset{N}{}}M_{m^{},n}^{(s)}(|\alpha |)G_{n,m}^{(s)}(|\alpha |)=\delta _{m^{},m},$$
(52)
for $`m,m^{}=0,1,\mathrm{},N_cs,\mathrm{}`$, which means that from the exact probabilities satisfying Eq. (47) the correct density matrix elements are obtained. Furthermore, combining Eqs. (46) and (51), we find
$$m+s|\widehat{\rho }_{ii}|m=\frac{1}{2}\underset{n=0}{\overset{N}{}}_0^{2\pi }𝑑\phi M_{m,n}^{(s)}(|\alpha |)e^{is\phi }P^{(i)}(n,\alpha ),$$
(53)
which may be regarded as the formula for the direct sampling of the cyclotron density matrix. In particular, Eq. (53) clearly shows that the determination of the cyclotron state requires only the value of $`\phi `$ to be varied.
We now only need to reconstruct the off-diagonal parts of the total density matrix (7), i.e. $`\widehat{\rho }_{12}=\widehat{\rho }_{21}^{}=|\psi _1\psi _2|`$. This can easily be done under the assumption that the initial unknown electron state is pure, as in Eq. (6). Then, we have
$$(\rho _{12})_{n,m}=\frac{_{i=0}^N(\rho _{11})_{n,i}(\rho _{22})_{i,m}}{\psi _1|\psi _2}.$$
(54)
Writing $`|\psi _1=_{i=0}^Na_n|n`$ and $`|\psi _2=_{i=0}^Nb_n|n`$ in the Fock basis, we can obtain the desired coefficients from the recursive relation
$$a_n=\left[\frac{(\rho _{11})_{n1,n}}{a_{n1}}\right]^{},$$
(55)
where without loss of generality we can set $`a_1=[(\rho _{11})_{1,1}]^{1/2}R`$. A similar relation yields the coefficients $`b_n`$ of $`|\psi _2`$. Finally, the scalar product in Eq. (54) may be determined as
$$\psi _1|\psi _2=\underset{i=1}{\overset{N}{}}a_i^{}b_i.$$
(56)
## VI Simulations and Results
In this section we show the results of numerical Monte-Carlo simulations of the method presented above, which allow us to state that this technique may be quite accurate also in the experimental implementation. To account for actual experimental conditions, we have considered the effects of a non-unit quantum efficiency $`\eta `$ in the counting of cyclotronic excitations. When $`\eta <1`$, the actually measured distribution $`𝒫(k,\alpha )`$ is related to the ideal distribution $`P(n,\alpha )`$ by the binomial convolution
$$𝒫(k,\alpha )=\underset{n=k}{\overset{\mathrm{}}{}}_{k,n}(\eta )P(n,\alpha ),$$
(57)
where
$$_{k,n}(\eta )=\left(\begin{array}{c}n\\ k\end{array}\right)\eta ^k(1\eta )^{nk}.$$
(58)
Eq. (47) is then modified as
$$𝒫^{(s,i)}(n,|\alpha |)=\underset{m=0}{\overset{N_cs}{}}𝒢_{n,m}^{(s)}(|\alpha |,\eta )m+s|\widehat{\rho }_{ii}|m,$$
(59)
where $`𝒫^{(s,i)}(n,|\alpha |)`$ is again defined according to Eq. (46), but now with $`𝒫^{(i)}(n,\alpha )`$ in place of $`P^{(i)}(n,\alpha )`$. In addition, $`𝒢_{n,m}^{(s)}(|\alpha |,\eta )`$ is defined as
$$𝒢_{n,m}^{(s)}(|\alpha |,\eta )=\underset{k=0}{\overset{\mathrm{}}{}}_{k,n}(\eta )G_{k,m}^{(s)}(|\alpha |).$$
(60)
The matrices $`𝒢_{n,m}^{(s)}(|\alpha |,\eta )`$ can then be inverted in the way described in the previous section to obtain the matrices $`_{n,m}^{(s)}(|\alpha |,\eta )`$ that can be used to reconstruct the density matrix.
In the above reconstruction procedure, we have to consider two possible sources of errors associated to any actual measurement process. First, we have noticed a strong correlation between the statistical error in the simulated reconstruction and the absolute value $`|\alpha |`$ of the coherent field amplitude applied to drive the cyclotron motion. When $`|\alpha |`$ is small, the density matrix elements near to the main diagonal are accurately reconstructed. Progressively increasing $`|\alpha |`$, the reconstruction of the off-diagonal elements becomes more accurate, while the elements close to the diagonal present significant fluctuations: to compensate for this effect, it is necessary to increase the number of measurements as much as possible. Another source of error stems from the truncation of the reconstructed density matrix at the value $`N_c`$ in the Fock space. Neglecting the terms with $`n>N_cs`$ in Eq. (47) causes a systematic error. This error can be reduced by increasing $`N_c`$, which however may also cause an increase of the statistical error. This suggests that for a given number of measurement events there is an optimal value of $`N_c`$ for which the systematic error is reduced below the statistical error.
In the following we will present two examples of application of the above method. We shall show the simulated tomographic reconstructions of an entangled electronic state of the type
$$|\mathrm{\Psi }=c_1|\gamma |+c_2e^{i\theta }|\gamma e^{i\xi }|,$$
(61)
in which $`c_1`$, $`c_2`$, $`\theta `$, $`\gamma `$, and $`\xi `$ are real parameters, and $`|\gamma `$ is a coherent state of the cyclotron oscillator. When $`\xi 0`$, the state (61) is entangled, and when $`\xi =\pi `$ (with $`|\gamma |1`$) it represents the most “genuine” example of a Schrödinger-cat state . We shall display the results of our simulations both in terms of the density matrix of Eq. (7) and of the Wigner-function matrix of Eq. (8). The latter can be derived from the density matrix elements $`(\rho _{ij})_{n,m}`$ through the relation
$`W_{ij}(x,y)`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}{\displaystyle \underset{n}{}}(\rho _{ij})_{n,n}(1)^ne^{2(x^2+y^2)}L_n[4(x^2+y^2)]`$ (62)
$`+`$ $`{\displaystyle \frac{4}{\pi }}{\displaystyle \underset{nm}{}}(\rho _{ij})_{n,m}(1)^n\sqrt{{\displaystyle \frac{n!}{m!}}}[2(x+iy)]^{mn}`$ (63)
$`\times `$ $`e^{2(x^2+y^2)}L_n^{mn}[4(x^2+y^2)].`$ (64)
In Figs. 1 and 2 we show the numerical reconstruction of an asymmetric superposition of the type of Eq. (61) with $`c_1=1/2`$, $`c_2=\sqrt{3}/2`$, $`\theta =\pi `$, $`\xi =\pi `$, $`\eta =0.9`$ and $`\gamma =1`$. In Fig. 1 we plot the results concerning the density matrix, while in Fig. 2 those concerning the Wigner function. In each figure, the true distributions are depicted on the left next to the corresponding reconstructed ones. Both the density matrices and the Wigner functions are well reconstructed.
In Figs. 3 and 4 analogous results are shown for a symmetric superposition state (61) with $`c_1=c_2=\sqrt{2}/2`$, $`\theta =0`$, $`\xi =\pi `$, $`\eta =0.9`$ and $`\gamma =1.5`$. Again, the reconstruction is faithful. We would like to emphasize the particular shape of $`\rho _{12}`$ and $`W_{12}`$ in both the examples above. It is due to the quantum interference given by the entanglement between the two degrees of freedom: in fact, in absence of entanglement ($`\xi =0`$) $`\rho _{12}`$ would just be a replica of the diagonal parts $`\rho _{11}`$ and $`\rho _{22}`$. In both the cases considered (Figs. 1 and 3) the off-diagonal density matrix elements are real, due to the particular choice of the cyclotron states. The imaginary parts of the reconstructed density matrices turn out to be smaller than 10<sup>-3</sup>.
We have performed a large number of simulations with different states and several values of the parameters, which confirm that the present method is quite stable and accurate. In addition, and for all the cases considered, the values of the parameters $`c_1`$, $`c_2`$, and $`\theta `$ (which obviously do not enter the plots of Figs. 14) are very well recovered, with a relative error of the order of $`10^5`$.
## VII Conclusions
In this paper we have proposed a technique suitable to reconstruct the (entangled) state of the cyclotron and spin degrees of freedom of an electron trapped in a Penning trap. It is based on the magnetic bottle configuration, which allows simultaneous measurements of the spin component along the $`z`$ axis and of the cyclotron excitation number. The cyclotron state is reconstructed with the use of a tomographic-like method, in which the phase of a reference driving field is varied. The numerical results based on Monte-Carlo simulations indicate that even in the case of a non-unit quantum efficiency the reconstructed density matrices and Wigner functions are almost identical to the ideal distributions. An experimental implementation of the proposed method might yield new insight in the foundations of quantum mechanics and allow further progress in the field of quantum information .
###### Acknowledgements.
We gratefully thank G. M. D’Ariano for help with the numerical simulations. This work has been partially supported by INFM (through the 1997 Advanced Research Project “CAT”), by the European Union in the framework of the TMR Network “Microlasers and Cavity QED”, and by MURST under the “Cofinanziamento 1997”.
|
warning/0007/cond-mat0007207.html
|
ar5iv
|
text
|
# Role of the unstable directions in the equilibrium and aging dynamics of supercooled liquids.
## Abstract
The connectivity of the potential energy landscape in supercooled atomic liquids is investigated through the calculation of the instantaneous normal modes spectrum and a detailed analysis of the unstable directions in configuration space. We confirm the hypothesis that the mode-coupling critical temperature is the $`T`$ at which the dynamics crosses over from free to activated exploration of configuration space. We also report the observed changes in the local connectivity of configuration space sampled during aging, following a temperature jump from a liquid to a glassy state.
Understanding the microscopic mechanism for the incredible slowing down of the dynamics in supercooled glass forming liquids is one of the hot topics in condensed matter physics. In recent years, the combined effort of high level experimental techniques, computational analysis and sophisticated theoretical approaches has provided an enormous amount of novel information. In particular, it appears more and more clearly that in addition to the melting temperature and the calorimetric glass transition temperature, $`T_g`$, another temperature plays a relevant role. This temperature, located between 1.2 and 2 $`T_g`$ depending on the fragility of the liquid, signals a definitive change in the microscopic processes leading to structural relaxation. Mode Coupling Theory (MCT) was first in identifying the role of this cross-over temperature, $`T_c`$. According to MCT, for $`TT_c`$ the molecular dynamics is controlled by the statistics of the orbits in phase space , while below $`T_c`$, the dynamics becomes controlled by phonon assisted processes. Studies based on disordered mean field $`p`$-spin models have also stressed the role of such a cross-over temperature.
Computer simulation studies of realistic models of liquids have addressed the issue of the structure of configuration space in supercooled states . Two different techniques have provided relevant information on phase space structure: the instantaneous normal mode approach (INM), which focuses on the properties of the finite temperature Hessian, allowing the calculation of the curvature of the potential energy surface (PES) along $`3N`$ independent directions and (ii) the inherent structure (IS) approach, which focuses on the local minima of the potential energy. Different from mean field models, computer simulation analysis provides a description based on the system’s potential energy, the free energy entering only via the equilibrium set of analyzed configurations.
In principle, the INM approach should be well suited for detecting a change in the structure of the PES visited above and below $`T_c`$. A plot of the $`T`$-dependence of the fraction of directions in configuration space with negative (unstable) curvature, $`f_u`$, should reveal the presence of $`T_c`$. Unfortunately, as it was soon found out, anharmonic effects play a non negligible role and several of the negative curvature directions are observed even in crystalline states, where diffusivity is negligible. A similar situation is seen in p-spin models, where the number of negative eigenvalues of the Hessian is not zero at $`T_c`$ . Bembenek and Laird suggested inspecting the energy profile along the unstable directions to partition the negative eigenmodes in shoulder modes ($`sh`$) (i.e. anharmonic effects) and double-well ($`dw)`$ modes (i.e. directions connecting different basins). While for a particular molecular system the fraction of double well modes $`f_{dw}`$ has been shown to go to zero close to $`T_c`$, for atomic system, $`f_{dw}`$ is significantly different from zero even in crystalline and glassy states. The existence of $`dw`$ directions in thermodynamic conditions where the diffusivity is zero, i.e. in situations where the system is constrained in a well defined basin, strongly suggests that the two minima joined by the unstable double-well directions may lead to the same inherent structure IS. Such possibility has been demonstrated very clearly in Ref..
The aim of this Letter is to present a detailed evaluation of: (i) the number of escape directions, $`N_{escape}`$, leading to a basin different from the starting one; (ii) the number of distinct basins, $`N_{distinct}`$, which are connected on average to each configuration via a $`dw`$ direction. This analysis, based on a computationally demanding procedure, shows that indeed the PES regions visited in thermal equilibrium above and below $`T_c`$ are clearly different. We also study the evolution of the sampled PES as a function of time following a quench from above to below $`T_c`$. We show that, in analogy with the equilibrium case, two qualitatively different dynamical regimes exist during aging, related to different properties of the sampled PES.
The system we study is composed of a binary (80:20) mixture of $`N=1000`$ Lennard Jones atoms. The dynamics of this system is well described by MCT, with a critical temperature $`T_c`$ equal to $`0.435`$.
We begin by considering the equilibrium case. At each $`T`$, for each of the 50 analyzed configurations, we calculate the INM spectrum and — by rebuilding the potential energy profile along straight paths following directions with negative curvature— we classify the unstable modes into $`dw`$ and $`sh`$. It is important to notice that the classification is done by studying the shape of the PES along one eigenvector, i.e. beyond the point in configuration space where the eigenvector was calculated. In principle, to identify a $`dw`$ or a $`sh`$ mode, one should follow a curvilinear path. Indeed the use of straight paths guarantees only the identification of the $`dw`$ modes whose one-dimensional saddle energy is close to the potential energy of the system. As discussed in more length in those $`dw`$ modes are the ones relevant for describing motion in configuration space.
Figure 1 shows the $`T`$ dependence of the unstable and $`dw`$ modes for the studied system. In agreement with the analogous calculation of Ref., even at the lowest temperature where equilibration is feasible within our computer facilities, $`f_{dw}`$ is significantly different from zero. To estimate roughly the number of double wells which do not contribute to diffusion, we have calculated the IS associated with the T=0.446 equilibrium configurations, and we have heated them back to various temperatures below $`T_c`$. The corresponding $`f_u`$ and $`f_{dw}`$ are also shown in Fig.1. Both these quantities depend linearly on the temperature. If we extrapolate $`f_{dw}`$ for the equilibrium and for these non-diffusive cases, we find that the two curves cross close to the MCT critical temperature $`T_c`$. Thus $`T_c`$ seems to be the temperature at which the number of directions leading to different basins goes to zero, leaving only local activated processes as residual channels for structural relaxation.
To estimate in a less ambiguous way the number of different basins which can be accessed from each configuration we apply the following procedure: (i) calculate the $`dw`$ directions via INM calculation; (ii) follow each straight $`dw`$ direction climbing over the potential energy barrier and down on the other side until a new minimum is found; (iii) perform a steepest descent path starting from the new minimum found along the $`dw`$ direction, as first suggested by Gezelter et al.; (iv) save the resulting $`IS`$ configuration; (v) repeat (ii-iv) for each $`dw`$ mode. This procedure produces a list of $`IS`$s which can be reached from the initial configuration crossing a one-dimensional $`dw`$. We note that our procedure only guarantees that the two starting configurations for the quenches are on different sides of the double well, since — as discussed in — there is some arbitrariness in the location of the two minima.
We calculate the relative distance $`d_{ij}`$
$$d_{ij}=\sqrt{\frac{1}{N}\underset{l=1}{\overset{N}{}}(x_l^ix_l^j)^2+(y_l^iy_l^j)^2+(z_l^iz_l^j)^2}$$
(1)
between all IS pairs in the list to determine the number of distinct basins, $`N_{distinct}`$, connected to the starting configuration. Here $`x_m^k`$, $`y_m^k`$, and $`z_m^k`$ are the coordinates of the $`m`$-th particle in the $`k`$-th IS of the list ($`IS_k`$). We also calculate the distance $`d_{0i}`$ — where $`0`$ indicates the IS associated to the starting configuration — to enumerate the number of escape directions, $`N_{escape}`$, leading to a basin different from the original one ($`IS_0`$).
In Fig. 2 (upper panel) we show a plot of the distributions of the distances $`d_{0i}`$ between the starting IS and the minima identified with this procedure for different temperatures. At high temperatures, the distribution has a single peak centered approximately at $`d=0.3`$. For lower temperatures, the peak moves to the left and decreases in height. At the same time, a second but distinct peak, centered around $`d10^4`$, appears. Since our sample is composed of $`1000`$ particles, an average distance of the order of $`10^4`$ means that, also in the case that only one particle has a different position between the two different configurations, this particle has moved less that $`0.003`$ interparticle distance. Thus, we consider the IS $`i`$ and $`j`$ as coincident if $`d_{ij}<10^{2.5}`$.
The upper panel of Fig.3 reports the $`T`$-dependence of the number of $`dw`$ directions $`N_{dw}`$, the number of escape directions and the number of distinct basins which can be reached by following a $`dw`$ direction. We find that at high temperature nearly all $`dw`$ directions leads to a different IS, thus fully contributing to the diffusion process. On lowering the temperature, a large fraction of the directions leads to the same minimum, and, close to $`T_c`$, almost all $`dw`$ directions lead to intra-basin motion. A basin change becomes a rare event. Data in Fig.3 support the view that $`T_c`$ is associated to a change in the PES sampled by the system, and, consequently, in the type of dynamics that the system experiences. Above $`T_c`$ the system can change basins in configuration space by moving freely along an accessible $`dw`$ direction, while below $`T_c`$ the thermally driven exploration of configuration space favors the exploration of the interior of the basins. Diffusion in configuration space, i.e. basin changes, requires an local activated process.
We next turn to the study of out-of-equilibrium dynamics. Configurations are equilibrated at high temperature ($`T=5.0`$) and at time $`t_w=0`$ are brought to low temperature $`T_f=0.2`$ by instantaneously changing the control temperature of the thermostat. Within 100 molecular dynamics steps, the kinetic energy of the system reaches the value corresponding to $`T_f`$. Configurations are recorded for different waiting times $`t_w`$ after the quench. We then repeat the same analysis done for the equilibrium configurations, as described before.
Figure 2 (lower panel) shows the distribution of the distances $`d_{0i}`$ between the minima $`IS_0`$ and $`IS_i`$. For short $`t_w`$ the distribution has a single peak centered approximately at $`d_{0i}=0.3`$, as found in the equilibrium configurations at high temperature. As the system ages, the peak shifts to smaller distances and a second distinct broad peak centered at approximately $`d_{0i}=5\times 10^4`$ appears. The observed behavior is similar to the equilibrium one, on substituting $`T`$ with $`t_w`$.
The lower panel of Fig. 3 shows $`N_{dw}`$, $`N_{escape}`$ and $`N_{distinct}`$ as functions of $`t_w`$. At small $`t_w`$, following a double well direction always brings to a basin that is distinct from the original one, while at long times all $`dw`$ directions leads to intra-basin motion. Again, the behavior is very reminiscent of the equilibrium one, on substituting $`T`$ with $`t_w`$. From these results, we suggest that two qualitatively different dynamical processes control the aging processes, respectively for short and long $`t_w`$. For short waiting times the system is always close to saddle points that allow it to move from one basin to a distinct one. As the system ages, it is confined to wells that are deeper and deeper, and the number of basins that it can visit by following a $`dw`$ direction decays to zero. Thus, for long $`t_w`$ the dynamics of aging proceeds through local activated processes, that allow the system to pass over potential energy barriers. As shown in Fig. 3, beyond $`t_w=10^5`$ the number of distinct basins is almost zero, which, in analogy with the equilibrium results, supports the view that a cross-over from MCT-like dynamics to hopping dynamics may take place at a cross-over time during the aging process.
In summary, this Letter confirms the hypothesis that $`T_c`$ can be considered as the $`T`$ at which dynamics crosses over from the free-exploration to the locally activated exploration of configuration space. In this respect, the analogy between the slowing down of dynamics of liquids and the slowing down of the dynamics in mean field $`p`$-spin model is strengthened. This Letter also shows that changes in dynamical processes associated with changes in the explored local connectivity of configuration space are observed even during the aging process, in the time window accessed by molecular dynamics experiments. MCT-based models of aging could be able to describe the early part of the aging process — i.e. the saddle-dominated dynamics — but may not be adequate for describing the locally activated dynamical region.
We acknowledge financial support from the INFM PAIS 98, PRA 99 and Iniziativa Calcolo Parallelo and from MURST PRIN 98. We thank W. Kob, G. Parisi and T. Keyes for discussions.
|
warning/0007/astro-ph0007394.html
|
ar5iv
|
text
|
# The Galactic disk: study of four low latitude Galactic Fields
## 1 Introduction
During the past years many efforts have been undertaken to gain information about the structure of the Milky Way. However, several unresolved questions still remain. A program has been recently started to study the structure of the Galaxy by means of deep photometry of stellar populations. Several fields in the direction of the Galactic center has been been analyzed with the aim of studying the Galactic structure; deriving the stellar extinction along the line of sight; obtaining information about the age and the metallicity of the stars in various Galactic components (halo, bulge and disc); deriving the past history of star formation and chemical enrichment.
In brief, Ng et al. (1995) studied the color magnitude diagram (CMD) and luminosity functions (LF) of the field #3 of the Palomar Groningen Survey (PG3) located at the periphery of the Galactic Bulge (l=0, b=-10) and set up the basic tool to deconvolve the contribution from many stellar population to highly composite CMDs. Bertelli et al. (1995) analyzed three regions of the Bulge located near the clusters NGC6603 (l=12.9 b=-2.8 ), Lynga 7 (l=328.8 b=-2.8) and Terzan 1 (l=357.6 b=1.0). First they show that the extinction law greatly varies with the direction towards the center. Second, they trace the position of a molecular ring between 3.5 and 4 kpc in the direction of NGC6603, of a stellar ring between 3.0 and 4.0 kpc, of the Sagittarius arm in the direction of NGC6603 (5.0 to 7.0 kpc) and Lynga 7 (4.2 to 7.0 kpc). Third, they recognize the presence near the Galactic Center of an old and metal rich population (see also the detailed study of the CMD of Terzan 1 by Bertelli et al. 1996 in which evidences for the existence of old, high metallicity stars in the so-called hot horizontal branch stage are found), and finally give a hint that the Galactic Bar points its nearest side toward positive galactic longitude.
This paper is devoted to the study of four low latitude fields in the Galactic disk in the direction of the Coalsack-Carina region. Studying the CMDs and luminosity functions it is possible to determine the age and star formation history of the thin disk components, the extinction along the line of sight derive hints about the scale height and length.
In section 2 and in section 3 the observations and the data reduction respectively are described. In section 4 the CMDs are presented. In section 5 the model and the input parameters are given. In section 6 the results are discussed. The presence of a spiral arm at l$``$ 292 and l$``$ 305 is analyzed in section 7. Finally, the conclusions are drawn in section 8.
## 2 Data
For this work, four fields have been observed in Bessel V and Gunn I with the 1.54 m Danish telescope at the European Southern Observatory (ESO), La Silla (Chile) on February $`10^{\mathrm{th}}13^{\mathrm{th}}`$, 1997 (camera C1W11/4 with 2k$`\times `$2k pixels and a nominal field of view of $`13^{}\times 13^{}`$). The coordinates of the fields, the number of averaged exposures, and the total exposure time per filter are given in Table 1.
The fields have been chosen on the large scale surface photometry maps of the Milky Way (Hoffmann et al., 1998 and Kimeswenger et al., 1993) and corresponding term maps which display all points of the Milky Way having the same colour (Schlosser et al., 1995). These maps show large scale structures and are therefore suitable for the selection of interesting areas. The integrated colours of the fields we select show they are affected by low interstellar absorption. Two of the fields point in the direction of the Sgr–Car arm and their integrated UBVR colors are consistent with the presence of a significantly fraction young population. In Fig. 1, the positions of the observed fields are plotted over the relevant part of the Galactic V map.
## 3 Reductions
All of the reductions have been carried out with the image processing system MIDAS. For the stellar photometry, the DAOPHOT and ALLSTAR code have been used. The raw data has been corrected for bias and flat-fields. Due to technical problems with the filter wheel resulting in vignetting, the flat-fields are not only very inhomogeneous but do also vary in time. Therefore some of the data couldn’t be used at all. The dimensions of the unvignetted section of each field are given in table 1.
For the calibration, at least 10 standard stars have been measured at the beginning, the middle and the end of every night at different air-masses. They have been selected from Landolt (1992). From the measured standard stars we derive the following relation between instrumental v, (v-i) and calibrated V , (V-I) for 1 sec of exposure time:
$$\begin{array}{cc}\hfill VI=& 1.034(vi)+0.9824\hfill \\ \hfill V=& v+0.007(VI)1.26\hfill \end{array}$$
(1)
The completeness corrections $`\mathrm{\Lambda }_V`$ and $`\mathrm{\Lambda }_I`$ defined as the ratios of recovered to injected stars in V and I frames respectively are derived by means of the usual artificial stars experiments. They are plotted for each field and filter in Fig. 2. The photometric errors are 0.01,0.05,0.12,0.20 for V=10,21,22,23 respectively and 0.01,0.02,0.05,0.2 for I=10,18,20,21.5 mag.
## 4 Colour-Magnitude Diagrams (CMDs)
In Fig. 3 the resulting V–(V-I)–CMDs are presented for all the fields. In these diagrams the main sequence population is the most prominent feature. In F3 and F4 a red branch parallel to the main sequence is visible. It is produced by He-burning evolved stars distributed along the line of sight with increasing reddening. In the fields F1 and F2 this branch is scarcely populated. The CMDs of F3 and F4 show a small group of stars brighter than V $``$ 15–15.5 on the blue side of the main sequence. In section 7, these stars will be shown to be tracers for the inner spiral arm (I).
## 5 Modeling the Galaxy
The description of the Galaxy is done with the code already described by Bertelli et al. (1995) and revised as described in the following sections. First a synthetic population is generated at varying the parameters age, metallicity range, star formation law and initial mass function. Second, the stars are distributed along the line of side following a model of the Galaxy, where all the components are taken into account. Finally, the photometric completeness of the data is taken into account dividing the simulated CMD in magnitude-colour bins and then subtracting from each bin having N<sub>th</sub> stars, (1-$`\mathrm{\Lambda }`$)N<sub>th</sub>, where $`\mathrm{\Lambda }`$ is the smallest of the V and I completeness factors given in Fig.2.
The generation of the synthetic population makes use of the set of stellar tracks by Girardi et al. (1996) for Z=0.0001, Bertelli et al. (1990) for Z=0.001, Bressan et al. (1993) for Z=0.020, Fagotto et al. (1994a,b,c) for Z=0.0004, 0.004, 0.008, 0.05, 0.10.
### 5.1 The star formation rate
The history of the star formation in the solar neighborhood has been derived using various methods. Several authors suggest that a constant star formation rate can be appropriate for the disk (see among the others Twarog 1980, Haywood et al. 1997). On the basis of the Hipparcos data a star formation rate increasing towards younger ages is derived (Bertelli et al 1999). In the following, several star formation rates going from constant to increasing or decreasing in time are adopted and the corresponding CMDs and luminosity functions are compared with the data (see section 6).
### 5.2 The position of the sun
The position of the sun above the Galactic disk mid-plane is found to range from 10 to 42 pc, the upper limit being obtained by star count method (Stobie & Ishida 1987). On the basis of star-counts in 12 fields in the North and South hemispheres, Humphreys & Larsen (1995) suggest that values as low as 20.5$`\pm `$3.5 pc are more appropriate. Even lower values of 15$``$14 pc are found by Binney et al. (1997), Haywood et al. (1997), Cohen (1995), Hammersley et al. (1995). However, as pointed out by Haywood et al. (1997) the expected offset of the sun deduced from star-counts is found to depend on the scale height of the disk, in the sense that small scale height value favors small offsets: a scale height of 350 pc is compatible with an offset of 20 pc, while a scale height of 200 pc suggests an offset of 15 pc. In the following we adopt an offset of 15 pc.
The distance of the sun from the Galactic center is discussed from 7 to 8.5 Kpc. We adopt 8 Kpc, as a mean value.
### 5.3 Thin disk
#### 5.3.1 Mass distribution
Two kind of mass distributions are usually adopted:
1) a double exponential law of the form:
$$\rho _{disk}=\rho _0e^{r/h_r}e^{(zz_{})/h_z}$$
(2)
where $`h_r`$ and $`h_z`$ are the scale length and scale height of the disk, respectively;
2) an exponential distribution on the plane and sech<sup>2</sup> perpendicularly to the plane:
$$\rho _{disk}=\rho _0e^{r/h_r}sech^2(zz_{}/h_z)$$
(3)
For z $`>>`$ $`h_z`$ the two functions are often considered equivalent. However, in our case, this condition is not always met. In particular, using DIRBE data, Freundenreich (1996) find that the sech<sup>2</sup> distribution is greatly superior to the exponential at low latitudes ($`|b|<5^{}`$), as in our case. Both distributions will be considered and discussed. In our formulation, the constant $`\rho _0`$ can be calculated for every model imposing that the total number of stars in a selected region of the observational CMD is reproduced by the simulations. From $`\rho _0`$ the local mass density distribution will be derived (see following sections).
#### 5.3.2 Scale height and scale length
The scale height $`h_z`$ of the thin disk varies from 325 pc (Gilmore & Reid 1983, Reid & Majewski 1993, among others) to 200 pc (Haywood et al. 1997). In our model, this value is assumed as a free parameter.
The scale length $`h_r`$ is found to vary from 3.0 Kpc (Eaton et al. 1984, Freudenreich 1997) to 2.0 Kpc (Jones et al. 1981). Intermediate values are derived by Ruelas-Mayorga (1991), Robin et al. (1992) who give $`h_r`$=2.5 Kpc.
In our description, $`h_r`$ and $`h_z`$ are assumed to be constant with the Galactic radius. This may not be valid in the outer Galaxy, where the dark matter might dominate the mass. In fact radio observations suggest that the thickness of the HI layer reaches about 400 pc at 13 Kpc, while the H2 layer is considerably flatter, reaching 200 pc at 12.5 Kpc of galactocentric distance (see Combes 1991 for a review). However the angle of maximum displacement above the plane is believed to range between 80<sup>0</sup>(Burton 1988) and 110<sup>0</sup> (Diplas & Savage 1991), measured counterclockwise from the direction l=0. The effect in the fields under discussion is negligible.
#### 5.3.3 The Metallicity
Suggestions have arisen in literature that no age-metallicity relation is present in the thin disk. Edvardson et al. (1993) find a spread of about 0.6 dex in $`\mathrm{\Delta }[M/H]`$ among main sequence stars of similar age in the solar neighborhood: the abundance spread for stars born at roughly the same galactocentric distance is similar in magnitude to the increase in metallicity during the lifetime of the disk. Studies of Galactic open clusters as well as of B stars in the solar neighborhood came to the same conclusion (Friel & Janes 1993, Carraro & Chiosi 1994, Cunha & Lambert 1992). In our simulations, a stochastic age metallicity relation for the disk has been adopted, with Z going from 0.008 to 0.03.
#### 5.3.4 Age components
Up to now the main source of information about the age of this component comes from the open cluster system. If all the open clusters are member of the thin disk, a limiting value might be given by NGC 6791 which is the oldest open cluster with well-determined age. Its age is going from 7 to 10 Gyr (Tripicco et al. 1995). However, Scott et al. (1995) find a somewhat peculiar kinematics for this object. The age of Berkeley 17, which is believed to be one of the oldest disk clusters with 12 Gyr (Phelps 1997) has recently been substantially revised using near-IR photometry to 8-9 Gyr (Carraro et al 1999). A lower limit to the age of the disk is given by the white dwarfs: recent determinations suggest an age of 6-8 Gyr (Ruiz et al. 1996). This value agrees quite well with the oldest age of field population based on Hipparcos data which is 11$`\pm 1`$ (Jimenez et al 1998). While the youngest age $`\tau _f`$ of the thin disk component is constrained by the CMD, the oldest age is assumed to be 10 Gyr. $`\tau _f`$ turns out to be in the range 1–5 $`\times 10^8`$ yr in all the fields.
The velocity dispersion of the disk stars suggest the presence of more age components, having different scale heights (Bessel & Stringfellow 1993). In the following in addition to the old component with large scale height, a second one is taken into account , having scale height $`h_{z1}=100`$ pc and ages ranging from $`\tau _B`$ to the final age $`\tau _f`$. $`\tau _B`$ is assumed to be 2 Gyr. This assumption has been verified on the luminosity functions of the observed fields. When $`\tau _B`$ is 3 Gyr, the luminosity functions are difficultly reproduced unless a quite high percentage of thick disk ($`>`$ 8%) is assumed.
### 5.4 Thick Disc
#### 5.4.1 The age and the metallicity of the thick disk
The mean metallicity is believed to be compatible with the one of the disk globular clusters (-0.6 to -0.7), even if a metal rich tail is expected up to -0.5 together with a metal poor component down to -1.5 dex (Morrison et al. 1990). Gilmore et al. (1995) find a peak of the iron abundance distribution of F/G thick disk stars at -0.7 dex, with no vertical gradient.
Edvardson et al. (1993) estimate an old age for the thick disk component of the solar neighborhood (about 12 Gyr), with no age gap between the thin-thick disk stars. Gilmore et al. (1995) confirm this result. Since the abundance ratios of the thick disk stars reflect incorporation of iron from type I supernovae, these authors suggest a star formation time scale larger than the time scale of SNI which is about 1 Gyr. The kinematics of the thin and thick disks suggest an evolutionary connection between the two components (Wyse & Gilmore 1992, Twarog & Antony-Twarog 1994). It cannot be ruled out that the thick disk is the chemical precursor of the thin disk and is formed through a dissipational collapse after the halo formation and before the end of the thin disk collapse.
However, other scenarios are presented in literature, suggesting that the thick disk formed after the thin disk, either for secular kinematic diffusion of the thin disk stars, or as the result of a violent thin disk heating due to the accretion of a satellite galaxy (Quinn et al. 1993, Robin et al. 1995). At present, it is quite difficult to discriminate between these two models. In the following, we assume the thick disk has an age range from 12 to 8 Gyr, with constant star formation and a metallicity ranging from Z=0.0006 to 0.008.
#### 5.4.2 The mass distribution
As in the case of the thin disk, two density laws are adopted: either double exponential, or exponential in the plane and following sech(z/z) perpendicularly to the plane.
The local density of the thick disk population is poorly known, in spite of the attempt made to measure it. From high proper motion stars, Sandage & Fouts (1987), Casertano et al. (1990) find values ranging from 2% to 10% of the total disk density. Large uncertainties are found also from star-counts, since such estimates are correlated with the determination of the scale height. Small local densities are correlated with large scale heights. The values are going from 1% to 10% of the total density. Robin et al. (1995), Haywood et al. (1997) suggest an intermediate value of 5.6% of the local disk population is due to thick disk. In the following we accept only solutions compatible with a thick disk percentage between 2% and 5%.
#### 5.4.3 The scale height and scale length
Reid & Majewski (1993) summarized the results on the scale height of the thick disk. Gilmore (1984) find a scale height of 1300 pc. Norris (1987), Chen (1996), Gilmore & Reid (1983) found values of 1100 pc, 1170 pc, and 1450 pc respectively. Lower values are suggested by Gould et al. (1995) on the basis of HST counts, by Robin et al. (1995), Haywood et al. (1997) using ground based star counts: 760$`\pm 50`$ pc.
Concerning the scale length of the thick disk Robin et al. (1992), Fux & Martinet (1994), Robin et al. (1995) favor a value of 2.5-2.8 kpc. In our simulations we consider scale height and scale length of the thick disk as free parameters.
## 6 The results
In this Section, the fields will be analyzed separately, deriving the reddening along the line of sight and the possible solutions. For the best fitting solutions the simulated CMDs are displayed in Fig.4. In the simulations of F3 and F4 a young spiral-arm-like population is also included, as described in Section 7. After discussing the single results for each field in the following subsections, they will be compared to derive the best fitting solution for all the fields. Since the studied fields are located at low Galactic latitude, we expect them to be more sensitive to the star formation rate and to the scale height than to the scale lenght of the disk components. For the same reason the parameters of the thick disk component would not be strongly constrained.
### 6.1 Discussing the CMD: Extinction determinations
Ng et al. (1995), Bertelli et al. (1995) proved that the slope of the main sequence in the CMD of the disk population is mainly governed by the extinction along the line of sight. At each magnitude $`V`$ the bluest stars on the main sequence can be interpreted as the envelope of the main sequence turnoffs of the population having absolute magnitude $`M_{tur}`$, shifted towards fainter magnitudes and redder colours by the increasing distance and corresponding extinction. Starting from an initial guess, the amount of extinction at increasing distances is adjusted until a satisfactory agreement between the main sequence blue edge location in the data and in the theoretical simulations is reached. The comparison between data and simulations is made using a $`\chi ^2`$ test. The results are given in Fig. 5.
We point out that F2 and F3 show a relatively modest increase of the interstellar absorption along the line of sight ($`\mathrm{\Delta }`$Ȧ<sub>v</sub> less than 0.5 mag) between 1.5 and 2 kpc distance from the Sun for F2 and between 2 and 5 kpc for F3. Whereas this increment in F2 is probably due to some local density increase of the interstellar matter, the increment in F3 can be explained on larger scales since the direction of F3 at l=292 is crossing the inner spiral arm, if the description of the spiral pattern given by Taylor & Cordes (1993) is adopted. At 1.5 Kpc distance from the Sun, the spiral arm is reached (see section 7), at a height of 40 pc above the Galactic plane. At about 5 Kpc distance, the spiral arm is left at a height of 150 pc above the plane. For this field the increase of extinction is well correlated with the intersection of the spiral arm pattern. In Field F4 (l=305) which does also point towards the inner arm, no special increment of the extinction is noticeable. This is not entirely surprising, since the the fields are chosen on the basis of their integrated colours as having low total absorption (see Section 2).
Finally, we would like to address the question whether the trends in the extinction versus the distance shown in Fig. 5 are real. It is quite difficult to give an estimate of the uncertainty on these determinations. However, the simulations show that by changing the extinction of $`\pm 0.2`$ mag at a given distance d a significant shift in the location of the main sequence edge at magnitudes fainter than $`M_{tur}5+5\times logd+A_V`$ is produced. We can safely assume that the $`A_V`$ determination of Fig. 5 cannot have an internal error larger than 0.2 mag.
We point out that these determinations of extinction are dependent on the adopted age and metallicity range of the population. To estimate the uncertainty due to the combined effect of different age and metallicity distributions, we derive the extinction separately for the three disk components, namely the young thin, the old thin and the thick disk described in Section . The determination of the extinction along the line of sight $`A_V`$ turns out to be different from the adopted values at maximum of 0.2. To further check these results, we make a comparison with the values derived from reddening maps by Mendez & van Altena (1998). Taking into account the errors on Mendez & van Altena determinations (0.23 mag in E<sub>(B-V)</sub>, with A<sub>v</sub>= 3.2 E<sub>(B-V)</sub>) , the agreement is reasonable up to distances of about 8 Kpc, at (l,b)=(292.4,1.63) (field F3), and (305.0,-4.87) (field F4). At (l,b)=(265.54, -3.08) (field F1) and (276.4,-0.54) (field F2) our and Mendez & van Altena (1998) A<sub>v</sub> determinations are consistent up to 3–4 Kpc, our values being higher of a factor $`2`$ at larger distances.
When the maximum reddening in each direction is compared with the maps by Schlegel et al. (1998) derived using DIRBE data, the agreement is excellent, in spite of the fact the authors claim their values should not be trusted for $`|b|<5^{}`$.
### 6.2 Discussing the CMD: the star formation rate
To infer the star formation rate (SFR) of the thin disk component, simulated CMD and LF are calculated with constant, increasing or decreasing rate and then compared with observational CMDs using a $`\chi ^2`$ test. The thick disc is assumed to have a SFR constant from 11 Gyr to 8 Gyr. Due to the low galactic latitude of the observed fields, it is not possible to distinguish between constant or slightly increasing/decreasing rate for this component.
From the analysis of the Hipparcos data Bertelli et al. (1999) derive a disk SFR constant from 10 Gyr to 4.5 Gyr, and then increasing by a factor of 1.5-2 from 4.5 Gyr to 0.1 Gyr. However it is not straightforward that the SFR found in the solar neighborhood is representative of the whole disk, as has already been suggested by Bertelli et al. (1999).
To assess this point, we simulate the CMDs and luminosity functions of the fields using the parameterization of the SFR derived from the Hipparcos data by Bertelli et al. (1999). While the luminosity functions are not inconsistent with this SFR, the CMDs of F3 and F4 cannot be reproduced, since too many young stars brighter than V=19.5 and bluer than the main sequence observational edge are produced. This is evident comparing the simulated and the observed CMDs of F3 and F4 in Fig. 6 and Fig.7. The extinction cannot be responsible of this discrepancy: if the extinction along the line of sight at closer distances is increased to match the observational location of the blue edge of the main sequence at brighter magnitudes, then the faint main sequence turns out to be too red.
The case of F1 and F2 is substantially different. In these fields the blue edge of the main sequence is reproduced using the Hipparcos SFR (see Fig. 8 for F2). However, too many evolved stars are expected in F2 (740 stars brighter than V=22) in comparison with the data (80 stars). No additional information is coming from F1. Due to the poor statistic, the observed and the expected number of stars in this field (63 and 74 respectively) are compatible inside the errors.
However, it cannot be excluded that this result is dependent on the adopted parameterization of the solar neighborhood star formation. To make a further check we use the observed Hipparcos population, and we distribute it along the line of sight in the disk. The resulting CMD is presented in Fig. 9 for F3. The previous conclusions are substantially unchanged. Similar result can be reached in the case of F4, not shown for conciseness.
From this investigation we conclude that the solar neighborhood cannot be considered representative of the properties of the whole disk. An analogous discussion can be made at varying SFR. The simulations show that any assumption of an increasing or even constant SFR yields a too high number of young stars on the blue side of the main sequence. Actually, the most convincing result is obtained with a SFR constant from 10 Gyr to 2 Gyr, then declining by a factor of 10 between 2 Gyr and 0.1 Gyr (see Fig.4).
As a final comment, the CMDs of F3 and F4 show an additional sprinkle of stars brighter that V $``$ 15.5 mag and bluer than the mean location of the main sequence (see Fig. 3). These stars cannot be reproduced unless a young burst of star formation well confined in distance is assumed. This feature will be proved to be consistent with the presence of a spiral arm (see section 7).
### 6.3 The results for F3
Various combinations of the disk parameters, at changing scale height and scale length of the thick disk, scale length of the thin disk. Both spatial distributions (sech<sup>2</sup> and double exponential) for the disk are tested. The observational luminosity functions are compared with the simulations, using a $`\chi `$ square test. Fig.10 presents the comparison of the observational luminosity function with four simulations.
For every set of models the most convincing solutions for decreasing stars formation are listed in Table 2 together with the $`\chi ^2=1/\nu _i(a_ib_i)^2/a_i`$ where $`a_i`$ and $`b_i`$ are the observed and the expected number of stars per magnitude bin, and $`\nu `$ is the number of degrees of freedom (which is equal to the number of bins minus 1, since the in the simulations we impose that the total number of model stars is equal to the observed total number of stars)
As expected, the scale height of the thick disk h<sub>z</sub> is poorly constrained by the data. Any value of h<sub>z</sub> in the range 700–1500 pc, turns out to be acceptable, higher values resulting in a slightly lower percentage of the thick disk component. The most convincing fits are obtained for a thick disk percentage of 2-3%. The sech<sup>2</sup> mass distribution seems to be favored by the data, resulting in more convincing luminosity function fits. Values of thin disk h<sub>z</sub> of 220$`\pm `$30 pc seem to be more consistent with the data.
Imposing a cut of the disk at 14 Kpc, the fit is not substantially improved, as is expected due to the low Galactic latitude.
### 6.4 The results for F4
Table 3 presents the solutions and Fig.11 shows the comparison of the observational luminosity function with four simulations. Analogously to F3, the sech<sup>2</sup> mass distribution is slightly favored, again resulting in higher percentages of the thick disk component. Compared to Field 3, the percentage of thick disk is generally higher in this field: to reproduce the data with constant SFR at least 3–4% of thick disk is needed. The data are not sensitive to the scale lenght and scale height of the thick disk.
Concerning the scale height of the think disk, the most convincing solutions are for 270 $`\pm `$ 30 pc. Scale lenght values as low as 1.1 Kpc result is a less good fit of the luminosity function, while all the values in the range 1.2–1.5 Kpc are consistent with the data.
Analogously to F3, introducing a cut of the disk at 14 Kpc, we do not improve the fit of the luminosity function.
### 6.5 Hunting for common solutions
The scale height of the thin disk derived for F4 are in the mean higher than the ones derived for F3. However they are consistent at the 2-$`\sigma `$ level, if using the best value of 280$`\pm `$30 pc for F4 and 220$`\pm `$30 for F3. The most convincing common solution with decreasing star formation rate is found about 2–4% of thick disk, a scale height around $`h_z=250\pm 60\mathrm{pc}`$ and the sech<sup>2</sup> distribution.
Due to the poor statistics, F1 and F2 (see Table 1) do not set further constraints on the determination of the scale height and lenght of the disk. The errors are relatively large and all the solutions, we found for the fields F3 and F4 do also fit the fields F1 and F2. Hence we can only conclude, that F1 and F2 are consistent within the uncertainties with the other fields.
### 6.6 Discussing the mass distribution
We use the value of the central value of the mass distribution $`\rho _0`$ (see Eqs.2 and 3) to derive the local mass density. $`\rho _0`$ can be derived for each field imposing the total number of stars in a selected region of the CMD. So, we expect that inhomogeneities of the mass distribution reflect in different constants in different fields.
Taking into account all the disk components, there is a slight evidence that the total mass density might be higher of a factor 1.5-2 in F4 and F1 than in F2 and F3, the main difference residing in the mass of the component older than 2 Gyr. However it is not clear whether this effect is real then reflecting inhomogeneities of the disk on small scale, or it must simply be interpreted as due to the uncertainty on the mass determination. From this constant, we can derive the mass density in the local neighborhood in stars more massive than 0.1 M, calculated using Kroupa (2000) IMF. For the best solutions, we derive a total local star density of 0.025 M pc <sup>-3</sup> in F3 and 0.036 in F4, 0.034 in F1 and 0.022 in F2. These results are in agreement with previous determinations of the local density. Oort (1960) find 0.18 M pc <sup>-3</sup> as total local mass density, where the total mass density in stars and interstellar matter would not exceed 0.08 M pc <sup>-3</sup>. Lower values are derived by Creze et al (1998) who estimate that the local mass density in stars can be 0.042-0.048 M pc <sup>-3</sup>, including 0.015 in stellar remnants, while the total dynamical mass density cannot exceed 0.065-0.10 M pc <sup>-3</sup>.
## 7 Discussing the presence of a spiral arm
As already pointed out in the previous sections, the CMDs of fields F3 and F4 (see Fig. 3) show the presence of a few stars brighter than V $`1515.5`$ mag for F4 and F3 respectively and bluer than the average main sequence. Stars in this part of the CMD can be interpreted as tracers of a very young population located in a spiral arm. The idea of a spiral arm is strongly supported by the fact that we find this young population in the fields F3 and F4, which are in the direction of the Sgr-Car arm (see Fig. 1 and 12), whereas the fields F1 and F2, which are not expected to cross the spiral arm, do not show any evidence for such a population.
The spiral arm has been parameterized following Vallee (1995) where however the parameters of the spiral pattern are set to reproduce our data. We adopt an arm/inter-arm density ratio of 2 as found in external galaxies (Rix & Zaritsky 1995) and suggested for our own Galaxy by Efremov (1997). The spiral arm is supposed to have a Gaussian distribution with a $`\sigma `$ of 300 pc as suggested by Taylor & Cordes (1993), and Vallee (1995). We impose that the age of the spiral arm population is younger than 1 $`\times 10^8`$ y. From CMD simulations, we find out that the faintest magnitude of the blue population gives hints about the distance of the spiral arm from us (see Fig 4). We find out that the distance of the spiral arm in the direction of F3 is about 1.3 $`\pm 0.1`$ kpc, with the maximum of the distribution at 2 kpc. In the direction of F4, we find that the spiral arm is at a distance of 1.5 $`\pm 0.2`$ kpc. These results are of course dependent on the adopted parameterization of the spiral arm. They are consistent with the spiral arm pattern defined by the pulsar distribution (Taylor & Cordes (1993)) or derived from optical observations for the local solar environment (Humphreys (1976)).
## 8 Conclusions
Deep CCD photometry in the V and I pass-bands is presented for 4 low latitude Galactic fields. The studied fields are located between l=265 and l=305.
The scale height and the scale lenght of the various populations are derived. The data seem to favor a disk scale height around $`h_z=250\pm 0.60\mathrm{pc}`$ and a scale lenght $`>`$ 1100 pc. The star formation seems to have been decreasing in time, while a constant or increasing star formation rate as the one found using Hipparcos data in the solar neighborhood, cannot reproduce the features of the CMDs: too many blue stars are expected.
From the star-counts in the studied fields we derive the local mass density in stars more massive than 0.1 M, found to be in the range 0.036-0.022 M pc$`{}_{}{}^{}3`$.
In two fields, namely F3 and F4 located at l=292 and l=305 evidence is found of the presence of a very young population. The data can be reproduced imposing the presence of a spiral arm. In these direction the Sgr-Car arm is crossed at a distance of 1.3$`\pm 0.1`$ and 1.5 $`\pm 0.2`$ Kpc respectively. These results are consistent with previously defined spiral arm patterns (Taylor & Cordes (1993); Humphreys (1976)).
Due to the small field of view, two of the studied fields, F1 and F2 do not set strong constrains on the scale height and the scale lenght of the disk. A larger field of view (see e.g. the WFI at the 2.2m ESO telescope having 30$`{}_{}{}^{}\times `$ 30) would allow us to have good statistics down to faint magnitudes. At l=305, b=-4.8 more than 10<sup>4</sup> disk stars down to V=21 are expected in such a large field. In a forthcoming paper we will discuss the Galactic structure towards the center making use of the good statistics of the WFI data.
###### Acknowledgements.
The paper was written while A.V. was visiting scientist of the Graduierten Kollege at the Sternwarte der Universität (Bonn). L.S. was supported by the Deutsche Forschungsgemeinschaft DFG, under the grant N<sup>0</sup> Schm 1444/1-1.
|
warning/0007/hep-ph0007113.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
One of the key issues at the frontier between particle physics and cosmology is the nature of the non-baryonic dark matter that apparently dominates the matter density of the Universe. This is probably dominated by cold dark matter, with a density that probably falls within the range $`0.2<\mathrm{\Omega }_{CDM}<0.5`$ , and may be in the form of massive weakly-interacting particles. It is therefore particularly important to search for such dark matter particles , and one of the most direct strategies is the search for relic particle scattering on nuclei in a laboratory detector . Many experiments around the world are engaged in this search, largely motivated by the cross sections calculated assuming that the cold dark matter is dominated by the lightest neutralino $`\chi `$ of the minimal supersymmetric extension of the Standard Model (MSSM) .
We recently re-evaluated the spin-dependent and spin-independent cross sections for neutralino scattering on protons and neutrons , assuming universality for all of the soft supersymmetry-breaking mass parameters of the MSSM including the Higgs multiplets, incorporating the latest available LEP constraints on the MSSM parameter space, and assuming that the cosmological density of the relic neutralino falls within the range $`0.1<\mathrm{\Omega }_\chi h^2<0.3`$, corresponding to the favoured range of $`\mathrm{\Omega }_{CDM}`$ and a Hubble expansion rate $`0.6h0.8`$ in units $`H_0100\times h`$ km/s/Mpc. We used the latest information from chiral symmetry , low-energy $`\pi p,n`$ scattering and deep-inelastic lepton-nucleon scattering to fix the hadronic matrix elements. Our calculations fell considerably below the present experimental sensitivities , as well as the highest theoretical estimates available in the literature , some of which used less restrictive assumptions. There are, however, some other recent lower estimates: see , for example, which is in good agreement with our previous work .
Shortly after our paper appeared, the DAMA collaboration confirmed their previous evidence for the annual modulation of energy deposits in their scintillation detector, which they interpret as due to the scattering of some cold dark matter particle with mass between about 50 and 100 GeV, and spin-independent cross section on a proton between about $`10^6`$ and $`10^5`$ pb. This cross section range is considerably larger than we found previously , though consistent with the range allowed by some previous cross section estimates. Subsequent to the DAMA paper, the CDMS collaboration has reported negative results from their experiment, establishing an upper limit on the spin-independent cross section that excludes most, but not all, of the range suggested by DAMA.
This unresolved situation motivates us to explore more widely the possible neutralino-proton cross sections in the MSSM, including both the spin-dependent and spin-independent (scalar) contributions. As before, we impose the latest constraints on the MSSM parameter space imposed by the LEP and other experiments , such as measurements of $`bs\gamma `$ decay. It is important to note that the LEP limits we use here have been updated significantly compared to what we used in . Notably, the chargino and particularly the Higgs mass limits we use here are stronger. The latter has a substantial effect at $`\mathrm{tan}\beta =3`$: in addition to the consequent direct reduction in the Higgs-exchange contribution to the scalar cross section, the improved lower limit on the Higgs mass further restricts $`m_0`$ and $`m_{1/2}`$ from below, because of their contribution to $`m_h`$ via radiative corrections. Also, previously we did not use the $`bs\gamma `$ constraint, which we implement here by requiring $`m_A>300`$ GeV for $`\mu <0`$. All of these effects tend to remove some of the higher cross sections that we found previously, particularly at low $`m_\chi `$.
The main thrust of this paper, however, is to relax two of the theoretical assumptions made in our previous work.
$``$ The absence of large flavour-changing neutral interactions suggests that the soft super- symmetry-breaking scalar mass parameters $`m_{0_i}`$ of the MSSM may be universal for different quark and lepton flavours. However, there is no strong phenomenological or theoretical reason why the $`m_{0_i}`$ should be the same for the Higgs multiplets as for squarks and sleptons, and we relax this universality assumption in this work. It is known that, in this case, the lightest neutralino $`\chi `$ might be mainly a Higgsino, but this particular option is greatly restricted by LEP data .
$``$ Neutralinos might not constitute all the cold dark matter, but might be complemented by other particles such as axions or superheavy relics. In this case, $`\mathrm{\Omega }_\chi <\mathrm{\Omega }_{CDM}`$, and $`\mathrm{\Omega }_\chi h^2<0.1`$ becomes a possibility. For any given neutralino mass, $`\mathrm{\Omega }_\chi `$ may be decreased by increasing the $`\chi `$ annihilation cross sections, which is often correlated with an enhanced elastic $`\chi `$-proton scattering cross section. Before concluding that cold dark matter detection becomes easier in this case, however, one must consider what fraction of our galactic halo density $`\rho _{halo}`$ could be composed of neutralinos. Since the process of halo formation is essentially independent of the nature of the cold dark matter, as long as it is non-relativistic and weakly interacting, one should expect that
$$\rho _\chi =\rho _{halo}\times \left(\frac{\mathrm{\Omega }_\chi }{\mathrm{\Omega }_{CDM}}\right),$$
(1)
In an effort to be as optimistic as is reasonable, we assume that $`\rho _\chi =\rho _{halo}`$ if $`\mathrm{\Omega }_\chi h^20.1`$, and rescale: $`\rho _\chi =\rho _{halo}\times (\mathrm{\Omega }_\chi h^2/0.1)`$ if $`\mathrm{\Omega }_\chi h^20.1`$.
In our previous work , in which we assumed universality for the Higgs masses (UHM) at the conventional supersymmetric GUT scale $`10^{16}`$ GeV, and the canonical range $`0.1<\mathrm{\Omega }_\chi h^2<0.3`$, we found that the possible ranges of elastic scattering cross sections were very narrow for any fixed values of $`m_\chi ,\mathrm{tan}\beta `$ and the sign of $`\mu `$, even allowing for plausible uncertainties in the hadronic inputs , and that they were always orders of magnitude below the present sensitivities , even for the smallest allowed values of $`m_\chi 50`$ GeV . Specifically, the maximum value we found for the spin-dependent $`\chi p`$ elastic scattering cross section for $`3\mathrm{tan}\beta 10`$ was well below $`10^3`$ pb, attained for $`m_\chi 60`$ GeV, and the maximum value we found for the spin-independent $`\chi p`$ elastic scattering cross section for $`3\mathrm{tan}\beta 10`$ was $`10^7`$ pb, again attained for $`m_\chi 60`$ GeV. The corresponding experimental sensitivities are $`1`$ pb and $`3\times 10^6`$ pb, respectively. At higher neutralino masses, the predicted cross-sections were significantly smaller still.
In the constrained version of the MSSM, when all soft scalar masses, including the Higgs masses, are set equal at the unification scale (UHM), there are four independent parameters, the soft scalar masses, $`m_0`$, the gaugino masses, $`m_{1/2}`$, the soft trilinear mass terms, $`A`$ (assumed to be universal), and $`\mathrm{tan}\beta `$. In addition, there is the freedom to choose the sign of the Higgs mixing mass $`\mu `$. Previously we scanned the $`m_0m_{1/2}`$ parameter space for fixed $`\mathrm{tan}\beta `$ and $`sgn(\mu )`$. Our results were not very sensitive to $`A`$.
Now that we relax the universal Higgs-mass assumption (nUHM), we find much broader ranges of elastic scattering cross sections for any fixed values of $`m_\chi ,\mathrm{tan}\beta `$ and the sign of $`\mu `$. As previously, we perform a systematic scan of the region of the $`m_0,m_{1/2}`$ parameter space of the MSSM that is consistent with accelerator constraints. Here, $`m_0`$ refers only to a common squark and slepton mass, and the two Higgs soft masses $`m_1`$ and $`m_2`$ are fixed by the conditions of electroweak symmetry breaking, since we allow $`\mu `$ and the Higgs pseudoscalar mass $`m_A`$ to be free parameters. Thus, we scan over $`m_0,m_{1/2},\mu ,m_A`$, and $`A`$ for fixed $`\mathrm{tan}\beta `$. The details of these scans are given below, where we document which parameter choices fail which LEP constraint and/or the cosmological relic density requirement.
We find that the elastic scattering cross sections may be somewhat larger than we found before in the UHM case, particularly for larger $`m_\chi `$. However, the absolute values are still well below the present experimental sensitivities , at least for the canonical range $`0.1<\mathrm{\Omega }_\chi h^2<0.3`$ for the relic neutralino density. This remains true when we consider $`\mathrm{\Omega }_\chi h^2<0.1`$, but rescale the halo density as described above.
We cannot exclude the possibility that there might be some variant of the MSSM that could accommodate the cold dark matter scattering interpretation of the DAMA data, but this would require an extension of the framework discussed here. One possibiliity might be to adopt a larger value of $`\mathrm{tan}\beta `$ : we restrict our attention to $`\mathrm{tan}\beta 10`$ to avoid some uncertainties in the treatment of radiative corrections in the renormalization-group evolution of the MSSM parameters which affect the relic density calculations. Another possibility might be to relax further the universality assumptions for soft supersymmetry-breaking masses, either in the scalar or the gaugino sector. In particular, models in which $`m_{\stackrel{~}{q}}/m_\stackrel{~}{\mathrm{}}`$ is smaller than in the models discussed here might be able to accommodate larger elastic $`\chi `$-proton rates for any given value of $`\mathrm{\Omega }_\chi `$. Another way to reduce $`m_{\stackrel{~}{q}}/m_\stackrel{~}{\mathrm{}}`$, with a similar effect, could be to postulate universality at a lower, intermediate renormalization scale, below the conventional supersymmetric GUT scale .
## 2 Theoretical and Phenomenological Background
We review in this Section relevant aspects of the MSSM . The neutralino LSP is the lowest-mass eigenstate combination of the Bino $`\stackrel{~}{B}`$, Wino $`\stackrel{~}{W}`$ and Higgsinos $`\stackrel{~}{H}_{1,2}`$, whose mass matrix $`N`$ is diagonalized by a matrix $`Z`$: $`diag(m_{\chi _1,..,4})=Z^{}NZ^1`$. The composition of the lightest neutralino may be written as
$$\chi =Z_{\chi 1}\stackrel{~}{B}+Z_{\chi 2}\stackrel{~}{W}+Z_{\chi 3}\stackrel{~}{H_1}+Z_{\chi 4}\stackrel{~}{H_2}$$
(2)
As previously, we neglect CP violation in this paper, so that there are no CP-violating phases in the neutralino mass matrix and mixing. For the effects of CP-violating phases on the neutralino scattering cross-section see -. We assume universality at the supersymmetric GUT scale for the $`U(1)`$ and $`SU(2)`$ gaugino masses: $`M_{1,2}=m_{1/2}`$, so that $`M_1=\frac{5}{3}\mathrm{tan}^2\theta _WM_2`$ at the electroweak scale.
We also assume GUT-scale universality for the soft supersymmetry-breaking scalar masses $`m_0`$ of the squarks and sleptons, but NOT for the Higgs bosons, in contrast to . We further assume GUT-scale universality for the soft supersymmetry-breaking trilinear terms $`A`$. Our treatment of the sfermion mass matrices $`M`$ follows those in , and we refer the interested reader to for further details and notation. It suffices here to recall that, CP being conserved, the sfermion mass-squared matrix for each flavour $`f`$ is diagonalized by a rotation through an angle $`\theta _f`$. We treat as free parameters $`m_{1/2}`$ (we actually use $`M_2`$ which is equal to $`m_{1/2}`$ at the unification scale), the soft supersymmetry-breaking scalar mass scale $`m_0`$ (which in the present context refers only to the universal sfermion masses at the unification scale), $`A`$ and $`\mathrm{tan}\beta `$. In addition, we treat $`\mu `$ and the pseudoscalar Higgs mass $`m_A`$ as independent parameters, and thus the two Higgs soft masses $`m_1`$ and $`m_2`$, are specified by the electroweak vacuum conditions, which we calculate using $`m_t=175`$ GeV <sup>1</sup><sup>1</sup>1We have checked that varying $`m_t`$ by $`\pm 5`$ GeV has a negligible effect on our results..
The MSSM Lagrangian leads to the following low-energy effective four-fermion Lagrangian suitable for describing elastic $`\chi `$-nucleon scattering :
$$=\overline{\chi }\gamma ^\mu \gamma ^5\chi \overline{q_i}\gamma _\mu (\alpha _{1i}+\alpha _{2i}\gamma ^5)q_i+\alpha _{3i}\overline{\chi }\chi \overline{q_i}q_i+\alpha _{4i}\overline{\chi }\gamma ^5\chi \overline{q_i}\gamma ^5q_i+\alpha _{5i}\overline{\chi }\chi \overline{q_i}\gamma ^5q_i+\alpha _{6i}\overline{\chi }\gamma ^5\chi \overline{q_i}q_i$$
(3)
This Lagrangian is to be summed over the quark generations, and the subscript $`i`$ labels up-type quarks ($`i=1`$) and down-type quarks ($`i=2`$). The terms with coefficients $`\alpha _{1i},\alpha _{4i},\alpha _{5i}`$ and $`\alpha _{6i}`$ make contributions to the elastic scattering cross section that are velocity-dependent, and may be neglected for our purposes. In fact, if the CP-violating phases are absent as assumed here, $`\alpha _5=\alpha _6=0`$ . The coefficients relevant for our discussion are:
$`\alpha _{2i}`$ $`=`$ $`{\displaystyle \frac{1}{4(m_{1i}^2m_\chi ^2)}}\left[\left|Y_i\right|^2+\left|X_i\right|^2\right]+{\displaystyle \frac{1}{4(m_{2i}^2m_\chi ^2)}}\left[\left|V_i\right|^2+\left|W_i\right|^2\right]`$ (4)
$`{\displaystyle \frac{g^2}{4m_Z^2\mathrm{cos}^2\theta _W}}\left[\left|Z_{\chi _3}\right|^2\left|Z_{\chi _4}\right|^2\right]{\displaystyle \frac{T_{3i}}{2}}`$
and
$`\alpha _{3i}`$ $`=`$ $`{\displaystyle \frac{1}{2(m_{1i}^2m_\chi ^2)}}Re\left[\left(X_i\right)\left(Y_i\right)^{}\right]{\displaystyle \frac{1}{2(m_{2i}^2m_\chi ^2)}}Re\left[\left(W_i\right)\left(V_i\right)^{}\right]`$ (5)
$`{\displaystyle \frac{gm_{qi}}{4m_WB_i}}[Re\left(\delta _{1i}[gZ_{\chi 2}g^{}Z_{\chi 1}]\right)D_iC_i({\displaystyle \frac{1}{m_{H_1}^2}}+{\displaystyle \frac{1}{m_{H_2}^2}})`$
$`+Re\left(\delta _{2i}[gZ_{\chi 2}g^{}Z_{\chi 1}]\right)({\displaystyle \frac{D_i^2}{m_{H_2}^2}}+{\displaystyle \frac{C_i^2}{m_{H_1}^2}})]`$
where
$`X_i`$ $``$ $`\eta _{11}^{}{\displaystyle \frac{gm_{q_i}Z_{\chi 5i}^{}}{2m_WB_i}}\eta _{12}^{}e_ig^{}Z_{\chi 1}^{}`$
$`Y_i`$ $``$ $`\eta _{11}^{}\left({\displaystyle \frac{y_i}{2}}g^{}Z_{\chi 1}+gT_{3i}Z_{\chi 2}\right)+\eta _{12}^{}{\displaystyle \frac{gm_{q_i}Z_{\chi 5i}}{2m_WB_i}}`$
$`W_i`$ $``$ $`\eta _{21}^{}{\displaystyle \frac{gm_{q_i}Z_{\chi 5i}^{}}{2m_WB_i}}\eta _{22}^{}e_ig^{}Z_{\chi 1}^{}`$
$`V_i`$ $``$ $`\eta _{22}^{}{\displaystyle \frac{gm_{q_i}Z_{\chi 5i}}{2m_WB_i}}+\eta _{21}^{}\left({\displaystyle \frac{y_i}{2}}g^{}Z_{\chi 1}+gT_{3i}Z_{\chi 2}\right)`$ (6)
where $`y_i,T_{3i}`$ denote hypercharge and isospin, and
$`\delta _{1i}=Z_{\chi 3}(Z_{\chi 4})`$ , $`\delta _{2i}=Z_{\chi 4}(Z_{\chi 3}),`$
$`B_i=\mathrm{sin}\beta (\mathrm{cos}\beta )`$ , $`A_i=\mathrm{cos}\beta (\mathrm{sin}\beta ),`$
$`C_i=\mathrm{sin}\alpha (\mathrm{cos}\alpha )`$ , $`D_i=\mathrm{cos}\alpha (\mathrm{sin}\alpha )`$ (7)
for up (down) type quarks. We denote by $`m_{H_2}<m_{H_1}`$ the two scalar Higgs masses, and $`\alpha `$ denotes the Higgs mixing angle <sup>2</sup><sup>2</sup>2We note that (4, 5) is taken from and agree with ..
As discussed in , the elastic cross section for scattering off a nucleus can be decomposed into a scalar (spin-independent) part obtained from the $`\alpha _{2i}`$ term in (3), and a spin-dependent part obtained from the $`\alpha _{3i}`$ term. Each of these can be written in terms of the cross sections for elastic scattering for scattering off individual nucleons. We re-evaluated the relevant matrix elements in . Here we limit ourselves to recalling that:
$``$ There are uncertainties in the scalar part of the cross section associated with the ratios of the light-quark masses, which we take from :
$$\frac{m_u}{m_d}=0.553\pm 0.043,\frac{m_s}{m_d}=18.9\pm 0.8$$
(8)
and information from chiral symmetry applied to baryons. Here the principal uncertainty is associated with the experimental value of the $`\pi `$-nucleon $`\sigma `$ term and the corresponding values of the ratios of the $`B_q<p|\overline{q}q|p>`$. Following , we use
$$z\frac{B_uB_s}{B_dB_s}=1.49$$
(9)
with a negligible experimental error, and
$$y\frac{2B_s}{B_d+B_u}=0.2\pm 0.1,$$
(10)
which yields
$$\frac{B_d}{B_u}=0.73\pm 0.02$$
(11)
The difference between the scalar parts of the cross sections for scattering off protons and neutrons are rather small.
$``$ The spin-dependent part of the elastic $`\chi `$-nucleus cross section can be written in terms of axial-current matrix elements $`\mathrm{\Delta }_i^{(p,n)}`$ that parametrize the quark spin content of the nucleon. We extract from a recent global analysis the values
$$\mathrm{\Delta }_u^{(p)}=0.78\pm 0.04,\mathrm{\Delta }_d^{(p)}=0.48\pm 0.04,\mathrm{\Delta }_s^{(p)}=0.15\pm 0.04$$
(12)
where the errors are essentially 100% correlated for the three quark flavours. In the case of the neutron, we have $`\mathrm{\Delta }_u^{(n)}=\mathrm{\Delta }_d^{(p)},\mathrm{\Delta }_d^{(n)}=\mathrm{\Delta }_u^{(p)}`$, and $`\mathrm{\Delta }_s^{(n)}=\mathrm{\Delta }_s^{(p)}`$.
## 3 Cosmological and Experimental Constraints
As noted in , several convergent measures of cosmological parameters have suggested that the cold dark matter density $`\mathrm{\Omega }_{CDM}=0.3\pm 0.1`$ and that the Hubble expansion rate $`Hh\times 100`$ km/s/Mpc: $`h=0.7\pm 0.1`$, leading to our preferred range $`0.1\mathrm{\Omega }_{CDM}h^20.3`$. The recent data on the spectrum of cosmic microwave background fluctuations from BOOMERANG and MAXIMA are consistent with this range, but do not significantly constrain it further. The upper limit on $`\mathrm{\Omega }_{CDM}`$ can be translated directly into the corresponding upper limit on $`\mathrm{\Omega }_\chi `$. However, it is possible that there is more than one component in the cold dark matter, for example axions and/or superheavy relics as well as the LSP $`\chi `$, opening up the possibility that $`\mathrm{\Omega }_\chi <0.1`$. For a given value of $`m_\chi `$, values of the MSSM parameters which lead to $`\mathrm{\Omega }_\chi <0.1`$ tend to have larger $`\chi `$ annihilation cross sections, and hence larger elastic scattering cross sections. Note, however, that the upper bound, $`\mathrm{\Omega }_\chi h^2<0.3`$, is a firm upper bound relying only on the lower limit to the age of the Universe, $`t_U>1.2\times 10^{10}`$ years (with $`\mathrm{\Omega }_{\mathrm{total}}1`$).
However, in such a ‘shared’ cold dark matter scenario, the packing fraction of neutralinos in the galactic halo must be reduced. As discussed in , for example, dark matter particles are taken into the halo in ‘sheets’ in phase space, whose thicknesses are determined by their initial (thermal) velocity. The ‘sheets’ of cold dark matter particles are of negligible thickness, so the ratios of their densities in the halo are identical with their cosmological densities, and therefore
$$\frac{\rho _\chi }{\rho _{CDM}}=\frac{\mathrm{\Omega }_\chi }{\mathrm{\Omega }_{CDM}}$$
(13)
On the other hand, the ‘sheets’ of hot dark matter particles are of finite thickness related to their thermal velocities at the onset of structure formation, which limits the possible phase-space density of hot dark matter particles, so that $`\rho _{HDM}/\rho _{CDM}<\mathrm{\Omega }_{HDM}/\mathrm{\Omega }_{CDM}`$ in general . Moreover, a large ratio $`\mathrm{\Omega }_{HDM}/\mathrm{\Omega }_{CDM}`$ is currently not expected.
The LSP detection rate also must be reduced correspondingly to (13). Accordingly, when we consider MSSM parameter choices that have $`\mathrm{\Omega }_\chi h^20.1`$, we rescale the calculated scattering rate by a factor $`\mathrm{\Omega }_\chi h^2/0.1`$. This rescaling by the minimal acceptable value of $`\mathrm{\Omega }_{CDM}h^2`$ is relatively optimistic.
For the calculation of the relic LSP density, we have included radiative corrections to the neutralino mass matrix and include all possible annihilation channels . In the MSSM, it is well known that there are large regions of the $`M_2,\mu `$ parameter plane for which the LSP and the next lightest neutralino (NLSP) and/or chargino are nearly degenerate, namely in the Higgsino portion of the plane when $`M_2\mu `$. It was shown that, in these regions, coannihilations between the LSP, NLSP, and charginos are of particular importance in determining the final relic density of LSPs, and these have been included in the present calculation. Inclusion of these coannihilation channels has the important consequence that, in the Higgsino regions where one expects larger elastic scattering cross sections, the relic abundance is substantially reduced. On the other hand, we do not include here coannihilations between the LSP and the sleptons $`\stackrel{~}{\mathrm{}}`$ , in particular the lighter stau $`\stackrel{~}{\tau }_1`$, which were shown to play an important role in models with scalar mass universality also for the Higgs multiplets (UHM). These are known, in particular, to be important for determining the maximum possible generic value of $`m_\chi `$ in the UHM case, but are of less generic importance than $`\chi \chi ^{}\chi ^\pm `$ coannihilations in the non-universal nUHM case considered here. For the same reason, we have also not implemented $`\chi \stackrel{~}{t}`$ coannihilation . The neglect of such $`\chi \stackrel{~}{f}`$ coannihilation processes is generally conservative as far as the elastic scattering rates are concerned, since any reduction they cause in $`\mathrm{\Omega }_\chi h^2`$ is unlikely to be compensated by a corresponding enhancement in the elastic scattering cross section. We also do not pay any particular attention to the narrow parameter slice of mixed gaugino/Higgsino dark matter where $`|\mu |m_{1/2}`$ and $`m_\chi `$ may become large , because this requires an adjustment of parameters at the % level, and is hence not generic. However, these are sampled, with the appropriate weighting, in our general randomized scan of the parameter space.
The lower limit on $`m_\chi `$ depends on the sparticle search limits provided by LEP and other experiments . The most essential of these for our current purposes are those provided by the experimental lower limits on the lighter chargino mass $`m_{\chi ^\pm }`$ and the lighter scalar Higgs mass $`m_{H_2}`$. As discussed in , here we assume a lower limit $`m_{\chi ^\pm }101`$ GeV. The impact of the recently-improved lower limit on the Higgs mass is potentially more significant , particularly for $`\mathrm{tan}\beta =3`$, as displayed in Figs. 6 of . The present experimental lower limit for $`\mathrm{tan}\beta =3`$ approaches $`m_{H_2}>107`$ GeV . In implementing this constraint, we allow a safety margin of $`3`$ GeV in the MSSM calculations of $`m_{H_2}`$ , and hence require the MSSM calculation to yield $`m_{H_2}>104`$ GeV for $`\mathrm{tan}\beta =3`$. In the case of $`\mathrm{tan}\beta =10`$, the LEP constraint on the MSSM Higgs mass is weaker (see Fig. 6 of ), and we require only $`m_{H_2}>86`$ GeV, which includes again a 3 GeV margin of uncertainty. The corresponding limit on $`m_0`$ and $`m_{1/2}`$ in this case may be ignored . The other two constraints that we implement are on sfermion masses, which we require to be (i) larger than 92 GeV, and (ii) larger than that of the lightest neutralino. We recall also the importance of the $`bs\gamma `$ constraint , which we implement in an approximate way, by requiring $`m_A>300`$ GeV for $`\mu <0`$ . As also discussed in , requiring our present electroweak vacuum to be stable against transitions to a lower-energy state in which electromagnetic charge and colour are broken (CCB) would remove a large part of the cosmologically-favoured domain of MSSM parameter space. We have not implemented this optional requirement in the present study. In the next section, we will show the effect of the various expermental constraints on our scan of the parameter plane.
## 4 MSSM Parameter Scan
We have scanned systematically the MSSM parameter space, taking into account the cosmological and experimental constraints enumerated in the previous Section and implementing the MSSM vacuum conditions for the representative choices $`\mathrm{tan}\beta =3`$ and 10. As discussed in , lower values of $`\mathrm{tan}\beta `$ are almost entirely excluded by LEP. Our parameter scan was over the following ranges of parameters:
$`0`$ $`<m_0<`$ $`1000`$ (14)
$`80`$ $`<|\mu |<`$ $`2000,`$ (15)
$`80`$ $`<M_2<`$ $`1000,`$ (16)
$`0`$ $`<m_A<`$ $`1000,`$ (17)
$`1000`$ $`<A<`$ $`1000.`$ (18)
The main scan, which covers $`m_0,\mu `$ and $`M_2>100`$ GeV and $`m_A>300`$ GeV, was supplemented with smaller but significant subscans, to cover the smaller values of these four parameters as described below. The values of $`m_0`$ we use are fixed at the unification scale $`10^{16}`$ GeV, while the values of the remaining parameters, $`\mu ,M_2,m_A,`$ and $`A`$ are evaluated at the electroweak scale. The lower cut off on both $`M_2`$ and $`\mu `$ is due to the lower limit on the chargino mass. As we indicated above, we impose a lower limit $`m_A>300`$ GeV for $`\mu <0`$ to avoid problems with $`bs\gamma `$. However, it should be noted that this restriction is quite conservative as, even for $`m_A=350`$ GeV, there are regions included in the above scan which are not allowed by $`bs\gamma `$ . Similarly, even for $`\mu >0`$, where we impose no cut off on $`m_A`$, we have incuded some points which should be excluded on the basis of $`bs\gamma `$.
As can be seen in the Table, the overall scan was divided into three (four) specific regions for each value of $`\mathrm{tan}\beta `$ and $`\mu `$ negative (positive), each with the number of points listed. The subscans with lower thresholds were designed to scour carefully the regions of MSSM parameter space close to the LEP exclusions, with the aim of ensuring that we sampled points close to their boundaries. For each subscan, we show the number of points which survive all the LEP experimental constraints discussed above, and we see that lower fractions of the low-threshold subscans survive them, in particular because they tend to yield excluded values of the chargino mass. Fig. 1 provides some insight into the impacts of the different LEP constraints for the case $`\mathrm{tan}\beta =3`$ and $`\mu >0`$. We plot in Fig. 1 the points scanned in the $`M_2\mu `$ parameter plane. In making this scatter plot, we show a randomly chosen subset of 5000 of the 90000 points sampled <sup>3</sup><sup>3</sup>3We have checked that there is no qualitative difference between this plot and the much denser plot with all points shown., since it is much easier to pick out the relevant physical effects of the cuts in such a subset of points, the full plot being extremely dense.
We see that the chargino cut removes points at low values of $`\mu `$ and $`M_2`$, denoted by (green) pluses, that the Higgs cut then removes many more points with low $`M_2`$, denoted by (red) crosses, that the sfermion cut removes still more points with low $`M_2`$, denoted by (violet) triangles (this occurs at high $`A`$ and/or $`\mu `$ when there is a sizeable off-diagonal component in the sfermion mass matrix), and that the LSP cut tends to remove points at higher $`M_2`$ denoted by (golden) diamonds. The surviving (blue) squares are spread over the $`\mu ,M_2`$ plane, except for small values. Note that some points may fail to survive more than one of the above cuts. These are only denoted by the first cut tested and failed in the order listed above. The scans for the opposite sign of $`\mu `$ and for $`\mathrm{tan}\beta =10`$ exhibit similar features, and are omitted here. The only noticeable difference when $`\mu >0`$ is that not so many were points eliminated by the Higgs cut at large values of $`M_2`$ for $`\mu <0`$, because we imposed the limit $`m_A>300`$ GeV: for $`\mu >0`$, many more points were run with low $`m_A`$ and hence low $`m_{H_2}`$. Also, for $`\mathrm{tan}\beta =10`$, more points survive at low $`\mu `$ and/or $`M_2`$ due to the relaxed contraint on the Higgs mass.
The last column of the Table shows how many of the points that survive the LEP constraints and have relic densities in the cosmologically preferred range $`0.1\mathrm{\Omega }_\chi h^20.3`$. It is apparent that most of the preferred points emerge from the first scan with $`M_2,\mu >100`$, as the lower values which were explored thoroughly in the second subscan generally failed the chargino cut. More details of the scan over cosmological relic densities for $`\mathrm{tan}\beta =3`$ and $`\mu >0`$ are shown in Fig. 2. As in Fig. 1, we show only a randomly selected subset of 5000 points out of the total of approximately 22000 points which survived the LEP cuts. We see that, among the points that survived the previous LEP constraints, those with a small ratio of $`\mu /M_2`$ generally have too small a relic density, denoted by (green) pluses, as a result of over-efficient $`\chi \chi ^{}\chi ^\pm `$ coannihilation, whereas points with $`\mu /M_21`$ to 5 tend to have too large a relic density, denoted by (red) crosses, particularly if $`\mu `$ and $`M_2`$ are individually large. The points with a relic density in the preferred range $`0.1\mathrm{\Omega }_\chi h^20.3`$, denoted by (blue) squares, tend to accumulate around $`\mu /M_21/2`$ or low $`M_2`$. The former points are in the transition region between over-efficient $`\chi \chi ^{}\chi ^\pm `$ coannihilation and under-efficient annihilation at large $`\mu `$ and $`M_2`$, whereas the latter are in the region of low $`M_2`$ where careful implementation of the LEP constraints is essential. However, it is apparent from Fig. 2 that there are exceptions to these general trends. We do not discuss them in detail, but remark that we have made an attempt to understand at least those exceptions that lead to ‘unusual’ elastic scattering cross sections.
## 5 Elastic Scattering Cross Sections
We now discuss the values of the elastic scattering cross sections that are attainable, bearing in mind the LEP and cosmological relic density constraints. Fig. 3 illustrates the allowed ranges of elastic scattering cross sections for the points included in our scan for the particular case $`\mathrm{tan}\beta =3,\mu >0`$, as it was described in the previous Section. Plotted is a subset of 3000 of the 90000 points scanned, indicating which points survive all the LEP cuts, and which other points fail which LEP cut. We find similar results for $`\mathrm{tan}\beta =10`$ and/or the opposite sign of $`\mu `$, with the exception that when $`\mu <0`$ we find some points trickling down below the apparent boundary at $`10^{10}`$ pb in Fig. 3(b), because of cancellations similar to those discussed in .
We note, in particular, that the LEP chargino and Higgs cuts remove many points with low $`m_\chi `$ and/or large elastic scattering cross sections. The sfermion mass cut is less important. The constraint that $`\chi `$ be the LSP removes quite a large number of points, populated more or less evenly in these cross section plots. The somewhat sparse set of points with very small cross sections give some measure of how low the cross section may fall in some special cases. These reflect instances where particular cancellations take place, examples of which were discussed in , and should not be regarded as generic. The lower boundary of the densely occupied region in Fig. 3 offers an answer to the question how low the elastic scattering cross sections may reasonably fall, roughly $`\sigma 10^9`$ pb for the spin-dependent cross section and $`10^{10}`$ pb for the spin-independent cross section.
We would like to draw particular attention to the spin-independent cross-section shown in Fig. 3(b). Notice that there are parameter choices with very large scattering cross sections. In this random selection, the cross-section may be as high as a few $`\times 10^4`$ pb, and could even be larger than that claimed by DAMA. Indeed, in the full set of 90000 points scanned, there are even a few points which surpass $`10^3`$ pb. However, all of these points have been excluded by LEP (primarily by the Higgs mass cut). The largest surviving cross section is slightly over $`10^7`$ pb, in both the randomly selected subset and the full scan. For $`\mu <0`$, the upper boundary in the scalar cross section is about an order of magnitude lower, as was the case in the model with universal Higgs masses . Note also that, for $`\mu <0`$, the limit $`m_A>300`$ GeV we impose removes the points with large cross sections (in this case with $`\sigma _{\mathrm{scalar}}>10^8`$ pb).
The next step is to implement the cosmological relic density constraints. We show in Fig. 4 the cross sections obtained for a representative subsample of points with $`\mathrm{tan}\beta =3,\mu >0`$ that survive the LEP cuts, sorted according to the calculated values of $`\mathrm{\Omega }_\chi h^2`$. Spin-dependent cross sections are plotted in panels (a) and (c), and spin-independent cross sections are plotted in panels (b) and (d). We include in panels (a) and (b) the cross sections calculated for unrealistic models with $`\mathrm{\Omega }_\chi h^2>0.3`$, and without making any rescaling correction for points with $`\mathrm{\Omega }_\chi h^2<0.1`$. The over-dense points with $`\mathrm{\Omega }_\chi h^2>0.3`$, denoted by (red) crosses, have been removed in panels (c) and (d), and the cross sections for under-dense points with $`\mathrm{\Omega }_\chi h^2<0.1`$, denoted by (green) pluses, have been rescaled by the appropriate halo density fraction (1). As could be expected, the over-dense points tend to have smaller cross sections, and the under-dense points larger cross sections before applying the rescaling correction. After rescaling, the under-dense points yield cross sections in the range found for the favoured points with $`0.1\mathrm{\Omega }_\chi h^20.3`$, denoted by (blue) boxes. For $`\mathrm{tan}\beta =10`$ and $`\mu >0`$, the scalar cross section is about an order of magnitude higher for points which survive all cuts. Relative to the cases with $`\mu >0`$, the $`\mu <0`$ cases have a scalar cross section which is 1-2 orders of magnitude smaller.
A comparison with Fig. 2 shows that the largest cross sections displayed in Figs. 4(a,b), are almost all for Higgsino-like states whose elastic cross section is mediated by $`Z`$ exchange. These are cosmologically under-dense, due to a combination of large annihilation and coannihilation cross sections. The cosmologically over-dense regions with relatively low elastic cross sections are mainly for gaugino-like states, and are for the most part more massive than 300 GeV, which is the oft-quoted upper bound on the bino mass in the MSSM .
Our resulting predictions for the spin-dependent elastic neutralino-proton cross section for $`\mathrm{tan}\beta =3`$ and $`\mu >0`$, after taking into account the LEP and cosmological constraints, are shown in Fig. 5(a), where a comparison with the UHM case is also made <sup>4</sup><sup>4</sup>4In contrast to , here we have taken into account the updated LEP constraints.. The raggedness of the upper and lower boundaries of the dark (blue) shaded allowed region reflect the coarseness of our parameter scan, and the relatively low density of parameter choices that yield cross sections close to these boundaries. We see that, at low $`m_\chi `$ close to the LEP limit, the spin-dependent cross section may be as much as an order of magnitude greater than in the UHM case considered previously , shown by the concave (red and turquoise) strip. However, even for low $`m_\chi `$, the attainable range is far below the present experimental sensitivity, which is to $`\sigma _{spin}1`$ pb, and could be many orders of magnitude lower. As $`m_\chi `$ increases, the maximum allowed value of $`\sigma _{spin}`$ decreases, though not as rapidly as in the previous UHM case . The hadronic uncertainties are basically negligible for this spin-dependent cross section, as seen from the light (yellow) shading. Turning now to the option $`\mathrm{tan}\beta =3`$ and $`\mu <0`$ shown in Fig. 5(b), we see that the allowed range of the spin-dependent cross section is similar to that in the $`\mathrm{tan}\beta =3,\mu >0`$ option. This is in contrast to the situation in the UHM , where the spin-dependent cross section at low $`m_\chi `$ is much smaller for $`\mu <0`$ than for $`\mu >0`$. However, the cross section is still three or more orders of magnitude away from the present experimental upper limit. In the option $`\mathrm{tan}\beta =10`$ and $`\mu >0`$ shown in Fig. 5(c), we see that the attainable range of the spin-dependent cross section is again similar to the previous option. This again contrasts with the UHM case, where the narrow allowed band for large $`m_\chi 500`$ GeV was somewhat higher than for the option $`\mathrm{tan}\beta =3`$ and $`\mu >0`$. As shown in Fig. 5(d), our results for $`\mathrm{tan}\beta =10`$ and $`\mu <0`$ are very similar to those for $`\mu >0`$.
The analogous results for the spin-independent elastic neutralino-proton cross section, after taking into account the LEP and cosmological constraints, are shown in Fig. 6, where comparisons with the UHM case are also made.
We see in panel Fig. 6(a) for $`\mathrm{tan}\beta =3`$ and $`\mu >0`$ a pattern that is similar to the spin-dependent case. For small $`m_\chi `$, the spin-independent scalar cross section, shown by the dark (blue) shaded region, may be somewhat higher than in the UHM case, shown by the (red and turquoise) diagonal strip, whilst it could be much smaller. For large $`m_\chi `$, the cross section may be rather larger than in the UHM case, but it is always far below the present sensitivity. The case shown in panel (b) of $`\mathrm{tan}\beta =3`$ and $`\mu <0`$ is somewhat different: the cross section never gets to be significantly larger than the UHM value at small $`m_\chi `$. The reason for the anomalous extension of the UHM band outside the more general range is that the newer analysis reflected in the (blue and yellow) shaded region incorporates updated LEP constraints , that are significantly stronger for small $`\mathrm{tan}\beta `$ and small $`m_\chi `$ than those used in . This ‘anomaly’ is absent in panel (c) for $`\mathrm{tan}\beta =10`$ and $`\mu <0`$, which closely resembles panel (a), and also panel (d) for $`\mathrm{tan}\beta =10`$ and $`\mu >0`$. We note in panel (d) a lesser reappearance of the ‘anomalous’ outdated UHM region at small $`m_\chi `$. The dip in the (red and turquoise) UHM band for $`m_\chi 230`$ GeV in panel (d) reflects rather special cancellations that are absent in the more general case. Overall, we note that the hadronic uncertainties, denoted by the light (yellow) bands, are somewhat larger in the spin-independent case than in the spin-dependent case.
## 6 Summary and Prospects
In this paper we have extended the analysis of to consider a more general sampling of supersymmetric models, relaxing the UHM assumption we made previously. For each of two choices of $`\mathrm{tan}\beta `$ and $`\mu `$ negative (positive), we have sampled 70000 (90000) sets of MSSM parameters, 30000 in general scans and 20000 each in two (three) special subscans over lower values of $`M_2,\mu ,m_0`$ (and $`m_A`$). We have implemented the current LEP constraints on MSSM parameters , discussing in detail which scan points survive which of these constraints. We have further discussed which of the remaining scan points yield a cosmological relic density in the allowed range $`\mathrm{\Omega }_\chi h^20.3`$, and which of these are in the preferred range $`\mathrm{\Omega }_\chi h^20.1`$. We exclude from further consideration the over-dense points with $`\mathrm{\Omega }_\chi h^2>0.3`$, and rescale the predicted cross sections for under-dense points with $`\mathrm{\Omega }_\chi h^2<0.1`$ as in (1).
The cross sections we predict for spin-dependent and spin-independent elastic neutralino-proton scattering for different values of $`\mathrm{tan}\beta `$ and the sign of $`\mu `$ are shown in Figs. 5 and 6, respectively. We provide in Fig. 7 a compilation of our results, compared with the present experimental upper limits on the cross sections and the detection of spin-independent scattering reported by the DAMA Collaboration .
We see that our predicted cross sections are well below the experimental upper limits for both the spin-dependent and -independent cases. We are unable to find MSSM parameter sets consistent with our relaxed universality assumptions that come close to explaining the DAMA measurements. Our assumptions would need to be questioned if the neutralino scattering interpretation of the DAMA data is confirmed. For example, we have restricted our attention to models with $`\mathrm{tan}\beta 10`$. Alternatively, the DAMA data might favour models with values of $`m_{\stackrel{~}{q}}/m_\stackrel{~}{\mathrm{}}`$, obtained either by relaxing the input universality assumption, or by imposing it at some renormalization scale below the conventional supersymmetric GUT scale .
In the future, we plan to improve the available relic density calculations by extending them to larger $`\mathrm{tan}\beta `$ and incorporating consistently all coannihilation processes. On the experimental side, we expect that other Collaborations will soon be able to confirm or exclude definitively the DAMA interpretation of their annual modulation signal as being due to neutralino scattering. Looking further ahead, we interpret our results as indicating a high priority for a new generation of direct dark matter detection experiments with a much higher sensitivity.
Acknowledgments
We thank Toby Falk and Gerardo Ganis for many related discussions. The work of K.A.O. was supported in part by DOE grant DE–FG02–94ER–40823.
|
warning/0007/cond-mat0007264.html
|
ar5iv
|
text
|
# The Evolution of Quasiparticle Charge in the Fractional Quantum Hall Regime
\[
## Abstract
The charge of quasiparticles in a fractional quantum Hall (FQH) liquid, tunneling through a partly reflecting constriction with transmission $`t`$, was determined via shot noise measurements. In the $`\nu =1/3`$ FQH state, a charge smoothly evolving from $`e^{}=e/3`$ for $`t_{1/3}1`$ to $`e^{}=e`$ for $`t_{1/3}1`$ was determined, agreeing with chiral Luttinger liquid theory. In the $`\nu =2/5`$ FQH state the quasiparticle charge evolves smoothly from $`e^{}=e/5`$ at $`t_{2/5}1`$ to a maximum charge less than $`e^{}=e/3`$ at $`t_{2/5}1`$. Thus it appears that quasiparticles with an approximate charge $`e/5`$ pass a barrier they see as almost opaque.
\]
The fractional quantum Hall (FQH) effect is a manifestation of the prominent and unique effects resulting from the Coulomb interactions between electrons in a two-dimensional electron gas (2DEG) under the influence of a strong magnetic field . In this regime the lowest Landau level is partially populated. Laughlin’s seminal explanation of the FQH effect involved the emergence of intriguing fractionally charged quasiparticles. Recently, shot noise measurements confirmed the existence of such quasiparticles with charge $`e/3`$ and $`e/5`$ at filling factors $`\nu =1/3`$ and $`\nu =2/5`$ , respectively. These experiments relied on the fact that shot noise, resulting from the granular nature of the quasiparticles, is proportional to their charge. Since current flowing in an ideal Hall state is noiseless a quantum point contact (QPC) constriction was used to weakly reflect the incoming current, leading to partitioning of the incoming carriers and hence to shot noise. A charge $`e^{}`$ was then deduced from the shot noise expression derived for non-interacting particles . In this paper, we extend the range of QPC reflection to the strong back-scattering limit, where the apparent noise-producing quasiparticle charge is expected to be different. Specifically, an opaque barrier is expected to allow only the tunneling of electrons, as both sides of the barrier should be quantized in units of the electronic charge. How this charge evolves is an important question in the understanding of the behavior of quasiparticles, and here we explore the evolution of the charge of the $`e/3`$ and $`e/5`$ quasiparticles. We first briefly describe the expected dependence of shot noise on charge and transmission.
At zero temperature $`(T=0)`$, the shot noise contribution of the $`p`$’th channel is :
$$S_{T=0}=2e^{}Vg_pt_p(1t_p),$$
(1)
where $`S`$ is the low frequency $`(f<<eV/h)`$ spectral density of current fluctuations $`(S\mathrm{\Delta }f=i^2)`$, $`V`$ the applied source-drain voltage, $`g_p`$ the conductance of the fully transmitted $`p`$’th channel in the QPC, and $`t_p`$ is its transmission coefficient. This reduces to the well known classical Poissonian expression for shot noise when $`t_p1`$ (the ’Schottky equation’), $`S_{T=0}=2eI`$, with $`I=Vg_pt_p`$ the DC current in the QPC.
The justification for the use of Eq. (1) comes from current theoretical studies of shot noise in the FQH regime, based on the chiral Luttinger liquid model. They are applicable only for Laughlin’s fractional states, $`\nu =1/3`$, $`1/5`$, etc. (where the edge is composed of one channel only) and not for more general filling factors. They predict the following:
$`S_{T=0}`$ $`=`$ $`2e^{}Vg_p(1t_p)=2e^{}I_r;t_p1,`$ (2)
$`S_{T=0}`$ $`=`$ $`2eVg_pt_p=2eI_t;t_p0,`$ (3)
where $`I_r`$ and $`I_t`$ are the reflected and transmitted DC currents, respectively. The most important result of Eq. (2) is that the tunneling of quasiparticles with charge $`e/3`$, $`e/5`$, etc. in Laughlin states, at weak reflection $`(t_p1)`$, changes to that of electrons at strong reflection $`(t_p0)`$.
One can gain insight into the characteristics of the expected shot noise in the FQH regime , and some insight into Eq. (1), by considering the Composite Fermion (CF) model . In the simplest approximation for the CF model the fractionally filled electronic Landau level with $`\nu =p/(2p+1)`$ is identified as $`p`$ filled Landau levels of CFs, $`\nu _{CF}=p`$, with each CF consisting of an electron with two attached magnetic flux quanta $`\varphi _0=h/e`$. The effective magnetic field sensed by the CFs is $`B2n_sh/e`$, with $`n_s`$ the density of the 2DEG. Under this weaker effective magnetic field the CFs are approximated as weakly interacting quasiparticles, flowing in separate and non-interacting edge channels, hence justifying the application of the above-mentioned formulae for the noise. When the QPC constriction is reduced in width and the conductance is in a transition between two different FQH plateaus of the series $`p/(2p+1)`$ only one edge channel is partitioned. The others can be approximated as being perfectly transmitted. Consequently, in Eqs. (1) and (2), $`p`$ designates the CF edge channel that is being partitioned. As examples, for the transition between $`\nu =1/3`$ and the insulator: $`p=1`$; $`g_1=g_0/3`$ and $`t_1=3g/g_0`$; while for the transition between $`\nu =2/5`$ and $`\nu =1/3`$: $`p=2`$; $`g_2=(2/51/3)g_0`$ and $`t_2=\frac{g/g_01/3}{2/51/3}`$, with $`g`$ being the total conductance and $`g_0=e^2/h`$ the quantum conductance. The dependence of ithe charge on transmission, in the simplest model, can be evaluated by considering the added current due to the two flux quanta attached to the electron. Doing this, de Picciotto predicted the quasiparticle charge to vary from $`e^{}=e/(2p+1)`$ at $`t_p1`$ to $`e^{}=e/(2p1)`$ at $`t_p0`$ as a linear function of $`t_p`$, namely, for $`p=1`$ $`e/3e`$ and for $`p=2`$ $`e/5e/3`$.
In order to apply the above principles in a realistic experiment a more general expression for the shot noise applicable at finite temperatures, has to be used :
$$S_T=2e^{}Vg_pt_p(1t_p)\left[\mathrm{coth}\left(\frac{e^{}V}{2k_BT}\right)\frac{2k_BT}{e^{}V}\right]+4k_BTg.$$
(4)
This equation leads to a finite noise at zero applied voltage, $`S=4k_BTg`$ \- the Johnson-Nyquist formula. When $`V>V_T2k_BT/e^{}`$ the noise approaches the linear behavior predicted by Eqs. (1) and (2).
Measuring quasiparticle charge in the strong back-scattering limit is difficult, and results so far were inconclusive . As the QPC constriction is closed to reflect a larger portion of the incident current, the conductance exhibits the familiar impurity resonances as a function of constriction width (, and see also in Fig. 1). Moreover, the $`IV`$ characteristic becomes highly nonlinear ($`g`$ and $`t`$ depend on current), making the analysis difficult. Measuring a large number of samples across the full range of the transmission coefficient in the first two CF channels, $`\nu =1/3`$ and $`\nu =2/5`$, we found relatively resonant-free samples. Moreover, we extended Eq. (4) to cases of nonlinear $`IV`$ characteristics allowing also the charge to change with the transmission coefficient. Consequently, we have found a universal behavior of the charge as a function of transmission in the $`\nu =1/3`$ channel, and qualitatively quite different behavior for the charge in $`\nu =2/5`$ channel.
Our samples were 2DEG’s embedded in GaAs-AlGaAs heterostructures with a low-temperature concentration of $`9.8\times 10^{10}cm^2`$ and a mobility of $`4\times 10^6cm^2/Vs`$. A perpendicular magnetic field of $`12.15T`$ is needed to reach the center of the $`\nu =1/3`$ plateau. The left-hand insert in Fig. 1 shows the schematic of the two-terminal Hall samples with source (S), drain (D) and a QPC. The Hall sample’s width was $`100\mu m`$ and the QPC opening width was $`300nm`$. The QPC gate’s potential was used to control the partitioning of the incoming current. Measurements were made in a dilution refrigerator at a lattice temperature of $`55mK`$ and a measured electron temperature of $`85mK`$ (see for details). Noise was measured within a bandwidth of $`30kHz`$ around a frequency of $`1.6MHz`$, chosen to be above the $`1/f`$-noise knee and much lower than $`eV/h`$. An LRC circuit determined the central frequency and bandwidth, with $`R`$ dominated by the resistance of the QPC and $`C`$ by the capacitance of the coaxial lines. A cold preamplifier, with a current noise of $`3\times 10^{29}A^2/Hz`$, amplified the noise signal.
We present here results from four samples $`(\mathrm{\#}1\mathrm{\#}4)`$: three measured in the $`\nu =1/3`$ FQH state and two in the $`\nu =2/5`$ FQH state. The bare samples (without applied gate voltage) exhibit, as a function of magnetic field, an accurate $`\nu =1/3`$ quantization of the resistance but deviate at the $`\nu =2/5`$ plateau due to finite bulk longitudinal resistance. The measurements in the $`\nu =2/5`$ state were conducted at two different bulk filling factors: $`\nu _{bulk}=2/5`$ and $`\nu _{bulk}=1/2`$ (see sample $`\mathrm{\#}1`$ in Fig. 1), while for the measurements in the $`\nu =1/3`$ state the bulk filling factors were $`\nu _{bulk}=1/3`$ and $`\nu _{bulk}=1/2`$ (see sample $`\mathrm{\#}4`$ in Fig. 1). Typical problems are seen in Fig. 1: sample $`\mathrm{\#}1`$ shows a single large ’resonance’ - the large spike on the left-hand side of the graph - which prohibits further measurement into the $`1/3`$ state; and the reduction of the transmission of the $`1/3`$ state in sample $`\mathrm{\#}4`$, although much smoother, saturates at about $`0.1e^2/h`$, presumably due to leakage across the QPC. The open circles on the graphs show where noise and $`IV`$ measurements were made.
In our experiment we measured two quantities: the differential conductance $`g`$ and the shot noise. Using $`ge^{}t`$ and $`S_T`$ from Eq. (4) we extracted the transmission probability $`t`$ and the quasiparticle’s charge $`e^{}`$. However, the analysis is complicated by the strong dependence of the conductance on the current - see the right-hand insert in Fig. 1. This insert shows the differential conductance of the QPC as a function of DC current for three different conductances indicated by points A, B, and C. While at point A, where $`t`$ is relatively large, the conductance is almost constant with current $`(\mathrm{\Delta }g/g_{I=0}=0.05)`$, at point C, where $`t`$ is very small, there is a significant change in the differential conductance at large currents $`(\mathrm{\Delta }g/g_{I=0}=0.3)`$. To account for this non-linearity, the energy independent Eq. (4) was modified by resorting to the integral over energy used in its derivation . However, the dependence of conductance on the current (in a small range), for a fixed QPC width, was all attributed to a changing $`t`$, i.e., the charge $`e^{}`$ was approximated not to vary with current. Transforming from the integration over energy to a sum over discreet current points, and substituting $`t`$ in terms of $`g`$ and $`e^{}`$ in Eq. (4), $`t_{p=1}=\frac{(g_i/g_0)}{e^{}/e}`$, we get for $`\nu =1/3`$:
$`S_T(I)=2e^{}I{\displaystyle \frac{1}{N}}{\displaystyle \underset{i=1}{\overset{N}{}}}(1{\displaystyle \frac{g_i/g_0}{e^{}/e}})[\mathrm{coth}\left({\displaystyle \frac{e^{}V}{2k_BT}}\right)`$ (6)
$`{\displaystyle \frac{2k_BT}{e^{}V}}]+4k_BTg.`$
Here $`i`$ runs over the measured points $`(N)`$ up to current $`I`$ and $`g_i`$ is the differential conductance at each point. In the $`\nu =2/5`$ state we substitute for the total current $`I_T`$ only that fraction which flows through the 2nd edge channel (using the CF model), $`I_{p=2}=\frac{(g/g_0)1/3}{g/g_0}I_T`$, and for the transmission $`t_{p=2}=\frac{(g/g_0)1/3}{(2/51/3)5e^{}/e}`$. Indeed, if $`e^{}=e/5`$, $`t_{p=2}`$ is the expected bare transmission of the 2nd CF channel given above. The noise expression now contains a single fitting parameter $`e^{}`$.
Figure 2 shows noise results for a partitioned $`\nu =1/3`$ channel in sample $`\mathrm{\#}4`$. There is no noise on the $`\nu =1/3`$ plateau. The top part of the graph shows the differential conductance of the QPC against DC current $`I`$ at points A, B, and C shown in Fig. 1. The current range we used for the extraction of the charge is $`\mathrm{\Delta }g/g_{I=0}=00.2`$ in order to reduce the effect of the charge variation with current while still being able to fit the curves to Eq. (6). The measured noise, with the background thermal noise subtracted, is shown in the lower part of Fig. 2. The curves are offset for clarity. Also shown is the behavior of Eq. (6) with $`e^{}=e/3`$ (solid lines) and $`e^{}=e`$ (dashed lines). For each width of the QPC constriction we find the best fitting quasiparticle charge $`e^{}`$ and consequently the channel transmission $`t`$ near $`I=0`$. In previously published high-$`t`$ data the noise is that of $`e/3`$ charges . As the transmission is reduced the apparent charge increases to a maximum around charge $`e`$. Consistent results were obtained for the two other samples (as seen in Fig. 4). Similarly, Fig. 3 shows similar graphs for the measurements in the $`\nu =2/5`$ state in sample $`\mathrm{\#}1`$ (points A’, B’, and C’ in Fig. 1). Again, no noise is measured on the $`\nu =2/5`$ plateau. The theoretical lines correspond to charges $`e^{}=e/5`$ (solid lines) and $`e^{}=e/3`$ (dashed lines). The other sample provided similar results.
The dependences of the quasiparticle charge on transmission coefficient for all four samples are summarized in Fig. 4. All results approximately collapse onto two separate curves. While in the $`\nu =1/3`$ case the deduced charge changes smoothly from $`e/3`$ at weak reflection (large $`t`$) to around $`e`$ at strong reflection ($`t0.1`$), the deduced charge in the $`\nu =2/5`$ case stays near $`e/5`$ over almost the full range of transmission. There is an apparent slight increase of $`e^{}`$ at lower transmissions. Although scattering of the data due to the small signal prevents a more accurate determination of the charge for $`t<0.3`$, it clearly does not show the steep rise to $`e^{}=e`$ observed at $`\nu =1/3`$.
Adopting the CF picture in accordance with Ref. 11, the difference between the two channels can be understood by considering how much charge crosses the constriction when a composite fermion, composed of an electron and two flux quanta, traverses it. In the $`\nu =1/3`$ case, a strongly closed constriction, reflecting almost all the incident current, is almost an insulator and the extra charge induced by the fluxes is negligible, leading to a quasiparticle charge approximately $`e`$. In contrast, in the $`\nu =2/5`$ case only one of the edge channels is strongly reflected, and consequently the constriction is not an insulator. Thus the extra transferred charge is finite and the quasiparticle’s charge is not $`e`$. Eqs. 1-4 are based on a picture in which the noise is produced by independent quasiparticles whose partitioning obeys binomial statistics. In fact the noise can be interpreted also as being generated by quasiparticles of fixed charge whose partitioning statistics are not binomial. For example, the measured charge of $`e^{}=e`$ could be interpreted as a quasiparticle of charge $`e`$ (a single electron) or as three quasiparticles of charge $`e^{}=e/3`$ bunched together. For the $`\nu =2/5`$ channel, we may conclude that the $`e^{}=e/5`$ quasiparticles traverse an opaque barrier without fully bunching, which would produce a charge $`e^{}=e`$. As yet, there is no rigorous theory for the $`\nu =2/5`$ case.
We thank M. Reznikov and R. de Picciotto for useful discussions and guidance. We also thank F. von Oppen for instructive discussions. The work was partly supported by the Israeli Academy of Science, the Israel-USA Binational Science Foundation and the Israel-Germany DIP grant.
|
warning/0007/physics0007043.html
|
ar5iv
|
text
|
# The use of GPS-arrays in detecting shock-acoustic waves generated during rocket launchings
## 1 Introduction
A large number of publications, including a variety of thorough reviews (Karlov et al., 1980; Mendillo, 1981; 1982; Nagorsky, 1998; 1999) are devoted to the study of the ionospheric response of the shock wave produced during launchings of powerful rockets, among them launch vehicles (LV). Scientific interest in this problem is due to the fact that such launchings can be regarded as active experiments in the Earth’s atmosphere, and they can be used in solving a wide variety of problems in the physics of the ionosphere and radio wave propagation.
These investigations have also important practical implications since they furnish a means of substantiating reliable signal indications of technogenic effects (among them, unauthorized), which is necessary for the construction of an effective global radiophysical subsystem for detection and localization of these effects. Essentially, existing global systems of such a purpose use different processing techniques for infrasound and seismic signals. However, in connection with the expansion of the geography and of the types of technogenic impact on the environment, very challenging problems heretofore have been those of improve the sensitivity of detection and the reliability of measured parameters of the sources of impacts, based also on using independent measurements of the entire spectrum of signals generated during such effects.
To solve the above problems requires reliable information about fundamental parameters of the ionospheric response of the shock wave, such as the amplitude and the form, the period, the phase and group velocity of the wavetrain, as well as angular characteristics of the wave vector. Note that for the designating the ionospheric response of the shock wave, the literature uses terminology incorporating a different physical interpretation, among them the term ‘shock-acoustic wave’ (SAW) — Nagorsky (1998). For convenience, in this paper we shall use this term despite the fact that it does not reflect essentially the physical nature of the phenomenon.
Published data on fundamental parameters of the SAW differ greatly. According to a review by Karlov et al. (1980), the oscillation period of the ionospheric response of the SAW, recorded during launchings of the Apollo mission rockets, varied from 6 to 90 min, and the propagation velocity was in the range from 600 to 1670 m/s. The added complication is also that following the first SAW response arriving at the detection point, a ‘second’ wave propagating with the sound velocity at the ground level is often recorded (Karlov et al., 1980; Calais and Minster, 1996). A ‘third’ wave with large periods and subsonic propagation velocities is occasionally also detected (Nagorsky, 1998; 1999). This wave is identified as an exciting Acoustic-Gravity Wave (AGW) during a rocket launching.
There is an even greater uncertainty in the localization of the region of SAW generation, which is necessary both for an understanding of the physical mechanisms of the phenomenon, and for a scientific justification of the solution of applied problems. Published data are discussed in more detail in Section 5, by comparison with our obtained results.
The lack of comprehensive, reliable data on SAW parameters is due primarily to the limitations of existing experimental methods and detection facilities. The main body of data was obtained by measuring the frequency Doppler shift at vertical and oblique-incidence ionospheric soundings in the HF range (Jacobson and Carlos, 1994; Nagorsky, 1998). In some instances the sensitivity of this method is sufficient to detect the SAW reliably; however, difficulties emerge when localizing the region where the detected signal is generated. These problems are caused by the multiple-hop character of HF signal propagation. This gives no way of deriving reliable information about the phase and group velocities of SAW propagation, estimating angular characteristics of the wave vector, and, further still, of localizing the SAW source.
Using the method of transionospheric sounding with VHF radio signals from geostationary satellites, a number of experimental data on SAW parameters were obtained in measurements of the Faraday rotation of the plane of signal polarization which is proportional to a total electron content (TEC) along the line connecting the satellite-borne transmitter with the receiver (Mendillo, 1981; 1982; Li et al., 1994). A serious limitation of methods based on analyzing signals from geostationary satellites is that their number is too small and ever decreasing with time, and the distribution in longitude is non-uniform; this is especially true in regard to the eastern hemisphere. Published data from incoherent scatter (IS) radars are also scarce; as far as our knowledge goes, there is only one publication (Noble, 1990) reporting the detection of long-period (gravity) waves over the Arecibo IS station, following the launching of the Space Shuttle LV on October 18, 1993.
A common limitation of the above-mentioned methods when determining the SAW phase velocity is the necessity of knowing the time of a rocket launching since this velocity is inferred from the SAW delay with respect to the launch time assuming that the velocity is constant along the propagation path, which is quite contrary to fact.
For determining the above-mentioned reasonably complete set of SAW parameters, it is necessary to have appropriate spatial and temporal resolution which cannot be ensured by existing, very sparse, networks of ionozondes, oblique-incidence radio sounding paths, and IS radars. Furthermore, the creation of these facilities involves developing dedicated equipment, including powerful radio transmitters which contaminate the radio medium. As a result, the above systems cannot serve as an efficient basis for a global radiophysical subsystem for detection and localization of technogenic effects which must provide a continuous, global monitoring.
The advent and evolution of a Global Positioning System, GPS, and also the creation on its basis of widely branched networks of GPS stations (at least 600 sites at the end of 1999, the data from which are placed on the INTERNET) opened up a new era in remote ionospheric sensing. In the immediate future this network will undergo a substantial expansion by integration with the GLONASS positioning system (Klobuchar, 1997). In addition, researchers have also access to a standard software which permits them to unify a preprocessing of the data from multichannel two-frequency GPS receivers. In this context, a major effort must go into experimental and theoretical explorations of the possibilities of detecting technogenic perturbations of the ionosphere, as well as into the development of appropriate algorithms and software for a reprocessing of the data from GPS networks.
Currently some authors have embarked on an intense development of methods for detecting the ionospheric response of strong earthquakes (Calais and Minster, 1995), rocket launchings (Calais and Minster, 1996), and industrial surface explosions (Fitzgerald, 1997; Calais et al., 1998a; Calais et al., 1998b). In the cited references the SAW phase velocity was determined by the ‘crossing’ method by estimating the time delay of SAW arrival at subionospheric points corresponding to different GPS satellites observed at a given time. However, the accuracy of such a method is rather low because the altitude at which the subionospheric points are specified, is determined in a crude way.
The goal of this paper is to develop a method for determining SAW parameters (including the phase velocity, angular characteristics of the SAW wave vector, the direction towards the source, and the source location) using GPS-arrays whose elements can be chosen out of a large set of GPS stations of a global IGS network. Contrary to existing radio techniques, this method estimates SAW parameters without a priori information about the location and time of a rocket launching.
The emphasis in this paper is on the ionospheric response of the SAW which is recorded as a ‘first’ wave, although our developed methods make it also possible to analyze other wave manifestations of rocket launchings. This will be the subject of future work.
Section 2 presents a description of the geometry of experiments and general information about the LV launchings from the Baikonur and Kennedy Space Center (KSC) cosmodromes which are analyzed in this study. A brief account of the proposed method is presented in Section 3. Measurements of SAW parameters at different GPS-arrays during rocket launchings are discussed in Section 4. Section 5 discusses experimental results and compares them with the data reported by other authors. It is beyond the scope of this paper to consider the physical mechanisms of SAW generation and propagation.
## 2 The geometry and general characterization of experiments
This paper presents the results derived from determining the key SAW parameters during ascents of the Proton and Space Shuttle launch vehicles (LV) from cosmodromes Baikonur (45.6$`{}_{}{}^{}N`$, 63.3$`{}_{}{}^{}E`$) and KSC (28.5$`{}_{}{}^{}N`$, 279.3$`{}_{}{}^{}E`$) in 1998 and 1999 (a total of five launchings).
In spite of the large number of GPS stations, the selection of GPS-arrays for detecting SAW generated during rocket launchings is made difficult by the fact that the LV flight paths for both cosmodromes pass either over the Atlantic ocean (Space Shuttle LV) or over sparsely populated regions of Kazakhstan (Proton LV). Nevertheless, it was possible to utilize a sufficient number of GPS stations in order for this method to be implemented. Table 1 presents the geographic coordinates of the GPS stations used as GPS-array elements.
Fig. 1 portrays the geometry of experiments during Proton LV launchings from the Baikonur cosmodrome. The dash line roughly corresponds to the horizontal projection of the rocket flight trajectory with orbital inclination $`\psi =51.6^{}`$ (Nagorsky, 1999). The symbol $``$ marks the pad site. Bold dots mark the position of GPS stations, and upper-case lettering corresponds to their names. The scaling of the coordinate axes is selected from considerations of an approximate equality of the linear dimensions along the latitude and longitude.
Below is a summary of results derived from detecting one standard launching (on November 20, 1998) and two abortive launchings (July 5 and October 27, 1999) of the Proton LV. The following information about rocket launchings from the Baikonur Cosmodrome as extracted from the INTERNET sites: http//www.flatoday.com; http://www.spacelaunchnews.com; and http://www.ksc.nasa.gov. Official data on abortive Proton LV launches on July 5 and October 27, 1999 were taken from the English version of the press release issued by the Khrunichev Center which is available on: http://www.isllaunch.com.
A Proton launch vehicle failed during launch on July 5, 1999 (Day 186). Initial analyses indicated a malfunction of the Proton’s second stage. Liftoff took place at 13:32 UT from a launch pad of the Baikonur Cosmodrome. First stage flight was normal, with second stage separation and ignition on time. Telemetry first reported anomalous data at approximately 280 seconds into the flight, with deviations from the planned trajectory appearing on ground tracking systems by 330 seconds. By 390 seconds the LV was 14 km below its planned trajectory. Ground systems tracked the upper stages and payload to impact. The Proton’s third stage, with the upper stage and its Raduga spacecraft payload, fell in the Karaganda region of Kazakhstan.
The Proton LV was launched from a launch pad at Baikonur Cosmodrome on October 27 (Day 300), 1999 at 16:16 UT. Based on preliminary data, failure of one of the main second stage engines occurred 222 seconds after liftoff, followed by failure of all engines of the stage. The second stage together with the upper stage and satellite fell back to earth in the Karaganda region of Kazakhstan.
In a typical Proton launch, the vehicle’s six first-stage engines ignite 1.6 s before liftoff. Stage two ignition occurs approximately two minutes into flight, four seconds prior to jettison of the first stage. Stage three vernier engine ignition occurs at 330 seconds, with separation of the second and third stages taking place 3.5 s later. The stage three main engine ignition occurs 2.5 s after separation.
Unfortunately, we were unable to obtain more detailed information about the LV launch schedule on the above days. However, for a further analysis we need only know that these failures occurred after the second stage was in operation longer than 200 s and 100 s for the July 5 and October 27, 1999 launches, respectively.
General information regarding these launches, as well as the scheduled launch of November 20, 1998 (Day 324) is summarized in Table 2 (including the start time $`t_0`$ of UT, the start time of LT for one of the GPS-array sites, the orbital inclination $`\psi `$ to the equatorial plane, and the level of geomagnetic disturbance according to $`Dst`$-variation data). It was found that the deviation of $`Dst`$ for the selected days was quite moderate thus enabling the SAW to be identified.
Fig. 2 presents the experimental geometry during Space Shuttle LV launchings from the KSC Cosmodrome (the designations are the same as in Fig. 1). The Space Shuttle Discovery spacecraft (STS-90) and Space Shuttle Columbia (STS-95) were launched on April 17 (Day 107) and on October 29 (Day 302), respectively. General information about these launches is also summarized in Table 2. The horizontal projection of the rocket trajectory for the October 29 launching is constructed using the INTERNET data. The level of geomagnetic disturbance as deduced from the $`Dst`$-variation data was also found to be quite moderate for these launches, which made it possible to reliably identify the SAW.
## 3 Methods of determining shock-acoustic wave characteristics using GPS-arrays
The standard GPS technology provides a means for wave disturbances detecion based on phase measurements of TEC at each of spaced two-frequency GPS receivers. A method of reconstructing TEC variations from measurements of the ionosphere-induced additional increment of the group and phase delay of the satellite radio signal was detailed and validated in a series of publications (Calais and Minster, 1995, 1996; Fitzgerald, 1997). We reproduce here only the final formula for phase measurements
$$I=\frac{1}{40.308}\frac{f_1^2f_2^2}{f_1^2f_2^2}[(L_1\lambda _1L_2\lambda _2)+const+nL]$$
(1)
where $`L_1\lambda _1`$ and $`L_2\lambda _2`$ are additional paths of the radio signal caused by the phase delay in the ionosphere, (m); $`L_1`$ and $`L_2`$ represent the number of phase rotations at the frequencies $`f_1`$ and $`f_2`$; $`\lambda _1`$ and $`\lambda _2`$ stand for the corresponding wavelengths, (m); $`const`$ is the unknown initial phase ambiguity, (m); and $`nL`$ are errors in determining the phase path, (m).
Phase measurements in the GPS can be made with a high degree of accuracy corresponding to the error of TEC determination of at least $`10^{14}`$ <sup>-2</sup> when averaged on a 30-second time interval, with some uncertainty of the initial value of TEC, however (Hofmann-Wellenhof et al., 1992). This makes possible detecting ionization irregularities and wave processes in the ionosphere over a wide range of amplitudes (up to $`10^4`$ of the diurnal TEC variation) and periods (from 24 hours to 5 min). The unit of TEC, which is equal to $`10^{16}`$ <sup>-2</sup> ($`TECU`$) and is commonly accepted in the literature, will be used in the following.
Fig. 3a gives a schematic representation of the transionospheric sounding geometry. The axes $`z`$, $`y`$, and $`x`$ are directed, respectively, zenithward, northward ($`N`$) and eastward ($`E`$). P — point of intersection of Line-of-Sight (LOS) to the satellite with the maximum of the ionospheric $`F2`$-region; S — subionospheric point; $`\alpha _s`$ —the azimuthal angle, counted off from the northward in a clockwise direction; and $`\theta _s`$ —the angle of elevation between the direction $`r`$ along LOS and the terrestrial surface at the reception site.
In some instances a convenient way of detecting and determining the ionospheric response delay of the shock wave involves inferring the frequency Doppler shift $`F`$ from TEC series obtained by formula (1). Such an approach is also useful in comparing TEC response characteristics from the GPS data with those obtained by analyzing VHF signals from geostationary satellites, as well as in detecting the shock wave in the HF range. To an approximation sufficient for the purpose of our investigation, a corresponding relationship was obtained by K. Davies (1969)
$$F=13.5\times 10^8I_t^{}/f$$
(2)
where $`I_t^{}`$ stands for the time derivative of TEC. Relevant results derived from analyzing the $`F(t)`$-variations calculated for the ‘reduced’ frequency of 136 MHz are discussed in Section 4.
The correspondence of space-time phase characteristics, obtained through transionospheric soundings, with local characteristics of disturbances in the ionosphere was considered in detail in a wide variety of publications (Bertel et al., 1976; Afraimovich et al., 1992; Mercier and Jacobson, 1997) and is not analyzed at length in this study. The most important conclusion of the cited references is the fact that, as for the extensively exploited model of a ‘plane phase screen’ disturbances $`\mathrm{\Delta }I(x,y,t)`$ of TEC faithfully copy the horizontal part of the corresponding disturbance $`\mathrm{\Delta }N(x,y,z,t)`$ of local concentration, independently of the angular position of the source, and can be used in experiments on measuring the wave disturbances of TEC.
However, the TEC response amplitude experiences a strong azimuthal dependence caused by the integral character of a transionospheric sounding. As a first approximation, the transionospheric sounding method is responsive only to TIDs with the wave vector $`K_t`$ perpendicular to the direction $`r`$. A corresponding condition for elevation $`\theta `$ and azimuth $`\alpha `$ of an arbitrary wave vector $`K_t`$ normal to the direction $`r`$, has the form
$$\theta =\mathrm{arctan}(\mathrm{cos}(\alpha _s\alpha )/\mathrm{tan}\theta _s)$$
(3)
We used formula (3) in determining the elevation $`\theta `$ of $`K_t`$ from the known mean value of azimuth $`\alpha `$ by Afraimovich et al. (1998) – see Sections 3.2 and 4.
### 3.1 Detection and determination of the horizontal phase velocity $`V_h`$ and the direction $`\alpha `$ of the SAW phase front along the ground by GPS-arrays
In the simplest form, space-time variations of the TEC $`\mathrm{\Delta }I(t,x,y)`$ in the ionosphere, at each given time $`t`$ can be represented in terms of the phase interference pattern that moves without a change in its shape (the solitary, plane travelling wave)
$$\mathrm{\Delta }I(t,x,y)=\delta \mathrm{sin}(\mathrm{\Omega }tK_xxK_yy+\phi _0)$$
(4)
where $`\delta `$, $`K_x`$, $`K_y`$, $`\mathrm{\Omega }`$ — are the amplitude, the x- and y-projections of the wave vector $`𝑲`$, and the angular frequency of the disturbance, respectively; $`T=2\pi /\mathrm{\Omega }`$, $`\mathrm{\Lambda }=2\pi /|K|`$ is its period and wavelength; and $`\phi _0`$ is the initial phase of the disturbance. The vector $`K`$ is a horizontal projection of the full vector $`K_t`$ (Fig. 3a).
At this point, it is assumed that in the case of small spatial and temporal increments (the distances between GPS-array sites are less than the typical spatial scale of TEC variation, and the time interval between counts is less than the corresponding time scale), the influence of second derivatives can be neglected. The following choices of GPS-arrays all meet these requirements.
For the SAW which have mostly the character of an $`N`$-wave (Li et al., 1994; Calais and Minster, 1995, 1996; Fitzgerald, 1997; Calais et al., 1998), the amplitude $`\delta `$ in (4) can be specified as (with an appropriate choice of the model parameters)
$$\delta (t)=\mathrm{exp}[(\frac{tt_{max}}{t_d})^2]$$
(5)
where $`t_{max}`$ is the time when the disturbance has a maximum amplitude, and $`t_d`$ is the half-thickness of the ‘wave packet’. We employed this model in simulating the SAW detected during Proton LV launch on July 5, 1999 (at the left of Fig. 4, and at the right of Fig. 5).
We now summarize briefly the sequence of data processing procedures. Out of a large number of GPS stations, three sites (A,B,C) are selected, the distances between which do not exceed about one-half the expected wavelength $`\mathrm{\Lambda }`$ of the perturbation. Site B is taken to be the center of a topocentric reference frame whose axis $`x`$ is directed eastward, and the axis $`y`$ is directed northward. The receivers in this frame of reference have the coordinates ($`x_A`$, $`y_A`$), (0,0), ($`x_C`$, $`y_C`$). Parallel lines in Fig. 3b are lines of equal TEC; the arrow indicates the direction $`\alpha `$ of a normal to these lines, that is, the direction of the wave vector $`K`$. Such a configuration of the GPS receivers represents the GPS-array with a minimum number of the required elements. In regions with a dense network of GPS sites, we can obtain a large variety of GPS-arrays of a different configuration enabling the acquired data to be checked for reliability; in this paper we have exploited such a possibility.
The input data include series of the ‘oblique’ value of TEC $`I_A(t)`$, $`I_B(t)`$, $`I_(t)`$, as well as corresponding series of values of the elevation $`\theta _s(t)`$ and the azimuth $`\alpha _s(t)`$ of the beam to the satellite calculated using our developed CONVTEC software program which converts the RINEX-files, standard for the GPS system, from the INTERNET (Gurtner, 1993). For determining SAW characteristics, continuous series of measurements of $`I_A(t)`$, $`I_B(t)`$, $`I_C(t)`$ are selected with a length of at least a one-hour interval which includes the start time.
To eliminate variations of the regular ionosphere, as well as trends introduced by orbital motion of the satellite, a procedure is used to remove the linear trend involving a preliminary smoothing of the initial series with the selected time window. This procedure is best suited to the detection of the signal such as a single pulse ($`N`$-wave) when compared to the frequently used band-pass filter (Li et al., 1994; Calais and Minster, 1995, 1996; Fitzgerald, 1997; Calais et al., 1998). A limitation of the band-pass filter is the delay and the oscillatory character of the response which gives no way of reconstructing the form of the $`N`$-wave.
Series of the values of the elevation $`\theta _s(t)`$ and azimuth $`\alpha _s(t)`$ of the beam to the satellite are used to determine the location of the subionospheric point S (see Fig. 3a), as well as to calculate the elevation $`\theta `$ of the wave vector $`K_t`$ of the disturbance from the known azimuth $`\alpha `$ (see formula (3) and Section 4.2).
The most reliable results from the determination of SAW parameters correspond to high values of elevations $`\theta _s(t)`$ of the beam to the satellite because sphericity effects become reasonably small. In addition, there is no need to recalculate the ‘oblique’ value of TEC $`\mathrm{\Delta }I(t)`$ to the ‘vertical’ value. In this paper, all results were obtained for elevations $`\theta _s(t)`$ larger than 30.
Since the distance between GPS-array elements (from several tens of kilometers to a few thousand kilometers) is much smaller than that to the GPS satellite (over 20000 km), the array geometry at the height $`h_{max}`$ is identical to that on the ground.
Fig. 4a shows typical time dependencies of an ‘oblique’ TEC $`\mathrm{\Delta }I(t)`$ at the GPS-array CHUM station near the Baikonur Cosmodrome for the start day, July 5, 1999 (heavy curve), one day before and after the start (thin lines). For the same days; panel b shows TEC variations $`\mathrm{\Delta }I(t)`$ with the removed linear trend and a smoothing with the 5-min time window. Variations in frequency Doppler shift $`F(t)`$; ‘reduced’ to the sounding signal frequency of 136 MHz, for three sites of the array (SELE, CHUM, SHAS) for the launch day, July 5, 1999, are presented in panel c. Day numbers, GPS station names and GPS PRN satellite numbers are indicated in all panels. The arrows at the abscissa axis indicate the start time $`t_0`$.
It is evident from Fig. 4 that fast $`N`$-shaped oscillations with a typical period $`T`$ of about 300 s are clearly distinguished among slow TEC variations. The oscillation amplitude (up to 0.5 $`TECU`$) is far in excess of the TEC fluctuation intensity during ‘background’ days. Variations in frequency Doppler shift $`F(t)`$ for spatially separated sites (SELE, CHUM, SHAS) are well correlated but are shifted relative to each other by an amount well below the period, which permits the SAW propagation velocity to be unambiguously determined. A time resolution of 30 s used in our study, which is standard for the GPS data, is not quite sufficient for determining small shifts of such signals with an adequate accuracy for different sites of the array (Fig. 4c). Therefore, we used a parabolic approximation of the $`F(t)`$-oscillations in the neighborhood of minimum $`F(t)`$, which is quite acceptable when the signal/noise ratio is high. In addition, we used different combinations for three sites of the possible number of GPS-array sites in the area of Baikonur and KSC Cosmodromes. Relevant results for all combinations are presented in Table 3.
With proper account of a good signal/noise ratio (larger than 1), we determine the horizontal projection of the phase velocity $`V_h`$ with the known coordinates of array sites A, B, C from time $`t_p`$ shifts of a maximum deviation of the frequency Doppler shift $`F(t)`$. Preliminarily measured shifts are subjected to a linear transformation with the purpose of calculating shifts for sites spaced relative to the central site northward $`N`$ and eastward $`E`$. This is followed by a calculation of the $`E`$\- and $`N`$-components of $`V_x`$ and $`V_y`$, as well as the direction $`\alpha `$ in the range of angles 0–360 and the modulus $`V_h`$ of the horizontal component of the SAW phase velocity
$$\begin{array}{cc}\hfill \alpha & =\mathrm{arctan}(V_y/V_x)\hfill \\ \hfill V_h& =|V_xV_y|(V_x^2+V_y^2)^{1/2}\hfill \end{array}$$
(6)
where $`V_y`$, $`V_x`$ are the velocities with which the phase front crosses the axes $`x`$ and $`y`$. The orientation $`\alpha `$ of the wave vector $`K`$, which is coincident with the propagation azimuth of the SAW phase front, is calculated unambiguously in the range 0–360 subject to the condition that arctan$`(V_y/V_x)`$ is calculated having regard to the sign of the numerator and denominator.
The above method for determining the SAW phase velocity neglects the correction for orbital motion of the satellite because the estimates of $`V_h`$ obtained below exceed an order of magnitude as a minimum the velocity of the subionospheric point at the height $`h_{max}`$ for elevations $`\theta _s>30^{}`$ (Afraimovich et al., 1998).
From the delay $`\mathrm{\Delta }t=t_pt_0`$ and the known path length between the launch pad and the subionospheric point we calculated also the SAW mean velocity $`V_a`$ in order to compare our obtained estimates of the SAW phase velocity with the usually used method of measuring this quantity.
### 3.2 Determination of the elevation of the wave vector $`\theta `$ and the velocity modulus $`V_t`$ of the shock wave
Afraimovich et al. (1992) showed that for an exponential ionization distribution the TEC disturbance amplitude ($`M`$) is determined by the aspect angle $`\gamma `$ between the vectors $`K_t`$ and $`r`$ (see Fig. 3a), as well as by the ratio of the wavelength of the disturbance $`\mathrm{\Lambda }`$ to the half-thickness of the ionization maximum $`h_d`$
$$M\mathrm{exp}\left(\frac{\pi ^2h_d^2\mathrm{cos}^2\gamma }{\mathrm{\Lambda }^2\mathrm{cos}^2\theta _s}\right)$$
(7)
where $`\theta _s`$ is the elevation of the wave vector $`r`$.
In the case under consideration (see below), for the phase velocity of order 1 km/s and for the period of about 200 s, the wavelength $`\mathrm{\Lambda }`$ is comparable with the half-thickness of the ionization maximum $`h_d`$. When the elevations $`\theta _s`$ are 30, 45, 60 , the ‘beam-width’ $`M(\gamma )`$ at 0.5 level is, respectively, 25, 22 and 15. If $`h_d`$ is twice as large as the wavelength, then the beam tapers to 14, 10 and 8, respectively.
The beam-width is sufficiently small that the aspect condition (3) restricts the number of beam trajectories to the satellite, for which it is possible to detect reliably the SAW response in the presence of noise (near the angles $`\gamma =90^{}`$ ). On the other hand, formula (LABEL:LAUNCH-fig-3) can be used to determine the elevation $`\theta `$ of the wave vector $`K_t`$ of the shock wave at the known value of the azimuth $`\alpha `$ (Afraimovich et al., 1998). Hence the phase velocity modulus $`V_t`$ can be defined as
$$V_t=V_h\times sec(\theta )$$
(8)
The above values of the width $`M(\gamma )`$ determine the error of calculation of the elevations $`\theta `$ (of order 20 to the above conditions (Section 3.3).
### 3.3 Determining the position and ‘switch-on’ time of the SAW source without regard for refraction corrections
The ionospheric region that is responsible for the main contribution to TEC variations lies in the neighborhood of the maximum of the ionospheric $`F`$-region, which does determine the height $`h_{max}`$ of the subionospheric point. When selecting $`h_{max}`$, it should be taken into consideration that the decrease in electron density with height above the main maximum of the $`F_2`$-layer proceeds much more slowly than is the case below the maximum. Since the density distribution with height is essentially a ‘weight function’ of the TEC response to a wave disturbance (Afraimovich et al., 1992), it is appropriate to use, as $`h_{max}`$, the value exceeding the true height of the layer $`h_{F2}`$ maximum by about 100 km. $`h_{F2}`$ varies over a reasonably wide range (250–350 km) depending on the time of day and on some geophysical factors which, when necessary, can be taken into account if corresponding additional experimental data and current ionospheric models are used. In all calculations that follow, $`h_{max}=400`$ km is used.
To a first approximation, it can be assumed that it is at this altitude where the imaginary detector is located, which records the ionospheric SAW response in TEC variations. The ‘horizontal size’ of the detection region, which can be inferred from the propagation velocity of the subionospheric point as a consequence of the orbital motion of the GPS satellite (of order 50 m/s), and from the SAW period (of order 300 s — see Section 4), does not exceed 15–20 km, which is far smaller than the ‘vertical size’.
From the GPS data we can determine the coordinates $`X_s`$ and $`Y_s`$ of the subionospheric point in the horizontal plane $`X0Y`$ of a topocentric frame of reference centered on the point B(0,0) at the time of a maximum TEC deviation caused by the arrival of the SAW at this point (see Fig. 3a). Since we know the angular coordinates $`\theta `$ and $`\alpha `$ of the wave vector $`K_t`$, it is possible to determine the location of the point at which this vector intersects the horizontal plane $`X^{}0Y^{}`$ at the height $`h_w`$ of the assumed source. Assuming a rectilinear propagation of the SAW from the source to the subionospheric point and neglecting the sphericity the coordinate $`X_w`$ and $`Y_w`$ of the source in a topocentric frame of reference can be defined as
$$X_w=X_p\left(h_{max}h_w\right)\frac{\mathrm{cos}\theta \mathrm{sin}\alpha }{\mathrm{sin}\theta }$$
(9)
$$Y_w=Y_p\left(h_{max}h_w\right)\frac{\mathrm{cos}\theta \mathrm{cos}\alpha }{\mathrm{sin}\theta }$$
(10)
The coordinates $`X_w`$ and $`Y_w`$, thus obtained, can readily be recalculated to the values of the latitude and longitude ($`\varphi _w`$ and $`\lambda _w`$) of the source.
For SAW generated during earthquakes, industrial explosions and underground tests of nuclear devices, $`h_w`$ is taken to be equal to 0 (the source lying at the ground level). When recording SAW produced by launchings of powerful rockets, the region of SAW generation can lie at heights $`h_w`$ of order 100 km or higher (Li et al., 1994; Nagorsky, 1998).
Using this approximation we neglect the possible refraction at the SAW propagation from the source to the height $`h_{max}`$. In some publications (Calais et al., 1998) this problem is solved by performing trajectory calculations using standard ‘$`raytracing`$’ procedures and neutral atmosphere models. In doing so, ‘$`raytracing`$’ calculations were carried out from the source. In this study it is also possible to perform such calculations, not from the source but from the subionospheric point (return trajectory).
It should be noted, however, that according to our data (Section 4) and to the conclusions of some authors (Li et al., 1994; Nagorsky, 1998), the source lies at 100 km altitude at least. Thus the region of SAW generation is within the narrow height range (90–150 km) where the sound velocity gradient is maximal (Li et al., 1994). Within the approximation of ‘short’ waves (geometrical optics), this condition would mean the significance of refraction effects. However, the radial size of the region of SAW formation, like the extent of the disturbance itself which it generated by the rocket jet, is about 30–50 km at the above-mentioned heights. This value is comparable with the typical scale of variation of the sound velocity with height, hence the wave is no longer a ‘short’ one, and the geometrical optics approximation is inapplicable. One would therefore expect that the refraction effect in the neighborhood of the SAW source will be smaller than anticipated in terms of geometrical optics.
Given the coordinates of the subionospheric point and of the disturbance source, the mean value of the SAW propagation velocity between the source and the subionospheric point, and the arrival time of the SAW at this point, then within the approximation of a rectilinear propagation it is easy to determine the ‘switch-on’ delay $`\mathrm{\Delta }t_w`$ of the anticipated SAW source with respect to the start. This would make it possible to obtain additional information about the SAW source which is needed to understand the mechanism of SAW generation. The estimates of $`\mathrm{\Delta }t_w`$ made below assume that the propagation velocity is taken equal to 700 m/s (see Li et al., 1994). Note that by the ‘switch-on’ time of the source is meant here the time of a maximum disturbance of the background state of the medium when the SAW is generated.
## 4 Results of measurements
Hence, using the transformations described in Section 3, we obtain the following parameters determined from TEC variations and characterizing the SAW: $`t_0`$ — start time; $`t_p`$ — time of a maximum deviation of the frequency Doppler shift $`F(t)`$; $`\mathrm{\Delta }t`$ — delay of $`t_p`$ with respect to $`t_0`$; $`T`$ — SAW period; $`A_I`$ — TEC disturbance amplitude; $`A_F`$ — amplitude of a maximum frequency Doppler shift at the ‘reduced’ frequency of 136 MHz; $`\alpha `$ and $`\theta `$ — azimuth and elevation of the wave vector $`K_t`$; $`V_h`$ and $`V_t`$ — horizontal component and modulus of the phase velocity; $`V_a`$ — mean wave velocity calculated from the delay $`\mathrm{\Delta }t`$ and from the known path-length between the launch pad and the subionospheric point; $`\varphi _w`$ and $`\lambda _w`$ — latitude and longitude of the source at 100km altitude; and $`\mathrm{\Delta }t_w`$ — ‘switch-on’ delay of the assumed SAW source with respect to the start.
Corresponding values of the SAW parameters, and also site names of the GPS-array and GPS satellite PRN numbers are presented in Table 3 for Baikonur Cosmodrome and Table 4 for the KSC Cosmodrome.
It should be noted that the estimates of $`A_I`$ and $`A_F`$ are obtained by filtering ‘oblique’ TEC series; therefore, equivalent estimates for ‘vertical’ TEC are smaller by a factor varying from 1 to 2 depending on the elevation $`\theta _s`$ of the beam to the satellite.
Solid curves in Figs. 1 and 2 show trajectories of subionospheric points for each of the GPS satellites at the height $`h_{max}=400`$ km. Dark diamonds along the trajectories correspond to the coordinates of subionospheric points at the time $`t_p`$ of a maximum deviation of the frequency Doppler shift $`F`$ (Fig. 4c). Asterisks designate the source location at 100 km altitude inferred from the GPS-array data. Numbers at the asterisks refer to the corresponding day numbers and to the ‘switch-on’ delay of the source with respect to the start time. Dashed straight lines connecting the anticipated source with the subionospheric point show the horizontal projection of the wave vector $`K_t`$ (see Fig. 3a).
### 4.1 SAW parameters during Proton LV launches from Baikonur Cosmodrome
Let us consider the results derived from analyzing the ionospheric effect of SAW during failed Proton LV launch on July 5, 1999 (line 2 in Table 2) obtained at the array (SELE, CHUM,SHAS) for PRN14 (at the left of Fig. 4, and line 1 in Table 3).
In this case the delay of the SAW response with respect to the start time is 12 min. The SAW has the form of an $`N`$-wave with a period $`T`$ of about 300 s and an amplitude $`A_I=0.5TECU`$, which is an order of magnitude larger than TEC fluctuations for background days. It should be noted, however, that this time interval was characterized by a very low level of geomagnetic activity (11 nT).
The amplitude of a maximum frequency Doppler shift $`A_F`$ at the ‘reduced’ frequency of 136 MHz was found to be 0.12 Hz. In view of the fact that the shift $`F`$ is inversely proportional to the sounding frequency squared (Davies, 1969), this corresponds to a Doppler shift at the working frequency of 13.6 MHz and the equivalent oblique-incidence sounding path of about $`A_F=12`$ Hz.
The azimuth and elevation $`\alpha `$ and $`\theta `$ of the wave vector $`K_t`$ whose horizontal projection is shown in Fig. 1 by a dashed line and is marked by $`K_1`$, are 153 and 59, respectively. The horizontal component and the modulus of the phase velocity were found to be $`V_h=1808`$ m/s and $`V_t=931`$ m/s. The source coordinates at 100 km altitude were determined as $`\varphi _w=48^{}`$ and $`\lambda _w=66^{}`$.
The ‘switch-on’ delay of the SAW source $`\mathrm{\Delta }t_w`$ with respect to the start time was 264 s. The ‘mean’ velocity of about $`V_a=1000`$ m/s, determined in a usual manner from the response delay with respect to the start, was close the phase velocity $`V_t`$.
Similar results for the array (POL2, CHUM, SHAS) and PRN09 were also obtained for failed launch of October 27, 1999. They correspond to the projection of the vector $`K_2`$ in Fig. 1, to the time dependencies in Fig. 4 (at the right), and to line 7 in Table 3. One can only note that the SAW amplitude was by a factor of 4–5 smaller than that for launch of July 5, 1999. At an increased level of magnetic activity (–80 nT) this led to a smaller (compared to July 5, 1999) signal/noise ratio, which, however, did not interfere with obtaining reliable estimates of SAW parameters.
Fig. 5 (at the right) and line 12 of Table 3 present data for standard launch of the Proton LV on November 20, 1998. A corresponding projection of the vector $`K_3`$ is shown in Fig. 1. A comparison of the data for standard launch November 20, 1998 and failed launches showed that SAW parameters were reasonably similar, irrespective of the level of geomagnetic disturbance, the season, and the local time.
### 4.2 SAW parameters during Space Shuttle launches from the KSC Cosmodrome
Let us consider the results derived from analyzing the ionospheric effect of the SAW during Space Shuttle LV launch on October 29, 1998 (line 3 of Table 2) obtained at the array (AOML, KYW1, EKY1) for PRN01 (at the left of Fig. 5, and line 1 in Table 4).
As in the case of Proton LV launches, the delay of the SAW response with respect to the start time is 12 min. The SAW has the form of an $`N`$-wave with a period $`T`$ of about 210 s and an amplitude $`A_I=0.3TECU`$ (for AOML), which is an order of magnitude larger than TEC fluctuations for background fields. The amplitude of a maximum frequency Doppler shift $`A_F`$ at the ‘reduced’ frequency of 136 MHz was found to be 0.07 Hz (for AOML). It should be noted, however, that this time interval was characterized by a very low level of geomagnetic activity (–15 nT).
The azimuth and elevation $`\alpha `$ and $`\theta `$ of the wave vector $`K_t`$ whose horizontal projection is shown in Fig. 2 by a dashed line and is marked by $`K_1`$ are 214 and 34.7, respectively. The horizontal component and the modulus of the phase velocity, and the mean velocity $`V_a`$ are similar for those of the Proton LV (see Table 4). The source coordinates at 100 km altitude are determined as $`\varphi _w=28.5^{}`$ and $`\lambda _w=284^{}`$. The ‘switch-on’ delay of the SAW source $`\mathrm{\Delta }t_w`$ with respect to the start time is 204 s.
### 4.3 Verifying the reliability of measurements of SAW parameters
To convince ourselves that the determination of the main parameters of the SAW form and dynamics is reliable for the launches analyzed here, in the area of the Baikonur and KSC Cosmodromes we selected different combinations of three sites out of the sets of GPS stations available to us, and these data were processed with the same processing parameters. Relevant results for Baikonur (including the average results for the sets $`\mathrm{\Sigma }`$), presented in Table 3 and in Fig. 1 (SAW source position), show that the values of SAW parameters are similar, which indicates a good stability of the data obtained, irrespective of the GPS-array configuration.
The relative position of the Proton LV flight path and of the GPS-array stations suitable for SAW detection was very convenient for our experiment because the subionospheric points of the GPS satellites were close to the portion of the trajectory where the anticipated SAW source was located (Li et al., 1994; Nagorsky, 1998), and away from the launch pad. In addition, the aspect condition (3) corresponding to a maximum amplitude of the SAW response to the SAW passage was satisfied quite well for this geometry for all array stations simultaneously. This is confirmed by a high degree of correlation of SAW responses at the array elements (Fig. 4c), which made it possible to obtain different sets of triangles from the six GPS stations available to us.
Unlike he Baikonur Cosmodrome, the relative position of the Space Shuttle LV flight path and of the GPS-array stations was inconvenient for SAW detection because the subionospheric points of the GPS satellites were near the launch pad. As a consequence the aspect condition (3) was not satisfied simultaneously for all GPS stations located near the Cosmodrome. This is especially true for stations CCV1, CCV3 and EKY1 located in the neighborhood of the launch pad along a direction similar to that of the LV flight path. This resulted in a low degree of correlation and a very different amplitude of SAW responses at the array elements (Fig. 5f). Some of the stations listed in Table 1 are too close to each other (CCV1, CCV3; MIA1, MIA3; KYW1, KYW2) and cannot be used in mutual verification of the reliability of measurements. As a result, we were able to obtain, for the October 29, 1998 launch, only two sets of triangles out of the nine GPS stations available to us (see Table 4).
Because of the low correlation of the responses and the inconvenient GPS-array geometry for the April 17, 1998 launch, we were also unable to obtain reliable estimates of the SAW wave vector parameters. Therefore, we did not plot the relevant data in Fig. 2 and limited ourselves to mentioning them in Table 4 (lines 4 and 5).
We availed ourselves also of another method to check the data from the GPS-array for reliability, namely, a modeling using the algorithm for calculating TEC along the ‘receiver-satellite’ beam for a typical model of the regular ionosphere, a disturbed SAW with specified properties described by Afraimovich et al. (1998). A peculiarity of this algorithm is that it is possible to calculate TEC for a particular selected array and for a real trajectory of the satellite determined by the initial navigational RINEX-file.
As an example, Fig. 5 (left) shows the computed data for the array (SELE, CHUM, SHAS) and for the SAW model in the form of wave packet (5) with a period of 240 s, the packet’s duration of 240 s, a maximum amplitude of 0.5 $`TECU`$, the horizontal phase velocity $`V_h=1200`$ m/s, and the azimuth $`\alpha `$ and the elevation $`\theta `$ of the wave vector $`K_t`$ equal to 153 and 60, respectively. A comparison of the parameters specified in the modeling procedure with those obtained by processing the data of TEC $`I(t)`$ model series shows that these values agree with an accuracy no worse than 10%. Noteworthy also is a good agreement of experimental (at the left of Fig. 4) and model (at the left of Fig. 5) TEC variations $`\mathrm{\Delta }I(t)`$ and $`F(t)`$ for the array (SELE, CHUM, SHAS).
## 5 Discussion
Here we discuss the main results and compare them with findings reported by other authors. Within the context of this study we deliberately pass over a physical interpretation of our obtained results because of no information available about the dynamics and energetics of Proton and Space Shuttle LV launches. Our intention is to obtain more trustworthy and reliable data ensuring the new possibilities of a global GPS monitoring.
### 5.1 The form of response
It was found that in spite of the difference of LV characteristics, the local time, the season, and the level of geomagnetic disturbance, the ionospheric response for all launches has the character of an $`N`$-wave. The SAW period $`T`$ is 270–360 s, and the amplitude exceeds the standard deviation of background TEC fluctuations in this range of periods under quiet and moderate geomagnetic conditions, by a factor of 2–5 as a minimum.
Our measurements of the period and amplitude of the SAW response are in good agreement with frequency Doppler shifts measured in the HF range during Space Shuttle LV launches on February 28, 1990 and April 28, 1991 (Jacobson and Carlos, 1994), as well as wit corresponding estimates of a maximum shift $`F`$ reported by Nagorsky (1998, 1999) for oblique HF radio path during LV launches from Baikonur Cosmodrome. They are also similar to the estimates obtained by Li et al. (1994) using the transionospheric VHF radio signal from geostationary satellite MARECS-B2 during Space Shuttle LV launches on October 18, 1993 (STS-58) and February 3, 1994 (STS-60).
According to measurements reported by Calais and Minster (1996), during Space Shuttle (STS-58) launch on October 18, 1993 two series of TEC fluctuations were recorded, one of which had also the form of an $`N`$-wave with a maximum amplitude of 0.25 TECU, which is also consistent with our data.
It was pointed out in the literature that SAW of a similar form and with a similar amplitude were detected during powerful industrial explosions (Afraimovich et al., 1984; Jacobson et al., 1988; Blanc and Jacobson, 1989; Fitzgerald, 1997; Calais et al., 1998a; 1998b; Nagorsky, 1998).
### 5.2 Angular characteristics of the wave vector, and the phase velocity of SAW
As pointed out in the Introduction, some researchers report markedly different values of the SAW propagation velocity — up to several thousand m/s, which exceeds the sound velocity at SAW propagation heights in the atmosphere. According to the data from a review by Karlov et al. (1980), the propagation velocity of the SAW ionospheric response recorded during launches of the Apollo mission rockets, varied from 600 to 1670 m/s.
Arendt (1971) suggested that the shock wave produced by rocket flight is divided in the ionosphere at about 160 km altitude into the ion-acoustic mode (with a velocity as high as 1.3 km/s) and a normal acoustic mode (with the velocity of up to 500 m/s). Arendt (1971) explains the difference in propagation velocities of the first and second disturbances observed during Apollo-14 and Apollo-15 launches at the same distance of 1440 km from the start site, by a difference in atmospheric conditions and propagation paths of the waves because of seasonal variations of the ionosphere.
Noble (1990) describes the observations of long-period waves over the Arecibo incoherent scatter station during Space Shuttle STS-4 launch on June 27, 1982. At a large distance from the path (up to 1000 km) the group velocity of wave propagation was found to be 600–700 m/s.
According to GPS measurements (Calais and Minster, 1996), the SAW phase velocity at ionospheric heights is about 1000–1300 m/s.
A common limitation of existing methods for determining the SAW phase velocity is the need to know the launch time of the rocket since this velocity is calculated from the SAW delay with respect to the start time assuming that the velocity along the propagation path is constant, which is by far contrary to fact. Furthermore, only the horizontal component of the phase velocity $`V_h`$ was in essence determined in such studies. At different values of elevation of the wave vector $`K_t`$, the velocity $`V_h`$ corresponds to markedly differing values of the modules of the phase velocity $`V_t`$.
The use of the method proposed in this paper makes possible determining angular characteristics of the wave vector $`K_t`$ and, accordingly, estimating $`V_t`$. According to our data (Tables 3 and 4), the elevation of the SAW wave vector varied from 30–60, and the SAW phase velocity was in the range of from 900 to 1200 m/s. We determine the phase velocity of the line of equal TEC at the height of the ionospheric $`F`$-region maximum which makes the main contribution to variations in TEC between the receiver and the GPS satellite and corresponds to the region of maximum sensitivity of the method. Since $`V_t`$ approaches the sound velocity at these altitudes (Li et al., 1994), this makes it possible to identify the sound nature of a TEC perturbation.
### 5.3 The location and ‘switch-on’ delay of the SAW source
The position of the SAW source, calculated by neglecting refraction corrections, corresponds to the segment of the LV trajectory (Figs. 1 and 2) within distances of at least 700–900 km from the launch pad for the Proton LV and of at least 200–500 km for the Space Shuttle LV. This is consistent with the ‘switch-on’ delay $`\mathrm{\Delta }t_w`$ of the source which is 250–300 s for the Proton LV and 200 s for the Space Shuttle LV. As is evident from our data, the calculated position of the SAW source for rocket launches does not coincide with the position of the launch pad. At the same time the source location is in reasonably good agreement with that of horizontal projections of LV trajectories (Figs. 1 and 2).
Kaschak et al. (1970) analyzed the data from the infra sound measuring arrays located on the USA north-eastern coast which observed strong acoustic signals from launch and reentry areas of the Saturn-5 LV. The authors identified three types of signals: one includes early signals whose arrival time corresponded to supersonic values of their velocity equal to 500–1000 m/s; the other type represents normal signals with the group velocity approximately equal to the normal velocity of sound propagation in the air; and the last type involves late signals with subsonic velocities in the range of from 190 to 240 m/s. Kaschak et al. (1970) and Balachandran et al. (1971) suggested that the so-called ‘early’ signals during rocket launches are caused by SAW which are generated during the reentry of the first stage at distances from the launch pad exceeding 500 km.
However, our data are in better agreement with the mechanism substantiated by Li et al. (1994), Calais and Minster (1996), and Nagorsky (1998, 1999). They believe that the generation of SAW occurs during a nearly horizontal travel of the rocket with the operating engine along the acceleration segment of the trajectory, at the lower atmospheric heights of 100–130 km. The rocket travels this segment with supersonic velocity at 100–300 s of its flight at a distance of at least 500 km from the launch pad (see the data on the Proton LV flight schedule in Section 2). As soon as it ascends to an altitude of about 100 km, the SAW source is ‘switched on’.
## 6 Conclusions
In this paper we have investigated the form and dynamics of shock-acoustic waves (SAW) generated during rocket launches. We have developed a method for determining SAW parameters (including angular characteristics of the wave vector and the SAW phase velocity, as well as the direction to the source) using the GPS-arrays whose elements can be chosen out of a large set of GPS stations forming part of a global IGS network. Unlike existing radio techniques, the proposed method estimates SAW parameters without a priory information about the site and time of rocket launch. The implementation of the method is illustrated by analyzing ionospheric effects from launches of the Proton and Space Shuttle LV from Baikonur and KSC Cosmodromes in 1998 and 1999 (totaling five launches).
The results reported in this study suggest the following conclusions:
1. In spite of the difference of LV characteristics, for all launches the ionospheric response has the character of an $`N`$-wave corresponding to the form of a shock wave, irrespective of the type of disturbance source (rocket launch, industrial explosion).
2. The SAW period $`T`$ is 270–360 s, and its amplitude (from 0.1 to 0.5 $`TECU`$) exceeds the standard deviation of $`TECU`$ background fluctuations in this range of periods under quiet and moderate geomagnetic conditions by a factor of 2–5 as a minimum.
3. The elevation of the SAW wave vector varies within 35–60, and the SAW phase velocity (900–1200 m/s) approaches the sound velocity at heights of the ionospheric $`F`$-region minimum. This makes it possible to identify the sound nature of a TEC perturbation.
4. The position of the SAW source, calculated by neglecting refraction corrections, corresponds to a segment of the Proton LV trajectory at a distance of at least 700–900 km from the launch pad and to the LV flight altitude of at least 100 km. For the Space Shuttle LV the position of the SAW source corresponds to the segment of the trajectory at a distance of at least 200–500 km from the launch pad and to the LV flight altitude of at least 100 km.
Hopefully, our investigation would provide additional useful insights into the physical processes occurring during flights of rockets in the Earth’s atmosphere on the acceleration segment of the trajectory. In addition, this would make it possible to identify more reliable signal indications of technogenic effects which are necessary for constructing an effective global radio subsystem for detection and localization of these effects by processing the data from an international network of two-frequency receivers of the GPS-GLONASS navigation systems.
## 7 Acknowledgements
We are grateful to E. A. Ponomarev, V. V. Yevstafiev, A. M. Uralov, P. M. Nagorsky, N. N. Klimov, and A. D. Kalikhman for their interest in this study, helpful advice and active participation in discussions. Authors are grateful to K. S. Palamartchouk and O. S. Lesuta for preparing the input data. Thanks are also due to V. G. Mikhalkosky for his assistance in preparing the English version of the manuscript. This work was done with support from the Russian foundation for Basic Research (grants 97-02-96060 and 99-05-64753), as well as under grant 1999 of the RF Ministry of Education (Minvuz), under direction of B. O. Vugmeister.
|
warning/0007/astro-ph0007060.html
|
ar5iv
|
text
|
# COBE- AND CLUSTER-NORMALIZED CDM SIMULATIONS
## 1 Introduction
$`N`$-body simulations are an essential tool for probing LSS and galaxy formation. As large, high-resolution simulations are computationally costly, one has to carefully consider the added effort resulting from increasing the simulations’ volume or from improving their resolution. In this context, most simulations gravitate towards a design stressing either of these two conflicting goals. The largest three-dimensional redshift surveys currently available are situated somewhere in between these two extremes: a $`M_{}`$ galaxy in the CfA2 survey is visible out to $`100h^1\text{Mpc}`$. A simulation designed to match these surveys must have both the required resolution to identify the dark matter (DM) halos associated with such galaxies, *and* this moderately large volume.
The other essential consideration of simulation design is the choice of cosmological models probed. Ideally, one would want to examine a certain range of the relevant cosmological parameters ($`H_0`$, $`\mathrm{\Omega }_0`$, $`\lambda _0`$, $`\mathrm{\Omega }_{\mathrm{B0}}`$, $`n`$, $`\sigma _8`$), but this is often not an attainable goal. In this work we focus on cosmological models with currently favored values of $`H_0`$, $`\mathrm{\Omega }_0`$ (and $`\lambda _0`$), and require that all models will be COBE-normalized. The above constraints can still be fulfilled through a variety of primordial power spectrum tilt $`n`$ and $`\sigma _8`$ combinations. We attempt to achieve also cluster normalization, hence determining the value of $`\sigma _8`$ (and thus fixing a tilt value).
## 2 Models
These simulations were designed with the goal of maximizing their volume while still being able to resolve DM halos associated with $`L_{}`$ galaxies. Adopting a mass-to-light ratio of $`100M_{}/L_{}`$, we thus require that we will be able to identify $`M_{\mathrm{halo}}>10^{12}M_{}`$ halos. A simulation with $`140^3`$ particles in a $`100h^1\text{Mpc}`$ box will have $`M_{\mathrm{particle}}=1.01\times 10^{11}\mathrm{\Omega }_0h^1M_{}`$. Identifying all DM halos with 20 particles or more, this parameter specification matches the stated requirements.
Failure to identify halos all the way down to this resolution renders such simulations inappropriate for the study of many LSS features. Specifically, it would be impossible to identify correctly voids in such simulations. Such studies often originate from DM simulations, and use some form of halo populating scheme in order to match the observed properties of the distribution of galaxies. These populating schemes (also known as bias recipes) assign a number of galaxies to each halo. But for such schemes to have a fighting chance at reproducing the distribution of galaxies, one must initially know the locations of the halos that should be populated—including halos that would be populated by just one galaxy.
We estimated that 20 particles per halo is the minimal number required in order to reliably identify halos. As this is getting close to the simulation’s mass resolution limit, we tested the lower end of our DM halo mass function by constructing a matching set of smaller simulation boxes with $`M_{\mathrm{particle}}`$ an order of magnitude smaller $`^\mathrm{?}`$ (see Table 1). We can then compare the number density of DM halos in the two sets of boxes and see if indeed we manage to recover the correct number of small halos in the larger simulation boxes.
All simulations were started at an initial redshift $`z_i=24`$ and evolved over 600–1000 timesteps using a $`\mathrm{P}^3\mathrm{M}`$ code over a $`256^3`$ grid. More details on the simulations can be found elsewhere. $`^{\mathrm{?},\mathrm{?}}`$
In Table 2 we summarize the cosmological parameters defining our models. Column 1 indicates the models’ acronyms. Columns 2–5 detail the models’ values of the Hubble parameter $`h`$, density parameter $`\mathrm{\Omega }_0`$, cosmological constant $`\lambda _0`$, and tilt $`n`$. Column 6 lists the theoretical $`\sigma _8^{\mathrm{cluster}}`$ based on X-ray cluster temperatures. Column 7 lists $`\sigma _8^{\mathrm{cont}}`$, the actual value corresponding to each cosmological model. If these two values match, we state that the model is cluster-normalized (Column 8).
We fixed all models to be COBE-normalized with $`T_{\mathrm{CMB}}=2.7K`$, assuming no contribution from tensor modes. Also, we used $`\mathrm{\Omega }_{\mathrm{B0}}=0.015h^2`$ throughout, $`^\mathrm{?}`$ in concordance with the Texas $`\mathrm{P}^3\mathrm{M}`$ Database. When practical (all models but TCDM) we adopt a Hubble constant $`^\mathrm{?}`$ $`H_0=65\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$. In addition to COBE normalization, we have also tried to achieve cluster normalization. Using the X-ray cluster $`M`$$`T`$ relation $`^\mathrm{?}`$ we derived the required $`\sigma _8`$ value for each of our models. For each combination of $`\mathrm{\Omega }_0`$, $`\lambda _0`$, and $`H_0`$, we computed the tilt required in order to achieve cluster normalization and examined whether it is acceptable in view of the limits allowed by the 4-year COBE data.
For two of the models—LCDM and TOCDM—we found an acceptable tilt value and managed to achieve cluster normalization. The two other models are not cluster-normalized. The OCDM model was designed as a direct companion to the LCDM model, where all the cosmological parameters in both these models—except the value of $`\lambda _0`$—are the same. The TCDM model is our best attempt with an $`\mathrm{\Omega }_0=1`$ model, where we used the lowest possible $`\sigma _8`$ value which does not require $`h<0.55`$ or $`n<0.7`$. For comparison, we have also included in Table 2 (and in Fig. 1) the familiar SCDM model, although it is not one of the cosmological models simulated.
## 3 Results
In Fig. 1 we compare the theoretical CMB angular power spectra of the models simulated here with the recent BOOMERanG $`^\mathrm{?}`$ anisotropy measurements. In Fig. 2 we present cumulative halo mass functions for the two sets of models simulated. Our two cluster-normalized models, LCDM and TOCDM, reproduce similar mass functions. The observational point in the figure $`^\mathrm{?}`$ $`n(T>4.0\mathrm{keV})=1.5\pm 0.4\times 10^6h^3\mathrm{Mpc}^3`$ is in good agreement with the LCDM cluster abundance. The TOCDM curve follows closely the LCDM curve, but for the former cosmology the observational point would be shifted along the horizontal axis by a factor $`0.3^{1/3}`$. However, it should be noted that there are still significant uncertainties associated with both the observational measurements of cluster abundance and the theoretical modeling of the $`M`$$`T`$ relation. $`^\mathrm{?}`$
There are two curves for each cosmological model in Fig. 2—one representing the mass function as measured in the $`100h^1\text{Mpc}`$ box, the other measured in the $`40h^1\text{Mpc}`$ box. As illustrated in the figure, for each model there is excellent agreement between the two curves.
## 4 Summary
In this paper we introduce two matching sets of four cosmological models. We derive halo mass functions for all models and use the small box, high resolution simulations in order to verify the validity of the mass function in the large box for halos as small as $`10^{12}\mathrm{\Omega }_0h^1M_{}`$. The simulations presented here are unique as they both cover a volume comparable to current large three-dimensional redshift surveys and at the same time resolve cluster masses down to $`M_{}`$. While the simulations were designed mostly in order to achieve COBE—and, when possible (LCDM and TOCDM), also cluster—normalization, they also serve to demonstrate the attractiveness of the LCDM model. Requests for the simulations presented in this paper, or for other data from the Texas $`\mathrm{P}^3\mathrm{M}`$ Database, should be sent to database@galileo.as.utexas.edu.
## Acknowledgments
We are indebted to Mike Gross for his stimulating help and friendship. We thank Ue-Li Pen for helpful discussions and comments. HE was supported by a Smithsonian Predoctoral Fellowship. This work was supported by NASA grants NAG5-7363 and NAG5-7812; NSF grant ASC 9504046; and the Texas Advance Research Program grant 3658-0624-1999.
## References
|
warning/0007/astro-ph0007334.html
|
ar5iv
|
text
|
# CO and Near-Infrared Observations of High-Redshift Submillimeter Galaxies
## 1 Introduction
The discovery of an ultraluminous population of high-redshift galaxies with deep submillimeter surveys has revolutionized our understanding of the distant universe. The current data show that the sub-mm population has a mixture of AGN and starburst characteristics with properties which are roughly consistent with the local population of ultraluminous ($`L>10^{12}L_{}`$) infrared galaxies (ULIGs). The relative importance of AGN and starburst activity in powering the high luminosities of the sub-mm population is still an open question, but the growing consensus is that the majority of the luminosity of the population is powered by star formation. The early CO and X-ray data on the sub-mm population support the starburst nature of the population by showing the presence of sufficient molecular gas to fuel the star-formation activity and the lack of expected X-ray emission if mostly dominated by AGN.
Although the redshift distribution of the sub-mm population is still uncertain, the majority of the sub-mm galaxies are believed to be at high redshifts ($`z>2`$) based on their radio and near-infrared data. The early redshift distributions based on optical imaging and spectroscopy suggested somewhat lower redshifts, but several of the original candidate optical counter-parts have turned out to be incorrect. Despite their ultra-high luminosities, many sub-mm galaxies are undetected at ultraviolet/optical wavelengths due to extinction by dust. For these highly obscured galaxies, follow-up radio and/or mm interferometric observations are required in order to uncover the proper counter-part.
In order to understand the nature of the sub-mm population, we have been carrying out multi-wavelength observations of individual systems in the SCUBA Cluster Lens Survey. This survey represents sensitive sub-mm mapping of seven massive, lensing clusters which uncovered 15 background sub-mm sources. The advantages of this sample are that the amplification of the background sources allows for deeper source frame observations and that lensing by cluster potentials does not suffer from differential lensing. We have concentrated our efforts on the nine background galaxies detected at the highest signal–to–noise (Table 1). Only three sources have spectroscopic redshifts, and the redshift lower limits shown in Table 1 are based on their sub-mm/radio flux ratios.
## 2 CO Results
At OVRO we have conclusively detected CO emission from two sub-mm systems, SMM J02399$``$0136 at $`z=2.8`$ (SMM J02399) and SMM J14011+0252 at $`z=2.6`$ (SMM J14011) (Fig. 1). A third system SMM J02399$``$0134, which is associated with a ring-galaxy containing a Seyfert nucleus at $`z=1.06`$, has recently been detected in CO at the PdB. We have also tentatively confirmed the PdB detection at OVRO. To date, these three galaxies are the only sub-mm sources with known redshifts, and it is promising that all three have already been detected in CO. The early CO results suggest that the sub-mm population contains massive reservoirs of molecular gas and are among the most CO luminous galaxies in the universe.
The strongest sub-mm source, SMM J02399, shows an AGN component in its optical spectrum, while SMM J14011 shows only evidence for starburst activity at optical/NIR wavelengths. Although the optical characteristics of these two galaxies are vastly different, their radio, sub-mm, and CO properties are fairly similar and are consistent with a high level of star formation activity (SFRs of a few$`\times 10^2`$ M yr<sup>-1</sup> to more than $`10^3`$ M yr<sup>-1</sup>, depending on the IMF and AGN contamination). After correcting for lensing, we derive CO luminosities of 3–4$`\times 10^{10}`$ K km s<sup>-1</sup> pc<sup>2</sup> (H$`{}_{o}{}^{}=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup>; $`q_o=1/2`$) in these two systems. These CO luminosities correspond to molecular gas masses of about $`5\times 10^{10}`$$`2\times 10^{11}`$ M, depending on the exact value of the CO to H<sub>2</sub> conversion factor. Both SMM J02399 and SMM J14011 appear to be associated with a merger event. Given that mergers of gas-rich galaxies at low-redshift result in massive starbursts, we expect star-formation to be an important component for powering the far-infrared luminosities in both of these systems. In fact, the large molecular gas masses of SMM J02399 and SMM J14011 are sufficient to form the stars of an entire L galaxy, which suggests that the sub-mm population may represent the formative phase of massive galaxies.
SMM J02399 is unresolved in CO, while SMM J14011 is extended over a large spatial scale in its source frame ($`>10`$ kpc). Figure 1 contains low resolution OVRO data which showed tentative evidence for extended CO emission in the north-south direction in SMM J14011. The extended morphology of the molecular gas in SMM J14011 has been recently confirmed with higher-resolution data from OVRO and BIMA. These results may suggest that SMM J14011 is in an early stage of its merger event, unlike the majority of ULIGs in the local universe whose CO emission is mostly contained within the central kpc. It is currently unknown what fraction of the sub-mm sources are compact or are extended over large spatial scales as is SMM J14011. If the progenitors of the sub-mm systems are more gas rich than those of local ULIGs, we could expect the sub-mm sources to have larger gas fractions and to be more extended than their low-redshift analogs. A large sample of sub-mm sources need to be observed in CO before statistical comparisons could be made between the CO properties of local ULIGs and the high-redshift sub-mm galaxies.
## 3 Near-Infrared Results
Many sub-mm galaxies are too obscured by dust to be detected at ultraviolet/optical wavelengths. At least four of the nine sources in our sample which were undetected at optical wavelengths ($`I>25`$–26) have faint near-infrared, $`K`$-band counter-parts. Two of these are bright enough in $`K`$-band ($`K=19.1`$, 19.6 mag) to be classified as extremely red objects (EROs). An additional faint ($`K=21`$) galaxy was found associated with a relatively bright (0.5 mJy) radio counter-part. The fourth and faintest galaxy with a near-infrared counter-part is SMM J00266+1708.
The sub-mm position of SMM J00266+1708 is located between three optically visible galaxies. We imaged the field at OVRO at 1.3 mm and determined its position to be offset from all three optical sources. Deep, follow-up near-infrared observations with NIRC on Keck uncovered a new faint galaxy at $`K=22.5`$ located at the position of the 1.3 mm source. Although SMM J00266+1708 is the second brightest sub-mm source in the SCUBA Cluster Lens Survey, it is currently the faintest known near-infrared counter-part of a sub-mm galaxy discovered to date (Fig. 2).
Only two of the nine sources in the sample still require deep $`K`$-band imaging and currently have uncertain counter-parts. The galaxy SMM J02400$``$0134 has no optically detected galaxies near the sub-mm position. For SMM J22471-0206 there are several optical galaxies which could be the sub-mm counter-part, but given previous results it will be interesting to test whether or not any new candidate galaxies are uncovered with deep $`K`$-band imaging. Depending on the results for these last two unknown systems, the current data suggest that approximately 40%–70% (4/9–6/9) of the sub-mm population as a whole have faint near–infrared counter-parts which are undetected at optical wavelengths. Only 30%–40% (3/9–4/9) of the sample have optical counterparts ($`I<`$26–27, correcting for source lensing).
The $`K`$-band magnitudes of the sub-mm counter-parts in the SCUBA Cluster Lens Survey range over 6 magnitudes which reflects the wide diversity of colors and redshifts for the population (Table 1). The magnitudes listed in Table 1 have not been corrected for lensing; unlensed sources would be about a magnitude fainter on average. Currently, only the three brightest optical ($`I`$-band) sources have spectroscopic redshifts. The other sources are much fainter and would require near-infrared spectroscopy to obtain redshifts, which will be challenging even with 8m/10m class ground base telescopes. Our early spectroscopic results with NIRSPEC on Keck suggest that lines may be detectable in the brighter sources ($`K<20`$), while many of the fainter sub-mm sources ($`K>22`$) may have to wait for the Next Generation Space Telescope.
## 4 Conclusions
Most sub-mm sources are too red and/or faint to be detected at optical wavelengths. There is very little overlap, if any, between the ultraluminous sub-mm population and the less luminous, optically-selected Lyman Break population of galaxies. This highlights the importance of radio, millimeter, and near-infrared observations of the sub-mm population for our general understanding of the evolution and formation of galaxies.
Potentially, we do not need to wait for future optical/near-infrared space-based missions to obtain redshifts for the bulk of the sub-mm population. Redshifts could be determined directly from the CO lines themselves with planned ground-based millimeter telescopes, such as the LMT and ALMA. Both ALMA and the LMT will have sufficient sensitivities and broad-bandwidth spectrometer capabilities to make large CO redshift surveys practical. The proposed 30 GHz spectrometer for the LMT would be an excellent redshift machine for the sub-mm population of galaxies.
The two best studied sub-mm galaxies (SMM J02399$``$0136 and SMM J14011+0252) share many of the same properties of the local population of ULIGs, such as high infrared luminosities, the association with mergers, massive molecular gas reservoirs, comparable CO line widths, and similar IR/radio and IR/CO luminosity ratios. Future CO observations of large samples of ultraluminous galaxies with ALMA and the LMT will enable us to study the evolution of the molecular gas properties as a function of redshift which will be crucial for our understanding of the star-formation history of the universe.
## Acknowledgments
I thank the work of my collaborators Nick Scoville, Rob Ivison, Ian Smail, Andrew Blain, Aaron Evans, Min Yun, and Jean-Paul Kneib. I appreciate the efforts of the OVRO and Keck staff who have made these observations a success. I acknowledge support from NSF grant AST 9981546 made to the OVRO Millimeter Array which is operated by the California Institute of Technology. I thank the organizers at the University of Massachusetts and INAOE for planning the conference and providing support to attend the conference.
|
warning/0007/math0007082.html
|
ar5iv
|
text
|
# New recursions for genus-zero Gromov-Witten invariants
## 0. Introduction
The localization theorem for equivariant cohomology has recently been used with great success to compute the genus-zero Gromov-Witten invariants relevant to the mirror conjecture . Zero-point invariants count expected numbers of rational curves on a projective manifold $`X`$, while the more general $`m`$-point invariants count expected numbers of rational curves meeting $`m`$ given submanifolds (or cohomology classes). For the the mirror conjecture, only the zero and one-point invariants are computed, though for the construction of the quantum product (even the small version), one needs more general invariants.
In this paper we will apply the localization theorem to study genus-zero Gromov-Witten invariants involving any number of marked points. A straightforward generalization of Givental’s (one variable) $`J`$-function yields homology-valued $`J`$-functions in any number of variables $`t_1`$,…,$`t_m`$ which encode all the (generalized) $`m`$-point genus-zero invariants. Our main theorem is a collection of relations among these $`J`$-functions expressing a part of the $`J`$-function for a fixed curve class and number of variables in terms of the $`J`$-functions involving fewer variables and/or “smaller” curve classes. When the cohomology of $`X`$ is generated by divisor classes, or, more generally, when every class orthogonal to the subring generated by divisor classes annihilates (via cap product) all one-variable $`J`$-functions, then these new relations completely determine all $`m`$-point genus-zero Gromov-Witten invariants (of classes generated by divisor classes) in terms of one-point invariants. That is, in this setting, the one-variable $`J`$-function determines all the others. A complete intersection in $`𝐏^n`$ has this orthogonality property, and in that case we exhibit a formula expressing “mixed” two-point invariants in terms of one-point invariants. We apply this new formula to compute previously unknown quantum products of cohomology classes on Fano complete intersections, where the one-variable $`J`$-function is known. Since our recursions do not require any positivity of $`X`$, they would apply just as well to general-type complete intersections. Unfortunately, in those cases, the one-variable $`J`$-function is not known.
The idea is the following. Given a stable map with several marked points, a copy of $`𝐏^1`$ is attached to each of the marked points (at $`0`$ on the $`𝐏^1`$). This allows one to embed the moduli space of $`m`$-pointed stable maps into the moduli space of stable maps with no marked points and $`m`$ parametrizations. We will call the latter space the graph space. There is a natural torus action on the graph space, one of whose fixed loci is the given moduli space of stable maps. There are equivariant forgetful morphisms among the graph spaces, and by comparing residues of some carefully chosen equivariant cohomology classes along the fixed loci, we obtain the recursive formulas for the $`J`$-functions. A startling (to us) feature of this approach is that it is much simpler than the computation of one-point invariants, since our argument requires no analysis of the boundary of graph spaces.
## 1. Kontsevich-Manin spaces
We recall the basic properties of the genus-zero stable map spaces and some results on Gromov-Witten invariants, and give an instance of our formula (to be proved in greater generality later).
###### Definition 1.
A morphism $`f:(C;p_1,\mathrm{},p_m)X`$ from a connected $`m`$-pointed complex rational curve $`C`$ to a complex projective manifold $`X`$ is prestable if $`C`$ has only nodes as singularities and $`p_1`$,…,$`p_mC`$ are nonsingular. If in addition every irreducible component of $`C`$ collapsed by $`f`$ has three or more distinguished points—a distinguished point is a node or marked point—we say that $`f`$ is stable.
###### Remark .
This notion of stability is analogous to Deligne-Mumford stability for pointed curves. Indeed, a stable map to a point is a stable pointed curve.
The moduli of stable maps has been extensively studied, ever since stable maps were introduced by Kontsevich and Manin . (See also and as good references for the following properties.)
Given $`\beta \mathrm{H}_2(X,𝐙)`$, there is a proper Deligne-Mumford stack $`\overline{M}_{0,m}(X,\beta )`$, representing flat families of genus-zero stable maps with $`m`$ marked points and image homology class $`\beta `$. Each of the moduli stacks $`\overline{M}_{0,m}(𝐏^n,d)`$ is smooth (as a stack) of the expected dimension $`n+(n+1)d+(m3)`$. For general $`X`$, there is a virtual class $`\left[\overline{M}_{0,m}(X,\beta )\right]^{\mathrm{vir}}`$ in the Chow group of $`\overline{M}_{0,m}(X,\beta )`$ of the expected dimension $`n\mathrm{deg}_{K_X}(\beta )+(m3)`$.
There are forgetful maps and evaluation maps:
$$\begin{array}{ccc}\overline{M}_{0,m+1}(X,\beta )& \stackrel{e_i}{}& X\\ \pi _i& & & & \\ \overline{M}_{0,m}(X,\beta )& & \end{array}$$
where $`\pi _i`$ “forgets” the marked point $`p_i`$ (and collapses components, if necessary), and $`e_i`$ evaluates the stable map at $`p_i`$. When $`i=m+1`$, this diagram can be taken as part of the “universal stable map” over $`\overline{M}_{0,m}(X,\beta )`$. The rest of the universal stable map consists of sections
$$\rho _i:\overline{M}_{0,m}(X,\beta )\overline{M}_{0,m+1}(X,\beta )$$
of $`\pi _{m+1}`$ corresponding to the marked points.
In case $`X𝐏^n`$ is the transverse zero locus of a section of a vector bundle $`E`$ on $`𝐏^n`$ which is generated by global sections, the refined top Chern class $`c_{\mathrm{top}}(\pi _{m+1}^{}{}_{}{}^{}e_{m+1}^{}E)`$ on $`\overline{M}_{0,m}(𝐏^n,d)`$ produces the virtual class on $`\overline{M}_{0,m}(X,d)`$.
Morphisms $`\varphi :XY`$ give rise to morphisms of stable map spaces
$$\overline{M}_{0,m}(X,\beta )\overline{M}_{0,m}(Y,\varphi _{}\beta ).$$
Finally, the “boundary” of $`\overline{M}_{0,m}(X,\beta )`$ is covered by the images of the gluing maps:
$$\delta _{S,\alpha }:\overline{M}_{0,k+1}(X,\alpha )\times _X\overline{M}_{0,mk+1}(X,\beta \alpha )\overline{M}_{0,m}(X,\beta )$$
where $`S\{1,\mathrm{},m\}`$ is a subset of cardinality $`k`$, which, together with the curve class $`\alpha `$, describes how the stable map breaks into (at least) two components.
The Gromov-Witten invariants are usually interpreted as intersection numbers on the Kontsevich-Manin spaces of stable maps. Given cohomology classes $`\gamma _1`$,…,$`\gamma _m`$ on $`X`$, one defines the “ordinary” invariants
$$\gamma _1,\mathrm{},\gamma _m_\beta ^X:=\mathrm{deg}(\pi _1^{}\gamma _1\mathrm{}\pi _m^{}\gamma _m\mathrm{ev}_{}\left[\overline{M}_{0,m}(X,\beta )\right]^{\mathrm{vir}}),$$
where
$$\mathrm{ev}:=(e_1,\mathrm{},e_m):\overline{M}_{0,m}(X,\beta )X^m$$
is the total evaluation map and $`\pi _i:X^mX`$ are the projections.
The general invariants are defined using the cotangent classes
$$\psi _i:=c_1(\rho _i^{}\omega _{\pi _{m+1}}),$$
where $`\omega _{\pi _{m+1}}`$ is the relative dualizing sheaf. The general invariants are:
$$\begin{array}{c}\gamma _1\psi ^{a_1},\mathrm{},\gamma _m\psi ^{a_m}_\beta ^X:=\hfill \\ \hfill \mathrm{deg}\left(\pi _1^{}\gamma _1\mathrm{}\pi _m^{}\gamma _m\mathrm{ev}_{}(\psi _1^{a_1}\mathrm{}\psi _m^{a_m}\left[\overline{M}_{0,m}(X,\beta )\right]^{\mathrm{vir}})\right),\end{array}$$
where $`a_1`$,…,$`a_m`$ are non-negative integers.
The following is a very useful way to package $`2`$-point “mixed” invariants:
$$\gamma _1,\frac{\gamma _2}{t\psi }_\beta ^X:=t^1\gamma _1,\gamma _2_\beta ^X+t^2\gamma _1,\gamma _2\psi _\beta ^X+t^3\gamma _1,\gamma _2\psi ^2_\beta ^X+\mathrm{},$$
where $`t`$ is a variable. Similarly, for general $`1`$-point invariants:
$$\frac{\gamma }{t(t\psi )}_\beta ^X:=t^2\gamma _\beta ^X+t^3\gamma \psi _\beta ^X+t^4\gamma \psi ^2_\beta ^X+\mathrm{}.$$
The simplest of our formulas is expressed in terms of these packages (extended $`t`$-linearly):
###### Formula 1.1.
Suppose $`X𝐏^n`$ is a complete intersection of dimension $`r3`$ and degree $`l`$. Then for all $`0a,bm`$ and $`d0`$,
$$\begin{array}{c}H^a,\frac{H^b}{t\psi }_d^X+\frac{H^a(Hdt)^b}{t(t\psi )}_d^X+\hfill \\ \hfill \underset{e=1}{\overset{d1}{}}\underset{c=0}{\overset{r}{}}\frac{1}{l}H^a,\frac{H^c}{t\psi }_{de}^X\frac{H^{rc}(Het)^b}{t(t\psi )}_e^X𝐐[t]\end{array}$$
.
This formula implies a special case of our reconstruction theorem 5.2:
###### Corollary 1.2.
The mixed two-point invariants of complete intersections in $`𝐏^n`$ involving only powers of $`H`$ are determined by the one-point invariants.
###### Proof.
The first term in the formula is clearly determined by the others. By induction on $`d`$, the mixed invariants of degree $`d`$ are therefore determined by the one-point invariants of degree $`d`$ or less. ∎
In the appendix, we use Formula 1.1 to compute small quantum products of cohomology classes on Fano complete intersections. For now, we point out the identities that follow from the formula when the classes (in the second slot) are of codimensions $`0`$, $`1`$ and $`2`$.
$`H^a,1_d^X=0`$.
$`H^a,\psi _1^X=H^a_d^X`$ and $`H^a,H_d^X=dH^a_d^X`$.
$`H^a,\psi ^2_d^X`$ $`=H^a\psi _d^X{\displaystyle \frac{1}{l}}{\displaystyle \underset{e=1}{\overset{d1}{}}}{\displaystyle \underset{c=0}{\overset{r}{}}}H^a,H^c_{de}^XH^{rc}_e^X`$
$`H^a,H\psi _d^X`$ $`=H^{a+1}_d^XdH^a\psi _d^X+{\displaystyle \frac{1}{l}}{\displaystyle \underset{e=1}{\overset{d1}{}}}{\displaystyle \underset{c=0}{\overset{r}{}}}eH^a,H^c_{de}^XH^{rc}_e^X`$
$`H^a,H^2_d^X`$ $`=2dH^{a+1}_d^X+d^2H^a\psi _d^X{\displaystyle \frac{1}{l}}{\displaystyle \underset{e=1}{\overset{d1}{}}}{\displaystyle \underset{c=0}{\overset{r}{}}}e^2H^a,H^c_{de}^XH^{rc}_e^X.`$
Notice that the codimension two identities are not self-contained, since they inductively involve classes of higher codimension.
###### Remark .
The codimension $`0`$ and the two codimension $`1`$ identities are special cases of the string, dilaton and divisor equations, respectively. The identities for codimension $`2`$ classes, however, are not special cases of any general equations that we are aware of (though we’ve been informed that Lee and Pandharipande have another method for producing such identities).
To state our main theorem, we will use $`J`$-functions of several variables, generalizing Givental’s one-variable definition.
###### Definition 2.
$`J_\beta ^X(t_1,\mathrm{},t_m)`$ $`:=\mathrm{ev}_{}\left({\displaystyle \frac{\left[\overline{M}_{0,m}(X,\beta )\right]^{\mathrm{vir}}}{t_1(t_1\psi _1)\mathrm{}t_m(t_m\psi _m)}}\right)`$
$`:=\mathrm{ev}_{}\left({\displaystyle \underset{i=1}{\overset{m}{}}}t_i^2\left(1+{\displaystyle \frac{\psi _i}{t_i}}+{\displaystyle \frac{\psi _i^2}{t_i^2}}+\mathrm{}\right)\left[\overline{M}_{0,m}(X,\beta )\right]^{\mathrm{vir}}\right)`$
$`\mathrm{H}_{}(X^m,𝐐)[t_1^1,\mathrm{},t_m^1]`$
with initial conditions:
$$J_0^X(t_1):=[X]$$
and
$$J_0^X(t_1,t_2):=\frac{\mathrm{\Delta }}{t_1t_2(t_1+t_2)},$$
where $`\mathrm{\Delta }\mathrm{H}_{}(X\times X,𝐐)`$ is the diagonal class.
###### Remark .
When $`m=1`$, our $`J`$-function is the Poincaré dual of Givental’s.
The $`J`$-functions encode all genus-zero Gromov-Witten invariants. The following result concerning the one-variable $`J`$-function was first proved in ; see and for alternate approaches.
###### Theorem 1.3 (Givental).
1. If $`X𝐏^n`$ is a complete intersection of type $`(l_1,\mathrm{},l_m)`$ which is Fano of index two or more (i.e. $`l_1+\mathrm{}+l_m<n`$), let $`H`$ be the hyperplane class. Then
$$J_d^X(t)=I_d^X(t):=\frac{_{i=1}^m_{k=1}^{dl_i}(l_iH+kt)}{_{k=1}^d(H+kt)^{n+1}}[X].$$
2. If $`l_1+\mathrm{}+l_m=n`$ or $`n+1`$, then the following generating functions coincide after an explicit “mirror transformation” (see , or ):
$$J^X(q):=\underset{d=0}{\overset{\mathrm{}}{}}J_d^X(t)q^d\text{and}I^X(q):=\underset{d=0}{\overset{\mathrm{}}{}}I_d^X(t)q^d.$$
We introduce a tool for manipulating $`J`$-functions of several variables:
###### Definition 3.
Given classes $`\mathrm{\Gamma }_1\mathrm{H}_{}(X^k\times X,𝐐)`$, $`\mathrm{\Gamma }_2\mathrm{H}_{}(X\times X^{mk},𝐐)`$ and $`\gamma \mathrm{H}^{}(X,𝐐)`$, we use the Künneth formula and Poincaré duality to regard the tensor product as a $`𝐐`$-linear map:
$$\mathrm{\Gamma }_1\mathrm{\Gamma }_2:\mathrm{H}^{}(X^2,𝐐)\mathrm{H}_{}(X^m,𝐐)$$
where the factors of $`X^2`$ are the distinguished factors of $`X^k\times X`$ and $`X\times X^{mk}`$. We then define the twisted product $`\mathrm{\Gamma }_1_\gamma \mathrm{\Gamma }_2\mathrm{H}_{}(X^m,𝐐)`$ by setting:
$$\mathrm{\Gamma }_1_\gamma \mathrm{\Gamma }_2:=(\mathrm{\Gamma }_1\mathrm{\Gamma }_2)(\delta \pi ^{}\gamma )$$
where $`\delta `$ is the diagonal class and $`\pi :X^2X`$ is either of the two projections.
###### Examples .
Let $`\gamma _1`$, $`\gamma _2`$ be Poincaré dual to $`\mathrm{\Gamma }_1`$, $`\mathrm{\Gamma }_2`$.
1. If $`k=m=0`$, then $`\mathrm{\Gamma }_1_\gamma \mathrm{\Gamma }_2𝐐`$ is the triple intersection:
$$_X\gamma \gamma _1\gamma _2.$$
More generally, if $`k=m`$, then $`\mathrm{\Gamma }_1_\gamma \mathrm{\Gamma }_2=\pi _{}^X(\pi _X^{}(\gamma \gamma _2)\mathrm{\Gamma }_1)`$, where $`\pi _X:X^m\times XX`$ and $`\pi ^X:X^m\times XX^m`$ are the two projections.
2. If $`k=m1`$, then $`\mathrm{\Gamma }_1_\gamma \mathrm{\Delta }=\pi _X^{}\gamma \mathrm{\Gamma }_1`$ for $`\pi _X:X^{m1}\times XX`$.
###### Theorem 1.4 (The main theorem—rank one case).
If $`H`$ is an ample divisor class generating $`\mathrm{H}^2(X,𝐐)`$ on a complex projective manifold $`X`$, then for each choice of $`m>0`$, $`d=\mathrm{deg}_H(\beta )0`$ and $`0bdim(X)`$,
$$\begin{array}{c}t(t_1+t)(\underset{1S[m]}{}\underset{e=0}{\overset{d}{}}J_{de}^X(\stackrel{}{t}_S,t)_{(Het)^b}J_e^X(t,\stackrel{}{t}_{S^c})+\hfill \\ \hfill \underset{j=2}{\overset{m}{}}J_d^X(\stackrel{}{t}_{\widehat{j}},t_j)_{H^b}J_0^X(t_j,t))\mathrm{H}_{}(X^m,𝐐)[t_1,t_1^1,\mathrm{},t_m,t_m^1,t].\end{array}$$
###### Notation .
We set $`[m]:=\{1,\mathrm{},m\}`$. For subsets $`S=\{s_1,\mathrm{},s_k\}[m]`$, we define $`\stackrel{}{t}_S:=(t_{s_1},\mathrm{},t_{s_k})`$. Since the $`J`$-functions are symmetric in their variables, the expressions $`J_d^X(\stackrel{}{t}_S,t)`$ are well-defined. We also set $`\widehat{j}:=[m]\{j\}`$.
We will prove the main theorem later, as well as a more general version where the rank one condition on $`\mathrm{H}^2(X,𝐐)`$ is removed. To finish this section, we show how Formula 1.1 follows from the main theorem.
###### Proof of Formula 1.1.
We apply the main theorem in the case $`m=1`$. Only the double sum appears, in this case as the single sum
(1)
$$\underset{e=0}{\overset{d}{}}J_{de}^X(t_1,t)_{(Het)^b}J_e^X(t).$$
The first and last of the terms are:
$$J_d^X(t_1,t)_{H^b}J_0^X(t)=J_d^X(t_1,t)_{H^b}[X]=\pi _{1}^{}{}_{}{}^{}(J_d^X(t_1,t)\pi _2^{}H^b)$$
and
$$J_0^X(t_1,t)_{(Hdt)^b}J_d^X(t)=\frac{\mathrm{\Delta }_{(Hdt)^b}J_d^X(t)}{t_1t(t_1+t)}=\frac{(Hdt)^bJ_d^X(t)}{t_1t(t+t_1)}.$$
Multiply (1) through by $`t_1t(t_1+t)`$, and the main theorem tells us we obtain an element of $`𝐐[t_1,t_1^1,t]`$ when we integrate against $`H^a`$. For example, by the projection formula, the first term gives
$$t_1t(t_1+t)\mathrm{deg}\left(H^a\pi _{1}^{}{}_{}{}^{}(\pi _2^{}H^bJ_d^X(t_1,t))\right)=(t_1+t)\frac{H^a}{t_1\psi },\frac{H^b}{t\psi }_d^X$$
and similarly for the other the terms. When we consider only the terms that are constant in $`t_1`$, we obtain the following formula:
$$\begin{array}{c}H^a,\frac{H^b}{t\psi }_d^X+\frac{H^a(Hdt)^b}{t(t\psi )}_d^X+\hfill \\ \hfill \underset{e=1}{\overset{d1}{}}\underset{i,j=1}{\overset{N}{}}H^a,\frac{\gamma _i}{t\psi }_{de}^Xg^{ij}\frac{\gamma _j(Het)^b}{t(t\psi )}_e^X𝐐[t],\end{array}$$
where $`\gamma _1,\mathrm{},\gamma _N\mathrm{H}^{}(X,𝐐)`$ are a basis, with respect to which $`g^{ij}`$ is the inverse of the intersection matrix. This much holds for any ample $`H`$ generating $`\mathrm{H}^2(X,𝐐)`$.
The fact that $`X`$ is a complete intersection tells us that all the one-point invariants of the form $`\gamma \psi ^c_\beta ^X`$ vanish when $`\gamma `$ is a primitive cohomology class. This can either be seen using Givental’s formulas, or by a monodromy argument. Since a basis for the cohomology may be chosen consisting of powers of $`H`$ and (orthogonal) primitive classes, this tells us that we may replace the basis $`\{\gamma _i\}`$ by the smaller set $`\{H^c\}`$ of powers of $`H`$, resulting in Formula 1.1. ∎
## 2. Graph spaces
Graph spaces are particular Kontsevich-Manin spaces which come equipped with a natural torus action. In this section, will describe some of the fixed loci under this torus action in order to eventually apply the Atiyah-Bott localization theorem to prove the main theorem.
###### Definition 4.
The $`m`$-parametrized graph space is
$$\overline{G}_{0,m}(X,\beta ):=\overline{M}_{0,0}(X\times (𝐏^1)^m,(\beta ,1^m)).$$
It is often useful to think of the graph space in the following way. Let a parametrization of a rational curve $`C`$ be an isomorphism from $`𝐏^1`$ to one of the irreducible components of $`C`$. A morphism $`f:(C;𝐏_1^1,\mathrm{},𝐏_m^1)X`$ from a connected rational curve with $`m`$ parametrizations isprestable if $`C`$ has only nodes as singular points. If in addition every component of $`C`$ is either parametrized (possibly in several ways) or has at least three nodes (or both), we say the $`f`$ is stable.
Then $`\overline{G}_{0,m}(X,\beta )`$ is the moduli stack of stable maps with $`m`$ parametrizations and no marked points. This stack admits the action of the torus $`(𝐂^{})^m`$ via its action on $`(𝐏^1)^m`$. To carefully give this torus action, we need some more precise notation.
Fix vector spaces $`W_i𝐂^2`$ for $`i=1`$,…,$`m`$. On $`W_i`$, choose coordinates $`x_i,y_iW_i^{}`$, and fix the action of $`𝐂^{}=𝐂_i^{}`$ via
$$\mu _i(x_i,y_i)=(x_i,\mu _iy_i).$$
Let $`0_i:=(0:1)`$ and $`\mathrm{}_i:=(1:0)𝐏_i^1`$ be the fixed points of the action of $`𝐂_i^{}`$ on $`𝐏_i^1=𝐏(W_i)`$. Let $`𝐓:=𝐂_i^{}`$ acting diagonally on $`𝐏_i^1`$ and hence on each of the graph spaces $`\overline{G}_{0,m}(X,\beta )`$.
###### Important Special Case .
The two-parametrized graph space of a point
$$\overline{G}_{0,2}(\mathrm{pt},0)=\overline{M}_{0,0}(𝐏^1\times 𝐏^1,(1,1))𝐏^3.$$
A stable map to $`𝐏^1\times 𝐏^1`$ either embeds $`C`$ as a smooth curve of type $`(1,1)`$ or as a pair of intersecting rulings. Thus the stable map space is the linear series.
###### Observation .
A stable parametrized map $`[f]\overline{G}_{0,m}(X,\beta )`$ is a fixed point for the action of $`𝐓`$ described above exactly when:
* $`f`$ is constant on each parametrized component
* Each parametrized component is uniquely parametrized
* Each node on a parametrized component is at $`0`$ or $`\mathrm{}`$.
For our purposes, we will only need to consider the following “types” of fixed loci for the action of $`𝐓=𝐂_i^{}`$ on the graph space $`\overline{G}_{0,m+1}(X,\beta )`$:
### Type 1
A single copy of $`\overline{M}_{0,m+1}(X,\beta )`$ “embedded at zeroes”. (See Figure 1.)
Let $`Y=X\times _{i=1}^{m+1}𝐏_i^1`$, and consider the gluing morphism:
$$\overline{M}_{0,m+1}(Y,\beta )\times _Y\underset{i=1}{\overset{m+1}{}}\overline{M}_{0,1}(Y,1_i)\overline{G}_{0,m+1}(X,\beta ),$$
where $`\beta =(\beta ,0^{m+1})\mathrm{H}_2(Y,𝐙)`$ and likewise for $`1_i`$. Each $`\overline{M}_{0,1}(Y,1_i)Y`$ and $`\overline{M}_{0,m+1}(Y,\beta )\overline{M}_{0,m+1}(X,\beta )\times 𝐏_i^1`$, and we obtain a regular embedding
$$i_{[m],\beta }:F_{[m],\beta }:=\overline{M}_{0,m+1}(X,\beta )\overline{G}_{0,m+1}(X,\beta )$$
by embedding $`\overline{M}_{0,m+1}(X,\beta )\times 0_i\overline{M}_{0,m+1}(X,\beta )\times 𝐏_i^1`$ and using the gluing morphism above to further embed in the graph space.
### Type 2
$`\overline{M}_{0,k+1}(X,\alpha )\times _X\overline{M}_{0,mk+1}(X,\beta \alpha )`$ “with $`𝐏_{m+1}^1`$ in the middle.” (Copies indexed by subsets $`1S[m]`$ with $`|S|=k`$ and $`\alpha \mathrm{H}_2(X,𝐙)`$). (See Figure 2.)
In this case, we consider the composition of gluing maps taking:
$$\begin{array}{c}\underset{s_iS}{}\overline{M}_{0,1}(Y,1_i)\times _Y\overline{M}_{0,k+1}(Y,\alpha )\times _Y\overline{M}_{0,2}(Y,1_{m+1})\times _Y\hfill \\ \hfill \times _Y\overline{M}_{0,mk+1}(Y,\beta \alpha )\times _Y\underset{s_i^cS^c}{}\overline{M}_{0,1}(Y,1_i)\overline{G}_{0,m}(X,\beta ),\end{array}$$
assuming $`(S,\alpha )(\{1\},0)`$, $`([m],\beta )`$ or any $`(\widehat{j},\beta )`$. (These appear as other types!) The product is isomorphic to
$$\underset{s_iS}{}𝐏_{s_i}^1\times 𝐏_{m+1}^1\times \overline{M}_{0,k+1}(X,\alpha )\times _X\overline{M}_{0,mk+1}(X,\beta \alpha )\times 𝐏_{m+1}^1\times \underset{s_i^c}{}𝐏_{s_i^c}^1$$
and we identify the embedding:
$$i_{S,\alpha }:F_{S,\alpha }=\overline{M}_{0,k+1}(X,\alpha )\times _X\overline{M}_{0,mk+1}(X,\beta \alpha )\overline{G}_{0,m+1}(X,\beta )$$
with $`(0_S,0_{m+1})\times \overline{M}_{0,k+1}(X,\alpha )\times _X\overline{M}_{0,mk+1}(X,\beta \alpha )\times (\mathrm{}_{m+1},0_{S^c})`$.
### Type 3
$`\overline{M}_{0,m}(X,\beta )`$ “with $`𝐏_{m+1}^1`$ in various places” (three subtypes).
1. $`𝐏_1^1`$ as a tail off of $`𝐏_{m+1}^1`$ (a single copy). Let
$$i_{\{1\},0}:F_{\{1\},0}:=\overline{M}_{0,m}(X,\beta )\overline{G}_{0,m+1}(X,\beta )$$
be the embedding associated to Figure 3.
2. $`𝐏_j^1`$ as a tail off of $`𝐏_{m+1}^1`$ (one for each $`1<jm`$). Let
$$i_{\widehat{j},\beta }:F_{\widehat{j},\beta }:=\overline{M}_{0,m}(X,\beta )\overline{G}_{0,m+1}(X,\beta )$$
be the embedding associated to Figure 4.
3. $`𝐏_{m+1}^1`$ as a tail off of $`𝐏_j^1`$ (indexed by $`1<jm`$). Let
$$i_j:F_j:=\overline{M}_{0,m}(X,\beta )\overline{G}_{0,m+1}(X,\beta )$$
be the embedding associated to Figure 5.
###### Lemma 2.1.
There is a $`𝐓`$-equivariant birational morphism:
$$\mathrm{\Phi }:\overline{G}_{0,m+1}(X,\beta )\overline{G}_{0,m}(X,\beta )\times 𝐏^3$$
which (when projected onto the first factor) forgets the last parametrization and (when projected onto the second factor) forgets the map to $`X`$ and all parametrizations except for the first and last.
###### Proof.
The existence of $`\mathrm{\Phi }`$ follows from the functoriality of Kontsevich-Manin spaces. The two projections are just the two maps:
$$\overline{M}_{0,0}(X\times \underset{i=1}{\overset{m+1}{}}𝐏_i^1,(\beta ,1^{m+1}))\overline{M}_{0,0}(X\times \underset{i=1}{\overset{m}{}}𝐏_i^1,(\beta ,1^m))$$
and
$$\overline{M}_{0,0}(X\times \underset{i=1}{\overset{m+1}{}}𝐏^1e_i,(\beta ,1^{m+1}))\overline{M}_{0,0}(𝐏_1^1\times 𝐏_{m+1}^1,(1,1))$$
which clearly commute with the action of $`𝐓`$.
Over the open subset of $`𝐏^3`$ consisting of smooth curves, $`\mathrm{\Phi }`$ is an isomorphism. The last parametrization is of the same component as the first, and is given by the correspondence $`𝐏_1^1\stackrel{}{}𝐏_{m+1}^1`$ induced by the curve in $`𝐏_1^1\times 𝐏_{m+1}^1`$. ∎
###### Lemma 2.2.
Let $`(0,0)\overline{G}_{0,2}(\mathrm{pt},0)=𝐏^3`$ be the fixed point corresponding to the “coordinate axes.” Then the embeddings of Types 1–3 listed above are a complete list of the fixed loci that are contained in
$$\mathrm{\Phi }^1\left(F_{[m1],\beta }\times (0,0)\right)\overline{G}_{0,m+1}(X,\beta ).$$
Moreover, the induced maps $`\mathrm{\Phi }|_F:FF_{[m1],\beta }\overline{M}_{0,m}(X,\beta )`$ are:
(Type 1) $`\pi _{m+1}:\overline{M}_{0,m+1}(X,\beta )`$ $`\overline{M}_{0,m}(X,\beta )`$
(Type 2) $`\delta _{S,\alpha }:\overline{M}_{0,k+1}(X,\alpha )\times _X\overline{M}_{0,mk+1}(X,\beta \alpha )`$ $`\overline{M}_{0,m}(X,\beta )`$
(Type 3) $`\overline{M}_{0,m}(X,\beta )`$ $`\stackrel{id}{}\overline{M}_{0,m}(X,\beta ).`$
###### Proof.
Given a stable map $`f:CX`$, represent $`C`$ by a tree with vertices and edges corresponding to the nodes and components of $`C`$, respectively. For an $`f\overline{G}_{0,m+1}(X,\beta )`$ to map to $`F_{[m1],\beta }\overline{G}_{0,m}(X,\beta )`$ under the forgetful map, each $`𝐏_i^1`$ must parametrize a different curve (edge of the tree) mapping with degree $`0`$ to $`X`$, and each $`0_i`$ must be a node (vertex) for $`i=1`$,…,$`m`$. Also, the shortest path between two such vertices of the tree cannot contain any such edges, and if one of those edges is removed, the tree either stays connected or it has two components, one of which is an edge corresponding to $`𝐏_{m+1}^1`$ mapping with degree $`0`$ to $`X`$. In order for $`f`$ to map to $`(0,0)`$ under the other forgetful map to $`𝐏^3`$, $`𝐏_1^1`$ and $`𝐏_{m+1}^1`$ must represent different edges, $`0_1`$ and $`0_{m+1}`$ must represent vertices (possibly the same one) and the shortest path from $`0_1`$ to $`0_{m+1}`$ may not contain either of the two edges.
The only fixed points under the torus action which satisfy both conditions are those of types 1–3. This proves the first part of the lemma.
Under the map to $`F_{[m1],\beta }`$, the parametrization of $`𝐏_{m+1}^1`$ is forgotten. This may result in an unparametrized component with $`1`$ or $`2`$ nodes, which is then collapsed. Moreover, in the $`1`$ node case, the resulting marked point must also be forgotten. This gives the second part of the lemma. ∎
## 3. Localization and the main theorem for $`𝐏^n`$.
When a complex Lie group $`G`$ acts on a complex manifold $`X`$, there is an equivariant cohomology ring:
$$H_G^{}(X,𝐐)$$
which is an algebra over the cohomology ring of the classifying space $`BG`$. If $`G=𝐓=(𝐂^{})^m`$, then $`\mathrm{H}^{}(BG,𝐐)𝐐[t_1,\mathrm{},t_m]`$. The equivariant cohomology ring for a trivial action of $`𝐓`$ is the polynomial algebra $`\mathrm{H}^{}(X,𝐐)[t_1,\mathrm{},t_m]`$, but in general it is more complicated. Linearized vector bundles $`E`$ on $`X`$ have equivariant Chern classes $`c_d^G(E)`$ taking values in equivariant cohomology, and equivariant cohomology pulls back and pushes forward (for proper maps).
The set-up is similar for smooth Deligne-Mumford stacks. In this setting, there is an equivariant Chow ring (see ) $`\mathrm{CH}_G^{}(X)`$ which always pulls back and pushes forward under equivariant proper maps. A linearized vector bundle $`E`$ in this setting has equivariant Chern classes $`c_d^G(E)\mathrm{CH}_G^{}(X)`$.
A basic result in either setting is the theorem of Atiyah-Bott (see ):
###### Theorem 3.1 (Localization).
Each fixed substack $`i:FX`$ of a torus action on a proper smooth Deligne-Mumford stack is a regularly embedded proper smooth Deligne-Mumford stack, its normal bundle is canonically linearized, its Euler class $`\epsilon _𝐓(F)`$ (the top equivariant Chern class of the normal bundle) is invertible in
$$\mathrm{CH}^{}(F,𝐐)_𝐐𝐐(t_1,\mathrm{},t_m),$$
and any element $`c\mathrm{CH}_𝐓^{}(X)`$ is uniquely recovered (modulo torsion) via the following *localization formula:*
$$c=\underset{F}{}i_{}\frac{i^{}c}{\epsilon _𝐓(F)}.$$
Our main interest is in the following simple corollary:
###### Corollary 3.2 (Correspondence of residues).
Suppose $`f:XX^{}`$ is a $`𝐓`$-equivariant map of smooth proper Deligne-Mumford stacks with $`𝐓`$-actions. If $`i^{}:F^{}X^{}`$ is a fixed substack and $`c\mathrm{CH}_𝐓^{}(X)`$, let $`f_F:FF^{}`$ be the restriction of $`f`$ to each of the fixed substacks $`Ff^1(X^{})`$. Then
$$\underset{Ff^1(F^{})}{}f_{F}^{}{}_{}{}^{}\frac{i^{}c}{\epsilon _𝐓(F)}=\frac{i_{}^{}{}_{}{}^{}f_{}c}{\epsilon _𝐓(F^{})}$$
###### Proof.
The two sides of the formula represent the contribution of $`F^{}`$ to localization formulas for $`f_{}c`$ which, by uniqueness, must coincide. ∎
To prove the main theorem for general $`X`$ we will need virtual classes. For now we will prove it in the case $`X=𝐏^n`$, where the basic idea and most of the computations are the same, and are not obscured by the presence of virtual classes.
###### Proof of the main theorem for $`𝐏^n`$.
Let $`c\mathrm{CH}_𝐓^{}(\overline{G}_{0,m+1}(𝐏^n,d))`$. Then applying correspondence of residues to the map $`\mathrm{\Phi }`$ of smooth Deligne-Mumford stacks (here we use $`X=𝐏^n`$) of Lemma 2.1, and using the enumeration of fixed loci in Lemma 2.2, we get:
(2)
$$\begin{array}{c}\pi _{m+1}^{}{}_{}{}^{}\left(\frac{i_{[m],d}^{}c}{\epsilon _𝐓(F_{[m],d})}\right)+\underset{1S}{}\underset{e=0}{\overset{d}{}}\delta _{S,e}^{}{}_{}{}^{}\left(\frac{i_{S,de}^{}c}{\epsilon _𝐓(F_{S,de})}\right)+\frac{i_{\{1\},0}^{}c}{\epsilon _𝐓(F_{\{1\},0})}+\hfill \\ \hfill \underset{j=2}{\overset{m}{}}\left(\frac{i_{\widehat{j},d}^{}c}{\epsilon _𝐓(F_{\widehat{j},d})}+\frac{i_j^{}c}{\epsilon _𝐓(F_j)}\right)=\frac{i_{[m1],d}^{}\mathrm{\Phi }_{}c}{\epsilon _𝐓(F_{[m1],d})}\frac{1}{t_1t_{m+1}(t_1+t_{m+1})}.\end{array}$$
(The computation $`\epsilon _𝐓(0,0)=t_1t_{m+1}(t_1+t_{m+1})`$ is easily made.)
Now, the equivariant Euler classes $`\epsilon _𝐓(F)`$ appearing in the denominators depend entirely on the nodes of the domain of a general representative $`fF`$. Essentially, there are two types of nodes: those of type I, where at the point $`p_i\{0_i,\mathrm{}_i\}`$, the $`i`$th parametrized component meets a component mapping in positive degree $`\alpha `$ to $`X`$, and those of type II, where at the point $`p_i\{0_i,\mathrm{}_i\}`$, the $`i`$th parametrized component meets the point $`p_j\{0_j,\mathrm{}_j\}`$ of the $`j`$th parametrized component. See Figure 6.
Any type I node is a codimension $`2`$ condition—one for the node and one for specifying $`p_i`$—while a type II node is a codimension $`3`$ condition—one for the node and two more for specifying $`p_i`$ and $`p_j`$. Set
$$\nu _i:=\{\begin{array}{cc}1\hfill & \text{if }p_i=0_i\hfill \\ 1\hfill & \text{if }p_i=\mathrm{}_i\hfill \end{array}.$$
Then the type I node contributes the factor
$$\nu _it_i(\nu _it_i\psi _i)$$
to $`\epsilon _𝐓(F)`$, while the type II node contributes
$$\nu _it_i\nu _jt_j(\nu _it_i+\nu _jt_j).$$
(For this type of computation, see .)
Thus, if we let
$$t:=t_{m+1},$$
the following computations are valid on any $`\overline{G}_{0,m+1}(X,\beta )`$:
(3.1) $`\epsilon _𝐓(F_{[m],\beta })`$ $`=t(t\psi _{m+1}){\displaystyle \underset{i[m]}{}}t_i(t_i\psi _i).`$
(3.2) $`\epsilon _𝐓(F_{S,\alpha })`$ $`=t(t\psi _{k+1})(t)(t\psi _1^{}){\displaystyle \underset{S}{}}t_{s_i}(t_{s_i}\psi _i){\displaystyle \underset{S^c}{}}t_{s_i^c}(t_{s_i^c}\psi _{i+1}^{})`$
where $`\psi _i,\psi _i^{}`$ are the cotangent classes on $`\overline{M}_{0,k+1}(X,\alpha )`$ and $`\overline{M}_{0,mk+1}(X,\beta \alpha )`$.
(3.3a) $`\epsilon _𝐓(F_{\{1\},0})`$ $`=t_1t(t_1+t)(t)(t\psi _1){\displaystyle \underset{i=2}{\overset{m}{}}}t_i(t_i\psi _i).`$
(3.3b) $`\epsilon _𝐓(F_{\widehat{j},\beta })`$ $`=(t)t_j(t+t_j){\displaystyle \underset{S=\widehat{j}}{}}t_{s_i}(t_{s_i}\psi _i)t(t\psi _m).`$
(3.3c) $`\epsilon _𝐓(F_j)`$ $`=(t_j)t(t_j+t){\displaystyle \underset{S=\widehat{j}}{}}t_{s_i}(t_{s_i}\psi _i)t_j(t_j\psi _m).`$
Finally, let $`𝐏^n=𝐏(V)`$ and consider $`H_𝐓`$, the equivariant hyperplane class on the linear space $`𝐏(\mathrm{Hom}(\mathrm{Sym}^d(W_{m+1}),V))`$. There is an equivariant morphism
$$v:\overline{G}_{0,m+1}(𝐏^n,d)\overline{G}_{0,1}(𝐏^n,d)𝐏(\mathrm{Hom}(\mathrm{Sym}^d(W_{m+1}),V))$$
which is a composition of the forgetful map remembering only $`𝐏^n`$ and the last parametrization, and the “map to the linear sigma model.” (The geometry of this second (birational) map was used in to give a proof of the mirror theorem.) This $`H_𝐓`$ pulls back to the fixed loci as follows:
(4.1) $`i_{[m],d}^{}v^{}H_𝐓`$ $`=e_{m+1}^{}H`$
(4.2) $`i_{S,de}^{}v^{}H_𝐓`$ $`=e_{k+1}^{}Het`$
(4.3a) $`i_{\widehat{j},d}^{}v^{}H_𝐓`$ $`=e_j^{}H`$
(4.3b) $`i_{\{1\},0}^{}v^{}H_𝐓`$ $`=e_1^{}Hdt`$
(4.3c) $`i_j^{}v^{}H_𝐓`$ $`=e_j^{}H.`$
To see this, note that under the morphism $`v`$, the fixed loci map to various copies of $`𝐏^n`$ sitting as the fixed loci in $`𝐏(\mathrm{Hom}(\mathrm{Sym}^d(W_{m+1}),V))`$. More specifically, set
$$\begin{array}{c}(𝐏^n)_e:=\{x_{m+1}^{de}y_{m+1}^e\}\times 𝐏(V)𝐏\left(\mathrm{Sym}^d(W_{m+1}^{})\right)\times 𝐏(V)\stackrel{Segre}{}\hfill \\ \hfill 𝐏(\mathrm{Hom}(\mathrm{Sym}^d(W_{m+1}),V))\end{array}$$
which are all fixed under the $`𝐓`$ action. One easily computes that $`H_𝐓`$ restricts to $`\mathrm{H}_𝐓^{}((𝐏^n)_e,𝐐)`$ as $`Het`$. Finally, fixed loci of types 1, 3a, and 3c map under $`v`$ to $`(𝐏^n)_0`$, the fixed loci of type 3b map to $`(𝐏^n)_d`$, and the loci of type 2 map to $`(𝐏^n)_e`$ for appropriate $`1ed1`$.
Substitute in the summands of (2) for the equivariant Euler classes and for the choice $`c=v^{}H_𝐓^b`$ and push forward under the total evaluation map $`\mathrm{ev}`$. Then by the projection formula and the computations above we obtain the following:
(5.1) $`\mathrm{ev}_{}\pi _{m+1}^{}{}_{}{}^{}{\displaystyle \frac{i_{[m],d}^{}c}{\epsilon _𝐓(F_{[m],d})}}`$ $`=J_d^{𝐏^n}(t_1,\mathrm{},t_m,t)_{H^b}J_0^{𝐏^n}(t)`$
(5.2) $`\mathrm{ev}_{}\delta _{S,\alpha }^{}{}_{}{}^{}{\displaystyle \frac{i_{S,\alpha }^{}c}{\epsilon _𝐓(F_{S,\alpha })}}`$ $`=J_{de}^{𝐏^n}(\stackrel{}{t}_S,t)_{(Het)^b}J_e^{𝐏^n}(t,\stackrel{}{t}_{S^c})`$
(5.3a) $`\mathrm{ev}_{}{\displaystyle \frac{i_{\{1\},0}^{}c}{\epsilon _𝐓(F_{\{1\},0})}}`$ $`=J_0^{𝐏^n}(t_1,t)_{(Hdt)^b}J_d^{𝐏^n}(t,\stackrel{}{t}_{\widehat{1}})`$
(5.3b) $`\mathrm{ev}_{}{\displaystyle \frac{i_{\widehat{j},d}^{}c}{\epsilon _𝐓(F_{\widehat{j},d})}}`$ $`=J_d^{𝐏^n}(\stackrel{}{t}_{\widehat{j}},t)_{H^b}J_0^{𝐏^n}(t,t_j)`$
(5.3c) $`\mathrm{ev}_{}{\displaystyle \frac{i_j^{}c}{\epsilon _𝐓(F_j)}}`$ $`=J_d^{𝐏^n}(\stackrel{}{t}_{\widehat{j}},t_j)_{H^b}J_0^{𝐏^n}(t_j,t)`$
We get the theorem by multiplying both sides of (2) by $`t_1t(t_1+t)`$, collecting types 1, 2, 3a, and 3b under the double sum, and noting that
$$\mathrm{ev}_{}\frac{i_{[m1],d}^{}\mathrm{\Phi }_{}c}{\epsilon _𝐓(F_{[m1],d})}\mathrm{H}^{}((𝐏^n)^m,𝐐)[t_1,t_1^1,\mathrm{},t_m,t_m^1,t]$$
because $`i_{[m1],d}^{}\mathrm{\Phi }_{}c`$ is polynomial in $`t_1`$,…,$`t_m,t`$ and the inverse to $`\epsilon _𝐓(F_{[m1],d})`$ belongs to $`\mathrm{H}^{}(\overline{M}_{0,m}(𝐏^n,d),𝐐)[t_1^1,\mathrm{},t_m^1]`$. ∎
## 4. Virtual classes
In order to prove our main theorem in the general case, we will need to establish a simple property of equivariant virtual classes. We begin by briefly recalling the construction of the virtual class on Kontsevich-Manin spaces, following Behrend and Fantechi (but see also Li-Tian ) and of the equivariant virtual class on graph spaces, following Graber-Pandharipande .
Fix a complex projective manifold $`X`$ and an embedding $`X𝐏^n`$. For each $`\beta \mathrm{H}_2(X,𝐙)`$, let $`d`$ be the degree of the image of $`\beta `$ in $`𝐏^n`$. Then there is a commuting diagram of stacks:
$$\begin{array}{ccccc}\overline{M}_{0,m}(X,\beta )& & \stackrel{𝑖}{}& & \overline{M}_{0,m}(𝐏^n,d)\\ & & & \rho \\ & & 𝔐_{0,m}\end{array}$$
where $`𝔐_{0,m}`$ is the Artin (not Deligne-Mumford!) stack of prestable $`m`$-marked curves. The map $`i`$ is a closed embedding (let $``$ be the associated ideal sheaf) and $`\rho `$ is smooth. It follows that the relative intrinsic normal cone of Behrend-Fantechi is the cone stack associated to the following map of sheaves on $`\overline{M}_{0,m}(X,\beta )`$:
$$/^2i^{}\mathrm{\Omega }_{\overline{M}_{0,m}(𝐏^n,d)/𝔐_{0,m}}$$
This relative normal cone, which we denote by $`_{\overline{M}_{0,m}(X,\beta )/𝔐_{0,n}}`$, embeds in the smooth $`h^1/h^0`$ cone stack $`V_{\overline{M}_{0,m}(X,\beta )/𝔐_{0,m}}`$ associated to the object
$$(R^1\pi _{}e^{}TX)^{}$$
of the derived category of coherent sheaves on $`\overline{M}_{0,m}(X,\beta )`$. (The dual is the Verdier dual and $`\pi :𝒞\overline{M}_{0,m}(X,\beta )`$ and $`e:𝒞X`$ come from the universal stable map). The virtual class $`\left[\overline{M}_{0,m}(X,\beta )\right]^{\mathrm{vir}}`$ is then obtained by pulling back the class of $`_{\overline{M}_{0,m}(X,\beta )/𝔐_{0,m}}`$ via the zero section of $`V_{\overline{M}_{0,m}(X,\beta )/𝔐_{0,m}}`$.
Similarly, for graph spaces, there is a diagram of $`𝐓`$-invariant morphisms
$$\begin{array}{ccccc}\overline{G}_{0,m}(X,\beta )& & \stackrel{𝑖}{}& & \overline{G}_{0,m}(𝐏^n,d)\\ & & & \rho \\ & & 𝔊_{0,m}\end{array},$$
where $`𝔊_{0,m}`$ is the stack of prestable zero-pointed maps to $`(𝐏^1)^m`$ of multi-degree $`(1,\mathrm{},1)`$. The (equivariant) intrinsic relative normal cone $`_{\overline{G}_{0,m}(X,\beta )/𝔊_{0,m}}`$ and $`h^1/h^0`$ cone $`V_{\overline{G}_{0,m}(X,\beta )/𝔊_{0,m}}`$ are defined exactly as before, and the (equivariant) virtual class $`\left[\overline{G}_{0,m}(X,\beta )\right]_𝐓^{\mathrm{vir}}`$ may also be defined as before, using equivariant Chow groups. This definition is simpler than the definition in , but is equivalent. The simplification in our case comes from the existence of the $`𝐓`$-invariant embedding $`i`$ into a relatively smooth graph space. The significance of the equivariant virtual class is in the following “virtual” version of the localization theorem:
###### Theorem 4.1 (Graber-Pandharipande).
In the equivariant Chow group of the graph space $`\overline{G}_{0,m}(X,\beta )`$ the virtual class satisfies
$$\left[\overline{G}_{0,m}(X,\beta )\right]_𝐓^{\mathrm{vir}}=\underset{F}{}i_{}\frac{i^{}\left[\overline{G}_{0,m}(X,\beta )\right]_𝐓^{\mathrm{vir}}}{\epsilon _𝐓(F)}$$
where $`i:F\overline{G}_{0,m}(X,\beta )`$ are the (regular) embeddings of the fixed substacks.
In order to use this theorem, we need the following:
###### Lemma 4.2.
1. The forgetful map
$$\varphi :\overline{G}_{0,m}(X,\beta )\overline{M}_{0,0}(X,\beta )$$
is flat and equivariant for the trivial action of $`𝐓`$ on $`\overline{M}_{0,0}(X,\beta )`$.
2. The equivariant virtual class satisfies
$$\left[\overline{G}_{0,m}(X,\beta )\right]_𝐓^{\mathrm{vir}}=\varphi ^{}\left[\overline{M}_{0,0}(X,\beta )\right]^{\mathrm{vir}},$$
where $`\left[\overline{M}_{0,0}(X,\beta )\right]^{\mathrm{vir}}`$ is the ordinary virtual class, regarded as an equivariant class for the trivial action of $`𝐓`$. In particular, each $`i^{}\left[\overline{G}_{0,0}(X,\beta )\right]_𝐓^{\mathrm{vir}}=\left[F\right]^{\mathrm{vir}}`$ in the theorem above, where $`\left[F\right]^{\mathrm{vir}}`$ is the “ordinary” virtual class on $`F`$, thought of as a fiber product of Kontsevich-Manin spaces.
###### Proof.
It suffices by induction to prove the lemma for the case $`m=1`$. In that case, we will consider a (non-commuting!) diagram of stacks:
$$\begin{array}{ccccc}\overline{M}_{0,3}(X,\beta )& \stackrel{g}{}& \overline{G}_{0,1}(X,\beta )& \stackrel{\varphi }{}& \overline{M}_{0,0}(X,\beta )\\ & & & & & & \\ 𝔐_{0,3}& \stackrel{𝔤}{}& 𝔊_{0,1}& & \end{array}$$
where the horizontal maps are the “cross-ratio” maps defined as follows. The universal curve $`𝒞\overline{M}_{0,4}(X,\beta )`$ over $`\overline{M}_{0,3}(X,\beta )`$ maps to $`\overline{M}_{0,4}𝐏^1`$ via the forgetful map. Together with the evaluation map to $`X`$, this defines $`g`$. If $`f:CX`$ is a stable map with $`3`$ marked points $`p`$, $`q`$, $`rC`$, then the map $`C𝐏^1`$ defined by $`g(f)`$ may be taken to be the unique map with the property that $`f(p)=0,f(q)=1`$ and $`f(r)=\mathrm{}`$. This is the cross-ratio if $`p`$, $`q`$, $`r`$ belong to the same component of $`C`$, but is well-defined even if they lie on different components.
For $`𝔤`$, we apply the prestabilization map $`𝒞𝔐_{0,4}`$ (see ) to the universal curve over $`𝔐_{0,3}`$ followed by the stabilization map $`𝔐_{0,4}\overline{M}_{0,4}𝐏^1`$. This map has the same pointwise description as $`g`$.
The diagram doesn’t commute because $`g`$ stabilizes unstable maps to $`X\times 𝐏^1`$, while $`𝔤`$ does not. On the other hand, there is a “good” open substack
$$U:=\{f:C𝐏^1f\text{ is an isomorphism over }0\text{}1\text{}\mathrm{}\}𝔊_{0,1}$$
with the following properties:
* $`g`$ and $`𝔤`$ are both isomorphisms over $`U`$.
* The diagram above is Cartesian when restricted to $`U`$.
* Translates of $`U`$ by elements $`m\mathrm{PGL}(2,𝐂)`$ cover $`𝔊_{0,1}`$,
If $`fU`$, then $`p`$, $`q`$, and $`r`$ are the preimages of $`0`$, $`1`$, and $`\mathrm{}`$, so $`𝔤`$ is invertible at $`f`$. If $`f\overline{M}_{0,3}(X,\beta )`$ lies over $`U`$, then $`p`$, $`q`$, and $`r`$ all belong to same component $`C_0C`$ of the curve associated to $`f`$, and $`g(f)`$ imposes the unique parametrization on $`C_0`$ taking $`p`$, $`q`$, and $`r`$ to $`0`$, $`1`$, and $`\mathrm{}`$. Clearly, then, $`g`$ and $`𝔤`$ are isomorphisms over $`U`$ and the diagram is Cartesian over $`U`$. Since every prestable map $`f:C𝐏^1`$ of degree one is generically an isomorphism over $`𝐏^1`$, it follows that the translates of $`U`$ cover $`𝔊_{0,1}`$.
We finish the proof now by comparing $`\overline{G}_{0,1}(X,\beta )`$ with $`\overline{M}_{0,3}(X,\beta )`$. Suppose $`f\overline{G}_{0,1}(X,\beta )`$ lies over $`U`$. Then $`g`$ is an isomorphism at $`f`$, so since $`\varphi g`$ is flat everywhere (it is a composition of the flat forgetful maps), it follows that $`\varphi `$ is flat at $`f`$. But an arbitrary $`f\overline{G}_{0,1}(X,\beta )`$ lies over some translate $`mU`$, over which the composition of $`g`$ with translation by $`m`$ is an isomorphism, and we similarly conclude that $`\varphi `$ is flat at an arbitrary $`f`$. This gives us (a).
Thus $`\varphi `$ is flat, and we may use the flat pull-back to define $`\varphi ^{}\left[\overline{M}_{0,0}(X,\beta )\right]^{\mathrm{vir}}`$. Behrend showed that the relative intrinsic normal cone $`_{\overline{M}_{0,0}(X,\beta )/𝔐_{0,0}}`$ pulls back under $`\varphi g`$ to $`_{\overline{M}_{0,3}(X,\beta )/𝔐_{0,3}}`$ and the same trick we employed in the previous paragraph shows that it pulls back under $`\varphi `$ to $`_{\overline{G}_{0,1}(X,\beta )/𝔊_{0,1}}`$. The flatness of $`\varphi `$ also tells us that $`R^1\pi _{}e^{}TX`$ pulls back to the corresponding element of the derived category of sheaves on $`\overline{G}_{0,1}(X,\beta )`$, and we get (b). The last sentence of (b) is a consequence of Behrend’s work, since the induced maps $`F\overline{M}_{0,0}(X,\beta )`$ are always gluing maps of Kontsevich-Manin spaces. ∎
## 5. The main theorem and reconstruction.
We now return to Theorem 1.4 and its generalizations and consequences.
###### Proof of the main theorem (rank one case):.
We may assume that $`H`$ is very ample. Indeed, suppose the polynomiality condition holds for the expression
$$\underset{1S[m]}{}\underset{e=0}{\overset{d}{}}J_{de}^X(\stackrel{}{t}_S,t)_{(lHet)^b}J_e^X(t,\stackrel{}{t}_{S^c})+\underset{j=2}{\overset{m}{}}J_d^X(\stackrel{}{t}_{\widehat{j}},t_j)_{(lH)^b}J_0^X(t_j,t)$$
for some $`l>0`$. Only the $`e`$’s divisible by $`l`$ will produce non-zero terms, because the degree of every curve (measured against $`lH`$) is a multiple of $`l`$. But replacing $`lHet`$ by $`H\frac{e}{l}t`$ in the twisted tensor products simply multiplies the expression by $`l^b`$. If we now replace the subscript of each $`J`$ by the degree of the curve against $`H`$ (instead of against $`lH`$) we get the desired result for $`H`$.
The embedding $`X𝐏^n`$ defined by $`H`$ allows us to define a morphism
$$v:\overline{G}_{0,m+1}(X,d)\overline{G}_{0,1}(X,d)\overline{G}_{0,1}(𝐏^n,d)𝐏(\mathrm{Hom}(\mathrm{Sym}^d(W_{m+1})V))$$
and an equivariant Chern class $`v^{}(H_𝐓^b)`$ as in the $`𝐏^n`$ case. Applying Lemma 4.2 (a) to the map $`\overline{G}_{0,m+1}(X,\beta )\overline{G}_{0,m}(X,\beta )`$, we see that $`\mathrm{\Phi }`$ is a local complete intersection (l.c.i.) morphism, since it factors through the graph followed by a flat morphism:
$$\begin{array}{ccc}& & \overline{G}_{0,m+1}(X,d)\times 𝐏^3\\ & & \\ \overline{G}_{0,m+1}(X,d)& \stackrel{\mathrm{\Phi }}{}& \overline{G}_{0,m}(X,d)\times 𝐏^3\end{array}.$$
Then by Lemma 4.2 (b),
$$\mathrm{\Phi }^{}\left(\left[\overline{G}_{0,m}(X,d)\right]_𝐓^{\mathrm{vir}}\times [𝐏^3]\right)=\left[\overline{G}_{0,m+1}(X,d)\right]_𝐓^{\mathrm{vir}}.$$
It follows by the projection formula that the correspondence of residues holds for $`c\left[\overline{G}_{0,m+1}(X,d)\right]_𝐓^{\mathrm{vir}}`$ (and any equivariant Chern class $`c`$) with each $`i^{}c`$ replaced by $`i^{}c\left[F\right]^{\mathrm{vir}}`$, and $`i_{[m1],d}^{}\mathrm{\Phi }_{}c`$ replaced by $`i_{[m1],d}^{}\mathrm{\Phi }_{}c\left[F_{[m1],d}\right]^{\mathrm{vir}}`$ (again, using Lemma 4.2). The proof of the $`𝐏^n`$ case now carries over to prove the general rank one case. ∎
Next we turn to the theorem for arbitrary $`\mathrm{H}^2(X,𝐐)`$. It seems best to do this, not for $`J`$-functions defined intrinsically on $`X`$, but for $`J`$-functions defined in terms of a choice of (generalized) polarization on $`X`$. (See also .)
###### Definition 5.
1. A divisor $`H`$ on $`X`$ is eventually free if some positive multiple $`lH`$ defines a morphism $`X𝐏^n`$.
2. A collection $`H_1`$,…,$`H_k`$ (written $`H`$ for short) of eventually free divisors is ample if positive $`𝐙`$-linear combinations $`l_1H_1+\mathrm{}+l_kH_k`$ are ample.
3. The $`J`$-functions associated to an $`H`$ as in (b) are
$`J_d^{X,H}(t_1,\mathrm{},t_m)`$ $`=J_{(d_1,\mathrm{},d_k)}^{X,H_1,\mathrm{},H_k}(t_1,\mathrm{},t_m)`$
$`:={\displaystyle \underset{d(\beta )=(d_1,\mathrm{},d_k)}{}}J_\beta ^X(t_1,\mathrm{},t_m),`$
where $`d(\beta )`$ is the multi-degree $`(\mathrm{deg}_{H_1}(\beta ),\mathrm{},\mathrm{deg}_{H_k}(\beta ))`$.
###### Theorem 5.1 (The main theorem—general case).
If $`X`$ is a complex projective manifold and $`H=(H_1,\mathrm{},H_k)`$ is an ample collection of eventually free divisors, then
$$\begin{array}{c}t(t_1+t)(\underset{1S[m]}{}\underset{ed}{}J_{de}^{X,H}(\stackrel{}{t}_S,t)_{{\scriptscriptstyle (H_ie_it)^{b_i}}}J_e^{X,H}(t,\stackrel{}{t}_{S^c})+\hfill \\ \hfill \underset{j=2}{\overset{m}{}}J_d^{X,H}(\stackrel{}{t}_{\widehat{j}},t_j)_{{\scriptscriptstyle H_i^{b_i}}}J_0^{X,H}(t_j,t))\mathrm{H}_{}(X^m,𝐐)[t_1,t_1^1,\mathrm{},t_m,t_m^1,t]\end{array}$$
In this case, we sum over $`0e=(e_1,\mathrm{},e_k)d`$, meaning that $`0e_id_i`$.
###### Proof.
As in the proof of the rank one version, we may assume that $`H_1`$,…,$`H_k`$ are not just eventually free, but free, by replacing them with positive multiples (which can be taken to be the same multiple). The $`H_i`$ define a morphism
$$v:\underset{d(\beta )=d}{}\overline{G}_{0,m+1}(X,\beta )\underset{d(\beta )=d}{}\overline{G}_{0,1}(X,\beta )\underset{i=1}{\overset{k}{}}𝐏(\mathrm{Hom}(\mathrm{Sym}^{d_i}W_{m+1},V_i))$$
and the theorem results from applying the correspondence of residues to the class $`v^{}_{i=1}^kH_{i}^{}{}_{𝐓}{}^{b_i}`$, where the $`H_{i}^{}{}_{𝐓}{}^{}`$ are the equivariant hyperplane classes pulled back from $`𝐏(\mathrm{Hom}(\mathrm{Sym}^{d_i}W_{m+1},V_i))`$. ∎
Finally, we have the
###### Theorem 5.2 (Reconstruction).
Let $`R_H\mathrm{H}^{}(X,𝐐)`$ be the subring generated as a $`𝐐`$-algebra by $`1`$ and an ample collection $`H_1`$,…,$`H_k`$ of eventually free divisors. If the orthogonal complement to $`R_H`$ annihilates each of the one-variable $`J`$-functions $`J_d^{X,H}(t)`$, then the Gromov-Witten invariants of the form
$$\underset{d(\beta )=d}{}\gamma _1\psi ^{a_1},\mathrm{},\gamma _m\psi ^{a_m}_\beta ^X$$
for $`\gamma _iR_H`$ are completely determined by the one-point invariants, the intersection matrix on $`R_H`$, and the canonical class $`K_X`$.
###### Proof.
The only term in the main theorem involving a $`J`$-function of $`m+1`$ variables and curves of (multi) degree $`d`$ is
$$J_d^{X,H}(t_1,\mathrm{},t_m,t)_{{\scriptscriptstyle H_i^{b_i}}}J_0^{X,H}(t)=\pi _{}^X\left(\left(\pi _X^{}H_i^{b_i}\right)J_d^{X,H}(t_1,\mathrm{},t_m,t)\right).$$
The product $`t(t+t_1)J_d^{X,H}(t_1,\mathrm{},t_m,t)`$ is a polynomial in $`t^1`$, expanding as
$$t(t+t_1)J_d^{X,H}(t_1,\mathrm{},t_m,t)=(t+t_1)\underset{a=1}{\overset{N}{}}t^a\underset{d(\beta )=d}{}\mathrm{ev}_{}\frac{\psi _{m+1}^{a1}\left[\overline{M}_{0,m+1}(X,\beta )\right]^{\mathrm{vir}}}{_{i=1}^mt_i(t_i\psi _i)},$$
for some $`N`$ depending on $`K_X`$. It follows by downward induction on the power of $`t^1`$ and the main theorem that every term in the expansion of $`\pi _{}^X(J_d^{X,H}(t_1,\mathrm{},t_m,t)\pi _X^{}H_i^{b_i})`$ in $`t^1`$ is determined inductively by $`J`$-functions involving fewer variables and/or lower degrees. Note that by stopping the induction at the $`t^1`$ term, we determine the constant term, about which the main theorem tells us nothing.
This argument only proves the reconstruction theorem when all cohomology is generated by the $`H_i`$ since it (inductively) requires knowledge of the classes $`\pi _{}^X((\pi _X^{}\gamma )J_d^{X,H}(t_1,\mathrm{},t_m,t))`$ where $`\gamma `$ is an *arbitrary* cohomology class. This argument does, however, capture the main idea of the proof.
We now prove the following by induction on $`(m+1,d)`$:
###### Claim 1.
1. For all $`\gamma _1,\mathrm{},\gamma _mR_H`$ and $`\alpha R_H^{}`$,
$$\underset{d(\beta )=d}{}\gamma _1\psi ^{a_1},\mathrm{},\gamma _m\psi ^{a_m},\alpha \psi ^a_\beta ^X=0.$$
2. For all $`\gamma _1,\mathrm{},\gamma _mR_H`$, the invariants
$$\underset{d(\beta )=d}{}\gamma _1\psi ^{a_1},\mathrm{},\gamma _m\psi ^{a_m},\gamma _{m+1}\psi ^{a_{m+1}}_\beta ^X$$
are determined by the one-point invariants and the intersection matrix on $`R_H`$.
In terms of $`J`$-functions (using the symmetry), this claim is equivalent to
###### Claim 2.
1. If $`\gamma _1,\mathrm{},\gamma _mR_H`$ and $`\alpha R_H^{}`$ then
$$\mathrm{deg}\left(\left(\pi _1^{}\alpha \pi _2^{}\gamma _1\mathrm{}\pi _{m+1}^{}\gamma _m\right)J_d^{X,H}(t_1,\mathrm{},t_m,t)\right)=0.$$
2. For all $`\gamma _1,\mathrm{},\gamma _mR_H`$,
$$\mathrm{deg}\left(\left(\pi _1^{}\gamma _1\pi _2^{}\gamma _2\mathrm{}\pi _{m+1}^{}\gamma _{m+1}\right)J_d^{X,H}(t_1,\mathrm{},t_m,t)\right)$$
is determined by one-point invariants and the intersection matrix on $`R_H`$.
To start our induction, note that the claim holds for $`m=0`$ by assumption. Also, the claim holds for $`d=0`$:
$$\mathrm{deg}((\pi _1^{}\alpha \pi _2^{}\gamma )J_0^{X,H}(t_1,t_2))=\frac{1}{t_1t_2(t_1+t_2)}_X\alpha \gamma =0$$
by orthogonality, and
$$\mathrm{deg}((\pi _1^{}\gamma _1\pi _2^{}\gamma _2)J_0^{X,H}(t_1,t_2))=\frac{1}{t_1t_2(t_1+t_2)}_X\gamma _1\gamma _2$$
and hence is determined by the intersection matrix on $`R_H`$.
Using the argument at the beginning of this proof, the vanishing in Claim 2(a) will follow by induction (on the power of $`t^1`$), once we establish vanishing for all expressions of the form
$`I_a`$ $`:=\mathrm{deg}\left(\left(\pi _1^{}\alpha \pi _2^{}\gamma _1\mathrm{}\pi _m^{}\gamma _{m1}\right)\left(J_{de}^{X,H}(\stackrel{}{t}_S,t)_{{\scriptscriptstyle (H_ie_it)^{b_i}}}J_e^{X,H}(t,\stackrel{}{t}_{S^c})\right)\right)`$
and
$`I_b`$ $`:=\mathrm{deg}\left(\left(\pi _1^{}\alpha \pi _2^{}\gamma _1\mathrm{}\pi _m^{}\gamma _{m1}\right)\left(J_d^{X,H}(\stackrel{}{t}_{\widehat{j}},t_j)_{{\scriptscriptstyle H_i^{b_i}}}J_0^{X,H}(t_j,t)\right)\right).`$
But these expressions may be rewritten:
$$I_b=\mathrm{deg}\left(\left(\pi _1^{}\alpha \pi _2^{}\gamma _1\mathrm{}\pi _m^{}(\gamma _{m1}H_i^{b_i})\right)J_d^{X,H}(\stackrel{}{t}_{\widehat{j}},t_j)\right).$$
To rewrite $`I_a`$, choose an orthogonal basis $`\lambda _j,\alpha _l\mathrm{H}^{}(X,𝐐)`$ such that $`\lambda _jR_H`$ with intersection matrix $`g_{jj^{}}`$ and $`\alpha _lR_H^{}`$ with intersection matrix $`h_{ll^{}}`$. Then
$$\begin{array}{c}I_a=\underset{j,j^{}}{}\mathrm{deg}\left(\left(\pi _1^{}\alpha \mathrm{}\pi _{k+1}^{}((H_ie_it)^{b_i}\lambda _j)\right)J_{de}^{X,H}(\stackrel{}{t}_S,t)\right)g^{jj^{}}\hfill \\ \hfill \mathrm{deg}\left(\left(\pi _1^{}(\lambda _j^{})\pi _2^{}\gamma _k\mathrm{}\pi _{mk+1}^{}\gamma _{m1}\right)J_e^{X,H}(t,\stackrel{}{t}_{S^c})\right)\\ \hfill +\underset{l,l^{}}{}\mathrm{deg}\left(\left(\pi _1^{}\alpha \mathrm{}\pi _{k+1}^{}((H_ie_it)^{b_i}\alpha _l)\right)J_{de}^{X,H}(\stackrel{}{t}_S,t)\right)h^{ll^{}}\\ \hfill \mathrm{deg}\left(\left(\pi _1^{}(\alpha _l^{})\pi _2^{}\gamma _k\mathrm{}\pi _{mk+1}^{}\gamma _{m1}\right)J_e^{X,H}(t,\stackrel{}{t}_{S^c})\right).\end{array}$$
Now suppose Claim 2(a) holds for all $`(n+1,e)`$ such that either $`n<m`$ or $`n=m`$ and $`ed`$. Then $`I_b=0`$ (taking $`n=m1`$), and $`I_a=0`$ since the first factors in the first double sum and the second factors in the second sum vanish. This proves Claim 1(a) by induction. Similarly, assuming Claim 1(a), we prove 1(b) by induction, noting that in this case, the second double sum in $`I_a`$ (but not the first) vanishes. The first double sum and the $`I_b`$ terms are explicitly determined by the intersection matrix $`g_{jj^{}}`$ and Gromov-Witten invariants for lower $`(n+1,e)`$. ∎
## Appendix A Small quantum product for complete intersections
We may turn Formula 1.1 into an algorithm for producing structure constants for the small quantum product on Fano complete intersections in $`𝐏^n`$.
Given the type, $`(l_1,\mathrm{},l_m)`$ of the Fano complete intersection $`S𝐏^n`$, set
$`f`$ $`:=n+1l_1\mathrm{}l_m`$
the Fano index of $`S`$, and
$`d_{\mathrm{max}}`$ $`:={\displaystyle \frac{nm+1}{f}},`$
the maximal degree $`d`$ for which nonzero “unmixed” $`2`$-point invariants $`H^a,H^b_d^X`$ may occur (by a dimension count).
For $`d=1`$,…,$`d_{\mathrm{max}}`$, let $`v(d)`$ be the vector of one-point invariants, i.e., $`v(d)`$ is defined by
$$e_{}\left(\frac{\left[\overline{M}_{0,1}(X,d)\right]^{\mathrm{vir}}}{t(t\psi )}\right)=v(d)_0t^f+v(d)_1Ht^{f1}+\mathrm{}+v(d)_{nm}H^{nm}t^{fn+m}.$$
(These are computed by Givental’s formulas, Theorem 1.3.)
We define shift matrices of size $`(nm+1)\times (nm+1)`$:
$$S(d):=\left(\begin{array}{ccccc}d& 0& \mathrm{}& 0& 0\\ 1& d& \mathrm{}& 0& 0\\ & & \mathrm{}\\ 0& 0& \mathrm{}& d& 0\\ 0& 0& \mathrm{}& 1& d\end{array}\right).$$
Applying $`S(d)`$ to a vector corresponds to multiplication by $`(H+dt)`$.
We define the matrices of mixed invariants, also of size $`(nm+1)\times (nm+1)`$, indexed from $`0`$ to $`nm`$:
$$M(d)_{nma,b}:=\frac{(1)^{c1}}{l_i}H^a,H^b\psi ^c_d^X,$$
where $`c:=df+nmab`$. This is the matrix associated to the operator
$$H^be_{1}^{}{}_{}{}^{}\left(\frac{e_2^{}H^b\left[\overline{M}_{0,2}(X,d)\right]^{\mathrm{vir}}}{t\psi }\right).$$
It is important to note that $`M(d)_{nma,b}=0`$ when $`c<0`$.
In terms of these data structures, our formula becomes a recursive formula for the $`b`$th column of $`M(d)`$ in terms of the lower $`M(e)`$’s:
$$M(d)_{,b}=S(d)^bv(d)\underset{e=1}{\overset{d1}{}}M(de)S(e)^bv(e),$$
except that we must set $`M(d)_{nma,b}=0`$ whenever $`c<0`$. This amounts to truncating $`M(d)`$ at the upper right corner.
Finally, reading off all coefficients of $`M(d)`$ with $`c=0`$ yields the complete list of “unmixed” two-point invariants, which in turn yield the structure constants of thesmall quantum product (via the associativity).
This algorithm is very easy to implement. For example, when $`X`$ is a quintic hypersurface in $`𝐏^6`$, it gives the following products:
$$\begin{array}{ccccccccc}HH& =& H^2& +& 120q& & & & \\ HH^2& =& H^3& +& 770qH& & & & \\ HH^3& =& H^4& +& 1345qH^2& +& 211,200q^2& & \\ HH^4& =& H^5& +& 770qH^3& +& 692,500q^2H& & \\ HH^5& =& & & 120qH^4& +& 211,200q^2H^2& +& 31,320,000q^3\end{array}$$
As a typical application, note that the last number implies the following interesting bit of enumerative geometry:
###### Corollary A.1.
The expected number of twisted cubics through two general points of of a quintic five-fold $`X𝐏^6`$ is:
$$2,088,000.$$
One similarly may produce the expected numbers of rational normal curves of degree $`d`$ passing through $`2`$ general points of a hypersurface of degree $`2d1`$ in $`𝐏^{2d}`$ for any $`d`$.
|
warning/0007/hep-ph0007320.html
|
ar5iv
|
text
|
# Kaon vs. Bottom: Where to look for a general MSSM?
## Abstract
We analyze CP violation and Flavor Changing effects in a general Minimal Supersymmetric extension of the Standard Model with arbitrary non–universal soft–breaking terms. We show that, in this conditions, large FCNC effects are naturally expected in the Kaon system, even in the absence of quark–squark flavor misalignment. On the other hand, the B system is only sensitive to new supersymmetric contributions if the non–universality implies, not only different soft term for the three generations but also the presence of a quark–squark misalignment much larger that the corresponding CKM mixing. The only exception to this rule are processes where the chirality changing contributions proportional to $`\mathrm{tan}\beta `$ are leading (for instance $`bs\gamma `$).
preprint: SISSA/69/2000/EP
CP violation and Flavor Changing Neutral Current (FCNC) experiments are a main arena where searches of new physics beyond the Standard Model (SM) of electroweak interactions will take place in the near future. In these processes, SM contributions are small and hence new contributions are allowed to compete on equal grounds. In fact, the currently available experimental information allows us to set stringent bounds on any extension of the SM in the neighborhood of the electroweak scale. The so–called minimal supersymmetric extension of the SM (MSSM) constitutes one of the most interesting examples of this. Indeed, it is well–known that a general MSSM with completely arbitrary soft–breaking terms at the electroweak scale suffers the so–called supersymmetric flavor and CP problems. This means that such a model tends to over–produce flavor changing and CP violation effects and this is used to set very stringent bounds on the structure of the sfermion mass matrices . However, the size of FCNC and CP violation effects are strongly dependent on the structure of the supersymmetry soft breaking terms which finally define a particular MSSM. In a recent work , we showed that in the absence of a new flavor structures in the soft breaking terms no sizeable contributions to CP violation observables, as $`\epsilon _K`$, $`\epsilon ^{}/\epsilon `$ or hadronic $`B^0`$ CP asymmetries, can be expected. In contrast, the presence of non–universality in the trilinear terms, expected in string theories , is already enough to generate large supersymmetric contributions to $`\epsilon ^{}/\epsilon `$ . Unfortunately, the large theoretical uncertainties in the SM prediction prevent us from using this observable to identify these new supersymmetric contributions.
In this letter, we will address the problem of indirect SUSY discovery in the framework of a completely general MSSM (see our definition below). This analysis does not intend to exhaust every single possibility but rather to provide a general view of the “natural” SUSY effects. To our knowledge, such a study is so far absent in the literature. The common attitude towards this problem consists in a pure phenomenological analysis through the Mass Insertion (MI) approximation . However, we will show that the inclusion of very general and reasonable theoretical requirements has a deep impact on the expectations of these SUSY effects when compared with MI bounds. In particular, we will see that, contrary to common wisdom, while large supersymmetric effects in hadronic B CP asymmetries are possible only under some (interesting) special conditions , in the kaon system these are naturally expected in any model with non–universal soft–breaking terms. More precisely: i)if the new SUSY flavor structure is given by non–universal flavor diagonal soft terms, then only the kaon system is sizeably affected in FCNC and CP violation processes <sup>*</sup><sup>*</sup>*However, as already known , observables with a dominant chirality changing part whose SUSY contribution is proportional to $`\mathrm{tan}\beta `$ (for instance $`bs\gamma `$ and $`bsl^+l^{}`$) escape this rule, even in the absence of new flavor structure, for large $`\mathrm{tan}\beta `$, ii) if SUSY signals are to be seen in CP violating hadronic B decays, then it implies the presence, not only of flavor diagonal non–universality, but also a rather large quark–squark flavor misalignment (larger than the corresponding CKM mixings).
In this context, it is interesting to comment the new results for the $`BJ/\psi K_S`$ CP asymmetry recently obtained by BaBar and BELLE . A possible discrepancy with the SM prediction could be due either to a dominant new physics contribution in the kaon system (i.e. modifying the determination of $`\epsilon _K`$) or mainly to a new effect in the $`B^0`$ CP asymmetry itself . Taking into account the above statements, in a generic MSSM with non–universal soft–breaking terms it is very likely to have a sizeable effect in $`\epsilon _K`$ while a contribution to the $`B^0`$ CP asymmetry is possible only under special conditions and again usually associated to a large contribution in the kaon system. Hence, in an general MSSM as defined below, a low value in the $`BJ/\psi K_S`$ CP asymmetry is more likely to be due to a new SUSY contribution to $`\epsilon _K`$.
In first place, we define our MSSM through a set of four general conditions:
* Minimal particle content: we consider the MSSM, the minimal supersymmetrization of the Standard Model with no additional particles from $`M_W`$ to $`M_{GUT}`$.
* Arbitrary Soft–Breaking terms $`𝒪(m_{3/2})`$: The supersymmetry soft–breaking terms as given at the scale $`M_{GUT}`$ have a completely general flavor structure. However, all of them are of the order of a single scale, $`m_{3/2}`$.
* Trilinear couplings originate from Yukawa couplings: Although trilinear couplings are a completely new flavor structure they are related to the Yukawas in the usual way: $`Y_{ij}^A=A_{ij}Y_{ij}`$, with all $`A_{ij}𝒪(m_{3/2})`$.
* Gauge coupling unification at $`M_{GUT}`$.
In this framework, to define our MSSM, all we have to do is to write the full set of soft–breaking terms. This model includes, in the quark sector, 7 different structures of flavor, $`M_{\stackrel{~}{Q}}^2`$, $`M_{\stackrel{~}{U}}^2`$, $`M_{\stackrel{~}{D}}^2`$, $`Y_d`$, $`Y_u`$, $`Y_d^A`$ and $`Y_u^A`$. However, these 7 matrices are not completely observable, only relative misalignments among them are physical. Hence, we have the freedom to choose a convenient basis. At the supersymmetry breaking scale, $`M_{GUT}`$, the natural basis, specially from the point of view of the SUSY breaking theory, is the basis where all the squark mass matrices, $`M_{\stackrel{~}{Q}}^2,M_{\stackrel{~}{U}}^2,M_{\stackrel{~}{D}}^2`$, are diagonal. In this basis, the Yukawa matrices are, $`v_1Y_d=K_{}^{D_L}{}_{}{}^{}M_dK^{D_R}`$ and $`v_2Y_u=K_{}^{D_L}{}_{}{}^{}K^{}M_uK^{U_R}`$, with $`M_d`$ and $`M_u`$ diagonal quark mass matrices, $`K`$ the Cabibbo–Kobayashi–Maskawa (CKM) mixing matrix and $`K^{D_L}`$, $`K^{U_R}`$, $`K^{D_R}`$ unknown, completely general, $`3\times 3`$ unitary matrices.
In a basis independent language the matrices $`K^{D_L}`$, $`K^{D_R}`$, $`K^{U_L}`$ and $`K^{U_R}`$ measure the flavor misalignment among, $`d_L`$$`\stackrel{~}{Q}_L`$, $`d_R`$$`\stackrel{~}{d}_R`$, $`u_L`$$`\stackrel{~}{Q}_L`$ and $`u_R`$$`\stackrel{~}{u}_R`$ respectively. Unfortunately, we have not much information on the structure of these matrices. All we know is that there is, indeed, a minimum degree of misalignment among $`K^{D_L}`$ and $`K^{U_L}`$, which necessarily implies that these two matrices can not be simultaneously diagonal. This misalignment is described by the CKM mixing matrix $`K^{U_L}K_{}^{D_L}{}_{}{}^{}=K`$, and we use this relation to fix one of these matrices. There remain then, three unknown matrices, $`K^{D_L}`$, $`K^{D_R}`$ and $`K^{U_R}`$ . From the phenomenological point of view, we could have two complementary pictures. If the source of flavor structure in the SUSY breaking sector is completely independent from the source of flavor in the Yukawa sector, these matrices are completely arbitrary and large mixings are possible. However, if the flavor structure in the Yukawa sector and the soft–breaking terms have a common origin or are somehow related they can be expected to be close to the identity, or more exactly, close to the CKM mixing matrix which is the minimum degree of misalignment experimentally required. As we will see below these two scenarios have very different phenomenological consequences. Finally, the trilinear matrices, $`Y_d^A`$ and $`Y_u^A`$, are specified in this basis by the SUSY breaking theory as $`Y_{ij}^A=A_{ij}Y_{ij}`$.
Once we specify these 7 matrices, our MSSM model is fully defined at the $`M_{GUT}`$ scale. However, experimentally measurable quantities involve the sfermion mass matrices at the electroweak scale. So, the next step is to use the MSSM Renormalization Group Equations (RGE) to evolve these matrices down to the electroweak scale. Below the electroweak scale, it is more convenient to work in the SCKM basis where the same unitary transformation is applied to both quarks and squarks so that quark mass matrices are diagonalized. The main RGE effects from $`M_{GUT}`$ to $`M_W`$ are those associated with the gluino mass and third generation Yukawa couplings. Regarding squark mass matrices, it is well–known that diagonal elements receive important RGE contributions proportional to gluino mass that dilute the mass eigenstate non–degeneracy, $`m_{\stackrel{~}{D}_{A_i}}^2(M_W)c_A^im_{\stackrel{~}{g}}^2+m_{\stackrel{~}{D}_{A_i}}^2`$ , with $`c_L^{1,2}(7,7)`$, $`c_L^3(5.8,5.0)`$, $`c_R^{1,2}(6.6,6.6)`$ and $`c_R^3(6.6,4.5)`$ for $`(\mathrm{tan}\beta =2,\mathrm{tan}\beta =40)`$. In the limit of complete degeneracy of the initial diagonal values at $`M_{GUT}`$ (equivalent to the Constrained MSSM) off–diagonal elements after RGE running are necessarily proportional to masses and CKM elements and hence small. This means that any larger new contribution to off–diagonal elements, possibly due to a large quark–squark misalignment in the different $`K^A`$ matrices, must be, at the same time, proportional to the difference of mass eigenstates at $`M_{GUT}`$ (for an example see Ref.). Then, in the SCKM basis the off–diagonal elements in the sfermion mass matrices will be given by $`(K^AM_{\stackrel{~}{A}}^2K_{}^{A}{}_{}{}^{})_{ij}`$ up to smaller RGE corrections. On the other hand, gaugino effects in the trilinear RGE are always proportional to the Yukawa matrices, not to the trilinear matrices themselves and so they are always diagonal to extremely good approximation in the SCKM basis. Once more, the off–diagonal elements will be approximately given by $`(K^AY_{u,d}^AK_{}^{A}{}_{}{}^{})_{ij}`$.
After RGE running, flavor–changing effects in the SCKM basis can be estimated by the insertion of flavor–off–diagonal components of the mass–squared matrices normalized by an average squark mass, the so–called mass insertions (MI). In first place, we will analyze the $`LR`$ MI. Due to the trilinear terms structure, the $`LR`$ sfermion matrices are always suppressed by $`m_q/m_{\stackrel{~}{q}}`$, with $`m_q`$ a quark mass and $`m_{\stackrel{~}{q}}`$ the average squark mass. In any case, this suppression is compulsory to avoid charge and color breaking and directions unbounded from below . In particular, it is required that $`(Y_d^A)_{ij}<\sqrt{3}m_{\stackrel{~}{q}}max\{m_i,m_j\}/v_1`$, which in turn implies that $`(\delta _{LR}^d)_{ij}<\sqrt{3}max\{m_i,m_j\}/m_{\stackrel{~}{q}}`$. So we must impose, as model independent upper bounds,
$`(\delta _{LR}^d)_{12}`$ $`<`$$``$ $`\sqrt{3}{\displaystyle \frac{m_s}{m_{\stackrel{~}{q}}}}3.2\times 10^4\left({\displaystyle \frac{500GeV}{m_{\stackrel{~}{q}}}}\right)`$ (1)
$`(\delta _{LR}^d)_{13}`$ $`<`$$``$ $`\sqrt{3}{\displaystyle \frac{m_b}{m_{\stackrel{~}{q}}}}0.01\left({\displaystyle \frac{500GeV}{m_{\stackrel{~}{q}}}}\right)`$ (2)
where we take all quark masses evaluated at $`M_Z`$ . The phenomenological MI bounds given in Table I are stronger only for $`Im(\delta _{LR}^d)_{12}`$. Hence, this means that there is still room to saturate $`\epsilon ^{}/\epsilon `$ with $`LR`$ mass insertions. However, these bounds imply that, under general circumstances, it is not possible to saturate simultaneously $`\epsilon _K`$ and $`\epsilon ^{}/\epsilon `$ with a $`|(\delta _{LR}^d)_{12}|=3\times 10^3`$ as suggested in Refs Unless the universe lives in a metastable vacuum for cosmologically long interval times.
In any case, we must remember that these are only upper bounds and a definite model in the framework of our four general conditions gives rise to somewhat smaller flavor changing effects. This can be seen explicitly in a type I string inspired example . In these models, we can write the trilinear couplings in matrix notation as,
$`Y_{d(u)}^A(M_{GUT})=\left(\begin{array}{ccc}a_1^Q& 0& 0\\ 0& a_2^Q& 0\\ 0& 0& a_3^Q\end{array}\right)Y_{d(u)}+Y_{d(u)}\left(\begin{array}{ccc}a_1^{D(U)}& 0& 0\\ 0& a_2^{D(U)}& 0\\ 0& 0& a_3^{D(U)}\end{array}\right)`$ (9)
As discussed above, gluino RGE effects are again diagonal in the SCKM basis and off-diagonal elements are basically given by the initial conditions. So, using unitarity of $`K^{D_L}`$ and $`K^{D_R}`$ it is straight–forward to get,
$`(\delta _{LR}^d)_{ij}`$ $`={\displaystyle \frac{1}{m_{\stackrel{~}{q}}^2}}(m_j(a_2^Qa_1^Q)K_{i2}^{D_L}K_{j2}^{D_L}{}_{}{}^{}+m_j(a_3^Qa_1^Q)K_{i3}^{D_L}K_{j3}^{D_L}{}_{}{}^{}`$ (10)
$`+m_i(a_2^Da_1^D)K_{i2}^{D_R}K_{j2}^{D_R}{}_{}{}^{}+m_i(a_3^Da_1^D)K_{i3}^{D_R}K_{j3}^{D_R}{}_{}{}^{})`$ (11)
In the kaon system, we can make a simple estimate neglecting $`m_d`$ and assuming $`K^{D_L}K`$, which is the expected size if quark and squark masses have a common flavor origin (case i in the above introductory discussion),
$`(\delta _{LR}^d)_{12}`$ $`{\displaystyle \frac{m_s}{m_{\stackrel{~}{q}}}}{\displaystyle \frac{(a_2^Qa_1^Q)}{m_{\stackrel{~}{q}}}}K_{12}K_{22}^{}{}_{}{}^{}a4\times 10^5\left({\displaystyle \frac{500GeV}{m_{\stackrel{~}{q}}}}\right)`$ (12)
$`a`$ is a constant typically between $`0.1`$ and $`1`$. Comparing with the bounds on the MI in Table I we can see that indeed this value could give a very sizeable contribution to $`\epsilon ^{}/\epsilon `$. It is important to notice that the phase of $`(a_2^Qa_1^Q)`$ is actually unconstrained by electric dipole moment (EDM) experiments as emphasized in . This result is very important: it means that even if the relative quark–squark flavor misalignment is absent, the presence of non–universal flavor–diagonal trilinear terms is enough to generate large FCNC effects in the Kaon system.
Similarly, in the neutral $`B`$ system, $`(\delta _{LR}^d)_{13}`$ contributes to the $`B_d\overline{B}_d`$ mixing parameter, $`\mathrm{\Delta }M_{B_d}`$. However, in the common flavor origin scenario (case i), i.e. $`K^{D_L}K`$ (a direct extension of Ref. mechanism), we obtain,
$`(\delta _{LR}^d)_{13}`$ $`{\displaystyle \frac{m_b}{m_{\stackrel{~}{q}}}}{\displaystyle \frac{(a_3^Qa_1^Q)}{m_{\stackrel{~}{q}}}}K_{13}K_{33}^{}{}_{}{}^{}a4.8\times 10^5\left({\displaystyle \frac{500GeV}{m_{\stackrel{~}{q}}}}\right),`$ (13)
clearly too small to generate sizeable $`\stackrel{~}{b}`$$`\stackrel{~}{d}`$ transitions, as can be seen comparing to the MI bounds in Table I.
Larger effects are still possible in our second scenario where flavor origin in the Yukawa sector and the soft–breaking sectors are unrelated (case ii). In this framework, a large mixing could be possible, and with a maximal value, $`|K_{13}^{D_L}K_{33}^{D_L}{}_{}{}^{}|=1/2`$, we would approach
$`(\delta _{LR}^d)_{13}`$ $`{\displaystyle \frac{m_b}{m_{\stackrel{~}{q}}}}{\displaystyle \frac{(a_3^Qa_1^Q)}{m_{\stackrel{~}{q}}}}{\displaystyle \frac{1}{2}}a3\times 10^3\left({\displaystyle \frac{500GeV}{m_{\stackrel{~}{q}}}}\right).`$ (14)
Even in this limiting situation, this result is roughly one order of magnitude too small to saturate $`\mathrm{\Delta }M_{B_d}`$, though it could still be observed through the CP asymmetries. Hence in the $`B`$ system we arrive to a very different result: it is not at all enough to have non–universal trilinear terms, but, it is also required to have large flavor misalignment among quarks and squarks.
The situation for the $`LL`$ and $`RR`$ mass insertions is less defined due to the absence of any theoretical prejudice on these mass matrices at $`M_{GUT}`$. In any case we can write these MI as,
$`(\delta _A^d)_{ij}`$ $`={\displaystyle \frac{1}{m_{\stackrel{~}{q}}^2}}\left((m_{\stackrel{~}{D}_{A_2}}^2m_{\stackrel{~}{D}_{A_1}}^2)K_{i2}^{D_A}K_{j2}^{D_A}{}_{}{}^{}+(m_{\stackrel{~}{D}_{A_3}}^2m_{\stackrel{~}{D}_{A_1}}^2)K_{i3}^{D_A}K_{j3}^{D_A}{}_{}{}^{}\right)`$ (15)
However, due to the gluino dominance in the squark eigenstates at $`M_W`$ we can say that $`m_{\stackrel{~}{q}}^2(M_W)7m_{\stackrel{~}{g}}^2(M_{GUT})`$. Hence, $`(m_{\stackrel{~}{D}_{A_i}}^2m_{\stackrel{~}{D}_{A_1}}^2)/m_{\stackrel{~}{q}}^20.1bm_{\stackrel{~}{q}}^2`$ with $`b`$ a number $`𝒪(1)`$. Replacing these values in Eq (15), this means, for the kaon system and assuming CKM–like mixing (case i),
$`(\delta _A^d)_{12}`$ $`0.1bK_{12}0.02b.`$ (16)
Hence, in the presence of non–universality, sizeable SUSY contributions to the kaon mass difference are expected. Regarding CP violation observables, with large phases in the $`K^{D_A}`$ matrices, big contributions to $`\epsilon _K`$ arise (notice $`m_{\stackrel{~}{D}_{A_i}}^2`$ are always real numbers). This is again equivalent to our result for the $`LR`$ transitions: even if the sfermion mass matrices at $`M_{GUT}`$ are flavor diagonal, large FCNC effects are produced in the $`K`$ system with family non–universal masses.
In the B system, we must again distinguish our two possible scenarios. In case i), $`K^{D_L}K^{D_R}K`$,
$`(\delta _A^d)_{13}`$ $`0.1bK_{13}8\times 10^4b,`$ (17)
clearly too small to be observed at the B factories. Nevertheless, in the presence of large mixing, still possible when flavor in the soft breaking terms has an independent source, we can get $`K_{13}^{D_A}K_{33}^{D_A}{}_{}{}^{}1/2`$. In this limiting situation $`(\delta _A^d)_{13}<0.05`$, still reachable through CP asymmetries at the B factories in the presence of a sizeable phase in $`K^{D_A}`$. So, we find again that to have observable effects in the $`B`$ system it is required to have not only the presence of non–universality but also large quark–squark flavor misalignment.
In summary, in gluino mediated transitions large effects are expected in the Kaon system in the presence of non–universal squark masses even with a “natural” CKM–like mixing. However in the $`B`$ system, due to the much lower sensitivity to supersymmetric contributions, observable effects are expected only with approximately maximal $`\stackrel{~}{b}`$$`\stackrel{~}{d}`$ mixings.
Finally we add a short comment on chargino contributions. The chargino sector is more involved due to the presence of the CKM mixing matrix and the additional mixing wino–higgsino. However, for large and similar squark mixings in the up and down sector, chargino and gluino contributions are expected to be of the same order, barring special situations as light stop ($`M_{\stackrel{~}{t}}<300GeV`$) and chargino. Indeed, in this case, the effects of the CKM matrix in chargino couplings are sufficiently small and can be neglected. Then, the flavor changing chargino and gluino vertices have similar mixing matrices and the difference in the couplings is compensated through the RGE relation $`\alpha _3/m_{\stackrel{~}{g}}^2=\alpha _2/m_{\stackrel{~}{w}}^2`$. The only additional ingredient in chargino vertices is the presence of higgsino. Higgsino couplings are suppressed by small Yukawa couplings with the exception of the top quark with $`Y_t1`$. Hence, the main difference between gluino and chargino contributions is the possible presence of a light stop due to the the large top quark Yukawa coupling. The analysis of these new SUSY flavor contributions to $`B_d`$$`\overline{B}_d`$ mixing in the presence of light stop and chargino was considered in Ref For a discussion in the flavor universal case see . It was shown there that for a light stop of $`140GeV`$ and a chargino of $`100GeV`$ and with a large $`\stackrel{~}{u}_L`$$`\stackrel{~}{t}_L`$ mixing, $`K_{13}^{U_L}\lambda _c`$ it is still possible to produce contributions to $`\mathrm{sin}2\beta `$ as large as $`0.78`$ with $`\delta _{CKM}=0`$. However, this is a limiting case, light stop and chargino and, more important, a sizeable mixing in the $`\stackrel{~}{u}_L`$$`\stackrel{~}{t}_L`$ are required. Hence, we conclude that also chargino contributions fit on our general scheme and only in an scenario with different flavor origin in the Yukawa and soft–breaking sectors sizeable effects appear.
In conclusion we have shown that in the kaon system large SUSY effects are naturally expected in any model with non–universal soft–breaking terms. In this direction, several works can be found in the literature in the mass insertion context, and a complete analysis will be presented in short time . On the other hand, a discovery of a supersymmetric contribution to hadronic $`B^0`$ CP asymmetries would signal the presence of a different origin of flavor in the breaking of SUSY.
We thank T. Kobayashi, E. Lunghi, M. Misiak and J.C. Pati for enlightening discussions. The work of A.M. was partially supported by the European TMR Project “Beyond the Standard Model” contract N. ERBFMRX CT96 0090; O.V. acknowledges financial support from a Marie Curie EC grant (TMR-ERBFMBI CT98 3087) and partial support from spanish CICYT AEN-99/0692.
|
warning/0007/hep-ph0007157.html
|
ar5iv
|
text
|
# hep-ph/0007157 IFIC/00-45 UWThPh-2000/23 Neutralino Phenomenology at LEP2 in Supersymmetry with Bilinear Breaking of R-parity
## 1 Introduction
The search for supersymmetry (SUSY) plays an important role in the experimental programs of existing high energy colliders like LEP2, HERA and the Tevatron. It will play an even more important role at future colliders like LHC or a linear $`e^+e^{}`$ collider. So far most of the effort in searching for supersymmetric signatures has been confined to the framework of R–parity-conserving realizations. Recent data on solar and atmospheric neutrinos strongly indicate the need for neutrino conversions . Motivated by this there has been in the last few years a substantial interest in R-parity violating models . The violation of R-parity could arise explicitly as a residual effect of some larger unified theory , or spontaneously, through nonzero vacuum expectation values (VEV’s) for scalar neutrinos . In the first case there is a large number of unknown parameters characterizing the superpotential of these models, so that for simplicity these effects are usually studied assuming in an *ad hoc* way that only a few dominant terms break R-parity explicitly, usually only one.
We prefer theoretical scenarios which break R–parity only as a result of the properties of the vacuum . There are two generic cases of spontaneous R-parity breaking models. In the first case lepton number is part of the gauge symmetry and there is a new gauge boson $`Z^{}`$ which gets mass via the Higgs mechanism . In this model the lightest SUSY particle (LSP) is in general a neutralino which decays, therefore breaking R-parity. The LSP decays mostly to visible states such as
$$\stackrel{~}{\chi }_1^0f\overline{f}\nu ,$$
(1)
where $`f`$ denotes a charged fermion. These decays are mediated by the $`Z`$-boson or by the exchange of scalars. In the second class of models there appears a physical massless Nambu-Goldstone boson, called majoron. The latter arises in $`SU(2)U(1)`$ models where the breaking of R-parity occurs spontaneously. In this case the majoron is the LSP, which is stable because it is massless (or nearly so). It leads to an additional invisible decay mode $`\stackrel{~}{\chi }_1^0\nu +J`$, which is R-parity conserving since the majoron has a large R-odd singlet sneutrino component . This decay is absent if lepton number is gauged, as the majoron is eaten up by a massive additional Z boson.
Although models with spontaneous R-parity breaking usually contain additional fields not present in the MSSM in order to drive the violation of R-parity (expected to lie in the TeV range), they are characterized by much fewer parameters than models with explicit breaking of R–parity. Most phenomenological features of these models are reproduced by adding three explicit bilinear R-parity breaking terms to the MSSM superpotential . This renders a systematic way to study R-parity breaking signals and leads to effects that can be large enough to be experimentally observable, even in the case where neutrino masses are as small as indicated by the simplest interpretation of solar and atmospheric neutrino data . Moreover, R-parity violating interactions follow a specific *pattern* which can be easily characterized. These features have been exploited in order to describe the R-parity violating signals expected for chargino production at LEP II .
Here we consider the phenomenology of the lightest neutralino in the simplest and well motivated class of models with an effective explicit R-parity breaking characterized by a single bilinear superpotential term . Apart from the absence of the majoron-emitting $`\stackrel{~}{\chi }_1^0`$ decays (which are absent in majoron-less models with spontaneous breaking of R-parity) this bilinear model mimics all features of neutralino decay properties relevant for our analysis. For simplicity and for definiteness we consider supergravity scenarios where the lightest supersymmetric particle (LSP) is the lightest neutralino. We present a detailed study of the LSP decay properties and general features of the corresponding signals expected at LEP2. In the following we denote the minimal SUGRA scenario with conserved R-parity by mSUGRA. It is well known that in models with gauge mediated supersymmetry breaking (GMSB) the lightest neutralino decays , because the gravitino is the LSP. We therefore also discuss the possibilities to distinguish between GMSB and our R-parity breaking model.
## 2 The model
Here we will adopt a supersymmetric Lagrangian specified by the following superpotential
$$W=\epsilon _{ab}\left[h_U^{ij}\widehat{Q}_i^a\widehat{U}_j\widehat{H}_2^b+h_D^{ij}\widehat{Q}_i^b\widehat{D}_j\widehat{H}_1^a+h_E^{ij}\widehat{L}_i^b\widehat{R}_j\widehat{H}_1^a\mu \widehat{H}_1^a\widehat{H}_2^b\right]+\epsilon _{ab}ϵ_i\widehat{L}_i^a\widehat{H}_2^b,$$
(2)
where $`i,j=1,2,3`$ are generation indices, $`a,b=1,2`$ are $`SU(2)`$ indices, and $`\epsilon `$ is a completely anti-symmetric $`2\times 2`$ matrix, with $`\epsilon _{12}=1`$. The symbol “hat” over each letter indicates a superfield, with $`\widehat{Q}_i`$, $`\widehat{L}_i`$, $`\widehat{H}_1`$, and $`\widehat{H}_2`$ being $`SU(2)`$ doublets with hypercharges $`1/3`$, $`1`$, $`1`$, and $`1`$ respectively, and $`\widehat{U}`$, $`\widehat{D}`$, and $`\widehat{R}`$ being $`SU(2)`$ singlets with hypercharges $`\frac{4}{3}`$, $`\frac{2}{3}`$, and $`2`$ respectively. The couplings $`h_U`$, $`h_D`$, and $`h_E`$ are $`3\times 3`$ Yukawa matrices, and $`\mu `$ and $`ϵ_i`$ are parameters with units of mass.
Supersymmetry breaking is parametrized by the standard set of soft supersymmetry breaking terms
$`V_{soft}`$ $`=`$ $`M_Q^{ij2}\stackrel{~}{Q}_i^a\stackrel{~}{Q}_j^a+M_U^{ij2}\stackrel{~}{U}_i^{}\stackrel{~}{U}_j+M_D^{ij2}\stackrel{~}{D}_i^{}\stackrel{~}{D}_j+M_L^{ij2}\stackrel{~}{L}_i^a\stackrel{~}{L}_j^a+M_R^{ij2}\stackrel{~}{R}_i^{}\stackrel{~}{R}_j`$ (3)
$`+m_{H_1}^2H_1^aH_1^a+m_{H_2}^2H_2^aH_2^a`$
$`[\frac{1}{2}M_3\lambda _3\lambda _3+\frac{1}{2}M\lambda _2\lambda _2+\frac{1}{2}M^{}\lambda _1\lambda _1+h.c.]`$
$`+\epsilon _{ab}[A_U^{ij}h_U^{ij}\stackrel{~}{Q}_i^a\stackrel{~}{U}_jH_2^b+A_D^{ij}h_D^{ij}\stackrel{~}{Q}_i^b\stackrel{~}{D}_jH_1^a+A_E^{ij}h_E^{ij}\stackrel{~}{L}_i^b\stackrel{~}{R}_jH_1^a`$
$`B\mu H_1^aH_2^b+B_iϵ_i\stackrel{~}{L}_i^aH_2^b],`$
Note that, in the presence of soft supersymmetry breaking terms the bilinear terms $`ϵ_i`$ can not be rotated away, since the rotation, that eliminates it, re-introduces an R–parity violating trilinear term, as well as a sneutrino vacuum expectation value . This happens even in the case where universal boundary conditions are adopted for the soft breaking terms at the unification scale, since universality will be effectively broken at the weak scale due to calculable renormalization effects. For definiteness and simplicity we will adopt this assumption throughout this paper.
Although for the discussion of flavour–changing processes, such as neutrino oscillations involving all three generations, it is important to consider the full three-generation structure of the model, for the following discussion of neutralino decay properties it will suffice to assume R-parity Violation (RPV) only in the third generation, as a first approximation. In this case we will omit the labels $`i,j`$ in the superptotential and the soft breaking terms
$`W`$ $`=`$ $`h_t\widehat{Q}_3\widehat{U}_3\widehat{H}_2+h_b\widehat{Q}_3\widehat{D}_3\widehat{H}_1+h_\tau \widehat{L}_3\widehat{R}_3\widehat{H}_1\mu \widehat{H}_1\widehat{H}_2+ϵ_3\widehat{L}_3\widehat{H}_2`$ (4)
$`V_{soft}`$ $`=`$ $`\epsilon _{ab}[A_th_t\stackrel{~}{Q}_3^a\stackrel{~}{U}_3H_2^b+A_bh_b\stackrel{~}{Q}_3^b\stackrel{~}{D}_3H_1^a+A_\tau h_\tau \stackrel{~}{L}_3^b\stackrel{~}{R}_3H_1^a`$ (5)
$`B\mu H_1^aH_2^b+B_3ϵ_3\stackrel{~}{L}_3^aH_2^b]+\mathrm{mass}\mathrm{terms}.`$
This amounts to neglecting the $`R_p/`$ effects in the two first families.
The bilinear terms in Eqs. (4) and (5) lead to a mixing between the charginos and the $`\tau `$–lepton which is described by the mass matrix
$$𝐌_𝐂=\left[\begin{array}{ccc}M_2& \frac{1}{\sqrt{2}}gv_2& 0\\ \frac{1}{\sqrt{2}}gv_1& \mu & \frac{1}{\sqrt{2}}h_\tau v_3\\ \frac{1}{\sqrt{2}}gv_3& ϵ_3& \frac{1}{\sqrt{2}}h_\tau v_1\end{array}\right],$$
(6)
where $`v_1`$, $`v_2`$, and $`v_3`$ are the vevs of $`H_1^0`$, $`H_2^0`$, and $`\stackrel{~}{\nu }_\tau `$, respectively. As in the MSSM, the chargino mass matrix is diagonalized by two rotation matrices $`𝐔`$ and $`𝐕`$
$$𝐔^{}𝐌_𝐂𝐕^1=\left[\begin{array}{ccc}m_{\stackrel{~}{\chi }_1^\pm }& 0& 0\\ 0& m_{\stackrel{~}{\chi }_2^\pm }& 0\\ 0& 0& m_\tau \end{array}\right].$$
(7)
The lightest eigenstate of this mass matrix must be the tau lepton ($`\tau ^\pm `$) and so the mass is constrained to be 1.7771 GeV. As explained in , the tau Yukawa coupling becomes a function of the SUSY parameters appearing in the mass matrix.
The neutralino mass matrix is given by:
$$𝐌_𝐍=\left[\begin{array}{ccccc}M_1& 0& g_1v_1& g_1v_2& g_1v_3\\ 0& M_2& g_2v_1& g_2v_2& g_2v_3\\ g_1v_1& g_2v_1& 0& \mu & 0\\ g_1v_2& g_2v_2& \mu & 0& ϵ_3\\ g_1v_3& g_2v_3& 0& ϵ_3& 0\end{array}\right],$$
(8)
where $`g_1=g^{}/2`$ and $`g_2=g/2`$ denote gauge couplings. This matrix is diagonalised by a $`5\times 5`$ unitary matrix N,
$$\chi _i^0=N_{ij}\psi _j^0,$$
(9)
where $`\psi _j^0=(i\stackrel{~}{B},i\stackrel{~}{W}_3,\stackrel{~}{H}_d,\stackrel{~}{H}_u,\nu _\tau `$).
The up squark mass matrix is given by
$$M_{\stackrel{~}{u}}^2=\left[\begin{array}{cc}M_Q^2+\frac{1}{2}v_2^2h_{u}^{}{}_{}{}^{2}+\mathrm{\Delta }_{UL}& \frac{h_u}{\sqrt{2}}\left(v_2A_u\mu v_1+ϵ_3v_3\right)\\ \frac{h_u}{\sqrt{2}}\left(v_2A_u\mu v_1+ϵ_3v_3\right)& M_U^2+\frac{1}{2}v_2^2h_{u}^{}{}_{}{}^{2}+\mathrm{\Delta }_{UR}\end{array}\right]$$
(10)
and the down squark mass matrix by
$$M_{\stackrel{~}{d}}^2=\left[\begin{array}{cc}M_Q^2+\frac{1}{2}v_1^2h_{d}^{}{}_{}{}^{2}+\mathrm{\Delta }_{DL}& \frac{h_d}{\sqrt{2}}\left(v_1A_d\mu v_2\right)\\ \frac{h_d}{\sqrt{2}}\left(v_1A_d\mu v_2\right)& M_D^2+\frac{1}{2}v_1^2h_{d}^{}{}_{}{}^{2}+\mathrm{\Delta }_{DR}\end{array}\right]$$
(11)
with $`\mathrm{\Delta }_{UL}=\frac{1}{8}\left(g^2\frac{1}{3}g_{}^{}{}_{}{}^{2}\right)\left(v_1^2v_2^2+v_3^2\right)`$, $`\mathrm{\Delta }_{DL}=\frac{1}{8}\left(g^2\frac{1}{3}g_{}^{}{}_{}{}^{2}\right)\left(v_1^2v_2^2+v_3^2\right)`$, $`\mathrm{\Delta }_{UR}=\frac{1}{6}g_{}^{}{}_{}{}^{2}(v_1^2v_2^2+v_3^2)`$, and $`\mathrm{\Delta }_{DR}=\frac{1}{12}g_{}^{}{}_{}{}^{2}(v_1^2v_2^2+v_3^2)`$. The sum of the $`v_i^2`$ is given by $`m_W^2=g^2(v_1^2+v_2^2+v_3^2)/2`$. The mass eigenstates are given by $`\stackrel{~}{q}_1=\stackrel{~}{q}_L\mathrm{cos}\theta _{\stackrel{~}{q}}+\stackrel{~}{q}_R\mathrm{sin}\theta _{\stackrel{~}{q}}`$ and $`\stackrel{~}{q}_2=\stackrel{~}{q}_R\mathrm{cos}\theta _{\stackrel{~}{q}}\stackrel{~}{q}_L\mathrm{sin}\theta _{\stackrel{~}{q}}.`$ The sfermion mixing angle is given by
$$\mathrm{cos}\theta _{\stackrel{~}{q}}=\frac{M_{\stackrel{~}{q}_{LR}}^2}{\sqrt{(M_{\stackrel{~}{q}_{LL}}^2m_{\stackrel{~}{q}_1}^2)^2+(M_{\stackrel{~}{q}_{LR}}^2)^2}},\mathrm{sin}\theta _{\stackrel{~}{q}}=\frac{M_{\stackrel{~}{q}_{LL}}^2m_{\stackrel{~}{q}_1}^2}{\sqrt{(M_{\stackrel{~}{q}_{LL}}^2m_{\stackrel{~}{q}_1}^2)^2+(M_{\stackrel{~}{q}_{LR}}^2)^2}}.$$
(12)
Here $`M_{\stackrel{~}{q}_{ij}}^2`$ are the corresponding entries of the mass matrices in Eqs. (10) and (11).
In addition the charged Higgs bosons mix with charged sleptons and the real (imaginary) parts of the sneutrino mix the scalar (pseudoscalar) Higgs bosons. The formulas can be found in and are reproduced, for completeness, in the appendix. The corresponding mass eigenstates are denoted by $`S_i^+`$ for the charged scalars, $`S_j^0`$ for the neutral scalars, and $`P_k^0`$ for the pseudoscalars.
## 3 Numerical results
In this section we present numerical predictions for the lightest and second lightest neutralino production cross sections in $`e^+e^{}`$ collisions, namely, $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0,\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_2^0`$. Moreover we will characterize in detail all branching ratios for the lightest neutralino decays, which violate R-parity.
The relevant parameters include the $`R_p/`$ parameters and the standard mSUGRA parameters $`M_{1/2}`$, $`m_0`$, $`\mathrm{tan}\beta `$, where $`M_{1/2}`$ is the common gaugino mass, $`m_0`$ the common scalar mass, and $`\mathrm{tan}\beta =v_2/v_1`$ is the ratio of the vacuum expectation values of the Higgs fields. The absolute value of $`\mu `$ is fixed by radiative breaking of electroweak symmetry. We take $`\mu `$ positive to be in agreement with the $`bs\gamma `$ decay . As representative values of $`\mathrm{tan}\beta `$ we take $`\mathrm{tan}\beta =3`$ and 50. It is a feature of models with purely spontaneous breaking of R–parity that neutrinos acquire a mass only due to the violation of R-parity . This feature also applies to models characterized by purely bilinear breaking of R–parity, like our reference model charaterized by Eqs. (4) and (5). As a result the $`R_p/`$ violating parameters are directly related with $`m_{\nu _3}`$, the mass of the neutrino $`\nu _3`$, which is generated due to the mixing implicit in Eq. (8).
### 3.1 Neutralino Production
While the violation of R–parity would allow for the single production of supersymmetric particles , for the assumed values of the $`R_p/`$ violation parameters indicated by the simplest interpretation of solar and atmospheric neutrino data , these cross sections are typically too small to be observable. As a result neutralino production at LEP2 in our model typically occurs in pairs with essentially the same cross sections as in the mSUGRA case. In Fig. 1a and b we show the maximum and minimum attainable values for the $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0`$ and $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_2^0`$ production cross sections as a function of $`m_{\stackrel{~}{\chi }_1^0}`$ at $`\sqrt{s}=205`$ GeV. We compare the cases $`\mathrm{tan}\beta =3`$ and $`\mathrm{tan}\beta =50`$, varying $`M_{1/2}`$ between 90 GeV and 260 GeV and $`m_0`$ between 50 GeV and 500 GeV. One can see that, indeed, these results are identical to those obtained in the mSUGRA. The $`\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0`$ production cross section can reach approximately 1 pb. In our calculation we have used the formula as given in and, in addition, we have included initial state radiation (ISR) using the formula given in . Note that $`\stackrel{~}{e}_L`$ and $`\stackrel{~}{e}_R`$ are exchanged in the $`t`$\- and $`u`$-channel implying that a large fraction of the neutralinos will be produced in the forward and backward directions.
In order to show more explicitly the dependence of the cross sections on the parameters $`m_0`$ and $`M_{1/2}`$ we plot in Fig. 2a and b the contour lines of $`\sigma (e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0)`$ in the $`m_0`$-$`M_{1/2}`$ plane at $`\sqrt{s}=205`$ GeV for $`\mathrm{tan}\beta =3`$ and $`\mathrm{tan}\beta =50`$. The contour lines for $`\sigma (e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_2^0)`$ are given in Fig. 2c and d.
### 3.2 Neutralino Decay Length
If unprotected by the ad hoc assumption of R–parity conservation the LSP will decay as a result of gauge boson, squark, slepton and Higgs boson exchanges. The relevant contributions to these decays are given in Table 1. The Feynman diagrams for the decays not involving taus, i.e. $`\stackrel{~}{\chi }_1^0\nu _3f\overline{f}`$ ($`f=e`$, $`\nu _e`$, $`\mu `$, $`\nu _\mu `$, $`u`$, $`d`$, $`c`$, $`s`$, $`b`$) are shown explicitly in Fig. 3.
For the observability of the R-parity violating effects it is crucial that with this choice of parameters the LSP will decay most of the time inside the detector. The neutralino decay path expected at LEP2 depends crucially on the values of $`R_p/`$ violating parameters or, equivalently, on the value of the heaviest neutrino mass, $`m_{\nu _3}`$. We fix the value of $`m_{\nu _3}`$ as indicated by the analyses of the atmospheric neutrino data . It is important to note that, as explained in , due to the projective nature of the neutrino mass matrix , only one of the three neutrinos picks up a mass in tree approximation. This means that, neglecting radiative corrections which give small masses to the first two neutrinos in order to account for the solar neutrino data, the neutralino decay length scale is set mainly by the tree–level value of $`m_{\nu _3}`$. In ref. we have explicitly shown that this is a good approximation for most points in parameter space.
In Fig. 4 we plot the $`\stackrel{~}{\chi }_1^0`$ decay length in cm expected at LEP2 for $`\sqrt{s}=205`$ GeV. Here and later on we consider the neutralinos stemming from the process $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0`$ when discussing the decay length. In Fig. 4a we plot the $`\stackrel{~}{\chi }_1^0`$ decay length in cm as a function of neutrino mass $`m_{\nu _3}`$, for different $`m_{\stackrel{~}{\chi }_1^0}`$ between 60 and 90 GeV, with $`m_0=100`$ GeV, and $`\mathrm{tan}\beta =3`$. As can be seen the expected neutralino decay length is typically such that the decays occur inside the detector, leading to a drastic modification of the mSUGRA signals. An equivalent way of presenting the neutralino decay path at LEP2 is displayed in Fig. 4b, which gives the decay length of $`\stackrel{~}{\chi }_1^0`$ as a function of $`m_{\stackrel{~}{\chi }_1^0}`$ for $`m_{\nu _3}=0.01`$, 0.1, and 1 eV.
Finally, we show the dependence of the neutralino decay path on the supergravity parameters fixing the magnitude of $`R_p/`$ violating parameters or, equivalently, the magnitude of the heaviest neutrino mass, $`m_{\nu _3}`$. In Fig. 5a and b we plot the contour lines of the decay length of $`\stackrel{~}{\chi }_1^0`$ in the $`m_0`$-$`M_{1/2}`$ plane for $`m_{\nu _3}=0.06`$ eV, $`\mathrm{tan}\beta =3`$ and 50. Note that the decay length is short enough that it may happen inside typical high energy collider detectors even for the small neutrino mass values $`0.06`$ eV indicated by the atmospheric neutrino data . For large values of $`\mathrm{tan}\beta `$ the total decay width increases and, correspondingly, the decay path decreases due to the tau Yukawa coupling and the bottom Yukawa coupling.
### 3.3 Neutralino Branching Ratios
As discussed in the beginning of this section, the lightest neutralino $`\stackrel{~}{\chi }_1^0`$ will typically decay in the detector. In the following we present our results for the branching ratios of all R-parity violating 3-body decay of $`\stackrel{~}{\chi }_1^0`$, and of the radiative decay $`\stackrel{~}{\chi }_1^0\nu _3\gamma `$. The Feynman diagrams for the decays $`\stackrel{~}{\chi }_1^0\nu _3f\overline{f}`$ ($`f=e`$, $`\nu _e`$, $`\mu `$, $`\nu _\mu `$, $`u`$, $`d`$, $`c`$, $`s`$, $`b`$) are shown in Fig. 3. For this class of decays we have $`Z^0`$, $`P_i^0`$, and $`S_j^0`$ exchange in the direct channel (Fig. 3a and b) and $`\stackrel{~}{f}`$ exchange in the crossed channels (Fig. 3c and d). In particular in the case $`f=b`$ the $`P_i^0`$ and $`S_j^0`$ exchange contributions are significant. This is quite analogous to the results found in for $`\stackrel{~}{\chi }_2^0\stackrel{~}{\chi }_1^0f\overline{f}`$ decays. The particles exchanged in the $`s`$-, $`t`$-, and $`u`$-channel for the decays $`\stackrel{~}{\chi }_1^0\tau ^\pm l^{}\nu _l`$, ($`l=e,\mu `$), $`\stackrel{~}{\chi }_1^0\tau ^\pm q\overline{q}^{}`$ ($`q,q^{}=u,d,s,c`$), $`\stackrel{~}{\chi }_1^0\tau ^{}\tau ^+\nu _l`$, and $`\stackrel{~}{\chi }_1^03\nu _3`$ are given in Table 1 .
In the calculations we have included all mixing effects, in particular the standard MSSM $`\stackrel{~}{f}_L\stackrel{~}{f}_R`$ mixing effects and those induced by the bilinear R-parity violating terms, i.e. $`Re(\stackrel{~}{\nu }_\tau )h^0H^0`$, $`Im(\stackrel{~}{\nu }_3)A^0G^0`$, , $`\stackrel{~}{\tau }_{L,R}^\pm H^\pm G^\pm `$ , $`\nu _\tau `$ \- $`\stackrel{~}{\chi }_i^0`$ , and $`\tau `$ \- $`\stackrel{~}{\chi }_j^{}`$ mixings . These mixing effects are particularly important in the calculations of the various R-parity violating decay rates of $`\stackrel{~}{\chi }_1^0`$, which are discussed below.
In the following plots Fig. 6 \- 13 we show contour lines in the $`m_0`$-$`M_{1/2}`$ plane for the branching ratios in % of the various $`\stackrel{~}{\chi }_1^0`$ decays, in (a) for $`\mathrm{tan}\beta =3`$ and in (b) for $`\mathrm{tan}\beta =50`$. We have fixed the mass of the heaviest neutrino to $`m_{\nu _3}=0.06`$ eV . It turns out, that in the range $`10^2`$ eV $`m_{\nu _3}1`$ keV all the $`\stackrel{~}{\chi }_1^0`$ decay branching ratios are rather insensitive to the actual value of $`m_{\nu _3}`$. This is an important feature of our supergravity–type R-parity violating model. It is a consequence of the fact that, as a result of the universal supergravity boundary conditions on the soft breaking terms, all R-parity violating couplings are proportional to a unique common parameter which may be taken as $`ϵ_3/\mu `$. For a more detailed discussion on this proportionality the reader is referred to ref. . Also note that for $`M_{1/2}>`$ 220 GeV the neutralino mass becomes larger than $`m_W`$ and $`m_Z`$ so that $`\stackrel{~}{\chi }_1^0`$ decays into real $`W`$ and $`Z`$ are possible. The effects of these real decays can be seen for $`M_{1/2}>`$ 220 GeV in most of the following plots. For the large $`\mathrm{tan}\beta `$ case ($`\mathrm{tan}\beta =50`$) and $`M_{1/2}M_0`$ the mass of the lighter charged boson $`S_1^\pm `$ is smaller than $`m_{\stackrel{~}{\chi }_1^0}`$ (upper left corner of Fig. 6b - 13b). In this region of the parameter space the two 2-body decays $`\stackrel{~}{\chi }_1^0W^\pm \tau _\pm `$ and $`\stackrel{~}{\chi }_1^0S_1^\pm \tau _\pm `$ compete. The first one is R-parity violating, but has more phase space than the second one which is R-parity conserving, since $`S_1^\pm `$ is mainly a stau. For this reason, the most import final state is $`\tau ^+\tau ^{}\nu _3`$, followed by $`\tau ^\pm q\overline{q}^{}`$ and $`\tau ^\pm l^{}\nu _i`$ ($`l=e,\mu `$) as shown in Figs. 12, 10, and 9, respectively. All other final states have nearly vanishing branching ratios in this corner of the parameter space.
Fig. 6a and b exhibit the contour lines for the branching ratio of the invisible decay $`\stackrel{~}{\chi }_1^03\nu `$. This branching ratio can reach 7% for the parameters chosen. In Fig. 7 and 8 we show the branching ratio for the decays $`\stackrel{~}{\chi }_1^0\nu _3l^+l^{}`$ and $`\stackrel{~}{\chi }_1^0\nu _3q\overline{q}`$ where $`l`$ and $`q`$ denote the leptons and quarks of the first two generations, summed over all flavors. These branching ratios can go up to 3% and 15%, respectively. Notice that the sneutrino, slepton, and squark exchange contributions to the $`\stackrel{~}{\chi }_1^0`$ decays become larger with increasing $`m_0`$, despite the fact that the increase of the scalar masses $`m_{\stackrel{~}{\nu }}`$, $`m_{\stackrel{~}{l}}`$, $`m_{\stackrel{~}{q}}`$ suppresses these exchange contributions. This trend can also be observed in Fig. 8, 9 and 10. This happens because the tadpole equations correlate $`\mu `$ to $`m_0`$. Increasing $`\mu `$ while keeping $`M_1`$ and $`M_2`$ fixed implies increasing the gaugino content of $`\stackrel{~}{\chi }_1^0`$ and, hence, enhancing the $`\stackrel{~}{\chi }_1^0`$-$`f`$-$`\stackrel{~}{f}`$ couplings.
In Fig. 9 and 10 we show the contour lines for the branching ratios of the LSP decays involving a single tau, namely $`\stackrel{~}{\chi }_1^0\nu _l\tau ^\pm l^{}`$ and $`\stackrel{~}{\chi }_1^0\tau ^\pm q\overline{q}^{}`$, where $`l`$, $`q`$, and $`q^{}`$ are summed over the first two generations. The branching for these decay modes can reach up to 20% and 60% respectively. For $`M_{1/2}>220`$ GeV decays into real $`W^\pm `$ dominate. If this is the case and if both $`\stackrel{~}{\chi }_1^0`$ produced in $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0`$ decay according to these modes this would lead to very distinctive final states, such as $`4j\tau ^+\tau ^+`$, $`\tau ^+\tau ^+l^{}l^{}`$ ($`l=e,\mu `$), or $`\tau ^+\tau ^+e^{}\mu ^{}`$. The full list of expected signals is given in Table 2 . The first column in this table specifies the two pairs of $`\stackrel{~}{\chi }_1^0`$ decay modes, while the second one gives the corresponding signature. In the last column we state whether the corresponding signature exists for $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_2^0`$ production within mSUGRA.
The LSP decays involving only third generation fermions, namely, $`\stackrel{~}{\chi }_1^0\nu _3b\overline{b}`$ and $`\stackrel{~}{\chi }_1^0\nu _3\tau ^+\tau ^{}`$ are different from those into the first and second generation fermion pairs, because the Higgs boson exchanges and the Yukawa terms play a very important rôle. This can be seen in Fig. 11 and 12 , where we plot the contour lines for these decays. The branching ratio of $`\stackrel{~}{\chi }_1^0\nu _3b\overline{b}`$ can reach up to 97%. The decay rate is large because the scalar exchange contributions ($`S_j^0,P_j^0,\stackrel{~}{b}_k`$) are large for $`M_{1/2}<200`$ GeV. Note that this is also the case for $`\mathrm{tan}\beta =3`$, because not only the neutrino-neutralino mixing proportional to $`m_{\nu _3}`$ is important but also the neutrino-higgsino mixing proportional to $`ϵ_3/\mu `$. The decrease of the branching ratio with increasing $`m_0`$ is due to the decrease of the higgsino component of $`\stackrel{~}{\chi }_1^0`$ and the increase of the Higgs boson masses. For $`M_{1/2}>200`$ GeV the decays into real $`W^+`$ and $`Z^0`$ are possible, reducing the branching ratio of $`\stackrel{~}{\chi }_1^0\nu _3b\overline{b}`$. As shown in Fig. 12 the branching ratio for $`\stackrel{~}{\chi }_1^0\nu _3\tau ^+\tau ^{}`$ is very small for $`\mathrm{tan}\beta =3`$ and $`M_{1/2}<200`$ GeV. This is due to the destructive interference between $`Z^0`$ contribution and the contributions of the exchanged charged scalar particles (mainly due to the stau components of $`S_k^\pm `$).
Finally we have also considered the radiative LSP decay mode $`\stackrel{~}{\chi }_1^0\nu _3\gamma `$ . In Fig. 13 the branching ratio for this mode is shown. This decay proceeds only at one-loop level and therefore is in general suppressed compared to the three-body decay modes. However, for $`M_{1/2}<125`$ GeV and large $`\mathrm{tan}\beta `$ it exceeds 1%, leading to interesting signatures like $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0\tau ^\pm \mu ^{}\gamma `$ \+ $`p_T/`$ . Due to initial state radiation it can easily happen that a second photon is observed in the same event.
The complete list of possible signatures stemming from LSP decays in our bilinear $`R_p/`$ model is shown in Table 2 . In this table we also indicate whether the same signatures could also arise in mSUGRA as a result of $`e^+e^{}\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_2^0`$ followed by the MSSM decay modes of $`\stackrel{~}{\chi }_2^0`$ if its production is kinematically allowed. The final states 4 jets + $`p_T/`$ , $`\tau `$ \+ 2 jets + $`p_T/`$ , and $`\tau `$ \+ ($`e`$ or $`\mu `$) + $`p_T/`$ would also occur in mSUGRA via the decay of $`\stackrel{~}{\chi }_2^0`$ into $`\stackrel{~}{\chi }_1^\pm `$. However one expects in general that these decay modes are suppressed within mSUGRA. In contrast in the R-parity violating case these signatures can be rather large as can be seen from Fig. 9 and 10. Note moreover, that some of the $`R_p/`$ signatures are practically background free. For example, due to the Majorana nature of $`\stackrel{~}{\chi }_1^0`$, one can have two same–sign $`\tau `$ leptons + 4 jets + $`p_T/`$ . Other interesting signals are: $`\tau `$ \+ 3 ($`e`$ and/or $`\mu `$) + $`p_T/`$ , 3 $`\tau `$ \+ ($`e`$ or $`\mu `$) + $`p_T/`$ , $`\tau `$ \+ ($`e`$ or $`\mu `$) + 2 jets + $`p_T/`$ , $`\tau `$ \+ 4 jets + $`p_T/`$ , $`\tau ^\pm \tau ^\pm `$ \+ ($`e`$ or $`\mu `$) + 2 jets + $`p_T/`$ , or $`\tau ^\pm \tau ^\pm `$ \+ $`l^{}l_{}^{}{}_{}{}^{}`$ \+ $`p_T/`$ with $`l=e,\mu `$. In Table 3 we give masses and branching ratios for typical examples.
As it is well known, also in gauge mediated supersymmetry breaking models (GMSB) the neutralino can decay inside the detector, because the gravitino $`\stackrel{~}{G}`$ is the LSP. It is therefore an interesting question if the R-parity violating model can be confused with GMSB. To answer this question let us have a look at the dominant decay modes of the lightest neutralino in GMSB. If the lightest neutralino is the NLSP, its main decay mode in GMSB is
$`\stackrel{~}{\chi }_1^0\gamma \stackrel{~}{G},`$
where $`\stackrel{~}{G}`$ is the gravitino. For the case where at least one of the sleptons is lighter than the lightest neutralino the latter has the following decay chain $`\stackrel{~}{\chi }_1^0\stackrel{~}{l}^\pm l^{}l^\pm l^{}\stackrel{~}{G},`$. In principle three-body decay modes mediated by virtual photon, virtual Z-boson and virtual sfermions also exist. However, in the neutralino mass range considered here these decays are phase–space–supressed . This implies that the R-parity violating model can not be confused with GMSB, because (i) in GMSB the final states containing quarks are strongly suppressed, and (ii) GMSB with conserved R-parity implies lepton flavour conservation, and therefore there are no final states like $`e^+e^+\tau ^{}\tau ^{}`$ \+ $`p_T/`$ . A further interesting question would be how the neutralino phenomenology changes in a GMSB scenario with broken R-parity. The main consequence would be an enhancement of final states containing photons and/or leptons. A detailed study of this question is, however, beyond the scope of the present paper.
## 4 Conclusions
We have studied the production of the lightest neutralino $`\stackrel{~}{\chi }_1^0`$ at LEP2 and the resulting phenomenology in models where an effective bilinear term in the superpotential parametrizes the explicit breaking of R-parity. We have considered supergravity scenarios which can be explored at LEP2 in which the lightest neutralino is also the lightest supersymmetric particle. We have presented a detailed study of the LSP $`\stackrel{~}{\chi }_1^0`$ decay properties and studied the general features of the corresponding signals expected at LEP2. A detailed investigation of the possible detectability of the signals discussed in Table 2 taking into account realistic detector features is beyond the scope of this paper. Clearly, existing LEP2 data are already probing the part of the parameter region which corresponds to approximately $`m_{\stackrel{~}{\chi }_1^0}<40`$ GeV. Finally, we note that, in addition to important modifications in the $`\stackrel{~}{\chi }_1^0`$ decay properties, R-parity violating decay models lead also to new interesting features in other decays, such as charged and neutral Higgs boson and slepton decays, stop decays , and gluino cascade decays . In addition we have shown that the R-parity violating model can not be confused with gauge mediated supersymmetry breaking and conserved R-parity due to the absence of several final states in the GMSB case.
## Acknowledgments
This work was supported by “Fonds zur Förderung der wissenschaftlichen Forschung” of Austria, project No. P13139-PHY, by Spanish DGICYT grants PB98-0693 and by the EEC under the TMR contract HPRN-CT-2000-00148. W.P. was supported by a fellowship from the Spanish Ministry of Culture under the contract SB97-BU0475382. D.R. was supported by Colombian COLCIENCIAS fellowship.
## Appendix A Scalar Mass Matrices
The mass matrix of the charged scalar sector follows from the quadratic terms in the scalar potential .
$$V_{quadratic}=𝐒_{}^{}{}_{}{}^{}𝐌_{𝐒^\pm }^\mathrm{𝟐}𝐒_{}^{}{}_{}{}^{+}$$
(13)
where$`𝐒_{}^{}{}_{}{}^{}=[H_1^{},H_2^{},\stackrel{~}{\tau }_L^{},\stackrel{~}{\tau }_R^{}]`$. For convenience reasons we will divide this $`4\times 4`$ matrix into $`2\times 2`$ blocks in the following way:
$$𝐌_{𝐒^\pm }^\mathrm{𝟐}=\left[\begin{array}{ccc}𝐌_{\mathrm{𝐇𝐇}}^\mathrm{𝟐}& 𝐌_{𝐇\stackrel{~}{\tau }}^{\mathrm{𝟐}}{}_{}{}^{T}& \\ 𝐌_{𝐇\stackrel{~}{\tau }}^\mathrm{𝟐}& 𝐌_{\stackrel{~}{\tau }\stackrel{~}{\tau }}^\mathrm{𝟐}& \end{array}\right]$$
(14)
where the charged Higgs block is
$`𝐌_{\mathrm{𝐇𝐇}}^\mathrm{𝟐}=`$ (15)
$`\left[\begin{array}{ccc}B\mu \frac{v_2}{v_1}+\frac{1}{4}g^2(v_2^2v_3^2)+\mu ϵ_3\frac{v_3}{v_1}+\frac{1}{2}h_\tau ^2v_3^2+\frac{t_1}{v_1}& B\mu +\frac{1}{4}g^2v_1v_2& \\ B\mu +\frac{1}{4}g^2v_1v_2& B\mu \frac{v_1}{v_2}+\frac{1}{4}g^2(v_1^2+v_3^2)B_3ϵ_3\frac{v_3}{v_2}+\frac{t_2}{v_2}& \end{array}\right]`$ (18)
and $`h_\tau `$ is the tau Yukawa coupling.
$`𝐌_{\stackrel{~}{\tau }\stackrel{~}{\tau }}^\mathrm{𝟐}=`$ (19)
$`\left[\begin{array}{ccc}\frac{1}{2}h_\tau ^2v_1^2\frac{1}{4}g^2(v_1^2v_2^2)+\mu ϵ_3\frac{v_1}{v_3}B_3ϵ_3\frac{v_2}{v_3}+\frac{t_3}{v_3}& \frac{1}{\sqrt{2}}h_\tau (A_\tau v_1\mu v_2)& \\ \frac{1}{\sqrt{2}}h_\tau (A_\tau v_1\mu v_2)& m_{R_3}^2+\frac{1}{2}h_\tau ^2(v_1^2+v_3^2)\frac{1}{4}g^2(v_1^2v_2^2+v_3^2)& \end{array}\right]`$ (22)
The mixing between the charged Higgs sector and the stau sector is given by the following $`2\times 2`$ block:
$$𝐌_{𝐇\stackrel{~}{\tau }}^\mathrm{𝟐}=\left[\begin{array}{ccc}\mu ϵ_3\frac{1}{2}h_\tau ^2v_1v_3+\frac{1}{4}g^2v_1v_3& B_3ϵ_3+\frac{1}{4}g^2v_2v_3& \\ \frac{1}{\sqrt{2}}h_\tau (ϵ_3v_2+A_\tau v_3)& \frac{1}{\sqrt{2}}h_\tau (\mu v_3+ϵ_3v_1)& \end{array}\right]$$
(23)
As we see the charged Higgs bosons mix with charged sleptons.
In a similar way the real (imaginary) parts of the sneutrino mix the scalar (pseudoscalar) Higgs bosons. The quadratic scalar potential responsible for the neutral Higgs sector mass matrices includes
$$V_{quadratic}=\frac{1}{2}𝐏_{}^{}{}_{}{}^{\mathrm{𝟎}}{}_{}{}^{T}𝐌_{𝐏^\mathrm{𝟎}}^\mathrm{𝟐}𝐏_{}^{}{}_{}{}^{\mathrm{𝟎}}+𝐒_{}^{}{}_{}{}^{\mathrm{𝟎}}{}_{}{}^{T}𝐌_{𝐒^\mathrm{𝟎}}^\mathrm{𝟐}𝐒_{}^{}{}_{}{}^{\mathrm{𝟎}}+\mathrm{}$$
(24)
where $`𝐏_{}^{}{}_{}{}^{\mathrm{𝟎}}{}_{}{}^{T}=[\phi _1^0,\phi _2^0,\stackrel{~}{\nu }_\tau ^I]`$, $`𝐒_{}^{}{}_{}{}^{\mathrm{𝟎}}{}_{}{}^{T}=\frac{1}{2}[\chi _1^0,\chi _2^0,\stackrel{~}{\nu }_\tau ^R]`$ and the CP-odd neutral scalar mass matrix is
$$𝐌_{𝐏^\mathrm{𝟎}}^\mathrm{𝟐}=\left[\begin{array}{ccc}B\mu \frac{v_2}{v_1}+\mu ϵ_3\frac{v_3}{v_1}+\frac{t_1}{v_1}& B\mu & \mu ϵ_3\\ B\mu & B\mu \frac{v_1}{v_2}B_3ϵ_3\frac{v_3}{v_2}+\frac{t_2}{v_2}& B_3ϵ_3\\ \mu ϵ_3& B_3ϵ_3& \mu ϵ_3\frac{v_1}{v_3}B_3ϵ_3\frac{v_2}{v_3}+\frac{t_3}{v_3}\end{array}\right]$$
(25)
The neutral CP-even scalar sector mass matrix in eq. (24) is given by
$`𝐌_{𝐒^\mathrm{𝟎}}^\mathrm{𝟐}=`$ (26)
$`\left[\begin{array}{ccc}B\mu \frac{v_2}{v_1}+\frac{1}{4}g_Z^2v_1^2+\mu ϵ_3\frac{v_3}{v_1}+\frac{t_1}{v_1}& B\mu \frac{1}{4}g_Z^2v_1v_2& \mu ϵ_3+\frac{1}{4}g_Z^2v_1v_3\\ B\mu \frac{1}{4}g_Z^2v_1v_2& B\mu \frac{v_1}{v_2}+\frac{1}{4}g_Z^2v_2^2B_3ϵ_3\frac{v_3}{v_2}+\frac{t_2}{v_2}& B_3ϵ_3\frac{1}{4}g_Z^2v_2v_3\\ \mu ϵ_3+\frac{1}{4}g_Z^2v_1v_3& B_3ϵ_3\frac{1}{4}g_Z^2v_2v_3& \mu ϵ_3\frac{v_1}{v_3}B_3ϵ_3\frac{v_2}{v_3}+\frac{1}{4}g_Z^2v_3^2+\frac{t_3}{v_3}\end{array}\right]`$ (30)
where we have defined $`g_Z^2g^2+g^2`$. Note that, as a result of CP invariance, the CP–even and CP–odd parts of the scalar mass matrices are disjoint and do not mix with each other.
The three mass matrices in eqs. (14), (25), and (26) are diagonalized by rotation matrices which define the eigenvectors
$$𝐒^+=𝐑_{𝐒^\pm }𝐒_{}^{}{}_{}{}^{+},𝐏^\mathrm{𝟎}=𝐑_{𝐏^\mathrm{𝟎}}𝐏_{}^{}{}_{}{}^{\mathrm{𝟎}},𝐒^\mathrm{𝟎}=𝐑_{𝐒^\mathrm{𝟎}}𝐒_{}^{}{}_{}{}^{\mathrm{𝟎}},$$
(31)
and the eigenvalues $`\mathrm{diag}(0,\mathrm{m}_{\mathrm{S}_2^\pm }^2,\mathrm{m}_{\mathrm{S}_3^\pm }^2,\mathrm{m}_{\mathrm{S}_4^\pm }^2)=𝐑_{𝐒^\pm }𝐌_{𝐒^\pm }^\mathrm{𝟐}𝐑_{𝐒^\pm }^𝐓`$ for the charged scalars, $`\mathrm{diag}(0,\mathrm{m}_{\mathrm{P}_2^0}^2,\mathrm{m}_{\mathrm{P}_3^0}^2)`$ $`=𝐑_{𝐏^\mathrm{𝟎}}𝐌_{𝐏^\mathrm{𝟎}}^\mathrm{𝟐}𝐑_{𝐏^\mathrm{𝟎}}^𝐓`$ for the CP–odd neutral scalars, and $`\mathrm{diag}(\mathrm{m}_{\mathrm{S}_1^0}^2,\mathrm{mS}_{2}^{0}{}_{}{}^{2},\mathrm{m}_{\mathrm{S}_3^0}^2)`$ $`=𝐑_{𝐒^\mathrm{𝟎}}𝐌_{𝐒^\mathrm{𝟎}}^\mathrm{𝟐}𝐑_{𝐒^\mathrm{𝟎}}^𝐓`$ for the CP–even neutral scalars.
The matrices $`R_{S^\pm }`$, $`R_{P^0}`$ and $`R_{S^0}`$ specify the mixing between the Higgs sector and the stau sector.
If a $`3\times 3`$ matrix $`𝐌`$ has a zero eigenvalue, then the other two eigenvalues satisfy
$$m_\pm =\frac{1}{2}\mathrm{Tr}𝐌\pm \frac{1}{2}\sqrt{\left(\mathrm{Tr}𝐌\right)^24(M_{11}M_{22}M_{12}^2+M_{11}M_{33}M_{13}^2+M_{22}M_{33}M_{23}^2)}$$
(32)
The CP-odd neutral scalar mass matrix eq. (25) has a zero determinant, so that its eigenvalues $`m_{P_2^0}^2`$ and $`m_{P_3^0}^2`$ ($`m_A^2`$ and $`m_{\stackrel{~}{\nu }_\tau ^R}^2`$ in the MSSM limit) can be calculated exactly with the previous formula.
|
warning/0007/astro-ph0007448.html
|
ar5iv
|
text
|
# MBM12: A Younger Version of the TW Hydrae Association?
## 1. Introduction
MBM 12 is a dark, high latitude cloud located about 65 pc from the sun (Hobbs, Blitz & Magnani 1986, but see Hearty et al. 2000). From the early 1970s through early 1990s, five Classical T Tauri stars (cTTs) and three naked T Tauri stars (nTTs) were discovered associated with the cloud (see Hearty et al. 1999 and references therein). These sources have been discovered through a combination of H$`\alpha `$ objective prism surveys and X-ray surveys. Such a close, young, isolated association would be in a league with the TW Hydrae Association (Kastner et al. 1997) and Eta Cha (Mamajek et al. 1999 and Lawson et al. this volume). However, the co–existence of the cloud and several stars with disks implies this is a very young association of stars which are very close to the earth. The 10-Myr-old TW Hydrae group has proved a fruitful region for high resolution imaging of disks and close binary companions (Jayawardhana 2000). At an estimated age of about 1 Myr, in MBM 12 brown dwarfs/Jupiters would be brighter and disks should be more numerous.
As a close, possibly active region of star formation, still nestled near its natal cloud, MBM 12 vitally requires a census of its young stellar population . Here, we present the results of our study of the photometric properties of stars in this region.
## 2. Optical Photometry
We observed the stars near the cloud in optical light during December 1998 with the 48<sup>′′</sup> telescope at the Whipple Observatory using the “four-shooter” CCD detector. There were a total of 6 pointings made, each centered on a known T Tauri star. The field of view for each image was about 13.5. Five of the pointings coincided with the edges of the MBM 12 cloud. The sixth was 2 away, closer to MBM 13. Observations were made in U, B, V, R and I. There was a series of short 60 second V, R and I exposures, along with a deeper 120 to 450 second exposure in all five Kron-Cousin’s bands.
The data were corrected in the usual way using IRAF. The only complication was that the “four–shooter” uses four separated CCDs for data taking. For processing, the data from the individual chips were segregated. Aperture photometry was performed on each night’s worth of data on a per chip basis. This way, each chip had its own set of standard stars so that effects induced by a chip would be confined to that data set. Cross–referencing of stars observed on multiple chips showed chip to chip variations of less than 1%. The overall photometric accuracy was about 3% fainter than V=21. The data are complete to about V=21 and I =17.5.
The main results of these observations are shown in figure 1. About 10,000 stars were detected in V, R and I bands. Brighter than V=21, all but about 40 of these stars fall well below the demarcation for 10 Myr old stars at 65pc. We define these as our PMS candidates. There is a small ($``$0.5 mag.) break between the main body of stars and the PMS candidates. This could be the effect of clustering but is most likely an effect of volume limiting. The volume of space a star could be in, and occupy a given location on the color–magnitude diagram becomes relatively small this close to the sun.
Spatially the candidates are spread among the 6 fields fairly evenly. The field near MBM 13 does have fewer PMS candidates than the other 5 fields, but this number is still within 2$`\sigma `$ of the mean. We do not see any candidates along the edge of the cloud. This could be due to high extinction. But a very small fraction of the observed regions (only about 5%) overlaps the MBM 12 cloud. So this result could simply be happenstance.
## 3. 2MASS Photometry
We obtained several nights to survey MBM 12 from the Whipple Observatory using the STELLIRCAM near–infrared camera. Unfortunately the weather did not cooperate. Fortunately, the second incremental release of the 2MASS database covers the MBM 12 region. We queried the database in the region of the optical surveys (within 3600 arcsec radius of RA:3h57m DEC:20<sup>o</sup>15 J2000). This returned about 5500 sources. We then filtered out all the sources with errors greater than 10% or any bad quality flags. This left a little under 4000 sources with which we correlated with our catalog of $``$40 PMS candidates looking for matches within 3<sup>′′</sup>. We found 24 matches. To this we added the previously known TTs in this region; these stars were too bright to be measured with our optical photometry. Six of these 8 were in the 2MASS catalog. We also used 3 HIPPARCHOS stars as reference for normal stars along the line of sight.
The results are given in figure 2. Two stars are found in the lower–right region of the diagram, which indicates the existence of circumstellar disks. About half of the PMS candidates are found in the region of the diagram which indicates a normal photosphere with significant (A$`{}_{V}{}^{}>1`$) reddening. Being found in this region does not mean that these stars do not have disks; in fact the four known cTTs are found in this region as well. Nor does the substantial reddening found in the IR mean that the optical data are misleading. The reddening vector in figure 1 is parallel to the (pre–) main sequence, so reddening only affects the mass estimate. For this reason, we refrain from attempting any sort of mass function at this point. The remaining stars lie close to the main sequence locus.
The large sky coverage of the 2MASS database allowed us to study the spatial distribution of sources. Using the second incremental release point source catalog, we created pencil beam samples of the infrared sky at various locations along the same latitude as MBM 12. Each pencil beam was 20<sup>′′</sup> in radius and so covered 0.35<sup>∘2</sup>. One pencil beam was centered on the MBM 12 cloud itself, one on the field near MBM 13, and others were spread further out, but always on areas with warm dust as seen by the IRAS 60$`\mu `$m survey. Figure 3 shows that the pencil beam centered on the MBM 12 dark cloud has a much greater density of sources than the other regions. Using a 20<sup>′′</sup> pencil beam, the enhancement along MBM 12 is about 10$`\sigma `$. However, this result becomes even more significant if smaller pencil beams are used. The total enhancement is about 70 objects more than expected in the 20<sup>′′</sup> pencil beam.
## 4. Conclusions
The proximity and extreme youth of stars near the MBM 12 cloud makes them of great interest. While the known membership of the cluster is small, we have found evidence for dozens of additional cluster members near the cloud and perhaps 70 embedded within the cloud. Over the coming observing seasons, we hope to be able to follow up on this photometric survey with spectra which will confirm the PMS nature of the candidates. These new data will help us better ascertain the number and age of these nearby stars.
### Acknowledgments.
This work was supported in part by NASA contract NAS8-39073.
## References
D’Antona, F.& Mazzitelli, I. 1997 Mem. S.A.It., 68, 807
Hearty, T. et al. 2000, A&A, 357, 681
Hearty, T. et al. 2000, A&A, 353, 1044
Hobbs, L.M. et al. 1986, ApJ, 306, 380L
Jayawardhana, R. 2000, Science, 288, 64
Kastner, J.H., et al. 1997, Science, 277, 67
Mamajek et al. 1999, ApJ, 516, L77
Lawson et al. 2000, this volume
|
warning/0007/hep-lat0007019.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The temperature driven first order phase transition of the three-dimensional, three-state Potts model has been analyzed in great detail in the absence of a symmetry breaking external field. This transition remains first order in the presence of a non-vanishing external field and for a critical field strength, $`h_c`$, it ends in a second order critical point. Although it is expected that this critical point belongs to the universality class of the 3-d Ising model, the universal behaviour in its vicinity has not been analyzed in detail so far. A first estimate for the location of the critical endpoint is given in .
Our interest in properties of the 3-d, 3-state Potts model in the vicinity of the second order critical point is motivated by its importance for the analysis of lines of first order transitions in lattice gauge field models. The universal behaviour in the vicinity of the second order endpoint of the line of first order transitions has recently been investigated in the U(1)-Higgs and SU(2)-Higgs models. Lines of first order transitions with critical endpoints do, however, also occur in QCD with light as well as heavy quarks. In the heavy quark mass limit of finite temperature QCD a line of first order phase transitions (deconfinement transition) occurs which is closely related to the phase transition in the 3-d, 3-state Potts model in an external field . Like in the Potts model the deconfinement transition is first order in the limit of infinitely heavy quarks and ends in a second order transition at some finite value of the quark mass. This critical point is expected to belong to the universality class of the 3-d Ising model . Another line of first order transitions occurs in the case of QCD with three light quark flavours (chiral symmetry restoration). This line also ends in a second order endpoint at some critical value of the quark mass. In order to analyze the critical behaviour at these endpoints it will be important to disentangle the relevant energy- and ordering field like directions and identify the Ising-like observables at the critical points. We will address this problem here first in the simpler case of the Potts model and will explore methods used for the analysis of the liquid-gas phase transition as well as lattice gauge models .
In this letter we present an accurate determination of the critical couplings at the endpoint of the line of first order phase transitions in the 3-d, 3-state Potts model with an external ordering field. Moreover, we will determine the relevant energy-like and ordering field like directions at this endpoint as well as the related operators from which critical exponents and other universal constants (Binder cumulants) can be extracted. In the next section we fix our notation for the Potts model and introduce the new couplings and operators which control the critical behaviour at the endpoint. In section 3 we present our numerical results for the location of the critical endpoint. A more detailed discussion of the universal properties at this endpoint is given in Section 4. Finally we present our conclusions in Section 5.
## 2 The Model and Simulation Parameters
The three-state Potts model is described in terms of spin variables $`\sigma _i\{1,2,3\}`$, which are located at sites $`i`$ of a cubic lattice of size $`V=L^3`$. The Hamiltonian of the model is given by,
$$H=\beta EhM,$$
(2.1)
where $`E`$ and $`M`$ denote the energy and magnetization,
$$E=\underset{i,j}{}\delta (\sigma _i,\sigma _j),M=\underset{i}{}\delta (\sigma _i,\sigma _g).$$
(2.2)
Here the first sum runs over all nearest neighbour pairs of sites $`i`$ and $`j`$. A non-vanishing field $`h>0`$ favours magnetization in the direction of the ghost spin $`\sigma _g`$. On a finite lattice of size $`L^3`$ the partition function of the model is then given by
$$Z(\beta ,h,L)=\underset{\{\sigma _i\}}{}\mathrm{e}^H.$$
(2.3)
For vanishing external field the model is known to have a first order phase transition for $`\beta _c(h=0)=0.550565(10)`$ . In the presence of a non-vanishing external field $`h`$ this first order transition weakens and ends in a second order critical endpoint, which is expected to belong to the universality class of the $`3`$-d Ising model. A first estimate of the critical endpoint $`(\beta _c,h_c)`$ has been given in Ref. . We will give here a more precise determination of $`(\beta _c,h_c)`$ and analyze the universal critical behaviour at this point.
At $`(\beta _c,h_c)`$ the original operators for the energy and magnetization, $`E`$ and $`M`$, loose their meaning as operators being conjugate to the temperature-like and symmetry breaking couplings. One rather has to determine the new relevant directions at the critical endpoint, which take over the role of temperature-like and symmetry breaking directions and allow the determination of the two relevant critical exponents at the second order endpoint. This also fixes the new order parameter and energy-like observables as mixed operators in terms of the original variables $`E`$ and $`M`$. Following the discussion of the corresponding problem for the liquid-gas transition we introduce new operators
$$\stackrel{~}{M}=M+sE,\stackrel{~}{E}=E+rM,$$
(2.4)
as superpositions of the original variables $`E`$ and $`M`$. The Hamiltonian of the Potts model can then be rewritten in terms of these new operators,
$$H=\tau \stackrel{~}{E}\xi \stackrel{~}{M},$$
(2.5)
where the new couplings are given by
$$\xi =\frac{1}{1rs}(hr\beta ),\tau =\frac{1}{1rs}(\beta sh).$$
(2.6)
We note that the general ansatz given by Eqs. 2.4 to 2.6 does allow for the possibility that the new couplings $`\tau `$ and $`\xi `$ do not define orthogonal directions in the space of the original couplings $`\beta `$ and $`h`$, i.e. they need not result from a rotation of the couplings $`\beta `$ and $`h`$.
In the presence of a non-vanishing external field the line of first order phase transitions, $`\beta _c(h)`$, singles out a direction which corresponds to the low temperature, symmetry broken part of a temperature-driven transition. The new temperature-like direction $`\tau `$ thus is identified with $`\beta _c(h)`$. In the vicinity of the critical endpoint the slope of this line determines the mixing parameter $`r`$,
$$r^1=\left(\frac{\mathrm{d}\beta _c(h)}{\mathrm{d}h}\right)_{h=h_c}.$$
(2.7)
The second mixing parameter, $`s`$, is determined by demanding that the energy-like fluctuations and those of the ordering field are uncorrelated,
$$\delta \stackrel{~}{M}\delta \stackrel{~}{E}0,$$
(2.8)
with $`\delta \stackrel{~}{X}=\stackrel{~}{X}\stackrel{~}{X}`$ for $`X=M`$ and $`E`$. This insures that the expectation value of the new ordering field operator, $`\stackrel{~}{M}`$, fulfills a basic property of an order parameter, i.e. for $`\xi =\xi _c`$ it stays $`\tau `$-independent in the symmetric phase. Eqs. 2.7 and 2.8 give two independent conditions which are sufficient to determine the mixing parameters $`r`$ and $`s`$.
All our simulations have been performed on lattices of size $`L^3`$ with $`L=40,50,60`$ and $`70`$. We use periodic boundary conditions and perform simulations with a cluster algorithm described in Ref. . Independent of the lattice size we call a new configuration the spin configuration obtained after 1000 cluster updates. Typically we have for each value of the external field $`h`$ performed simulations at 3 to 4 $`\beta `$-values in the vicinity of the critical point. For each pair of couplings $`(\beta ,h)`$ we generated about 10000 configurations which then have been used in a Ferrenberg-Swendsen reweighting analysis to calculate observables at intermediate parameter values. Autocorrelation times have been estimated by monitoring the time evolution of the energy $`E`$. Close to the pseudo-critical points they vary between 7 and 25 configurations on the smallest and largest lattices, respectively. All errors quoted below have been obtained from a jackknife analysis.
## 3 Determination of the critical endpoint
The basic observables for the determination of the pseudo-critical couplings on finite lattices are susceptibilities constructed from the Hamiltonian $`H`$, and the magnetization $`M`$,
$`c_L`$ $`=`$ $`{\displaystyle \frac{1}{L^3}}\left(H^2H^2\right),`$ (3.1)
$`\chi _L`$ $`=`$ $`{\displaystyle \frac{1}{L^3}}\left(M^2M^2\right).`$ (3.2)
The location of the maxima in these observables define pseudo-critical couplings, $`\beta _{c,L}(h)`$, which may differ for different observables. We generally find that the locations of $`c_{L,max}`$ and $`\chi _{L,max}`$ differ by about one standard deviation of the statistical errors on $`\beta _{c,L}(h)`$ for our smaller lattices and agree within errors for the largest lattice, $`L=70`$. The maxima have been obtained from a combined Ferrenberg-Swendsen analysis which takes into account all data sets at fixed values of $`h`$ at $`\beta `$-values in the vicinity of the pseudo-critical points. In Tab. 1 we quote results for $`\beta _{c,L}(h)`$ obtained from $`\chi _{L,max}`$ for couplings, $`h`$, in the vicinity of the critical endpoint.
A first indication for the location of the critical region is obtained from an analysis of the volume dependence of the peak heights in the susceptibilities. In the region of first order phase transitions ($`h<h_c`$) $`\chi _{L,max}`$ is expected to increase proportional to the volume, while for $`h>h_c`$ the peak heights will approach a finite value in the infinite volume limit. Results for $`\chi _{L,max}`$ obtained on different size lattices are shown in Fig. 1. From this figure as well as from a similar analysis of $`c_{L,max}`$ it is clear that the critical endpoint will be located in the interval $`0.0005h0.001`$. In this interval we find that the dependence of the pseudo-critical couplings $`\beta _{c,L}(h)`$ on the external field $`h`$ is well approximated by a leading order Taylor-expansion,
$$\beta _{c,L}(h_1)\beta _{c,L}(h_2)=\frac{1}{r_L}(h_1h_2),h_1,h_2[0.0005,0.001].$$
(3.3)
For our largest lattice, $`L=70`$, this dependence is shown in Fig. 2. A straight line fit to these data yields, $`r_{70}=0.689(8)`$. Results from other lattice sizes are summarized in Tab. 2. The slope parameter $`r_L`$ is slightly volume dependent which, of course, reflects the volume dependence of the pseudo-critical couplings. We thus have extrapolated $`r_L`$ to the infinite volume limit using the ansatz $`r_L=r_{\mathrm{}}+c/L^3`$. This yields $`r_{\mathrm{}}=0.685(10)`$ and fixes the mixing parameter $`r`$ defined in Eq. 2.4.
The mixing parameter $`r_L`$ determined above may be used in connection with Eq. 2.8 to determine the second mixing parameter $`s_L`$. We have done so and within statistical errors we found $`s_L=r_L`$, i.e. the couplings $`(\tau ,\xi )`$ can be obtained from a rotation in the space of the original couplings $`(\beta ,h)`$. An alternative approach thus is to assume $`s_L=r_L`$ and use Eq. 2.8 to determine $`s_L`$. In this way we also can make use of the information on diagonal correlations $`(\delta X)^2`$, $`X=E,M`$ and determine $`s_L`$ from a diagonalization of the fluctuation matrix . The results obtained in this way for $`s_L`$ on different size lattice are also given in Tab. 2. We note that within errors this approach is consistent with the determination of $`r_L`$ from the slope of the critical line. The statistical errors are, however, significantly smaller. This shows that the transformation of variables, $`(\beta ,h)(\tau ,\xi )`$ defined in Eq. 2.6 indeed is a rotation. In the following we will use for the mixing parameters $`r=s`$ with $`s=0.690`$.
Having fixed the mixing parameters we can perform an analysis of the critical behaviour in the vicinity of the critical endpoint using standard techniques to study temperature driven second order phase transitions in the absence of an external symmetry breaking field. In particular, we can use Binder cumulants to locate the critical endpoint,
$$B_{3,L}=\frac{(\delta \stackrel{~}{M})^3}{(\delta \stackrel{~}{M})^2^{3/2}},B_{4,L}=\frac{(\delta \stackrel{~}{M})^4}{(\delta \stackrel{~}{M})^2^2}.$$
(3.4)
For given values of $`h`$ and $`L`$ we find that the cumulant $`B_{3,L}(\beta ,h)`$ vanishes and $`B_{4,L}(\beta ,h)`$ acquires a minimum at values of the coupling $`\beta `$ which agree with each other within statistical errors. This, in turn, defines a pseudo-critical coupling, which again agrees within errors with the pseudo-critical couplings given in Tab. 1. The minima of the second Binder cumulant, $`B_{4,L}`$, calculated on different size lattices should have a unique crossing point when plotted, for instance, versus the external field $`h`$. This is indeed the case, as can be seen in Fig. 3. The crossing point yields the critical field $`h_c`$ at the second order endpoint.
From the crossing points of the Binder cumulants we find for the critical endpoint
$$\beta _c=0.54938(2),h_c=0.000775(10).$$
(3.5)
or equivalently in terms of the rotated couplings,
$$(\tau _c,\xi _c)=(0.37182(2),0.25733(2)).$$
(3.6)
At this point the Binder cumulant takes on the value $`B_4=1.609(4)(10)`$, where the first error only takes into account the fluctuation of $`B_{4,L}`$ on the different size lattices at $`(\beta _c,h_c)`$ and the second error estimates the uncertainties arising from the errors on the location of the endpoint. Our result agrees within errors with the value found for the three dimensional Ising model, $`B_4=1.604(1)`$ .
## 4 Universality class of the endpoint
The value of the Binder cumulant $`B_4`$ determined at the critical endpoint strongly suggests that the critical endpoint belongs to the universality class of the 3-d Ising model. Rather impressive support for Ising-like behaviour at the critical endpoint, which at the same time also demonstrates the importance of the correct choice of energy- and ordering field like variables, is given by the contour plots obtained from the joined probability distributions (two-dimensional histograms) of the operators $`(E,M)`$ and $`(\stackrel{~}{E},\stackrel{~}{M})`$, respectively. These are shown in Fig. 4 for the $`70^3`$ lattice at $`(\beta _c,h_c)`$. Although the contour plot for $`(\stackrel{~}{E},\stackrel{~}{M})`$ is still slightly asymmetric, a comparison with the corresponding contour plot of the 3-d Ising model clearly shows the same characteristic fingerprint. We have checked that this asymmetry is well within the uncertainties of our determinations of $`(\beta _c,h_c)`$ as well as the mixing parameters $`r`$ and $`s`$.
Further support for the Ising universality class of the critical endpoint comes from a more conventional finite size scaling analysis performed for the newly defined mixed operators $`\stackrel{~}{M}`$, $`\stackrel{~}{E}`$ and the related susceptibilities. In particular we have considered the order parameter
$$\stackrel{~}{m}(\tau ,\xi )=\frac{1}{L^3}\left(\stackrel{~}{M}(\tau ,\xi )\stackrel{~}{M}(\tau _c,\xi _c)\right),$$
(4.1)
and the susceptibility
$$\stackrel{~}{\chi }_L=L^3\left(\stackrel{~}{m}^2|\stackrel{~}{m}|^2\right).$$
(4.2)
These observables indeed show the usual finite size behaviour in the vicinity of the critical point. For fixed $`\xi \xi _c`$ the susceptibility $`\stackrel{~}{\chi }_L`$ reaches a maximum at a pseudo-critical coupling $`\tau _{pc}`$. Critical exponents can then be determined from the scaling of the order parameter, $`|\stackrel{~}{m}|L^{\beta /\nu }`$, the peak height of the susceptibility, $`\stackrel{~}{\chi }_{L,max}L^{\gamma /\nu }`$, and the pseudo-critical couplings, $`(\tau _{pc}\tau _c)L^{1/\nu }`$. In our analysis of the critical behaviour we have first fixed the critical point $`(\tau _c,\xi _c)`$ and determined critical exponents from two-parameter fits. Errors have then be determined by also varying $`(\tau _c,\xi _c)`$ within the errors given in Eq. 3.6. In Tab. 3 we summarize the results of this analysis and compare the calculated exponents with known values for the 3-d Ising model. As can be seen the agreement is quite satisfactory although in our analysis the lattices have not yet been large enough to reach an accuracy similar to that obtained for the 3-d Ising model .
Most sensitive to the correct choice of the new, mixed observables are the new energy-like observable, $`\stackrel{~}{e}=\stackrel{~}{E}/L^3,`$ and the corresponding susceptibility, $`c_e(\stackrel{~}{e}/\tau )_\xi `$. Their critical behaviour is controlled by the thermal critical exponent ($`y_t`$) which fixes the critical exponent $`\alpha `$. In particular, the energy density calculated at $`(\tau _c,\xi _c)`$ is expected to scale like $`\stackrel{~}{e}=c_0+c_1L^{(1\alpha )/\nu }`$ at $`(\tau _c,\xi _c)`$. It is obvious from Eq. 2.4 that this scaling behaviour can hold only if in the mixed observable $`\stackrel{~}{e}`$ the leading singular behaviour of $`E/L^3`$ and $`M/L^3`$, which in general is proportional to $`L^{\beta /\nu }`$, gets canceled. For the 3-d Ising model the exponent $`\alpha `$ is small ($`\alpha =0.11`$). As a consequence the dominant singular behaviour in $`\stackrel{~}{e}`$ is expected to behave like $`L^{1.41}`$ rather than $`L^{0.52}`$ as would be the case for $`E/L^3`$ and $`M/L^3`$ separately. As can be seen from Tab. 3 this is supported by our analysis of the finite size scaling behaviour of the energy-like observable $`\stackrel{~}{e}`$, although in this case the critical exponent shows the largest deviation from the corresponding value of the 3-d Ising model.
## 5 Conclusions
We have determined the critical endpoint of the 3-d, 3-state Potts model in the presence of an external ordering field and have analyzed the critical behaviour in the vicinity of this second order phase transition. New energy-like and ordering field like couplings and observables have been obtained through a rotation of the original couplings and fields. At the critical endpoint of the line of first order phase transitions the joined probability distribution of these new observables shows a correlation pattern which is characteristic for the 3-d Ising model. A finite size scaling analysis performed with the newly defined rotated observables further supports that the critical endpoint belongs to the universality class of the 3-d Ising model.
Acknowledgements:
We thank Jürgen Engels and Andreas Peikert for helpful discussions. This work has been supported by the TMR network ERBFMRX-CT-970122 and the DFG under grant Ka 1198/4-1.
|
warning/0007/astro-ph0007045.html
|
ar5iv
|
text
|
# Detecting Dark Matter in High Velocity Clouds
## 1 INTRODUCTION
Observations of the Galactic halo make a compelling case that the formation of halos continues to the present day (Wyse 1999). The halo appears to have built up through a process of accretion and merging of low-mass structures which is still going on at a low level. Hierarchical cold dark matter (CDM) simulations, however, predict that the Galactic halo should have many more satellites than are actually observed (Klypin et al. 1999; Moore et al. 1999). Observations reveal that much of the sky is peppered with high-velocity H$`\mathrm{I}`$ clouds (HVCs) which do not conform to orderly Galactic rotation. These are interesting accretion candidates $``$ particularly if they are associated with dark matter ‘mini halos’ $``$ except that their distances, $`d`$, are unknown for all but a few sources. As a result, fundamental physical quantities $``$ size ($`d`$) and mass ($`d^2`$) $``$ are unconstrained which has encouraged wide speculation as to the nature of HVCs (Wakker & van Woerden 1997; 1999).
The current renaissance in HVC studies can be traced in part to one paper. Blitz et al. (1999) have shown that the velocity centroids and groupings of positive/negative velocity clouds on the sky may be understood within a reference frame centered on the Local Group barycenter. They interpret HVCs as gas clouds accreting onto the Local Group over a megaparsec sphere. Braun & Burton (2000) have identified specific examples of compact clouds that have ‘rotation curves’ consistent with CDM mass profiles. For sources at 700 kpc, the kinematic signatures imply a high dark-to-visible mass ratio of 10$``$50.
However, Zwaan & Briggs (2000) note from the Arecibo H$`\mathrm{I}`$ Strip Survey that of 300 galaxies and 14 galaxy groups, none appear to have properties resembling HVCs associated with the Milky Way or the Local Group. A possible interpretation is that the clouds are somewhat closer to the galaxy or group barycenters.
Reliable distance indicators for HVCs are presently hard to come by. Oort (1966) proposed virial distances on the assumption that HVCs are self-gravitating. The mass inferred from the H$`\mathrm{I}`$ column density has a different distance dependence to the gravitationally inferred mass from the H$`\mathrm{I}`$ line width. This crude method puts some HVCs at megaparsec distances (Blitz et al. 1999; Braun & Burton 1999). An alternative method is to use bright sources with well-calibrated distances along the sight line to gas clouds. If the cloud produces absorption in the spectrum of one source but not in the other, the cloud distance can be bracketed. This method establishes that two HVCs are within 10 kpc of the Sun (van Woerden et al. 1999).
We have recently developed the H$`\alpha `$ distance method which has the potential to reach much greater distances (Bland-Hawthorn et al. 1998; Bland-Hawthorn & Maloney 1999a,b). If a gas cloud is optically thick to ionizing radiation from a source with known luminosity, the H$`\alpha `$ flux can be used to infer the external field strength and therefore the cloud’s distance from the source. Two groups have now used this method to show that many HVCs are faint or invisible in H$`\alpha `$ and appear to be distributed on scales of tens of kiloparsecs throughout the Galactic halo. Some of these clouds have H$`\mathrm{I}`$ masses in excess of 10<sup>6</sup>M (Weiner et al. in preparation; Putman et al. in preparation).
In establishing whether the HVCs are truly candidates for merging ‘mini halos’, their dark matter content needs to be determined. By considering their potential gravitational lensing properties, we now describe a test to establish whether compact HVCs do indeed contain such dark matter halos.
## 2 GRAVITATIONAL LENSING
### 2.1 MASS PROFILE
Utilizing an extensive sample of numerical simulations, Navarro, Frenk and White (1997) (NFW) recently suggested that the density profile of dark matter halos formed within the framework of cold dark matter cosmology can be simply described as
$$\frac{\rho (r)}{\rho _{crit}}=\frac{\delta _s}{(r/r_s)(1+r/r_s)^2}$$
(1)
where $`r_s`$ is a scale radius which is related to the $`r_{200}`$ (the radius of the halo within which the average is $`200\times `$ the critical density of the Universe, $`\rho _{crit}`$), by a ‘concentration’ $`c_s`$, such that $`r_{200}=c_sr_s`$. The (dimensionless) characteristic density, $`\delta _s`$, is simply related to $`c_s`$ by
$$\delta _s=\left(\frac{200}{3}\right)\frac{c_s^3}{\left[\mathrm{log}(1+c_s)c_s/\left(1+c_s\right)\right]}.$$
(2)
We adopt this profile in describing the dark matter distribution in HVCs. For a particular halo mass, defined as $`M_{200}=M(<r_{200})`$, the concentration $`c_s`$ depends on the halo collapse redshift, and hence the cosmology, power spectrum etc. Assuming $`\mathrm{\Omega }_o=1`$ & $`\mathrm{\Lambda }_o=0`$, we employ the recipe provided by NFW to determine the halo profiles in a standard CDM cosmology. A numerical routine for calculating these properties was kindly provided by Prof. Navarro.
Bartelmann (1996) and Wright & Brainerd (2000) have considered the gravitational lensing properties of massive galaxy clusters with NFW mass profiles. Rather than the radial density distribution (Equation 1), the important quantity in such an analysis is the projected surface mass density. This is given by
$$\kappa (x)=2\kappa _s\frac{f(x)}{x^21}$$
(3)
where $`x=r/r_s`$. The normalizing factor is given by $`\kappa _s=\delta _sr_s\rho _{crit}/\mathrm{\Sigma }_{crit}`$, where $`\mathrm{\Sigma }_{crit}`$ is the critical surface mass density to gravitational lensing;
$$\mathrm{\Sigma }_{crit}=\frac{c^2}{4\pi G}\frac{D_{os}}{D_{ol}D_{ls}}$$
(4)
where $`D_{ij}`$ are the angular diameter distances between the observer $`(o)`$, lens $`(l)`$ and source $`(s)`$. The auxiliary function $`f(x)`$ is given by
$$f(x)=\{\begin{array}{cc}1\frac{2}{\sqrt{1x^2}}arctanh\sqrt{\frac{1x}{1+x}}\hfill & (x<1)\hfill \\ 0\hfill & (x=1)\hfill \\ 1\frac{2}{\sqrt{x^21}}arctan\sqrt{\frac{x1}{x+1}}\hfill & (x>1)\hfill \end{array}$$
(5)
Defining $`g(x)=f(x)/(x^21)`$, the normalised surface mass density is given by;
$$\kappa (x)=3.33\times 10^{13}\frac{\delta _s}{c_s}r_{200}D_{ol}\left(1\frac{D_{ol}}{D_{os}}\right)g(x)h^2$$
(6)
where all distances are in kpc, and it is assumed that only objects in the local Universe are considered, such that $`D_{ls}=D_{os}D_{ol}`$. The Hubble parameter, $`h`$, is defined such that $`H_o=100hkm/s/Mpc`$.
### 2.2 LENSING PROPERTIES
Armed with these various tools, we can now examine the gravitational lensing properties of dark matter halos of HVCs. Considering fiducial masses of $`M_{200}=10^6,10^7,10^8`$ & $`10^9M_{}`$, corresponding to the potential dark matter masses of HVCs, we determined the NFW parameters; these are summarized in Table 1. Firstly, it is important to determine whether the projected surface density of these dark matter halos is sufficient to induce macrolensing effects, resulting in multiple images; in the absence of strong shearing, this requires the normalized surface mass density, $`\kappa `$ to exceed unity. An examination of Equations 3 and 5, reveals that $`\kappa (x)`$ becomes singular at $`x=0`$ and must meet the criterion for producing multiple images at some radius. At small $`x`$, the surface density becomes
$$\kappa (x)2\kappa _s\mathrm{log}\left(\frac{2}{e}\frac{1}{x}\right).$$
(7)
Considering the parameters in Table 1, and placing the HVC at any reasonable distance, $`\kappa (x)`$ does not exceed unity until very small (and unphysical) radii. Therefore, such halos are unable to produce observable macrolensing effects and the dark matter cannot be probed by looking for ‘image splitting’ of distant sources. However, if the dark matter in HVC is in the form of compact objects, these will introduce microlensing variability into the observations of background sources. As potential source are extragalactic, namely galaxies, the resulting microlensing is unlike that in the Magellanic Clouds and the Galactic Bulge, where an individual star is seen to brighten and fade as a compact object crosses the line-of-sight. Rather, as a patch of light from a distant galaxy represents an unresolved population of stars, any microlensing will be seen as against this smooth background. Hence, microlensing will be detected as fluctuations in surface brightness over the source galaxy. The framework of this ‘pixel-lensing’ was laid down by Crotts (1992) and Gould (1996), and has recently proved successful in a search for compact objects along the lines of sight to M31 and the Galactic Bulge (Crotts & Tomaney 1996; Tomaney & Crotts 1996; Alcock et al. 1999). Pixel lensing has also been proposed as a tool to search for both intracluster compact objects (Gould 1995; Lewis, Ibata & Wyithe 2000) and cosmologically distributed dark matter (Lewis & Ibata 2000).
Using Equation 6, and considering the NFW parameters presented in Table 1, the microlensing optical depths for the dark matter halos can be calculated. By assuming that the halo lies at a distance of 100kpc from the Earth, in front of a source at 3Mpc, Table 2 presents the optical depths at several angular radii. It is immediately apparent that the optical depths for all the models are small, in the regime probed by the microlensing searches towards the Galactic Bulge and halo. Given the subcritical value of the microlensing optical depth seen through the dark matter halos associated with the HVCs, the expected number of ‘pixel-lensing’ events simply scales as $`\mathrm{\Gamma }\kappa `$ (see Binney 2000). Hence, in searching for HVC compact dark matter objects objects, a simple test presents itself; in monitoring the surface brightness distributions of galaxies, regions that overlap with the dark matter halos of HVCs will display pixel-lensing variability. Moreover, as the optical depth increases towards the center of the HVC systems, the resultant number of microlensing events should also increase towards the center. Given the simple linear scaling between the number of events and the surface mass density, the identification of microlensing of systems viewed through HVCs will not only detect the dark matter component, but will also provide a (non-kinematic) map of the dark matter mass over tens to hundreds of arcseconds, depending upon the mass of the halo. A detailed calculation of the optical depth distribution for specific HVCs and source galaxies is beyond the scope of this current paper and will be presented elsewhere. A simple estimate of the expected number of events can be found by extrapolating the analysis of Binney (2000), who determines that for an optical depth of $`\kappa =10^6`$ and sources at 50 Mpc, monitoring with a 4-m class (diffraction limited) telescope will uncover three or four events per $`10^6`$ resolution elements per week. For a similar source distance, and a HVC located at 100 kpc, the optical depth in the central regions of the halos presented in this paper are greater than $`10^6`$, with the $`10^9M_{}`$ exceeding this by a factor of $`50`$ over a region 10<sup>′′</sup> in radius. Increasing the distance to the HVC will similarly increase the number of expected microlensing events. This analysis indicates that if HVCs are enshrouded in halos of compact dark matter, then a substantial number of microlensing events should be detectable.
One potential contaminant is microlensing of the sources by compact objects within our own Galactic halo, or that of the potential source. Binney (2000) recently examined the sky distribution of halo microlensing optical depth to distant sources within several models for the mass distribution of the Galaxy. The optical depth is greatest in the disk of the Galaxy, and falls rapidly with Galactic latitude, falling below $`10^6`$ for $`|b|\stackrel{>}{}12^o`$. At these higher galactic latitudes even the least massive halo considered in this paper will dominate the microlensing optical depth along a line-of-sight by a factor of ten. Similarly, the optical depth through the halo of the source will also be of order $`10^6`$, and show a similar distribution over the galaxy. Microlensing by material in the HVC halo will enhance this value, and will be spatially correlated with the HVC core, making in discernible from any intrinsic ‘self-lensing’.
Recent results from the MACHO (Alcock et al. 2000) and EROS (Lasserre et al. 2000) studies towards the Magellanic Clouds suggest that only $`20\%`$ (at most) of the dark matter halo of the Galaxy resides in the form of compact objects. If this is the case, and the distribution of dark matter is universal, then the optical depths presented for the HVCs would have to be scaled by a similar factor. However, if the dark matter in the Galactic halo is clumped on large scales then our view to the Magellanic Clouds may be through a relatively empty region. Such a picture, which is consistent with the hierarchical accretion model discussed in this paper (see Klypin et al. 1999), may explain why our view towards the Galactic Bulge appears to be relatively overdense in MACHOs (Binney, Bissantz & Gerhard 2000; Alcock et al. 2000a).
## 3 PROPOSED EXPERIMENT
The practical limitations to the pixel lensing technique described in Section 2 are, of course, set primarily by the availability, or lack thereof, of (purported) dark matter-dominated HVCs suitably aligned with background galaxies. To a 5 $`\sigma `$ limiting H I column density of 7$`\times `$10<sup>17</sup> cm<sup>-2</sup>, Murphy et al. (1995) claim an HVC sky covering fraction of $``$37%. The high detection rate of high-velocity Galactic Mg II gas seen toward background quasars (Savage et al. 1993) implies a covering fraction of $``$50% at column densities of 2$`\times `$10<sup>17</sup> cm<sup>-2</sup>. Clearly, high-velocity gas does exist, at some level, along virtually all extragalactic sightlines. Unfortunately, one needs to exercise restraint before proclaiming that the pixel lensing experiment will therefore be a trivial one.
Of greatest concern is the potential contamination due to the inclusion of ‘‘non-dark matter-dominated HVCs’’ in the above sky covering fractions. Eliminating large, diffuse, HVCs from the above sample (e.g. Magellanic Stream, Complexes A, C, and M), and restricting the analysis to unresolved ($`<`$1 deg<sup>2</sup>), isolated, HVCs, immediately reduces the covering fraction by a factor of $``$50 -- to $``$0.7% -- Blitz & Robishaw (2000). An even more stringent sampling was adopted by Braun & Burton (1999) in constructing their Compact High-Velocity Cloud (CHVC) catalog, resulting in a compilation of only 65 candidates. Subsequent high-resolution imaging of a subset of these CHVCs (Braun & Burton 2000) shows that each is $``$0.2 deg<sup>2</sup> in areal extent, for a total sky covering fraction of $``$0.03% -- more than three orders of magnitude lower than that found by Murphy et al. (1995). The southern sample of CHVCs in the Braun & Burton catalog was necessarily limited to the older H I survey by Bajaja et al. (1985); this has since been supplanted by the Morras et al. (2000) survey (a direct southern analog to the Leiden-Dwingeloo Survey \- Hartmann & Burton 1997) and the H I Parkes All Sky Survey (HIPASS -- Putman & Gibson 1999a,b). This latter survey is particularly attuned to the discovery of CHVCs, as it is the first of its kind to sub-Nyquist sample the southern sky. A visual inspection of several random HIPASS data cubes demonstrates that a factor of two increase in the number of known southern CHVCs (with H I column densities $`>`$10<sup>18</sup> cm<sup>-2</sup>) can be expected.
As Blitz & Robishaw (2000) demonstrate, the probability of a chance alignment of a CHVC with a nearby, background galaxy is $`1\%`$. Even allowing for the aforementioned expected increase in the number of catalogued CHVCs, this probability will remain $`<`$2%. On the other hand, if the simulations of Klypin et al. (1999) and Moore et al. (1999) are correct, one might expect there to be 300$``$500 CHVCs in the Local Group, of which only $``$100 have been accounted for. Perhaps deeper H I surveys will uncover this missing population, or perhaps they will only be discovered serendipitously through studies of background quasars. Regardless, the inclusion of this additional hidden population of CHVCs, would increase the probability of finding a CHVC-background galaxy alignment to 3$``$5%.
Obviously, this will be a challenging experiment, but not an impossible one. We are initiating a search through the HIPASS data cubes, in an attempt to uncover prospective candidates. Several possibilities currently exist, although we stress this is neither a finalized nor complete list – HIPASS 1328$``$30 (Banks et al. 1999) and ESO 383-G087, both in the Cen A Group, NGC 3109, in the Antlia-Sextans Grouping, and NGC55 & AM0106-382 in Sculptor, each have CHVCs lying within 15-30, yet kinematically separated, from the galaxy.
## 4 CONCLUSIONS
Cold dark matter models for the formation of universal structure predict that the Galaxy should be be surrounded by many infalling ‘clumps’ of material. While the Galaxy is accompanied by a number of dwarf galaxies, they cannot account for the total expected population of objects. Recently, it has been suggested that high-velocity clouds are accompanied by halos of dark matter and hence represent, and trace, the missing dark matter population.
In this paper we have demonstrated that while HVCs are too diffuse to produce macrolensing splitting of distant sources, if their dark matter consists of compact object, then this adds to the microlensing optical depth. This signal is detectable with a 4m class telescope and a modest observing campaign over a year, and provide a map of the underlying dark matter distribution. While chance alignments of HVCs with more distant galaxies are not common, several potential sources already present themselves. This situation is likely to improve as current data is scanned and future observations are undertaken. We eagerly await the monitoring of a ‘mini halo’ candidate HVC along the sight line to a nearby galaxy.
|
warning/0007/gr-qc0007032.html
|
ar5iv
|
text
|
# Spin-1/2 Particles in Non-Inertial Reference Frames: Low- and High-Energy Approximations
## 1 Introduction
The behaviour of quantum systems in inertial and gravitational fields is of interest in investigations regarding the structure of spacetime at the quantum level . Quantum objects are in fact finer and more appropriate probes of structures that appear classically as results of limiting procedures. Though a definitive answer to questions regarding the fundamental structure of spacetime may only come from a successful quantum theory of gravity, the extrapolation of general relativity from planetary lengths, over which it is well established, to Planck’s length requires a leap of faith in its validity of over forty orders of magnitude and the resolution of difficult quantization problems. The alternative, performing experiments at Planck’s length, appears remote indeed. A more modest, but realistic approach consists in verifying the theory at intermediate lengths. This may be accomplished, to some extent, by considering the interaction of classical inertial and gravitational fields with quantum objects. A vast array of effects can be predicted in this instance and a unified treatment is afforded by Einstein’s theory. General relativity incorporates the equivalence principle from the outset and observations, where feasible, do confirm that inertia and gravity interact with quantum systems in ways that are compatible with Einstein’s views. This is borne out of measurements on superconducting electrons and on neutrons which are certainly not tests of general relativity per se, but offer tangible evidence that the effect of inertia and Newtonian gravity on wave functions down to lengths of $`10^3`$ and $`10^{13}`$ cm respectively is that predicted by wave equations compliant with general relativity .
Inertial effects must be identified with great accuracy. This is dictated by their unavoidable presence in Earth-bound and near-space experiments of ever increasing accuracy aimed at testing fundamental theories. They also provide a guide in the study of relativity because, in all instances where non-locality is not an issue, the equivalence principle, in some of its forms , ensures the existence of a gravitational effect for each inertial effect.
Among the quantum mechanical probes, spin-1/2 particles play a prominent role and in reality some of the most precise experiments in physics involve Dirac particles. They are very versatile tools that can be used in a variety of experimental situations and energy ranges while still retaining essentially a non-classical behaviour. Within the context of general relativity comprehensive studies of the Dirac equation were conducted by De Oliveira and Tiomno and Peres . More recently, spin-inertia and spin-gravity interactions have been shown to have non-trivial physical and astrophysical consequences. This is the case for neutrinos whose inertia/gravity interactions are just starting to be studied . Superconducting and neutron interferometers of large dimensions hold great promise in many of these investigations. They can provide accurate measurements of quantum phases, whose role is important in gyroscopy, and possibly in testing general relativity . It is anticipated that similar studies will be performed with particle accelerators . The forerunner of this second group of investigations is the work of Bell and Leinaas in which evidence was found for the coupling of spin to rotation.
For most studies involving non-relativistic Dirac particles, the inertial/gravitational phase is of paramount importance. For other problems, such as the interaction of inertia/gravitation at the atomic level , the knowledge of the Hamiltonian is of greater importance. The derivation of the Hamiltonian is usually accomplished by following a sequence of Foldy-Wouthuysen (FW) transformations . It has been recently presented in comprehensive form by Hehl and Ni purely within the framework of special relativity and in the local frame of the fermion. In the present work, the non-relativistic case is tackled by means of a procedure that renders the quantum phase manifest and can, therefore, be applied to both types of problems.
For tests involving accelerators, a high-energy approximation corresponding to the FW-transformation was given long ago by Cini and Touschek (CT) for free Dirac particles. Their work is extended here to include external electromagnetic and gravitational fields and quantum phases to first order. The derivation of the Hamiltonian can then be accomplished in a standard way.
The method and derivation of the Hamiltonian are given in Section 2. The low-energy approximation in Section 3 uses the FW-transformation and follows the standard textbook approach . In contrast, the high-energy approximation in Section 4 uses the CT-transformation and involves a non-standard procedure.
## 2 Derivation of the Dirac Hamiltonian
We use the formalism of general relativity, that encompasses both inertial and true gravitational fields, to derive the Hamiltonian of a fermion in a non-inertial frame.
The starting point is represented by the covariant Dirac equation
$`\left(i\gamma ^\mu (x)D_\mu {\displaystyle \frac{mc}{\mathrm{}}}\right)\psi (x)`$ $`=`$ $`0,`$ (2.1)
where the generalized matrices $`\gamma ^\mu (x)`$ satisfy the relation $`\{\gamma ^\mu (x),\gamma ^\nu (x)\}=2g^{\mu \nu }(x)`$ and are related to the usual Dirac matrices $`\gamma ^{\widehat{\alpha }}`$ by means of the vierbeins $`e^\mu {}_{\widehat{\alpha }}{}^{}(x)`$. Caratted indices refer to the observer’s local inertial frame. In (2.1), $`D_\mu _\mu +i\mathrm{\Gamma }_\mu ,\psi (x)`$ is the wavefunction defined for a general co-ordinate frame, $`_\mu `$ represents the usual covariant derivative, $`\mathrm{\Gamma }_\mu `$ is the spinor connection which follows from $`D_\mu \gamma _\nu (x)=0`$ and is given by
$`\mathrm{\Gamma }_\mu ={\displaystyle \frac{i}{4}}\gamma ^\nu (_\mu \gamma _\nu )={\displaystyle \frac{1}{4}}\sigma ^{\widehat{\alpha }\widehat{\beta }}e^\nu {}_{\widehat{\alpha }}{}^{}(_\mu e_{\nu \widehat{\beta }}),`$ (2.2)
where $`\sigma ^{\widehat{\alpha }\widehat{\beta }}=\frac{i}{2}[\gamma ^{\widehat{\alpha }},\gamma ^{\widehat{\beta }}].`$ Obviously, $`_\mu \psi (x)=_\mu \psi (x),`$ where $`_\mu `$ indicates partial differentiation.
It is possible to define a local co-ordinate frame according to an orthonormal tetrad with three-acceleration $`𝒂`$ along a particle’s world-line and three-rotation $`𝝎`$ of the spatial triad, subject to Fermi-Walker transport. This tetrad $`𝒆_{\widehat{\mu }}`$, is related to the general co-ordinate tetrad $`𝒆_\mu `$ by
$`𝒆_{\widehat{0}}`$ $`=`$ $`\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)^1\left[𝒆_0{\displaystyle \frac{1}{c}}(𝝎\times 𝒙)^k𝒆_k\right],`$ (2.3)
$`𝒆_{\widehat{ı}}`$ $`=`$ $`𝒆_i.`$ (2.4)
The corresponding vierbeins relating the two frames are then
$`e^0_{\widehat{0}}`$ $`=`$ $`\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)^1,`$ (2.5)
$`e^k_{\widehat{0}}`$ $`=`$ $`{\displaystyle \frac{1}{c}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)^1ϵ^{ijk}\omega _ix_j,`$ (2.6)
$`e^0_{\widehat{ı}}`$ $`=`$ $`0,`$ (2.7)
$`e^k_{\widehat{ı}}`$ $`=`$ $`\delta ^k{}_{i}{}^{}.`$ (2.8)
Similarly, by inverting (2.3) and (2.4), we find the inverse vierbeins
$`e^{\widehat{0}}_0`$ $`=`$ $`\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)`$ (2.9)
$`e^{\widehat{k}}_0`$ $`=`$ $`{\displaystyle \frac{1}{c}}ϵ^{ijk}\omega _ix_j,`$ (2.10)
$`e^{\widehat{0}}_i`$ $`=`$ $`0,`$ (2.11)
$`e^{\widehat{k}}_i`$ $`=`$ $`\delta ^k{}_{i}{}^{}.`$ (2.12)
The vierbeins satisfy the orthonormality conditions
$`\delta ^{\widehat{\alpha }}_{\widehat{\mu }}`$ $`=`$ $`e^\nu {}_{\widehat{\mu }}{}^{}e_{}^{\widehat{\alpha }}{}_{\nu }{}^{},`$ (2.13)
$`\delta ^\alpha _\mu `$ $`=`$ $`e^{\widehat{\nu }}{}_{\mu }{}^{}e_{}^{\alpha }{}_{\widehat{\nu }}{}^{}.`$ (2.14)
It follows that the metric tensor components are
$`g_{00}`$ $`=`$ $`\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)^2+{\displaystyle \frac{2}{c^2}}\left[\left(𝝎𝝎\right)\left(𝒙𝒙\right)\left(𝝎𝒙\right)^2\right],`$ (2.15)
$`g_{0j}`$ $`=`$ $`{\displaystyle \frac{1}{c}}\left(𝝎\times 𝒙\right)_j,`$ (2.16)
$`g_{jk}`$ $`=`$ $`\eta _{jk},`$ (2.17)
where $`\eta _{\mu \nu }`$ represents the Minkowski metric of signature $`2`$. Equation (2.1) has an exact solution to first order in the weak-field approximation defined by $`g_{\mu \nu }(x)=\eta _{\mu \nu }+\gamma _{\mu \nu }(x),`$ where the metric deviation $`\gamma _{\mu \nu }`$ is a small quantity of first order.
In fact, a new spinor $`\stackrel{~}{\psi }(x)`$ defined by
$`\stackrel{~}{\psi }(x)e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}\psi (x),`$ (2.18)
where
$`\mathrm{\Phi }_\mathrm{S}`$ $``$ $`\mathrm{}𝒫{\displaystyle _X^x}𝑑z^\lambda \mathrm{\Gamma }_\lambda (z),`$ (2.19)
satisfies the equation
$`\left(i\stackrel{~}{\gamma }^\mu (x)_\mu {\displaystyle \frac{mc}{\mathrm{}}}\right)\stackrel{~}{\psi }(x)=0,`$ (2.20)
where $`𝒫`$ refers to path ordering and
$`\stackrel{~}{\gamma }^\mu (x)e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}\gamma ^\mu (x)e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}.`$ (2.21)
By multiplying (2.20) on the left by $`(i\stackrel{~}{\gamma }^\nu (x)_\nu mc/\mathrm{})`$, we obtain the equation
$`\left(g^{\mu \nu }_\mu _\nu +{\displaystyle \frac{m^2c^2}{\mathrm{}^2}}\right)\stackrel{~}{\psi }(x)=0.`$ (2.22)
This last equation can be solved, in the weak-field approximation, for every component of $`\stackrel{~}{\psi }(x)`$ . Here, $`\stackrel{~}{\psi }(x)`$ is a solution of a second-order equation and does, of course, contain redundant solutions. These can be eliminated by writing the appropriate solution as
$`\stackrel{~}{\psi }(x)=\left(i\stackrel{~}{\gamma }^\mu (x)_\mu {\displaystyle \frac{mc}{\mathrm{}}}\right)e^{i\mathrm{\Phi }_\mathrm{G}/\mathrm{}}\mathrm{\Psi }_0,`$ (2.23)
where
$`\mathrm{\Phi }_\mathrm{G}`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle _X^x}𝑑z^\lambda \gamma _{\alpha \lambda }(z)P^\alpha `$ (2.24)
$`{\displaystyle \frac{1}{4}}{\displaystyle _X^x}𝑑z^\lambda (\gamma _{\alpha \lambda ,\beta }(z)\gamma _{\beta \lambda ,\alpha }(z))L^{\alpha \beta }(z),`$
and $`\mathrm{\Psi }_0`$ satisfies the Klein-Gordon equation for a free particle in Minkowski space.
For (2.24),
$`[L^{\alpha \beta }(z),\mathrm{\Psi }_0]`$ $``$ $`\left\{(x^\alpha z^\alpha )P^\beta (x^\beta z^\beta )P^\alpha \right\}\mathrm{\Psi }_0,`$ (2.25)
$`[P^\alpha ,\mathrm{\Psi }_0]`$ $``$ $`\{i\mathrm{}^\alpha \}\mathrm{\Psi }_0,`$ (2.26)
where $`P^\mu `$ is the momentum operator of the free particle. In (2.23), $`\mathrm{\Phi }_\mathrm{G}`$ plays the role of Berry’s phase because spacetime co-ordinates are just parameters and spacetime becomes simply Berry’s parameter space . The solution of (2.1) is, therefore, $`\psi (x)=\mathrm{exp}\left(i\mathrm{\Phi }_\mathrm{S}/\mathrm{}\right)\stackrel{~}{\psi }(x)`$ and is exact to first-order in the metric deviation.
It is interesting to notice that, by multiplying (2.1) on the left with $`(i\gamma ^\nu (x)D_\nu mc/\mathrm{}),`$ we obtain the second-order equation
$`\left(\gamma ^\mu (x)\gamma ^\nu (x)D_\mu D_\nu +{\displaystyle \frac{m^2c^2}{\mathrm{}^2}}\right)\psi (x)=0,`$ (2.27)
which, on using the relations $`[D_\mu ,D_\nu ]=\frac{i}{4}\sigma ^{\alpha \beta }R_{\alpha \beta \mu \nu }`$ and $`\sigma ^{\mu \nu }\sigma ^{\alpha \beta }R_{\mu \nu \alpha \beta }=2R`$ , reduces to
$`\left(g^{\mu \nu }D_\mu D_\nu {\displaystyle \frac{R}{4}}+{\displaystyle \frac{m^2c^2}{\mathrm{}^2}}\right)\psi (x)=0.`$ (2.28)
This equation does not contain any spin-curvature coupling for pure gravitational fields ($`R`$ = 0) and the gyro-gravitational g-factor is, therefore, zero . When $`R=\frac{8\pi G}{c^4}T,`$ the g-factor (coefficient of $`R`$) is $`\frac{1}{4}`$ . An equation that also yields a value for the orbital g-factor can be obtained from (2.1) by means of the transformation
$`\psi ^{}(x)e^{i\mathrm{\Phi }_\mathrm{G}/\mathrm{}}\psi (x).`$ (2.29)
It is easy to show that $`\psi ^{}(x)`$ satisfies the equation
$`\left[i\gamma ^\mu (x)\left(D_\mu +{\displaystyle \frac{i}{\mathrm{}}}(_\mu \mathrm{\Phi }_\mathrm{G})\right){\displaystyle \frac{mc}{\mathrm{}}}\right]\psi ^{}(x)`$ $`=`$ $`0.`$ (2.30)
Equation (2.30) can be immediately extended to include the electromagnetic fields by adding to $`\mathrm{\Phi }_\mathrm{G}`$ the phase
$`\mathrm{\Phi }_{\mathrm{EM}}`$ $``$ $`{\displaystyle \frac{e}{c}}{\displaystyle _X^x}𝑑z^\lambda A_\lambda (z),`$ (2.31)
and can be written in the form
$`\left[i\gamma ^\mu (x)\left(_\mu +{\displaystyle \frac{i}{\mathrm{}}}(_\mu \mathrm{\Theta })\right){\displaystyle \frac{mc}{\mathrm{}}}\right]\psi ^{}(x)`$ $`=`$ $`0,`$ (2.32)
where $`\mathrm{\Theta }\mathrm{\Phi }_\mathrm{G}+\mathrm{\Phi }_{\mathrm{EM}}+\mathrm{\Phi }_\mathrm{S}.`$
By operating from the left with $`\left[i\gamma ^\nu (x)\left(D_\nu +\frac{i}{\mathrm{}}(_\nu \mathrm{\Phi }_\mathrm{G})\right)mc/\mathrm{}\right]`$ on both sides of (2.30), we obtain to first-order in the metric deviation
$`\left[g^{\mu \nu }D_\mu D_\nu {\displaystyle \frac{i}{2}}\sigma ^{\mu \nu }\left([D_\mu ,D_\nu ]+iG_{\mu \nu }\right)+{\displaystyle \frac{m^2c^2}{\mathrm{}^2}}\right]\psi ^{}(x)=0,`$ (2.33)
where $`\frac{1}{\mathrm{}}(_\mu \mathrm{\Phi }_\mathrm{G})\kappa _\mu `$ and $`G_{\mu \nu }\kappa _{\nu ;\mu }\kappa _{\mu ;\nu }.`$ On the other hand, $`\frac{1}{2}G_{\mu \nu }=\frac{1}{4}R_{\mu \nu \alpha \beta }L^{\alpha \beta }`$ and the resulting orbital g-factor is also $`\frac{1}{4}.`$ This confirms that the gyro-gravitational ratio of a spin-1/2 particle is 1 as shown in .
Equation (2.32) can now be used to derive a Hamiltonian in general co-ordinates, taking the form
$`i\mathrm{}c_0\psi ^{}(x)`$ $`=`$ $`(g^{00}(x))^1[c\gamma ^0(x)\gamma ^j(x)(i\mathrm{}_j)+mc^2\gamma ^0(x)`$ (2.34)
$`+c\gamma ^0(x)\gamma ^\mu (x)(_\mu \mathrm{\Theta })]\psi ^{}(x)=H\psi ^{}(x),`$
where
$`g^{00}(x)`$ $`=`$ $`(e^0{}_{\widehat{0}}{}^{})^2\eta ^{\widehat{0}\widehat{0}}=(1+{\displaystyle \frac{𝒂𝒙}{c^2}})^2,`$ (2.35)
$`\gamma ^0(x)`$ $`=`$ $`e^0{}_{\widehat{0}}{}^{}\gamma _{}^{\widehat{0}}=\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)^1\beta ,`$ (2.36)
$`\gamma ^0(x)\gamma ^j(x)`$ $`=`$ $`e^0{}_{\widehat{0}}{}^{}(\gamma ^{\widehat{0}}\gamma ^{\widehat{ȷ}}+e^j{}_{\widehat{0}}{}^{})`$ (2.37)
$`=`$ $`\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)^2\left[\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)\alpha ^{\widehat{ȷ}}{\displaystyle \frac{1}{c}}ϵ^{jkl}\omega _kx_l\right].`$
Since $`\mathrm{\Phi }_\mathrm{G}`$ is correct only to first-order, this is also a constraint on the validity of (2.34). In what follows, terms of higher order in the metric deviation will be dropped. Explicit evaluation of $`_\mu \mathrm{\Theta }`$ shows that
$`(_\mu \mathrm{\Theta })`$ $`=`$ $`_\mu (\mathrm{\Phi }_{\mathrm{EM}}+\mathrm{\Phi }_\mathrm{S}+\mathrm{\Phi }_\mathrm{G})={\displaystyle \frac{e}{c}}A_\mu +\mathrm{}\mathrm{\Gamma }_\mu +(_\mu \mathrm{\Phi }_\mathrm{G}),`$ (2.38)
where
$`\mathrm{\Gamma }_0`$ $`=`$ $`{\displaystyle \frac{i}{2c^2}}(𝒂𝜶){\displaystyle \frac{1}{2c}}𝝎𝝈,`$ (2.39)
$`\mathrm{\Gamma }_j`$ $`=`$ $`0,`$ (2.40)
and
$`(_\mu \mathrm{\Phi }_\mathrm{G})`$ $`=`$ $`{\displaystyle \frac{1}{2}}\gamma _{\alpha \mu }(x)p^\alpha {\displaystyle \frac{1}{2}}{\displaystyle _X^x}𝑑z^\lambda (\gamma _{\mu \lambda ,\beta }(z)\gamma _{\beta \lambda ,\mu }(z))p^\beta ,`$ (2.41)
where $`p^\mu `$ is the momentum eigenvalue of the free particle.
It follows that, to first-order in $`𝒂`$ and $`𝝎`$, the Dirac Hamiltonian in the general co-ordinate frame is
$`Hc(𝜶𝒑)+mc^2\beta +V(𝒙),`$ (2.42)
where
$`V(𝒙)`$ $`=`$ $`{\displaystyle \frac{1}{c}}(𝒂𝒙)(𝜶𝒑)+m(𝒂𝒙)\beta 𝝎(𝑳+𝑺){\displaystyle \frac{i\mathrm{}}{2c}}(𝒂𝜶)`$ (2.43)
$`e\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)(𝜶𝑨)+{\displaystyle \frac{e}{c}}𝝎(𝒙\times 𝑨)+e\phi `$
$`+c𝜶(\mathbf{}\mathrm{\Phi }_\mathrm{G})+c(_0\mathrm{\Phi }_\mathrm{G}),`$
the $`𝜶,\beta ,𝝈`$ matrices are those of Minkowski space, and $`𝑳=𝒙\times 𝒑`$ and $`𝑺=\mathrm{}𝝈/2`$ are the orbital and spin angular momenta, respectively.
## 3 Low-Energy Approximation
Although the Dirac Hamiltonian as described by (2.42) and (2.43) is useful as is, there is some benefit in considering approximations which emphasize both the low- and high-energy limits in a particle’s range of motion. In this section, the low-energy approximation is being considered.
According to the FW transformation technique, it is possible to group the Dirac Hamiltonian into the form
$`H`$ $`=`$ $`mc^2\beta +𝒪+,`$ (3.1)
where the “odd” and “even” operators $`𝒪`$ and $`,`$ respectively satisfy $`\{𝒪,\beta \}=[,\beta ]=0`$. For this derivation, we introduce by hand the anomalous magnetic moment
$`{\displaystyle \frac{\kappa e\mathrm{}}{2mc}}\sigma ^{\mu \nu }F_{\mu \nu }={\displaystyle \frac{\kappa e\mathrm{}}{2mc}}(i𝜶𝑬𝝈𝑩),`$ (3.2)
with $`\kappa (g2)/2,`$ as another term in $`V(𝒙)`$, by means of the substitution
$`mc^2\beta \beta \left[mc^2+{\displaystyle \frac{\kappa e\mathrm{}}{2mc}}(i𝜶𝑬𝝈𝑩)\right].`$ (3.3)
Then, given (2.42) and (2.43), it is possible to identify with (3.1), where
$`𝒪`$ $`=`$ $`c𝜶\left[\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝅+(\mathbf{}\mathrm{\Phi }_\mathrm{G}){\displaystyle \frac{i\kappa e\mathrm{}}{2mc^2}}\beta 𝑬{\displaystyle \frac{i\mathrm{}}{2c^2}}𝒂\right],`$ (3.4)
$``$ $`=`$ $`\left[m(𝒂𝒙){\displaystyle \frac{\kappa e\mathrm{}}{2mc}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)(𝝈𝑩)\right]\beta `$ (3.5)
$`𝝎(𝒙\times 𝝅)𝝎\left({\displaystyle \frac{\mathrm{}}{2}}𝝈\right)+e\phi +c(_0\mathrm{\Phi }_\mathrm{G}),`$
where $`𝝅=𝒑e𝑨/c.`$
Following the procedure given by Bjorken and Drell , the transformed Hamiltonian is represented by a series expansion of $`S,`$ according to a unitary transformation
$`H^{}`$ $`=`$ $`UHU^1`$ (3.6)
$``$ $`H+i[S,H]{\displaystyle \frac{1}{2}}[S,[S,H]]{\displaystyle \frac{i}{6}}[S,[S,[S,H]]]`$
$`+{\displaystyle \frac{mc^2}{24}}[S,[S,[S,[S,\beta ]]]]\mathrm{}\dot{S}{\displaystyle \frac{i\mathrm{}}{2}}[S,\dot{S}],`$
and $`S=\mathrm{O}(1/m)`$ is the Hermitian exponent of a unitary transformation operator $`U\mathrm{exp}(iS).`$ By three successive applications of (3.6) for the choice
$`SS_{\mathrm{FW}}={\displaystyle \frac{i}{2mc^2}}\beta 𝒪,`$ (3.7)
the final transformed Hamiltonian becomes
$`H_{\mathrm{FW}}`$ $`=`$ $`mc^2\beta +^{}`$ (3.8)
$`=`$ $`\beta \left(mc^2+{\displaystyle \frac{1}{2mc^2}}𝒪^2{\displaystyle \frac{1}{8m^3c^6}}𝒪^4\right)+`$
$`{\displaystyle \frac{1}{8m^2c^4}}[𝒪,[𝒪,]]{\displaystyle \frac{i\mathrm{}}{8m^2c^4}}[𝒪,\dot{𝒪}].`$
To determine the gravitational corrections in (3.8), it is necessary to isolate the external electromagnetic potentials within the definition of odd and even operators. This implies that $`𝒪𝒪_0+𝒪_1`$ and $`_0+_1,`$ where
$`𝒪_0`$ $`=`$ $`c𝜶𝝅,`$ (3.9)
$`_0`$ $`=`$ $`e\phi .`$ (3.10)
Therefore,
$`𝒪_1`$ $`=`$ $`c𝜶\left[\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝅+(\mathbf{}\mathrm{\Phi }_\mathrm{G}){\displaystyle \frac{i\kappa e\mathrm{}}{2mc^2}}\beta 𝑬{\displaystyle \frac{i\mathrm{}}{2c^2}}𝒂\right],`$ (3.11)
$`_1`$ $`=`$ $`\left[m(𝒂𝒙){\displaystyle \frac{\kappa e\mathrm{}}{2mc}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)(𝝈𝑩)\right]\beta `$ (3.12)
$`𝝎(𝒙\times 𝝅)𝝎\left({\displaystyle \frac{\mathrm{}}{2}}𝝈\right)+c(_0\mathrm{\Phi }_\mathrm{G}).`$
Neglecting the $`𝒪^4`$ contribution and considering only terms up to first-order in $`𝒂,𝝎,`$ and $`1/m^2,`$ it follows that
$`𝒪^2`$ $`=`$ $`𝒪_0^2+𝒪_1^2+\{𝒪_0,𝒪_1\},`$ (3.13)
$`[𝒪,[𝒪,]]`$ $``$ $`[𝒪_0,[𝒪_0,_0]]+[𝒪_0,[𝒪_1,_0]]+[𝒪_0,[𝒪_0,_1]]+[𝒪_1,[𝒪_0,_0]],`$ (3.14)
$`[𝒪,\dot{𝒪}]`$ $``$ $`[𝒪_0,\dot{𝒪}_0]+[𝒪_0,\dot{𝒪}_1]+[𝒪_1,\dot{𝒪}_0].`$ (3.15)
¿From the zeroth-order terms in (3.8), it can be shown that
$`H_{\mathrm{FW}(0)}`$ $`=`$ $`mc^2\beta +{\displaystyle \frac{1}{2mc^2}}\beta 𝒪_0^2+e\phi {\displaystyle \frac{1}{8m^2c^4}}[𝒪_0,[𝒪_0,_0]]{\displaystyle \frac{i\mathrm{}}{8m^2c^4}}[𝒪_0,\dot{𝒪}_0]`$ (3.16)
$`=`$ $`mc^2\beta +\left[{\displaystyle \frac{1}{2m}}\pi ^2{\displaystyle \frac{e\mathrm{}}{2mc}}𝝈𝑩\right]\beta {\displaystyle \frac{e\mathrm{}}{4m^2c^2}}𝝈(𝑬\times 𝝅)`$
$`{\displaystyle \frac{e\mathrm{}^2}{8m^2c^2}}\left[(\mathbf{}𝑬)+i𝝈(\mathbf{}\times 𝑬)\right]+e\phi ,`$
where the third term coupled to $`\beta `$ is the magnetic dipole energy, and the following term is the spin-orbit energy.
Neglecting the time-dependent contributions from (3.15) and considering only those terms up to second-order in $`𝝅`$, it follows that
$`𝒪_1^2`$ $`=`$ $`{\displaystyle \frac{i\kappa e\mathrm{}}{m}}\beta \left[\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)(𝑬𝝅)+𝑬(\mathbf{}\mathrm{\Phi }_\mathrm{G})\right]{\displaystyle \frac{\kappa e\mathrm{}^2}{2mc^2}}\beta (𝒂𝑬)`$ (3.17)
$`{\displaystyle \frac{\kappa e\mathrm{}^2}{2m}}\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)\beta (\mathbf{}𝑬){\displaystyle \frac{i\kappa e\mathrm{}^2}{2m}}\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)\beta 𝝈(\mathbf{}\times 𝑬),`$
$`\{𝒪_0,𝒪_1\}`$ $`=`$ $`(𝒂𝒙)\pi ^2+2c^2((\mathbf{}\mathrm{\Phi }_\mathrm{G})𝝅){\displaystyle \frac{i\kappa e\mathrm{}}{m}}\beta (𝑬𝝅)i\mathrm{}(𝒂𝝅)`$ (3.18)
$`+c^2𝝅\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝅i\mathrm{}c^2(^2\mathrm{\Phi }_\mathrm{G}){\displaystyle \frac{\kappa e\mathrm{}^2}{2m}}\beta (\mathbf{}𝑬){\displaystyle \frac{\kappa e\mathrm{}}{m}}\beta 𝝈(𝑬\times 𝝅)`$
$`+\mathrm{}𝝈(𝒂\times 𝝅)2e\mathrm{}c\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝈(\mathbf{}\times 𝑨){\displaystyle \frac{i\kappa e\mathrm{}^2}{2m}}\beta 𝝈(\mathbf{}\times 𝑬),`$
$`[𝒪_0,[𝒪_1,_0]]`$ $`=`$ $`ie\mathrm{}^2𝒂\mathbf{}\phi e\mathrm{}^2(𝒂𝒙)^2\phi 2e\mathrm{}(𝒂𝒙)𝝈(\mathbf{}\phi \times 𝝅),`$ (3.19)
$`[𝒪_0,[𝒪_0,_1]]`$ $`=`$ $`4imc^2\mathrm{}\beta (𝒂𝝅)\mathrm{}^2c^3^2(_0\mathrm{\Phi }_\mathrm{G})ie\mathrm{}^2c𝝈(𝝎\times (\mathbf{}\times 𝑨))`$ (3.20)
$`+2mc^2\mathrm{}\beta 𝝈(𝒂\times 𝝅)+4mc^2\beta (𝒂𝒙)\pi ^24emc\mathrm{}\beta (𝒂𝒙)𝝈(\mathbf{}\times 𝑨)`$
$`2\mathrm{}c^3𝝈(\mathbf{}(_0\mathrm{\Phi }_\mathrm{G})\times 𝝅)+2\mathrm{}c^2𝝈(𝝎\times 𝝅)\times 𝝅,`$
$`[𝒪_1,[𝒪_0,_0]]`$ $`=`$ $`\mathrm{}^2ec^2\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)^2\phi 2\mathrm{}ec^2\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝈(\mathbf{}\phi \times 𝝅)`$ (3.21)
$`+2\mathrm{}ec^2𝝈((\mathbf{}\mathrm{\Phi }_\mathrm{G})\times \mathbf{}\phi )i\mathrm{}^2e𝝈(𝒂\times \mathbf{}\phi ).`$
After neglecting the non-Hermitian terms in (3.17) - (3.21), it becomes evident that the low-energy approximation for the Dirac Hamiltonian is
$`H_{\mathrm{FW}}`$ $``$ $`mc^2\beta +\left[{\displaystyle \frac{1}{2m}}\pi ^2{\displaystyle \frac{e\mathrm{}}{2mc}}𝝈𝑩\right]\beta {\displaystyle \frac{(g1)e\mathrm{}}{4m^2c^2}}𝝈(𝑬\times 𝝅)`$ (3.22)
$`{\displaystyle \frac{e\mathrm{}^2}{8m^2c^2}}\left[(\mathbf{}𝑬)+i𝝈(\mathbf{}\times 𝑬)\right]+e\phi `$
$`+\left[m(𝒂𝒙){\displaystyle \frac{e\mathrm{}}{2mc}}\left\{\kappa \left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)+\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)\right\}(𝝈𝑩)\right]\beta `$
$`𝝎(𝒙\times 𝝅)𝝎\left({\displaystyle \frac{\mathrm{}}{2}}𝝈\right)+c(_0\mathrm{\Phi }_\mathrm{G})`$
$`+{\displaystyle \frac{1}{2m}}\beta 𝝅\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝅+{\displaystyle \frac{\mathrm{}}{4mc^2}}\beta 𝝈(𝒂\times 𝝅){\displaystyle \frac{\mathrm{}}{4m^2c^2}}𝝈(𝝎\times 𝝅)\times 𝝅`$
$`{\displaystyle \frac{\kappa e\mathrm{}^2}{4m^2c^2}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)\left[(\mathbf{}𝑬)+i𝝈(\mathbf{}\times 𝑬)\right]`$
$`{\displaystyle \frac{\kappa e\mathrm{}^2}{4m^2c^4}}(𝒂𝑬)+{\displaystyle \frac{e\mathrm{}^2}{4m^2c^2}}\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)^2\phi `$
$`+{\displaystyle \frac{1}{m}}\beta (\mathbf{}\mathrm{\Phi }_\mathrm{G})𝝅+{\displaystyle \frac{e\mathrm{}}{4m^2c^2}}𝝈\left(\mathbf{}\phi \times \left[(\mathbf{}\mathrm{\Phi }_\mathrm{G})+\left({\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝝅\right]\right)`$
$`+{\displaystyle \frac{\mathrm{}^2}{8m^2c}}^2(_0\mathrm{\Phi }_\mathrm{G})+{\displaystyle \frac{\mathrm{}}{4m^2c}}𝝈(\mathbf{}(_0\mathrm{\Phi }_\mathrm{G})\times 𝝅).`$
The occurrence of non-Hermitian terms, here neglected, is a well known phenomenon, likely connected with the breakdown of the single-particle interpretation of the Dirac equation in the presence of time-dependent inertial and gravitational fields .
## 4 High-Energy Approximation
Though used less often than the FW-transformation, the CT-transformation follows the same mathematical principles as the former to arrive at the high-energy limit for the Dirac Hamiltonian. Although the FW-transformation can be successfully applied when non-trivial potential energy terms are present, it is far from obvious that the CT-transformation can do the same. This is because it is not clear how to classify the Dirac Hamiltonian into a high-energy analogue of odd and even operators, as found in the FW approach, in order to systematically remove the undesired terms. An unfortunate consequence of this impasse is that it becomes impossible to analyze the motion and properties of fast-moving massive particles in the presence of fields without arbitrarily setting its mass equal to zero within the Hamiltonian. Clearly, this precludes any opportunity to compare the behaviour of these particles with that of strictly massless particles.
It is shown below, however, that it is possible to derive a high-energy approximation of the generalized Dirac Hamiltonian for a spin-1/2 particle moving in a potential defined by both electromagnetic and gravitational fields. For this derivation, the final expression for the Hamiltonian is also in terms of a general co-ordinate frame. The steps now proceed backward from the Dirac equation or corresponding second-order equation in locally Minkowski space.
Given that
$`i\mathrm{}c_{\widehat{0}}\mathrm{\Psi }_0`$ $`=`$ $`H_0\mathrm{\Psi }_0`$ (4.1)
$`=`$ $`\left[c\alpha ^{\widehat{ȷ}}P_{\widehat{ȷ}}+mc^2\beta \right]\mathrm{\Psi }_0,`$
it is possible to define a new wavefunction $`\mathrm{\Psi }_0^{}\mathrm{exp}(iS_{\mathrm{CT}})\mathrm{\Psi }_0,`$ where
$`S_{\mathrm{CT}}`$ $``$ $`{\displaystyle \frac{imc}{2p^2}}\beta (\alpha ^{\widehat{ȷ}}P_{\widehat{ȷ}}),`$ (4.2)
and $`[_{\widehat{\mu }},S_{\mathrm{CT}}]=0`$. It then follows from the CT-transformation that
$`i\mathrm{}c_{\widehat{0}}\mathrm{\Psi }_0^{}`$ $`=`$ $`\left[e^{iS_{\mathrm{CT}}}H_0e^{iS_{\mathrm{CT}}}\right]\mathrm{\Psi }_0^{}`$ (4.3)
$``$ $`H_{\mathrm{CT}(0)}\mathrm{\Psi }_0^{},`$
where
$`H_{\mathrm{CT}(0)}`$ $``$ $`{\displaystyle \frac{E}{p}}(\alpha ^{\widehat{ȷ}}P_{\widehat{ȷ}})`$ (4.4)
is the CT Hamiltonian for the free particle. Since $`\mathrm{\Psi }_0`$ can be related to $`\stackrel{~}{\psi }(x)`$ by (2.23), it can be shown that
$`\mathrm{\Psi }_0^{}`$ $`=`$ $`e^{iS_{\mathrm{CT}}}\mathrm{\Psi }_0`$ (4.5)
$`=`$ $`e^{iS_{\mathrm{CT}}}\left[e^{i\mathrm{\Theta }/\mathrm{}}\psi ^{\prime \prime }(x)\right],`$
where
$`\psi ^{\prime \prime }(x)`$ $`=`$ $`e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}\left(i\gamma ^\mu (x)_\mu {\displaystyle \frac{mc}{\mathrm{}}}\right)^1e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}\psi (x).`$ (4.6)
Substituting (4.5) into (4.3), we obtain
$`i\mathrm{}c\left[{\displaystyle \frac{E}{pc}}\left(e^{iS_{\mathrm{CT}}}\alpha ^{\widehat{ȷ}}e^{iS_{\mathrm{CT}}}\right)_{\widehat{ȷ}}+_{\widehat{0}}\right]e^{i\mathrm{\Theta }/\mathrm{}}\psi ^{\prime \prime }(x)`$ $`=`$ $`0.`$ (4.7)
By using the vierbeins, it becomes possible to describe (4.7) in terms of the general co-ordinate frame, so that
$`i\mathrm{}c[{\displaystyle \frac{E}{pc}}\left(e^{iS_{\mathrm{CT}}}\alpha ^{\widehat{ȷ}}e^{iS_{\mathrm{CT}}}\right)e^\mu {}_{\widehat{ȷ}}{}^{}+e^\mu {}_{\widehat{0}}{}^{}]_\mu \left(e^{i\mathrm{\Theta }/\mathrm{}}\psi ^{\prime \prime }(x)\right)`$ $`=`$ $`0.`$ (4.8)
It is a straightforward matter to evaluate the transformation of $`\alpha ^{\widehat{ȷ}}.`$ Though it is possible to perform the expansion for higher-order terms, it is only necessary to consider the zeroth- and first-order terms. It follows that
$`e^{iS_{\mathrm{CT}}}\alpha ^{\widehat{ȷ}}e^{iS_{\mathrm{CT}}}`$ $``$ $`\alpha ^{\widehat{ȷ}}i[S_{\mathrm{CT}},\alpha ^{\widehat{ȷ}}]`$ (4.9)
$``$ $`\alpha ^{\widehat{ȷ}}+\left({\displaystyle \frac{mc}{p^2}}\right)\beta P^{\widehat{ȷ}}.`$
Substituting (4.9) into (4.7), we obtain the expression
$`i\mathrm{}c_0\psi ^{\prime \prime }(x)`$ $`=`$ $`{\displaystyle \frac{E}{p}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)\left[𝜶^{\prime \prime }𝒑+𝜶^{\prime \prime }\mathbf{}\mathrm{\Theta }+\left({\displaystyle \frac{mc}{p^2}}\right)\beta ^{\prime \prime }\left[(𝒑𝒑)+𝒑\mathbf{}\mathrm{\Theta }\right]\right]\psi ^{\prime \prime }(x)`$ (4.10)
$`𝝎\left[𝒙\times \left(𝒑+\mathbf{}\mathrm{\Theta }\right)\right]\psi ^{\prime \prime }(x)+c_0\mathrm{\Theta }\psi ^{\prime \prime }(x)`$
$`=`$ $`H_{\mathrm{CT}}\psi ^{\prime \prime }(x),`$
where
$`𝜶^{\prime \prime }`$ $`=`$ $`e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}𝜶e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}},`$ (4.11)
$`\beta ^{\prime \prime }`$ $`=`$ $`e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}\beta e^{i\mathrm{\Phi }_\mathrm{S}/\mathrm{}}.`$ (4.12)
To first-order in $`𝒂`$ and $`𝝎`$, the $`_\mu \mathrm{\Theta }`$ contribution is given by (2.38). The only remaining term to be evaluated is $`𝒑\mathbf{}\mathrm{\Theta }.`$ It follows that
$`𝒑\mathbf{}\mathrm{\Theta }`$ $`=`$ $`{\displaystyle \frac{e}{c}}\left(𝒑𝑨\right)+𝒑(\mathbf{}\mathrm{\Phi }_\mathrm{G}).`$ (4.13)
Therefore, the final expression for the CT-Hamiltonian with electromagnetic and gravitational fields present is
$`H_{\mathrm{CT}}`$ $``$ $`{\displaystyle \frac{E}{p}}(𝜶^{\prime \prime }𝒑)+V(𝒙)_{\mathrm{CT}},`$ (4.14)
where
$`V(𝒙)_{\mathrm{CT}}`$ $``$ $`{\displaystyle \frac{E}{pc^2}}(𝒂𝒙)(𝜶^{\prime \prime }𝒑)+{\displaystyle \frac{E}{p}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)\left({\displaystyle \frac{mc}{p^2}}\right)\beta ^{\prime \prime }(𝒑𝒑)𝝎(𝑳+𝑺){\displaystyle \frac{i\mathrm{}}{2c}}(𝒂𝜶)`$ (4.15)
$`{\displaystyle \frac{Ee}{pc}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)𝜶^{\prime \prime }𝑨+{\displaystyle \frac{e}{c}}𝝎(𝒙\times 𝑨)+e\phi +{\displaystyle \frac{E}{p}}𝜶^{\prime \prime }(\mathbf{}\mathrm{\Phi }_\mathrm{G})+c(_0\mathrm{\Phi }_\mathrm{G})`$
$`+{\displaystyle \frac{E}{p}}\left(1+{\displaystyle \frac{𝒂𝒙}{c^2}}\right)\left({\displaystyle \frac{mc}{p^2}}\right)\beta ^{\prime \prime }\left[𝒑(\mathbf{}\mathrm{\Phi }_\mathrm{G}){\displaystyle \frac{e}{c}}\left(𝒑𝑨\right)\right].`$
¿From (2.19), (2.39) - (2.40), (4.11) and (4.12), the first-order expansions of $`𝜶^{\prime \prime }`$ and $`\beta ^{\prime \prime }`$ are
$`𝜶^{\prime \prime }`$ $``$ $`𝜶+𝒫{\displaystyle _X^x}𝑑z^0\left[{\displaystyle \frac{i}{c^2}}(𝒂\times 𝝈)+{\displaystyle \frac{1}{2c}}(𝝎\times 𝜶)\right],`$ (4.16)
$`\beta ^{\prime \prime }`$ $``$ $`\beta +{\displaystyle \frac{1}{c^2}}𝒫{\displaystyle _X^x}𝑑z^0\beta (𝒂𝜶).`$ (4.17)
## 5 Conclusions
The study of inertia/gravity requires in general knowledge of the phase factors $`\mathrm{\Phi }_\mathrm{G}`$, along with $`\mathrm{\Phi }_\mathrm{S}`$ and the Hamiltonian. The procedure followed gives both. It also leads to a solution of the Dirac equation that, as $`\mathrm{\Phi }_\mathrm{G}`$, is exact to first-order in the metric deviation. As well, $`\mathrm{\Phi }_\mathrm{S}`$ is exact.
The Hamiltonian for both the non-relativistic and extreme relativistic cases are represented by (2.42) – (2.43) and (4.14) – (4.15), respectively. Low- and high-energy approximations can be further developed by following well-known procedures. The low-energy approximation only has been derived in detail. A comparison of (2.43) with (4.15) indicates differences only in terms proportional to the mass. This is understandable in view of the different expansions of the energy implied by the corresponding approximations.
Some of the terms that appear in (2.43) and (4.15) are identical and can be identified with corresponding terms that appear in the Hamiltonian of Hehl and Ni. One such instance is represented by the term $`𝝎𝑺`$, the Mashhoon effect , wrongly interpreted by Bell and Leinaas as a version of the Unruh effect. The term $`𝝎𝑳`$ is the Page-Werner effect , while $`m(𝒂𝒙)\beta `$ represents the Bonse-Wroblewski effect . Both effects have been tested experimentally. Also present is the term $`(𝒂𝒙)(𝜶𝒑)/c`$, which is an energy-momentum redshift. While several terms in (4.15) vanish in the limit $`m0`$, the Mashhoon effect is not affected by this limit. This leads one to conclude that the rotation-helicity effects discussed by Cai and Papini for massive neutrinos persist in the vanishing mass limit. The Mashhoon effect is obviously a prime candidate for experiments with accelerators and interferometers. In the latter case, $`\mathrm{\Theta }`$ can be applied to a spacetime loop that is effectively closed because of the coherence of the particle wavefunctions. The result is manifestly gauge invariant and is given by
$`\mathrm{\Theta }={\displaystyle \frac{1}{4}}{\displaystyle _\mathrm{\Sigma }}R_{\mu \nu \alpha \beta }J^{\alpha \beta }𝑑\tau ^{\mu \nu }+{\displaystyle \frac{e}{c}}{\displaystyle _\mathrm{\Sigma }}F_{\mu \nu }𝑑\tau ^{\mu \nu },`$ (5.1)
where $`J^{\alpha \beta }`$ is the total angular momentum of the particle, $`R_{\mu \nu \alpha \beta }`$ is the linearized Riemann tensor, $`F_{\mu \nu }`$ is the electromagnetic field tensor and $`\mathrm{\Sigma }`$ is a surface bounded by the loop. Contributions by the second order derivatives of the metric therefore appear in measurable phases also in the case of inertial fields. Eq.(5.1) clearly indicates that the phase shifts of quantum interferometry depend on the masses of the particles involved and that a strong form of the equivalence principle cannot be present at the quantum level . Nonetheless Eq.(2.22), on which our solution is based, still implies that gravitational fields can be simulated locally by acceleration fields .
Eq.(3.22) for the low-energy Hamiltonian can be now compared with the results obtained by other authors. Let us neglect, for simplicity, the anomalous magnetic moment contributions introduced in Section 3. The Bonse-Wroblewski, Page-Werner and Mashhoon terms can be immediately recognized by inspection. They correspond to the eighth, tenth and eleventh terms respectively. The fifteenth term contains electromagnetic and momentum corrections to the Mashhoon effect. The thirteenth term represents the redshift effect of the kinetic energy already mentioned, but here in the company of its electromagnetic corrections. These also appear in the new inertial spin-orbit term found by Hehl and Ni (the fourteenth term). The fourth and sixth terms represent spin-orbit coupling and are discussed, for instance, by Bjorken and Drell. The third term is also wellknown and represents the magnetic dipole interaction. The Darwin term is the fifth and the nineteenth represents an acceleration correction to it. All remaining terms are proportional to the derivatives of $`\mathrm{\Phi }_\mathrm{G}`$ (see Eq.(2.41). Among these $`c(_0\mathrm{\Phi }_\mathrm{G})+\frac{1}{m}\beta (\mathbf{}\mathrm{\Phi }_\mathrm{G})𝝅`$ appear to dominate. The integral-dependent part of (2.41) yields contributions that are small for small paths and low particle momenta. They produce, in general, curvature contributions for closed spacetime paths, as in interferometry, and lead to (5.1) above. The largest contributions come from the first part of (2.41) which contains the terms $`\frac{1}{2}mc^2\gamma _{00}(x)`$ and $`mc\gamma _{0i}(x)p^i`$ already discussed by De Witt and Papini in connection with the behaviour of superconductors in weak inertial and gravitational fields.
In view of the above, Eq.(2.32) appears remarkably successful in dealing with all the inertial and gravitational effects discussed in the literature.
|
warning/0007/gr-qc0007033.html
|
ar5iv
|
text
|
# Radiation-reaction force on a particle plunging into a black hole
\[
## Abstract
We calculate the self force acting on a scalar particle which is falling radially into a Schwarzschild black hole. We treat the particle’s self-field as a linear perturbation over the fixed Schwarzschild background. The force is calculated by numerically solving the appropriate wave equation for each mode of the field in the time domain, calculating its contribution to the self force, and summing over all modes using Ori’s mode-sum regularization prescription. The radial component of the force is attractive at large distances, and becomes repulsive as the particle approaches the event horizon.
PACS number(s): 04.25-g, 04.70.-s, 04.70.Bw
\]
The problem of finding the equations of motion for a particle in curved spacetime is a long-standing open problem in General Relativity. Recently, this problem has also become timely and extremely important. The planned Laser Interferometer Space Antenna (LISA) is expected to detect (among other sources) the gravitational waves emitted from a compact object orbiting a supermassive black hole (BH). Accurate templates, which include also the radiation-reaction (RR) effects on the compact-object’s orbit, are essential for the detection of the signal.
The traditional approach for calculation of the orbital evolution under RR requires the calculation of the fluxes at infinity and through the BH’s event horizon (EH), of quantities which are constants of motion in the absence of RR. Then one uses balance arguments to relate these fluxes to the local quantities of the object . However, such techniques typically fail, because the evolution of the Carter constant, which is a non-additive constant of motion, cannot be found by balance arguments .
Several prescriptions to include the RR effects in the orbital evolution have been suggested. Quinn and Wald and Mino, Sasaki, and Tanaka recently proposed general approaches for the calculation of self forces. However, it is not clear how to practically apply these approaches for actual computations, the greatest problem being the calculation of the non-local “tail” contribution to the self force, which arises from the failure of the Huygens principle in curved spacetime. More recently, Ori proposed a practical approach for the calculation of the self force , which is based on decomposition of the self force into modes, and on a mode-sum regularization prescription (MSRP). MSRP has been developed in full in Refs. for a scalar particle in static spherically-symmetric spacetimes, and has been applied for several non-trivial cases including a static scalar charge outside a Schwarzschild black hole (SBH) and a scalar charge in circular orbit around a SBH . In addition, there is strong evidence that MSRP is applicable also for electric-field RR , and some evidence that it is applicable also for gravitational-field RR .
MSRP has been directly applied until now only for stationary problems, were the field was decomposed into Fourier-harmonic modes, and the analysis was done in the frequency domain. This was easy to be done in for the case of circular orbits around a SBH, because the RR in that case admits a discrete spectrum. However, in general one faces a time-dependent, evolutionary problem, and one expects the spectrum to be continuous rather than discrete. In this paper, we apply MSRP for the first time to a time-dependent, dynamical problem.
We consider a pointlike massless particle of scalar charge $`q`$, moving along a radial (timelike) geodesic outside a SBH of mass $`M|q|`$, where the metric is $`ds^2=F(r)dt^2+F^1(r)dr^2+r^2d\mathrm{\Omega }^2`$, $`d\mathrm{\Omega }^2`$ being the metric on the unit 2-sphere, and $`F(r)=12M/r`$. Let the particle’s worldline be represented by $`x^\mu =x_\mathrm{p}^\mu (\tau )`$, with $`\tau `$ being the proper time along the geodesic. For inward radial geodesic motion, to be considered here, we have (in Schwarzschild coordinates) $`\dot{\theta }_\mathrm{p}=\dot{\phi }_\mathrm{p}=0`$,
$$\dot{r}_\mathrm{p}=\left[E^2F(r_\mathrm{p})\right]^{1/2},\text{and}\dot{t}_\mathrm{p}=E/F(r_\mathrm{p}),$$
(1)
where a dot denotes $`d/d\tau `$, and $`E`$ is the energy parameter (which is a constant of motion in the absence of the self force). The scalar field $`\mathrm{\Phi }`$ coupled to the particle satisfies the inhomogeneous wave equation
$`\mathrm{}\mathrm{\Phi }=4\pi q{\displaystyle _{\mathrm{}}^{\mathrm{}}}\delta ^4\left[x^\mu x_\mathrm{p}^\mu (\tau )\right](g)^{1/2}𝑑\tau ,`$ (2)
$`g`$ being the metric determinant, and $`\mathrm{}`$ denoting the covariant wave operator. We next decompose $`\mathrm{\Phi }`$ into modes
$$\mathrm{\Phi }=\underset{\mathrm{}=0}{\overset{\mathrm{}}{}}\varphi ^{\mathrm{}}=2\pi q\underset{\mathrm{},m}{}Y_\mathrm{}m(\theta ,\phi )\left[Y_\mathrm{}m^{}(\theta _\mathrm{p},\phi _\mathrm{p})\frac{\psi ^{\mathrm{}}}{r}\right],$$
(3)
where $`Y_\mathrm{}m(\theta ,\phi )`$ are the standard scalar spherical harmonics, an asterisk denotes complex conjugation, and $`\psi ^{\mathrm{}}=\psi ^{\mathrm{}}(r,t;r_\mathrm{p},t_\mathrm{p})`$. By expanding the delta function in Eq. (2) as $`\delta (\theta \theta _\mathrm{p})\delta (\phi \phi _\mathrm{p})=_{\mathrm{},m}\mathrm{sin}\theta Y_\mathrm{}m(\theta ,\phi )Y_\mathrm{}m^{}(\theta _\mathrm{p},\phi _\mathrm{p})`$ and using the orthogonality of the $`Y_\mathrm{}m`$, we find that $`\psi ^{\mathrm{}}`$ satisfies
$$\psi _{,uv}^{\mathrm{}}+V^{\mathrm{}}(r)\psi ^{\mathrm{}}=S(r;r_\mathrm{p}).$$
(4)
Here, $`vt+r^{}`$ and $`utr^{}`$ (with $`r^{}r+2M\mathrm{ln}[(r2M)/2M]`$) are the ingoing and outgoing Eddington null coordinates, correspondingly, the effective potential is given by $`V^{\mathrm{}}(r)=(F/4)\left[\mathrm{}(\mathrm{}+1)r^2+2Mr^3\right]`$, and
$$S\frac{F}{2r}_{\mathrm{}}^{\mathrm{}}\delta (rr_\mathrm{p})\delta (tt_\mathrm{p})𝑑\tau =\frac{F^2}{2rE}\delta (rr_\mathrm{p}).$$
(5)
\[In the last equality we use $`d\tau =dt/\dot{t}=(F/E)dt`$, followed by integration over $`t`$.\] Finally, we express the modes $`\varphi ^{\mathrm{}}`$ in terms of the ($`m`$-independent) functions $`\psi ^{\mathrm{}}`$ by summing over the azimuthal numbers $`m`$ in Eq. (3). For radial motion we thus find that
$$\varphi ^{\mathrm{}}=q\left(\mathrm{}+\frac{1}{2}\right)\frac{\psi ^{\mathrm{}}}{r}.$$
(6)
The total regularized self force (including both the local and the tail parts) exerted on the scalar particle, $`f_\alpha ^{\mathrm{RR}}`$, can be calculated by
$$f_\alpha ^{\mathrm{RR}}\underset{\mathrm{}=0}{\overset{\mathrm{}}{}}\left[f_\alpha ^\mathrm{}\pm A_\alpha ^\pm \left(\mathrm{}+\frac{1}{2}\right)B_\alpha \right]$$
(7)
(evaluated on the particle’s worldline), where $`f_\alpha ^{\mathrm{}}=q\varphi _{,\alpha }^{\mathrm{}}`$ is the (covariant) self-force contribution associated with the $`\mathrm{}`$-mode of the particle’s self-field, and $`A_\alpha ^\pm `$ and $`B_\alpha `$ are regularization parameters, whose values are given by Eqs. (101) and (134) of Ref. . For the radial geodesics considered here, these parameters take the form
$$A_r^{}^\pm =\frac{q^2}{r^2}E,A_t^\pm =\pm \frac{q^2}{r^2}\dot{r},$$
(8)
$$B_r^{}=\frac{q^2}{2r^2}\left(2FE^2\right),B_t=\frac{q^2}{2r^2}\dot{r}E$$
(9)
(with all quantities evaluated at $`x^\mu =x_\mathrm{p}^\mu `$). One should use $`f_\alpha ^\mathrm{}+,A_\alpha ^+`$ ($`f_\alpha ^{\mathrm{}},A_\alpha ^{}`$) when one calculates the field’s gradient from the $`rr_\mathrm{p}^+`$ ($`rr_\mathrm{p}^{}`$) limit (in general $`f_\alpha ^\mathrm{}+f_\alpha ^{\mathrm{}}`$ ). (Of course, the physical quantity $`f_\alpha ^{\mathrm{RR}}`$ can be derived from either of these two values, or from any of their linear combinations.) In practice, we take below the $`rr_\mathrm{p}^{}`$ limit.
Thus, in practice, to derive the self force along any given radial geodesic (parametrized by $`E`$), one should first solve Eq. (4) for the various modes (with appropriately chosen initial data—see below), then construct the quantities $`f_\alpha ^{\mathrm{}}`$, and finally sum over the regularized self-force modes using Eq. (7). This sum over modes is expected to convergence at least as $`1/\mathrm{}`$, as the $`O(1/\mathrm{})`$ term in the $`1/\mathrm{}`$ expansion of $`f_\alpha ^{\mathrm{}}`$ vanishes .
To solve for $`\psi ^{\mathrm{}}`$, we integrate Eq. (4) numerically (in the time domain) on a double-null grid. This grid is spanned by $`v`$ and $`u`$, covering the entire exterior of the SBH (with the EH approached at $`u\mathrm{}`$). A characteristic initial-value problem for $`\psi ^{\mathrm{}}`$ is set up by specifying initial data on two null hypersurfaces $`v=v_0`$ and $`u=u_0`$, taken to intersect at some point along the particle’s worldline. As initial data we take the exact solution corresponding to a static particle held fixed at $`r_0^{}(v_0u_0)/2`$ . This solution is not the actual initial field of the geodesic particle considered here (which is unknown, in general). However, it does approximates the initial field if $`r_0^{}`$ is chosen to be a turning point of the geodesic (when such exists) (these initial data are inexact because the acceleration of the geodesic particle at $`t_0^{}`$ is not the one of a static particle, although its position and velocity are), or—for a marginally bounded particle (with $`E=1`$)—if $`r_0^{}`$ is taken large enough. The difference between the actual initial field and the static initial data results in the occurrence of spurious waves superposed on the actual field; however, one may expect such waves to quickly die off, unveiling the intrinsic behavior of the field. Numerical experiments showed that this is indeed the case: The spurious waves were found to decay fast in all cases examined (see Fig. 2, e.g., for $`r_0^{}=40M`$). For a marginally bounded particle it has been confirmed that the field left after the spurious waves decay becomes independent of $`r_0^{}`$—indicating that one indeed extracts the actual physical behavior. (In addition, we found that the larger $`r_0^{}`$, the smaller the amplitude of the spurious waves, and the quicker they decay.)
To construct the difference scheme for the numerical integration we use a method similar to that applied by Lousto and Price in Ref. . We integrate the field equation (4) over the unit cell shown in Fig. 1, which is centered at $`v,u`$ and whose sides are of length $`2h`$. Let $`\psi _1\psi ^{\mathrm{}}(vh,uh)`$, $`\psi _2\psi ^{\mathrm{}}(vh,u+h)`$, $`\psi _3\psi ^{\mathrm{}}(v+h,uh)`$, and $`\psi _4\psi ^{\mathrm{}}(v+h,u+h)`$, and suppose that $`\psi _1`$, $`\psi _2`$, and $`\psi _3`$ are already known, and we wish to calculate $`\psi _4`$. Integration over the $`\psi _{,uv}^{\mathrm{}}`$ term in Eq. (4) yields (exactly) $`\psi _1\psi _2\psi _3+\psi _4`$. Integration over the potential term yields $`(\psi _1+\psi _4)[1+h^2V^{\mathrm{}}(r)](\psi _2+\psi _3)[1h^2V^{\mathrm{}}(r)]+O(h^3)`$. (Note that because $`\psi ^{\mathrm{}}`$ is continuous across the worldline, the integration of the potential term here is much simpler than in , where the metric perturbations were studied using the Moncrief gauge, in which the wave function suffers a discontinuity across the worldline.) Finally, integrating over the source $`S`$ (which is most easily done by transforming to the $`r,t`$ coordinates, recalling that $`dvdu=2F^1drdt`$), we obtain $`ZS𝑑v𝑑u=0`$ if the worldline does not cross the cell, or, if it does,
$`Z=E^1\left[k(t_{\mathrm{out}})k(t_{\mathrm{in}})\right](t_{\mathrm{out}}t_{\mathrm{in}})+O(h^3).`$ (10)
Here, $`k(t)F[r_\mathrm{p}(t)]/r_\mathrm{p}(t)`$, and $`t_{\mathrm{in}}`$ ($`t_{\mathrm{out}}`$) is the $`t`$ value where the worldline enters (leaves) the cell. We can now extract the desired quantity $`\psi _4`$. To $`O(h^2)`$ we find
$$\psi _4=\psi _1+\left[12h^2V^{\mathrm{}}(r)\right](\psi _2+\psi _3)+Z.$$
(11)
Our code, which is second-order convergent, evolves the scalar field in a straightforward marching. At each grid cell, the code does the following: (i) it decides whether or not the given worldline crosses the cell; (ii) if the worldline crosses the cell, it determines the point where it leaves it and calculates $`t_{\mathrm{out}}`$ (given $`t_{\mathrm{in}}`$) to $`O(h^2)`$; (iii) it uses Eq. (11) to calculate the field $`\psi ^{\mathrm{}}`$ at the cell’s upper point; and (iv) at grid cells containing a section of the worldline, it constructs the quantities $`f_r^{\mathrm{}}`$ and $`f_t^{\mathrm{}}`$ by appropriately extrapolating the field gradients along the worldline (based on the already-derived values of the field at a few neighboring grid points).
We next present our results for a particle released from rest at $`r_0^{}=40M`$ (similar results are obtained also for other values of $`r_0^{}`$ and for the marginally-bound case). Figure 2 displays the behavior of the $`f_r^{}^{\mathrm{}}`$ (2A) and $`f_t^{\mathrm{}}`$ (2B) components of the RR force. The $`\mathrm{}=0`$ components are everywhere negative, whereas all the other modes ($`\mathrm{}1`$) are everywhere positive. Figure 2 also shows the decay of the spurious waves. Clearly, for values of $`r`$ smaller than $`25M`$ they are already too small to be noticed.
Three properties of the behavior of the individual modes are particularly interesting: First, the dipole ($`\mathrm{}=1`$) modes behave differently than the other modes, and the closer to the BH, the less important they are. Second, the relative importance of the higher modes increases approaching the BH. This will require care in the evaluation of the remainder of the series when we sum over all modes (see below). Third, as we approach the BH $`f_r^{}^{\mathrm{}}f_t^{\mathrm{}}`$. This latter property is obvious from the following consideration: The covariant components $`f_v^{\mathrm{}}`$ and $`f_U^{\mathrm{}}`$ \[where $`U`$ is the outgoing Kruskal coordinate, satisfying $`U\mathrm{exp}(u/4M)`$ near the EH\] assume finite values at the EH itself, as $`v`$ and $`U`$ are regular coordinates at the EH. Consequently, $`f_u^{\mathrm{}}`$ vanishes exponentially with $`u`$ approaching the EH, yielding $`f_t^{\mathrm{}}=f_v^{\mathrm{}}+f_u^{\mathrm{}}f_v^{\mathrm{}}`$ and $`f_r^{}^{\mathrm{}}=f_v^{\mathrm{}}f_u^{\mathrm{}}f_v^{\mathrm{}}`$ as we approach the EH. Thus, $`f_r^{}^{\mathrm{}}f_t^{\mathrm{}}`$. This is, in fact, a result of spatial gradients becoming comparable to temporal gradients near the EH. This behavior is shown in Figure 3(A) for two modes ($`\mathrm{}=0,1`$), but similar behavior is found also for all the other modes. Figure 3(B) displays the behavior of the modes as a function of the mode number $`\mathrm{}`$, for various values of $`r`$. The individual modes behave like $`\mathrm{}^2`$ for large values of $`\mathrm{}`$. Note, that the closer the particle to the BH, the later the asymptotic $`\mathrm{}^2`$ behavior starts. Most importantly, the detailed behavior of the modes confirms the expressions for the analytically-derived regularization parameters .
Next, we sum over all modes to find the total RR force. As noted above, the relative importance of the higher modes increases as we approach the horizon. This causes two problems: (i) it is crucial to include an accurate approximation of the remainder of the series due to our computation of only a finite number of modes, and (ii) the noise contribution from the $`\mathrm{}`$-mode to the overall force increases with $`\mathrm{}`$. The $`\mathrm{}^2`$ behavior of the modes indicates that we can sum over the modes and calculate the remainder as was done in Ref. . Specifically, the full RR force is
$$f_\alpha ^{\mathrm{RR}}=\underset{n=0}{\overset{\mathrm{}}{}}f_\alpha ^{n(\mathrm{reg})}+_\alpha ^{\mathrm{}+1},$$
(12)
where the remainder can be approximated by $`_\mu ^{\mathrm{}+1}\mathrm{}^2f_\mu ^{\mathrm{}\mathrm{tail}}\psi ^{(1)}(\mathrm{}+1)`$. Here, $`f_\alpha ^{n(\mathrm{reg})}`$ is the regularized $`\mathrm{}`$-mode of the force, and $`\psi ^{(1)}(x)x^1+x^2/2+x^3/6+O(x^5)`$ is the trigamma function. As we sum the series only up to $`\mathrm{}=17`$, this approximation for $`_\mu ^{\mathrm{}+1}`$ guarantees accuracy of $`7\times 10^4`$ (we neglect here the contribution to $`_\mu ^{\mathrm{}+1}`$ from terms which scale like $`\mathrm{}^3`$). Obviously, approaching the EH $`f_r^{}f_t`$ (as each of the individual modes does). Figure 4 shows the full RR force as a function of $`r`$ for two cases: Fig. 4(A) shows $`f_r^{}^{\mathrm{RR}}`$ for a marginally-bound trajectory ($`E=1`$). At large distances this force behaves like $`f_r^{}^{\mathrm{RR}}(G/c^2)\beta q^2M/r^3`$. The exponent of $`r`$ is found here to a $`1\%`$ accuracy, and we find the parameter $`\beta =(1.00\pm 0.15)\times 10^1`$. Figure 4(B) shows the case of fall from rest, starting from $`r_0^{}=40M`$. At large values of $`r`$ both components of the force vanish, in accord with the vanishing of the force for a static scalar charge. The covariant $`t`$ component, $`f_t^{\mathrm{RR}}`$, is everywhere positive and increases monotonically approaching the BH. This is a consequence of the particle losing energy by radiating, part of which escapes to infinity, and the rest being captured by the BH. The covariant $`r^{}`$ component, $`f_r^{}^{\mathrm{RR}}`$, is attractive at large distances. However, near the peak of the effective potential barrier (near $`r^{}0`$) its behavior changes, and near the EH it approaches the value of $`f_t^{\mathrm{RR}}`$, as expected. Note that both components arrive at the EH at a bounded value. Because $`f_t^{\mathrm{RR}}`$ is expected to be positive (the particle only loses energy by radiating), we infer that $`f_r^{}^{\mathrm{RR}}`$ would also be positive approaching the EH, under very general conditions. In particular, if this behavior persists also for charged BHs, and for an electrically-charged particle, then the properly-defined covariant spatial-component of the RR force at the EH would be repulsive. If this is indeed the case, then the RR force acts to reduce the parameter space for which a nearly-extreme spherical charged BH can be overcharged . The question of whether cosmic censorship for that case is saved by RR effects, however, awaits further considerations.
We thank Amos Ori, Lee Lindblom, and Kip Thorne for discussions. LMB wishes to thank the Technion Institute of Theoretical Physics, where part of this research was done, for hospitality. At Caltech this research was supported by NSF grants AST-9731698 and PHY-9900776 and by NASA grant NAG5-6840.
|
warning/0007/hep-th0007088.html
|
ar5iv
|
text
|
# Untitled Document
Fusion in coset CFT
from
admissible singular-vector decoupling
P. Mathieu<sup>1</sup> pmathieu@phy.ulaval.ca; work supported by NSERC (Canada) and FCAR (Québec)., J. Rasmussen<sup>2</sup> rasmussj@cs.uleth.ca; work supported by NSERC (Canada). and M.A. Walton<sup>3</sup> walton@uleth.ca; work supported by NSERC (Canada).
Département de Physique, Université Laval, Québec (Québec), Canada G1K 7P4
Physics Department, University of Lethbridge, Lethbridge, Alberta, Canada T1K 3M4
Abstract: Fusion rules for Wess-Zumino-Witten (WZW) models at fractional level can be defined in two ways, with distinct results. The Verlinde formula yields fusion coefficients that can be negative. These signs cancel in coset fusion rules, however. On the other hand, the fusion coefficients calculated from decoupling of singular vectors are non-negative. They produce incorrect coset fusion rules, however, when factorisation is assumed. Here we give two prescriptions that yield the correct coset fusion rules from those found for the WZW models by the decoupling method. We restrict to the Virasoro minimal models for simplicity, and because decoupling results are only complete in the $`\widehat{su}(2)`$ case.
1. Introduction
The formulation of WZW models in terms of an action is problematic at fractional admissible level. But their algebraic treatment seems to lead to well-defined conformal field theories (CFTs). One vexing point, however, is that the fusion rules do not seem to be uniquely defined. More precisely, there are two distinct ways by which they can be calculated. The Verlinde formula and the decoupling of singular vectors yield different results.
Although WZW models at fractional level may turn out to be pathological, certain coset CFTs built from them are consistent non-unitary models. For example, the minimal Virasoro models may be described by the diagonal cosets
$$\frac{\widehat{su}(2)_k\widehat{su}(2)_1}{\widehat{su}(2)_{k+1}}.$$
$`(1.1)`$
Here $`\widehat{su}(2)_k`$ indicates the non-twisted $`\widehat{su}(2)`$ Kac-Moody algebra that is the central extension of the simple Lie algebra $`su(2)`$, at fixed level $`k`$. Now, the fusion rules for the minimal models and similar coset theories are unambiguous. But while the WZW Verlinde fusions have already been shown to be compatible with them, the decoupling fusions have not. Here we attempt to fill this gap. We restrict consideration to the coset (1.1) for simplicity, and because complete results for decoupling fusion are only known in the $`\widehat{su}(2)`$ case.
The fusion rules for the minimal models $`M(p,p^{})`$ take a simple factorised form :
$$𝒩_{(\dot{r};\dot{s})(\ddot{r};\ddot{s})}^{[p,p^{}](r;s)}=𝒩_{\dot{r}\ddot{r}}^{[p^{}]r}𝒩_{\dot{s}\ddot{s}}^{[p]s},$$
where $`𝒩_{\dot{r}\ddot{r}}^{[p^{}]r}`$ stands for an $`\widehat{su}(2)`$ fusion coefficient at level $`p^{}2`$. The three labels $`\dot{r},\ddot{r}`$ and $`r`$ are the finite Dynkin labels of three integrable affine weights, plus one. Unitary models are described by the coset (1.1) with $`p=p^{}+1`$ and $`p^{}=k+2`$. In the non-unitary case, the same coset construction holds, but the level must be fractional : $`k=t/u`$. Here $`u`$ is a positive integer, while the integer $`t`$ is relatively prime to $`u`$ and bounded by $`t22u`$. These conditions define an admissible level. For the general minimal models, unitary and non-unitary both, we have
$$pp^{}=u,up^{}=k+2.$$
$`(1.2)`$
The unitary case is recovered when $`u=1`$.
The factorisation (1.2) will be our starting point. We stress that it does not rely on special properties of the minimal models other than the existence of a diagonal coset realisation for which one WZW component is at level 1, i.e. (1.1). In that case, the integrable weight at the level 1 can be dropped from the triplet of weights that together label a coset field. That is because the branching rules for the embedding $`\widehat{su}(2)_k\widehat{su}(2)_1\widehat{su}(2)_{k+1}`$ selecting these triplets are very simple. Once the two weights of $`\widehat{su}(2)_k`$ and $`\widehat{su}(2)_{k+1}`$ are fixed, the branching conditions just determine the unique level-one weight required as the third label.
In the non-unitary case, the fractional part of the branching condition leads to a simple constraint: the finite (horizontal) parts of the fractional parts of the two non-integrable weights must be equal. If this is satisfied, the original (though slightly modified) branching condition plays the same role as in the unitary cosets. It simply determines the level-one weight required in the triplet of weights labelling the coset field.
In non-unitary coset models, there is a large number of field identifications. These include types not present in their unitary analogues. It has been proved in that we can always choose a coset-field representative whose three weights have vanishing finite fractional parts. For fusions involving such weights, the two methods yield identical results. However, the coset fusion rules should be computable with any choice of coset representatives, and the result must be independent of this choice. It is natural then, to test the two methods of calculating admissible fusions by using coset representatives with non-vanishing finite fractional parts.
The Verlinde formula generates negative WZW fusion coefficients for WZW models at fractional admissible levels , perhaps indicating an ill-defined theory. In coset models, however, the Verlinde method always leads to positive fusion coefficients as all the negative signs cancel. Moreover, for the coset (1.1), it reproduces (1.2).
On the other hand, singular-vector decoupling leads to non-negative fusion coefficients. From that point of view then, the non-unitary WZW models look well-defined. The naive application of these fusion coefficients to the computation of coset fusion rules seems to fail, however: assuming factorisation, even the multiplicities are not correct.
But the relation between the WZW fusions and the coset ones may need to be modified in the fractional case. Here we search for that modification, and find a new product of component fields that yields results compatible with the Verlinde method. It is evocative (in being formulated as a sort of supertrace) of the $`\widehat{osp}(1|2)`$ Lie superalgebra pattern of the $`\widehat{su}(2)`$ singular-vector decoupling fusions, that has already been described in .
Alternatively, the coset fusion rules may be obtained by using certain truncations of the singular-vector decoupling fusion. Several possibilities exist. We discuss some of them, and we provide a natural motivation for a particular truncation, while relating it to Verlinde fusion.
2. Admissible representations of $`\widehat{su}(2)_k`$
The set of admissible representations contains the set of integrable representations. At fixed level, the characters of admissible representations form a finite-dimensional representation of the modular group . The admissible $`\widehat{su}(2)_k`$ representations are the building blocks in the algebraic formulation of the WZW models at admissible level $`k`$, and are in one-to-one correspondence with the WZW primary fields.
Admissible representations of $`\widehat{su}(2)_k`$ at fractional admissible level $`k=t/u`$ can be described rather simply. The highest weight $`\lambda `$ of such a representation can be written as
$$\lambda =\lambda ^I(k+2)\lambda ^F,$$
in terms of two integrable weights $`\lambda ^I`$ and $`\lambda ^F`$ at respective levels
$$\begin{array}{cc}\hfill k^I=& u(k+2)20,\hfill \\ \hfill k^F=& u10.\hfill \end{array}$$
The superscripts $`I`$ and $`F`$ refer to integer and fractional. Notice that although $`\lambda ^F`$ is responsible for the fractional part of $`\lambda `$, it is itself an integrable weight. We will use the Dynkin label description
$$\lambda =\lambda _0\omega _0+\lambda _1\omega _1,$$
$`(2.1)`$
where the $`\omega _i`$ are the affine fundamental weights.<sup>4</sup> Since in the present work only affine weights are used, the usual “hats” are omitted for simplicity. An affine weight $`\lambda `$ at level $`k`$ has
$$\lambda _0+\lambda _1=k,$$
so that if $`k`$ is fixed, $`\lambda _1`$ specifies the weight uniquely. Consequently, $`(\lambda _1^I,\lambda _1^F)`$ specifies an admissible $`\widehat{su}(2)_k`$ weight, and this shorthand notation will prove useful in the following.
3. WZW fusion rules from the Verlinde formula
The Verlinde formula leads to the following expression for the $`\widehat{su}(2)_k`$ fusion rules at fractional level $`k`$ :
$$\lambda \times \mu =\underset{\genfrac{}{}{0pt}{}{\nu _1^I\delta _{k^I}}{\nu _1^F\lambda _1^F+\mu _1^F(\mathrm{mod}u)}}{}(a^I)^{[(\lambda _1^F+\mu _1^F)/u]}\nu .$$
$`(3.1)`$
Here the set $`\delta _{k^I}`$ is defined as
$$\delta _{k^I}=\left\{\nu _1^I\right|\lambda _1^I+\mu _1^I+\nu _1^I0(\mathrm{mod}2);|\lambda _1^I\mu _1^I|\nu _1^Ik^I|k^I\lambda _1^I\mu _1^I|\},$$
$`(3.2)`$
and the square brackets denote the integer part:
$$[(\lambda _1^F+\mu _1^F)/u]=\{\begin{array}{cc}0& \mathrm{if}\lambda _1^F+\mu _1^F<u\\ 1& \mathrm{if}\lambda _1^F+\mu _1^Fu\end{array}.$$
$`a^I`$ is the outer automorphism with its action restricted to the integer part of $`\nu `$:
$$a^I\nu =a\nu ^I(k+2)\nu ^F.$$
$`(3.3)`$
$`a`$ itself simply interchanges the two Dynkin labels of an affine weight:
$$a\lambda =a(\lambda _0\omega _0+\lambda _1\omega _1)=\lambda _1\omega _0+\lambda _0\omega _1=k(\omega _0+\omega _1)\lambda ,$$
$`(3.4)`$
where (2.2) was used in the last equality. Of course,
$$a\lambda =a\left(\lambda ^I(k+2)\lambda ^F\right)=a\lambda ^I(k+2)a\lambda ^F.$$
In convenient notation,
$$a(\lambda _1^I,\lambda _1^F)=(k^I\lambda _1^I,k^F\lambda _1^F).$$
The result (3.1) is proved along the following lines. We first relate the ratio $`\gamma _\lambda ^{(\sigma )}=S_{\lambda ,\sigma }/S_{0,\sigma }`$ to the finite character $`\chi _\lambda `$ evaluated at $`\xi _\sigma =2\pi i(\sigma +\rho )/(k+2)`$. The line of the argument is then standard: the $`\gamma _\lambda ^{(\sigma )}`$’s satisfy the fusion rules, while the product of $`\chi _\lambda (\xi _\sigma )`$’s is decomposed using the tensor-product coefficients. The fusion coefficients at level $`k`$ are thus related to tensor-product coefficients and then to fusion coefficients at level $`k^I`$.<sup>5</sup> That it is $`k`$ and not $`k^I`$ that appears in the sine factor of the $`S`$ matrix prevents a more direct relation between $`𝒩^{(k)}`$ and $`𝒩^{(k^I)}`$. When $`\lambda _1^F+\mu _1^F<u`$, the above result is immediate. In that case the fusion coefficients are non-negative, and reduce to the $`\widehat{su}(2)_{k^I}`$ fusion coefficients with a $`\widehat{u}(1)`$ factor arising from the fractional parts (the $`\widehat{u}(1)`$ interpretation means that the fractional parts simply add ($`\mathrm{mod}u`$)). When $`\lambda _1^F+\mu _1^Fu`$, we can write
$$\lambda ^F+\mu ^F=\nu ^F+u\zeta .$$
$`(3.5)`$
Clearly, $`\zeta `$ can always be written in the form $`\zeta _1a\omega _0`$. This $`\zeta `$ piece is responsible for an extra phase factor. Up to a sign, this phase factor can be absorbed in the transformation $`\chi _{\nu ^I}\chi _{a\nu ^I}`$. The resulting minus sign is the one that makes fusion coefficients negative when $`\lambda _1^F+\mu _1^Fu`$. The fusion coefficients are now expressed in terms of $`𝒩_{\lambda ^I\mu ^I}^{a\nu ^I}`$.
Ignoring outer automorphisms and minus signs, the pattern of fusion rules as computed with the Verlinde formula is
$$𝒩[\widehat{su}(2)_k]𝒩[\widehat{su}(2)_{k^I}]𝒩[\widehat{u}(1)_{k^F}].$$
$`(3.6)`$
In the last factor we have indicated that the $`\widehat{u}(1)`$ may be interpreted to have level $`k^F=u1`$.
We should stress that the computations of fusion rules via BRST cohomology or vertex-operator methods agree with those just written when $`\lambda _1^F+\mu _1^F<u`$.
4. Coset fusion rules from the Verlinde formula
Now let $`\lambda `$ and $`\nu `$ stand for the admissible weights at levels $`k`$ and $`k+1`$, respectively, that label a coset field $`\{\lambda ,\nu \}`$. Let two other coset fields be labeled $`\{\dot{\lambda },\dot{\nu }\}`$, $`\{\ddot{\lambda },\ddot{\nu }\}`$ in a similar way. The coset fusion rules take the form
$$𝒩_{\{\dot{\lambda },\dot{\nu }\}\{\ddot{\lambda },\ddot{\nu }\}}^{\{\lambda ,\nu \}}=𝒩_{\dot{\lambda }\ddot{\lambda }}^{(k)\lambda }𝒩_{\dot{\nu }\ddot{\nu }}^{(k+1)\nu }.$$
Using (3.1), this becomes
$$𝒩_{\{\dot{\lambda },\dot{\nu }\}\{\ddot{\lambda },\ddot{\nu }\}}^{\{\lambda ,\nu \}}=\{\begin{array}{cc}𝒩_{\dot{\lambda }^I\ddot{\lambda }^I}^{(k^I)\lambda ^I}𝒩_{\dot{\nu }^I\ddot{\nu }^I}^{(k^I+u)\nu ^I}& \mathrm{if}\dot{\lambda }_1^F+\ddot{\lambda }_1^F<u\\ 𝒩_{\dot{\lambda }^I,\ddot{\lambda }^I}^{(k^I)a\lambda ^I}𝒩_{\dot{\nu }^I,\ddot{\nu }^I}^{(k^I+u)a\nu ^I}& \mathrm{if}\dot{\lambda }_1^F+\ddot{\lambda }_1^Fu.\end{array}$$
Notice that if $`\dot{\lambda }_1^F+\ddot{\lambda }_1^F<u`$, it also follows that $`\dot{\nu }_1^F+\ddot{\nu }_1^F<u`$ since $`\lambda `$ and $`\nu `$ are part of the same coset field, and hence have the same fractional part. However, the coset field identification
$$\{\lambda ,\nu \}=\{a\lambda ,a\nu \}$$
$`(4.1)`$
holds in the non-unitary case as well as in the unitary one. Therefore, the two expressions in (4.1) are equivalent. The result is equivalent to (1.2); the latter is expressed in terms of $`\rho `$-shifted weights. This result is manifestly independent of the choice of the coset-field representative since here we have made no choice of the value of the fractional part.
5. WZW fusion rules from singular-vector decoupling
Fusion rules can be calculated by enforcing the decoupling of singular vectors. In the integrable WZW case , this method leads to results that agree with those obtained by the Verlinde formula. Fractional levels present the novelty of two types of solutions, called A and B below. The admissible representations $`\nu `$ that appear in the product $`\lambda \times \mu `$ are <sup>6</sup> The parameters in are related to ours as follows: $`r_{\mathrm{AY}}=\lambda _1^I+1`$ and $`s_{\mathrm{AY}}=\lambda _1^F`$, $`t_{\mathrm{AY}}=k+2`$, $`p_{\mathrm{AY}}=t+2u`$ and $`q_{\mathrm{AY}}=u`$.:
| | $`\mathrm{Case}\mathrm{A}:`$ | $`|\lambda _1^I\mu _1^I|\nu _1^Ik^I|k^I\lambda _1^I\mu _1^I|`$ |
| --- | --- | --- |
| | | $`|\lambda _1^F\mu _1^F|\nu _1^Fk^F|k^F\lambda _1^F\mu _1^F|,`$ |
| | | |
| | $`\mathrm{Case}\mathrm{B}:`$ | $`|\lambda _1^I\mu _1^I|k^I\nu _1^Ik^I|k^I\lambda _1^I\mu _1^I|`$ |
| | | $`|\lambda _1^F\mu _1^F|+1k^F\nu _1^Fk^F1|k^F\lambda _1^F\mu _1^F|.`$ |
$`(5.1)`$
In all 4 lines just written, the bounded quantities take all values from their lower bounds to their upper bounds, increasing in steps of 2.
For later use, we separate the fusion coefficients determined by (5.1) into A and B parts:
$$𝒩_{\lambda \mu }^\nu =𝒜_{\lambda \mu }^\nu +_{\lambda \mu }^\nu .$$
Notice that a B-type solution is intrinsically fractional, that is, $`_{\lambda \mu }^\nu =0`$ when $`u=1`$. More precisely, the B-set of solutions can be non-empty only if the following bounds are satisfied:
$$1\lambda _1^F,\mu _1^F,\nu _1^Fu2.$$
$`(5.2)`$
Hence, not only do the fractional parts need to be non-zero, but $`u`$ must be greater than 2.
Case A describes an $`\widehat{su}(2)_{k^I}\times \widehat{su}(2)_{k^F}`$ fusion rule, i.e., a separate $`\widehat{su}(2)`$ fusion for each of the integer and the fractional parts:
$$𝒜_{\lambda \mu }^\nu =𝒩_{\lambda ^I\mu ^I}^{(k^I)\nu ^I}𝒩_{\lambda ^F\mu ^F}^{(k^F)\nu ^F}.$$
Similarly, case B may be factorised into two $`\widehat{su}(2)`$ fusions, though with shifted weights and level in the fractional sector:
$$_{\lambda \mu }^\nu =𝒩_{\lambda ^I\mu ^I}^{(k^I)a\nu ^I}𝒩_{\lambda ^F\omega _0\omega _1\mu ^F\omega _0\omega _1}^{(k^F2)a(\nu ^F\omega _0\omega _1)}.$$
A further observation is that the combined set of A and B fusions (5.1) displays an $`\widehat{osp}(1|2)_k`$ fusion pattern. We recall that $`\widehat{osp}(1|2)_k`$ fusion rules are equivalent to those for $`\widehat{su}(2)_k`$, except that the constraint $`\lambda _1+\mu _1\nu _10`$ (mod 2) is weakened to $`\lambda _1+\mu _1\nu _10`$ (mod 1) . That is, all integer intermediate values of the Dynkin label $`\nu _1`$ are allowed. The fusion rules obtained by the singular-vector decoupling method are seen to allow sets of weights isomorphic to those allowed by $`\widehat{osp}(1|2)`$ fusion rules. It is the shifts in the lower and upper bounds in case B of (5.1) that allow this simple interpretation. Thus the case-B solutions correspond to fermionic contributions in this analogy.
Let us inspect further the set of weights appearing in the fusion rule encoded in (5.1). For a fixed non-vanishing fusion $`\lambda \times \mu `$, there is always one more fractional A-type solution $`\nu ^F`$ than fractional B-type solution. Furthermore, as already indicated in (5.3) and (5.3), each of the four lines in (5.1) corresponds to an $`\widehat{su}(2)`$ fusion. Finally, in order to emphasise the similarity between the two types, one is written in terms of $`\nu `$ while the other is written in terms of $`a\nu `$. We also note that acting with $`ac`$ on a weight with non-vanishing fractional part has the effect of reducing the fractional part (of the first Dynkin label) by one while leaving the integer part unchanged. Here $`c`$ represents the action of charge conjugation as defined by the square of the modular matrix $`S`$: $`(S^2)_{\lambda ,\mu }=\pm \delta _{c\lambda ,\mu }`$. Explicitly, we have
$$c(\lambda _1^I,\lambda _1^F)=\delta _{\lambda _1^F,0}(\lambda _1^I,\lambda _1^F)+(1\delta _{\lambda _1^F,0})(k^I\lambda _1^I,u\lambda _1^F).$$
In conclusion, for fixed $`\nu ^I`$, the fractional parts paired with it form the pattern illustrated by the following diagram:
$$\begin{array}{cc}& \\ \hfill 𝒜_{\lambda \mu }^{(\nu _1^I,k^F|k^F\lambda _1^F\mu _1^F|)}& \\ & c\hfill \\ & _{\lambda \mu }^{a(\nu _1^I,k^F|k^F\lambda _1^F\mu _1^F|1)}\hfill \\ & c\hfill \\ \hfill 𝒜_{\lambda \mu }^{(\nu _1^I,k^F|k^F\lambda _1^F\mu _1^F|2)}& \\ & c\hfill \\ & \mathrm{}\hfill \\ & c\hfill \\ & _{\lambda \mu }^{a(\nu _1^I,|\lambda _1^F\mu _1^F|+1)}\hfill \\ & c\hfill \\ \hfill 𝒜_{\lambda \mu }^{(\nu _1^I,|\lambda _1^F\mu _1^F|)}& .\hfill \end{array}$$
All arrows indicate an action of the charge conjugation $`c`$ on the upper weight. All fusion coefficients $`𝒜`$ and $``$ depicted here have identical value, i.e. 0 or 1.
Thus, up to the action of the outer automorphism group, we have the following schematic structure
$$𝒩[\widehat{su}(2)_k]𝒩[\widehat{su}(2)_{k^I}]𝒩[\widehat{osp}(1|2)_{k^F}].$$
$`(5.3)`$
This is very different from the pattern that follows from the Verlinde formula (3.6). The occurrence of the $`\widehat{osp}(1|2)`$ pattern makes contact with the results obtained in .
We would like to stress that the result (5.1) obtained in is supported by calculations of four-point functions using a free-field realisation (see also for similar results from a somewhat different approach). These correlation functions satisfy the Knizhnik-Zamolodchikov equations and are invariant under projective and duality transformations, testifying for their soundness. Generalisations are reported in .
6. WZW fusion rules: comparing the two methods
To compare the results of the two methods at the WZW level, consider the special case $`k=4/3`$ ($`k^I=0`$ and $`k^F=2`$). There are $`(k^I+1)(k^F+1)=3`$ admissible representations. In the notation $`(\lambda _1^I,\lambda _1^F)`$, these are $`(0,0),(0,1)`$ and $`(0,2)`$. The Verlinde formula produces the fusion rules
$$\begin{array}{cc}& (0,1)\times (0,1)=(0,2),\hfill \\ & (0,1)\times (0,2)=(0,0),\hfill \\ & (0,2)\times (0,2)=(0,1).\hfill \end{array}$$
$`(6.1)`$
We see the minus signs arising in the cases where $`\lambda _1^F+\mu _1^Fu=3`$. On the other hand, the results of give (combining solutions from cases A and B, cf. (5.1)):
$$\begin{array}{cc}& (0,1)\times (0,1)=(0,0)+(0,1)+(0,2),\hfill \\ & (0,1)\times (0,2)=(0,1),\hfill \\ & (0,2)\times (0,2)=(0,0).\hfill \end{array}$$
$`(6.2)`$
Case B leads to a non-zero contribution only for the first fusion: this is the $`(0,1)`$ representation. We see the $`\widehat{osp}(1|2)`$-like structure showing up in the fractional sector: in the first example, there appears an intermediate weight between those predicted by the $`\widehat{su}(2)_2`$ fusion rule. Let us emphasise that since $`k^I=0`$ and $`k^F=2`$, $`a(0,1)=(k^I0,k^F1)=(0,1)`$.
One can easily verify that the fusion matrices defined by (6.2) are not diagonalisable by any symmetric matrix. On the other hand, the Verlinde formula guarantees that the symmetric modular-$`S`$ matrix diagonalises the fusion matrices encoded in (6.1).
This simple example illustrates important differences between the two methods of computing the WZW fusion rules. Due to the lack of a physical realisation of WZW models at fractional level, it seems difficult to favour one method over the other. However, given that the number of weights appearing in the fusion is quite different in the two cases, one might expect notable differences in the coset models.
7. Coset fusion rules from WZW singular-vector decoupling
Let us consider the simplest non-unitary minimal model, the Yang-Lee singularity. It has $`(p,p^{})=(5,2)`$, and hence $`k=4/3`$ in the diagonal coset description. There are 12 coset fields that can be grouped into two sets among which all 6 fields are identified. That is, there are two equivalence classes of coset labels under field identification, corresponding to the two Yang-Lee primary fields (see , reviewed in sect. 18.7.2 of , for details about the field identifications, and for an improved approach). In the notation $`\{(\lambda _1^I,\lambda _1^F),\mu _1;(\nu _1^I,\nu _1^F)\}`$, where the three weights are at levels $`k=4/3,\mathrm{\hspace{0.17em}1}`$ and $`k+1=1/3`$, respectively, they are:
$$\begin{array}{cc}\hfill h=0& \{(0,0),0;(0,0)\}\{(0,2),1;(3,2)\}\{(0,1),1;(0,1)\}\hfill \\ & \{(0,1),0;(3,1)\}\{(0,2),0;(0,2)\}\{(0,0),1;(3,0)\},\hfill \\ \hfill h=\frac{1}{5}& \{(0,0),1;(1,0)\}\{(0,2),0;(2,2)\}\{(0,1),0;(1,1)\}\hfill \\ & \{(0,1),1;(2,1)\}\{(0,2),1;(1,2)\}\{(0,0),0;(2,0)\}.\hfill \end{array}$$
For our example, we will focus on the $`h=0`$ identity field.
Let us pick $`\{(0,1),1;(0,1)\}`$ and $`\{(0,1),0;(3,1)\}`$ as representatives of the identity field, and omit the level-one weights, relabelling our choices as $`\{(0,1);(0,1)\}`$ and $`\{(0,1);(3,1)\}`$. We calculate the fusion rule of the identity field ($`I\times I=I`$) using these representatives, computing the level $`k`$ and the level $`k+1`$ fusions independently:
$$\begin{array}{cc}\hfill \{(0,1);(0,1)\}\times \{(0,1);(3,1)\}& =\{(0,1)\times (0,1);(0,1)\times (3,1)\}\hfill \\ & =\{(0,0)_\mathrm{A}+(0,1)_\mathrm{B}+(0,2)_\mathrm{A};(3,0)_\mathrm{A}+(0,1)_\mathrm{B}+(3,2)_\mathrm{A}\}.\hfill \end{array}$$
$`(7.1)`$
The subscripts indicate case A and B results. Next we combine the separate WZW fields into coset fields. A simple product of the different fields leads to 9 possibilities. Not all resulting combinations respect the coset branching rules, however. If we restrict ourselves to the ones that do, we get only three candidates:
$$\{(0,0)_\mathrm{A};(3,0)_\mathrm{A}\}+\{(0,1)_\mathrm{B};(0,1)_\mathrm{B}\}+\{(0,2)_\mathrm{A};(3,2)_\mathrm{A}\}.$$
All of these three fields are representatives of the identity, as they should be (cf. the list (7.1)). However, the fusion coefficient of the identity with itself is then 3 rather than 1.
A facetious cure for that discrepancy is to introduce a minus sign in front of the second coset field, transforming $`1+1+1`$ into $`11+1`$. But this illustrates our general proposal. We show that (i) only fusions that are of type A, or type B, in both labels should contribute, and suggest that (ii) the type B contributions should be negative. Because of property (i), the proposal is of a trace form. As already mentioned, case B is analogous to fermionic contributions in $`\widehat{osp}(1|2)`$ fusion. The minus sign of property (ii) thus gives our proposal a structure similar to a supertrace.
8. A coset “supertrace”
In this section, we will describe our general proposal explicitly, and prove it computes the correct minimal model fusion rules.
Consider the fusion of the two coset fields:
$$\begin{array}{cc}\hfill \{(\dot{\lambda }_1^I,\dot{\lambda }_1^F);(\dot{\nu }_1^I,\dot{\nu }_1^F)\},& \dot{\lambda }_1^F=\dot{\nu }_1^F,\hfill \\ \hfill \{(\ddot{\lambda }_1^I,\ddot{\lambda }_1^F);(\ddot{\nu }_1^I,\ddot{\nu }_1^F)\},& \ddot{\lambda }_1^F=\ddot{\nu }_1^F.\hfill \end{array}$$
Our main result is the following proposal:
$$\begin{array}{cc}\hfill \{(\dot{\lambda }_1^I& ,\dot{\lambda }_1^F);(\dot{\nu }_1^I,\dot{\nu }_1^F)\}\times \{(\ddot{\lambda }_1^I,\ddot{\lambda }_1^F);(\ddot{\nu }_1^I,\ddot{\nu }_1^F)\}\hfill \\ & =\mathrm{str}\{(\dot{\lambda }_1^I,\dot{\lambda }_1^F)\times (\ddot{\lambda }_1^I,\ddot{\lambda }_1^F);(\dot{\nu }_1^I,\dot{\nu }_1^F)\times (\ddot{\nu }_1^I,\ddot{\nu }_1^F)\}\hfill \\ & =\underset{\lambda }{}\underset{\nu }{}\left(𝒜_{\dot{\lambda }\ddot{\lambda }}^\lambda 𝒜_{\dot{\nu }\ddot{\nu }}^\nu _{\dot{\lambda }\ddot{\lambda }}^\lambda _{\dot{\nu }\ddot{\nu }}^\nu \right)\delta _{\lambda _1^F,\nu _1^F}\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\nu _1^F)\}.\hfill \end{array}$$
The delta function is due to the branching conditions. It is the only condition to impose here as we do not specify the weight $`\mu `$. The latter is given by the remaining branching condition
$$\mu _1\nu _1^I\lambda _1^I\lambda _1^F(\mathrm{mod}2).$$
Similar branching conditions apply to the coset fields (8.1). The finite double-summation in (8.1) may be written
$$\underset{\lambda ^F}{}\underset{\lambda ^I,\nu ^I}{}\left(𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda ^I,\lambda ^F)}𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu ^I,\lambda ^F)}_{\dot{\lambda }\ddot{\lambda }}^{(\lambda ^I,\lambda ^F)}_{\dot{\nu }\ddot{\nu }}^{(\nu ^I,\lambda ^F)}\right)\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\}.$$
Our proof that this reduces to (1.2) relies on the structure of field identifications . Those that are important here can be written in terms of the outer (diagram) automorphism $`a`$, and the operation $`c`$, related to charge conjugation, as defined by $`S^2`$. Their actions on (first Dynkin labels of) coset weights are
$$\begin{array}{cc}\hfill a\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\}=& \{(k^I\lambda _1^I,k^F\lambda _1^F);(k^I+u\nu _1^I,k^F\lambda _1^F)\},\hfill \\ \hfill c\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\}=& \delta _{\lambda _1^F,0}\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\}\hfill \\ & +(1\delta _{\lambda _1^F,0})\{(k^I\lambda _1^I,u\lambda _1^F);(k^I+u\nu _1^I,u\lambda _1^F)\}.\hfill \end{array}$$
The idea behind the proof is the following. First, we notice that for the product $`𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda ^I,\lambda ^F)}𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu ^I,\lambda ^F)}`$ to be non-vanishing, $`\lambda ^F`$ needs to be allowed by the A-type fusion of $`\dot{\lambda }`$ with $`\ddot{\lambda }`$ as well as by the A-type fusion of $`\dot{\nu }`$ with $`\ddot{\nu }`$. However, due to the branching conditions (8.1), these prerequisites are equivalent, and we conclude
$$\begin{array}{cc}& 𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda ^I,\lambda ^F)}𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu ^I,\lambda ^F)}0\mathrm{requires}\hfill \\ & \lambda _1^F=|\dot{\lambda }_1^F\ddot{\lambda }_1^F|,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+2,\mathrm{},k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|.\hfill \end{array}$$
Similarly, for the B-type contributions we find
$$\begin{array}{cc}& _{\dot{\lambda }\ddot{\lambda }}^{(\lambda ^I,\lambda ^F)}_{\dot{\nu }\ddot{\nu }}^{(\nu ^I,\lambda ^F)}0\mathrm{requires}\hfill \\ & k^F\lambda _1^F=|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+1,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+3,\mathrm{},k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|1.\hfill \end{array}$$
Second, from (5.3) we read off the action of the combination $`ac`$ and its powers as
$$(ac)^m\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\}=\{(\lambda _1^I,\lambda _1^Fm);(\nu _1^I,\lambda _1^Fm)\},m\lambda _1^F0.$$
Using these results, we can mimic the pattern of (5.3) in the fusion (8.1), establishing that for any allowed $`\lambda ^F`$ the summation over integer parts ($`\lambda ^I`$ and $`\nu ^I`$) gives the same result as for any other allowed $`\lambda ^F`$:
$$\begin{array}{cc}& \underset{\lambda ^I,\nu ^I}{}𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}\hfill \\ & \times \{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|);(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|)\}\hfill \\ & \\ & =\underset{\lambda ^I,\nu ^I}{}_{\dot{\lambda }\ddot{\lambda }}^{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|1)}_{\dot{\nu }\ddot{\nu }}^{(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|1)}\hfill \\ & \times \{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|1);(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|1)\}\hfill \\ & \\ & \mathrm{}\hfill \\ & \\ & =\underset{\lambda ^I,\nu ^I}{}_{\dot{\lambda }\ddot{\lambda }}^{(\lambda _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+1)}_{\dot{\nu }\ddot{\nu }}^{(\nu _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+1)}\{(\lambda _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+1);(\nu _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+1)\}\hfill \\ & \\ & =\underset{\lambda ^I,\nu ^I}{}𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}\{(\lambda _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|);(\nu _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|)\}.\hfill \end{array}$$
From these identities it follows immediately that (8.1) reduces to
$$\begin{array}{cc}\hfill \underset{\lambda ^I,\nu ^I}{}& 𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda _1^I,\lambda _1^F)}𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu _1^I,\lambda _1^F)}\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\},\hfill \\ & \mathrm{for}\mathrm{any}\mathrm{of}\lambda _1^F=|\dot{\lambda }_1^F\ddot{\lambda }_1^F|,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+2,\mathrm{},k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|.\hfill \end{array}$$
Note that the fractional branching conditions in (8.1) still apply to the expressions (8.1) and (8.1). Of course, according to (8.1) we might as well have chosen to represent the final double-summation (8.1) in terms of B-fusion coefficients, with $`\lambda ^F`$ subject to (8.1). Though, this would only be possible provided $`|\dot{\lambda }_1^F\ddot{\lambda }_1^F|+2k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|`$, while (8.1) only requires $`|\dot{\lambda }_1^F\ddot{\lambda }_1^F|k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|`$.
Now, substituting (5.3), the double summation (8.1) becomes
$$\underset{\lambda ^I,\nu ^I}{}𝒩_{\dot{\lambda }^I\ddot{\lambda }^I}^{(k^I)\lambda ^I}𝒩_{\dot{\lambda }^F\ddot{\lambda }^F}^{(k^F)\lambda ^F}𝒩_{\dot{\nu }^I\ddot{\nu }^I}^{(k^I+u)\nu ^I}𝒩_{\dot{\lambda }^F\ddot{\lambda }^F}^{(k^F)\lambda ^F}\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\},$$
with $`\lambda ^F`$ subject to (8.1), in which case the two fusion coefficients at level $`k^F`$ are simply 1 and we reproduce the product form (1.2). This concludes the proof that the “supertrace” reproduces the minimal models fusion rules.
The “trace” part of the proposal (8.1) can actually be proved rather simply: the branching conditions on coset fields rule out the mixed combinations AAB and BAA. Here we have included the A-label for the level-one integrable weights. Consider for instance the AAB situation. The fusion rules imply the parity conditions
$$\begin{array}{cc}\hfill (1)^{\dot{\lambda }_1^I+\ddot{\lambda }_1^I+\lambda _1^I}=(1)^{\dot{\mu }_1^I+\ddot{\mu }_1^I+\mu _1^I}=(1)^{\dot{\nu }_1^I+\ddot{\nu }_1^I+\nu _1^I+k^I+u}=1& ,\hfill \\ \hfill (1)^{\dot{\lambda }_1^F+\ddot{\lambda }_1^F+\lambda _1^F}=(1)^{\dot{\nu }_1^F+\ddot{\nu }_1^F+\nu _1^F+u}=1& ,\hfill \end{array}$$
while the branching conditions require
$$\begin{array}{cc}\hfill (1)^{\dot{\lambda }_1^I+\dot{\mu }_1^I+\dot{\nu }_1^I+\dot{\lambda }_1^F}=(1)^{\ddot{\lambda }_1^I+\ddot{\mu }_1^I+\ddot{\nu }_1^I+\ddot{\lambda }_1^F}=(1)^{\lambda _1^I+\mu _1^I+\nu _1^I+\lambda _1^F}=1& ,\hfill \\ \hfill (1)^{\dot{\lambda }_1^F+\dot{\nu }_1^F}=(1)^{\ddot{\lambda }_1^F+\ddot{\nu }_1^F}=(1)^{\lambda _1^F+\nu _1^F}=1& .\hfill \end{array}$$
Firstly, substitute in the fusion condition $`(1)^{\dot{\mu }_1^I+\ddot{\mu }_1^I}=(1)^{\mu _1^I}`$ the integer branching conditions and compare the two sides using the fusion conditions. From this we obtain $`(1)^{k^I+u}=1`$. Secondly, a comparison of the fractional fusion and branching conditions yields $`(1)^u=1`$. It follows that both $`k^I`$ and $`u`$ must be even (and it is recalled that $`u=k^F+1`$). However, that contradicts
$$k^F\mathrm{odd}k^I\mathrm{odd}$$
$`(8.1)`$
which is a consequence of (2.1) and the fact that $`t`$ and $`u`$ are relatively prime. The analysis of the case BAA is similar.
Therefore, the proposal boils down to the introduction of minus signs in front of the BB-type contributions, cf. (8.1).
We should also stress that neglecting the intermediate weights at level 1 is not an assumption as they can be reinserted everywhere using the branching rule (8.1) and its dotted versions.
Our proposal can also be stated by specifying what must replace the naive factorisation in order to recover the correct coset fusions. From (8.1), we find
$$𝒩_{\dot{\lambda }\ddot{\lambda }}^{(k)\lambda }𝒩_{\dot{\mu }\ddot{\mu }}^{(1)\mu }𝒩_{\dot{\nu }\ddot{\nu }}^{(k+1)\nu }𝒜_{\dot{\lambda }\ddot{\lambda }}^{(k)\lambda }𝒩_{\dot{\mu }\ddot{\mu }}^{(1)\mu }𝒜_{\dot{\nu }\ddot{\nu }}^{(k+1)\nu }_{\dot{\lambda }\ddot{\lambda }}^{(k)\lambda }𝒩_{\dot{\mu }\ddot{\mu }}^{(1)\mu }_{\dot{\nu }\ddot{\nu }}^{(k+1)\nu }.$$
In the $`\widehat{osp}(1|2)`$ analogy, this can be termed a “super-factorisation”.
9. A coset truncation
As an offshot of the supertrace analysis, we identify another way of recovering the coset fusion rules from the singular-vector decoupling method, namely via a truncation.
In (8.1), the fractional part can be chosen among any of the listed, possible values. Take for instance the upper limit. Then, field identifications allow us to write (8.1) as
$$\begin{array}{cc}\hfill \underset{\lambda ^I,\nu ^I}{}𝒜_{\dot{\lambda }\ddot{\lambda }}^{(\lambda _1^I,\lambda _1^F)}& 𝒜_{\dot{\nu }\ddot{\nu }}^{(\nu _1^I,\lambda _1^F)}\{(\lambda _1^I,\lambda _1^F);(\nu _1^I,\lambda _1^F)\}=\underset{\lambda ^I,\nu ^I}{}𝒜_{(\dot{\lambda }_1^I,\dot{\lambda }_1^F),(\ddot{\lambda }_1^I,\ddot{\lambda }_1^F)}^{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}𝒜_{(\dot{\nu }_1^I,\dot{\lambda }_1^F),(\ddot{\nu }_1^I,\ddot{\lambda }_1^F)}^{(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}\hfill \\ & (aca)^{[(\dot{\lambda }_1^F+\ddot{\lambda }_1^F)/u]}\{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|);(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|)\},\hfill \end{array}$$
This clearly resembles the result (4.1) obtained from the Verlinde formula.<sup>7</sup> The precise form of the formula depends upon the specific choice made for $`\lambda _1^F`$. For instance, with
$$\lambda _1^F=k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|2n\mathrm{with}2n=0,2,\mathrm{},k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F||\dot{\lambda }_1^F\ddot{\lambda }_1^F|,$$
$`(9.1)`$ the part $`(aca)^{[(\dot{\lambda }_1^F+\ddot{\lambda }_1^F)/u]}\{\mathrm{}\}`$ would be replaced by
$$(aca)^{[(\dot{\lambda }_1^F+\ddot{\lambda }_1^F)/u]}(ca)^{2n}\{(\lambda _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|2n);(\nu _1^I,k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|2n)\}$$
The upper limit ($`n=0`$) is singled out as the more natural choice when comparing with the Verlinde formula as it is the only case where there are no required field identifications when $`\dot{\lambda }_1^F+\ddot{\lambda }_1^F<u`$. The conclusion concerning associativity is general in that it holds for all values of $`n`$. This is an example of a truncation of the singular-vector decoupling fusions that is sufficient to reproduce the correct coset fusions (see also Section 10).
It should be stressed, however, that truncating a set of fusion rules known to satisfy fundamental, albeit delicate, compatibility requirements such as associativity, is bound to put these properties in peril. Indeed, the truncated fusions presented above are not associative. Technically, this is a reflection of the non-identity
$$\delta _{\nu _1^F,k^F|\mu _1^F|k^F\lambda _1^F\rho _1^F||}\delta _{\nu _1^F,k^F|\lambda _1^F|k^F\mu _1^F\rho _1^F||}$$
This indicates the restricted usefulness of such a truncation.
Nevertheless, this simple analysis provides a second prescription for recovering the coset fusions: they may be expressed in terms of a (non-associative) truncation of the singular-vector decoupling fusions. The truncation amounts to considering only type A fusions and moreover, only with a fixed value of the fractional part of the evaluated weights. The truncation can in most cases also be expressed in terms of type B fusions (cf. discussion following (8.1)).
10. WZW fusion rules: relating the two methods
In this section, we discuss a correspondence between other truncated versions of the fusion obtained by singular-vector decoupling, and the fusion computed using the Verlinde formula. This correspondence relies on the notion of highest- vs lowest-weight condition as applied to a triple product, and is motivated by the note added to .
The decoupling method is based on a study of three-point chiral blocks of the form $`\nu |\varphi _\mu |\lambda `$. In that context, it makes sense to discuss the possible representations carried by the middle field and use those to characterise the associated fusions. Thus, we have the following highest- and lowest-weight conditions (see and )
| | $`\mathrm{hwc}:`$ | $`{\displaystyle \frac{1}{2}}(\lambda _1+\mu _1\nu _1)Z_{},`$ |
| --- | --- | --- |
| | $`\mathrm{lwc}:`$ | $`{\displaystyle \frac{1}{2}}(\lambda _1+\mu _1+\nu _1)Z_{}.`$ |
$`(10.1)`$
A fusion satisfying neither of these two possibilities belongs to the so-called continuous series. Note that the resulting fusion rules (5.1) are symmetric despite the asymmetric starting point $`\nu |\varphi _\mu |\lambda `$ and its associated assignments (10.1).
We shall demonstrate that B-fusions belong to the continuous series. First we rule out the possibility that a B-fusion can satisfy the highest-weight condition. In terms of integer and fractional parts, this condition reads
$$\frac{1}{2}(\lambda _1^I+\mu _1^I\nu _1^I)\frac{1}{2}(k+2)(\lambda _1^F+\mu _1^F\nu _1^F)Z_{}.$$
$`(10.2)`$
It follows immediately that
$$\lambda _1^F+\mu _1^F\nu _1^F=0\mathrm{or}u.$$
$`(10.3)`$
A comparison with the fractional B-fusion yields that $`k^F`$ is odd in the first case ($`\lambda _1^F+\mu _1^F\nu _1^F=0`$) while a contradiction is obtained in the second ($`\lambda _1^F+\mu _1^F\nu _1^F=u`$). To rule out the first case, we consider the integer part from which it follows that $`k^I`$ is even, in contradiction with (8.1).
Second, the lowest-weight condition reads
$$\frac{1}{2}(\lambda _1^I+\mu _1^I+\nu _1^I)\frac{1}{2}(k+2)(\lambda _1^F+\mu _1^F+\nu _1^F)Z_{}.$$
$`(10.4)`$
and as before, there are two possibilities for the fractional part
$$\lambda _1^F+\mu _1^F+\nu _1^F=0\mathrm{or}u.$$
$`(10.5)`$
Again, the first case is ruled out by (8.1), while the second case contradicts the fractional fusion rule itself.
A highest- or lowest-weight fusion can thus only be of type A. But not all type-A fusions respect the highest- or lowest-weight conditions. Again, the highest-weight condition (10.3) is compatible with the A-type fusion rules only if $`\nu _1^F=\lambda _1^F+\mu _1^F`$ (in which case $`\lambda _1^F+\mu _1^F<u`$), while the lowest-weight condition (10.5) is compatible with the A-type fusion rules only if $`\nu _1^F=\lambda _1^F\mu _1^F`$ (in which case $`\mu _1^F\lambda _1^F`$). It is seen that the highest- and lowest-weight fusions correspond to the upper and lower bounds, respectively, of the fractional part of the type-A fusion. The integer parts are not constrained further than by the A-fusion rule.
We note that a fusion can satisfy both the highest- and the lowest-weight conditions. That happens precisely when $`\mu _1^F=0`$ (and $`\lambda _1^F=\nu _1^F`$), in which case the integer part of the type-A fusion ensures that both the conditions in (10.1) are satisfied. We may therefore group the highest- and lowest-weight fusions into two disjoint sets: the highest-weight fusions and the purely lowest-weight fusions, respectively. The latter are thus characterised by $`\mu _1^F=\lambda _1^F\nu _1^F>0`$.
It is natural to symmetrise the notion of lowest-weight fusion:
$$\mathrm{slwc}:\frac{1}{2}(\lambda _1+\mu _1+\nu _1)Z_{}\mathrm{or}\frac{1}{2}(\lambda _1\mu _1+\nu _1)Z_{},$$
$`(10.6)`$
and denote the two cases lowest-weight fusions with respect to $`\mu `$ or $`\lambda `$, respectively. The latter case corresponds to studying the three-point chiral block $`\nu |\varphi _\lambda |\mu `$. It follows immediately that purely $`\lambda `$-lowest-weight fusions are of type A and characterised by $`\nu _1^F=\lambda _1^F+\mu _1^F`$ (in which case $`\lambda _1^F\mu _1^F`$). A symmetrised lowest-weight fusion (10.6) is then characterised by $`\nu _1^F=|\lambda _1^F\mu _1^F|`$, while an associated purely lowest-weight fusion respects in addition $`\lambda _1^F,\mu _1^F1`$. Hence, the symmetrised lowest-weight condition corresponds precisely to the lower bound of the fractional part of the type-A fusion.
We now construct a precise correspondence between the set of highest- or lowest-weight fusions from singular-vector decouplings (fusions known to be of type A) and the Verlinde formula. Highest-weight fusions correspond to positive Verlinde fusions, whereas purely lowest-weight fusions correspond to negative Verlinde fusions. In the first case, the correspondence is obvious as it is precisely the condition $`\lambda _1^F+\mu _1^F<u`$ that ensures the positivity in the Verlinde formula. The integer parts also match.
The correspondence between lowest-weight fusions and negative Verlinde coefficients is less direct. It is established through simple manipulations of the Verlinde fusion rules. The “negative part” of the Verlinde fusion rule (3.1) can be written as (with the understanding that $`\stackrel{~}{\lambda }_1^F+\stackrel{~}{\mu }_1^Fu0`$)
$$(\stackrel{~}{\lambda }_1^I,\stackrel{~}{\lambda }_1^F)\times (\stackrel{~}{\mu }_1^I,\stackrel{~}{\mu }_1^F)\underset{\stackrel{~}{\nu }_1^I\delta _{k^I}}{}(k^I\stackrel{~}{\nu }_1^I,\stackrel{~}{\lambda }_1^F+\stackrel{~}{\mu }_1^Fu).$$
$`(10.7)`$
where the tildes have been introduced to avoid confusion later. The set $`\delta _{k^I}`$ is defined as in (3.2), in terms of $`\stackrel{~}{\lambda },\stackrel{~}{\mu }`$ and $`\stackrel{~}{\nu }`$. Replacing $`\stackrel{~}{\nu }_1^I`$ by $`k^I\stackrel{~}{\nu }_1^I`$ modifies the range $`\delta _{k^I}`$ to
$$\delta _{k^I}^{}=\{\stackrel{~}{\nu }_1^I|\stackrel{~}{\lambda }_1^I+\stackrel{~}{\mu }_1^I\stackrel{~}{\nu }_1^I+k^I2Z_{};|k^I\stackrel{~}{\lambda }_1^I\stackrel{~}{\mu }_1^I|\stackrel{~}{\nu }_1^Ik^I|\stackrel{~}{\lambda }^I\stackrel{~}{\mu }_1^I|\}.$$
$`(10.8)`$
On the other hand, recall that a purely $`\mu `$-lowest-weight fusion respects $`1\mu _1^F\lambda _1^F`$. This means that the sum of the fractional parts of $`\lambda `$ and $`c\mu `$ is $`\lambda _1^F+(u\mu _1^F)u`$. This corresponds to the condition for a negative Verlinde fusion. This suggests that instead of (10.7), we should consider
$$(\lambda _1^I,\lambda _1^F)\times c(\mu _1^I,\mu _1^F)\underset{\nu _1^I\delta _{k^I}^{(c)}}{}(\nu _1^I,\lambda _1^F\mu _1^F).$$
$`(10.9)`$
where
$$\delta _{k^I}^{(c)}=\{\nu _1^I|\lambda _1^I\mu _1^I\nu _1^I+2k^I2Z_{};|\lambda _1^I\mu _1^I|\nu _1^Ik^I|k^I\lambda ^I\mu _1^I|\}.$$
$`(10.10)`$
The conditions in $`\delta _{k^I}^{(c)}`$ imply that $`\lambda _1^I+\mu _1^I+\nu _1^I2Z_{}`$; since $`\nu _1^F=\lambda _1^F\mu _1^F`$, we have all the characteristics of a purely $`\mu `$-lowest-weight fusion (see (10.1)). The analogous correspondence between $`\lambda `$-lowest-weight fusions and the negative Verlinde fusions is based on $`c\lambda \times \mu `$ instead of $`\lambda \times c\mu `$.
The correspondence between highest-weight fusions and the positive Verlinde fusions is obviously one-to-one. This is not the case for the correspondence between one of the two types of purely lowest-weight fusions and the negative Verlinde fusions. A natural way of defining a one-to-one map between the negative Verlinde fusions and the purely lowest-weight fusions is to group the latter in equivalence classes:
$$[\lambda ,\mu ][c\lambda ,c\mu ].$$
$`(10.11)`$
Note that $`c`$ always maps a pure and symmetrised lowest-weight onto a pure and symmetrised lowest-weight (while that is not a property of $`a`$). We also observe that when $`\lambda \mu `$, (10.11) relates a $`\mu `$-lowest-weight fusion to a $`\lambda `$-lowest-weight fusion. When $`\lambda =\mu `$, both fusions are lowest-weight fusions with respect to both $`\mu `$ and $`\lambda `$. The sought one-to-one map is between these equivalence classes of symmetrised, purely lowest-weight fusions and negative Verlinde fusions.<sup>8</sup> Alternatively, one may work with the $`\mu `$-lowest-weight fusions alone, for example. In that case, one must choose a set of representatives for the equivalence classes. When $`\lambda \mu `$, the choice is unique, while for $`\lambda =\mu `$ there are several. Natural choices are either $`\lambda _1^F+\mu _1^Fu`$ or $`\lambda _1^F+\mu _1^Fu`$. Having made such a choice, the correspondence is one-to-one between the accordingly reduced (or truncated) purely $`\mu `$-lowest-weight fusions and the negative Verlinde fusions.
We have thus established a one-to-one correspondence between different truncations of the A-type decoupling fusions and the full set of Verlinde fusions. This clarifies and corrects the corresponding statements made in a note added to .
Finally, we want to comment on yet another truncation of the WZW fusions that reproduces the coset fusions. It combines the notion of highest- and lowest-weight fusions with the result of the previous section.
First we observe that for each possible value of the sum $`\dot{\lambda }_1^F+\ddot{\lambda }_1^F`$, we may choose a representative of (8.1) – see (9.1) and the subsequent footnote. A particular set of choices is reflected in the expression
$$\begin{array}{cc}\hfill \left(1[(\dot{\lambda }_1^F+\ddot{\lambda }_1^F)/u]\right)& \underset{\lambda ^I,\nu ^I}{}𝒜_{(\dot{\lambda }_1^I,\dot{\lambda }_1^F),(\ddot{\lambda }_1^I,\ddot{\lambda }_1^F)}^{(\lambda _1^I,\dot{\lambda }_1^F+\ddot{\lambda }_1^F)}𝒜_{(\dot{\nu }_1^I,\dot{\lambda }_1^F),(\ddot{\nu }_1^I,\ddot{\lambda }_1^F)}^{(\nu _1^I,\dot{\lambda }_1^F+\ddot{\lambda }_1^F)}\{(\lambda _1^I,\dot{\lambda }_1^F+\ddot{\lambda }_1^F);(\nu _1^I,\dot{\lambda }_1^F+\ddot{\lambda }_1^F)\}\hfill \\ \hfill +[(\dot{\lambda }_1^F+\ddot{\lambda }_1^F)/u]& \underset{\lambda ^I,\nu ^I}{}𝒜_{(\dot{\lambda }_1^I,\dot{\lambda }_1^F),(\ddot{\lambda }_1^I,\ddot{\lambda }_1^F)}^{(\lambda _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}𝒜_{(\dot{\nu }_1^I,\dot{\lambda }_1^F),(\ddot{\nu }_1^I,\ddot{\lambda }_1^F)}^{(\nu _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|)}\{(\lambda _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|);(\nu _1^I,|\dot{\lambda }_1^F\ddot{\lambda }_1^F|)\},\hfill \end{array}$$
Note that due to the prefactor of the first term, $`\lambda _1^F=k^F|k^F\dot{\lambda }_1^F\ddot{\lambda }_1^F|`$ is reduced to $`\dot{\lambda }_1^F+\ddot{\lambda }_1^F`$.
Let H and L stand for highest-weight fusions and purely lowest-weight fusions, respectively, while L represents an L subject to $`\dot{\lambda }_1^F+\ddot{\lambda }_1^Fu`$. It is emphasised that L simply corresponds to choosing a particular equivalence class representative. From (10.12) we thus see, that one need only truncate the full set of triple-field products obtained from the separate WZW fusions to those of type HHH or LHL, where the three labels refer to the representations at respective levels $`k`$, $`1`$ and $`k+1`$ (the middle field is associated to a highest-weight fusion H, as L denote a purely lowest-weight fusion, or more precisely, an equivalence class representative). It is also natural to end up with a requirement specifying that in coset fields, highest- (respectively, lowest-) weight fusions are to be combined with highest- (respectively, lowest-) weight fusions. Note, however, that only a subset of the lowest-weight fusions are used.
Incidentally, restricting to the set of symmetrised (but not necessarily purely) lowest-weight fusions, likewise reproduces the coset fusions. This situation occurs when choosing $`\lambda _1^F=|\dot{\lambda }_1^F\ddot{\lambda }_1^F|`$ in (8.1).
11. Conclusion
Our results aim to fill an obvious gap in the coset description of non-unitary rational CFTs. Such coset CFTs involve WZW models at fractional, admissible levels. Their fusion rules are ambiguous: Verlinde fusion differs from (singular-vector) decoupling fusion. Only the Verlinde fusion rules were known to be consistent with the unambiguous coset fusion, in the simplest case of diagonal cosets involving one integer level. Here we have shown how the decoupling fusion can yield the correct fusion in the Virasoro minimal models, described by the diagonal coset (1.1). We restricted to this coset for simplicity, and because decoupling fusion has only been worked out completely in the $`\widehat{su}(2)`$ case.<sup>9</sup> Complete results for certain levels are now known for $`\widehat{su}(3)`$ .
We should emphasise again that WZW models at fractional levels may not be consistent, except in coset theories. After all, the Verlinde fusion coefficients are sometimes negative integers. These negative signs do cancel in the diagonal coset CFTs, however. The key input is the factorisation (4.1), a consequence of the factorisation of the corresponding coset modular $`S`$ matrix.
We find here that although the decoupling fusion coefficients are non-negative, they are too big for the cosets. That is, they lead to coset fusion coefficients that are too large, when factorisation is assumed. By introducing minus signs at the coset level, we found a prescription that led to the correct coset fusions.
These minus signs cause cancellations, and so are equivalent to a truncation. We also found direct descriptions for such truncations. Particular attention was paid to a truncation motivated by the notion of highest and lowest weights in the derivation of the singular-vector decoupling fusions.
The first prescription is partly motivated by its interpretation as a “super-trace” or “super-factorisation”, in the $`\widehat{osp}(1|2)`$ analogy found previously. The second is natural in the coset description, since it pairs highest weights with highest weights, and lowest weights with lowest weights.
Of course, a deeper motivation should be found for our recipes. Perhaps the relation between the singular vectors of algebra and subalgebra can provide it. We leave that to future work.
Acknowledgements
Part of this work was done during the workshop Quantum integrability at the CRM, Université de Montréal. We thank the CRM for its generous hospitality.
REFERENCES
relax1.P. Mathieu and M.A. Walton, Prog. Theor. Phys., suppl. 102 (1990) 229.
relax2.H. Awata and Y. Yamada, Mod. Phys. Lett. A7 (1992) 1185.
relax3.A.A. Belavin, A.M. Polyakov and A.B. Zamolodchikov, Nucl. Phys. B241 (1984) 333.
relax4.D. Gepner, Nucl. Phys. B287 (1987) 111.
relax5.A. Kent, Ph.D. thesis (Cambridge University, 1986).
relax6.P. Mathieu, D. Sénéchal and M.A. Walton, Int. J. Mod. Phys. A7 (1992) 731.
relax7.P. Mathieu and M.A. Walton, Nucl. Phys. B553 (1999) 533.
relax8.I.G. Koh and P. Sorba, Phys. Lett. B215 (1988) 723.
relax9.B. Feigen and F. Malikov, Modular functor and representation theory of $`\widehat{sl}_2`$ at fractional level, q-alg/9511011.
relax10.V. Kac and M. Wakimoto, Proc. Nat. Acad. Sci. USA 85 (1988) 4956, Adv. Ser. Math. Phys. 7 (World Scientific, 1988) p.138.
relax11.D. Bernard and G. Felder, Commun. Math. Phys. 127 (1990) 145.
relax12.C. Dong, H. Li and G. Mason, Commun. Math. Phys. 184 (1997) 65.
relax13.A.B. Zamolodchikov and V.A. Fateev, Sov. J. Nucl. Phys. 43 (1986) 657.
relax14.J-B. Fan and M. Yu, Modules over affine Lie superalgebras, hep-th/9304122; I.P. Ennes, A.V. Ramallo and J.M. Sanchez de Santos, Nucl. Phys. B491 (1997) 574.
relax15.J.L. Petersen, J. Rasmussen and M. Yu, Nucl. Phys. B457 (1995) 309, Nucl. Phys. B481 (1996) 577; J. Rasmussen, Ph.D. thesis (Niels Bohr Institute, 1996) hep-th/9610167.
relax16.P. Furlan, A.Ch. Ganchev and V.B. Petkova, Nucl. Phys. B491 (1997) 635.
relax17.J.L. Petersen, J. Rasmussen and M. Yu, Nucl. Phys. B502 (1997) 649; J. Rasmussen, Mod. Phys. Lett. A13 (1998) 1281, Int. J. Mod. Phys. A14 (1999) 1225.
relax18.P. Di Francesco, P. Mathieu and D. Sénéchal, Conformal field theory (Springer, 1997).
relax19.P. Furlan, A.Ch. Ganchev and V.B. Petkova, Nucl. Phys. B518 (1998) 645, Commun. Math. Phys. 202 (1999) 701; A.Ch. Ganchev, V.B. Petkova and G. Watts, Nucl. Phys. B571 (2000) 457.
|
warning/0007/hep-ph0007264.html
|
ar5iv
|
text
|
# High Energy Nucleus-Nucleus Collisions In View Of The Multi-Peripheral Model
## 1 Introduction
In the last few years the research work was concentrated on possible existence of the quark-gluon plasma phase, considering of unconfined quarks and gluons at high temperature or high density. In the laboratory, nucleus-nucleus collisions at very high energies provide a promising way to produce high temperature or high-density matter. As estimated by Bajorken the energy density can be so high that these reactions might be utilized to explore the existence of the quark-gluon plasma. One of the many factors that lead to an optimistic assessment that matter at high density and high temperature may be produced with nucleus-nucleus collisions is the occurrence of multiple collisions. By this mean, a nucleon of one nucleus may collide with many nucleons in the other and deposits a large amount of energy in the collision region. In the nucleon-nucleon center of mass system, the longitudinal inter-nucleon spacing between target nucleons are Lorentz contracted and can be smaller than 1 fm in high-energy collisions. On the other hand, particle production occurs only when a minimum distance of about 1 fm separates the leading quark and antiquark in the nucleon-nucleon center of mass system . Therefore when the projectile nucleon collides with many target nucleons, particles production arising from the first N-N collision is not finished before the collision of the projectile with another target nucleon begins. There are models \[4-11\] that describe how is the second collision is affected by the first one. Nevertheless, the fundamental theory of doing that remains one of the unsolved problems. Experimental data suggest that after the projectile nucleon makes a collision, the projectile-like object that emerges from the first collision appears to continue to collide with other nucleons in the target nucleus on its way through the target nucleus. In each collision the object that emerges along the projectile nucleon direction has a net baryon number of unity because of the conservation of the baryon number. One can speak loosely of this object as the projectile or baryon-like object and can describe the multiple collision process in terms of the projectile nucleon making many collisions with the target nucleons. Then, losing energy and momentum in the process, and emerging from the other side of the target with a much diminished energy. The number of collisions depends on the thickness of the target nucleus. Experimental evidence of occurrence of multiple collision process can be best illustrated with the data of p-A reactions in the projectile fragmentation region . In the present work we shall investigate the particle production mechanism in heavy ion collision by extending the multi-peripheral model \[13-15\] to the nucleon-Nucleus and then generalize to the nucleus-nucleus case. The paper is organized so that in section 2, we present the experimental features of high-energy events. This will serve in setting the boundary condition of the extended problem and building up the model in a realistic form. The hypothesis of the multi-peripheral model is demonstrated in section 3. The nucleon-nucleon case is briefly described in 3.1 and the extension to the nucleon-nucleus and the nucleus-nucleus collisions are followed in section 3.2 and 3.3 respectively. Concluding remarks are given in section 4.
## 2 Experimental Features of High Energy Events
The heavy ion collisions initiate the chance of processing multiple collisions of single nucleon-nucleon (N-N) events. This behavior allows the magnification of any small signal that may appear unclear in (N-N) event. The magnification increases as the size of the interacting nuclei increases. Nevertheless, the relation is not simply linear but many other factors are expected to contribute the complexity of the problem. The study of the multiplicity distribution at different rapidity intervals carries valuable information in this sense . The rapidity range is classified into regions that characterizes the type of the interaction. The target and the projectile fragmentation regions contain the fast particles characterizing the forward and backward production from the fragmentation of both the target and the projectile nuclei in their center of mass system. The central rapidity interval is the hot region of the reaction. It sends global information about the strong interactions inside the nuclear bowl. Moreover, signals about (QGP) may be extracted from the study of this region. Fig. (1) shows the experimental data of the rapidity distributions of fast particles produced in the interaction of sulfur-projectile with proton, sulfur, silver and gold as target nuclei at 200 A GeV incident energy . The average multiplicity at each rapidity interval depends on the incident projectile energy (which is fixed in this case) and the size of the interacting nuclei. The nucleus-nucleus (A-A) collision may be seen as N-N base or N-A base. To put the problem in a quantitative measurable form, we define a multiplication factor MF (N\_N) as a base of (N-N);
$$MF_{(NN)}=\frac{Yield,in(AA)}{Yield,in(NN)}$$
(1)
And the multiplication factor MF (N-A) as a base of (N-A);
$$MF_{(NA)}=\frac{Yield,in(AA)}{Yield,in(NA)}$$
(2)
Fig. (2) and Fig. (3) show the multiplication factors based on (N-N) and (N-A) for the different reactions. The multiplication factor has maximum value at the central rapidity region (Y~0) in both cases. The experimental data of p-S and the S-S collisions show constant value allover the rapidity range. Nevertheless, it is difficult to estimate the number of base-collisions in each reaction from this approach. An alternative approach is to renormalize the rapidity distribution to get a scaled function that is independent on the projectile and target size. The average number of encountered nucleons from the target by an incident hadron is a good measure of the number of collisions inside a target nucleus. This is defined as,
$$\nu =A_t\frac{\sigma _{hN}}{\sigma _{hA}}$$
(3)
Where $`\sigma _{hN}`$, and $`\sigma _{hA}`$ are the total inelastic cross section of (h-N) and (N-A) respectively. By analogy, the average number of binary (N-N) collisions inside the (A-A) collision may be defined as,
$$\nu _{NN}=A_pA_t\frac{\sigma _{hN}}{\sigma _{AA}}$$
(4)
$`\sigma _{AA}`$ is the total inelastic cross section of (A-A) collision at the same incident energy. On the other hand, the average number of collisions based to N-A is defined as,
$$\nu _{NA}=A_p\frac{\sigma _{hAt}}{\sigma _{AA}}$$
(5)
is total inelastic cross section of (h-At) collision. The scaling function is then obtained by dividing the rapidity distribution by the average number of collisions or . The resultant are displayed in Fig. (4) which shows the overlapped distributions corresponding to the reactions of different nuclear sizes. Following now a geometric approach which assumes that the nuclear radius is linearly proportional to the one third power of its mass number, then Eqs.( 3-5) becomes,
$$\nu =A_t^{1/3}$$
$$\nu _{NN}=\frac{A_pA_t}{A_t^{2/3}+A_p^{2/3}+2A_t^{1/3}A_p^{1/3}}$$
(6)
$$\nu _{NA}=\frac{A_pA_t^{2/3}}{A_t^{2/3}+A_p^{2/3}+2A_t^{1/3}A_p^{1/3}}$$
Let $`\overline{m}`$ be the average multiplicity produced in N-N collision and $`\overline{m}_{NA}`$ is the corresponding figure for N-A collision at the same incident energy. So that the average multiplicity produced in A-A collisions may be calculated in base of N-N as,
$$\overline{m}_{AA}=\nu _{NN}\overline{m}$$
(7)
and in base of N-A as,
$$\overline{m}_{AA}=\nu _{NA}\overline{m}_{NA}$$
(8)
Table (1) shows the values of $`\overline{m}_{AA}`$ as calculated by Eqs.(7 & 8) compared with the experimental data. The comparison of the last three columns shows that the predicted values of $`\overline{m}\nu _{NN}`$ are more close to the experimental data which implies that the A-A collision behaves as a base of N-N not N-A collision.
Table (1)
| Reaction | $`\nu _{NN}`$ | $`\nu _{NA}`$ | $`\overline{m}\nu _{NN}`$ | $`\overline{m}_{NA}`$ $`\nu _{NA}`$ | $`\overline{m}_{AA}(\mathrm{exp})`$ |
| --- | --- | --- | --- | --- | --- |
| p-S | 1.83 | 0.25 | 5.87 | 1.48 | 5.90 |
| p-Au | 4.20 | 0.42 | 13.5 | 4.14 | 9.90 |
| d-Au | 7.8 | 0.83 | 25.2 | 11.3 | 23.0 |
| O-Au | 45.3 | 6.70 | 145.0 | 90.0 | 137.0 |
| S-S | 25.4 | 8.0 | 81.3 | 47.2 | 90.0 |
| S-Ag | 54.8 | 27.0 | 175.0 | 160.0 | 162.0 |
| S-Au | 77.9 | 49.2 | 250.0 | 290.0 | 225.0 |
## 3 The Multi-Peripheral Model
### 3.1 Hadron-Nucleon Collision
The multi-peripheral technique was found to be a good tool in dealing with the problems of multi particle production in hadron proton (h-p) interactions. In this technique, the many body-system is expanded into subsystems, each concerns a two body collision. It is assumed that each hadron in the final state is produced at a specific peripheral surface that is characterized by a peripheral parameter. The phase space integral $`I_n(s)`$ of the produced hadrons is a measure of the probability of producing n particle in the final state at center of mass energy $`\sqrt{s}`$. It depends mainly on the volume in phase space and the transition matrix element $`T`$, and defined as,
$$I_n(s)=\mathrm{}\underset{i}{\overset{n}{}}\frac{d^3p_i}{2E_i}\delta ^4(\sqrt{s}\underset{j}{}p_j)|T|^2$$
(9)
Eq. (9) may be simplified as if expressed as a sequence of two particle decay Fig.(5). Accordingly, the integration in Eq. (9) becomes,
$$I_n(s)=_{\mu _{n1}^2}^{(M_nm_n)^2}𝑑M_{n1}^2I_2(k_n^2,k_{n1}^2,p_n^2)I_{n1}(M_{n1}^2)|T|^2$$
(10)
$$=_{\mu _{n1}^2}^{(M_nm_n)^2}𝑑M_{n1}^2𝑑\mathrm{\Omega }_{n1}\frac{\lambda ^{1/2}(M_n^2,M_{n1}^2,m_n^2}{8M_n^2}I_{n1}(M_{n1}^2)|T|^2$$
Where $`M_i`$ is the invariant mass of the subsystem composed of $`i`$particles. The integration over all possible values of $`M_{n1}`$ concerns the first vertex in the chain of Fig.(5). To proceed further, we iterate Eq.(10) for the $`M_{n2},M_{n3},M_2`$ to obtain the entire chain, so that,
$$I_n(s)=_{\mu _{n1}}^{(M_nm_n)}𝑑M_{n1}𝑑\mathrm{\Omega }_{n1}\frac{1}{2}P_{n1}|T(P_{n1})|^2\mathrm{}_{\mu _2}^{(M_3m_3)}𝑑M_2𝑑\mathrm{\Omega }_2\frac{1}{2}P_3|T(P_2)|^2$$
$$d\mathrm{\Omega }_1\frac{1}{2}P_2|T(P_1)|^2$$
(11)
Where $`P_i=\lambda ^{1/2}(M_i^2,M_{i1}^2,m_i^2)`$ is the three-vector momentum of the ith particle and $`T(p_i)`$ is the transition matrix element at the i<sup>th</sup> vertex. For the case of strong interactions $`T(p_i)`$ has a parametric form which cut the magnitude of the particles momenta according to their physical region ,
$$T(p_i)=\mathrm{exp}(\alpha _ip_i)$$
(12)
And $`\alpha _i`$ is a peripheral parameter, which fits the experimental data. The multiple integration in Eq. (11) may be solved by the Monte Carlo technique \[20-22\]. At extremely high energy, Eq. (11) has an asymptotic limit in the form;
$$I_n(s)=\frac{(\pi /2)^{n1}}{(n1)!(n2)!}s^{n2}|T|^{2(n1)}$$
(13)
### 3.2 Hadron-Nucleus Collisions
On extending the model to the hadron-nucleus or nucleus-nucleus collisions, we follow the NN-base super position as expected from the features of the experimental data that are illustrated in section 2. We should consider the possible interactions with the nucleons forming the target nucleus $`A_t`$. The incident hadron makes successive collisions inside the target. The energy of the incident hadron (leading particle) slows down after each collision, producing number of created hadrons each time that depends on the available energy. The phase space integral $`I_n^{NA}`$ in this case has the form,
$$I_n^{NA}(s)=\underset{\nu }{\overset{A_t}{}}I_{n_\nu }(s_\nu )P(\nu ,A_t)\delta (n\underset{i}{\overset{A_t}{}}n_i)$$
(14)
Where $`P(\nu ,A_t)`$ is the probability that $`\nu `$ nucleons out of $`A_t`$ will interact with the leading particle and $`I_{n_\nu }(s_\nu )`$ is the phase space integral of $`NN`$ collision that produces hadrons at energy $`\sqrt{s_\nu }`$. The delta function in Eq.(14) is to conserve the number of particle in the final state. Treating all nucleons identically, and that $`X_{NN}`$ is the $`NN`$ phase shift function \[18,23-25\], then, according to the eikonal approximation,
$$P(l,A_t)=\left(\begin{array}{c}A_t\\ l\end{array}\right)\underset{j=0}{\overset{l}{}}(1)^j\left(\begin{array}{c}l\\ j\end{array}\right)\{1\mathrm{exp}(2Rei(A_tl+j)X_{NN})\}$$
(15)
The working out of this approach is to put the multi-dimension integration of Eq. (11) and the generated kinematical variables into a Monte Carlo subroutine . This in turn is restored $`\nu `$ times, where $`\nu `$ is the number of collisions inside the target nucleus that is generated by a Monte Carlo Generator according to the probability distribution Eq.(15). In the first one, the incident hadron has its own incident energy $`E_0`$ and moves parallel to the collision axis (z-axis) $`\theta _0=0`$. The output of the subroutine determines the number of created hadrons n1 as well as the energy $`E_1(<E_0)`$ and the direction q1 of the leading particle. The leading particle leads the reaction in its second round with $`E_1`$ and $`\theta _1`$ as input parameters and creates new number of particles $`n2`$ and so on. The number nj is determined according to a multiplicity generator which depends on the square of the center of mass energy sj in the round number j.
$$s_j=2m_N^2+2m_NE_j$$
(16)
Fig. (6) demonstrates the particle rapidity distribution produced in the first three- collisions as predicted by the model for the p- <sup>32</sup>S at 200 GeV. The yield (as measured by the area under the curve) decreases with the order of the collision because of the appreciable drop in the energy after the successive collisions. On the other hand, the family of curves concerning the different number of collisions acquires gradually decreased rapidity range. The overall distribution is compared with the experimental data SLAC-NA-035 in Fig. (7). The comparison shows good agreement.
### 3.3 Nucleus-Nucleus Collisions
The extension of the multi peripheral model to the nucleus- nucleus case is more complicated. The number of available collisions is multi-folded due to the contribution of the projectile nucleons. By analogy to the N-A collision, it is possible to define the phase space integral $`I_n^{AA}`$ in $`AA`$ collisions as,
$$I_n^{AA}(s)=\underset{j}{\overset{A_p}{}}\underset{k}{\overset{A_t}{}}I_{n_{j,k}}(s_{j,k})P_{AA}(j,A_p,k,A_t)\delta (n\underset{j,k}{\overset{A_t}{}}n_{j,k})$$
(17)
Where $`I_{n_{j,k}}(s_{j,k})`$ is the phase space integral due to the knocked on nucleon number $`j`$ from the projectile and that, number $`k`$ from the target. The probability that the $`AA`$ collision encounters $`\nu _p`$ collisions from the projectile and $`\nu _t`$ collisions from the target is treated as independent events. So that,
$$P_{AA}(j,A_p,k,A_t)=P(\nu _p,A_p)P(\nu _t,A_t)$$
(18)
Another modification is carried out on this calculation that is to consider the screening effect of the projectile nucleons upon the interaction. This effect is summarized as follows. The first projectile nucleon faces the target nucleus as a whole, i.e. it can see the complete $`A_t`$ nucleons and interacts with only $`\nu _{t1}`$ of them. The next projectile nucleon will see that target as partially screened by the first. The target size in this case is $`A_t\nu _{t1}`$ and it interacts with only $`\nu _{t2}`$ nucleons according to a probability function, $`P(\nu _{t2},A_t\nu _{t1})`$ , by simple iteration, the i<sup>th</sup> projectile nucleon will see a target as $`A_t_k^{i1}\nu _{tk}`$and so on. The prediction of the model is applied to the <sup>32</sup>S-<sup>32</sup>S collisions at 200 A GeV incident energy. The rapidity distribution of the produced particles in the first 3- projectile nock on nucleons is demonstrated in Fig. (8). While the overall rapidity distribution is compared with the data of the experiment SLAC-NA-035 in Fig. (9). The fair agreement obtained for $`NA,`$ Fig. (7) and $`AA`$ collisions Fig. (9) supports the success of the multi-peripheral model. A small defect in the model for the $`NN`$ case will be multi-folded and would be seen clearly in $`NA`$ and $`AA`$ collisions.
## 4 Summary and Conclusive Remarks
1- A scaling rapidity function is obtained for particles produced in h-N, $`AA`$ and $`AA`$ collisions, which assumes that the reaction is built on $`NN`$ base.
2- The multi-peripheral model is extended to the nucleon-nucleus and the nucleus-nucleus interactions on bases of nucleon-nucleon collisions.
3- The phase space integral of the nucleon-nucleon collision is folded several times according to the number of encountered nucleons from the target.
4- The probability that $`\nu `$ nucleons from the target are encountered by a projectile nucleon is calculated in terms of the nucleon-nucleon phase shift according to the eikonal approximation.
5- The number of created particles in each collision is summed over to get the production in the nucleon-nucleus case, where the conservation of number of particles in the final state is taken into consideration.
6- In nucleus-nucleus collisions, we followed the statistics of independent events. The screening effect among the interacting projectile nucleons is also considered.
References
Figure Captions
Fig. (1) The rapidity distribution of particles produced in nucleus collisions at the same incident energy.
Fig. (2) The multiplicity multiplication factor in nucleus-nucleus collisions in terms of the yield in pp collision at the same incident energy. The p-S collision is represented by the black- plus line, S-S (crossed), S- Ag (triangle) and S-Au (circles).
Fig. (3) The multiplicity multiplication factor in nucleus-nucleus collisions in terms of the yield in p-S collision at the same incident energy. The black- plus line represents the S-S collision, S-Ag (crossed) and S-Au (triangles).
Fig. (4) The scaled rapidity distribution produced from the nucleus- nucleus collisions at the same incident energy.
Fig. (5) A tree diagram for the multi-peripheral process, producing n- particles in the final state.
Fig. (6) The particle rapidity distribution produced in p-S at multiple order collisions as predicted by the MPM.
Fig. (7) The particle rapidity distribution produced in p-S collision at 200 GeV incident proton energy as predicted by the MPM and compared with the experimental data SLAC-NA-035.
Fig. (8) The particle rapidity distribution produced in S-S at multiple order collisions as predicted by the MPM.
Fig. (9) The particle rapidity distribution produced in S-S collision at 200 GeV incident proton energy as predicted by the MPM and compared with the experimental data SLAC-NA-035.
|
warning/0007/cond-mat0007114.html
|
ar5iv
|
text
|
# Library Design in Combinatorial Chemistry by Monte Carlo Methods
## I Introduction
The goal of combinatorial materials discovery is to find compositions of matter that maximize a specific material property , such as superconductivity , magnetoresistance , luminescence , ligand specificity , sensor response , or catalytic activity . This problem can be reformulated as one of searching a multi-dimensional space, with the material composition, impurity levels, and synthesis conditions as variables. The property to be optimized, the figure of merit, is generally an unknown function of the variables and can be measured only experimentally.
Present approaches to combinatorial library design and screening invariably perform a grid search in composition space, followed by a “steepest-ascent” maximization of the figure of merit. This procedure becomes inefficient in high-dimensional spaces or when the figure of merit is not a smooth function of the variables, and its use has limited most combinatorial chemistry experiments to ternary or quaternary compounds.
In this paper, we suggest new experimental protocols for searching the space of variables in combinatorial chemistry, exploiting an analogy between combinatorial materials discovery and Monte Carlo computer modeling methods. In Section II we discuss several of these strategies for library design and redesign. In Section III we introduce the Random Phase Volume Model that we will use to compare the different methods. The effectiveness of different strategies is discussed in Section IV. We conclude in Section V.
## II Sampling the Space of Variables in Materials Discovery
Several variables can be manipulated in order to seek the material with the optimal figure of merit. Material composition is certainly a variable. But also, film thickness and deposition method are variables for materials made in thin film form. The processing history, such as temperature, pressure, pH, and atmospheric composition, is a variable. The guest composition or impurity level can greatly affect the figure of merit . In addition, the “crystallinity” of the material can affect the observed figure of merit . Finally, the method of nucleation or synthesis may affect the phase or morphology of the material and so affect the figure of merit .
We assume that the composition and non-composition variables of each sample can be changed independently . Then, instead of a grid search on the composition and non-composition variables, we consider choosing the variables *at random* from the allowed values. We also consider choosing the variables in a fashion that attempts to maximize the amount of information gained from the limited number of samples screened, via a quasi-random, low-discrepancy sequence .
We further consider performing multiple rounds of screening, incorporating feedback as the experiment proceeds by treating the combinatorial chemistry experiment as a Monte Carlo in the laboratory. This leads to sampling the experimental figure of merit, $`E`$, proportional to $`\mathrm{exp}(\beta E)`$. If $`\beta `$ is large, then the Monte Carlo procedure will seek out values of the composition and non-composition variables that maximize the figure of merit. If $`\beta `$ is too large, however, the Monte Carlo procedure will get stuck in relatively low-lying local maxima. The first round is initiated by choosing the composition and non-composition variables at random from the allowed values. The variables are changed in succeeding rounds as dictated by the Monte Carlo procedure.
Two ways of changing the variables are considered: randomly changing the variables of a randomly chosen sample a small amount and exchanging a subset of the variables between two randomly chosen samples. These moves are repeated until all the samples in a round have been modified. The values of the figure of merit for the proposed new samples are then measured. Whether to accept the newly proposed samples or to keep the current samples for the next round is decided according to the detailed balance acceptance criterion. For the random change of one sample, we find the Metropolis acceptance probability:
$$\mathrm{acc}(cp)=\mathrm{min}\{1,\mathrm{exp}\left[\beta \left(E_{\mathrm{proposed}}E_{\mathrm{current}}\right)\right]\}.$$
(1)
Proposed samples that increase the figure of merit are always accepted; proposed samples that decrease the figure of merit are accepted with the Metropolis probability. Allowing the figure of merit occasionally to decrease is what allows samples to escape from local maxima. The random displacement of the $`d`$ mole fraction variables, $`x_i`$, is done in the $`(d1)`$-dimensional subspace orthogonal to the $`d`$-dimensional vector $`(1,1,\mathrm{},1)`$. This procedure ensures that the constraint $`_{i=1}^dx_i=1`$ is maintained. This subspace is identified by the Gram-Schmidt procedure. Moves that violate the constraint $`x_i0`$ are rejected. Moves that lead to invalid values of the non-composition variables are also rejected. For the swapping move applied to samples $`i`$ and $`j`$, we find the modified acceptance probability:
$`\mathrm{acc}(cp)`$ $`=`$ $`\mathrm{min}\{1,\mathrm{exp}[\beta (E_{\mathrm{proposed}}^i+E_{\mathrm{proposed}}^j`$ (3)
$`E_{\mathrm{current}}^iE_{\mathrm{current}}^j)]\}.`$
Fig. 1a shows one round of a Monte Carlo procedure. The parameter $`\beta `$ is not related to the thermodynamic temperature of the experiment and should be optimized for best efficiency. The characteristic sizes of the random changes in the composition and non-composition variables are also parameters that should be optimized.
If the number of composition and non-composition variables is too great, or if the figure of merit changes with the variables in a too-rough fashion, normal Monte Carlo will not achieve effective sampling. Parallel tempering is a natural extension of Monte Carlo that is used to study statistical , spin glass , and molecular systems with rugged energy landscapes. Our most powerful protocol incorporates the method of parallel tempering for changing the system variables. In parallel tempering, a fraction of the samples are updated by Monte Carlo with parameter $`\beta _1`$, a fraction by Monte Carlo with parameter $`\beta _2`$, and so on. At the end of each round, samples are randomly exchanged between the groups with different $`\beta `$’s, as shown in Fig. 1b. The acceptance probability for exchanging two samples is
$$\mathrm{acc}(cp)=\mathrm{min}\{1,\mathrm{exp}\left[\mathrm{\Delta }\beta \mathrm{\Delta }E\right]\},$$
(4)
where $`\mathrm{\Delta }\beta `$ is the difference in the values of $`\beta `$ between the two groups, and $`\mathrm{\Delta }E`$ is the difference in the figures of merit between the two samples. It is important to notice that this exchange step does not involve any extra screening compared to Monte Carlo and is, therefore, “free” in terms of experimental costs. This step is, however, dramatically effective at facilitating the protocol to escape from local maxima. The number of different systems and the temperatures of each system are parameters that must be optimized.
To summarize, the first round of combinatorial chemistry consists of the following steps: constructing the initial library of samples, measuring the initial figures of merit, changing the variables of each sample a small random amount or swapping subsets of the variables between pairs of samples, constructing the proposed new library of samples, measuring the figures of merit of the proposed new samples, accepting or rejecting each of the proposed new samples, and performing parallel tempering exchanges. Following rounds of combinatorial chemistry repeat these steps, starting with making changes to the current values of the composition and non-composition variables. These steps are repeated for as many rounds as desired, or until maximal figures of merit are found.
We have chosen to sample the figure of merit by Monte Carlo, rather than to optimize it globally by some other method, for several reasons. First, Monte Carlo is an effective stochastic optimization method. Second, simple global optimization may be misleading since concerns such as patentability, cost of materials, and ease of synthesis are not usually included in the experimental figure of merit. Moreover, the screen that is most easily performed in the laboratory, the “primary screen,” is usually only roughly correlated with the true figure of merit. Indeed, after finding materials that look promising based upon the primary screen, experimental secondary and tertiary screens are usually performed to identify that material which is truly optimal. Third, it might be advantageous to screen for several figures of merit at once. For all of these reasons, sampling by Monte Carlo to produce several candidate materials is preferred over global optimization.
## III The Random Phase Volume Model
The effectiveness of these protocols is demonstrated by combinatorial chemistry experiments as simulated by the Random Phase Volume Model. The Random Phase Volume Model is not fundamental to the protocols; it is introduced as a simple way to test, parameterize, and validate the various searching methods. The model relates the figure of merit to the composition and non-composition variables in a statistical way. The model is fast enough to allow for validation of the proposed searching methods on an enormous number of samples, yet possesses the correct statistics for the figure-of-merit landscape. The $`d`$-dimensional vector of composition mole fractions is denoted by $`𝐱`$. The composition mole fractions are non-negative and sum to unity, and so the allowed compositions are constrained to lie within a simplex in $`d`$ dimensions. For the familiar ternary system, this simplex is an equilateral triangle. The composition variables are grouped into phases centered around $`N_x`$ points $`𝐱_\alpha `$ randomly placed within the allowed composition range (the phases form a Voronoi diagram , see Fig. 2). The model is defined for any number of composition variables, and the number of phase points is defined by requiring the average spacing between phase points to be $`\xi =0.25`$. To avoid edge effects, additional points are added in a belt of width $`2\xi `$ around the simplex of allowed compositions. The figure of merit should change dramatically between composition phases. Moreover, within each phase $`\alpha `$, the figure of merit should also vary with $`𝐲=𝐱𝐱_\alpha `$ due to crystallinity effects such as crystallite size, intergrowths, defects, and faulting . In addition, the non-composition variables should also affect the measured figure of merit. The non-composition variables are denoted by the $`b`$-dimensional vector $`𝐳`$, with each component constrained to fall within the range $`[1,1]`$ without loss of generality. There can be any number of non-composition variables. The figure of merit depends on the composition and non-composition variables in a correlated fashion, and so the non-composition variables also fall within $`N_z`$$`z`$-phases” defined in the space of composition variables. There are a factor of 10 fewer non-composition phases than composition phases. The functional form of the model when $`𝐱`$ is in composition phase $`\alpha `$ and non-composition-phase $`\gamma `$ is
$`E(𝐱,𝐳)`$ $`=`$ $`U_\alpha `$ (5)
$`+\sigma _x{\displaystyle \underset{k=1}{\overset{q}{}}}`$ $`{\displaystyle \underset{i_1\mathrm{}i_k=1}{\overset{d}{}}}f_{i_1\mathrm{}i_k}\xi _x^kA_{i_1\mathrm{}i_k}^{(\alpha k)}y_{i_1}y_{i_2}\mathrm{}y_{i_k}`$ (6)
$`+{\displaystyle \frac{1}{2}}(W_\gamma `$ (7)
$`+\sigma _z{\displaystyle \underset{k=1}{\overset{q}{}}}`$ $`{\displaystyle \underset{i_1\mathrm{}i_k=1}{\overset{b}{}}}f_{i_1\mathrm{}i_k}\xi _z^kB_{i_1\mathrm{}i_k}^{(\gamma k)}z_{i_1}z_{i_2}\mathrm{}z_{i_k}),`$ (8)
where $`f_{i_1\mathrm{}i_k}`$ is a constant symmetry factor, $`\xi _x`$ and $`\xi _z`$ are constant scale factors, and $`U_\alpha `$, $`W_\gamma `$, $`A_{i_1\mathrm{}i_k}^{(\alpha k)}`$, and $`B_{i_1\mathrm{}i_k}^{(\gamma k)}`$ are random Gaussian variables with unit variance. In more detail, the symmetry factor is given by
$$f_{i_1\mathrm{}i_k}=\frac{k!}{_{i=1}^lo_i!},$$
(9)
where $`l`$ is the number of distinct integer values in the set $`\{i_1,\mathrm{},i_k\}`$, and $`o_i`$ is the number of times that distinct value $`i`$ is repeated in the set. Note that $`1lk`$ and $`_{i=1}^lo_i=k`$. The scale factors are chosen so that each term in the multinomial contributes roughly the same amount: $`\xi _x=\xi /2`$ and $`\xi _z=(z^6/z^2)^{1/4}=(3/7)^{1/4}`$. The $`\sigma _x`$ and $`\sigma _z`$ are chosen so that the multinomial, crystallinity terms contribute 40% as much as the constant, phase terms on average. For both multinomials $`q=6`$. As Fig. 2 shows, the Random Phase Volume Model describes a rugged figure of merit landscape, with subtle variations, local maxima, and discontinuous boundaries.
## IV Results
Six different search protocols are tested with increasing numbers of composition and non-composition variables. The total number of samples whose figure of merit will be measured is fixed at $`M=100,000`$, so that all protocols have the same experimental cost. The single pass protocols Grid, Random, and LDS are considered. For the Grid method, we define $`M_x=M^{(d1)/(d1+b)}`$ and $`M_z=M^{b/(d1+b)}`$. The grid spacing of the composition variables is $`\zeta _x=\left(V_d/M_x\right)^{1/(d1)}`$, where
$$V_d=\frac{\sqrt{d}}{(d1)!}$$
(10)
is the volume of the allowed composition simplex. Note that the distance from the centroid of the simplex to the closest point on the boundary of the simplex is
$$R_d=\frac{1}{\left[d(d1)\right]^{1/2}}.$$
(11)
The spacing for each component of the non-composition variables is $`\zeta _z=2/M_z^{1/b}`$. For the LDS method, different quasi-random sequences are used for the composition and non-composition variables. The feedback protocols Monte Carlo, Monte Carlo with swap, and Parallel Tempering are considered. The Monte Carlo parameters were optimized on test cases. It was optimal to perform 100 rounds of 1,000 samples with $`\beta =2`$ for $`d=3`$ and $`\beta =1`$ for $`d=4`$ or 5, and $`\mathrm{\Delta }x=0.1R_d`$ and $`\mathrm{\Delta }z=0.12`$ for the maximum random displacement in each component. The swapping move consisted of an attempt to swap all of the non-composition values between the two chosen samples, and it was optimal to use $`P_{\mathrm{swap}}0.1`$ for the probability of a swap versus a regular random displacement. For Parallel Tempering it was optimal to perform 100 rounds with 1,000 samples, divided into three subsets: 50 samples at $`\beta _1=50`$, 500 samples at $`\beta _2=10`$, and 450 samples at $`\beta _3=1`$. The 50 samples at large $`\beta `$ essentially perform a “steepest-ascent” optimization and have smaller $`\mathrm{\Delta }x=0.01R_d`$ and $`\mathrm{\Delta }z=0.012`$.
The figures of merit found by the protocols are shown in Fig. 3. The Random and LDS protocols find better solutions than does Grid in one round of experiment. More importantly, the Monte Carlo methods have a tremendous advantage over one pass methods, especially as the number of variables increases, with Parallel Tempering the best method. The Monte Carlo methods, in essence, gather more information about how best to search the variable space with each succeeding round. This feedback mechanism proves to be effective even for the relatively small total sample size of 100,000 considered here. We expect that the advantage of the Monte Carlo methods will become even greater for larger sample sizes. Note that in cases such as catalytic activity, sensor response, or ligand specificity , the experimental figure of merit would likely be exponential in the values shown in Fig. 3, so that the success of the Monte Carlo methods would be even more dramatic. A better calibration of the parameters in Eq. 8 may be possible as more data becomes available in the literature.
## V Conclusion
To conclude, the experimental challenges in combinatorial chemistry appear to lie mainly in the screening methods and in the technology for the creation of the libraries. The theoretical challenges, on the other hand, appear to lie mainly in the library design and redesign strategies. We have addressed this second question via an analogy with Monte Carlo computer simulation, and we have introduced the Random Phase Volume Model to compare various strategies. We find the multiple-round, Monte Carlo protocols to be especially effective on the more difficult systems with larger numbers of composition and non-composition variables.
An efficient implementation of the search strategy is feasible with existing library creation technology. Moreover “closing the loop” between library design and redesign is achievable with the same database technology currently used to track and record the data from combinatorial chemistry experiments. These multiple-round protocols, when combined with appropriate robotic controls, should allow the practical application of combinatorial chemistry to more complex and interesting systems.
## Acknowledgment
This research was supported by the National Science Foundation through grant number CTS–9702403.
|
warning/0007/cs0007029.html
|
ar5iv
|
text
|
# Dimension-dependent behavior in the satisfiability of random 𝑘-Horn formulae
## 1 Introduction
Finding the ground state (state of minimum energy) of a physical system and computing an optimal solution to a combinatorial optimization problem are intuitively two very similar tasks. This simple observation, that motivated the development of simulated annealing , a simple general-purpose heuristic for combinatorial optimization, lies behind the recent birth of a new field at the crossroads of Statistical Mechanics, Theoretical Computer Science and Artificial Intelligence, that studies phase transitions in combinatorial problems (see for a readable introduction). The transfer of principles and methods from Physics (mainly from Spin Glass Theory ) to Computer Science has already been quite successful, and is responsible for a couple of interesting results, such as a better understanding of the factors that account for computational intractability , strikingly accurate predictions of the average running time of various algorithms , or of expected values of optimal solutions .
The need for a rigorous validation of these insights is quite obvious. The theory of spin glasses is a relatively young field, which still presents many heuristic, unsolved or plain controversial aspects (for example see for a debate on the validity and scope of the so-called Parisi solution of the Sherrington–Kirkpatrick model). Moreover, while physical intuition can guide the development of the theory for “physical” models, by corroborating (or falsifying) some of its predictions (e.g. see , for a discussion of the demise, on physical grounds, of the first formulation of the so-called replica method), such intuition is not available when applying this type of ideas to combinatorial problems. Given that rigorous results are hard to come by in the case of spin glasses proper, it is not surprising that while there has been recently some progress (see e.g. ), an analysis of most interesting combinatorial problems is still out of reach.
An approach that was popular in Statistical Mechanics was to gather intuition through the systematic study of exactly solved models . These are “toy” versions of the original models that are simple to deal with, but retain much of the properties of the former ones. We advocate such an approach for problems in Computer Science as well, and the purpose of this paper is to present a (hopefully nontrivial) “exactly solvable satisfiability model” that displays a dimension-dependent behavior fairly similar to the one observed previously in various contexts such as percolation , self-avoiding walks, and recently for $`k`$-satisfiability by Kirkpatrick and Selman . The problem we investigate is random Horn satisfiability, and the “dimensionality” of a formula is taken to be the maximum length of its clauses.<sup>1</sup><sup>1</sup>1for technical convenience, all over the paper random $`k`$-Horn satisfiability is understood as random at-most-$`k`$-Horn satisfiability.
## 2 Overview
There are actually two different notions of phase transition in a combinatorial problem. The first of them, called order-disorder phase transition applies to optimization problems and directly parallels the approach from Statistical Mechanics. Potential solutions for an instance of $`P`$ are viewed as “states” of a system. One defines an abstract Hamiltonian (energy) function, that measures the “quality” of a given solution, and applies methods from the theory of spin glasses to make predictions on the typical structure of optimal solutions. In this setting a phase transition is defined as non-analytical behavior of a certain “order parameter” called free energy, and a discontinuity in this parameter, manifest by the sudden emergence of a backbone of constrained “degrees of freedom” is responsible for the exponential slow-down of many natural algorithms.
The second definition is combinatorial and pertains to decision problems. It relies on the concept of threshold property from random graph theory, more precisely a restricted version of this notion, called sharp threshold. A satisfiability threshold always exists for monotone problems , but may or may not be sharp (we speak of a coarse threshold in the latter case).
The layout of the paper is as follows: in section 3 we review the results of Kirkpatrick and Selman, in particular discussing the concept of critical behavior, as well as some objectionable aspects of their results. We then define the type of dimension dependent behavior we are interested in, argue that it captures to a large extent the results presented in , and contrast it with critical behavior. Our results are presented and discussed in section 6, while in section 14 we further discuss their significance.
Finally for $`k=2`$, the one where the satisfaction probability has a singularity we are able to rigorously display another phenomenon that is believed to be characteristic of phase transitions: in many cases the “hardest on the average” instances appear at the transition point (even if we only consider satisfiable instances ); this feature is quite robust with respect to the choice of the particular algorithm . We are able to prove that for a particular problem, random at-most-2-Horn satisfiability, the average running time of a particular algorithm, when restricted to satisfiable instances (the ones that are statistically significant on both sides of the critical point) is finite outside the critical point, and it diverges as we approach this point, thus providing some evidence for the experimental wisdom.
## 3 Phase transitions and critical behavior
We first discuss, briefly and limited to our interests, threshold phenomena. Perhaps the best way to introduce them is through a concrete example. To do this, we will use one “canonical” NP-complete problem, $`k`$-CNF satisfiability.
To generate random formulas we use a model with one parameter, the constraint density $`c`$, defined as the ratio between the number of clauses $`m`$ and the number of variables $`n`$ of the formula. A random formula is obtained by choosing $`m`$ random clauses. If we plot the probability that such a random formula is satisfiable against the constraint density $`c`$, we notice the existence of a critical value $`c_k`$ such that the satisfaction probability drops (as $`n\mathrm{}`$) from one to zero at $`c_k`$. Such a “sudden change” is an illustration of the mathematical concept of sharp threshold, qualitatively illustrated in Figure 1. The existence of a critical value $`c_k`$ has not been rigorously established (except for $`c_2=1`$), even though Friedgut has shown that the transition is “sharp” for every $`k`$.
Of special interest will also be the width of the so-called scaling window (a.k.a. critical region). To define it consider, for $`0<\delta <1`$, $`\alpha _{}(n,\delta )`$, the supremum over $`\alpha `$ such that for $`m=\alpha n`$, the probability of a random formula being satisfiable is at least $`1\delta `$. Similarly, let $`\alpha _+(n,\delta )`$ be the infimum over $`\alpha `$ such that for $`m=\alpha n`$, the probability of a random formula being satisfiable is at most $`\delta `$. Then, for $`\alpha `$ within the $`\delta `$-scaling window
$$W(n,\delta )=(\alpha _{}(n,\delta ),\alpha _+(n,\delta )),$$
(1)
the probability that a random formula is satisfiable is between $`\delta `$ and $`1\delta `$.
We will be interested in the width of the window $`W(n,\delta )`$ as a function of $`n`$. It is generally believed that $`|W(n)|=\theta (n^{1/\nu })`$ for some $`\nu =\nu _k1`$ independent of $`\delta `$, even though the existence of $`\nu _k`$ has only been established for $`k=2`$ .
### 3.1 Order/disorder phase transitions
Statistical mechanics deals with the description of systems having a large number of degrees of freedom. One of its fundamental predictions concerns the fact that at thermal equilibrium each such state occurs with probability proportional to $`exp(\beta H(\sigma ))`$, where $`\beta `$ is an inverse temperature, and $`H`$ is a Hamiltonian function, describing the energy of the particular state $`\sigma `$. The resulting distribution is called the Gibbs distribution $`G_\beta `$ given by
$$\mathrm{Pr}[\sigma ]=\frac{exp(\beta H(\mathrm{\Phi };\sigma ))}{Z[\mathrm{\Phi }]},$$
where
$$Z[\mathrm{\Phi }]=\underset{\sigma \{0,1\}^n}{}exp(\beta H(\mathrm{\Phi };\sigma ))$$
is the so-called partition function.
Changes in the order properties of the system, which characterize order-disorder phase transitions, manifest themselves as non-analytical behavior of thermal averages (i.e. averages over the Gibbs distribution) of a certain order parameter. We want to emphasize that the physicists’ use of the term order parameter would be quite different from the one from combinatorics. An order parameter is a quantity that is zero on one side of the phase transition and becomes non-zero on the other side (for instance the satisfaction probability could be an order parameter).
One of the simplest illustrations of these concepts is the two-dimensional Ising model (see for a thorough treatment). In this model we have a number of spins, that are small magnets located on the vertices of the two-dimensional lattice, and pointing either up or down. The spins interact with their neighbors and with an external magnetic field $`h𝐑`$, which will tend to align the spins in one of the two directions. The energy of a state $`\sigma `$ is
$$H(\sigma )=\underset{ij}{}\sigma _i\sigma _j+h\left(\underset{i}{}\sigma _i\right).$$
The order parameter is called free energy, is a function of temperature, and is formally defined as
$$f=\frac{1}{\beta n}\mathrm{ln}Z[\mathrm{\Phi }].$$
It measures the fraction of spins that are “frozen” when the field is turned off.
We now briefly describe the essence of the phase transition: above a certain temperature $`T_c`$, the Curie-Weiss point, when the magnetic field is turned to zero the proportion of spins that point in each direction is about $`\frac{1}{2}`$ (the so-called disordered phase). But for temperatures below $`T_c`$ when we turn the field to zero some orientation still dominates (the ordered phase), and the proportion of spins pointing up(down) changes discontinuously as $`h`$ passes through zero.
The connection with combinatorial optimization follows from the observation that when $`\beta \mathrm{}`$ (that is the temperature approaches 0 K), the Gibbs distribution $`G_\beta `$ converges to a uniform distribution $`G`$ on the set of states of minimal energy (ground states). Thus, based on this analogy, one can hope that ideas from Statistical Mechanics are able to provide insight into the structure of optimal solutions to an instance of a problem in Combinatorial Optimization. Rather than providing a complete discussion (which would require to rigorously define the notion of optimization problem) we will discuss this in the context of MAX 3-SAT, the optimization version of satisfiability. For now it suffices to mention the three main ingredients of an optimization problem, its instances, solutions to instances of a problem, and an cost function, that measures the quality of a solution for a certain instance.
###### Example 1.
(MAX 3-SAT)
Input: A propositional formula $`\mathrm{\Phi }`$ in conjunctive normal form, such that every clause has length exactly 3.
Solution: A truth assignment $`\sigma `$ for the propositional variables in $`\mathrm{\Phi }`$ that maximizes the number of satisfied clauses.
Cost function: The cost $`C(\mathrm{\Phi },\sigma )`$ of a truth assignment $`\sigma `$ for an instance $`\mathrm{\Phi }`$ of MAX 3-SAT is the number of clauses of $`\mathrm{\Phi }`$ that are violated by $`\sigma `$.
Let $`Q`$ be an optimization problem and let $`\mathrm{\Phi }`$ be an instance of $`Q`$ “on $`n`$ variables” (i.e., all solutions have length $`n`$). We view the set of all assignments on $`\{0,1\}^n`$ as “states of a system.” To each such state $`\sigma `$ we associate the Hamiltonian (energy function)
$$H(\mathrm{\Phi };\sigma )=\text{ the cost of instance }(\mathrm{\Phi };\sigma )\text{ of }Q.$$
###### Example 2.
Let $`\mathrm{\Phi }`$ be a 3-CNF formula, and let $`\sigma `$ be an assignment. According to the previous definition $`H(\mathrm{\Phi };\sigma )=C(\mathrm{\Phi };\sigma )`$. $`H`$ can be formally expressed as
$$H(\mathrm{\Phi };\sigma )=\underset{l=1}{\overset{m}{}}\delta [\underset{i=1}{\overset{n}{}}C_{l,i}(1)^{\sigma _i};3],$$
where $`\delta [i;j]=1_{\{i=j\}}`$ is the Kronecker symbol and $`C_{l,i}`$ is 1 if the $`l`$th clause contains the literal $`x_i`$, $`1`$ if it contains $`\overline{x_i}`$ and zero otherwise.
For the case of problems of interest to Computer Science the instance $`\mathrm{\Phi }`$ is not fixed, but rather is a sample from a certain distribution. This is very similar to the context of spin-glass theory, a subfield of Statistical Mechanics. The extra ingredient of this theory is that the coupling coefficients are no longer considered fixed, but are rather independent samples from a certain distribution. In the language of the theory of spin glasses $`\mathrm{\Phi }`$ is called a quenched quantity).
As in the case of the Ising model, the order parameter is the ground state free energy, more precise its expected value
$$\overline{f}=\frac{1}{\beta n}\overline{\mathrm{ln}(Z)},$$
where $`\overline{(\mathrm{})}`$ stands for the average over the random distribution of $`\mathrm{\Phi }`$.
###### Definition 1.
A physical (order/disorder) phase transition in a combinatorial optimization problem is a point where $`\overline{f}`$ is not analytical.
Free energy has an especially crisp intuitive interpretation in the case of the problem MAX 3-SAT :
###### Example 3.
Let $`\mathrm{\Phi }_n`$ be an instance of MAX 3-SAT, let $`A`$ be the set of optimal assignments to $`\mathrm{\Phi }_n`$, endowed with the uniform measure $`\mu _n`$. Statistical Mechanics predicts that, as $`n\mathrm{}`$, $`\mu _n`$ is “close” to a product measure on $`\{0,1\}^n`$, $`\mu _{1,n}\mathrm{}\mu _{n,n}`$. The free energy per site $`f`$ is the fraction of variables $`x_i`$ that are (asymptotically) fully constrained (that is $`\mu _{i,n}`$ converges in distribution to a measure having all its weight on one of the two points 0,1.
## 4 Critical behavior and the mean-field approximation
An important feature that order/disorder phase transition share with the combinatorial notion of threshold properties (that are usually the type of phase transition of interest in combinatorics) is that the various quantities of interest, such as the satisfaction probability, the ground state energy, and the location of the phase transition are hard to compute. No general-purpose methods exist, and in some cases even obtaining good non-rigorous estimates is a challenging open problem.
A technique that often provides realistic approximate values for these quantities came to be known as the mean-field (annealed) approximation. In a nutshell a mean-field approximation assumes that we are trying to compute the average (over a certain discrete probability space) of a certain expression $`f(g_1,\mathrm{},g_n)`$. Then the mean field-approximation amounts to taking
$$E[f(g_1(x),\mathrm{},g_n(x)]f[E[g_1(x)],\mathrm{},E[g_n(x)]].$$
This technical definition of the mean-field approximation does not convey a useful intuition: suppose we want to solve a combinatorial problem whose objective function depends on simultaneously satisfy several “constraints” whose effects are usually not independent. The mean-field approximation ignores the dependencies between various constraints, and treat them as independent.
###### Example 4.
Let us return to the case of spin glasses. Each configuration of spins $`\sigma `$ has an energy specified by a Hamiltonian $`H(\sigma )`$. A typical expression for $`H(\sigma )`$ is
$$H(\sigma )=\underset{ij}{}a_{i,j}\sigma _i\sigma _j,$$
where the $`a_{i,j}`$’s are interaction coefficients between adjacent spins (according to some adjacency graph specific to the considered model). The quantity of interest, average free energy $`\overline{f}`$ is hard to compute directly because of the logarithmic function present in the definition of the free energy. In this context the mean-field approximation amounts to
$$\overline{f}\frac{1}{\beta n}\mathrm{ln}[\overline{Z[\mathrm{\Phi }]}].$$
The advantage of this heuristic is that the average on the right-hand side is one that is usually much easier to compute.
For combinatorial phase transitions, the mean-field approach usually amounts to an approximation using the so-called first-moment method
###### Example 5.
($`k`$-Satisfiability)
The reason that the satisfiability probability of a random formula is hard to compute is that, for two assignments $`A,B`$ the events $`A\mathrm{\Phi }`$ and $`B\mathrm{\Phi }`$ are not independent. One way to construct a mean-field theory for $`k`$-SAT is to ignore the dependencies between these events. More precisely, we have
$$1_{SAT}[\mathrm{\Phi }]=f(g_{A_1}[\mathrm{\Phi }],\mathrm{},g_{A_{2^n}}[\mathrm{\Phi }]),$$
where
$$f(x_1,x_2,\mathrm{},x_{2^n})=1\underset{i=1}{\overset{2^n}{}}x_i,$$
and
$$g_A[\mathrm{\Phi }]=\{\begin{array}{cc}1,\hfill & \text{ if }A\vDash ̸\mathrm{\Phi },\hfill \\ 0,\hfill & \text{ otherwise.}\hfill \end{array}$$
Define $`\gamma _k=12^k`$. The mean-field approximation amounts to
$$\mathrm{Pr}[\mathrm{\Phi }SAT]=E[1_{SAT}[\mathrm{\Phi }]]f(E_{g_1}[\mathrm{\Phi }],\mathrm{},E_{g_{2^n}}[\mathrm{\Phi }])$$
Since
$$E_{g_1}[\mathrm{\Phi }]=\mathrm{}=E_{g_{2^n}}[\mathrm{\Phi }])=1\gamma _k^{cn}$$
this reads,
$$\mathrm{Pr}[\mathrm{\Phi }SAT]1\left[1\gamma _k^{cn}\right]^{2^n}1e^{2^n\gamma _k^{cn}}=1e^{E[\mathrm{\#}_{SAT}[\mathrm{\Phi }]]}$$
where $`\mathrm{\#}_{SAT}[\mathrm{\Phi }]`$ is the number of satisfying assignments for $`\mathrm{\Phi }`$. Thus (neglecting the case $`E[\mathrm{\#}_{SAT}[\mathrm{\Phi }]]=1`$)
$$\mathrm{Pr}[\mathrm{\Phi }SAT]=\{\begin{array}{cc}1,\hfill & \text{ if }E[\mathrm{\#}_{SAT}[\mathrm{\Phi }]]\mathrm{},\hfill \\ 0,\hfill & \text{ if }E[\mathrm{\#}_{SAT}[\mathrm{\Phi }]]0.\hfill \end{array}$$
### 4.1 Critical exponents and behavior
A phenomenon that has been observed in various contexts is critical behavior. In these cases the class of problems under study has an intrinsic notion of dimensionality $`d`$, and in the limit $`d\mathrm{}`$ (or sometimes even when $`d`$ is greater than a so-called critical dimension) “the annealed approximation becomes exact”.
A way to give precise meaning to the above quote comes from the concept of universality. In Statistical Mechanics one define certain critical exponents, that describe the behavior of the system near the critical points; universality predicts that phase transitions with the same critical exponents are “structurally similar”.
Since critical exponents can be defined for the mean-field versions of the physical models too, critical behavior means that as $`d\mathrm{}`$ (or, sometimes, for $`d`$ larger than a value called the upper critical dimension) the critical exponents of the $`d`$-dimensional system coincide with the critical exponents of the $`d`$-dimensional mean-field model.
###### Example 6.
(Bond) percolation on the lattice $`𝐙^d`$. Percolation is a mathematical theory that models the flow of liquids in random porous media. In our case the flow is on the lattice $`𝐙^d`$ of dimension $`d`$, and the model has one parameter, the edge probability $`p[0,1]`$. Each bond (grid edge of the lattice $`𝐙^d`$) is considered open with probability $`p`$ (independently of the other bonds) and the order parameter is the probability $`P_d(p)`$ that the origin lies in an infinite cluster. $`P_d`$ is a monotonically increasing function of $`p`$. It is believed that $`P_d(p)`$ is zero up to a critical value $`p_c(d)`$ (known rigorously only for $`d=2`$), greater than zero beyond that point, and non-analytical but continuous (at least for $`d=2`$) at $`p_c(d)`$. It is also believed that above (and around the critical value) $`P_d(p)(pp_c(d))^\beta `$ where $`\beta `$ is a critical exponent that depends on $`d`$ but not on the explicit lattice considered (i.e. it would be the same if we choose another $`d`$-dimensional lattice instead of $`𝐙^d`$). This is only one of the several critical exponents that are believed to structurally characterize percolation on $`d`$-dimensional lattices (see ).
Without going into further details, we note that the “mean-field approximation” corresponds to considering percolation on the $`d`$-dimensional Bethe lattice, a nd the critical behavior amounts to the observation that for $`d`$ greater than a critical dimension (known to be at most 16 , and is believed to be 6) the critical exponents of percolation on $`𝐙^d`$ are those of percolation on the Bethe lattice.
### 4.2 Rescaling and critical behavior
A recent example of critical behavior has recently been observed experimentally by Kirkpatrick and Selman for satisfiability problems.
Their results does not mention critical exponents (although it is closely related). To explain them, we need to introduce first another concept from Statistical Mechanics: finite-size scaling. The intuition behind it is that “sufficiently close to a threshold or critical point, systems of all sizes are indistinguishable except for an overall change of scale.” In mathematical terms this amounts to defining a new order parameter that “opens up” the scaling window, the region where the probability decreases from 1 to 0.
###### Example 7.
Hamiltonian Cycle.
The random model has one parameter $`m`$, the number of edges. A random sample is obtained by choosing uniformly at random a set of $`m`$ distinct edges of a complete graph with $`n`$ vertices. The following result (obtained by Komlós and Szemerédi ) describes the phase transition in this problem:
Let $`m=m(n)=\frac{1}{2}n\mathrm{log}(n)+\frac{1}{2}n\mathrm{log}\mathrm{log}(n)+c_nn`$. Then
$$\underset{n\mathrm{}}{lim}Pr[G\text{ has a Hamiltonian cycle}]=\{\begin{array}{cc}0,\hfill & \text{ if }c_n\mathrm{}\text{,}\hfill \\ e^{e^{2c}},\hfill & \text{if }c_nc\text{,}\hfill \\ 1,\hfill & \text{ if }c_n\mathrm{}\text{.}\hfill \end{array}$$
A rescaled parameter for the Hamiltonian cycle problem can be defined by $`c_n=\frac{1}{n}[m\frac{1}{2}n\mathrm{log}(n)\frac{1}{2}n\mathrm{log}\mathrm{log}(n)]`$. This parameter yields a rescaled limit probability function $`f(c)=e^{e^{2c}}`$.
It is important to note that, since an annealed approximation yields an expression for the order parameter (in our case satisfaction probability) that will usually display a phase transition as well, a rescaled parameter can be defined for the mean-field version of the problem as well.
The definition of the rescaled parameter allows a precise formulation of the intuition that an annealed approximation becomes exact in the limit $`d\mathrm{}`$. Let $`P_d`$ be a class of satisfiability problems indexed by a dimensionality parameter $`d`$, let $`F_d`$ be the rescaled satisfaction probability graph of $`P_d`$, and let $`F_{ann,d}`$ be the rescaled graph corresponding to the annealed approximation. Kirkpatrick and Selman observe experimentally that as $`d\mathrm{}`$, the function sequences $`F_d`$, $`F_{ann,d}`$ converge punctually to a common limit $`F_{\mathrm{}}`$.
###### Example 8.
We present in detail the experimental results of Kirkpatrick and Selman. They define an (approximate) rescaled parameter for $`k`$-SAT
$$y_k=n^{1/\nu _k}\frac{(cc_k)}{c_k},$$
where $`c=m/n`$, $`c_k`$ is the critical threshold for $`k`$-SAT, and $`\nu _k`$ is the scaling width coefficient. Also, define the “annealed rescaled parameter”
$$y_{\mathrm{},k}=n\frac{(cc_k)}{c_k},$$
The rescaled limit probability graphs (and, see below, the rescaled versions of the mean-field versions) seem to converge (see Fig. 4 in that paper) to the “annealed limit”
$$f_{\mathrm{}}(y)=e^{2^y}.$$
###### Definition 2.
In this paper dimension-dependent behavior refers to the above-mentioned type phenomenon, convergence of the “rescaled” probability functions (and their annealed counterparts) to some common annealed limit.
###### Observation 1.
It is important to note that dimension-dependent behavior is at the same time more and less demanding than critical behavior.
It is more demanding since it requires that the annealed approximation be exact throughout the (rescaled version) of the critical region. In contrast, critical exponents only provide a qualitative picture of this region, rather than uniquely determine the limit probability throughout it; for instance the width of the scaling window $`\nu `$ is equal to $`2\beta +\gamma `$, where $`\beta `$ is the so-called order-parameter exponent, that characterizes the asymptotic behavior of the order parameter close to the transition point, and $`\gamma `$ is called susceptibility exponent (see e.g. ).
It is less demanding since it does not assume the existence of critical exponents, therefore it makes sense for problems having coarse thresholds, including those that have no singular/critical points.
Why should we expect critical behavior and the above form for the annealed limit ? The intuition is very simple: the major difficulty in computing the probability that a random $`kSAT`$ formula is satisfiable is the fact that, for two assignments $`A`$ and $`B`$, the events “$`A\mathrm{\Phi }`$” and “$`B\mathrm{\Phi }`$” are not generally independent, because there exist clauses of length $`k`$ that are falsified by both $`A`$ and $`B`$. On the other hand, qualitatively, as $`k\mathrm{}`$ clausal constraints become progressively “looser”, so that in the limit we can neglect such correlations.
As to the exact expression for $`f_{\mathrm{}}(y)`$, for a $`k`$-CNF formula the mean-field approximation implies
$$\mathrm{Pr}[\mathrm{\Phi }\overline{SAT}](1\gamma _k^{cn})^{2^n}e^{2^n\gamma _k^{cn}}.$$
But since $`c_k`$ is specified (in the mean-field approximation) by $`E[\mathrm{\#}SAT]1`$, i.e. $`2^n\gamma _k^{c_kn}1`$, or $`1+c_k\mathrm{log}_2\gamma _k=0`$, this implies that as $`k\mathrm{}`$
$$\mathrm{Pr}[\mathrm{\Phi }\overline{SAT}]e^{2^{n[1c/c_k]}}f_{\mathrm{}}(y_{\mathrm{},k}).$$
In other words, when plotted against the annealed order parameters $`y_{ann,k}`$ the rescaled satisfaction probability graphs (and their annealed counterparts) punctually converge to the graph of $`f_{\mathrm{}}`$.
## 5 Does critical behavior really exist ?
The intuitive argument sketched in the preceding paragraph seems to provide a beautiful explanation of the experimental results from . That this intuition is, however, problematic has been shown by Wilson . First note that if the previous argument were true, we would have $`\nu _k=1`$ for any large enough $`k`$, since this is the width of the scaling window that the mean-field versions of $`kSAT`$ predict. On the other hand Wilson presented a simple argument that implies that $`\nu _k2`$) Hence the above explanation is not rigorously valid.
We stress that Wilson’s observation does not rule out the existence of critical behavior: we, in fact, believe that the qualitative intuition that motivated , that versions of $`kSAT`$ become more and more “similar” as $`k`$ goes to infinity, is correct. It is the notion of annealed approximation that needs to be changed. And, certainly, his results do not rule the possibility that the rescaled limit probabilities converge, as $`k\mathrm{}`$, to a suitable-defined limit. Obtaining a rigorous example where this holds, that identifies a suitable “annealed approximation that becomes exact” and also obtains an explanation for this convergence, could hopefully offer insights on how to address this problem for random $`kSAT`$ as well. This is what our theorems in the next section provide.
## 6 Our results
A Horn clause is a disjunction of literals containing at most one positive literal. It will be called positive if it contains a positive literal and negative otherwise. A Horn formula is a conjunction of Horn clauses. Horn satisfiability (denoted by $`\mathrm{HORN}\text{-}\mathrm{SAT}`$) is the problem of deciding whether a given Horn formula has a satisfying assignment.
In this chapter we prove a result that displays dimension-dependent behavior for (at most) $`k`$-Horn satisfiability, the natural version of Horn satisfiability studied, parameterized by the maximum clause length. This problem is also of practical interest in Artificial Intelligence, mainly in connection to theory approximation . The results can be summarized as follows:
1. For an unbounded $`k=k(n)`$ the threshold phenomenon is essentially the one from the “uniform case” $`k(n)=n`$. In particular there exists a “rescaled” parameter that makes the graphs of the limit probabilities superimpose (Theorem 4).
2. For any constant $`k`$ the threshold phenomenon is qualitatively described by a suitably chosen queuing model (Theorem 6). This yields a closed-form expression for the satisfaction probability when $`k=2`$ (Theorem 5). This expression has a singularity (though $`k=2`$ is likely the only case that does so).
3. The rescaled limit probabilities from the cases when $`k`$ is a constant converge to the one from the “infinite” case, that can in turn be seen as the result of a mean-field approximation (thus the problem displays what we have called dimension-dependent behavior).
4. Somewhat surprisingly, the explanation for this convergence (an intrinsic feature of the problem) is a threshold property for the number of iterations of PUR (a particular algorithm) on random satisfiable Horn formulas “in the critical range.”
5. In the case when $`k=2`$ PUR displays an “easy-hard-easy” pattern for the average number of iterations on satisfiable instances, peaked at the point where the limit probability has a singularity (Theorem 8).
Note, however, the important difference between random $`k`$-SAT and random at-most-$`k`$-HORN-SAT: for every $`k2`$, $`k`$-SAT has a sharp threshold . All versions of HORN-SAT have coarse thresholds.
###### Definition 3.
Let $`k=k(n):𝐍𝐍`$ be monotonically increasing, $`1k(n)n`$. We define the following random model $`\mathrm{\Omega }(k,n,m)`$: formula $`\mathrm{\Phi }`$ on $`n`$ variables is obtained by selecting (uniformly at random and with repetition) $`m`$ clauses from the set of all (non-empty) Horn clauses in the given variables of length at most $`k(n)`$.
The following are our results (whose proofs are only sketched):
###### Theorem 4.
If $`k(n)\mathrm{}`$, $`c>0`$, $`H_{k(n)}`$ is the number of Horn clauses on $`n`$ variables having length at most $`k(n)`$, and $`m(n)=c\frac{H_{k(n)}}{n}`$ then
$$p_{\mathrm{}}(c):=\underset{n\mathrm{}}{lim}Pr_{\mathrm{\Phi }\mathrm{\Omega }(k(n),n,m)}(\mathrm{\Phi }\text{HORN-SAT})=1F_1(e^c).$$
(2)
###### Theorem 5.
If $`c>0`$, and $`F_2:(0,1)(1,\mathrm{})`$, $`F_2(x)=\mathrm{ln}x/(x1)`$, then
$$p_2(c):=\underset{n\mathrm{}}{lim}Pr_{\mathrm{\Phi }\mathrm{\Omega }(2,n,cn)}(\mathrm{\Phi }\text{HORN-SAT})=\{\begin{array}{cc}1,\hfill & \text{ if }c\frac{3}{2}\text{,}\hfill \\ F_2^1(2c/3),\hfill & \text{ otherwise.}\hfill \end{array}$$
(3)
More generally, define $`\lambda _k=\frac{k!}{k+1}`$ and $`S_j^i=\left(\genfrac{}{}{0pt}{}{i}{0}\right)+\left(\genfrac{}{}{0pt}{}{i}{1}\right)+\mathrm{}+\left(\genfrac{}{}{0pt}{}{i}{j}\right)`$ (with the usual convention $`\left(\genfrac{}{}{0pt}{}{i}{j}\right)=0`$ for $`i<j`$). Then
###### Theorem 6.
The limit probability $`p_k(c):=lim_n\mathrm{}Pr_{\mathrm{\Phi }\mathrm{\Omega }(k,n,cn^{k1})}(\mathrm{\Phi }\text{HORN-SAT})`$ is equal to the probability that the following Markov chain ever hits state zero:
$$\{\begin{array}{c}Q_0=1,\hfill \\ Q_{i+1}=Q_i\dot{}1+Po(c\lambda _kS_{k2}^{i+1}),\hfill \end{array}$$
(4)
To get a better intuition on the threshold phenomenon, as displayed by Theorems 4, 5 and 6, we have plotted (in Fig. 1) the limit probability functions $`p_2(),p_3(),p_{\mathrm{}}()`$, against the “rescaled” parameter (inspired by Theorem 4) $`\widehat{c}=\frac{mn}{H_{k(n)}}`$. This rescaling has the pleasant property that it simplifies the factor $`\lambda _k`$ from the right-hand side of 4, in particular mapping the critical point in Theorem 5 to $`\widehat{c}=1`$. The graphs of $`p_2`$ (continuous) and $`p_{\mathrm{}}`$ (dashed) are obtained from their formulas in the previous results, while $`p_3`$ (dotted) is obtained via simulations. The figure makes apparent that the graphs of $`p_2,p_3,\mathrm{},\mathrm{}`$ converge to the graph of $`p_{\mathrm{}}`$. This statement can be proved rigorously :
###### Theorem 7.
For every $`\widehat{c}>0`$, $`lim_n\mathrm{}p_n(\widehat{c})=p_{\mathrm{}}(\widehat{c})`$.
As a bonus our analysis yields the following result:
###### Theorem 8.
Let $`q`$ be the limit of the expected number of iterations of PUR on a random formula $`\mathrm{\Phi }\mathrm{\Omega }(2,n,cn)`$, conditional on $`\mathrm{\Phi }`$ being satisfiable. Then
$$q=\{\begin{array}{cc}\frac{1}{1p_2\lambda _2c}\hfill & \text{, if }c\frac{3}{2}\text{,}\hfill \\ \mathrm{},\hfill & \text{ otherwise.}\hfill \end{array}$$
(5)
This theorem suggests (see Fig.2) and explains the “easy-hard-easy” pattern for the average running time of PUR in this case. Experiments we performed confirm this prediction.
## 7 Preliminaries
Throughout this paper we use “with high probability” (w.h.p.) as a substitute for “with probability $`1o(1)`$”. We denote (sometimes abusing notation) by $`B(n,p)(Po(\lambda ))`$ a random variable having a binomial (Poisson) distribution with the corresponding parameter(s), and by $`a\dot{}b`$ the value $`max(ab,0)`$. We will use the following version of the Chernoff bound
###### Theorem 9.
If $`0<\theta <1/4`$ then $`\mathrm{Pr}[|B(n,p)np|>\theta np]e^{np\frac{\theta ^2}{4}}`$.
as well as the related inequality from :
###### Proposition 10.
Let $`P`$ have Poisson distribution with mean $`\mu `$. For $`ϵ>0`$,
$$\mathrm{Pr}[P\mu (1ϵ)]e^{ϵ^2\mu /2},$$
$$\mathrm{Pr}[P\mu (1+ϵ)][e^ϵ(1+ϵ)^{(1+ϵ)}]^\mu .$$
We also use the following inequality:
###### Proposition 11.
Let $`k𝐍`$ and $`p[0,1]`$. Then for every $`nk`$
$$1\underset{i=0}{\overset{k1}{}}\left(\genfrac{}{}{0pt}{}{n}{i}\right)p^i(1p)^{ni}\left(\genfrac{}{}{0pt}{}{n}{k}\right)p^k.$$
(6)
Proof: Define $`f:[0,1]R`$, $`f(p)=1_{i=0}^{k1}\left(\genfrac{}{}{0pt}{}{n}{i}\right)p^i(1p)^{ni}\left(\genfrac{}{}{0pt}{}{n}{k}\right)p^k`$. It is easy to see that $`f^{}(p)=n\left(\genfrac{}{}{0pt}{}{n1}{k1}\right)p^{k1}[(1p)^{nk}1]0`$, therefore $`f`$ is monotonically decreasing, and $`f(0)=0`$. $`\mathrm{}`$
We will also employ couplings of Markov chains (see ) to assert stochastic domination. The following is the definition of the type of coupling we employ in this paper:
###### Definition 12.
Let $`(X_t)_t`$ and $`(Y_t)_t`$ be two Markov chains on $`𝐙`$. A coupling of $`X`$ and $`Y`$ such that $`X_tY_t`$ is a Markov chain $`Z=(Z_{t,1},Z_{t,2})`$ such that:
* $`Z_{t,1}`$ is distributed like $`X_t`$ given $`X_0`$.
* $`Z_{t,2}`$ is distributed like $`Y_t`$ given $`Y_0`$.
* for every $`i0`$, $`Z_{i,1}Z_{i,2}`$.
We use such couplings to bound the probability that a Markov chain $`Y_t`$ ever decreases below a certain value $`a`$ by coupling it with a chain $`X_t`$ such that $`X_tY_t`$ and using the estimate $`\mathrm{Pr}[t:Y_ta]\mathrm{Pr}[t:X_ta]`$ (that follows from the coupling). The couplings we construct employ the following ideas:
* Suppose the recurrences describing $`\mathrm{\Delta }X_t`$ and $`\mathrm{\Delta }Y_t`$ are identical, except for one term, which is $`B(m_1,\tau )`$ in $`X_t`$ and $`B(m_2,\tau )`$ in $`Y_t`$, where $`m_1m_2`$ are positive integers and $`\tau (0,1)`$. Obtain a coupling by identifying $`B(m_1,\tau )`$ with the outcome of the first $`m_1`$ Bernoulli experiments in $`B(m_2,\tau )`$.
* Suppose now that $`\mathrm{\Delta }X_t`$ and $`\mathrm{\Delta }Y_t`$ differ by exactly one term which is $`B(m,p)`$ in $`\mathrm{\Delta }X_t`$ and $`B(m,q)`$ in $`\mathrm{\Delta }Y_t`$, $`pq`$. Let $`A_i`$ and $`B_i`$, $`i=1,m`$, be independent $`0/1`$ experiments with success probabilities $`p`$ and $`\frac{qp}{1p}`$ respectively. Define the pair $`(Z_{t,1},Z_{t,2})`$ so that
1. $`Z_{t,1}`$ is the number of times $`A_i`$ succeeds.
2. $`Z_{t,2}`$ is the number of times at least one of $`A_i`$ and $`B_i`$ succeeds.
We measure the distance between two probability distributions $`P`$ and $`Q`$ by the total variation distance, denoted by $`d_{TV}(P,Q)`$, and recall the following results, (see and , page 2 and Remark 1.4):
###### Lemma 13.
If $`n,p,\lambda ,\mu >0`$ then $`d_{TV}(B(n,p),Po(np))\mathrm{min}\{np^2,\frac{3p}{2}\}`$ and $`d_{TV}(Po(\lambda ),Po(\mu ))|\mu \lambda |`$.
We will also need the following simple lemma:
###### Lemma 14.
Let c be a fixed positive integer. For every $`t𝐍`$ let $`\xi _t`$, $`\eta _t`$ be two probability distributions. Define the Markov chains $`(X_t)_t`$ and $`(Y_t)_t`$ by recurrences
$$\{\begin{array}{c}X_{t+1}=X_t\dot{}c+\xi _t,\hfill \\ Y_{t+1}=Y_t\dot{}c+\eta _t.\hfill \end{array}$$
(7)
Then, for every $`t0`$, $`d_{TV}(X_t,Y_t)d_{TV}(X_0,Y_0)+_{i=0}^{t1}d_{TV}(\xi _i,\eta _i).`$
Proof.
The following result gives a more convenient inequality that immediately implies Lemma 14
###### Lemma 15.
Let c be a fixed positive integer. Let $`X`$, $`Y`$, $`\xi `$, $`\eta `$ be random variables with nonnegative integer values. Define the random variables $`Z`$ and $`T`$ by recurrences
$$\{\begin{array}{c}Z=X\dot{}c+\xi ,\hfill \\ T=Y\dot{}c+\eta .\hfill \end{array}$$
(8)
Then, for every $`d_{TV}(Z,T)d_{TV}(X,Y)+d_{TV}(\xi ,\eta ).`$
Proof.
To prove this result, we will denote (for the “generic” r.v. $`A`$) by $`A_i`$ the probability that $`A`$ takes value $`i`$. We also employ the following simple inequality, valid for $`a,b,c,d0`$: $`|adbc|a|dc|+|ab|c`$.
For every $`a0`$ we have:
$$Z_a=\underset{i=0}{\overset{c}{}}X_i\xi _a+\underset{i=c+1}{\overset{c+a}{}}X_i\xi _{a+ci},$$
$$T_a=\underset{i=0}{\overset{c}{}}Y_i\eta _a+\underset{i=c+1}{\overset{c+a}{}}Y_i\eta _{a+ci},$$
Applying the above-mentioned inequality and summing we get:
$`d_{TV}(Z,T)`$
$``$ $`{\displaystyle \frac{1}{2}}\{{\displaystyle \underset{i=0}{\overset{c}{}}}{\displaystyle \underset{a=0}{\overset{\mathrm{}}{}}}X_i|\xi _a\eta _a|+{\displaystyle \underset{i=0}{\overset{c}{}}}{\displaystyle \underset{a=0}{\overset{\mathrm{}}{}}}|X_iY_i|\eta _a+`$
$`+`$ $`{\displaystyle \underset{i=c+1}{\overset{c+a}{}}}{\displaystyle \underset{a=0}{\overset{\mathrm{}}{}}}X_i|\xi _{c+ai}\eta _{c+ai}|+{\displaystyle \underset{i=c+1}{\overset{c+a}{}}}{\displaystyle \underset{a=0}{\overset{\mathrm{}}{}}}|X_iY_i|\eta _{c+ai}\}.`$
Let A,B,C,D be the four terms of the sum. By simple algebraic manipulations we obtain:
$$\begin{array}{ccc}A=(_{i=0}^cX_i)d_{TV}(\xi ,\eta ),\hfill & & B=\frac{1}{2}_{i=0}^c|X_iY_i|,\hfill \\ C=(_{i=c+1}^{\mathrm{}}X_i)d_{TV}(\xi ,\eta ),\hfill & & D=\frac{1}{2}_{i=c+1}^{\mathrm{}}|X_iY_i|,\hfill \end{array}$$
and the result follows. $`\mathrm{}`$
Finally, we need the following trivial occupancy property:
###### Lemma 16.
Let $`a`$ white balls and $`b`$ black balls be thrown uniformly at random in $`n`$ bins.
1. if $`r=\mathrm{max}(a,b)=o(n^{1/2})`$ then the probability that there is a bin that contains both white and black balls is at most $`\frac{4r^2}{n}=o(1)`$.
2. if $`s=\mathrm{min}(a,b)=\omega (n^{1/2})`$ then the probability that there is a bin that contains both white and black balls is $`1o(1/poly)`$.
Proof. The first part is easy: the probability that two balls (of any color) end up in the same bin is at most $`\left(\genfrac{}{}{0pt}{}{a+b}{2}\right)\frac{1}{n}`$. For the second part, let $`A`$ be the event that no two balls of different colors end up in the same bin, and let $`B`$ the event that at least $`\sqrt{n}`$ bins contain white balls. We have:
$$\mathrm{Pr}[A]\mathrm{Pr}[A|B]+\mathrm{Pr}[\overline{B}].$$
But
$$\mathrm{Pr}[\overline{B}]\left(\genfrac{}{}{0pt}{}{n}{\sqrt{n}}\right)(\frac{1}{\sqrt{n}})^a=n^{\sqrt{n}a/2}=o(\frac{1}{poly}),$$
and
$$\mathrm{Pr}[A|B](1\frac{1}{\sqrt{n}})^be^{b/\sqrt{n}}=o(\frac{1}{poly}).$$
$`\mathrm{}`$
The algorithm PUR is displayed in Figure 3. We regard PUR as working in stages, indexed by the number of variables still left unassigned; thus, the stage number decreases as PUR moves on. We say that formula $`\mathrm{\Phi }`$ survives Stage $`t`$ if PUR on input $`\mathrm{\Phi }`$ does not halt at Stage $`t`$ or earlier. Let $`\mathrm{\Phi }_i`$ be the formula at the beginning of stage $`i`$, and let $`N_i`$ denote the number of its clauses. We will also denote by $`P_{i,t}(N_{i,t})`$, the number of clauses of $`\mathrm{\Phi }_t`$ of size $`i`$ and containing one (no) positive literal. Define $`\mathrm{\Phi }_{i,t}^P`$ ($`\mathrm{\Phi }_{i,t}^N`$) to be the subformula of $`\mathrm{\Phi }_t`$ containing the clauses counted by $`P_{i,t}(N_{i,t})`$.
The following lemmas were proved in , in the context of analyzing the behavior of PUR on $`\mathrm{\Phi }\mathrm{\Omega }(n,n,m)`$, $`m=c2^n`$.
###### Lemma 17.
1. Suppose $`\mathrm{PUR}`$ does not halt before stage $`t`$. Then, conditional on $`N_t`$, the clauses of $`\mathrm{\Phi }_t`$ are random and independent.
2. Suppose now that we condition on $`\mathrm{\Gamma }_t=(N_{1,t},N_{2,t},P_{1,t},P_{2,t}`$ and on the fact that $`\mathrm{\Phi }`$ survives Stage $`t`$ as well. Then we have
$$N_{t1}=N_t\mathrm{\Delta }_{1,P}(t)\mathrm{\Delta }_{2,P}(t),$$
(9)
where
* $`\mathrm{\Delta }_{1,P}(t)`$, the number of positive clauses that are satisfied at stage $`t`$, has the distribution $`1+B(P_{1,t}1,\frac{1}{t})`$.
* $`\mathrm{\Delta }_{2,P}(t)`$, the number of positive non-unit clauses that are satisfied at stage $`t`$, has the binomial distribution $`B(P_{2,t},\frac{1}{t})`$.
###### Lemma 18.
For every $`c>0`$ and every $`t,nc\sqrt{n}tn`$, the conditional probability that the inequality
$$N_n(nt)\left[1+\frac{2(N_n1)}{t}\right]N_jN_n$$
(10)
holds for all $`tjn`$, in the event that $`\mathrm{PUR}`$ reaches stage $`t`$, is $`1o(1)`$.
###### Lemma 19.
Let $`X_n[0,n]`$ be the r.v. denoting the number of iterations of PUR on a random satisfiable formula $`\mathrm{\Phi }\mathrm{\Omega }(n,c2^n)`$. Then $`X_n`$ converges in distribution to a distribution $`\rho `$ on $`[0,n]`$ having support on the nonnegative integers, $`\rho =(\rho _k)_{k0}`$, $`\rho _k=Prob[\rho =k]`$, given by
$$\rho _k=\frac{e^{2^kc}}{1F(e^c)}\underset{i=1}{\overset{k1}{}}(1e^{2^ic}).$$
## 8 The proof of Theorem 4
Let $`c_1<c_2<c_3`$ be arbitrary constants. Consider three random formulas $`\mathrm{\Phi }_1\mathrm{\Omega }(n,𝐤(𝐧),c_1\frac{H_{k(n)}}{n})`$,$`\mathrm{\Phi }_2\mathrm{\Omega }(n,𝐧,c_22^n)`$ and $`\mathrm{\Phi }_3\mathrm{\Omega }(n,𝐤(𝐧),c_3\frac{H_{k(n)}}{n})`$, and let $`\mathrm{\Phi }^{}`$ be the subformula of $`\mathrm{\Phi }_2`$ consisting of the clauses of size at most $`k(n)`$. By the Chernoff bound, with high probability, $`m^{}`$, the number of clauses of $`\mathrm{\Phi }^{}`$, is in the interval $`[c_1\frac{H_{k(n)}}{n},c_3\frac{H_{k(n)}}{n}]`$. When $`n\mathrm{}`$ the probability that $`\mathrm{\Phi }_2\mathrm{HORN}\text{-}\mathrm{SAT}`$ tends to $`1F_1(e^{c_2})`$.
From Lemma 19 we infer the following easy consequence
###### Claim 1.
The probability that PUR accepts $`\mathrm{\Phi }_2`$ after stage $`nk(n)+1`$ is $`o(1)`$.
Since in the first $`k(n)1`$ stages of PUR only the clauses of $`\mathrm{\Phi }^{}`$ can influence the algorithm acceptance/rejection of $`\mathrm{\Phi }_2`$ (because PUR accepts/rejects at Stage $`i`$ based only on the unit clauses, and each non-simplified clause loses at most one literal at each phase),
$$|\mathrm{Pr}[\mathrm{\Phi }_2\mathrm{HORN}\text{-}\mathrm{SAT}]\mathrm{Pr}[\mathrm{\Phi }^{}\mathrm{HORN}\text{-}\mathrm{SAT}]|=o(1).$$
By the monotonicity of SAT and the randomness of $`\mathrm{\Phi }_1,\mathrm{\Phi }_2,\mathrm{\Phi }^{^{}}`$ we have
$$\mathrm{Pr}[\mathrm{\Phi }_1\mathrm{HORN}\text{-}\mathrm{SAT}]o(1)\mathrm{Pr}[\mathrm{\Phi }_2\mathrm{HORN}\text{-}\mathrm{SAT}]\mathrm{Pr}[\mathrm{\Phi }_3\mathrm{HORN}\text{-}\mathrm{SAT}]+o(1).$$
Taking limits it follows that
$`\overline{lim}_n\mathrm{}\underset{\mathrm{\Phi }\mathrm{\Omega }(n,k(n),c_1H_{k(n)}/n)}{\mathrm{Pr}}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}]`$ $`1F(e^{c_2})`$
$`\underset{¯}{lim}_n\mathrm{}\underset{\mathrm{\Phi }\mathrm{\Omega }(n,k(n),c_3H_{k(n)}/n)}{\mathrm{Pr}}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}].`$
Since $`c_1,c_2,c_3`$ were chosen arbitrarily, by choosing $`c_1=c,c_2=c+ϵ`$, and $`c_2=cϵ,c_3=c`$, respectively, we infer that
$`1F_1(e^{(cϵ)})\underset{¯}{lim}_n\mathrm{}\underset{\mathrm{\Phi }\mathrm{\Omega }(n,k(n),cH_{k(n)}/n)}{\mathrm{Pr}}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}]`$
$`\overline{lim}_n\mathrm{}\underset{\mathrm{\Phi }\mathrm{\Omega }(n,k(n),cH_{k(n)}/n)}{\mathrm{Pr}}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}]1F_1(e^{(c+ϵ)}).`$
As $`ϵ`$ is arbitrary, we get the desired result. $`\mathrm{}`$
###### Observation 2.
One point about the previous proof that is intuitively clear, but gets somewhat obscured by the technical details of the proof, is that if $`\mathrm{\Phi }_2\mathrm{\Omega }(n,𝐧,c_22^n)`$ then $`\mathrm{\Phi }^{^{}}`$ behaves “for every practical purpose” as if it were a uniform formula in $`\mathrm{\Omega }(n,𝐤(𝐧),c_2\frac{H_{k(n)}}{n})`$. We will use a similar intuition in the proof of Proposition 7.
## 9 The uniformity lemma
The following lemma is the analog of Lemma 17 for the case $`k=2`$, and the basis for our analysis of this case:
###### Lemma 20.
Suppose that $`\mathrm{\Phi }`$ survives up to stage $`t`$. Then, conditional on $`(P_{1,t},N_{1,t},P_{2,t},N_{2,t})`$, the clauses in $`\mathrm{\Phi }_{1,t}^P,\mathrm{\Phi }_{1,t}^N,\mathrm{\Phi }_{2,t}^P,\mathrm{\Phi }_{2,t}^N`$ are chosen uniformly at random and are independent. Also, conditional on the fact that $`\mathrm{\Phi }`$ survives stage $`t`$ as well, the following recurrences hold:
$$\{\begin{array}{c}P_{1,t1}=P_{1,t}1\mathrm{\Delta }_{1,t}^P+\mathrm{\Delta }_{12,t}^P,\hfill \\ N_{1,t1}=N_{1,t}+\mathrm{\Delta }_{12,t}^N,\hfill \\ P_{2,t1}=P_{2,t}\mathrm{\Delta }_{12,t}^P\mathrm{\Delta }_{02,t}^P,\hfill \\ N_{2,t1}=N_{2,t}\mathrm{\Delta }_{12,t}^N,\hfill \end{array}$$
(11)
where (in distribution)
$$\{\begin{array}{c}\mathrm{\Delta }_{1,t}^P=B(P_{1,t}1,1/t),\hfill \\ \mathrm{\Delta }_{12,t}^P=B(P_{2,t},1/t),\hfill \\ \mathrm{\Delta }_{02,t}^P=B(P_{2,t}\mathrm{\Delta }_{12,t}^P,1/t),\hfill \\ \mathrm{\Delta }_{12,t}^N=B(N_{2,t},2/t).\hfill \end{array}$$
(12)
Proof. A formula will be represented by an $`m\times 2`$ table. The rows of the table correspond to clauses in the formula and the entries are its literals. They are gradually unveiled as the algorithm proceeds. We assume that when generating $`\mathrm{\Phi }`$ we mark those clauses containing only one literal (so that we know their location, but not their content). We say that a row (or a clause) is “blocked” either if the clause is already satisfied or the clause has been turned into the empty clause. Suppose $`\mathrm{PUR}`$ arrives at stage $`t`$ on $`\mathrm{\Phi }`$. Then in stages $`i=n,n1,\mathrm{},t+1`$, $`\mathrm{\Phi }_i`$ should contain a unit clause consisting of a positive literal but should not have contained complementary unit clauses of the same variable. To carry out the disclosure at stage $`i`$, let $`x`$ be the variable set to one in this stage. We assume that the formula unveils all occurrences of $`x`$ or $`\overline{x}`$ in $`\mathrm{\Phi }`$. For each clause we perform the following:
1. if it contains $`x`$ we unveil all its literals and block;
2. otherwise we do nothing.
The clauses of $`\mathrm{\Phi }_t`$ having size two correspond to the rows of $`\mathrm{\Phi }`$ that contain no unveiled literal. The clauses of size one are either the clauses of size one in $`\mathrm{\Phi }`$ that contain none of the chosen literals, or the clauses of size two that contain the negation of one chosen variable and another is yet to be chosen. Given these observations the uniformity and independence follow from the way we construct $`\mathrm{\Phi }`$.
To prove the recurrences, let $`x`$ be the variable set to one in stage $`t`$ (it exists since PUR does not halt at this stage). By uniformity and independence, each of the $`P_{1,t}1`$ positive unit clauses of $`\mathrm{\Phi }_t`$, other than the chosen one, is equal to $`x`$ with probability $`1/t`$ (since there are $`t`$ variables left at this stage). On the other hand, the positive unit clauses of $`\mathrm{\Phi }_{t1}`$ that are not present in $`\mathrm{\Phi }_t`$ can only come from clauses of size two of $`\mathrm{\Phi }_t`$ that contain $`\overline{x}`$ and a positive literal (therefore counted by $`P_{2,t}`$). Uniformity and independence imply therefore that $`\mathrm{\Delta }_1^P(t)`$ has the distribution claimed in (12). The other relations can be justified similarly (noting that, since PUR does not reject at this stage, every negative unit clause of $`\mathrm{\Phi }_t`$ is also present in $`\mathrm{\Phi }_{t1}`$).
It will be useful to consider the Markov chain (11) for all values of $`t=n,\mathrm{},0`$ (even when the algorithm halts). To accomplish that, the “minus” signs in the first equation of (11) and the definition of $`\mathrm{\Delta }_{1,t}^P`$ should be replaced by $`\dot{}`$. We also need to specify the distribution of each component of the tuple $`(P_{1,n},N_{1,n},P_{2,n},N_{2,n})`$. Let $`\mathrm{\Delta }_n`$ be a random variable having the Bernoulli distribution $`B(cn,\frac{2n}{2n+3\left(\genfrac{}{}{0pt}{}{n}{2}\right)})`$. It is easy to see that in distribution
$$\{\begin{array}{c}P_{1,n}=B(\mathrm{\Delta }_n,1/2),\hfill \\ N_{1,n}=\mathrm{\Delta }_nP_{1,n},\hfill \\ P_{2,n}=B(cn\mathrm{\Delta }_n,2/3)\hfill \\ N_{2,n}=cn\mathrm{\Delta }_nP_{2,n}.\hfill \end{array}$$
(13)
$`\mathrm{}`$
## 10 Proof of Theorem 5
The main intuition for the proof is that in “most interesting stages” $`\mathrm{\Delta }_{1,t}^P=0`$ and $`\mathrm{\Delta }_{12,t}^P`$ is approximately Poisson distributed. Therefore, $`P_{1,t}`$ qualitatively behaves like the Markov Chain $`(Q_t)_t`$ defined by
$$\{\begin{array}{c}Q_{n+1}=1,\hfill \\ Q_{t1}=Q_t\dot{}1+Po(\lambda ),\hfill \end{array}$$
(14)
where $`\lambda =2c/3`$. This explains the closed form of the limit probability: a well-known result states that $`\rho `$, the probability that the queuing chain $`Q_t`$ reaches state 0, satisfies the equation $`\rho =\mathrm{\Phi }(\rho )`$, where $`\mathrm{\Phi }(t)=e^{\lambda (t1)}`$ is the generating function of the arrival distribution $`Po(\lambda )`$. We will define a suitable value $`\omega _0`$ such that:
1. With high probability PUR does not reject in any of stages $`n,\mathrm{},n\omega _0`$.
2. PUR accepts “mostly before or at stage $`n\omega _0`$” (i.e. the probability that PUR accepts after stage $`n\omega _0`$, given that $`\mathrm{\Phi }`$ survives this far is $`o(1)`$).
3. With high probability, for every $`tn,\mathrm{},n\omega _0`$, $`\mathrm{\Delta }_{1,t}^P=0`$.
4. At stages $`n,\mathrm{},n\omega _0`$, $`P_{1,t}`$ is “very close” to $`Q_t`$, with respect to total variation distance.
This program can be accomplished as described if $`c<3/2`$. To prove Property 4 we make use of Lemmas 13 and 16. Property 2 is proved only implicitly: in this case (see ) the probability that $`Q_i=0`$ for some $`i`$ tends to one, and, in fact, by a technical result due to Frieze and Suen (Lemma 3.1 in ), $`\mathrm{Pr}[Q_i=0\text{ for some }in\mathrm{log}n]`$ is $`1o(1)`$.
Let us now concentrate on the case when $`c>3/2`$ (the case when $`c=3/2`$ will follow by a monotonicity argument). In the previous argument we only used the fact that $`c<3/2`$ when deriving the probability that $`Q_t`$ hits state 0, hence the arguments from above carry on, and the conclusion is that the probability that PUR accepts at one of the stages $`n,\mathrm{},n\omega _0`$ differs by $`o(1)`$ from the probability that $`Q_t=0`$ somewhere in this range. We now, however, have to consider the probability that PUR accepts at some stage later than $`n\omega _0`$ and aim to prove that this probability is $`o(1)`$. It is conceptually simpler to divide the interval $`[n\omega _0,0]`$ into two subintervals, $`[n\omega _0,n\omega _1]`$ and its complement, such that w.h.p. $`\mathrm{\Phi }_{n\omega _1}`$ (if defined) contains two opposite unit clauses, therefore the probability that PUR accepts after stage $`n\omega _1`$ is $`o(1)`$. In the range $`[n\omega _0,n\omega _1]`$ we would like to prove that “most of the time” $`\mathrm{\Delta }_{1,t}^P`$ is zero and $`P_{1,t}`$ is “close” to $`Q_t`$ and to reduce the problem to the analysis of $`Q_t`$. Unfortunately there are two problems with this approach: although the probability that each individual $`\mathrm{\Delta }_{1,t}^P>0`$ is fairly small, to make $`\mathrm{\Phi }_{n\omega _1}`$ unsatisfiable w.h.p., $`\omega _1`$ has to be $`\omega (\sqrt{n})`$. This implies that we cannot sum these probabilities over $`[n\omega _0,n\omega _1]`$ and expect the sum to be $`o(1)`$; a similar problem arises if we want to sum the upper bounds for $`d_{TV}(\mathrm{\Delta }_{12,t}^P,Po(\lambda ))`$.
Fortunately there is a way to circumvent this, avoiding the use of total variation distance altogether: although we cannot guarantee that w.h.p. each $`\mathrm{\Delta }_{1,t}^P=0`$, we can arrange that w.h.p. for every sequence of $`p`$ consecutive stages $`t,t1,\mathrm{}tp+1`$, $`\mathrm{\Delta }_{1,t}^P+\mathrm{\Delta }_{1,t1}^P+\mathrm{}+\mathrm{\Delta }_{1,tp+1}^P3`$ (\*). Intuitively, in any sequence of $`p`$ consecutive steps at most $`p+3`$ clients leave the queue, and the number of those who arrive is the sum of $`p`$ approximately Poisson variables, thus approximately Poisson with parameter $`p\lambda `$. Choosing $`p`$ large enough so that $`\lambda >1+\frac{3}{p}`$ ensures that in any $`p`$ steps the average number of customers that arrive is strictly larger than the number of customers that are served in this time span. Therefore we will seek to approximate $`P_{1,t}`$ by a queuing chain $`\overline{Q}_t`$ with this property. Since $`P_{1,n\omega _0}=\overline{Q}_{n\omega _0}`$ is “large,” an elementary analysis of the queuing chain implies that the probability that $`\overline{Q}_t`$ hits state 0 in the interval $`[n\omega _0,n\omega _1]`$ is exponentially small. So we obtain the desired result if $`\overline{Q}_t`$ is constructed so that it is stochastically dominated by $`P_{1,t}`$.
### 10.1 The case $`c<3/2`$
Define $`\omega _0=n^{0.1}`$. The following are the main steps of the proof in this case:
###### Lemma 21.
With probability $`1o(1/poly)`$ for every $`t[n,\mathrm{},n/2]`$ we have
$$\mathrm{\Delta }_{12,t}^P,\mathrm{\Delta }_{02,t}^P,\mathrm{\Delta }_{12,t}^N\frac{1}{2}n^{0.1}.$$
Proof. Use the coupling with $`m_1=P_{2,t}(N_{2,t})`$, $`m_2=cn`$, $`\tau =1/t`$, and apply Chernoff bound to $`B(cn,1/t)`$. $`\mathrm{}`$
###### Corollary 1.
Consider $`\omega n/2`$. If for every $`t[n,\mathrm{},n/2]`$, $`\mathrm{\Delta }_{12,t}^P,\mathrm{\Delta }_{02,t}^P,\mathrm{\Delta }_{12,t}^N\frac{1}{2}n^{0.1}`$ then, for all $`t[n,\mathrm{},n\omega ]`$, $`P_{1,t},N_{1,t},|P_{2,t}P_{2,n}|,|N_{2,t}N_{2,n}|<(nt)n^{0.1}`$.
###### Lemma 22.
If for all $`t[n,\mathrm{},n\omega ]`$, $`P_{1,t},N_{1,t},|P_{2,t}P_{2,n}|,|N_{2,t}N_{2,n}|<(nt)n^{0.1}`$ holds then w.h.p. $`\mathrm{\Delta }_{1,t}^P=0`$ for every $`t[n,\mathrm{},n\omega _0]`$.
Proof.$`\mathrm{Pr}[B(P_{1,t}1,\frac{1}{t})>0]=1\mathrm{Pr}[B(P_{1,t}1,\frac{1}{t})=0]=1(1\frac{1}{t})^{P_{1,t}1}<\frac{P_{1,t}1}{t}<n^{0.9}`$. $`\mathrm{}`$
###### Lemma 23.
W.h.p., $`|P_{2,n}\frac{2}{3}cn|,|N_{2,n}\frac{1}{3}cn|<n^{0.6}`$.
Proof. Directly from the Chernoff bounds on $`\mathrm{\Delta }_n`$ and $`P_{2,n}`$. $`\mathrm{}`$
###### Lemma 24.
If the events in the conclusions of Lemmas 1 and 23 hold for $`\omega =\omega _0`$, $`ϵ_1=1/6`$ and $`ϵ_2=0.1`$, then there exists a constant $`r>0`$ such that for every $`t=n,\mathrm{},n\omega _0`$, $`|\frac{P_{2,t}}{t}\frac{2}{3}c|rn^{0.4}`$.
Proof. We have
$$|\frac{P_{2,t}}{t}\frac{2}{3}c|P_{2,t}\left|\frac{1}{t}\frac{1}{n}\right|+\frac{|P_{2,t}P_{2,n}|}{n}+\left|\frac{P_{2,n}}{n}\frac{2}{3}c\right|P_{2,n}\frac{\omega _0}{n(n\omega _0)}+\frac{n^{0.2}}{n}+n^{0.61},$$
by Lemma 24 and $`n\omega _0tn,`$ and the result immediately follows. $`\mathrm{}`$
###### Lemma 25.
If the conclusions of Lemmas 24 and 22 are true then
$$\underset{t=n\omega _0}{\overset{n}{}}d_{TV}(P_{1,t},Q_t)=o(1/\omega _0).$$
Proof. By Lemma 24 and the inequalities on total variation distance there exist $`r_1,r_2>0`$ such that
$`d_{TV}(\mathrm{\Delta }_{12,t}^P,Po(\lambda ))`$ $``$ $`d_{TV}(\mathrm{\Delta }_{12,t}^P,Po\left({\displaystyle \frac{P_{2,t}}{t}}\right))+d_{TV}(Po\left({\displaystyle \frac{P_{2,t}}{t}}\right),Po\left({\displaystyle \frac{2}{3}}c\right))`$
$``$ $`r_1{\displaystyle \frac{1}{t}}+r_2n^{0.4}r_3n^{0.4},`$
where $`r_3=r_1+r_2`$. Employing Lemma 14 it follows that
$$\underset{t=n\omega _0}{\overset{n}{}}d_{TV}(P_{1,t},Q_t)r_3\underset{t=n\omega _0}{\overset{n}{}}tn^{0.4}r_3n^{0.4}\frac{\omega _0^2}{2},$$
and this amount is $`o(1/\omega _0)`$. $`\mathrm{}`$
###### Observation 3.
The probability that the conditions in the previous lemma are not fulfilled is at most $`\omega _0^4/n=n^{0.6}`$. Indeed, the events that ensure the applicability of the previous lemma are:
1. for every $`t[n,\mathrm{},n/2]`$, $`\mathrm{\Delta }_{12,t}^P,\mathrm{\Delta }_{02,t}^P,\mathrm{\Delta }_{12,t}^N\frac{1}{2}n^{0.1}`$,
2. for all $`t[n,\mathrm{},n\omega _0]`$, $`\mathrm{\Delta }_{1,t}^P=0`$, and
3. $`|P_{2,n}\frac{2}{3}cn|,|N_{2,n}\frac{1}{3}cn|,<n^{0.6}`$
The first and the third events have probability $`1o(1/poly)`$ (as they come from applying Chernoff bounds). The second fails (for a specific $`t`$) with probability at most $`\frac{P_{1,t}}{nt}\omega _0^2/(n\omega _0)`$, so its total probability is at most $`\omega _0\omega _0^2/(n\omega _0)`$. Both terms can be absorbed into $`\omega _0^4/n`$.
###### Lemma 26.
If the event in Lemma 1 holds then w.h.p. PUR does not reject at stage $`t`$, for every $`t`$ in the range $`n`$, $`n1,\mathrm{},n\omega _0`$, given that $`\mathrm{\Phi }`$ survives up to this stage.
Proof. To prove Lemma 26 we show that, with high probability the unit clauses of each $`\mathrm{\Phi }_t`$ involve different variables. This can be seen as follows: consider $`P_{1,t}+N_{1,t}`$ balls to be thrown into $`t`$ urns. The probability that two of them arrive in the same urn is at most $`\left(\genfrac{}{}{0pt}{}{P_{1,t}+N_{1,t}}{2}\right)\frac{1}{t}`$. This is upper bounded by $`\frac{(\omega _0n^{0.1})^2}{2(n\omega _0)}`$. Summing this for $`t=n,\mathrm{},n\omega _0`$ yields an upper bound, which is $`o(1)`$. $`\mathrm{}`$
The proof for the case $`c<3/2`$ follows easily from these results: with probability $`1o(1)`$ all the events in Lemmas 21, 1, 22, 23, 25, and 26 take place, therefore PUR does not reject at any of the stages $`n`$ to $`n\omega _0`$ and $`P_{1,t}`$ is close to $`Q_t`$ in the sense of Lemma 25. Therefore the probability that for some $`t`$ in this range $`P_{1,t}=0`$ (i.e. PUR accepts) differs by $`o(1)`$ from the corresponding probability for $`Q_t`$. But according to the result by Frieze and Suen this latter probability is $`1o(1)`$.
### 10.2 The case $`c>3/2`$
Define $`\omega _1=n^{0.51}`$. The following are the auxiliary results we use in this case:
###### Lemma 27.
Let $`A=n^{0.61}`$. For every $`k>0`$ there exists a constant $`c_k>0`$ such that for every $`r>0`$ the probability that there exists $`t[n\omega _0,n\omega _1]`$, $`\mathrm{\Delta }_{1,t}^P+\mathrm{\Delta }_{1,t1}^P+\mathrm{}+\mathrm{\Delta }_{1,tr+1}^Pk`$ is at most $`c_k(\omega _1\omega _0)(rA/n)^k`$.
Proof. By Corollary 1 we can assume that $`P_{1,t}A`$. Then for every $`i`$,
$$\mathrm{Pr}[\mathrm{\Delta }_{1,t}^Pi]=\mathrm{Pr}[B(P_{1,t}1,\frac{1}{t})i]\mathrm{Pr}[B(A,\frac{1}{t})i]$$
$$=1\underset{j=1}{\overset{i1}{}}\left(\genfrac{}{}{0pt}{}{A}{j}\right)\left(\frac{1}{t}\right)^j\left(1\frac{1}{t}\right)^{Aj}$$
$$\left(\genfrac{}{}{0pt}{}{A}{i}\right)(\frac{1}{t})^i$$
The event $`\mathrm{\Delta }_{1,t}^P+\mathrm{\Delta }_{1,t1}^P+\mathrm{}+\mathrm{\Delta }_{1,tr+1}^Pk`$ happens when:
* one of the factors is at least $`k`$, or
* one of the factors is at least $`k1`$, and another one is at least 1, or
*
* at least $`k`$ of the factors are at least one.
(a finite number of possibilities). Applying the previous inequality, and taking into account that $`r,k`$ are fixed immediately proves the lemma.
To flesh out the argument outlined before we construct a succession of Markov chains running along $`P_{1,t}`$, that provide better and better “approximations” to $`\overline{Q}_t`$. Our use of indices will be slightly nonstandard (to reflect the connection with $`P_{1,t}`$), in that the sequence of indices starts with $`n\omega _0`$ and is decreasing.
###### Definition 28.
Let $`X_{n\omega _0}=Y_{n\omega _0}=Z_{n\omega _0}=\overline{Q}_{n\omega _0}=P_{1,n\omega _0}`$ and
$$\{\begin{array}{c}X_{t1}=X_t(p+3)\chi _{p𝐙+1}(n\omega _0t)+\mathrm{\Delta }_{12,t}^P,\hfill \\ Y_{t1}=Y_t(p+3)\chi _{p𝐙+1}(n\omega _0t)+B(P_{2,n\omega _1},1/t),\hfill \\ Z_{t1}=Z_t(p+3)\chi _{p𝐙+1}(n\omega _0t)+B(P_{2,n\omega _1},\frac{1}{n}),\hfill \\ \overline{Q}_{t1}=\overline{Q}_{t1}1+B(p\frac{P_{2,n\omega _1}}{p+3},\frac{1}{n}).\hfill \end{array}$$
(15)
Let $`c=\mathrm{Pr}[(t[n\omega _0,n\omega _1]):P_{1,t}=0]`$. Note that the amount $`p+3`$ is subtracted from $`X_t,Y_t,Z_t`$ exactly once in every $`p`$ consecutive steps, so whenever the condition (\*) is satisfied it holds that $`X_tP_{1,t}`$ for every $`t[n\omega _0,n\omega _1]`$. By coupling $`\mathrm{\Delta }_{12,t}^P(=B(P_{2,t},1/t))`$ with $`B(P_{2,n\omega _1},1/t)`$ we deduce that we can couple $`X_t`$ and $`Y_t`$ so that $`Y_tX_t`$. We can also couple $`Y_t`$ and $`Z_t`$ such that $`Z_tY_t`$. Finally, notice that we can couple $`Z_{n\omega _0jp}`$ and $`\overline{Q}_{n\omega _0j(p+3)}`$ such that $`\overline{Q}_{n\omega _0j(p+3)}Z_{n\omega _0jp}`$. So an upper bound on $`\alpha `$ is $`\mathrm{Pr}[(t[0,n\omega _0]):\overline{Q}_t=0]`$. With high probability the Bernoulli distribution in the definition of the chain $`\overline{Q}_t`$ has the average strictly greater than one, (because the flow from $`P_{2,t}`$ is approximately Poisson), and $`\overline{Q}_{n\omega _0}=\mathrm{\Omega }(\omega _0)`$, therefore, by an elementary property of the queuing chain, the probability that $`\overline{Q}_t`$ hits state 0 is exponentially small. This yields the desired conclusion, that $`\alpha =o(1)`$.
One word about the way to prove the fact that $`\mathrm{\Phi }_{n\omega _1}`$ is unsatisfiable (if defined): one can prove that w.h.p. both $`P_{1,n\omega _1}`$ and $`N_{1,n\omega _1}`$ are $`\mathrm{\Omega }(\omega _1)`$. By the uniformity lemma 20 we are left with the following instance of the occupancy problem: there are $`P_{1,n\omega _1}`$ white balls, $`N_{1,n\omega _1}`$ black balls and $`n\omega _1`$ bins. The desired fact now follows from the second part of Lemma 16.
## 11 Proof of Theorem 6
Theorem 6 is proved along lines very similar to the proof of Theorem 5. The basis is the following generalization of Lemma 20:
###### Lemma 29.
Suppose that $`\mathrm{\Phi }`$ survives up to stage $`t`$. Then, conditional on the values $`(P_{1,t},N_{1,t},\mathrm{},P_{k,t},N_{k,t})`$, the clauses in $`\mathrm{\Phi }_{1,t}^P,\mathrm{\Phi }_{1,t}^N,\mathrm{},\mathrm{\Phi }_{k,t}^P,\mathrm{\Phi }_{k,t}^N`$ are chosen uniformly at random and are independent. Also, conditional on the fact that $`\mathrm{\Phi }`$ survives stage $`t`$ as well, the following recurrences hold:
$$\{\begin{array}{c}P_{1,t1}=P_{1,t}1\mathrm{\Delta }_{01,t}^P+\mathrm{\Delta }_{12,t}^P,\hfill \\ N_{1,t1}=N_{1,t}+\mathrm{\Delta }_{12,t}^N,\hfill \\ P_{i,t1}=P_{i,t}\mathrm{\Delta }_{0i,t}^P\mathrm{\Delta }_{(i1)i,t}^P+\mathrm{\Delta }_{i(i+1),t}^P\text{, for }i=\overline{2,k},\hfill \\ N_{i,t1}=N_{i,t}\mathrm{\Delta }_{(i1)i,t}^N+\mathrm{\Delta }_{i(i+1),t}^N\text{, for }i=\overline{2,k},\hfill \end{array}$$
(16)
where (in distribution)
$$\{\begin{array}{c}\mathrm{\Delta }_{01,t}^P=B(P_{1,t}1,1/t),\hfill \\ \mathrm{\Delta }_{(i1)i,t}^P=B(P_{i,t},(i1)/t),\hfill \\ \mathrm{\Delta }_{0i,t}^P=B(P_{i,t}\mathrm{\Delta }_{(i1)i,t}^P,1/t),\hfill \\ \mathrm{\Delta }_{(i1)i,t}^N=B(N_{i,t},i/t),\hfill \\ \mathrm{\Delta }_{k(k+1),t}^P=\mathrm{\Delta }_{k(k+1),t}^N=0.\hfill \end{array}$$
(17)
Proof.
The uniformity condition and the justification of the recurrences are absolutely similar to the ones from Lemma 5. The additional technical complication is that now there is a “positive flow into $`P_{2,t},N_{2,t}`$.” $`\mathrm{}`$
###### Lemma 30.
With high probability it holds that
$$P_{i,t}=(1+o(1))\frac{c}{n}\lambda _ki\left(\genfrac{}{}{0pt}{}{t}{i}\right)S_{ki}^{n+1t},$$
and
$$N_{i,t}=(1+o(1))\frac{c}{n}\lambda _k\left(\genfrac{}{}{0pt}{}{t}{i}\right)S_{ki}^{n+1t},$$
for every $`i2`$, and uniformly on $`t=no(n)`$.
Proof.
Let us first heuristically derive a formula for $`x_{i,t}`$, $`y_{i,t}`$, the expected values of $`P_{i,t}`$, $`N_{i,t}`$, obtained by replacing the binomial distributions in the equations by their expected values.
We have:
$$\{\begin{array}{c}x_{i,t1}=x_{i,t}\frac{x_{i,t}}{t}\frac{(i1)x_{i,t}}{t}+\frac{ix_{i+1,t}}{t}\text{, for }i=\overline{2,k},\hfill \\ y_{i,t1}=y_{i,t}\frac{iy_{i,t}}{t}+\frac{(i+1)y_{(i+1),t}}{t}\text{, for }i=\overline{2,k},\hfill \end{array}$$
(18)
Rearranging terms the recurrences become
$$\{\begin{array}{c}x_{i,t1}=x_{i,t}(1\frac{i}{t})+x_{i+1,t}\frac{i}{t}\text{, for }i=\overline{2,k},\hfill \\ y_{i,t1}=y_{i,t}(1\frac{i}{t})+y_{(i+1),t}\frac{(i+1)}{t}\text{, for }i=\overline{2,k}.\hfill \end{array}$$
(19)
Also,
$$\{\begin{array}{c}x_{i,n}=\frac{i\left(\genfrac{}{}{0pt}{}{n}{i}\right)}{H_k}c\lambda _k\frac{H_k}{n}=\frac{c}{n}\lambda _ki\left(\genfrac{}{}{0pt}{}{n}{i}\right),\hfill \\ y_{i,n}=\frac{\left(\genfrac{}{}{0pt}{}{n}{i}\right)}{H_k}c\lambda _k\frac{H_k}{n}=\frac{c}{n}\lambda _k\left(\genfrac{}{}{0pt}{}{n}{i}\right).\hfill \end{array}$$
(20)
A simple induction shows that these expected values are $`x_{i,t}=\frac{c}{n}\lambda _ki\left(\genfrac{}{}{0pt}{}{t}{i}\right)S_{ki}^{n+1t}`$, and $`y_{i,t}=\frac{c}{n}\lambda _k\left(\genfrac{}{}{0pt}{}{t}{i}\right)S_{ki}^{n+1t}`$.
The concentration property can be proved inductively, starting from $`i=k`$ towards $`3`$, by noting that the expected values of the binomial terms in the recurrence are $`\omega (n)`$, hence, by the Chernoff bound, the probabilities that they significantly deviate from their expected values is exponentially small.
Almost the same argument holds for $`P_{2,t}`$ and for $`N_{2,t})`$. The only amounts to be handled differently are “the clause flows out of $`P_{2,t},N_{2,t}`$,” but they are approximately Poisson distributed, hence “small” with high probability by Proposition 10. Therefore $`P_{2,t}=(1+o(1))\frac{c}{n}\lambda _k2\left(\genfrac{}{}{0pt}{}{t}{2}\right)S_{k2}^{n+1t}`$. $`\mathrm{}`$
The previous lemma implies that $`\mathrm{\Delta }_{2,t}^PPo(c\lambda _kS_{k2}^{n+1t})`$ (for $`t=no(n)`$); thus in this range $`P_{1,t1}P_{1,t}1+Po(c\lambda _kS_{k2}^{n+1t})`$. The proof follows exactly the same pattern as in the case $`c<3/2`$ for $`k=2`$: the conclusion for the stages $`[n,n\omega _0]`$ is that the probability that $`P_{1.t}`$ is zero somewhere in this range differs by $`o(1)`$ from the corresponding probability for the queuing chain in (4). The fact that the stages after $`[n,n\omega _0]`$ have a contribution of $`o(1)`$ to the final accepting probability can be seen by the fact that there is possible to couple the Markov $`M_1`$, describing the evolution of PUR on a random $`k`$-SAT formula, and $`M_2`$ that runs on the 2-CNF component of the formula, such that for every $`t`$ we have $`P_{1,t}^{M_2}P_{1,t}^{M_1}`$. Perhaps the most intuitive way to see this coupling is to “paint” the initial clauses of the formula having size at most two in red, and the other clauses in blue. At every step $`t`$ $`P_{1,t}^{M_2}`$ will count only red clauses having unit size at step $`t`$, while $`P_{1,t}^{M_1}`$ will count clauses of both colors.
Given the stochastic domination, the desired result follows from the corresponding proof in the case $`k=2`$. $`\mathrm{}`$
## 12 Proof of Proposition 7
The idea of the proof is to consider PUR on a random at-most-$`k`$-Horn formula $`\mathrm{\Phi }`$ with $`\widehat{c}\frac{H_k}{n}`$ clauses and prove that there exists a function $`\varphi (k)`$ with $`lim_k\mathrm{}\varphi (k)=0`$ such that
$$\underset{n\mathrm{}}{lim}\mathrm{Pr}[\mathrm{PUR}\text{ accepts in at least }k\text{ steps }]\varphi (k).$$
Indeed, from the previous proof it follows that $`lim_n\mathrm{}\mathrm{Pr}[\mathrm{PUR}\text{ accepts in }k\text{ steps }]`$ satisfies the recurrence:
$$x_{t+1}=x_{1,t}1+Po(\widehat{c}S_{k2}^{t+1+k}),$$
where
$$x_0=P_{1,k}1.$$
We define $`\varphi (k)`$ to be the probability that the sequence in the recurrence (12) hits zero. Trivially $`lim_k\mathrm{}S_{k2}^{k+1}=\mathrm{}`$, so the expected values of the Poisson distributions in (12) can be made larger than any given constant $`\lambda `$. Using the fact that the sum of two Poisson distributions with parameters $`a`$ and $`b`$ has a Poisson distribution with parameter $`a+b`$ it follows that, for large enough $`k`$, one can couple $`x_t`$ with the queuing chain
$$y_{t+1}=y_{1,t}1+Po(\lambda ),$$
$$y_0=1,$$
such that $`y_tx_t`$. It follows that, for large $`k`$, $`\varphi (k)\mathrm{Pr}[\text{ the chain }y_t\text{ hits state zero}]`$. Since $`\lambda `$ was arbitrary, it follows that $`lim_k\mathrm{}\varphi (k)=0`$.
Now consider a random uniform Horn formula $`\mathrm{\Phi }`$ with $`\widehat{c}\frac{H_n}{n}`$ clauses, and let $`\overline{\mathrm{\Phi }}`$ be its subformula consisting of clauses of size at most $`k`$. It is easily seen that the behavior of PUR on the first $`k1`$ steps depends only on the clauses of $`\overline{\mathrm{\Phi }}`$, so
$$\mathrm{Pr}[\mathrm{PUR}\text{ accepts }\mathrm{\Phi }\text{ in less than }k\text{ steps}]=\mathrm{Pr}[\mathrm{PUR}\text{ accepts }\overline{\mathrm{\Phi }}\text{ in less than }k\text{ steps}].$$
On the other hand we have
$$0\mathrm{Pr}[\mathrm{PUR}\text{ accepts }\mathrm{\Phi }\text{ in at least }k\text{ steps}]\mathrm{Pr}[\mathrm{PUR}\text{ accepts }\overline{\mathrm{\Phi }}\text{ in at least }k\text{ steps}].$$
The fact that “$`\overline{\mathrm{\Phi }}`$ is close to a random formula in $`\mathrm{\Omega }(n,k,c\frac{H_k}{n})`$” (see the discussion in Observation 2) implies that the right-hand side term can be made less than any fixed constant $`ϵ`$ (for $`n,k`$ big enough). It follows that
$$|\mathrm{Pr}[\mathrm{PUR}\text{ accepts }\mathrm{\Phi }]\mathrm{Pr}[\mathrm{PUR}\text{ accepts }\overline{\mathrm{\Phi }}]|2ϵ,$$
for large enough values of $`n,k`$. This immediately implies the desired result. $`\mathrm{}`$
## 13 Proof of Theorem 8
Theorem 8 is based on the proof of the Theorem 5 and an elementary property of the queuing chain $`Q_t`$ (the expected time to hit state zero, conditional on actually hitting it has the desired form).
The crucial point is to prove that the probabilities that any of the conditions we have employed in our analysis fails have a negligible effect on the running time.
This is easy to see for stages smaller than $`n\omega _0`$: since the probabilities that the various steps of the analysis are either exponentially small or can be made $`o(1/n)`$ (by choosing a large enough $`k`$ in Lemma 27, the probability that $`P_{1,t}`$ hits state zero after stage $`n\omega _0`$ is $`o(1/n)`$, therefore its influence on the average running time of PUR is $`o(1)`$. The corresponding observation is not true for stages before $`n\omega _0`$, but these stages can be handled directly, using the statement from Lemma 25.
$`\mathrm{}`$
## 14 Random Horn satisfiability as a mean-field approximation
What we have shown so far is to prove that (under a suitably rescaled picture) the rescaled probability graphs for random at-most-$`k`$ Horn satisfiability converge to the graph for random Horn satisfiability. To be able to argue that our results display critical behavior, we have to be able to show that this latter probability $`p_{\mathrm{}}`$, is indeed the one predicted by some mean-field approximation.
In the sequel we will show that this is indeed the case. However the mean-field approximation is not the one from , and incorporates a correction specific to the properties of random Horn satisfiability.
Let us first see that it is not accurate if no correction is taken into account. Indeed, were it true we would have
$$\underset{n\mathrm{}}{lim}Pr[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}]=1\underset{n\mathrm{}}{lim}\underset{A\{0,1\}^n}{}\left(1\mathrm{Pr}[A\mathrm{\Phi }]\right).$$
Since, for an assignment $`A`$ of Hamming weight $`i`$ there are exactly $`2^i1+(ni)2^i`$ Horn clauses that $`A`$ falsifies, we have
$$\mathrm{Pr}[A\mathrm{\Phi }]=\left(1\frac{2^i1+(ni)2^i}{(n+2)2^n1}\right)^{c2^n},$$
so the mean-field prediction reads
$$\underset{n\mathrm{}}{lim}\mathrm{Pr}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}]=1\underset{n\mathrm{}}{lim}\underset{j=0}{\overset{n}{}}\left(1\left(1\frac{2^j1+(nj)2^j}{(n+2)2^n1}\right)^{c2^n}\right)^{\left(\genfrac{}{}{0pt}{}{n}{j}\right)}.$$
All terms in the product are less than 1. Since the term corresponding to $`j=1`$ is $`\left(1\left(1\frac{1+2(n1)}{(n+2)2^n1}\right)^{c2^n}\right)^n`$ has limit 0, the mean-field prediction would imply that $`lim_n\mathrm{}\mathrm{Pr}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}]=1`$. On the other hand let us observe that, if we do not consider the power $`\left(\genfrac{}{}{0pt}{}{n}{j}\right)`$ in the infinite product we obtain the right result: it is a simple but tedious task to prove that
$$\underset{n\mathrm{}}{lim}\underset{j=0}{\overset{n}{}}\left(1\left(1\frac{2^j1+(nj)2^j}{(n+2)2^n1}\right)^{c2^n}\right)=\underset{j=0}{\overset{\mathrm{}}{}}\left(1e^{c2^j}\right).$$
Intuitively this means that “there exist a correction of the mean-field approximation that only considers a single assignment of each weight, and is accurate.” The following simple result gives a precise statement to the above intuition:
###### Lemma 31.
Suppose $`\mathrm{\Phi }`$ is given as a union of formulas $`\mathrm{\Phi }_1,\mathrm{},\mathrm{\Phi }_n`$, where $`\mathrm{\Phi }_i`$ contains all clauses of length exactly $`i`$. Then there is a set $`T=\{T_0,\mathrm{},T_{n1}\}`$ of assignments, with $`T_i`$ of Hamming weight exactly $`i`$ and depending only on $`\mathrm{\Phi }_1\mathrm{}\mathrm{\Phi }_{i+1}`$, such that $`\mathrm{\Phi }`$ is satisfiable if and only if it is satisfied by some assignment in $`T`$.
Proof.
Let $`\overline{y_1\mathrm{}y_k}`$ denote the assignment that makes $`y_1=\mathrm{}=y_k=1`$, and all the other variables equal to zero.
The set $`T`$ has two parts: the first is simply the set of assignments implicitly examined by the algorithm PUR in testing satisfiability. That is, if $`x_1,\mathrm{},x_k`$ are the variables assigned by PUR in this order, the first part includes the assignments $`00000`$, $`\overline{x_1},\mathrm{},\overline{x_1,\mathrm{},x_k}`$. The second part contains a random assignment for each remaining weight. $`\mathrm{}`$
The result has a “mean-field” interpretation: as before, define $`f(x_1,\mathrm{},x_n)=1_{i=1}^nx_i`$, and the function $`g_k[\mathrm{\Phi }]`$ to be the indicator function for the event “$`T_k\overline{)}\mathrm{\Phi }`$, given that event $`\overline{A}_n\mathrm{}\overline{A}_{nk+1}`$ happens,” i.e.
$$g_k[\mathrm{\Phi }]=\frac{1}{\mathrm{Pr}[\overline{A}_n\mathrm{}\overline{A}_{nk+1}]}\{\begin{array}{cc}1,\hfill & \text{ if }T_k\vDash ̸\mathrm{\Phi }\overline{A}_n\mathrm{}\overline{A}_{nk+1}\hfill \\ 0,\hfill & \text{ otherwise.}\hfill \end{array}$$
We have
$$E[g_k[\mathrm{\Phi }]]=\mathrm{Pr}[\overline{A_{nk}}|\overline{A}_n\mathrm{}\overline{A}_{nk+1}].$$
Indeed, $`g_k[\mathrm{\Phi }]0`$ exactly when $`R_n\mathrm{}R_{nk+1}`$ or $`T_k\vDash ̸\mathrm{\Phi }S_n\mathrm{}S_{nk+1}`$. The second event is equivalent to $`\overline{A_{nk}}S_n\mathrm{}S_{nk+1}`$, hence we have $`g_k[\mathrm{\Phi }]0`$ exactly when $`\overline{A_{nk}}\overline{A}_n\mathrm{}\overline{A}_{nk+1}`$ holds.
Thus we have, by the discussion in the previous chapter,
$$f(E[g_1[\mathrm{\Phi }]],\mathrm{},E[g_n[\mathrm{\Phi }]])=1\underset{k=0}{\overset{n}{}}\mathrm{Pr}[\overline{A_{nk}}|\overline{A}_n\mathrm{}\overline{A}_{nk+1}]=\mathrm{Pr}[\mathrm{\Phi }\mathrm{HORN}\text{-}\mathrm{SAT}].$$
The above correction seems to be specific to the random model for Horn satisfiability, which allows clauses of varying lengths.
To sum up: the mean-field approximation is true, modulo a correction that takes into account some particular features of the random model for Horn satisfiability.
## 15 Discussion
We have characterized the asymptotical satisfiability probability of a random $`k`$-Horn formula, and showed that it exhibits very similar behavior to the one uncovered experimentally in .
We have also displayed an “easy-hard-easy” pattern similar to the ones observed experimentally in the AI literature. In our case the pattern is fully explained by elementary properties of the queuing chain.
As for an explanation of the “critical behavior”, consider an intermediate stage $`i`$ of PUR and let $`C_j`$ be the set of clauses of $`\mathrm{\Phi }_{i,j}^P`$. It is clear that whether PUR accepts is dependent only on the number of clauses in $`C_1`$. The restriction on the clause length acts like a “dampening” perturbation (in that it eliminates the “clause flow into $`C_k`$”). The proof of Theorem 4 states that when $`k(n)\mathrm{}`$, with high probability PUR accepts (if $`\mathrm{\Phi }`$ is satisfiable) “before the perturbation reaches $`C_1`$”, therefore the satisfiability probability is the one from the uniform case. On the other hand, for any constant $`k`$, with probability greater than 0 PUR does not halt during the first $`k`$ iterations (for the exact value see ), and the dampening has a significant influence. Thus the explanation for the occurrence (and specific form of) critical behavior is a threshold property for the number of iterations of PUR on random satisfiable Horn formulas “in the critical region”.
A related, and somewhat controversial, open issue is whether random Horn satisfiability properly displays critical behavior. Problems with a sharp threshold display “critical” (i.e singular) behavior at least in one parameter, the satisfaction probability, which conceivably allows the definition of critical exponents. This is not so for random $`k`$-Horn satisfiability, that has a coarse threshold, and no criticality for $`k>2`$, hence the question seems not to be meaningful. Note, however, that the order parameter involved in the recent study of the phase transition in 2-SAT is not satisfaction probability, but the (expected size) of the so-called backbone (or its more tractable version spine) of a random formula. The “window” that we use to peek at the threshold behavior of random Horn satisfiability does not seem to be “naturally required” by any physical considerations, and it is possible in principle that the random Horn formulas display critical behavior if we take the spine as the order parameter.
## 16 Acknowledgments
This paper is part of the author’s Ph.D. thesis at the University of Rochester. Support for this work has come from the NSF CAREER Award CCR-9701911 and the NSF Grant 9725021.
|
warning/0007/cond-mat0007076.html
|
ar5iv
|
text
|
# Statistics of Velocity Gradients in Two-Dimensional Navier-Stokes and Ocean Turbulence
## I Introduction
Statistical properties of turbulent flows, such as probability density functions (pdfs), are important for characterizing turbulence. For instance, velocity gradients are directly related to velocity correlations, relative dispersion, and energy dissipation in the fluid Monin and Yaglom (1975). This study evaluates statistics of turbulence, as observed in recent satellite measurements of the upper ocean. Statistics of observed phenomena are compared with corresponding statistics for the forced two-dimensional Navier-Stokes equations. Our results show that, in comparison with unforced decaying turbulence, simple forced two-dimensional Navier-Stokes equations provide better agreement with ocean observations.
For this analysis, ocean velocities are derived from altimeter data collected by the TOPEX/POSEIDON satellite, which performs repeated measurements of the height $`\eta `$ of the ocean surface. We use only observations from the TOPEX altimeter, which has lower noise levels than the POSEIDON instrument. The geostrophic relation, $`v_x=(g/f)\eta /y`$, yields the velocity component perpendicular to the satellite ground track. Surface geostrophic velocities are characteristic of sub-surface flow in the ocean Wunsch and Stammer (1998). This geostrophic flow is typically well-represented by two-dimensional shallow-water equations and resembles two-dimensional turbulence Rhines (1979); McWilliams (1984). The derivative along the satellite track, $`_yv_x`$, yields the transverse velocity gradient. We compute velocities $`v`$ from consecutive high-frequency altimeter measurements Yale et al. (1995); Stewart et al. (1996); Llewellyn Smith and Gille (1998); Gille and Llewellyn Smith (2000), and then determine velocity gradients by computing along track differences over a distance of 12 km. For comparison, the first baroclinic Rossby radius ranges between 10 km and 80 km between 60 and 10 latitude Houry et al. (1987); Stammer (1997), so transverse gradients over 12 km distance are expected to be representative of mesoscale geostrophic motions. The cross-track, or longitudinal, derivative cannot be determined. Higher order derivatives are increasingly noisy.
Earlier results have shown that velocities typically have Gaussian distributions within small regions of the ocean. When satellite data from the global ocean were combined, the resulting pdfs were non-Gaussian, due to regional variations in velocity variance Llewellyn Smith and Gille (1998); Gille and Llewellyn Smith (2000). When velocities were normalized by their local variances, the pdfs were Gaussian Gille and Llewellyn Smith (2000), at least for well-sampled velocities within three standard deviations of the mean. Similar results were obtained for subsurface floats deployed in the North Atlantic Ocean, although analysis for velocities more than three standard deviations from the mean indicated non-Gaussian tails Bracco et al. (2000). The Lagrangian statistics of floats are however not directly comparable to the results from the TOPEX altimeter, which captures the Eulerian statistics. In this study, we specifically normalize velocities and velocity gradients by their local variances before computing pdfs and other statistics.
We compare observed oceanic pdfs with simulations of two-dimensional quasi-geostrophic flow. The equations of motion are
$$\left(\frac{}{t}+\frac{\psi }{y}\frac{}{x}\frac{\psi }{x}\frac{}{y}\right)q=D^2q+F,$$
(1)
where the potential vorticity $`q=^2\psi +\psi /R^2`$. The second term in the potential vorticity is neglected both in the quasi-geostrophic limit of the shallow water equations and when the Rossby radius $`R`$ is large in the homogeneous quasi-geostrophic equations. In either case, eq. (1) is equivalent to two-dimensional Navier-Stokes flow. In this study, we perform simulations of the Navier-Stokes equations. Rapidly varying (white-in-time) forcing $`F`$ is applied on large scales, through random stirring of the low vorticity modes. This forcing resembles wind forcing of the ocean, which varies rapidly in time but slowly in space Large et al. (1991); Wikle et al. (1999). We consider an isotropic, homogeneous, and statistically stationary state. Simulations use a conventional pseudo-spectral method and second-order dissipation. Results were obtained on a $`1024\times 1024`$ grid with long time averaging. Large-scale coherent vortices are clearly visible. Further details about the numerics and resulting velocity pdfs are described elsewhere Schorghofer (2000a, b).
In many instances the velocity pdf is approximately Gaussian Frisch (1995), and this is also the case for the simulated turbulence here Schorghofer (2000a). Far more conclusive than the velocity distribution turns out to be the statistics of velocity derivatives. A number of authors have investigated the velocity gradients of three-dimensional Navier-Stokes turbulence Kida and Murakami (1989); Kraichnan (1990); Vincent and Meneguzzi (1991); Yamamoto and Kambe (1991); Frisch and She (1991); Benzi et al. (1991); Chen et al. (1993); Belin et al. (1997); Shafi et al. (1997), but here we consider the far less studied two-dimensional case. Measurements of the transverse velocity derivatives are presented in section II. Section III discusses evidence that large eddies alone provide only a minor contribution to the observed statistics. Section IV briefly discusses pertinent differences between forced and unforced turbulence. The last section contains conclusions.
## II The probability distribution of velocity derivatives
Earlier work based on satellite altimeter data reported transverse velocity gradient pdfs in small parts of the ocean Llewellyn Smith and Gille (1998). These results differed from gradient pdfs derived for decaying two-dimensional turbulence, which showed an approximate Cauchy distribution during the late stage of the evolution Jimenez (1996); Min et al. (1996). The discrepancy is resolved by comparing to simulations of stationary turbulence.
Figure 1 shows velocity gradient pdfs derived from ocean observations and simulations. The solid line indicates the pdf of normalized velocity gradient data derived from global satellite altimetry. To determine the oceanic pdf, velocity gradient data drawn from latitudes between 10 and 60N and between 10 and 60S are sorted geographically into 2.5 by 2.5 boxes. Data near the equator are omitted because the geostrophic relationship is weak at low latitudes. The standard deviation of gradients in each latitude-longitude box varies from $`1.4\times 10^5`$ s<sup>-1</sup> near 60 latitude up to $`6.4\times 10^5`$ s<sup>-1</sup> near 10 latitude, with a median value of $`1.9\times 10^5`$ s<sup>-1</sup>. To compensate for this geographic variation, gradients are normalized to have unit standard deviation in each box, and then the pdf is computed from all of the normalized gradient data. For comparison, we also normalized our pdfs using the mean absolute value of the velocity gradient; this did not diminish the strong tails of the gradient pdf.
The dashed line in Fig. 1 represents the transverse velocity gradient pdf from two-dimensional Navier-Stokes turbulence. The dotted lines represent the narrow Gaussian distribution and the broader Cauchy distribution, $`P(x)=c/[\pi (c^2+x^2)]`$, with long tails. The tails contribute noticeably to the standard deviation of the pdf, and therefore the Gaussian is fitted to the data without requiring unit area and unit standard deviation. This is necessary to make the Gaussian closely approximate the central part of the pdf. Since the Cauchy distribution cannot be normalized by its standard deviation, the constant $`c`$ is chosen such that $`P(0)`$ matches. Both the simulated and observed gradient pdfs appear Gaussian for small velocity gradients, up to about one standard deviation. For large gradients they decay significantly more slowly than do Gaussian tails but substantially faster than the Cauchy distribution found in simulations of decaying turbulence Jimenez (1996); Min et al. (1996). There is good agreement between observed and simulated pdfs up to even the largest fluctuations measured in the simulation.
Error bars for the pdfs were estimated by grouping the data into $`N`$ groups and computing pdfs for each group. Error of the mean pdf is then taken to be the standard deviation of the pdf divided by $`\sqrt{N}`$. For this analysis, $`N`$ was the total number of 2.5 boxes for the surveyed ocean or the number of velocity snapshots for the simulation. Since many ocean observations are available, statistical errors are expected to be small compared to systematic errors. In fact, the statistical errors are frequently narrower than the line width in Fig. 1. Differences between the two distributions exceed the statistical errors and are likely to be due to a number of factors, including instrumental and atmospheric correction errors in the altimeter data, which make the measurements noisy, as well as differences in the physics of two-dimensional Navier–Stokes equations compared with the ocean.
The kurtosis (flatness), $`x^4/x^2^2`$, can serve as quantitative comparison of the shape of the pdf. In the simulation results, the velocity gradient pdf has a kurtosis of $`4.7`$, indicating clear deviation from Gaussian distribution. If the observed ocean pdf is terminated beyond the extent of the simulated one, at about four and a half standard deviations, its kurtosis is also $`4.7`$. This quantitative comparison confirms that the oceanic pdf is substantially better matched by the simulation than by either of the two ideal distributions.
Velocity gradient pdfs depend on the spatial separation between velocity measurements. The velocity correlations between two points decrease with distance, and velocities at points very far apart can be assumed to be statistically independent. The distribution of velocity differences across very large distances reduces to that of the velocity (with twice the variance). The 12 km separation of TOPEX observations is small compared with the decorrelation length scales of wind forcing, O(1000 km), and of mesoscale ocean features O(100 km), so velocities at adjacent observation points are expected to be strongly correlated. Therefore, to obtain comparable results from the numerical simulation, we have computed gradient pdfs from velocity differences over asymptotically small separations. For comparison, if we compute gradient pdfs over progressively larger distances in the simulation, then the distribution narrows from its original shape (dashed line in Fig. 1) and becomes close to Gaussian.
The basic simulations had a large-scale Reynolds number on the order of $`10^4`$, while for ocean turbulence a Reynolds number of $`10^7`$ might be typical Mellor (1996). Pdfs were also determined for simulations with lower and higher Reynolds numbers, using respectively lower and higher resolutions, but shorter sampling time. There is no significant change in the shape of the pdfs Takahashi and Gotoh (1996), although these data do not exclude a weak dependence on Reynolds number. The absence of any detectable Reynolds number dependence suggests that the simulation data are close to what they look like at substantially higher Reynolds number. The difference in the length of the tails in Fig. 1 may be due to the vast difference in Reynolds number, difference in sampling size, and errors from the numerical differentiation of data.
The real ocean differs from the forced Navier–Stokes system because of the addition of the $`\beta `$-effect, stratification, three-dimensional motions, and buoyancy. Hence it is surprising that there is such a close agreement between measurement and simulation. In any case, the agreement between observation and simulation suggests that the oceanic velocity statistics may be understood within the framework of two-dimensional turbulence.
## III The role of coherent vortices
Idealized models of point vortices predict a Cauchy distribution for the velocity gradients and a Gaussian distribution for the velocity Jimenez (1996); Min et al. (1996); Kuvshinov and Schep (2000); Chavanis and Sire (2000, 2001). This agrees with results from decaying two-dimensional turbulence Jimenez (1996); Arroyo et al. (1995). Hence, in the late stages of decay, the statistics of velocity gradients have been successfully understood to result from the far-field of well-separated vortices Jimenez (1996); Min et al. (1996). In contrast, pdfs of ocean surface velocity gradients are observed to have more rapidly decaying tails than do Cauchy distributions. Also as Fig. 1 makes evident, the simulations of stationary two-dimensional turbulence show far less pronounced tails than a Cauchy distribution.
The discrepancy arises not only in the shape of the distribution but also in its width. A straight-forward way of illustrating this is to calculate the velocity field produced by vortices with vorticities that exceed twice the root-mean-square value. Figure 2 shows the transverse velocity gradients produced by these coherent vortices (solid line). For comparison, the actual distribution is shown as a dashed line. Clearly the large coherent vortices do not generate enough intermediate gradients. (Nor, for that matter, do they account for most of the velocities.) Consequently, the distribution of gradients is poorly accounted for by large-scale coherent vortices.
The contribution of the small-scale turbulence is also relevant. This agrees with the basic physical picture, according to which the late stage of decaying turbulence consists of coherent vortices. Its statistics can therefore be understood in terms of them. In the stationary case, on the other hand, fluctuations over a wide spectrum of spatial scales contribute to the gradient statistics.
Available statistical variables from the altimeter are the velocity and the transverse velocity derivative. Hence, one can study the cross-correlation between these two quantities. Here, we examine the conditional average of the squared velocity gradient as a function of velocity, $`(_yv_x)^2||v_x|`$, which is the average of $`(_yv_x)^2`$ over all points with velocity component $`\pm v_x`$. The slope of $`(_yv_x)^2||v_x|`$ is a measure of the correlation between the velocity and the transverse velocity derivative. If there were no correlation between the velocity at a point and the gradient at the same point, the conditional average would be constant for all values of $`v_x`$ and would be exactly one if velocity gradients were normalized by their standard deviation. In contrast, if gradients and velocities were strongly correlated, as would be expected around an isolated vortex, then the graph for the conditional average would have a pronounced slope.
Figure 3 shows the square-root of the measured conditional average together with that for two-dimensional Navier-Stokes turbulence. For the ocean, we have normalized both $`v_x`$ and $`_yv_x`$ by their standard deviations in each 2.5 by 2.5 geographic box, because both quantities vary spatially. The axes are labeled in units of their respective standard deviations, $`\sqrt{v_x^2}`$ and $`\sqrt{(_yv_x)^2}`$. The correlation between velocity and its gradient is weak, but the gradients tend to be higher when the velocity is large. If oceanic gradients beyond four standard deviations are excluded, which is a fairer comparison with the simulation, the conditional average is closer to one. The longitudinal component of the conditional average (not shown) exhibits behavior similar to the transverse component. Also shown in Fig. 3 is the conditional average for the velocity field of vortices larger than twice the root-mean-square vorticity (dash-dotted line). As expected there is a comparatively strong correlation between velocities and velocity derivatives. At large velocities the slopes of the graphs for the coherent vortices and the ocean are similar. This may indicate influence by large eddies in regions where the velocities are high, although no corresponding evidence is found in the probability distribution of the gradients. The situation at high velocities is therefore somewhat ambiguous. For small velocities, which cover most of the area, conditional averages are near one for simulations (dashed line) and observations (solid line), indicating that at low velocity, gradients are almost uncorrelated with velocity. The inset in Fig. 3 shows the conditional average using $`|_yv_x|`$ instead of $`(_yv_x)^2`$, which is less sensitive to outliers. For small velocities, the agreement between observation and simulation is closer and the discrepancy between the large eddy field and the other two conditional averages is stronger.
The deviation of the observed conditional averages from that for coherent vortices strengthens the evidence that the gradient statistics are unaccounted for by the velocity field created by large-scale coherent eddies. The role of coherent vortices in generating the observed velocity statistics is minor. This conclusion cautions against attempts to model oceanic velocity fields by large eddies.
## IV Forced versus unforced turbulence
Although only the transverse velocity component can be determined from altimeter data, simulations also permit us to examine the longitudinal derivative, $`_xv_x`$. Figure 4 shows a clear difference between the behavior of the longitudinal and transverse components. In our forced simulations, the standard deviation of longitudinal fluctuations is about 60% of that for transverse fluctuations. In isotropic and incompressible turbulence there is an exact relation between the standard deviation of transverse and longitudinal component Batchelor (1953). With a calculation analogous to the well-known three-dimensional case Batchelor (1953), we find in the two-dimensional case $`(_yv_x)^2=3(_xv_x)^2`$. Hence, the standard deviation for the longitudinal component is $`1/\sqrt{3}58\%`$ of that for the transverse component. This agrees with the measured value of 60% in the simulation.
Longitudinal and transverse pdfs differ not only by a factor of $`\sqrt{3}`$ in their standard deviation, but also in their shape. While the transverse gradients strongly deviate from a Gaussian distribution, the longitudinal gradient pdf more closely approximates a Gaussian. The kurtosis of the longitudinal component is $`3.5`$, substantially closer to the Gaussian value of $`3`$ than the transverse component is, implying that large longitudinal gradients occur less frequently than do large transverse gradients. For simple point-vortex models both, transverse and longitudinal components, are distributed like Cauchy distributions (albeit with different standard deviations) Jimenez (1996). Also in the late stage of decaying turbulence, the transverse component is distributed in the same way as the longitudinal component Jimenez (1996). This is yet another difference between forced and unforced turbulence.
Overall, our study establishes a clear distinction between the gradient statistics of unforced (freely decaying) and forced (stationary) turbulence. The presence of forcing not only influences the properties of large-scale vortices but also changes the distribution of eddies over different scales. (Freely decaying turbulence has an inverse energy cascade while two-dimensional turbulence forced at large scales is governed by a direct enstrophy cascade.) Further study is needed to determine how the statistics may depend on the temporal and spatial structure of the forcing.
## V Conclusions
In conclusion, we find that transverse velocity derivative pdfs from observed upper-ocean turbulence agree closely with forced two-dimensional simulations but differ from previously reported unforced turbulence. The forcing diminishes the role of coherent vortices in the pertinent statistics. The distribution and cross-correlation of velocity derivatives provide clear evidence that large coherent eddies play only a minor role in generating the observed statistics. Further study of forced two-dimensional turbulence appears likely to shed light on the character of meso-scale turbulence in the ocean.
###### Acknowledgements.
The work of N.S. was supported by a grant from the Research Grants Council of the Hong Kong Special Administrative Region, China (RGC Ref. No. CUHK4119/98P), by a postdoctoral fellowship from the Chinese University of Hong Kong, and by the Massachusetts Institute of Technology. S.T.G. was supported by NASA through the Jason Altimeter Science Working Team (JPL contract 1204910).
|
warning/0007/nlin0007002.html
|
ar5iv
|
text
|
# References
A connection between the mKdV hierarchy and the sine-Gordon equation has been a recurrent theme in the soliton literature, see and references therein. As observed already in 1980 , the Hamiltonians of the mKdV hierarchy remain conserved also with respect to the sine-Gordon flow. This coincidence finds a natural explanation in the framework in which the mKdV hierarchy is embedded in the extended hierarchy consisting of mutually commuting positive and negative flows. The positive part of the hierarchy comprises of the mKdV hierarchy while its negative counterpart contains the sine-Gordon equation and its own hierarchy of differential equations. The existence of two mutually compatible family of flows for every integrable system is a reflection of the Riemann problem connected with two complementing solutions to the underlying linear spectral problem. One solution method uses an expansion in negative powers of the spectral parameter $`\lambda `$ and gives rise to the positive hierarchy while the other method uses an expansion in the positive powers of $`\lambda `$ and gives rise to the negative hierarchy. In the present letter we show how to construct the hierarchy of the negative flows and apply this method to the AKNS hierarchy. The negative hierarchy is shown in the latter case to contain the complex sine-Gordon equation, introduced in the context of the Lund-Regge model.
The approach we develop is a combination of the algebraic and pseudo-differential formalisms. In its general form it explains mutual commutativity of positive and negative flows in the framework of constrained KP hierarchy which contains AKNS model as a special case with the $`sl(2)`$ loop algebra and homogeneous gradation .
In the end of the letter we also comment on how our approach applies to the $`sl(n+1)`$ mKdV type of hierarchies and we obtain the Toda type of hierarchies among the negative flows.
Let $`\widehat{𝒢}=\widehat{s}l(2)`$ be a loop algebra with a graded structure $`\widehat{𝒢}=_n\text{}\widehat{𝒢}_n`$ given by a power series expansion in the spectral parameter $`\lambda `$. This expansion defines an integral homogeneous gradation with respect to the gradation operator $`d=\lambda d/d\lambda `$. The algebra $`𝒢=sl(2,\text{})`$ has a standard basis $`E_\alpha =\sigma _+`$, $`E_\alpha =\sigma _{}`$ and $`H=\sigma _3`$. We work within an algebraic approach to the integrable models based on the linear spectral problem $`L(\mathrm{\Psi })=0`$ with a matrix Lax operator $`L=D_x+E+A`$. Here, $`E=\lambda \sigma _3/2`$ is a semi-simple element of $`\widehat{𝒢}`$, chosen for simplicity to be a grade one element ($`EE^{(1)}\widehat{𝒢}_1`$, $`E^{(n)}=\lambda ^nH/2\widehat{𝒢}_n`$). The matrix $`A=q\sigma _++r\sigma _{}`$ is the grade zero component of Im (ad $`E`$) (see ). Accordingly, the matrix operator $`L`$ for the AKNS hierarchy reads:
$$L=\left(\begin{array}{cc}D+\lambda /2& q\\ r& D\lambda /2\end{array}\right)=ID+\frac{\lambda }{2}H+qE_\alpha +rE_\alpha $$
(1)
here $`D`$ is the derivative with respect to $`x`$ acting to the right as an operator according to the Leibniz rule. In the corresponding formalism based on the pseudo-differential calculus the equivalent spectral problem $`(\psi )=\lambda \psi `$ is given in terms of the pseudo-differential Lax operator $`=DrD^1q`$. The self-commuting isospectral flows ($`n>0`$): $`_nr=B_n(r)`$ and $`_nq=B_n^{}(q)`$ with $`B_n=(^n)_+`$ belong to the positive part of the AKNS hierarchy. The conjugation of $`B_n`$ is defined is such a way that $`D=D`$ and $`(AB)^{}=B^{}A^{}`$. The second flow of the hierarchy:
$$_2r=r_{xx}2qr^2;_2q=q_{xx}+2q^2r$$
(2)
gives the familiar vector non-linear Schrödinger equation.
To define a “negative part” of the hierarchy we need a matrix $`M`$ which arises as a formal solution of the linear spectral problem:
$$L(M)=\left(_x+E+A\right)M=0$$
(3)
given in terms of the path ordered exponential:
$$M=𝒫e^{^x\left(E+A\right)𝑑y}$$
(4)
where symbol $`𝒫`$ denotes a path ordering. Note, that all terms in the above exponential contain only positive (and zero) grade generators.
The negative flows are induced by conjugation with the matrix $`M`$. To the element $`X_n=X\lambda ^n`$ of $`\widehat{𝒢}_n`$ with $`n>0`$ we associate a flow:
$$\delta _X^{(n)}M(MX_nM^1)_+M$$
(5)
Direct calculation shows that these flows constitute a graded Borel loop algebra $`[\delta _X^{(n)},\delta _Y^{(m)}]=\delta _{[X,Y]}^{(nm)}`$. Their action on the grade-zero matrix $`A`$ is given by
$$\delta _X^{(n)}A=[(MX_nM^1)_{},L]=[(MX_nM^1)_1,E]$$
(6)
The flow generated by $`X_1=E^{(1)}`$ is of special interest and we now provide for it a zero curvature formulation. We choose the Gauss decomposition given by the following exponential of terms belonging to zero grade subalgebra $`\widehat{𝒢}_0=sl(2)`$:
$$B=e^{\chi E_\alpha }e^{RH}e^{\psi E_\alpha }$$
(7)
and define gauge potentials:
$$𝒜_{}=BE^{(1)}B^1;𝒜_+=_xBB^1E$$
(8)
In order to match the number of independent modes in the matrix $`A`$ we impose two “diagonal” constraints $`Tr\left(_xBB^1H\right)=0`$ and $`Tr\left(B^1\overline{}BH\right)=0`$ which effectively eliminate $`R`$ in terms of $`\psi `$ and $`\chi `$. In fact, those constraints reduce the zero grade subspace $`\widehat{𝒢}_0=sl(2)`$ into the coset $`sl(2)/U(1)`$. A more general and systematic construction for the affine non-abelian Toda models in terms of the coset $`sl(2)U(1)^{\mathrm{rank}𝒢}/U(1)`$ is discussed in reference where the models are constructed in terms of the two-loop WZWN models . Thus, after imposing these constraints $`𝒜_+`$ becomes equal to $`_xBB^1E=AE=_xMM^1`$ and the zero curvature condition:
$$[\overline{}+𝒜_{},_x+𝒜_+]=\overline{}𝒜_+_x𝒜_{}+[𝒜_{},𝒜_+]=0$$
(9)
holds for $`\overline{}=\delta _E^{(1)}`$ as a consequence of (6).
$`B`$ has been chosen so that after imposition of the constraints $`_xBB^1=qE_\alpha +rE_\alpha `$. Accordingly, we obtain the following representation for $`q`$ and $`r`$:
$$q=\frac{(_xu)}{\mathrm{\Delta }}e^R;r=(_x\overline{u})e^R$$
(10)
where
$$u=\psi e^R;\overline{u}=\chi e^R;\mathrm{\Delta }=1+u\overline{u}$$
(11)
with non-local field $`R`$ being determined in terms $`u`$ and $`\overline{u}`$ from the “diagonal” constraints :
$`Tr\left(_xBB^1H\right)`$ $`=`$ $`0_xR={\displaystyle \frac{\overline{u}_xu}{\mathrm{\Delta }}}`$ (12)
$`Tr\left(B^1\overline{}BH\right)`$ $`=`$ $`0\overline{}R={\displaystyle \frac{u\overline{}\overline{u}}{\mathrm{\Delta }}}`$ (13)
The zero curvature equations (9):
$`\overline{}q`$ $`=`$ $`\overline{}\left({\displaystyle \frac{_xu}{\mathrm{\Delta }}}e^R\right)=2ue^R`$ (14)
$`\overline{}r`$ $`=`$ $`\overline{}\left(_x\overline{u}e^R\right)=2\overline{u}\mathrm{\Delta }e^R`$ (15)
together with eqs. (12)-(13) take now a form of the complex sine-Gordon equations :
$`_x\overline{}u+{\displaystyle \frac{u^{}_xu\overline{}u}{1uu^{}}}+2u(1uu^{})`$ $`=`$ $`0`$ (16)
$`_x\overline{}u^{}+{\displaystyle \frac{u_xu^{}\overline{}u^{}}{1uu^{}}}+2u^{}(1uu^{})`$ $`=`$ $`0`$ (17)
after substitution $`uiu`$ and $`\overline{u}iu^{}`$. Notice, that with the identification from (10)-(13) the $`𝒜_+`$ component of gauge potentials are shared by the AKNS and complex sine-Gordon theories. Therefore, by gauge transforming $`𝒜_+`$ in (8) into the Ker (ad $`E`$) we obtain simultaneous Hamiltonians for both complex sine-Gordon and AKNS models.
We now sketch a pseudo-differential approach to the study of “negative” flows developed in . Here we work with the AKNS Lax operator $`=DrD^1q`$. First, note that $``$ can be described as a ratio of two ordinary monic differential operators as $`=L_2L_1^1`$, where $`L_1,L_2`$ denote monic operators $`L_1=(D+\phi _1^{}+\phi _2^{})`$ and $`L_2=(D+\phi _1^{})(D+\phi _2^{})`$ of, respectively, order $`1`$ and $`2`$. A monic differential operator $`L_2`$ is fully characterized by elements of its kernel, $`\varphi _1=\mathrm{exp}(\phi _2)`$ and $`\varphi _2=\mathrm{exp}(\phi _2)^x\mathrm{exp}(\phi _2\phi _1)`$. Its inverse $`L_2^1`$, is given by $`L_2^1=_{\alpha =1}^2\varphi _\alpha D^1\psi _\alpha `$, where $`\psi _1=\mathrm{exp}(\phi _1)^x\mathrm{exp}(\phi _2\phi _1)`$ and $`\psi _2=\mathrm{exp}(\phi _1)`$ are kernel elements of the conjugated operator $`L_2^{}=(D+\phi _2^{})(D+\phi _1^{})`$, see and references therein. In this notation, $`=D+L_2(\mathrm{exp}(\phi _1\phi _2))D^1\mathrm{exp}(\phi _1+\phi _2)`$ and accordingly:
$$q=\mathrm{exp}(\phi _1+\phi _2);r=\left(\phi _1^{\prime \prime }\phi _1^{}\phi _2^{}\right)\mathrm{exp}(\phi _1\phi _2)$$
(18)
Similarly, the inverse of $``$ is too given as a ratio of differential operators $`^1=L_1L_2^1=_{\alpha =1}^2L_1(\varphi _\alpha )D^1\psi _\alpha `$. The functions $`\mathrm{\Phi }_\alpha ^{(1)}L_1(\varphi _\alpha )`$ and $`\mathrm{\Psi }_\alpha ^{(1)}\psi _\alpha `$ satisfy the same flow equations as $`r`$ and $`q`$ with respect to the positive flows of the AKNS hierarchy. We now extend the AKNS hierarchy by the “negative” flows generated by the pseudo-differential operators :
$$_𝒜^{(n)}=\underset{\alpha ,\beta =1}{\overset{2}{}}𝒜_{\alpha \beta }\underset{s=1}{\overset{n}{}}\mathrm{\Phi }_\beta ^{(n+s1)}D^1\mathrm{\Psi }_\alpha ^{(s)};n=1,2,3,\mathrm{}$$
(19)
where $`\mathrm{\Phi }_\alpha ^{(n)}=^{n+1}(\mathrm{\Phi }_\alpha ^{(1)})`$ and $`\mathrm{\Psi }_\alpha ^{(n)}=(^{})^{n+1}(\mathrm{\Psi }_\alpha ^{(1)})`$ are expressed entirely by the phase variables $`\phi _1`$ and $`\phi _2`$ of the AKNS hierarchy. Furthermore, $`𝒜_{\alpha \beta }`$ is a constant $`2\times 2`$ matrix. The corresponding “negative” symmetry flows are defined by:
$$𝒟_𝒜^{(n)}=[_𝒜^{(n)},]$$
(20)
The following relations follow from (20) and determine flows on $`\mathrm{\Phi }_\alpha ^{(m)},\mathrm{\Psi }_\alpha ^{(m)}`$:
$`𝒟_𝒜^{(n)}(\mathrm{\Phi }_\alpha ^{(m)})`$ $`=`$ $`_𝒜^{(n)}\left(\mathrm{\Phi }_\alpha ^{(m)}\right){\displaystyle \underset{\beta =1}{\overset{2}{}}}𝒜_{\alpha \beta }\mathrm{\Phi }_\beta ^{(nm)}`$ (21)
$`𝒟_𝒜^{(n)}(\mathrm{\Psi }_\alpha ^{(m)})`$ $`=`$ $`\left(_𝒜^{(n)}\right)^{}\left(\mathrm{\Psi }_\alpha ^{(m)}\right)+{\displaystyle \underset{\beta =1}{\overset{2}{}}}𝒜_{\beta \alpha }\mathrm{\Psi }_\beta ^{(nm)}`$ (22)
These relations ensure that $`𝒟_𝒜^{(n)}`$ span a graded Borel loop algebra: $`[𝒟_𝒜^{(n)},𝒟_{}^{(m)}]=𝒟_{[𝒜,]}^{(nm)}`$. The flows $`𝒟_𝒜^{(n)}`$ preserve the constrained structure of the AKNS hierarchy and act on (adjoint) eigenfunctions $`q`$ and $`r`$ according to $`𝒟_𝒜^{(n)}(r)=_𝒜^{(n)}(r)`$ and $`𝒟_𝒜^{(n)}(q)=\left(_𝒜^{(n)}\right)^{}\left(q\right)`$, due to identities $`(\mathrm{\Phi }_\alpha ^{(1)})=0`$ and $`(^{})(\mathrm{\Psi }_\alpha ^{(1)})=0`$. It is interesting to note at this point that the generating functions $`F_\alpha (\lambda )=_{n=1}^{\mathrm{}}\lambda ^{n1}\mathrm{\Phi }_\alpha ^{(n)}`$ and $`G_\alpha (\lambda )=_{n=1}^{\mathrm{}}\lambda ^{n1}\mathrm{\Psi }_\alpha ^{(n)}`$ for $`\mathrm{\Phi }_\alpha ^{(n)}`$ and $`\mathrm{\Psi }_\alpha ^{(n)}`$ are the solutions of the spectral problems $`(F_\alpha (\lambda ))=\lambda F_\alpha (\lambda )`$, $`^{}(G_\alpha (\lambda ))=\lambda G_\alpha (\lambda )`$.
We now present two of the main results of this paper. First, the flows $`𝒟_𝒜^{(n)}`$ commute with the isospectral flows of the AKNS hierarchy. This follows from (20) and the fact that $`\mathrm{\Phi }_\alpha ^{(n)}`$, $`\mathrm{\Psi }_\alpha ^{(n)}`$ are (adjoint) eigenfunctions with respect to isospectral flows, i.e. $`_n\mathrm{\Phi }_\alpha ^{(m)}=B_n(\mathrm{\Phi }_\alpha ^{(n)})`$ and $`_n\mathrm{\Psi }_\alpha ^{(m)}=B_n^{}(\mathrm{\Psi }_\alpha ^{(m)})`$. Accordingly, the flows $`𝒟_𝒜^{(n)}`$ define the symmetry of the AKNS hierarchy. One can generalize this result to the case of the arbitrary constrained KP model associated with the loop algebra $`\widehat{s}l(N)`$ and with the Lax operator $`=()_++_{i=1}^M\mathrm{\Phi }_iD^1\mathrm{\Psi }_i`$ with $`M<N`$ . It holds in that case that the flows of the negative Borel loop algebra will commute with the flows of the positive Borel loop algebra, which has recently been defined for the constrained KP hierarchy in reference , which contains several technical lemmas helpful for completing the proofs ommitted here. The final result is that $`[𝒟_𝒜^{(n)},𝒟_{}^{(m)}]=0`$ where $`𝒟_{}^{(m)}`$ are flows corresponding to the positive Borel loop algebra defined in with $``$ being a constant $`M\times M`$ matrix, $`n,m>0`$, $`𝒜`$ is a constant $`N\times N`$ matrix appearing in a straightforward generalization of (19) .
Secondly, the flows $`𝒟_𝒜^{(n)}`$ defined in (20) for the AKNS hierarchy coincide with the flows $`\delta _X^{(m)}`$ defined by the matrix $`M`$ for $`m=n>0`$ and $`X=𝒜=\sigma _3`$. This observation provides an indirect proof for that the flows induced by the conjugation with the matrix $`M`$ in (5) and (6) are the symmetry flows of the AKNS hierarchy and in particular commute with the isospectral flows. We will ilustrate the identity $`𝒟_𝒜^{(n)}=\delta _𝒜^{(n)}`$ for $`𝒜=\sigma _3`$ and $`n=1`$. From (21)-(22) we find
$$𝒟_{\sigma _3}^{(1)}(\phi _1)=2^xe^{\phi _2\phi _1}^x\phi _1^{}e^{\phi _1\phi _2};𝒟_{\sigma _3}^{(1)}(\phi _2)=2^xe^{\phi _2\phi _1}^x\phi _2^{}e^{\phi _1\phi _2}$$
(23)
which, using expressions (18), leads to:
$`𝒟_{\sigma _3}^{(1)}(q)`$ $`=`$ $`2e^{2\phi _1}{\displaystyle ^x}e^{\phi _2\phi _1}`$ (24)
$`𝒟_{\sigma _3}^{(1)}(r)`$ $`=`$ $`2\phi _1^{}e^{\phi _1\phi _2}\left(1+\phi _1^{}e^{\phi _1\phi _2}{\displaystyle ^x}e^{\phi _2\phi _1}\right)`$ (25)
Comparing with equations (14)-(15) we see that equality $`\overline{}=𝒟_{\sigma _3}^{(1)}`$ holds provided we identify:
$$R=\phi _1;u=e^{\phi _1}^xe^{\phi _2\phi _1};\overline{u}=\phi _1^{}e^{\phi _2}$$
(26)
With representation (26) and transformations (24)-(25) the constraints (12)-(13) hold automatically and (14)-(15) are satisfied as well with $`\overline{}=\delta _{\sigma _3}^{(1)}=𝒟_{\sigma _3}^{(1)}`$. Similarly, we find that $`\delta _{\sigma _\pm }^{(1)}=𝒟_\sigma _{}^{(1)}`$ with:
$`𝒟_{\sigma _+}^{(1)}(q)`$ $`=`$ $`u^2;𝒟_{\sigma _+}^{(1)}(r)=e^{2\phi _1}\mathrm{\Delta }^2`$ (27)
$`𝒟_\sigma _{}^{(1)}(q)`$ $`=`$ $`e^{2\phi _1};𝒟_\sigma _{}^{(1)}(r)=\overline{u}^2`$ (28)
Due to the fact that we are dealing with a Borel loop algebra all the remaining symmetry flows can be found from the commutator relations involving known lower grade flows.
We now comment on the special case of the generalized mKdV model associated with the $`\widehat{s}l(n+1)`$ algebra with the principal gradation . In the algebraic approach the Lax matrix $`L=D+A+E`$ contains
$$E=E^{(1)}=\underset{j=1}{\overset{n}{}}E_{\alpha _j}^{(0)}+E_{(\alpha _1+\mathrm{}+\alpha _n)}^{(1)};A=\underset{i=1}{\overset{n}{}}(\phi _1^{}+\mathrm{}+\phi _i^{})\alpha _iH$$
(29)
with $`E`$ and $`A`$ possessing grade $`1`$ and zero according to the principal gradation defined by the charge $`Q=(n+1)d+_{i=1}^n\lambda _iH`$, where $`\lambda _i`$ are fundamental weights corresponding to the simple roots $`\alpha _i`$. The solution to the linear problem $`(D+A+E)(M)=0`$ is given by the path ordered exponentials :
$`M`$ $`=`$ $`e^{_{i=1}^n\left(\phi _1+\mathrm{}+\phi _i\right)\alpha _iH}𝒫e^{^x_{i=1}^n\left(f_iE_{\alpha _i}^{(0)}+f_0E_{(\alpha _1+\mathrm{}+\alpha _n)}^{(1)}\right)}`$ (30)
$`f_j`$ $`=`$ $`e^{_{j=1}^nK_{ji}\left(\phi _1+\mathrm{}+\phi _i\right)};f_0=e^{_{i=1}^n(K_{1i}+\mathrm{}+K_{ni})\left(\phi _1+\mathrm{}+\phi _i\right)}`$
where $`K_{ij}`$ is a Cartan matrix of $`sl(n+1)`$. Inserting $`X=E_{\alpha _j}^{(0)}`$ with grade $`1`$ into (6) with $`M`$ from (30) we obtain
$$\delta _{\alpha _j}^{(1)}(\phi _j^{})=e^{\phi _j+\phi _{j+1}};\delta _{\alpha _j}^{(1)}(\phi _{j+1}^{})=e^{\phi _j+\phi _{j+1}};\delta _{\alpha _j}^{(1)}(\phi _l^{})=0lj,j+1$$
(31)
while for $`X=E_{(\alpha _1+\mathrm{}+\alpha _n)}^{(1)}`$ we obtain
$$\delta _{(\alpha _1+\mathrm{}+\alpha _n)}^{(1)}(\phi _1^{})=e^{2\phi _1+\phi _2+\mathrm{}+\phi _n};\delta _{(\alpha _1+\mathrm{}+\alpha _n)}^{(1)}(\phi _l^{})=0l>1$$
(32)
These results give for the element $`X=E^{(1)}=_{j=1}^nE_{\alpha _j}^{(0)}+E_{(\alpha _1+\mathrm{}+\alpha _n)}^{(1)}`$ and $`\overline{}=\delta _E^{(1)}`$ the affine Toda equation for $`sl(n+1)`$:
$$_x\overline{}y_i=e^{_{j=1}^nK_{ij}y_j}e^{_{j=1}^nK_{0j}y_j};y_i=\underset{j=1}{\overset{i}{}}\phi _j$$
(33)
with the extended Cartan matrix $`K_{ab}`$. In the corresponding pseudo-differential approach the mKdV Lax operator $`=L_{n+1}=(D+\phi _1^{})\mathrm{}(D+\phi _{n+1}^{})`$ with the trace condition $`\phi _1+\mathrm{}+\phi _{n+1}=0`$) is an ordinary differential operator. Let $`\varphi _\alpha \mathrm{Ker}(L_{n+1})`$, $`\psi _\alpha \mathrm{Ker}(L_{n+1}^{})`$ with $`\varphi _1=\mathrm{exp}(\phi _{n+1})`$ and $`\psi _{n+1}=\mathrm{exp}(\phi _1)`$. For $`𝒜_{\alpha \beta }=\delta _{\alpha ,n+1}\delta _{\beta ,1}`$ the corresponding generator $`_𝒜^{(1)}=\varphi _1D^1\psi _{n+1}`$ will induce according to relation (20) :
$$𝒟_𝒜^1(\phi _1^{})=e^{\phi _1\phi _{n+1}};𝒟_𝒜^1(\phi _{n+1}^{})=e^{\phi _1\phi _{n+1}}$$
(34)
with $`𝒟_𝒜^1(\phi _j^{})=0`$ for $`1<j<n+1`$. We recognize in (34) the Toda structure of eq. (32). The other transformations of eqs. (31) follow by applying the Darboux-Bäcklund transformations $`\phi _i\phi _{i+j},\mathrm{\hspace{0.17em}1}jn`$ (modulo $`n+1`$).
Outlook. We presented a concept of (non-local) Toda-like symmetries occupying the “negative” sector of the $`sl(N)`$ constrained KP hierarchy and giving rise to the negative Borel $`sl(N)`$ loop algebra. The case of $`sl(2)`$ (both homogeneous and principal gradations) has been described in details for AKNS and mKdV hierarchies. Details of the corresponding Toda like models of the $`sl(3)`$ constrained KP hierarchies will be given elsewhere. It is also of interest to establish similar negative flow structure for the graded algebras connected with supersymmetric integrable models in order to obtain a new point of view on the supersymmetric Toda systems. We also plan to describe relation of the negative Borel additional symmetry loop algebra to the complete (centerless) Virasoro symmetry recently established for the arbitrary constrained KP models . It will also be of interest to establish a general tau-function realization valid for both positive and negative sectors of the integrable models.
Acknowledgements We are indebted to L. Dickey for useful comments on the manuscript. HA thanks Fapesp for financial support and IFT-Unesp for hospitality.
|
warning/0007/hep-ph0007049.html
|
ar5iv
|
text
|
# Instanton Effects on the Role of the Low-Energy Theorem for the Scalar Gluonic Correlation Function
## 1 Introduction
In the chiral limit of $`n_f`$ quarks, the low-energy theorem (LET) for scalar gluonic correlation functions is
$$\mathrm{\Pi }(0)=\underset{Q0}{lim}\mathrm{\Pi }\left(Q^2\right)=\frac{8\pi }{\beta _0}J,$$
(1)
where
$`\mathrm{\Pi }\left(Q^2\right)=i{\displaystyle }d^4xe^{iqx}O|T\left[J(x)J(0)\right]|O,Q^2=q^2>0`$ (2)
$`J(x)={\displaystyle \frac{\pi ^2}{\alpha \beta _0}}\beta \left(\alpha \right)G_{\mu \nu }^a(x)G_{\mu \nu }^a(x)`$ (3)
$`\beta \left(\alpha \right)=\nu ^2{\displaystyle \frac{d}{d\nu ^2}}\left({\displaystyle \frac{\alpha (\nu )}{\pi }}\right)=\beta _0\left({\displaystyle \frac{\alpha }{\pi }}\right)^2\beta _1\left({\displaystyle \frac{\alpha }{\pi }}\right)^3+\mathrm{}`$ (4)
$`\beta _0={\displaystyle \frac{11}{4}}{\displaystyle \frac{1}{6}}n_f,\beta _1={\displaystyle \frac{51}{8}}{\displaystyle \frac{19}{24}}n_f.`$ (5)
The current $`J(x)`$ is renormalization group (RG) invariant for massless quarks , and its normalization has been chosen so that to lowest order in $`\alpha `$
$$J(x)=\alpha G_{\mu \nu }^a(x)G_{\mu \nu }^a(x)\left[1+\frac{\beta _1}{\beta _0}\frac{\alpha }{\pi }+𝒪\left(\alpha ^2\right)\right]\alpha G^2(x)\left[1+\frac{\beta _1}{\beta _0}\frac{\alpha }{\pi }+𝒪\left(\alpha ^2\right)\right].$$
(6)
Most applications of dispersion relations in sum-rules are designed to remove dependence on low-energy subtraction constants. However, knowledge of the LET for gluonic correlation functions permits the possibility of sum-rules that contain explicit dependence on the LET subtraction constant $`\mathrm{\Pi }(0)`$. For example, the dispersion relation appropriate to the asymptotic (perturbative) behaviour of the correlation function (2) is
$$\mathrm{\Pi }\left(Q^2\right)=\mathrm{\Pi }(0)+Q^2\mathrm{\Pi }^{}(0)+\frac{1}{2}Q^4\mathrm{\Pi }^{\prime \prime }(0)Q^6\frac{1}{\pi }\underset{t_0}{\overset{\mathrm{}}{}}𝑑t\frac{\rho (t)}{t^3\left(t+Q^2\right)}.$$
(7)
where $`\rho (t)`$ is the hadronic spectral function with physical threshold $`t_0`$ appropriate to the quantum numbers of the current used to construct the correlation function.
Unfortunately, direct application of the dispersion relation (7) is not possible because the theoretical (perturbative) calculation of $`\mathrm{\Pi }\left(Q^2\right)`$ contains a field-theoretical divergence proportional to $`Q^4`$. A related problem is the significant contribution of excited states and the QCD continuum to the integral of $`\rho (t)`$ in (7). Enhancement of the lowest-lying resonance contribution in applications to light hadronic systems requires greater high-energy suppression of this integral.
The established technique for dealing with these issues is the Laplace sum-rules . A family of Laplace sum-rules can be obtained from the dispersion relation (7) through the Borel transform operator $`\widehat{B}`$,
$$\widehat{B}\underset{\stackrel{N,Q^2\mathrm{}}{N/Q^2\tau }}{lim}\frac{\left(Q^2\right)^N}{\mathrm{\Gamma }(N)}\left(\frac{d}{dQ^2}\right)^N$$
(8)
which has the following useful properties in the construction of the Laplace sum-rules:
$`\widehat{B}[a_0+a_1Q^2+\mathrm{}a_mQ^{2m}]=0,(m\mathrm{finite})`$ (9)
$`\widehat{B}\left[{\displaystyle \frac{Q^{2n}}{t+Q^2}}\right]=\tau (1)^nt^ne^{t\tau },n=0,1,2,\mathrm{}(n\mathrm{finite})`$ (10)
The theoretically-determined quantity
$`_k(\tau ){\displaystyle \frac{1}{\tau }}\widehat{B}\left[\left(1\right)^kQ^{2k}\mathrm{\Pi }\left(Q^2\right)\right],`$ (11)
leads to the following family of Laplace sum-rules, after application of $`\widehat{B}`$ to the dispersion relation (7) weighted by the appropriate power of $`Q^2`$:
$`_k(\tau )`$ $`=`$ $`{\displaystyle \frac{1}{\tau }}\widehat{B}\left[\left(1\right)^kQ^{2k}\mathrm{\Pi }(0)+\left(1\right)^kQ^{2k+2}\mathrm{\Pi }^{}(0)+\left(1\right)^k{\displaystyle \frac{1}{2}}Q^{2k+4}\mathrm{\Pi }^{\prime \prime }(0)\right]`$ (12)
$`{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{\mathrm{}}{}}}𝑑t{\displaystyle \frac{1}{\tau t^3}}\widehat{B}\left[\left(1\right)^k{\displaystyle \frac{Q^{2k+6}}{\left(t+Q^2\right)}}\right]\rho (t)`$
There are some important constraints on $`k`$ that will lead to sum-rules with predictive power. Since the perturbative prediction of $`\mathrm{\Pi }\left(Q^2\right)`$ contains divergent constants multiplied by $`Q^4`$, the sum-rules $`_k(\tau )`$ where this contribution is absent require $`k2`$. However, the low-energy constants $`\mathrm{\Pi }^{}(0)`$ and $`\mathrm{\Pi }^{\prime \prime }(0)`$ are not determined by the LET \[i.e. only the quantity $`\mathrm{\Pi }(0)`$ appears in (1)\]. Hence the sum-rules $`_k(\tau )`$ which will be independent of $`\mathrm{\Pi }^{}(0)`$ and $`\mathrm{\Pi }^{\prime \prime }(0)`$ must satisfy $`k1`$, and only the $`k=1`$ sum-rule will contain dependence on the LET-determined quantity $`\mathrm{\Pi }(0)`$:
$`_1(\tau )=\mathrm{\Pi }(0)+{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{\mathrm{}}{}}}𝑑t{\displaystyle \frac{1}{t}}e^{t\tau }\rho (t)`$ (13)
$`_k(\tau )={\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{\mathrm{}}{}}}dtt^ke^{t\tau }\rho (t),k>1`$ (14)
The “resonance(s) plus continuum” model is used to represent the hadronic physics phenomenology contained in $`\rho (t)`$ in (1314) . In this model, hadronic physics is (locally) dual to the theoretical QCD prediction for energies above the continuum threshold $`t=s_0`$:
$$\rho (t)\theta \left(s_0t\right)\rho ^{had}(t)+\theta \left(ts_0\right)\mathrm{Im}\mathrm{\Pi }^{QCD}(t)$$
(15)
The contribution of the QCD continuum to the sum-rules is denoted by
$$c_k(\tau ,s_0)=\frac{1}{\pi }\underset{s_0}{\overset{\mathrm{}}{}}𝑑tt^ke^{t\tau }\mathrm{Im}\mathrm{\Pi }^{QCD}(t).$$
(16)
Since the continuum contribution is determined by QCD, it is usually combined with the theoretical quantity $`_k\left(\tau \right)`$
$$𝒮_k(\tau ,s_0)_k\left(\tau \right)c_k(\tau ,s_0),$$
(17)
resulting in the following Laplace sum-rules relating QCD to hadronic physics phenomenology:
$`𝒮_1(\tau ,s_0)=\mathrm{\Pi }(0)+{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{s_0}{}}}𝑑t{\displaystyle \frac{1}{t}}e^{t\tau }\rho ^{had}(t)`$ (18)
$`𝒮_k(\tau ,s_0)={\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{s_0}{}}}dtt^ke^{t\tau }\rho ^{had}(t),k>1`$ (19)
The property
$$\underset{s_0\mathrm{}}{lim}c_k(\tau ,s_0)=0$$
(20)
implies that the sum-rules (17) and (14) are identical in the $`s_0\mathrm{}`$ limit.
$$\underset{s_0\mathrm{}}{lim}𝒮_k(\tau ,s_0)=_k\left(\tau \right)$$
(21)
The only appearance of the $`\mathrm{\Pi }(0)`$ term is in the $`k=1`$ sum-rule, and as first noted in , this LET term comprises a significant contribution in the $`k=1`$ sum-rule. From the significance of this scale-independent term one can ascertain the important qualitative role of the LET in sum-rule phenomenology. To see this role, we first model the hadronic contributions $`\rho ^{had}(t)`$ using the narrow resonance approximation
$$\frac{1}{\pi }\rho ^{had}(t)=\underset{r}{}F_r^2m_r^2\delta \left(tm_r^2\right),$$
(22)
where the sum over $`r`$ represents a sum over sub-continuum resonances of mass $`m_r`$. The quantity $`F_r`$ is the coupling strength of the resonance to the vacuum through the gluonic current $`J(0)`$, so the sum-rule for scalar gluonic currents probes scalar gluonium states. In the narrow-width approximation the Laplace sum-rules (1819) become
$`𝒮_1(\tau ,s_0)+\mathrm{\Pi }(0)={\displaystyle \underset{r}{}}F_r^2e^{m_r^2\tau }`$ (23)
$`𝒮_k(\tau ,s_0)={\displaystyle \underset{r}{}}F_r^2m_r^{2k+2}e^{m_r^2\tau },k>1.`$ (24)
Thus if the (constant) LET term is a significant contribution on the theoretical side of (23), then the left-hand side of (23) will exhibit reduced $`\tau `$ dependence relative to other theoretical contributions. To reproduce this diminished $`\tau `$ dependence, the phenomenological (i.e. right-hand) side must contain a light resonance with a coupling larger than or comparable to the heavier resonances. By contrast, the absence of the $`\mathrm{\Pi }(0)`$ (constant) term in $`k>1`$ sum-rules leads to stronger $`\tau `$ dependence which is balanced on the phenomenological side by suppression of the lightest resonances via the additional powers of $`m_r^2`$ occurring in (24). Thus if $`\mathrm{\Pi }(0)`$ is found to dominate $`𝒮_1(\tau ,s_0)`$, then one would expect qualitatively different results from analysis of the $`k=1`$ and $`k>1`$ sum-rules.
Such distinct conclusions drawn from different sum-rules can be legitimate. In the pseudoscalar quark sector, the lowest sum-rule is dominated by the pion, and the low mass of the pion is evident from the minimal $`\tau `$ dependence in the lowest sum-rule. By contrast, the first subsequent sum-rule has an important contribution from the $`\mathrm{\Pi }(1300)`$ , as the pion contribution is suppressed by its low mass, resulting in the significant $`\tau `$ dependence of the next-to-lowest sum-rule.
In the absence of instantons , explicit sum-rule analyses of scalar gluonium uphold the above generalization—those which include the $`k=1`$ sum-rule find a light (less than or on the order of the $`\rho `$ mass) gluonium state, and those which omit the $`k=1`$ sum-rule find a state with a mass greater than $`1\mathrm{GeV}`$. The prediction of a light gluonium state would have interesting phenomenological consequences as a state which could be identified with the $`f_0(4001200)/\sigma `$ meson . However, a detailed treatment of instanton contributions is essential in assessing the viability of such a two-resonance-scale scenario (as evident in the sum-rule analysis of the $`q\overline{q}`$ pseudoscalar channel ) in the scalar gluonium channel.
Seminal work by Shuryak in the instanton liquid model has indicated how an asymptotic ($`\rho _c/\sqrt{\tau }1`$) expression for the instanton contribution to the $`k=1`$ sum-rule may serve to compensate for that sum-rule’s LET component and bring the predicted scalar gluonium mass in line with subsequent lattice estimates ($`1.6\mathrm{GeV}`$ ). Recent work by Forkel has addressed in detail instanton effects on scalar gluonium mass predictions from higher-weight sum-rules and has also corroborated lattice estimates. However, the overall consistency of the $`k=1`$ sum-rule, which is sensitive to the low-energy theorem term, and $`k0`$ sum-rules, which are not, has not been addressed quantitatively.
In Section 2, we explicitly calculate the instanton contributions to Laplace sum-rules of scalar gluonic currents. We pay particular attention to the $`k=1`$ sum-rule and demonstrate that instanton contributions partially cancel against the LET constant $`\mathrm{\Pi }(0)`$ and serve to appreciably diminish its dominance of this leading order sum-rule. The phenomenological implications of this partial cancellation are investigated in Section 3, and a discussion relating our work to other analyses of instanton effects in the scalar gluonium channel is presented in Section 4.
## 2 Instanton Effects in the Laplace Sum-Rules
The field-theoretical (QCD) calculation of $`\mathrm{\Pi }\left(Q^2\right)`$ consists of perturbative (logarithmic) corrections known to three-loop order ($`\overline{\mathrm{MS}}`$ scheme) in the chiral limit of $`n_f=3`$ massless quarks , QCD vacuum effects of infinite correlation length parameterized by the power-law contributions from the QCD vacuum condensates , <sup>1</sup><sup>1</sup>1The calculation of one-loop contributions proportional to $`J`$ in have been extended non-trivially to $`n_f=3`$ from $`n_f=0`$, and the operator basis has been changed from $`\alpha G^2`$ to $`J`$. and QCD vacuum effects of finite correlation length devolving from instantons
$$\mathrm{\Pi }\left(Q^2\right)=\mathrm{\Pi }^{pert}\left(Q^2\right)+\mathrm{\Pi }^{cond}\left(Q^2\right)+\mathrm{\Pi }^{inst}\left(Q^2\right),$$
(25)
with
1. $`\mathrm{}`$ the perturbative contribution (ignoring divergent terms proportional to $`Q^4`$) given by
$`\mathrm{\Pi }^{pert}\left(Q^2\right)=Q^4\mathrm{log}\left({\displaystyle \frac{Q^2}{\nu ^2}}\right)\left[a_0+a_1\mathrm{log}\left({\displaystyle \frac{Q^2}{\nu ^2}}\right)+a_2\mathrm{log}^2\left({\displaystyle \frac{Q^2}{\nu ^2}}\right)\right]`$ (26)
$`a_0=2\left({\displaystyle \frac{\alpha }{\pi }}\right)^2\left[1+{\displaystyle \frac{659}{36}}{\displaystyle \frac{\alpha }{\pi }}+247.480\left({\displaystyle \frac{\alpha }{\pi }}\right)^2\right],a_1=2\left({\displaystyle \frac{\alpha }{\pi }}\right)^3\left[{\displaystyle \frac{9}{4}}+65.781{\displaystyle \frac{\alpha }{\pi }}\right],a_2=10.1250\left({\displaystyle \frac{\alpha }{\pi }}\right)^4`$
2. $`\mathrm{}`$ the condensate contributions given by
$`\mathrm{\Pi }^{cond}\left(Q^2\right)=\left[b_0+b_1\mathrm{log}\left({\displaystyle \frac{Q^2}{\nu ^2}}\right)\right]J+c_0{\displaystyle \frac{1}{Q^2}}𝒪_6+d_0{\displaystyle \frac{1}{Q^4}}𝒪_8`$ (27)
$`b_0=4\pi {\displaystyle \frac{\alpha }{\pi }}[1+{\displaystyle \frac{175}{36}}{\displaystyle \frac{\alpha }{\pi }}],b_1=9\pi \left({\displaystyle \frac{\alpha }{\pi }}\right)^2,c_0=8\pi ^2\left({\displaystyle \frac{\alpha }{\pi }}\right)^2,d_0=8\pi ^2{\displaystyle \frac{\alpha }{\pi }}`$ (28)
$`𝒪_6=gf_{abc}G_{\mu \nu }^aG_{\nu \rho }^bG_{\rho \mu }^c,𝒪_8=14\left(\alpha f_{abc}G_{\mu \rho }^aG_{\nu \rho }^b\right)^2\left(\alpha f_{abc}G_{\mu \nu }^aG_{\rho \lambda }^b\right)^2`$ (29)
3. $`\mathrm{}`$ and the instanton contribution given by
$$\mathrm{\Pi }^{inst}\left(Q^2\right)=32\pi ^2Q^4\rho ^4\left[K_2\left(\rho \sqrt{Q^2}\right)\right]^2𝑑n(\rho ),$$
(30)
where $`K_2(x)`$ represents a modified Bessel function .
The strong coupling constant $`\alpha `$ is understood to be the running coupling at the renormalization scale $`\nu `$, and renormalization group improvement of the Laplace sum-rules implies that $`\nu ^2=1/\tau `$ . The instanton contributions represent a calculation with non-interacting instantons of size $`\rho `$, with subsequent integration over the instanton density distribution $`n(\rho )`$. <sup>2</sup><sup>2</sup>2A factor of 2 to include the sum of instanton and anti-instanton contributions has been included in (25). The theoretical contributions to the Laplace sum-rules corresponding to (25) are
$$_k(\tau )=_k^{pert}(\tau )+_k^{cond}(\tau )+_k^{inst}(\tau ).$$
(31)
An alternative to the direct calculation of the Laplace sum-rules through the definition of $`\widehat{B}`$ in (8) is obtained through an identity relating the Borel and Laplace transform
$`f\left(Q^2\right)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑\tau F(\tau )e^{Q^2\tau }\left[F(\tau )\right]{\displaystyle \frac{1}{\tau }}\widehat{B}\left[f\left(Q^2\right)\right]=F(\tau )=^1\left[f\left(Q^2\right)\right]`$ (32)
$`^1\left[f\left(Q^2\right)\right]={\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{bi\mathrm{}}{\overset{b+i\mathrm{}}{}}}f\left(Q^2\right)e^{Q^2\tau }𝑑Q^2`$ (33)
where the real parameter $`b`$ in the definition (33) of the inverse Laplace transform must be chosen so that $`f\left(Q^2\right)`$ is analytic to the right of the contour of integration in the complex plane. Using the result (32), the Laplace sum-rules (11) can be obtained from an inverse Laplace transform of the theoretically-determined correlation function:
$$_k(\tau )=^1\left[\left(1\right)^kQ^{2k}\mathrm{\Pi }\left(Q^2\right)\right].$$
(34)
In the complex $`Q^2`$ plane where the inverse Laplace transform (33) is calculated, the QCD expression (25) for the correlation function $`\mathrm{\Pi }\left(Q^2\right)`$ is analytic apart from a branch point at $`Q^2=0`$ with a branch cut extending to infinity along the negative-real-$`Q^2`$ axis. Consequently, analyticity to the right of the contour in (33) implies that $`b>0`$. Consider the contour $`C(R)`$ in Figure 1; $`\mathrm{\Pi }\left(Q^2\right)`$ is analytic within and on $`C(R)`$ and so with $`z=Q^2`$
$$0=\frac{1}{2\pi i}\underset{C(R)}{}\left(z\right)^ke^{z\tau }\mathrm{\Pi }(z)𝑑z,$$
(35)
which leads to
$$\frac{1}{2\pi i}\underset{biR}{\overset{b+iR}{}}\left(z\right)^ke^{z\tau }\mathrm{\Pi }(z)𝑑z=\frac{1}{2\pi i}\underset{\mathrm{\Gamma }_1+\mathrm{}+\mathrm{\Gamma }_4}{}\left(z\right)^ke^{z\tau }\mathrm{\Pi }(z)𝑑z\frac{1}{2\pi i}\underset{\mathrm{\Gamma }_c+\mathrm{\Gamma }_ϵ}{}\left(z\right)^ke^{z\tau }\mathrm{\Pi }(z)𝑑z.$$
(36)
Taking the limit as $`R\mathrm{}`$, which requires use of the asymptotic behaviour of the modified Bessel function
$$K_2\left(z\right)\sqrt{\frac{\pi }{2z}}e^z;|z|1,\left|\mathrm{arg}(z)\right|\frac{\pi }{2}$$
(37)
the individual integrals over $`\mathrm{\Gamma }_{1\mathrm{}4}`$ are found to vanish, resulting in the following expression for the Laplace sum-rule.
$$_k\left(\tau \right)=\frac{1}{2\pi i}\underset{ϵ}{\overset{\mathrm{}}{}}t^ke^{t\tau }\left[\mathrm{\Pi }\left(te^{i\pi }\right)\mathrm{\Pi }\left(te^{i\pi }\right)\right]𝑑t+\frac{1}{2\pi }\underset{\pi }{\overset{\pi }{}}\left(1\right)^k\mathrm{exp}\left(ϵe^{i\theta }\tau \right)ϵ^{k+1}e^{i(k+1)\theta }\mathrm{\Pi }\left(ϵe^{i\theta }\right)𝑑\theta $$
(38)
Perturbative and QCD condensate contributions to the Laplace sum-rules are well known , and serve as a consistency check for the conventions used to determine the instanton contribution through (38). Keeping in mind the $`k1`$ constraint established previously, we see that the perturbative contributions to the $`\theta `$ integral in (38) are zero in the limit as $`ϵ0`$, leaving only the anticipated integral of the discontinuity across the branch cut \[i.e. $`\mathrm{Im}\mathrm{\Pi }^{pert}(t)`$\] to determine the following perturbative contributions to the Laplace sum-rule.
$$_k^{pert}\left(\tau \right)=\underset{0}{\overset{\mathrm{}}{}}t^{k+2}e^{t\tau }\left[a_02a_1\mathrm{log}\left(\frac{t}{\nu ^2}\right)+a_2\left(\pi ^23\mathrm{log}^2\left(\frac{t}{\nu ^2}\right)\right)\right]𝑑t$$
(39)
The QCD condensate terms proportional to $`b_0`$, $`c_0`$ and $`d_0`$ in the correlation function $`\mathrm{\Pi }(z)`$ do not have a branch discontinuity, so their contribution to the Laplace sum-rule arises solely from the contour $`\mathrm{\Gamma }_ϵ`$ (represented by the term in (38) with the $`\theta `$ integral), and can be evaluated using the result
$$\frac{1}{2\pi i}\underset{\mathrm{\Gamma }_ϵ}{}\frac{e^{z\tau }}{z^n}𝑑z=\{\begin{array}{c}0,n=0,1,2,\mathrm{}\hfill \\ \frac{\tau ^{n1}}{(n1)!},n=1,2,3,\mathrm{}\hfill \end{array}$$
(40)
The QCD condensate term proportional to $`b_1`$ requires a more careful treatment. If $`\mathrm{\Pi }(z)`$ is replaced with $`\mathrm{log}\left(z/\nu ^2\right)`$ in (38) then we find
$$\frac{1}{2\pi i}\underset{bi\mathrm{}}{\overset{b+i\mathrm{}}{}}\left(z\right)^ke^{z\tau }\mathrm{log}\left(\frac{z}{\nu ^2}\right)𝑑z=\underset{ϵ}{\overset{\mathrm{}}{}}t^ke^{t\tau }𝑑t+\frac{1}{2\pi }\underset{\pi }{\overset{\pi }{}}\left(1\right)^k\mathrm{exp}\left(ϵe^{i\theta }\tau \right)ϵ^{k+1}e^{i(k+1)\theta }\left(\mathrm{log}\left(\frac{ϵ}{\nu ^2}\right)+i\theta \right)𝑑\theta $$
(41)
The last term in this equation will be zero in the $`ϵ0`$ limit except when $`k=1`$. Similarly, the $`t`$ integral is well defined in the $`ϵ0`$ limit except when $`k=1`$. With $`\nu ^2=1/\tau `$, and with evaluation of the $`ϵ0`$ limit \[which, for $`k=1`$, involves cancellation between the two integrals in (41)\] we find
$$\frac{1}{2\pi i}\underset{bi\mathrm{}}{\overset{b+i\mathrm{}}{}}\left(z\right)^ke^{z\tau }\mathrm{log}\left(\frac{z}{\nu ^2}\right)𝑑z=\{\begin{array}{c}\underset{0}{\overset{\mathrm{}}{}}t^ke^{t\tau }dt,k>1\hfill \\ \gamma __E,k=1\hfill \end{array}$$
(42)
where $`\gamma __E0.5772`$ is Euler’s constant. It is easily verified that equations (42), (40), and (39) lead to the known results for the non-instanton contributions to the Laplace sum-rules for scalar gluonic currents.
To evaluate the instanton contributions to the Laplace sum-rule, we must calculate the following integral:
$`{\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{bi\mathrm{}}{\overset{b+i\mathrm{}}{}}}\left(z\right)^ke^{z\tau }z^2\left[K_2\left(\rho \sqrt{z}\right)\right]^2𝑑z`$ $`=`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{ϵ}{\overset{\mathrm{}}{}}}t^{k+2}e^{t\tau }\left[\left[K_2\left(\rho \sqrt{t}e^{i\pi /2}\right)\right]^2\left[K_2\left(\rho \sqrt{t}e^{i\pi /2}\right)\right]^2\right]𝑑t`$ (43)
$`+{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{\pi }{\overset{\pi }{}}}\left(1\right)^k\mathrm{exp}\left(ϵe^{i\theta }\tau \right)ϵ^{k+3}e^{i(k+3)\theta }\left[K_2\left(\rho \sqrt{ϵ}e^{i\theta /2}\right)\right]^2𝑑\theta `$
Simplification of (43) requires the following properties of the modified Bessel function $`K_2(z)`$
$`K_2\left(z\right){\displaystyle \frac{2}{z^2}},z0`$ (44)
$`K_2\left(z\right)=\{\begin{array}{c}i\frac{\pi }{2}H_2^{(1)}\left(ze^{i\pi /2}\right),\pi <\mathrm{arg}(z)\frac{\pi }{2}\hfill \\ i\frac{\pi }{2}H_2^{(2)}\left(ze^{i\pi /2}\right),\frac{\pi }{2}<\mathrm{arg}(z)\pi \hfill \end{array}`$ (47)
where $`H_2^{(1)}(z)=J_2(z)+iY_2(z)`$ and $`H_2^{(2)}(z)=J_2(z)iY_2(z)`$. The asymptotic behaviour (44) implies that the $`\theta `$ integral of (43) will be zero in the $`ϵ0`$ limit for $`k>1`$ and the identity (47) allows evaluation of the discontinuity in the $`t`$ integral of (43), leading to the following instanton contribution to the Laplace sum-rules:
$`_1^{inst}\left(\tau \right)`$ $`=16\pi ^3{\displaystyle dn(\rho )\rho ^4\underset{0}{\overset{\mathrm{}}{}}tJ_2\left(\rho \sqrt{t}\right)Y_2\left(\rho \sqrt{t}\right)\mathrm{e}^{t\tau }dt}128\pi ^2{\displaystyle dn(\rho )}`$ (48)
$`=64\pi ^2{\displaystyle dn(\rho )a\mathrm{e}^a\left[(1+a)aK_0(a)+(2+2a+a^2)K_1(a)\right]}`$
$`_k^{inst}\left(\tau \right)`$ $`=16\pi ^3{\displaystyle }\mathrm{d}n(\rho )\rho ^4{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}t^{k+2}J_2\left(\rho \sqrt{t}\right)Y_2\left(\rho \sqrt{t}\right)\mathrm{e}^{t\tau }\mathrm{d}t,k>1`$ (49)
$`=\{\begin{array}{cc}& 128\pi ^2{\displaystyle }\mathrm{d}n(\rho ){\displaystyle \frac{a^4\mathrm{e}^a}{\rho ^2}}[\mathrm{\hspace{0.33em}2}aK_0(a)+(1+2a)K_1(a)],k=0\hfill \\ & 256\pi ^2{\displaystyle }\mathrm{d}n(\rho ){\displaystyle \frac{a^5\mathrm{e}^a}{\rho ^4}}[(94a)aK_0(a)+(3+7a4a^2)K_1(a)],k=1\hfill \end{array}`$
where $`a\rho ^2/(2\tau )`$ and where $`K_n`$ is the modified Bessel function of the first kind of order $`n`$ (c.f. ). Observe the symmetry between (48) and (49) broken by the term $`128\pi ^2dn(\rho )`$ appearing in (48)—a term which corresponds to the second integral on the right-hand side of (43) and which is nonzero *only* for $`k=1`$. Conversely, we note that naively substituting $`k=1`$ into (49) leads to an incorrect expression for the instanton contribution to the leading order sum-rule. This asymmetric role of the instanton contributions to $`k=1`$ and $`k>1`$ sum-rules is also a property of the LET as illustrated in (13,14).
As discussed in Section 1, we wish to determine whether the leading order sum-rule $`_1`$ might support the existence of a lowest-lying resonance whose presence is mass-suppressed in subsequent higher order $`k>1`$ sum-rules. Such is indeed the case, for example, for the pion within sum-rules based on a pseudoscalar $`\overline{q}\gamma _5q`$ current. Correspondingly, one might anticipate the identification of the lowest-lying scalar gluonium state with a $`500\text{}600\mathrm{MeV}`$ $`\sigma `$ (i.e. the lower-mass range of the $`f_0(4001200)`$) resonance whose contribution to higher order sum-rules is suppressed by additional factors of $`m_\sigma ^2`$ \[i.e. the additional factors of $`m_r^2`$ in (24)\], a scenario analogous to the $`m_\pi ^2`$ suppression of pion contributions to the pseudoscalar $`\overline{q}\gamma _5q`$ sum-rules . As already noted in Section 1, the LET constant in the absence of instantons supports this scenario for a sub-GeV scalar glueball.
However, the instanton contribution to the $`k=1`$ sum-rule is opposite in sign and comparable in magnitude to the LET subtraction constant $`\mathrm{\Pi }(0)`$, thereby ameliorating this term’s dominance of the lowest order sum-rule. For example, the contribution of instanton and LET terms to the $`k=1`$ sum rule in the dilute instanton liquid (DIL) model ,
$$dn(\rho )=n_c\delta (\rho \rho _c)d\rho ;n_c=8\times 10^4\mathrm{GeV}^4,\rho _c=\frac{1}{600\mathrm{MeV}}$$
(50)
renders trivial the remaining integrations in (4849). If we approximate $`J`$ by $`\alpha G^2`$ and employ a recently determined value of the gluon condensate
$$\alpha G^2=(0.07\pm 0.01)\mathrm{GeV}^4$$
(51)
we obtain via (1) and (5) an $`n_f=3`$ estimate of the LET subtraction constant:
$$\mathrm{\Pi }(0)\frac{32\pi }{9}\alpha G^2(0.78\pm 0.11)\mathrm{GeV}^4.$$
(52)
In Figure 2, we use (48) and the central value of (52) \[with $`n_c`$ and $`\rho _c`$ given in (50)\] to plot
$$\frac{\mathrm{\Pi }(0)+_1^{inst}(\tau )}{\mathrm{\Pi }(0)}$$
(53)
as a function $`\tau `$. We note that, as anticipated, instanton effects do indeed significantly reduce the impact of the LET on the $`k=1`$ sum-rule: anywhere from 20–65% for $`\tau `$ ranging between $`0.6\mathrm{GeV}^2`$ and $`1.0\mathrm{GeV}^2`$. Recalling that the dominance of the LET over $`𝒮_1`$ is responsible for the discrepancy in gluonium mass scales in the analysis of the $`k=1`$ and $`k>1`$ sum-rules, we see that suppression of the LET by instanton effects could reconcile this discrepancy, a possibility which is investigated further in the next section.
## 3 Phenomenological Impact of Instanton Effects in the Laplace Sum-Rules
Ratios of Laplace sum-rules provide a simple technique for extracting the mass of the lightest (narrow) resonance probed by the sum-rules. If only the lightest resonance (of mass $`m`$) is included in (23) and (24), then for the first few sum-rules we see that
$`{\displaystyle \frac{𝒮_1(\tau ,s_0)}{𝒮_0(\tau ,s_0)}}=m^2`$ (54)
$`{\displaystyle \frac{𝒮_0(\tau ,s_0)}{𝒮_1(\tau ,s_0)+\mathrm{\Pi }(0)}}=m^2.`$ (55)
This method of predicting the mass $`m`$ requires optimization of $`s_0`$ to minimize the $`\tau `$ dependence that can occur in the sum-rule ratios. However, a qualitative analysis which avoids these optimization issues occurs in the $`s_0\mathrm{}`$ limit where bounds on the mass $`m`$ can also be obtained. These bounds originate from inequalities satisfied on the hadronic physics side of the sum-rule because of the positivity of $`\rho ^{had}(t)`$. For example,
$`{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{s_0}{}}}𝑑tte^{t\tau }\rho ^{had}(t)={\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{s_0}{}}}𝑑t\left(ts_0+s_0\right)e^{t\tau }\rho ^{had}(t)s_0{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{t_0}{\overset{s_0}{}}}𝑑te^{t\tau }\rho ^{had}(t)`$
$`𝒮_1(\tau ,s_0)s_0𝒮_0(\tau ,s_0).`$ (56)
Furthermore, positivity of $`\mathrm{Im}\mathrm{\Pi }^{QCD}(t)`$ leads to an inequality for the continuum.
$`{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{s_0}{\overset{\mathrm{}}{}}}𝑑tte^{t\tau }\mathrm{Im}\mathrm{\Pi }^{QCD}(t)={\displaystyle \frac{1}{\pi }}{\displaystyle \underset{s_0}{\overset{\mathrm{}}{}}}𝑑t\left(ts_0+s_0\right)e^{t\tau }\mathrm{Im}\mathrm{\Pi }^{QCD}(t)s_0{\displaystyle \frac{1}{\pi }}{\displaystyle \underset{s_0}{\overset{\mathrm{}}{}}}𝑑te^{t\tau }\mathrm{Im}\mathrm{\Pi }^{QCD}(t)`$
$`c_1(\tau ,s_0)s_0c_0(\tau ,s_0)`$ (57)
These inequalities can be extended to include the $`k=1`$ sum-rules and continuum.
$`𝒮_0(\tau ,s_0)s_0\left[𝒮_1(\tau ,s_0)+\mathrm{\Pi }(0)\right]`$ (58)
$`c_0(\tau ,s_0)s_0c_1(\tau ,s_0)`$ (59)
We then see that
$$\frac{_1\left(\tau \right)}{_0\left(\tau \right)}=\frac{𝒮_1(\tau ,s_0)+c_1(\tau ,s_0)}{𝒮_0(\tau ,s_0)+c_0(\tau ,s_0)}=\frac{𝒮_1(\tau ,s_0)}{𝒮_0(\tau ,s_0)}\left[\frac{1+\frac{c_1(\tau ,s_0)}{𝒮_1(\tau ,s_0)}}{1+\frac{c_0(\tau ,s_0)}{𝒮_0(\tau ,s_0)}}\right]\frac{𝒮_1(\tau ,s_0)}{𝒮_0(\tau ,s_0)}=m^2$$
(60)
where the final inequality of (60) follows from $`c_1/𝒮_1s_0c_0/𝒮_1c_0/𝒮_0`$ via (57) and (56). Similarly, we find from (58) and (59) that
$$\frac{_0\left(\tau \right)}{_1\left(\tau \right)+\mathrm{\Pi }(0)}=\frac{𝒮_0(\tau ,s_0)+c_0(\tau ,s_0)}{𝒮_1(\tau ,s_0)+\mathrm{\Pi }(0)+c_1(\tau ,s_0)}\frac{𝒮_0(\tau ,s_0)}{𝒮_1(\tau ,s_0)+\mathrm{\Pi }(0)}=m^2.$$
(61)
Thus the ratios of the $`s_0\mathrm{}`$ limit of the sum-rules provide bounds on the mass in this single narrow resonance approximation. Extending the analysis to many narrow resonances alters (5455) so that the sum-rule ratios are an upper bound on the lightest resonance, upholding the bounds (6061) on the mass $`m^2`$ of the lightest resonance.
$`{\displaystyle \frac{_1\left(\tau \right)}{_0\left(\tau \right)}}m^2`$ (62)
$`{\displaystyle \frac{_0\left(\tau \right)}{_1\left(\tau \right)+\mathrm{\Pi }(0)}}m^2`$ (63)
The sum-rule bounds in (6263) can now be employed to determine the phenomenological impact of the instanton contributions on the sum-rule estimates of the lightest gluonium state, and to assess whether the suppression of the LET contribution by the instanton effects is sufficient to reduce the discrepancy between sum-rule analyses containing or omitting the $`k=1`$ sum-rule. Collecting results from equations (27,3840,48,49) the first few sum-rules $`_k(\tau )`$ are
$`_1(\tau )`$ $`={\displaystyle \frac{1}{\tau ^2}}\left[a_0+a_1\left(2+2\gamma _E\right)+a_2\left({\displaystyle \frac{\pi ^2}{2}}+6\gamma _E3\gamma _E^2\right)\right]+\left(b_0+b_1\gamma _E\right)Jc_0\tau 𝒪_6d_0{\displaystyle \frac{\tau ^2}{2}}𝒪_8`$
$`64\pi ^2{\displaystyle dn(\rho )a\mathrm{e}^a\left[(1+a)aK_0(a)+(2+2a+a^2)K_1(a)\right]}`$ (64)
$`_0(\tau )`$ $`={\displaystyle \frac{1}{\tau ^3}}\left[2a_0+a_1\left(6+4\gamma _E\right)+a_2\left(\pi ^26+18\gamma _E6\gamma _E^2\right)\right]{\displaystyle \frac{b_1}{\tau }}J+c_0𝒪_6+d_0\tau 𝒪_8`$
$`+128\pi ^2{\displaystyle dn(\rho )\frac{a^4\mathrm{e}^a}{\rho ^2}\left[\mathrm{\hspace{0.33em}2}aK_0(a)+(1+2a)K_1(a)\right]}`$ (65)
$`_1(\tau )`$ $`={\displaystyle \frac{1}{\tau ^4}}\left[6a_0+a_1\left(22+12\gamma _E\right)+a_2\left(3\pi ^236+66\gamma _E18\gamma _E^2\right)\right]{\displaystyle \frac{b_1}{\tau ^2}}Jd_0𝒪_8`$
$`+256\pi ^2{\displaystyle dn(\rho )\frac{a^5\mathrm{e}^a}{\rho ^4}\left[(94a)aK_0(a)+(3+7a4a^2)K_1(a)\right]}.`$ (66)
Renormalization-group improvement has been achieved by setting $`\nu ^2=1/\tau `$ in the correlation function and in the (three-loop, $`n_f=3`$, $`\overline{\mathrm{MS}}`$) running coupling $`\alpha `$:
$`{\displaystyle \frac{\alpha _s(\nu )}{\pi }}`$ $`=`$ $`{\displaystyle \frac{1}{\beta _0L}}{\displaystyle \frac{\overline{\beta }_1\mathrm{log}L}{\left(\beta _0L\right)^2}}+{\displaystyle \frac{1}{\left(\beta _0L\right)^3}}\left[\overline{\beta }_1^2\left(\mathrm{log}^2L\mathrm{log}L1\right)+\overline{\beta }_2\right]`$ (68)
$`L=\mathrm{log}\left({\displaystyle \frac{\nu ^2}{\mathrm{\Lambda }^2}}\right),\overline{\beta }_i={\displaystyle \frac{\beta _i}{\beta _0}},\beta _0={\displaystyle \frac{9}{4}},\beta _1=4,\beta _2={\displaystyle \frac{3863}{384}}`$
with $`\mathrm{\Lambda }_{\overline{MS}}300\mathrm{MeV}`$ for three active flavours, consistent with current estimates of $`\alpha _s(M_\tau )`$ and matching conditions through the charm threshold .
The nonperturbative QCD parameters are needed for further analysis of the sum-rules. We employ the DIL model parameters summarized in (50), as well as vacuum saturation for the dimension-8 gluon condensate
$$𝒪_8=14\left(\alpha f_{abc}G_{\mu \rho }^aG_{\nu \rho }^b\right)^2\left(\alpha f_{abc}G_{\mu \nu }^aG_{\rho \lambda }^b\right)^2=\frac{9}{16}\left(\alpha G^2\right)^2$$
(69)
and instanton estimates of the dimension-six condensate
$$𝒪_6=gf_{abc}G_{\mu \nu }^aG_{\nu \rho }^bG_{\rho \mu }^c=\left(0.27\mathrm{GeV}^2\right)\alpha G^2.$$
(70)
Finally, again using the approximation $`J=\alpha G^2`$ and the central gluon condensate value \[see (51)\] from reference , we find that the role of instanton contributions to the sum-rules (6466) is as illustrated in Figures 3, 4, and 5. In particular, we see that the instanton contributions diminish $`_1(\tau )+\mathrm{\Pi }(0)`$. The LET term $`\mathrm{\Pi }(0)`$, which leads to the asymptotic flattening of $`_1(\tau )+\mathrm{\Pi }(0)`$ at a value substantially different from zero when instantons are absent, is clearly suppressed by instanton effects in the large $`\tau `$ region. As noted earlier, such flattening over the $`\tau 1.0\mathrm{GeV}^2`$ region would be indicative via (23) of a sub-GeV lowest-lying resonance (i.e., $`m_r^2\tau 1`$ over the physically relevant region of $`\tau `$) Instanton effects no only undo this flattening, but also increase $`_0(\tau )`$ and alter the shape of $`_1(\tau )`$. The corresponding effects of instantons on the sum-rule ratios (6263) is shown in Figures 6 and 7. As expected from the instanton’s impact of lowering $`_1(\tau )`$ and elevating $`_0(\tau )`$, the ratio $`_0(\tau )/\left[_1(\tau )+\mathrm{\Pi }(0)\right]`$ is increased substantially by inclusion of instanton effects, increasing the corresponding upper bound on the mass of the lightest gluonium state. Instanton effects also serve to lower the ratio $`_1(\tau )/_0(\tau )`$, decreasing the corresponding upper bound on the mass of the lightest gluonium state.
Figures 8 and 9 summarize the ratio (mass bound) analysis in the presence and in the absence of instanton effects. It is evident that instanton effects lead to a substantial increase in the mass bound on the lightest gluonium state, but other important features emerge. For example, the instanton suppression of the LET term $`\mathrm{\Pi }(0)`$ reduces the discrepancy between the ratios including or omitting the $`k=1`$ sum-rule. Furthermore, a $`\tau `$-minimum stability plateau crucial for establishing a credible upper mass bound is seen to occur at reasonable energy scales ($`1/\sqrt{\tau }1.0\mathrm{GeV}`$) only when instanton effects are included. The ratios with instanton effects included (see Figure 8) are remarkably flat, suggesting that the mass bounds could be close to the mass prediction that would be obtained from a full sum-rule analysis incorporating the QCD continuum (i.e. $`s_0<\mathrm{}`$) in the phenomenological model.
## 4 Discussion
We have calculated the instanton contribution to the Laplace sum-rules of scalar gluonium and demonstrated explicitly how, for the lowest order $`k=1`$ sum-rule, this instanton contribution cancels part of the dominant LET constant.
As noted in the Introduction, a discrepancy between the lowest lying states evident from the lowest and from the next-to-lowest Laplace sum-rules may be indicative of two distinct states. Such is found to be the case, for example, in the pseudoscalar channel in which the pion dominates the leading ($`k=0`$) Laplace sum rule, but the $`\mathrm{\Pi }(1300)`$ resonance is found to dominate the next-to-leading ($`k=1`$) Laplace sum rule, because of a mass-suppression of the lowest-lying (pion) state in the latter sum rule . Moreover, analyses of the scalar gluonium channel in the absence of explicit instanton contributions seem to exhibit a similar discrepancy between leading $`k=1`$ and non-leading $`k>1`$ sum-rules .
Prior QCD sum-rule analyses of the instanton contribution to the scalar gluonium channel have focused either on the $`k=1`$ sum rule exclusively or the LET-insensitive $`k>1`$ sum-rules . Although these two analyses (which are separated by almost two decades) are both indicative of a lowest lying-resonance mass near or above $`1.4\mathrm{GeV}`$, their input content (parameter values and levels of perturbation theory) are necessarily different, suggesting the need for a single consistent treatment of leading $`k=1`$ and non-leading $`k>1`$ Laplace sum-rules in the scalar gluonium channel that is inclusive of instanton effects. We have shown here that careful consideration of the contribution arising from instantons within the $`k=1`$ sum rule in this channel leads to consistency with higher sum-rules in the estimation of lowest-lying resonance masses in the scalar gluonium channel.<sup>3</sup><sup>3</sup>3Note also that we have incorporated the significant NNLO perturbative corrections, as opposed to LO corrections in and NLO corrections in .
Correspondence with the prior treatment of the $`k=1`$ sum rule can be obtained by examining the instanton contribution (48) to this sum rule in the high-energy limit of small $`\tau `$. This contribution is obtained through evaluation of the integral in (48) in the large-$`a`$ \[$`\tau =\rho ^2/(2a)`$\] limit:
$`_1^{inst}(\tau )`$ $`=`$ $`128\pi ^2{\displaystyle 𝑑n(\rho )}32\pi ^3{\displaystyle 𝑑n(\rho )\underset{0}{\overset{\mathrm{}}{}}x^3J_2\left(x\right)Y_2\left(x\right)e^{\frac{x^2}{2a}}𝑑x}`$ (71)
$`=`$ $`64\pi ^2{\displaystyle 𝑑n(\rho )\left[\frac{\rho ^4}{4\tau ^2}\left[1+\frac{\rho ^2}{2\tau }\right]K_0\left(\frac{\rho ^2}{2\tau }\right)+\frac{\rho ^2}{2\tau }\left[2+\frac{\rho ^2}{\tau }+\frac{\rho ^4}{4\tau ^2}\right]K_1\left(\frac{\rho ^2}{2\tau }\right)\right]e^{\frac{\rho ^2}{2\tau }}}`$
$``$ $`16\pi ^{\frac{5}{2}}{\displaystyle 𝑑n(\rho )e^{\frac{\rho ^2}{\tau }}\rho ^5\tau ^{\frac{5}{2}}\left[1+𝒪\left(\frac{\tau }{\rho ^2}\right)\right]},\tau \rho ^2.`$ (72)
In the instanton liquid model \[$`dn(\rho )=n_c\delta \left(\rho \rho _c\right)d\rho `$\], (72) is consistent with eq. (42) of reference for the instanton contribution to the $`k=1`$ sum-rule, which was utilized to anticipate the $`(1.6\mathrm{GeV})`$ lattice prediction of the scalar gluonium mass. Figure 10 provides a comparison of this asymptotic form with the exact expression (71) for $`_1^{inst}(\tau )`$ under the instanton-liquid assumption. Of particular interest is the difference between the two expressions over the range $`2a3`$, corresponding to the $`0.4\mathrm{GeV}^2\tau 0.6\mathrm{GeV}^2`$ range in Figure 8 for which $`\sqrt{_0/[_1+\mathrm{\Pi }(0)]}`$ is flat. Although both the asymptotic expression of and the exact expression provide negative contributions which mitigate dominance of the positive LET contribution over $`_1`$, the exact expression is substantially larger in magnitude over the region of phenomenological interest. We speculate that such an increase relative to the analysis of serves to compensate for the larger phenomenological value at present for the gluon condensate within the low-energy theorem term (52), although it may also compensate the non-leading perturbative contributions in (26) not known at the time of .
Of course, more sophisticated expressions for the instanton contributions have been utilized to generate consistent scalar gluonium phenomenology both within and complementary to a QCD sum rule framework, as noted above. The key point of the work presented here is the reconciliation of the $`k=1`$ sum rule with higher-$`k`$ sum-rules. We reiterate the approximate consistency between scalar-gluonium masses obtained from the flattened regions of the $`\sqrt{_0/[_1+\mathrm{\Pi }(0)]}`$ and $`\sqrt{_1/_0}`$ curves of Figure 8, as well as the drastic reduction of the scalar gluonium mass evident in the curves of Figure 9 when instanton contributions are omitted. Roughly speaking, such contributions account for half the lowest-lying scalar-gluonium mass within a sum-rule context.
Acknowledgements: The authors are grateful for research support from the Natural Sciences and Engineering Research Council of Canada (NSERC). We are also grateful to N. Kochelev for helpful correspondence.
|
warning/0007/hep-ph0007205.html
|
ar5iv
|
text
|
# On the masses, branching ratios, and full widths of heavy 𝜌', 𝜌'' and 𝜔', 𝜔'' resonances.
## Abstract
Based on the previous and recent fits of multiple data on the reactions of $`e^+e^{}`$ annihilation, $`\tau `$ lepton decay, and the reaction $`K^{}p\pi ^+\pi ^{}\mathrm{\Lambda }`$, the magnitude of the branching ratios and total widths of the isovector $`\rho ^{}`$, $`\rho ^{\prime \prime }`$, and the isoscalar $`\omega ^{}`$, $`\omega ^{\prime \prime }`$ resonances are calculated. Some topics on the spectroscopy of the $`\rho (1450)`$ and $`\omega (1420)`$ states are discussed.
The situation with the resonances $`\rho ^{}\rho _1^{}\rho (1450)`$, $`\rho ^{\prime \prime }\rho _2^{}\rho (1700)`$, with isospin one and those $`\omega ^{}\omega _1^{}\omega (1420)`$, $`\omega ^{\prime \prime }\omega _2^{}\omega (1600)`$, with isospin zero is still far from being clear. Although the characteristic peaks corresponding to these resonances were observed in a number of channels of the one-photon $`e^+e^{}`$ annihilation, $`\tau `$ lepton decays, $`N\overline{N}`$ annihilation, photoproduction, etc., the specific masses and branching ratios are not properly established and hence are not given in the Tables . Recently, the present authors undertook an attempt to fit then existing data on $`e^+e^{}`$ annihilation, $`\tau `$ lepton decays, and the reaction $`K^{}p\pi ^+\pi ^{}\mathrm{\Lambda }`$, where the above heavy resonances were observed, in the framework of the unified approach. The approach is based on the scheme that takes into account both energy dependence of the partial widths and the mixing via common decay modes among the latter and the ground state $`\rho (770)`$ and $`\omega (782)`$ resonances, in the respective isovector and isoscalar channels. When so doing, the masses and coupling constants of the resonances were taken as free parameters to be determined from the fit. This choice is convenient because, first, it is just these parameters that are essential in identifying the nature of heavy resonances with $`J^{CP}=1^{}`$. Second, the function $`\chi ^2`$ used in obtaining the intervals of the variation of the extracted parameters is represented in this case as the sum of independent contributions, so that the covariance matrix (see the statistics section in Ref. ) is diagonal. However, the results of the measurements are often represented in terms of the masses and branching ratios. In view of the excessive size of the publications we refrained at that time from the calculation of the partial widths with the parameters found in . Here we fill this gap and calculate the branching ratios and total widths of the $`\rho _{1,2}^{}`$ and $`\omega _{1,2}^{}`$ resonances using the new data on the cross section of the reactions $`e^+e^{}\pi ^+\pi ^{}\pi ^0`$ , $`e^+e^{}\omega \pi ^0`$ , and the data on the spectral function $`v_1`$ measured in the decays $`\tau ^{}\omega \pi ^{}\nu _\tau `$, $`\tau ^{}\pi ^+\pi ^{}\pi ^{}\pi ^0\nu _\tau `$ . We also comment on the issue of why the resonances with the masses greater than 1400 MeV should be wide in the conventional $`q\overline{q}`$ model, and why their bare masses get shifted towards the greater values as compared to the visible peak positions. The latter point, as will be shown below, is of the direct relevance to the hadronic spectroscopy of the $`\rho (1300)`$ resonance reported by the LASS group , and the $`\omega (1200)`$ resonance reported by the SND team .
First, let us try to understand why the situation with the resonances possessing the masses greater than 1400 MeV is so complicated despite the numerous experiments aimed at the determination of their masses and partial widths. As is known , the indications on the existance of the $`\rho _{1,2}^{}`$ and $`\omega _{1,2}^{}`$ resonances were obtained, in particular, in the $`VP`$ channels, where $`V`$, $`P`$ standing, respectively, for the vector, pseudoscalar meson. It is also known that the typical magnitude of the $`VVP`$ coupling constant is $`g_{\omega \rho \pi }14.3`$ GeV <sup>-1</sup>. All other $`VVP`$ coupling constants can be expressed through $`g_{\omega \rho \pi }`$ via the SU(3) Clebsh-Gordan coefficients or, equivalently, the quark model relations. From the point of view of the simplest $`q\overline{q}`$ quark model, there are no reasons to expect that the $`V_{1,2}^{}VP`$ coupling constants should be drastically suppressed as compared to the $`VVP`$ ones . Hence, taking $`g_{\rho _{1,2}^{}\omega \pi }g_{\omega _{1,2}^{}\rho \pi }10`$ GeV <sup>-1</sup>, and $`m_{\rho _1^{}}m_{\omega _1^{}}=1400`$ MeV, one finds
$`\mathrm{\Gamma }_{\rho _1^{}\omega \pi }`$ $`=`$ $`g_{\rho _1^{}\omega \pi }^2q_{\omega \pi }^3(m_{\rho _1^{}})/12\pi 280\text{ MeV},`$ (1)
$`\mathrm{\Gamma }_{\omega _1^{}\rho \pi }`$ $`=`$ $`g_{\omega _1^{}\rho \pi }^2q_{\rho \pi }^3(m_{\omega _1^{}})/4\pi 820\text{ MeV}.`$ (2)
Here
$`q_{bc}(m_a)`$ $`=`$ $`\{[m_a^2(m_b+m_c)^2]`$ (4)
$`\times [m_a^2(m_bm_c)^2]\}^{1/2}/2m_a`$
is the momentum of the final particle $`b`$ or $`c`$ in the rest frame system of the particle $`a`$, in the decay $`ab+c`$, and the three isotopic modes in the $`\omega _1^{}\rho \pi `$ decay are taken into account. To appreciate the rapid growth of the $`VP`$ widths with energy, their evaluation, assuming the masses 1200 MeV, gives, respectively, 92 and 295 MeV. Analogously, assuming that $`m_{\rho _2^{}}m_{\omega _2^{}}=1750`$ MeV, one finds
$`\mathrm{\Gamma }_{\rho _2^{}\omega \pi }`$ $``$ $`880\text{ MeV},`$ (5)
$`\mathrm{\Gamma }_{\omega _2^{}\rho \pi }`$ $``$ $`2600\text{ MeV}.`$ (6)
Since the $`VP`$ decay modes are not the only ones to which heavy resonances can decay , the resonances $`\rho _{1,2}^{}`$ and $`\omega _{1,2}^{}`$, in fact, should be rather wide. The large width of a resonance is one of the obstacles in its identification, because such resonance often reveals itself as the rather smooth feature in the energy behavior of the cross section.
The second obstacle, as was pointed out in , is the shift of the resonance peak position from the input value of the resonance mass. Indeed, let us consider, for simplicity, the single resonance $`R`$ with the bare mass $`m_R`$ observed in some channel $`f`$ of $`e^+e^{}`$ annihilation, $`e^+e^{}Rf`$. Then the cross section of the above process can be written as
$`\sigma (s)`$ $`=`$ $`12\pi m_R^3\mathrm{\Gamma }_{Rl^+l^{}}(m_R)g_{Rf}^2`$ (8)
$`\times {\displaystyle \frac{s^{3/2}W_{Rf}(s)}{(sm_R^2)^2+s\mathrm{\Gamma }_R^2(s)}},`$
where $`s`$ is the square of the total center-of-mass energy, $`\mathrm{\Gamma }_{Rl^+l^{}}(m_R)`$ is the leptonic width of the resonance evaluated at $`\sqrt{s}=m_R`$. The partial hadronic width of the decay $`Rf`$ is represented in the form
$$\mathrm{\Gamma }_{Rf}(s)=g_{Rf}^2W_{Rf}(s),$$
where $`g_{Rf}`$ is the coupling constant of $`R`$ with the final state $`f`$, and $`W_{Rf}(s)`$ being the dynamical phase space factor of the decay $`Rf`$ that includes the possible resonance intermediate states as, for example, in the decay $`\omega \rho \pi 3\pi `$. The total width of the resonance is
$$\mathrm{\Gamma }_R(s)=\underset{f}{}\mathrm{\Gamma }_{Rf}(s).$$
The peak position is given by the condition of the vanishing derivative of $`\sigma (s)`$ with respect to $`s`$:
$$sm_R^2=\frac{1}{G(s)}\left\{1\pm \left[1+s\mathrm{\Gamma }_R^2F(s)G(s)\right]^{1/2}\right\},$$
(9)
where
$`G(s)`$ $`=`$ $`\left[\mathrm{ln}\left(s^{3/2}W_{Rf}\right)\right]^{},`$ (10)
$`F(s)`$ $`=`$ $`\left[\mathrm{ln}\left(s^{5/2}\mathrm{\Gamma }_R^2/W_{Rf}\right)\right]^{}.`$ (11)
Hereafter, $`\mathrm{\Gamma }_R\mathrm{\Gamma }_R(s)`$, $`W_{Rf}W_{Rf}(s)`$, and prime denotes the differentiation with respect to $`s`$. Eq. (9) should match with the usual expression $`sm_R^2=0`$ in the limit of slow varying narrow width, hence the lower minus sign should be chosen in Eq. (9). The latter is still very complicated equation for the determination of the peak position, so the numerical methods should be invoked for its solution. However, all necessary qualitative conclusions can be drawn upon approximating the right hand side of this equation by taking its value at $`s=m_R^2`$. One can convince oneself that the dominant decay modes of heavy resonances have the phase space factors growing faster than the decrease of the leptonic width as $`s^{3/2}`$. Then the function $`G(s)`$ is positive. Also positive is the function $`F(s)`$. Hence, the factor following $`G^1(s)`$ in Eq. (9) is negative. To first order in the derivative of the phase space volume, one finds from Eq. (9) the peak position $`s_R`$:
$$s_Rm_R^2\frac{1}{2}s\mathrm{\Gamma }_R^2F(s)|_{s=m_R^2}.$$
(12)
One can see that in the case of the sufficiently narrow ($`\mathrm{\Gamma }_R<200`$ MeV) resonance such as $`\rho (770)`$, $`\omega (782)`$, and $`\varphi (1020)`$ one, the peak position, with a good accuracy, coincides with the bare mass $`m_R`$.
The situation changes when the resonance is wide, as it takes place in the case of the $`\rho _{1,2}^{}`$ and the $`\omega _{1,2}^{}`$ one. See Eqs. (2) and (6). The peak position is shifted towards the lower value as compared to the magnitude of the bare mass. This is just what was revealed in the fits . The more the width of the resonance width is, the more it is shifted, so that, say, the $`\rho _2^{}`$ and $`\omega _2^{}`$ resonances with the bare masses around 1900 MeV are revealed as the peaks at 1500 $``$ 1600 MeV.
After these preliminary remarks let us present the results of the calculation of the branching ratios and full widths. As in Refs. , the results are given for each channel where the indication on the specific resonance exists. The procedure presented here is as follows. First, we calculate the partial widths and their errors from the errors of the masses and coupling constants found in . Second, upon dividing each of these partial widths by the sum of their central values, the branching ratios and their errors are calculated. The leptonic widths of the $`\rho _{1,2}^{}`$ are taken from , while those of $`\omega _{1,2}^{}`$ are calculated from the extracted masses and leptonic coupling constants $`f_{\omega _{1,2}^{}}`$ in according to the expression
$$\mathrm{\Gamma }_{\omega _{1,2}^{}}=\frac{4\pi \alpha ^2m_{\omega _{1,2}^{}}}{3f_{\omega _{1,2}^{}}^2}.$$
Furthermore, as compared to Ref. , the recent data on the reactions $`e^+e^{}\pi ^+\pi ^{}\pi ^0`$ , $`e^+e^{}\omega \pi ^0`$ , and $`\tau `$ lepton decays are used to extract the necessary resonance parameters. The results of the evaluation of the branching ratios and full widths are shown in Tables I, II, III. One can see that the simple qualitative estimates displayed in Eqs. (2) and (6), assuming the modes besides the $`VP`$ one are included, agree with the results presented there.
Now, some remarks on the spectroscopy of the heavier vector mesons are in order. First, the $`\rho (1300)`$ state reported by the LASS detector team who studied the reaction $`K^{}p\pi ^+\pi ^{}\mathrm{\Lambda }`$, revived an old discussion concerning the possible existence of the $`\rho (1250)`$ meson, in addition to the $`\rho (1450)`$ claimed to be observed in $`e^+e^{}`$ annihilation. The results presented in Table I show that the corresponding peak observed by the LASS group should be attributed to the same state $`\rho (1450)`$ as that presented in Reviews of Particle Physics (RPP).
The similar situation is with the state $`\omega (1200)`$ observed recently by the SND team in the reaction $`e^+e^{}\pi ^+\pi ^{}\pi ^0`$ . Since the rather old data on the latter reaction were used in the fit , the new data on this reaction are included in the fit done to consider the $`\omega (1200)`$ state in the framework of the approach , however, upon neglecting the contributions of the $`\varphi _{1,2}^{}`$ resonances, because their couplings in the above reaction were found to be consistent with zero . The parameters of the $`\omega _{1,2}^{}`$ resonances extracted from this new fit coincide, within errors with those reported earlier . They are used in filling the corresponding entries in Table III. The corresponding curves are plotted in Fig. 1. Our conclusion is that the $`\omega (1200)`$ state observed by SND is the same state as $`\omega (1420)\omega _1^{}`$ presented in RPP . The shift of the visible peak as compared to the bare mass of the resonance should be attributed, as is explained above, to the rather large width and the rapid growth of the partial widths with the energy increase. As far as the $`\omega _2^{}`$ resonance is concerned, its huge width, see Table II, results, in accord with Eq. (12), in the shift of the peak to $`1600`$ MeV from the bare mass $`19002000`$ MeV. The large partial widths found in the analysis , are also in the qualitative agreement with the expectations presented in Eq. (2) and (6).
Using the recent SND data , one can draw some conclusions about the characteristics of the radial wave function of the bound $`q\overline{q}`$ state at the origin. Since the $`\omega _1^{}`$ resonance is usually attributed to the $`2^3S_1`$ state , one finds, using Ref. ,
$$|R_S(0)|^2=6\pi m_{\omega _1^{}}^3/f_{\omega _1^{}}^2=(25\pm 10)\times 10^3\text{ GeV}^3,$$
(13)
to be compared to $`(38\pm 4)\times 10^3`$ GeV<sup>3</sup> computed for $`\omega (782)`$ meson. Within errors, both above figures agree with each other. On the other hand, the $`\omega _2^{}`$ state is treated as the $`1^3D_1`$ one, hence the second derivative of the radial wave function at the origin is appropriate :
$$|R_D^{\prime \prime }(0)|^2=\frac{3\pi m_{\omega _2^{}}^7}{25f_{\omega _2^{}}^2}=(10\pm 8)\times 10^3\text{ GeV}^7.$$
(14)
Note that the S-wave characteristics of the $`\omega _2^{}`$ resonance analogous to Eq. (13) is $`|R_S(0)|^2=(35\pm 23)\times 10^3`$ GeV<sup>3</sup>.
The recent data on the reaction $`e^+e^{}\omega \pi ^0`$ have been included in the fit. The resonance parameters are found to agree within errors with those obtained in Ref. . Specifically, the $`\rho _1^{}`$ resonance is not revealed, since its extracted parameters are consistent with zero within very large errors. Hence we exclude $`\rho _1^{}`$ from the fit, leaving only the $`\rho (770)+\rho _2^{}`$ contribution. Corresponding curve is shown in Fig. 2. Note that the central value of the range parameter $`R`$ entering the formfactor
$$C_{\rho \omega \pi }(E)=\frac{1+(Rm_\rho )^2}{1+(RE)^2},$$
(15)
the latter introduced to restrict the growth of the $`\rho \omega \pi ^0`$ partial widths with the energy increase, $`\mathrm{\Gamma }_{\rho \omega \pi }(E)C_{\rho \omega \pi }^2(E)\mathrm{\Gamma }_{\rho \omega \pi }(E)`$, turns out to be zero in our fit. To be more precise, the data allow its variation within the interval from 0 to 0.37 GeV<sup>-1</sup>. The difference with the conclusion reached in Ref. that the best fit requires nonzero $`R`$ is, in our opinion, attributed to the fact that, first, only two decay modes, $`\pi ^+\pi ^{}`$ and $`\omega \pi ^0`$ were included in , while we include all known ones (see Tables I and II), and the mixing arising due to common decay modes neglected in Ref. . Second, the same range parameter $`R`$ was used for the $`\pi ^+\pi ^{}`$ and $`\omega \pi ^0`$ modes and for all the $`\rho `$ -like resonance, while we have assumed it to be different, and in the case of the $`\pi ^+\pi ^{}`$ state it was set to zero, because this mode grows rather moderately with the energy increase .
The joint fit of the CLEO data results in the considerably improved accuracy of the determination of the $`\rho _2^{}`$ resonance parameters as compared to the earlier fit of the ARGUS data . Here also the $`\rho _1^{}`$ resonance is unnecessary, and the data are well described by the only $`\rho _2^{}`$ resonance in addition to the $`\rho (770)`$. The results are shown in Table II and Fig. 3. The shift of the visible peak towards the lower invariant mass of the $`4\pi `$ system around 1500 MeV is explained by the discussed effect exemplified by Eq. (12).
Our conclusions are as follows. First, the states $`\rho _{1,2}^{}`$ and $`\omega _{1,2}^{}`$ turn out to be the wide resonance structures, as if the conventional quark picture of them as the radial excitations is implied. In this respect, the present results, having in mind their significant uncertainties, do not contradict to the assignment of $`\rho _1^{}`$ and $`\omega _1^{}`$ resonances to the state $`2^3S_1`$ . In the meantime, the resonances $`\rho _2^{}`$ and $`\omega _2^{}`$ are found to be wide, which contradicts to the assigning them to the state $`1^3D_1`$ predicted to be relatively narrow . Second, one should be careful in attributing the specific peak or structure in the cross section to the specific spectroscopy state, because the large width, the rapid growth of the phase space with the energy increase, and the mixing among the resonances result in the shift of the visible peaks in the cross sections. Third, the very large widths of the resonances found in the present paper may indirectly evidence in favor of some nonresonant contributions to the amplitudes. The accuracy of the existing data is still poor to isolate such contributions reliably. The forthcoming improvement of the accuracy of the data in the energy range 1400 $``$ 2000 MeV will hopefully permit one to specify the above nonresonant contributions (if any) and to test the whole resonance interpretation of the high mass states.
###### Acknowledgements.
We are grateful to V. P. Druzhinin, S. I. Eidelman, E. V. Pakhtusova and S. I. Serednyakov for the discussions. The present work is supported in part by the grant RFBR-INTAS IR-97-232.
|
warning/0007/math0007142.html
|
ar5iv
|
text
|
# An excursion from enumerative geometry to solving systems of polynomial equations with Macaulay 2
## 1. Introduction
A basic question to ask about a system of polynomial equations is its number of solutions. For this, the fundamental result is the following Bézout Theorem.
###### Theorem 1.1.
The number of isolated solutions to a system of polynomial equations
$$f_1(x_1,\mathrm{},x_n)=f_2(x_1,\mathrm{},x_n)=\mathrm{}=f_n(x_1,\mathrm{},x_n)=0$$
is bounded by $`d_1d_2\mathrm{}d_n`$, where $`d_i:=\mathrm{deg}f_i`$. If the polynomials are generic, then this bound is attained for solutions in an algebraically closed field.
Here, isolated is taken with respect to the algebraic closure. This Bézout Theorem is a consequence of the refined Bézout Theorem of Fulton and MacPherson \[11, §1.23\].
A system of polynomial equations with fewer than this degree bound or Bézout number of solutions is called deficient, and there are well-defined classes of deficient systems that satisfy other bounds. For example, fewer monomials lead to fewer solutions, for which polyhedral bounds on the number of solutions are often tighter (and no weaker than) the Bézout number, which applies when all monomials are present. When the polynomials come from geometry, determining the number of solutions is the central problem in enumerative geometry.
Symbolic computation can help compute the solutions to a system of equations that has only isolated solutions. In this case, the polynomials generate a zero-dimensional ideal $`I`$. The degree of $`I`$ ($`dim_kk[X]/I`$), which is the number of standard monomials in any term order gives an upper bound on the number of solutions, which is attained when $`I`$ is radical.
###### Example 1.2.
We illustrate this discussion with an example. Let $`f_1`$, $`f_2`$, $`f_3`$, and $`f_4`$ be random quadratic polynomials in the ring $`𝔽_{101}[y_{11},y_{12},y_{21},y_{22}]`$.
i1 : R = ZZ/101\[y11, y12, y21, y22\];
i2 : PolynomialSystem = apply(1..4, i -\>
random(0, R) + random(1, R) + random(2, R));
The ideal they generate has dimension 0 and degree $`16=2^4`$, which is the Bézout number.
i3 : I = ideal PolynomialSystem;
o3 : Ideal of R
i4 : dim I, degree I
o4 = (0, 16)
o4 : Sequence
If we restrict the monomials which appear in the $`f_i`$ to be among
$$1,y_{11},y_{12},y_{21},y_{22},y_{11}y_{22},\text{ and }y_{12}y_{21},$$
then the ideal they generate again has dimension 0, but its degree is now 4.
i5 : J = ideal (random(R^4, R^7) * transpose(
matrix{{1, y11, y12, y21, y22, y11\*y22, y12\*y21}}));
o5 : Ideal of R
i6 : dim J, degree J
o6 = (0, 4)
o6 : Sequence
If we further require that the coefficients of the quadratic terms sum to zero, then the ideal they generate now has degree 2.
i7 : K = ideal (random(R^4, R^6) * transpose(
matrix{{1, y11, y12, y21, y22, y11\*y22 - y12\*y21}}));
o7 : Ideal of R
i8 : dim K, degree K
o8 = (0, 2)
o8 : Sequence
In Example 4.2, we shall see how this last specialization is geometrically meaningful.
For us, enumerative geometry is concerned with enumerating geometric figures of some kind having specified positions with respect to general fixed figures. That is, counting the solutions to a geometrically meaningful system of polynomial equations. We use Macaulay 2 to investigate some enumerative geometric problems from this point of view. The problem of enumeration will be solved by computing the degree of the (0-dimensional) ideal generated by the polynomials.
## 2. Solving systems of polynomials
We briefly discuss some aspects of solving systems of polynomial equations. For a more complete survey, see the relevant chapters in .
Given an ideal $`I`$ in a polynomial ring $`k[X]`$, set $`𝒱(I):=\mathrm{Spec}k[X]/I`$. When $`I`$ is generated by the polynomials $`f_1,\mathrm{},f_N`$, $`𝒱(I)`$ gives the set of solutions in affine space to the system
(1)
$$f_1(X)=\mathrm{}=f_N(X)=0$$
a geometric structure. These solutions are the roots of the ideal $`I`$. The degree of a zero-dimensional ideal $`I`$ provides an algebraic count of its roots. The degree of its radical counts roots in the algebraic closure, ignoring multiplicities.
### 2.1. Excess intersection
Sometimes, only a proper (open) subset of affine space is geometrically meaningful, and we want to count only the meaningful roots of $`I`$. Often the roots $`𝒱(I)`$ has positive dimensional components that lie in the complement of the meaningful subset. One way to treat this situation of excess or improper intersection is to saturate $`I`$ by a polynomial $`f`$ vanishing on the extraneous roots. This has the effect of working in $`k[X][f^1]`$, the coordinate ring of the complement of $`𝒱(f)`$ \[9, Exer. 2.3\].
###### Example 2.1.
We illustrate this with an example. Consider the following ideal in $`𝔽_7[x,y]`$.
i9 : R = ZZ/7\[y, x, MonomialOrder=\>Lex\];
i10 : I = ideal (y^3\*x^2 + 2\*y^2\*x + 3\*x\*y, 3\*y^2 + x\*y - 3\*y);
o10 : Ideal of R
Since the generators have greatest common factor $`y`$, $`I`$ defines finitely many points together with the line $`y=0`$. Saturate $`I`$ by the variable $`y`$ to obtain the ideal $`J`$ of isolated roots.
i11 : J = saturate(I, ideal(y))
4 3 2
o11 = ideal (x + x + 3x + 3x, y - 2x - 1)
o11 : Ideal of R
The first polynomial factors completely in $`𝔽_7[x]`$,
i12 : factor(J\_0)
o12 = (x - 2)(x + 1)(x + 2)(x)(1)
o12 : Product
and so the isolated roots of $`I`$ are $`(0,1),(2,5),(5,4)`$, and $`(6,6)`$.
Here, the extraneous roots came from a common factor in both equations. A less trivial example of this phenomenon will be seen in Section 5.2.
### 2.2. Elimination, rationality, and solving
Elimination theory can be used to study the roots of a zero-dimensional ideal $`Ik[X]`$. A polynomial $`hk[X]`$ defines a map $`k[y]k[X]`$ (by $`yh`$) and a corresponding projection $`h:\mathrm{Spec}k[X]𝔸^1`$. The generator $`g(y)k[y]`$ of the kernel of the map $`k[y]k[X]/I`$ is called an eliminant and it has the property that $`𝒱(g)=h(𝒱(I))`$. When $`h`$ is a coordinate function $`x_i`$, we may consider the eliminant to be in the polynomial ring $`k[x_i]`$, and we have $`g(x_i)=Ik[x_i]`$. The most important result concerning eliminants is the Shape Lemma .
Shape Lemma. Suppose $`h`$ is a linear polynomial and $`g`$ is the corresponding eliminant of a zero-dimensional ideal $`Ik[X]`$ with $`\mathrm{deg}(I)=\mathrm{deg}(g)`$. Then the roots of $`I`$ are defined in the splitting field of $`g`$ and $`I`$ is radical if and only if $`g`$ is square-free.
Suppose further that $`h=x_1`$ so that $`g=g(x_1)`$. Then, in the lexicographic term order with $`x_1<x_2<\mathrm{}<x_n`$, $`I`$ has a Gröbner basis of the form:
(2)
$$g(x_1),x_2g_2(x_1),\mathrm{},x_ng_n(x_1),$$
where $`\mathrm{deg}(g)>\mathrm{deg}(g_i)`$ for $`i=2,\mathrm{},n`$.
When $`k`$ is infinite and $`I`$ is radical, an eliminant $`g`$ given by a generic linear polynomial $`h`$ will satisfy $`\mathrm{deg}(g)=\mathrm{deg}(I)`$. Enumerative geometry counts solutions when the fixed figures are generic. We are similarly concerned with the generic situation of $`\mathrm{deg}(g)=\mathrm{deg}(I)`$. In this case, eliminants provide a useful computational device to study further questions about the roots of $`I`$. For instance, the Shape Lemma holds for the ideal of Example 2.1. Its eliminant, which is the polynomial J\_0, factors completely over the ground field $`𝔽_7`$, so all four solutions are defined in $`𝔽_7`$. In Section 4.3, we will use eliminants in another way, to show that an ideal is radical.
Given a polynomial $`h`$ in a zero-dimensional ring $`k[X]/I`$, the procedure eliminant(h, k\[y\]) finds a linear relation modulo $`I`$ among the powers $`1,h,h^2,\mathrm{},h^d`$ of $`h`$ with $`d`$ minimal and returns this as a polynomial in $`k[y]`$. This procedure is included in the Macaulay 2 package realroots.m2.
i13 : load "realroots.m2"
i14 : code eliminant
o14 = -- code for eliminant:
-- realroots.m2:65-81
eliminant = (h, C) -\> (
Z := C\_0;
A := ring h;
assert( dim A == 0 );
F := coefficientRing A;
assert( isField F );
assert( F == coefficientRing C );
B := basis A;
d := numgens source B;
M := fold((M, i) -\> M ||
substitute(contract(B, h^(i+1)), F),
substitute(contract(B, 1\_A), F),
flatten subsets(d, d));
N := ((ker transpose M)).generators;
P := matrix {toList apply(0..d, i -\> Z^i)} * N;
(flatten entries(P))\_0
)
o14 : Net
Here, M is a matrix whose rows are the normal forms of the powers $`1`$, $`h`$, $`h^2`$, $`\mathrm{}`$, $`h^d`$ of $`h`$, for $`d`$ the degree of the ideal. The columns of the kernel N of transpose M are a basis of the linear relations among these powers. The matrix P converts these relations into polynomials. Since N is in column echelon form, the initial entry of P is the relation of minimal degree. (This method is often faster than naïvely computing the kernel of the map $`k[Z]A`$ given by $`Zh`$, which is implemented by eliminantNaive(h, Z).
Suppose we have an eliminant $`g(x_1)`$ of a zero-dimensional ideal $`Ik[X]`$ with $`\mathrm{deg}(g)=\mathrm{deg}(I)`$, and we have computed the lexicographic Gröbner basis (2). Then the roots of $`I`$ are
(3)
$$\{(\xi _1,g_2(\xi _1),\mathrm{},g_n(\xi _1))g(\xi _1)=0\}.$$
Suppose now that $`k=`$ and we seek floating point approximations for the (complex) roots of $`I`$. Following this method, we first compute floating point solutions to $`g(\xi )=0`$, which give all the $`x_1`$-coordinates of the roots of $`I`$, and then use (3) to find the other coordinates. The difficulty here is that enough precision may be lost in evaluating $`g_i(\xi _1)`$ so that the result is a poor approximation for the other components $`\xi _i`$.
### 2.3. Solving with linear algebra
We describe another method based upon numerical linear algebra. When $`Ik[X]`$ is zero-dimensional, $`A=k[X]/I`$ is a finite-dimensional $`k`$-vector space, and any Gröbner basis for $`I`$ gives an efficient algorithm to compute ring operations using linear algebra. In particular, multiplication by $`hA`$ is a linear transformation $`m_h:AA`$ and the command regularRep(h) from realroots.m2 gives the matrix of $`m_h`$ in terms of the standard basis of $`A`$.
i15 : code regularRep
o15 = -- code for regularRep:
-- realroots.m2:97-102
regularRep = f -\> (
assert( dim ring f == 0 );
b := basis ring f;
k := coefficientRing ring f;
substitute(contract(transpose b, f\*b), k)
)
o15 : Net
Since the action of $`A`$ on itself is faithful, the minimal polynomial of $`m_h`$ is the eliminant corresponding to $`h`$. The procedure charPoly(h, Z) in realroots.m2 computes the characteristic polynomial $`det(ZIdm_h)`$ of $`h`$.
i16 : code charPoly
o16 = -- code for charPoly:
-- realroots.m2:108-116
charPoly = (h, Z) -\> (
A := ring h;
F := coefficientRing A;
S := F\[Z\];
Z = value Z;
mh := regularRep(h) \** S;
Idz := S\_0 * id\_(S^(numgens source mh));
det(Idz - mh)
)
o16 : Net
When this is the minimal polynomial (the situation of the Shape Lemma), this procedure often computes the eliminant faster than does eliminant, and for systems of moderate degree, much faster than naïvely computing the kernel of the map $`k[Z]A`$ given by $`Zh`$.
The eigenvalues and eigenvectors of $`m_h`$ give another algorithm for finding the roots of $`I`$. The engine for this is the following result.
Stickelberger’s Theorem. Let $`hA`$ and $`m_h`$ be as above. Then there is a one-to-one correspondence between eigenvectors $`𝐯_\xi `$ of $`m_h`$ and roots $`\xi `$ of $`I`$, the eigenvalue of $`m_h`$ on $`𝐯_\xi `$ is the value $`h(\xi )`$ of $`h`$ at $`\xi `$, and the multiplicity of this eigenvalue (on the eigenvector $`𝐯_\xi `$) is the multiplicity of the root $`\xi `$.
Since the linear transformations $`m_h`$ for $`hA`$ commute, the eigenvectors $`𝐯_\xi `$ are common to all $`m_h`$. Thus we may compute the roots of a zero-dimensional ideal $`Ik[X]`$ by first computing floating-point approximations to the eigenvectors $`𝐯_\xi `$ of $`m_{x_1}`$. Then the root $`\xi =(\xi _1,\mathrm{},\xi _n)`$ of $`I`$ corresponding to the eigenvector $`𝐯_\xi `$ has $`i`$th coordinate satisfying
(4)
$$m_{x_i}𝐯_\xi =\xi _i𝐯_\xi .$$
An advantage of this method is that we may use structured numerical linear algebra after the matrices $`m_{x_i}`$ are precomputed using exact arithmetic. (These matrices are typically sparse and have additional structures which may be exploited.) Also, the coordinates $`\xi _i`$ are linear functions of the floating point entries of $`𝐯_\xi `$, which affords greater precision than the non-linear evaluations $`g_i(\xi _1)`$ in the method based upon elimination. While in principle only one of the $`\mathrm{deg}(I)`$ components of the vectors in (4) need be computed, averaging the results from all components can improve precision.
### 2.4. Real Roots
Determining the real roots of a polynomial system is a challenging problem with real world applications. When the polynomials come from geometry, this is the main problem of real enumerative geometry. Suppose $`k`$ and $`Ik[X]`$ is zero-dimensional. If $`g`$ is an eliminant of $`k[X]/I`$ with $`\mathrm{deg}(g)=\mathrm{deg}(I)`$, then the real roots of $`g`$ are in 1-1 correspondence with the real roots of $`I`$. Since there are effective methods for counting the real roots of a univariate polynomial, eliminants give a naïve, but useful method for determining the number of real roots to a polynomial system. (For some applications of this technique in mathematics, see .)
The classical symbolic method of Sturm, based upon Sturm sequences, counts the number of real roots of a univariate polynomial in an interval. When applied to an eliminant satisfying the Shape Lemma, this method counts the number of real roots of the ideal. This is implemented in Macaulay 2 via the command SturmSequence(f) of realroots.m2
i17 : code SturmSequence
o17 = -- code for SturmSequence:
-- realroots.m2:120-134
SturmSequence = f -\> (
assert( isPolynomialRing ring f );
assert( numgens ring f === 1 );
R := ring f;
assert( char R == 0 );
x := R\_0;
n := first degree f;
c := new MutableList from toList (0 .. n);
if n \>= 0 then (
c#0 = f;
if n \>= 1 then (
c#1 = diff(x,f);
scan(2 .. n, i -\> c#i = - c#(i-2) % c#(i-1));
));
toList c)
o17 : Net
The last few lines of SturmSequence construct the Sturm sequence of the univariate argument $`f`$: This is $`(f_0,f_1,f_2,\mathrm{})`$ where $`f_0=f`$, $`f_1=f^{}`$, and for $`i>1`$, $`f_i`$ is the normal form reduction of $`f_{i2}`$ modulo $`f_{i1}`$. Given any real number $`x`$, the variation of $`f`$ at $`x`$ is the number of changes in sign of the sequence $`(f_0(x),f_1(x),f_2(x),\mathrm{})`$ obtained by evaluating the Sturm sequence of $`f`$ at $`x`$. Then the number of real roots of $`f`$ over an interval $`[x,y]`$ is the difference of the variation of $`f`$ at $`x`$ and at $`y`$.
The Macaulay 2 commands numRealSturm and numPosRoots (and also numNegRoots) use this method to respectively compute the total number of real roots and the number of positive roots of a univariate polynomial.
i18 : code numRealSturm
o18 = -- code for numRealSturm:
-- realroots.m2:161-165
numRealSturm = f -\> (
c := SturmSequence f;
variations (signAtMinusInfinity \ c)
\- variations (signAtInfinity \ c)
)
o18 : Net
i19 : code numPosRoots
o19 = -- code for numPosRoots:
-- realroots.m2:170-174
numPosRoots = f -\> (
c := SturmSequence f;
variations (signAtZero \ c)
\- variations (signAtInfinity \ c)
)
o19 : Net
These use the commands signAt$``$(f), which give the sign of $`𝚏`$ at $``$. (Here, $``$ is one of Infinity, zero, or MinusInfinity. Also variations(c) computes the number of sign changes in the sequence c.
i20 : code variations
o20 = -- code for variations:
-- realroots.m2:187-195
variations = c -\> (
n := 0;
last := 0;
scan(c, x -\> if x =!= 0 then (
if last \< 0 and x \> 0 or last \> 0
and x \< 0 then n = n+1;
last = x;
));
n)
o20 : Net
A more sophisticated method to compute the number of real roots which can also give information about their location uses the rank and signature of the symmetric trace form. Suppose $`Ik[X]`$ is a zero-dimensional ideal and set $`A:=k[X]/I`$. For $`hk[X]`$, set $`S_h(f,g):=\mathrm{trace}(m_{hfg})`$. It is an easy exercise that $`S_h`$ is a symmetric bilinear form on $`A`$. The procedure traceForm(h) in realroots.m2 computes this trace form $`S_h`$.
i21 : code traceForm
o21 = -- code for traceForm:
-- realroots.m2:200-208
traceForm = h -\> (
assert( dim ring h == 0 );
b := basis ring h;
k := coefficientRing ring h;
mm := substitute(contract(transpose b, h * b \** b), k);
tr := matrix {apply(first entries b, x -\>
trace regularRep x)};
adjoint(tr * mm, source tr, source tr)
)
o21 : Net
The value of this construction is the following theorem.
###### Theorem 2.2 ().
Suppose $`k`$ and $`I`$ is a zero-dimensional ideal in $`k[x_1,\mathrm{},x_n]`$ and consider $`𝒱(I)^n`$. Then, for $`hk[x_1,\mathrm{},x_n]`$, the signature $`\sigma (S_h)`$ and rank $`\rho (S_h)`$ of the bilinear form $`S_h`$ satisfy
$`\sigma (S_h)`$ $`=`$ $`\mathrm{\#}\{a𝒱(I)^n:h(a)>0\}\mathrm{\#}\{a𝒱(I)^n:h(a)<0\}`$
$`\rho (S_h)`$ $`=`$ $`\mathrm{\#}\{a𝒱(I):h(a)0\}.`$
That is, the rank of $`S_h`$ counts roots in $`^n𝒱(h)`$, and its signature counts the real roots weighted by the sign of $`h`$ (which is $`1`$, $`0`$, or $`1`$) at each root. The command traceFormSignature(h) in realroots.m2 returns the rank and signature of the trace form $`S_h`$.
i22 : code traceFormSignature
o22 = -- code for traceFormSignature:
-- realroots.m2:213-224
traceFormSignature = h -\> (
A := ring h;
assert( dim A == 0 );
assert( char A == 0 );
S := QQ\[Z\];
TrF := traceForm(h) \** S;
IdZ := Z * id\_(S^(numgens source TrF));
f := det(TrF - IdZ);
\<\< "The trace form S\_h with h = " \<\< h \<\<
" has rank " \<\< rank(TrF) \<\< " and signature " \<\<
numPosRoots(f) - numNegRoots(f) \<\< endl
)
o22 : Net
The Macaulay 2 command numRealTrace(A) simply returns the number of real roots of $`I`$, given $`𝙰=k[X]/I`$.
i23 : code numRealTrace
o23 = -- code for numRealTrace:
-- realroots.m2:229-237
numRealTrace = A -\> (
assert( dim A == 0 );
assert( char A == 0 );
S := QQ\[Z\];
TrF := traceForm(1\_A) \** S;
IdZ := Z * id\_(S^(numgens source TrF));
f := det(TrF - IdZ);
numPosRoots(f)-numNegRoots(f)
)
o23 : Net
###### Example 2.3.
We illustrate these methods on the following polynomial system.
i24 : R = QQ\[x, y\];
i25 : I = ideal (1 - x^2\*y + 2\*x\*y^2, y - 2\*x - x\*y + x^2);
o25 : Ideal of R
The ideal $`I`$ has dimension zero and degree 5.
i26 : dim I, degree I
o26 = (0, 5)
o26 : Sequence
We compare the two methods to compute the eliminant of $`x`$ in the ring $`R/I`$.
i27 : A = R/I;
i28 : time g = eliminant(x, QQ\[Z\])
-- used 0.03 seconds
5 4 3 2
o28 = Z - 5Z + 6Z + Z - 2Z + 1
o28 : QQ \[Z\]
i29 : time g = charPoly(x, Z)
-- used 0.01 seconds
5 4 3 2
o29 = Z - 5Z + 6Z + Z - 2Z + 1
o29 : QQ \[Z\]
The eliminant has 3 real roots, which we test in two different ways.
i30 : numRealSturm(g), numRealTrace(A)
o30 = (3, 3)
o30 : Sequence
We use Theorem 2.2 to isolate these roots in the $`x,y`$-plane.
i31 : traceFormSignature(x\*y);
The trace form S\_h with h = x\*y has rank 5 and signature 3
Thus all 3 real roots lie in the first and third quadrants (where $`xy>0`$). We isolate these further.
i32 : traceFormSignature(x - 2);
The trace form S\_h with h = x - 2 has rank 5 and signature 1
This shows that two roots lie in the first quadrant with $`x>2`$ and one lies in the third. Finally, one of the roots lies in the triangle $`y>0`$, $`x>2`$, and $`x+y<3`$.
i33 : traceFormSignature(x + y - 3);
The trace form S\_h with h = x + y - 3 has rank 5 and signature -1
Figure 1 shows these three roots (dots), as well as the lines $`x+y=3`$ and $`x=2`$.
### 2.5. Homotopy methods
We describe symbolic-numeric homotopy continuation methods for finding approximate complex solutions to a system of equations. These exploit the traditional principles of conservation of number and specialization from enumerative geometry.
Suppose we seek the isolated solutions of a system $`F(X)=0`$ where $`F=(f_1,\mathrm{},f_n)`$ are polynomials in the variables $`X=(x_1,\mathrm{},x_N)`$. First, a homotopy $`H(X,t)`$ is found with the following properties:
1. $`H(X,1)=F(X)`$.
2. The isolated solutions of the start system $`H(X,0)=0`$ are known.
3. The system $`H(X,t)=0`$ defines finitely many (complex) curves, and each isolated solution of the original system $`F(X)=0`$ is connected to an isolated solution $`\sigma _i(0)`$ of $`H(X,0)=0`$ along one of these curves.
Next, choose a generic smooth path $`\gamma (t)`$ from 0 to 1 in the complex plane. Lifting $`\gamma `$ to the curves $`H(X,t)=0`$ gives smooth paths $`\sigma _i(t)`$ connecting each solution $`\sigma _i(0)`$ of the start system to a solution of the original system. The path $`\gamma `$ must avoid the finitely many points in $``$ over which the curves are singular or meet other components of the solution set $`H(X,t)=0`$.
Numerical path continuation is used to trace each path $`\sigma _i(t)`$ from $`t=0`$ to $`t=1`$. When there are fewer solutions to $`F(X)=0`$ than to $`H(X,0)=0`$, some paths will diverge or become singular as $`t1`$, and it is expensive to trace such a path. The homotopy is optimal when this does not occur.
When $`N=n`$ and the $`f_i`$ are generic, set $`G(X):=(g_1,\mathrm{},g_n)`$ with $`g_i=(x_i1)(x_i2)\mathrm{}(x_id_i)`$ where $`d_i:=\mathrm{deg}(f_i)`$. Then the Bézout homotopy
$$H(X,t):=tF(X)+(1t)G(X)$$
is optimal. This homotopy furnishes an effective demonstration of the bound in Bézout’s Theorem for the number of solutions to $`F(X)=0`$.
When the polynomial system is deficient, the Bézout homotopy is not optimal. When $`n>N`$ (often the case in geometric examples), the Bézout homotopy does not apply. In either case, a different strategy is needed. Present optimal homotopies for such systems all exploit some structure of the systems they are designed to solve. The current state-of-the-art is described in .
###### Example 2.4.
The Gröbner homotopy is an optimal homotopy that exploits a square-free initial ideal. Suppose our system has the form
$$F:=g_1(X),\mathrm{},g_m(X),\mathrm{\Lambda }_1(X),\mathrm{},\mathrm{\Lambda }_d(X)$$
where $`g_1(X),\mathrm{},g_m(X)`$ form a Gröbner basis for an ideal $`I`$ with respect to a given term order $``$, $`\mathrm{\Lambda }_1,\mathrm{},\mathrm{\Lambda }_d`$ are linear forms with $`d=dim(𝒱(I))`$, and we assume that the initial ideal $`\mathrm{in}_{}I`$ is square-free. This last, restrictive, hypothesis occurs for certain determinantal varieties.
As in \[9, Chapter 15\], there exist polynomials $`g_i(X,t)`$ interpolating between $`g_i(X)`$ and their initial terms $`\mathrm{in}_{}g_i(X)`$
$$g_i(X;1)=g_i(X)\text{and}g_i(X;0)=\mathrm{in}_{}g_i(X)$$
so that $`g_1(X,t),\mathrm{},g_m(X,t)`$ is a flat family with generic fibre isomorphic to $`I`$ and special fibre $`\mathrm{in}_{}I`$. The Gröbner homotopy is
$$H(X,t):=g_1(X,t),\mathrm{},g_m(X,t),\mathrm{\Lambda }_1(X),\mathrm{},\mathrm{\Lambda }_d(X).$$
Since $`\mathrm{in}_{}I`$ is square-free, $`𝒱(\mathrm{in}_{}I)`$ is a union of $`\mathrm{deg}(I)`$-many coordinate $`d`$-planes. We solve the start system by linear algebra. This conceptually simple homotopy is in general not efficient as it is typically overdetermined.
## 3. Some enumerative geometry
We use the tools we have developed to explore the enumerative geometric problems of cylinders meeting 5 general points and lines tangent to 4 spheres.
### 3.1. Cylinders meeting 5 points
A cylinder is the locus of points equidistant from a fixed line in $`^3`$. The Grassmannian of lines in 3-space is 4-dimensional, which implies that the space of cylinders is 5-dimensional, and so we expect that 5 points in $`^3`$ will determine finitely many cylinders. That is, there should be finitely many lines equidistant from 5 general points. The question is: How many cylinders/lines, and how many of them can be real?
Bottema and Veldkamp show there are 6 complex cylinders and Lichtblau observes that if the 5 points are the vertices of a bipyramid consisting of 2 regular tetrahedra sharing a common face, then all 6 will be real. We check this reality on a configuration with less symmetry (so the Shape Lemma holds).
If the axial line has direction $`𝐕`$ and contains the point $`𝐏`$ (and hence has parameterization $`𝐏+t𝐕`$), and if $`r`$ is the squared radius, then the cylinder is the set of points $`𝐗`$ satisfying
$$0=r𝐗𝐏\frac{𝐕(𝐗𝐏)}{𝐕^2}𝐕^2.$$
Expanding and clearing the denominator of $`𝐕^2`$ yields
(5)
$$0=r𝐕^2+[𝐕(𝐗𝐏)]^2𝐗𝐏^2𝐕^2.$$
We consider cylinders containing the following 5 points, which form an asymmetric bipyramid.
i34 : Points = {{2, 2, 0 }, {1, -2, 0}, {-3, 0, 0},
{0, 0, 5/2}, {0, 0, -3}};
Suppose that $`𝐏=(0,y_{11},y_{12})`$ and $`𝐕=(1,y_{21},y_{22})`$.
i35 : R = QQ\[r, y11, y12, y21, y22\];
i36 : P = matrix{{0, y11, y12}};
1 3
o36 : Matrix R \<--- R
i37 : V = matrix{{1, y21, y22}};
1 3
o37 : Matrix R \<--- R
We construct the ideal given by evaluating the polynomial (5) at each of the five points.
i38 : Points = matrix Points \** R;
5 3
o38 : Matrix R \<--- R
i39 : I = ideal apply(0..4, i -\> (
X := Points^{i};
r * (V * transpose V) +
((X - P) * transpose V)^2) -
((X - P) * transpose(X - P)) * (V * transpose V)
);
o39 : Ideal of R
This ideal has dimension 0 and degree 6.
i40 : dim I, degree I
o40 = (0, 6)
o40 : Sequence
There are 6 real roots, and they correspond to real cylinders (with $`r>0`$).
i41 : A = R/I; numPosRoots(charPoly(r, Z))
o42 = 3
### 3.2. Lines tangent to 4 spheres
We now ask for the lines having a fixed distance from 4 general points. Equivalently, these are the lines mutually tangent to 4 spheres. Since the Grassmannian of lines is four-dimensional, we expect there to be only finitely many such lines. Macdonald, Pach, and Theobald show that there are indeed 12 lines, and that all 12 may be real. This problem makes geometric sense over any field $`k`$ not of characteristic 2, and the derivation of the number 12 is also valid for algebraically closed fields not of characteristic 2.
A sphere in $`k^3`$ is $`𝒱(q(1,𝐱))`$, where $`q`$ is a quadratic form on $`k^4`$ and $`𝐱k^3`$. If our field does not have characteristic 2, then there is a symmetric $`4\times 4`$ matrix $`M`$ such that $`q(𝐮)=𝐮M𝐮^t`$.
A line $`\mathrm{}`$ having direction $`𝐕`$ and containing the point $`𝐏`$ is tangent to the sphere defined by $`q`$ when the univariate polynomial in $`s`$
$$q((1,𝐏)+s(0,𝐕))=q(1,𝐏)+2s(1,𝐏)M(0,𝐕)^t+s^2q(0,𝐕),$$
has a double root. Thus its discriminant vanishes, giving the equation
(6)
$$\left((1,𝐏)M(0,𝐕)^t\right)^2(1,𝐏)M(1,𝐏)^t(0,𝐕)M(0,𝐕)^t=0.$$
The matrix $`M`$ of the quadratic form $`q`$ of the sphere with center $`(a,b,c)`$ and squared radius $`r`$ is constructed by Sphere(a,b,c,r).
i43 : Sphere = (a, b, c, r) -\> (
matrix{{a^2 + b^2 + c^2 - r ,-a ,-b ,-c },
{ -a , 1 , 0 , 0 },
{ -b , 0 , 1 , 0 },
{ -c , 0 , 0 , 1 }}
);
If a line $`\mathrm{}`$ contains the point $`𝐏=(0,y_{11},y_{12})`$ and $`\mathrm{}`$ has direction $`𝐕=(1,y_{21},y_{22})`$, then tangentTo(M) is the equation for $`\mathrm{}`$ to be tangent to the quadric $`uMu^T=0`$ determined by the matrix $`M`$.
i44 : R = QQ\[y11, y12, y21, y22\];
i45 : tangentTo = (M) -\> (
P := matrix{{1, 0, y11, y12}};
V := matrix{{0, 1, y21, y22}};
(P * M * transpose V)^2 -
(P * M * transpose P) * (V * M * transpose V)
);
The ideal of lines having distance $`\sqrt{5}`$ from the four points $`(0,0,0)`$, $`(4,1,1)`$, $`(1,4,1)`$, and $`(1,1,4)`$ has dimension zero and degree 12.
i46 : I = ideal (tangentTo(Sphere(0,0,0,5)),
tangentTo(Sphere(4,1,1,5)),
tangentTo(Sphere(1,4,1,5)),
tangentTo(Sphere(1,1,4,5)));
o46 : Ideal of R
i47 : dim I, degree I
o47 = (0, 12)
o47 : Sequence
Thus there are 12 lines whose distance from those 4 points is $`\sqrt{5}`$. We check that all 12 are real.
i48 : A = R/I;
i49 : numRealSturm(eliminant(y11 - y12 + y21 + y22, QQ\[Z\]))
o49 = 12
Since no eliminant given by a coordinate function satisfies the hypotheses of the Shape Lemma, we took the eliminant with respect to the linear form $`y_{11}y_{12}+y_{21}+y_{22}`$.
This example is an instance of Lemma 3 of . These four points define a regular tetrahedron with volume $`V=9`$ where each face has area $`A=\sqrt{3^5}/2`$ and each edge has length $`e=\sqrt{18}`$. That result guarantees that all 12 lines will be real when $`e/2<r<A^2/3V`$, which is the case above.
## 4. Schubert calculus
The classical Schubert calculus of enumerative geometry concerns linear subspaces having specified positions with respect to other, fixed subspaces. For instance, how many lines in $`^3`$ meet four given lines? (See Example 4.2.) More generally, let $`1<r<n`$ and suppose that we are given general linear subspaces $`L_1,\mathrm{},L_m`$ of $`k^n`$ with $`dimL_i=nr+1l_i`$. When $`l_1+\mathrm{}+l_m=r(nr)`$, there will be a finite number $`d(r,n;l_1,\mathrm{},l_m)`$ of $`r`$-planes in $`k^n`$ which meet each $`L_i`$ non-trivially. This number may be computed using classical algorithms of Schubert and Pieri (see ).
The condition on $`r`$-planes to meet a fixed $`(nr+1l)`$-plane non-trivially is called a (special) Schubert condition, and we call the data $`(r,n;l_1,\mathrm{},l_m)`$ (special) Schubert data. The (special) Schubert calculus concerns this class of enumerative problems. We give two polynomial formulations of this special Schubert calculus, consider their solutions over $``$, and end with a question for fields of arbitrary characteristic.
### 4.1. Equations for the Grassmannian
The ambient space for the Schubert calculus is the Grassmannian of $`r`$-planes in $`k^n`$, denoted $`𝐆_{r,n}`$. For $`H𝐆_{r,n}`$, the $`r`$th exterior product of the embedding $`Hk^n`$ gives a line
$$k^rH^rk^nk^{\left(\genfrac{}{}{0pt}{}{n}{r}\right)}.$$
This induces the Plücker embedding $`𝐆_{r,n}^{\left(\genfrac{}{}{0pt}{}{n}{r}\right)1}`$. If $`H`$ is the row space of an $`r`$ by $`n`$ matrix, also written $`H`$, then the Plücker embedding sends $`H`$ to its vector of $`\left(\genfrac{}{}{0pt}{}{n}{r}\right)`$ maximal minors. Thus the $`r`$-subsets of $`\{0,\mathrm{},n1\}`$, $`𝕐_{r,n}:=\mathrm{𝚜𝚞𝚋𝚜𝚎𝚝𝚜}(𝚗,𝚛)`$, index Plücker coordinates of $`𝐆_{r,n}`$. The Plücker ideal of $`𝐆_{r,n}`$ is therefore the ideal of algebraic relations among the maximal minors of a generic $`r`$ by $`n`$ matrix.
We create the coordinate ring $`k[p_\alpha \alpha 𝕐_{2,5}]`$ of $`^9`$ and the Plücker ideal of $`𝐆_{2,5}`$. The Grassmannian $`𝐆_{r,n}`$ of $`r`$-dimensional subspaces of $`k^n`$ is also the Grassmannian of $`r1`$-dimensional affine subspaces of $`^{n1}`$. Macaulay 2 uses this alternative indexing scheme.
i50 : R = ZZ/101\[apply(subsets(5,2), i -\> p\_i )\];
i51 : I = Grassmannian(1, 4, R)
o51 = ideal (p p - p p + p p , p $`\mathrm{}`$
{2, 3} {1, 4} {1, 3} {2, 4} {1, 2} {3, 4} {2, 3} $`\mathrm{}`$
o51 : Ideal of R
This projective variety has dimension 6 and degree 5
i52 : dim(Proj(R/I)), degree(I)
o52 = (6, 5)
o52 : Sequence
This ideal has an important combinatorial structure \[26, Example 11.9\]. We write each $`\alpha 𝕐_{r,n}`$ as an increasing sequence $`\alpha :\alpha _1<\mathrm{}<\alpha _r`$. Given $`\alpha ,\beta 𝕐_{r,n}`$, consider the two-rowed array with $`\alpha `$ written above $`\beta `$. We say $`\alpha \beta `$ if each column weakly increases. If we sort the columns of an array with rows $`\alpha `$ and $`\beta `$, then the first row is the meet $`\alpha \beta `$ (greatest lower bound) and the second row the join $`\alpha \beta `$ (least upper bound) of $`\alpha `$ and $`\beta `$. These definitions endow $`𝕐_{r,n}`$ with the structure of a distributive lattice.Figure 2 shows $`𝕐_{2,5}`$.
We give $`k[p_\alpha ]`$ the degree reverse lexicographic order, where we first order the variables $`p_\alpha `$ by lexicographic order on their indices $`\alpha `$.
###### Theorem 4.1.
The reduced Gröbner basis of the Plücker ideal with respect to this degree reverse lexicographic term order consists of quadratic polynomials
$$g(\alpha ,\beta )=p_\alpha p_\beta p_{\alpha \beta }p_{\alpha \beta }+\text{lower terms in },$$
for each incomparable pair $`\alpha ,\beta `$ in $`𝕐_{r,n}`$, and all lower terms $`\lambda p_\gamma p_\delta `$ in $`g(\alpha ,\beta )`$ satisfy $`\gamma \alpha \beta `$ and $`\alpha \beta \delta `$.
The form of this Gröbner basis implies that the standard monomials are the sortable monomials, those $`p_\alpha p_\beta \mathrm{}p_\gamma `$ with $`\alpha \beta \mathrm{}\gamma `$. Thus the Hilbert function of $`𝐆_{r,n}`$ may be expressed in terms of the combinatorics of $`𝕐_{r,n}`$. For instance, the dimension of $`𝐆_{r,n}`$ is the rank of $`𝕐_{r,n}`$, and its degree is the number of maximal chains. From Figure 2, these are 6 and 5 for $`𝕐_{2,5}`$, confirming our previous calculations.
Since the generators $`g(\alpha ,\beta )`$ are linearly independent, this Gröbner basis is also a minimal generating set for the ideal. The displayed generator in o51,
$$p_{\{2,3\}}p_{\{1,4\}}p_{\{1,3\}}p_{\{2,4\}}p_{\{1,2\}}p_{\{3,4\}},$$
is $`g(23,14)`$, and corresponds to the underlined incomparable pair in Figure 2. Since there are 5 such incomparable pairs, the Gröbner basis has 5 generators. As $`𝐆_{2,5}`$ has codimension 3, it is not a complete intersection. This shows how the general enumerative problem from the Schubert calculus gives rise to an overdetermined system of equations in this global formulation.
The Grassmannian has a useful system of local coordinates given by $`\mathrm{Mat}_{r,nr}`$ as follows
(7)
$$Y\mathrm{Mat}_{r,nr}\mathrm{rowspace}[I_r:Y]𝐆_{r,n}.$$
Let $`L`$ be a ($`nr+1l`$)-plane in $`k^n`$ which is the row space of a $`nr+1l`$ by $`n`$ matrix, also written $`L`$. Then $`L`$ meets $`X𝐆_{r,n}`$ non-trivially if
$$\text{maximal minors of }\left[\begin{array}{c}L\\ X\end{array}\right]=0.$$
Laplace expansion of each minor along the rows of $`X`$ gives a linear equation in the Plücker coordinates. In the local coordinates (substituting $`[I_r:Y]`$ for $`X`$), we obtain multilinear equations of degree $`\mathrm{min}\{r,nr\}`$. These equations generate a prime ideal of codimension $`l`$.
Suppose each $`l_i=1`$ in our enumerative problem. Then in the Plücker coordinates, we have the Plücker ideal of $`𝐆_{r,n}`$ together with $`r(nr)`$ linear equations, one for each $`(nr)`$-plane $`L_i`$. By Theorem 4.1, the Plücker ideal has a square-free initial ideal, and so the Gröbner homotopy of Example 2.4 may be used to solve this enumerative problem.
###### Example 4.2.
$`𝐆_{2,4}^5`$ has equation
(8)
$$p_{\{1,2\}}p_{\{0,3\}}p_{\{1,3\}}p_{\{0,2\}}+p_{\{2,3\}}p_{\{0,1\}}=0.$$
The condition for $`H𝐆_{2,4}`$ to meet a 2-plane $`L`$ is the vanishing of
(9)
$$p_{\{1,2\}}L_{34}p_{\{1,3\}}L_{24}+p_{\{2,3\}}L_{14}+p_{\{1,4\}}L_{23}p_{\{2,4\}}L_{13}+p_{\{3,4\}}L_{12},$$
where $`L_{ij}`$ is the $`(i,j)`$th maximal minor of $`L`$.
If $`l_1=\mathrm{}=l_4=1`$, we have 5 equations in $`^5`$, one quadratic and 4 linear, and so by Bézout’s Theorem there are two 2-planes in $`k^4`$ that meet 4 general 2-planes non-trivially. This means that there are 2 lines in $`^3`$ meeting 4 general lines. In local coordinates, (9) becomes
$$L_{34}L_{14}y_{11}+L_{13}y_{12}L_{24}y_{21}+L_{23}y_{22}+L_{12}(y_{11}y_{22}y_{12}y_{21}).$$
This polynomial has the form of the last specialization in Example 1.2.
### 4.2. Reality in the Schubert calculus
Like the other enumerative problems we have discussed, enumerative problems in the special Schubert calculus are fully real in that all solutions can be real . That is, given any Schubert data $`(r,n;l_1,\mathrm{},l_m)`$, there exist subspaces $`L_1,\mathrm{},L_m^n`$ such that each of the $`d(r,n;l_1,\mathrm{},l_m)`$ $`r`$-planes that meet each $`L_i`$ are themselves real.
This result gives some idea of which choices of the $`L_i`$ give all $`r`$-planes real. Let $`\gamma `$ be a fixed rational normal curve in $`^n`$. Then the $`L_i`$ are linear subspaces osculating $`\gamma `$. More concretely, suppose that $`\gamma `$ is the standard rational normal curve, $`\gamma (s)=(1,s,s^2,\mathrm{},s^{n1})`$. Then the $`i`$-plane $`L_i(s):=\gamma (s),\gamma ^{}(s),\mathrm{},\gamma ^{(i1)}(s)`$ osculating $`\gamma `$ at $`\gamma (s)`$ is the row space of the matrix given by oscPlane(i, n, s).
i53 : oscPlane = (i, n, s) -\> (
gamma := matrix {toList apply(1..n, i -\> s^(i-1))};
L := gamma;
j := 0;
while j \< i-1 do (gamma = diff(s, gamma);
L = L || gamma;
j = j+1);
L);
i54 : QQ\[s\]; oscPlane(3, 6, s)
o55 = | 1 s s2 s3 s4 s5 |
| 0 1 2s 3s2 4s3 5s4 |
| 0 0 2 6s 12s2 20s3 |
3 6
o55 : Matrix QQ \[s\] \<--- QQ \[s\]
(In o55, the exponents of $`s`$ are displayed in line: $`s^2`$ is written s2. Macaulay 2 uses this notational convention to display matrices efficiently.)
###### Theorem 4.3 ().
For any Schubert data $`(r,n;l_1,\mathrm{},l_m)`$, there exist real numbers $`s_1,s_2,\mathrm{},s_m`$ such that there are $`d(r,n;l_1,\mathrm{},l_m)`$ $`r`$-planes that meet each osculating plane $`L_i(s_i)`$, and all are real.
The inspiration for looking at subspaces osculating the rational normal curve to study real enumerative geometry for the Schubert calculus is the following very interesting conjecture of Boris Shapiro and Michael Shapiro, or more accurately, extensive computer experimentation based upon their conjecture .
Shapiros’s Conjecture. For any Schubert data $`(r,n;l_1,\mathrm{},l_m)`$ and for all real numbers $`s_1,s_2,\mathrm{},s_m`$ there are $`d(r,n;l_1,\mathrm{},l_m)`$ $`r`$-planes that meet each osculating plane $`L_i(s_i)`$, and all are real.
In addition to Theorem 4.3, (which replaces the quantifier for all by there exist), the strongest evidence for this Conjecture is the following result of Eremenko and Gabrielov .
###### Theorem 4.4.
Shapiros’s Conjecture is true when either $`r`$ or $`nr`$ is $`2`$.
We test an example of this conjecture for the Schubert data $`(3,6;1^3,2^3)`$, (where $`a^b`$ is $`a`$ repeated $`b`$ times). The algorithms of the Schubert calculus predict that $`d(3,6;1^3,2^3)=6`$. The function spSchub(r, L, P) computes the ideal of $`r`$-planes meeting the row space of $`L`$ in the Plücker coordinates $`P_\alpha `$.
i56 : spSchub = (r, L, P) -\> (
I := ideal apply(subsets(numgens source L,
r + numgens target L), S -\>
fold((sum, U) -\> sum +
fold((term,i) -\> term\*(-1)^i, P\_(S\_U) * det(
submatrix(L, sort toList(set(S) - set(S\_U)))), U),
0, subsets(#S, r))));
We are working in the Grassmannian of 3-planes in $`^6`$.
i57 : R = QQ\[apply(subsets(6,3), i -\> p\_i )\];
The ideal $`I`$ consists of the special Schubert conditions for the 3-planes to meet the 3-planes osculating the rational normal curve at the points 1, 2, and 3, and to also meet the 2-planes osculating at 4, 5, and 6, together with the Plücker ideal Grassmannian(2, 5, R). Since this is a 1-dimensional homogeneous ideal, we add the linear form p\_{0,1,5} - 1 to make the ideal zero-dimensional. As before, Grassmannian(2, 5, R) creates the Plücker ideal of $`𝐆_{3,6}`$.
i58 : I = fold((J, i) -\> J +
spSchub(3, substitute(oscPlane(3, 6, s), {s=\> 1+i}), p) +
spSchub(3, substitute(oscPlane(2, 6, s), {s=\> 4+i}), p),
Grassmannian(2, 5, R), {0,1,2}) +
ideal (p\_{0,1,5} - 1);
o58 : Ideal of R
This has dimension 0 and degree 6, in agreement with the Schubert calculus.
i59 : dim I, degree I
o59 = (0, 6)
o59 : Sequence
As expected, all roots are real.
i60 : A = R/I; numRealSturm(eliminant(p\_{2,3,4}, QQ\[Z\]))
o61 = 6
There have been many checked instances of this conjecture , and it has some geometrically interesting generalizations .
The question remains for which numbers $`0dd(r,n;l_1,\mathrm{},l_m)`$ do there exist real planes $`L_i`$ with $`d(r,n;l_1,\mathrm{},l_m)`$ $`r`$-planes meeting each $`L_i`$, and exactly $`d`$ of them are real. Besides Theorem 4.3 and the obvious parity condition, nothing is known in general. In every known case, every possibility occurs—which is not the case in all enumerative problems, even those that are fully real<sup>1</sup><sup>1</sup>1For example, of the 12 rational plane cubics containing 8 real points in $`^2`$, either 8, 10 or 12 can be real, and there are 8 points with all 12 real \[8, Proposition 4.7.3\].. Settling this (for $`d=0`$) has implications for linear systems theory .
### 4.3. Transversality in the Schubert calculus
A basic principle of the classical Schubert calculus is that the intersection number $`d(r,n;l_1,\mathrm{},l_m)`$ has enumerative significance—that is, for general linear subspaces $`L_i`$, all solutions appear with multiplicity 1. This basic principle is not known to hold in general. For fields of characteristic zero, Kleiman’s Transversality Theorem establishes this principle. When $`r`$ or $`nr`$ is 2, then Theorem E of establishes this principle in arbitrary characteristic. We conjecture that this principle holds in general; that is, for arbitrary infinite fields and any Schubert data, if the planes $`L_i`$ are in general position, then the resulting zero-dimensional ideal is radical.
We test this conjecture on the enumerative problem of Section 4.2, which is not covered by Theorem E of . The function testTransverse(F) tests transversality for this enumerative problem, for a given field $`F`$. It does this by first computing the ideal of the enumerative problem using random planes $`L_i`$.
i62 : randL = (R, n, r, l) -\>
matrix table(n-r+1-l, n, (i, j) -\> random(0, R));
and the Plücker ideal of the Grassmannian $`𝐆_{3,6}`$ Grassmannian(2, 5, R).) Then it adds a random (inhomogeneous) linear relation 1 + random(1, R) to make the ideal zero-dimensional for generic $`L_i`$. When this ideal is zero dimensional and has degree 6 (the expected degree), it computes the characteristic polynomial g of a generic linear form. If g has no multiple roots, 1 == gcd(g, diff(Z, g)), then the Shape Lemma guarantees that the ideal was radical. testTransverse exits either when it computes a radical ideal, or after limit iterations (which is set to 5 for these examples), and prints the return status.
i63 : testTransverse = F -\> (
R := F\[apply(subsets(6, 3), i -\> q\_i )\];
continue := true;
j := 0;
limit := 5;
while continue and (j \< limit) do (
j = j + 1;
I := fold((J, i) -\> J +
spSchub(3, randL(R, 6, 3, 1), q) +
spSchub(3, randL(R, 6, 3, 2), q),
Grassmannian(2, 5, R) +
ideal (1 + random(1, R)),
{0, 1, 2});
if (dim I == 0) and (degree I == 6) then (
lin := promote(random(1, R), (R/I));
g := charPoly(lin, Z);
continue = not(1 == gcd(g, diff(Z, g)));
));
if continue then \<\< "Failed for the prime " \<\< char F \<\<
" with " \<\< j \<\< " iterations" \<\< endl;
if not continue then \<\< "Succeeded for the prime " \<\<
char F \<\< " in " \<\< j \<\< " iteration(s)" \<\< endl;
);
Since 5 iterations do not show transversality for $`𝔽_2`$,
i64 : testTransverse(ZZ/2);
Failed for the prime 2 with 5 iterations
we can test transversality in characteristic 2 using the field with four elements, $`𝔽_4=`$ GF 4.
i65 : testTransverse(GF 4);
Succeeded for the prime 2 in 3 iteration(s)
We do find transversality for $`𝔽_7`$.
i66 : testTransverse(ZZ/7);
Succeeded for the prime 7 in 2 iteration(s)
We have tested transversality for all primes less than 100 in every enumerative problem involving Schubert conditions on 3-planes in $`k^6`$. These include the problem above as well as the problem of 42 3-planes meeting 9 general 3-planes.<sup>2</sup><sup>2</sup>2After this was written, we discovered an elementary proof of transversality for the enumerative problems $`d(r,n;1^{r(nr)})`$, where the conditions are all codimension 1 .
## 5. The 12 lines: reprise
The enumerative problems of Section 3 were formulated in local coordinates (7) for the Grassmannian of lines in $`^3`$ (Grassmannian of 2-dimensional subspaces in $`k^4`$). When we formulate the problem of Section 3.2 in the global Plücker coordinates of Section 4.1, we find some interesting phenomena. We also consider some related enumerative problems.
### 5.1. Global formulation
A quadratic form $`q`$ on a vector space $`V`$ over a field $`k`$ not of characteristic 2 is given by $`q(𝐮)=(\phi (𝐮),𝐮)`$, where $`\phi :VV^{}`$ is a symmetric linear map, that is $`(\phi (𝐮),𝐯)=(\phi (𝐯),𝐮)`$. Here, $`V^{}`$ is the linear dual of $`V`$ and $`(,)`$ is the pairing $`VV^{}k`$. The map $`\phi `$ induces a quadratic form $`^rq`$ on the $`r`$th exterior power $`^rV`$ of $`V`$ through the symmetric map $`^r\phi :^rV^rV^{}=(^rV)^{}`$. The action of $`^rV^{}`$ on $`^rV`$ is given by
(10)
$$(𝐱_1𝐱_2\mathrm{}𝐱_r,𝐲_1𝐲_2\mathrm{}𝐲_r)=det|(𝐱_i,𝐲_j)|,$$
where $`𝐱_iV^{}`$ and $`𝐲_jV`$.
When we fix isomorphisms $`Vk^nV^{}`$, the map $`\phi `$ is given by a symmetric $`n\times n`$ matrix $`M`$ as in Section 3.2. Suppose $`r=2`$. Then for $`𝐮,𝐯k^n`$,
$$^2q(𝐮𝐯)=det\left[\begin{array}{cc}𝐮M𝐮^t& 𝐮M𝐯^t\\ 𝐯M𝐮^t& 𝐯M𝐯^t\end{array}\right],$$
which is Equation (6) of Section 3.2.
###### Proposition 5.1.
A line $`\mathrm{}`$ is tangent to a quadric $`𝒱(q)`$ in $`^{n1}`$ if and only if its Plücker coordinate $`^2\mathrm{}^{\left(\genfrac{}{}{0pt}{}{n}{2}\right)1}`$ lies on the quadric $`𝒱(^2q)`$.
Thus the Plücker coordinates for the set of lines tangent to 4 general quadrics in $`^3`$ satisfy 5 quadratic equations: The single Plücker relation (8) together with one quadratic equation for each quadric. Thus we expect the Bézout number of $`2^5=32`$ such lines. We check this.
The procedure randomSymmetricMatrix(R, n) generates a random symmetric $`n\times n`$ matrix with entries in the base ring of $`R`$.
i67 : randomSymmetricMatrix = (R, n) -\> (
entries := new MutableHashTable;
scan(0..n-1, i -\> scan(i..n-1, j -\>
entries#(i, j) = random(0, R)));
matrix table(n, n, (i, j) -\> if i \> j then
entries#(j, i) else entries#(i, j))
);
The procedure tangentEquation(r, R, M) gives the equation in Plücker coordinates for a point in $`^{\left(\genfrac{}{}{0pt}{}{n}{r}\right)1}`$ to be isotropic with respect to the bilinear form $`^rM`$ (R is assumed to be the coordinate ring of $`^{\left(\genfrac{}{}{0pt}{}{n}{r}\right)1}`$). This is the equation for an $`r`$-plane to be tangent to the quadric associated to $`M`$.
i68 : tangentEquation = (r, R, M) -\> (
g := matrix {gens(R)};
(entries(g * exteriorPower(r, M) * transpose g))\_0\_0
);
We construct the ideal of lines tangent to 4 general quadrics in $`^3`$.
i69 : R = QQ\[apply(subsets(4, 2), i -\> p\_i )\];
i70 : I = Grassmannian(1, 3, R) + ideal apply(0..3, i -\>
tangentEquation(2, R, randomSymmetricMatrix(R, 4)));
o70 : Ideal of R
As expected, this ideal has dimension 0 and degree 32.
i71 : dim Proj(R/I), degree I
o71 = (0, 32)
o71 : Sequence
### 5.2. Lines tangent to 4 spheres
That calculation raises the following question: In Section 3.2, why did we obtain only 12 lines tangent to 4 spheres? To investigate this, we generate the global ideal of lines tangent to the spheres of Section 3.2.
i72 : I = Grassmannian(1, 3, R) +
ideal (tangentEquation(2, R, Sphere(0,0,0,5)),
tangentEquation(2, R, Sphere(4,1,1,5)),
tangentEquation(2, R, Sphere(1,4,1,5)),
tangentEquation(2, R, Sphere(1,1,4,5)));
o72 : Ideal of R
We compute the dimension and degree of $`𝒱(I)`$.
i73 : dim Proj(R/I), degree I
o73 = (1, 4)
o73 : Sequence
The ideal is not zero dimensional; there is an extraneous one-dimensional component of zeroes with degree 4. Since we found 12 lines in Section 3.2 using the local coordinates (7), the extraneous component must lie in the complement of that coordinate patch, which is defined by the vanishing of the first Plücker coordinate, $`p_{\{0,1\}}`$. We saturate $`I`$ by $`p_{\{0,1\}}`$ to obtain the desired lines.
i74 : Lines = saturate(I, ideal (p\_{0,1}));
o74 : Ideal of R
This ideal does have dimension 0 and degree 12, so we have recovered the zeroes of Section 3.2.
i75 : dim Proj(R/Lines), degree(Lines)
o75 = (0, 12)
o75 : Sequence
We investigate the rest of the zeroes, which we obtain by taking the ideal quotient of $`I`$ and the ideal of lines. As computed above, this has dimension 1 and degree 4.
i76 : Junk = I : Lines;
o76 : Ideal of R
i77 : dim Proj(R/Junk), degree Junk
o77 = (1, 4)
o77 : Sequence
We find the support of this extraneous component by taking its radical.
i78 : radical(Junk)
2 2 2
o78 = ideal (p , p , p , p + p + p )
{0, 3} {0, 2} {0, 1} {1, 2} {1, 3} {2, 3}
o78 : Ideal of R
From this, we see that the extraneous component is supported on an imaginary conic in the $`^2`$ of lines at infinity.
To understand the geometry behind this computation, observe that the sphere with radius $`r`$ and center $`(a,b,c)`$ has homogeneous equation
$$(xwa)^2+(ywb)^2+(zwc)^2=r^2w^2.$$
At infinity, $`w=0`$, this has equation
$$x^2+y^2+z^2=0.$$
The extraneous component is supported on the set of tangent lines to this imaginary conic. Aluffi and Fulton studied this problem, using geometry to identify the extraneous ideal and the excess intersection formula to obtain the answer of 12. Their techniques show that there will be 12 isolated lines tangent to 4 quadrics which have a smooth conic in common.
When the quadrics are spheres, the conic is the imaginary conic at infinity. Fulton asked the following question: Can all 12 lines be real if the (real) four quadrics share a real conic? We answer his question in the affirmative in the next section.
### 5.3. Lines tangent to real quadrics sharing a real conic
We consider four quadrics in $`_{}^3`$ sharing a non-singular conic, which we will take to be at infinity so that we may use local coordinates for $`𝐆_{2,4}`$ in our computations. The variety $`𝒱(q)_{}^3`$ of a nondegenerate quadratic form $`q`$ is determined up to isomorphism by the absolute value of the signature $`\sigma `$ of the associated bilinear form. Thus there are three possibilities, 0, 2, or 4, for $`|\sigma |`$.
When $`|\sigma |=4`$, the real quadric $`𝒱(q)`$ is empty. The associated symmetric matrix $`M`$ is conjugate to the identity matrix, so $`^2M`$ is also conjugate to the identity matrix. Hence $`𝒱(^2q)`$ contains no real points. Thus we need not consider quadrics with $`|\sigma |=4`$.
When $`|\sigma |=2`$, we have $`𝒱(q)S^2`$, the 2-sphere. If the conic at infinity is imaginary, then $`𝒱(q)^3`$ is an ellipsoid. If the conic at infinity is real, then $`𝒱(q)^3`$ is a hyperboloid of two sheets. When $`\sigma =0`$, we have $`𝒱(q)S^1\times S^1`$, a torus. In this case, $`𝒱(q)^3`$ is a hyperboloid of one sheet and the conic at infinity is real.
Thus either we have 4 ellipsoids sharing an imaginary conic at infinity, which we studied in Section 3.2; or else we have four hyperboloids sharing a real conic at infinity, and there are five possible combinations of hyperboloids of one or two sheets sharing a real conic at infinity. This gives six topologically distinct possibilities in all.
###### Theorem 5.2.
For each of the six topologically distinct possibilities of four real quadrics sharing a smooth conic at infinity, there exist four quadrics having the property that each of the 12 lines in $`^3`$ simultaneously tangent to the four quadrics is real.
###### Proof.
By the computation in Section 3.2, we need only check the five possibilities for hyperboloids. We fix the conic at infinity to be $`x^2+y^2z^2=0`$. Then the general hyperboloid of two sheets containing this conic has equation in $`^3`$
(11)
$$(xa)^2+(yb)^2(zc)^2+r=0,$$
(with $`r>0`$). The command Two(a,b,c,r) generates the associated symmetric matrix.
i79 : Two = (a, b, c, r) -\> (
matrix{{a^2 + b^2 - c^2 + r ,-a ,-b , c },
{ -a , 1 , 0 , 0 },
{ -b , 0 , 1 , 0 },
{ c , 0 , 0 ,-1 }}
);
The general hyperboloid of one sheet containing the conic $`x^2+y^2z^2=0`$ at infinity has equation in $`^3`$
(12)
$$(xa)^2+(yb)^2(zc)^2r=0,$$
(with $`r>0`$). The command One(a,b,c,r) generates the associated symmetric matrix.
i80 : One = (a, b, c, r) -\> (
matrix{{a^2 + b^2 - c^2 - r ,-a ,-b , c },
{ -a , 1 , 0 , 0 },
{ -b , 0 , 1 , 0 },
{ c , 0 , 0 ,-1 }}
);
We consider $`i`$ quadrics of two sheets (11) and $`4i`$ quadrics of one sheet (12). For each of these cases, the table below displays four 4-tuples of data $`(a,b,c,r)`$ which give 12 real lines. (The data for the hyperboloids of one sheet are listed first.)
| $`i`$ | Data |
| --- | --- |
| $`0`$ | $`(5,3,3,16),(5,4,2,1),(3,1,1,1),(2,7,0,1)`$ |
| $`1`$ | $`(3,2,3,6),(3,7,6,7),(6,3,5,2),(1,6,2,5)`$ |
| $`2`$ | $`(6,4,6,4),(1,3,3,6),(7,2,3,3),(6,7,2,5)`$ |
| $`3`$ | $`(1,4,1,1),(3,3,1,1),(7,6,2,9),(5,6,1,12)`$ |
| $`4`$ | $`(5,2,1,25),(6,6,2,25),(7,1,6,1),(3,1,0,1)`$ |
We test each of these, using the formulation in local coordinates of Section 3.2.
i81 : R = QQ\[y11, y12, y21, y22\];
i82 : I = ideal (tangentTo(One( 5, 3, 3,16)),
tangentTo(One( 5,-4, 2, 1)),
tangentTo(One(-3,-1, 1, 1)),
tangentTo(One( 2,-7, 0, 1)));
o82 : Ideal of R
i83 : numRealSturm(charPoly(promote(y22, R/I), Z))
o83 = 12
i84 : I = ideal (tangentTo(One( 3,-2,-3, 6)),
tangentTo(One(-3,-7,-6, 7)),
tangentTo(One(-6, 3,-5, 2)),
tangentTo(Two( 1, 6,-2, 5)));
o84 : Ideal of R
i85 : numRealSturm(charPoly(promote(y22, R/I), Z))
o85 = 12
i86 : I = ideal (tangentTo(One( 6, 4, 6, 4)),
tangentTo(One(-1, 3, 3, 6)),
tangentTo(Two(-7,-2, 3, 3)),
tangentTo(Two(-6, 7,-2, 5)));
o86 : Ideal of R
i87 : numRealSturm(charPoly(promote(y22, R/I), Z))
o87 = 12
i88 : I = ideal (tangentTo(One(-1,-4,-1, 1)),
tangentTo(Two(-3, 3,-1, 1)),
tangentTo(Two(-7, 6, 2, 9)),
tangentTo(Two( 5, 6,-1,12)));
o88 : Ideal of R
i89 : numRealSturm(charPoly(promote(y22, R/I), Z))
o89 = 12
i90 : I = ideal (tangentTo(Two( 5, 2,-1,25)),
tangentTo(Two( 6,-6, 2,25)),
tangentTo(Two(-7, 1, 6, 1)),
tangentTo(Two( 3, 1, 0, 1)));
o90 : Ideal of R
i91 : numRealSturm(charPoly(promote(y22, R/I), Z))
o91 = 12
In each of these enumerative problems, we have checked that every possible number of real solutions (0, 2, 4, 6, 8, 10, or 12) can occur.
### 5.4. Generalization to higher dimensions
We consider lines tangent to quadrics in higher dimensions. First, we reinterpret the action of $`^rV^{}`$ on $`^rV`$ described in (10) as follows. The vectors $`𝐱_1,\mathrm{},𝐱_r`$ and $`𝐲_1,\mathrm{},𝐲_r`$ define maps $`\alpha :k^rV^{}`$ and $`\beta :k^rV`$. The matrix $`[(𝐱_i,𝐲_j)]`$ is the matrix of the bilinear form on $`k^r`$ given by $`𝐮,𝐯:=(\alpha (𝐮),\beta (𝐯))`$. Thus (10) vanishes when the bilinear form $`,`$ on $`k^r`$ is degenerate.
Now suppose that we have a quadratic form $`q`$ on $`V`$ given by a symmetric map $`\phi :VV^{}`$. This induces a quadratic form and hence a quadric on any $`r`$-plane $`H`$ in $`V`$ (with $`H𝒱(q)`$). This induced quadric is singular when $`H`$ is tangent to $`𝒱(q)`$. Since a quadratic form is degenerate only when the associated projective quadric is singular, we see that $`H`$ is tangent to the quadric $`𝒱(q)`$ if and only if $`(^r\phi (^rH),^rH)=0`$. (This includes the case $`H𝒱(q)`$.) We summarize this argument.
###### Theorem 5.3.
Let $`\phi :VV^{}`$ be a linear map with resulting bilinear form $`(\phi (𝐮),𝐯)`$. Then the locus of $`r`$-planes in $`V`$ for which the restriction of this form is degenerate is the set of $`r`$-planes $`H`$ whose Plücker coordinates are isotropic, $`(^r\phi (^rH),^rH)=0`$, with respect to the induced form on $`^rV`$.
When $`\phi `$ is symmetric, this is the locus of $`r`$-planes tangent to the associated quadric in $`(V)`$.
We explore the problem of lines tangent to quadrics in $`^n`$. From the calculations of Section 5.1, we do not expect this to be interesting if the quadrics are general. (This is borne out for $`^4`$: we find 320 lines in $`^4`$ tangent to 6 general quadrics. This is the Bézout number, as $`\mathrm{deg}𝐆_{2,5}=5`$ and the condition to be tangent to a quadric has degree 2.) This problem is interesting if the quadrics in $`^n`$ share a quadric in a $`^{n1}`$. We propose studying such enumerative problems, both determining the number of solutions for general such quadrics, and investigating whether or not it is possible to have all solutions be real.
We use Macaulay 2 to compute the expected number of solutions to this problem when $`r=2`$ and $`n=4`$. We first define some functions for this computation, which will involve counting the degree of the ideal of lines in $`^4`$ tangent to 6 general spheres. Here, $`X`$ gives local coordinates for the Grassmannian, $`M`$ is a symmetric matrix, tanQuad gives the equation in $`X`$ for the lines tangent to the quadric given by $`M`$.
i92 : tanQuad = (M, X) -\> (
u := X^{0};
v := X^{1};
(u * M * transpose v)^2 -
(u * M * transpose u) * (v * M * transpose v)
);
nSphere gives the matrix $`M`$ for a sphere with center V and squared radius r, and V and r give random data for a sphere.
i93 : nSphere = (V, r) -\>
(matrix {{r + V * transpose V}} || transpose V ) |
( V || id\_((ring r)^n)
);
i94 : V = () -\> matrix table(1, n, (i,j) -\> random(0, R));
i95 : r = () -\> random(0, R);
We construct the ambient ring, local coordinates, and the ideal of the enumerative problem of lines in $`^4`$ tangent to 6 random spheres.
i96 : n = 4;
i97 : R = ZZ/1009\[flatten(table(2, n-1, (i,j) -\> z\_(i,j)))\];
i98 : X = 1 | matrix table(2, n-1, (i,j) -\> z\_(i,j))
o98 = | 1 0 z\_(0,0) z\_(0,1) z\_(0,2) |
| 0 1 z\_(1,0) z\_(1,1) z\_(1,2) |
2 5
o98 : Matrix R \<--- R
i99 : I = ideal (apply(1..(2\*n-2),
i -\> tanQuad(nSphere(V(), r()), X)));
o99 : Ideal of R
We find there are 24 lines in $`^4`$ tangent to 6 general spheres.
i100 : dim I, degree I
o100 = (0, 24)
o100 : Sequence
The expected numbers of solutions we have obtained in this way are displayed in the table below. The numbers in boldface are those which are proven.
| $`n`$ | 2 | 3 | 4 | 5 | 6 |
| --- | --- | --- | --- | --- | --- |
| \# expected | 4 | 12 | 24 | 48 | 96 |
## Acknowledgements
We thank Dan Grayson and Bernd Sturmfels; Some of the procedures in this chapter were written by Dan Grayson and the calculation in Section 5.2 is due to Bernd Sturmfels.
## Index
* Macaulay 2, 3, 4, 7, 15, 18
* Aluffi, P., 23
* artinian, see also ideal, zero-dimensional, 2
* Bézout number, 1, 21, 26
* Bézout Theorem, 1, 11, 17
* bilinear form, 21
+ signature, 8, 23
+ symmetric, 8
* Bottema, O., 12
* cylinder, 12
* discriminant, 13
* eliminant, 4–6, 14
* elimination theory, 4
* ellipsoid, 23
* enumerative geometry, 1, 3, 4, 10, 11, 14
+ real, 7, 18
* enumerative problem, 3, 11, 14, 20
+ fully real, 17, 19
* Eremenko, A., 18
* field
+ algebraically closed, 1, 13
+ splitting, 4
* Fulton, W., 23
* Gabrielov, A., 18
* Gröbner basis, 4
+ reduced, 15
* Grassmannian, 12–14, 26
+ local coordinates, 16
+ not a complete intersection, 16
* Grayson, D., 27
* Hilbert function, 15
* homotopy
+ Bézout, 11
+ Gröbner, 11, 17
+ optimal, 11
* homotopy continuation, 10
* hyperboloid, 23, 24
* ideal
+ degree, 2, 3
+ radical, 2, 4, 19
+ zero-dimensional, 2, 18, 22
* initial ideal
+ square-free, 11, 17
* Lichtblau, D., 12
* Macdonald, I., 13
* Pach, J., 13
* Plücker coordinate, 15, 20, 21, 26
* Plücker embedding, 15
* Plücker ideal, 15
* polynomial equations, 1, 3
+ deficient, 1
+ overdetermined, 16
* quadratic form, 13, 21
* rational normal curve, 17, 18
* saturate, 3, 22
* Schubert calculus, 14, 17–19
* Shape Lemma, 4, 19
* Shapiro, B., 18
* Shapiro, M., 18
* Shapiros’s Conjecture, 18
* solving polynomial equations, 3
+ real solutions, 6
+ via eigenvectors, 6
+ via elimination, 4
+ via numerical homotopy, 10
* sphere, 13, 22, 26
* Stickelberger’s Theorem, 6
* Sturm sequence, 7
* Sturmfels, B., 27
* symbolic computation, 2
* Theobald, Th., 13
* trace form, 8
* Veldkamp, G., 12
|
warning/0007/hep-ph0007057.html
|
ar5iv
|
text
|
# Testing Lorentz invariance violations in the tritium beta-decay anomaly
## 1 Introduction
Recent research on the determination of neutrino mass by studying the low-energy beta decay spectrum of tritium has produced a best fit value for $`m_\nu ^2`$ which is significantly negative . This unphysical value is caused by an anomalous excess of electron events at the end of the spectrum, at about 20 eV below the end point. The origin of this anomaly is not known. It seems clear that there is some systematic effect not taken into account, most probably of experimental nature. However, in this letter we want to point that an apparent excess of electron events near the end of the spectrum is compatible with a certain deviation of the relativistic dispersion relation for the neutrino. In this way, the tritium experimental results could be used to put bounds on the parameters characterizing this violation of Lorentz symmetry.
The idea that Lorentz invariance is not an exact symmetry, but an approximate one which works extremely well in our low-energy experiments, is not new. In the third section of this letter we will review the main proposals for modifications of the energy-momentum relativistic dispersion relation. There we will argue that a plausible dispersion relation for the neutrino, and only for this particle, is the following one:
$$E_\nu ^2=𝒑_𝝂^2+m_\nu ^2+2\lambda |𝒑_𝝂|,$$
(1)
where $`|𝒑|`$ means the module of the three-momentum $`𝒑`$, $`m_\nu `$ is the neutrino mass, and $`\lambda `$ is some mass scale, to be determined afterwards. We will consider $`\lambda >0`$ in order to have a positive contribution to the energy squared.
## 2 Kurie plot
Let us see how the usual Kurie plot of the tritium beta decay is modified by the dispersion relation Eq. (1). The phase-space factor is
$$\mathrm{d}\mathrm{\Pi }=p_e^2\mathrm{d}p_ep_\nu ^2\mathrm{d}p_\nu \delta (QEE_\nu ),$$
(2)
where $`p_e`$ is the momentum of the electron, $`E`$ is the kinetic energy of the electron ($`E_e=m_e+E`$) and $`Q`$ is the energy available to distribute between the neutrino energy and the kinetic energy of the electron. $`Q`$ is given by
$$MM^{}=Q+m_e,$$
(3)
$`M`$ and $`M^{}`$ being the masses of the initial and final nuclei. We are neglecting here the kinetic energy of the nucleus, which is typically of order $`10^1`$ eV. The Kurie plot is proportional to the function $`K(E)`$ given by
$$K(E)=\left[\delta (QEE_\nu )p_\nu ^2dp_\nu \right]^{1/2}.$$
(4)
For the usual relativistic dispersion relations, $`E_e^2=𝒑_𝒆^2+m_e^2`$ and $`E_\nu ^2=𝒑_𝝂^2+m_\nu ^2`$, one gets
$$K(E)=\left[(QE)\sqrt{(QE)^2m_\nu ^2}\right]^{1/2},$$
(5)
which is a straight line, $`(QE)`$, in the case $`m_\nu =0`$.
If one takes the new dispersion relation for the neutrino, Eq. (1), then the function $`K(E)`$ becomes (from now on, we will use the notation $`mm_\nu `$)
$`K(E)`$ $`=`$ $`(QE)^{1/2}\{[(QE)^2+\lambda ^2m^2]^{1/4}`$ (6)
$`{\displaystyle \frac{\lambda }{\left[(QE)^2+\lambda ^2m^2\right]^{1/4}}}\}.`$
It is easy to check that the end point of this curve is $`E=Qm`$, just as in the standard case Eq. (5). On the other hand, for large $`QE`$ values, the curve is very well approximated by a straight line. Considering a point $`E_l`$ so that $`QE_lm,\lambda `$, we introduce the linear approximation
$$\overline{K}_{E_l}(E)=K(E_l)+\frac{K(E)}{E}|_{E=E_l}(EE_l),$$
(7)
i.e., it is the tangent to $`K(E)`$ at the point $`E=E_l`$. Expanding $`\overline{K}_{E_l}(E)`$ in powers of $`\lambda /(QE_l)`$, $`m/(QE_l)`$, we get
$$\overline{K}_{E_l}(E)=(QE)\lambda +\frac{\lambda ^2m^2}{4(QE_l)}+\frac{\lambda ^2m^2}{4(QE_l)^2}(EE_l)+\mathrm{}$$
(8)
so that, instead of having a straight line of slope $`1`$ ending at $`E=Q`$, which is the standard case with $`m=\lambda =0`$, we obtain a straight line of slope
$$1+\frac{1}{4}\frac{\lambda ^2m^2}{(QE_l)^2}+\mathrm{}$$
(9)
which ends at the point
$$EE_0=Q\lambda +\frac{1}{2}\frac{\lambda ^2m^2}{QE_l}+\mathrm{}$$
(10)
On the other hand, one can calculate the exact value of the slope at the end point:
$$\frac{K(E)}{E}|_{E=Qm}=\left(\frac{m}{\lambda }\right)^{3/2}.$$
(11)
Two cases are clearly distinguished: $`\lambda >m`$ and $`\lambda <m`$. These are shown in Fig. 1. We see that near the end of the spectrum, the curve $`K(E)`$ is above the linear approximation $`\overline{K}_{E_l}(E)`$ when $`\lambda >m`$, which corresponds to an apparent excess of electrons at high energies. Indeed it is only apparent, because the curve lies always below the corresponding curve of a relativistic dispersion relation for a massless neutrino, which is also indicated in the figure. In the $`\lambda <m`$ case, we get the oposite situation: the effect due to the neutrino mass dominates over the $`\lambda `$ term (responsible for the “apparent excess”) and we get a reduction on the number of electrons at high energies.
The tritium anomaly consists in an excess of electron events at high energies, where “excess” here means that the data stay above the straight line which is the linear approximation to the curve at low energies. A possible measure of the anomaly could be given by using the following quantity
$$R_\mathrm{\Delta }=4\frac{_{E_0\mathrm{\Delta }/2}^{E_0}dEK(E)}{_{E_0\mathrm{\Delta }}^{E_0}dEK(E)},$$
(12)
so that $`R_\mathrm{\Delta }>1`$ indicates an apparent excess of electrons at high energies. This quantity could be measured experimentally, and a comparison with the theoretical prediction would put bounds on the values of the parameters $`\lambda ,m`$. We will show now the orders of magnitude of these bounds in a hypothetical but realistic example.
Introducing the variable $`x=(QE)/\lambda `$, and taking $`E_0Q\lambda `$, $`R_\mathrm{\Delta }`$ is rewritten as follows:
$$R_\mathrm{\Delta }=4\frac{_1^{1+\frac{\mathrm{\Delta }}{2\lambda }}dx\sqrt{x}\left[(x^2+\alpha )^{1/4}(x^2+\alpha )^{1/4}\right]}{_1^{1+\frac{\mathrm{\Delta }}{\lambda }}dx\sqrt{x}\left[(x^2+\alpha )^{1/4}(x^2+\alpha )^{1/4}\right]},$$
(13)
where $`\alpha 1m^2/\lambda ^2`$. Fig. 2 shows $`R_\mathrm{\Delta }`$ as a function of $`\mathrm{\Delta }/2\lambda `$ for different values of $`\alpha `$, that is, for different values of the quotient $`m/\lambda `$.
Experimentally, one could put a bound on the excess of electron events, that is, write $`R_\mathrm{\Delta }<1+ϵ_\mathrm{\Delta }`$ for a certain value of $`\mathrm{\Delta }`$, which should be chosen in an appropiate way. $`\mathrm{\Delta }`$ should be large enough, so that all the anomaly is contained in the region $`E_0\mathrm{\Delta }/2<E<E_0`$, but it should not be too large, because in that case $`R_\mathrm{\Delta }1`$. Let us do an estimate of orders of magnitude. From Table 1 of Ref. , we see that the fit which gives $`m_\nu ^2<0`$, signal of the anomaly, is stable and has a good $`\chi ^2/\mathrm{d}.\mathrm{o}.\mathrm{f}.`$ in a region between 200 and 400 eV before the end point, which is $`E_018570`$ eV. Let us take $`\mathrm{\Delta }/2=200`$ eV. An approximate value of $`R_\mathrm{\Delta }`$ comes then from Fig. 2 of Ref. , which contains the experimental data for the Kurie plot. $`R_\mathrm{\Delta }`$ is approximately given by $`4(s+t)/(4s+t)`$, where $`s`$ is the area of a triangle of base and height $`\mathrm{\Delta }/2`$ (the straight line of the Kurie plot has slope $`1`$), and $`t`$ is the area delimited by the data above the straight line, spread in a range of energies of around 20 eV, and the line itself. A rough estimate of this area gives us a conservative value of $`ϵ_\mathrm{\Delta }=1.510^3`$. For the four curves of Fig. 2, corresponding to four different values of the quotient $`m/\lambda `$, we obtain bounds for $`\mathrm{\Delta }/2\lambda `$, which, recalling that we have taken $`\mathrm{\Delta }/2=200`$ eV, produce bounds for $`\lambda `$ which go from $`\lambda <12.5`$ eV to $`\lambda <6.5`$ eV for the extreme curves. Again, this translates into bounds for the neutrino mass, depending on the value of $`\alpha `$. For $`\lambda =2m`$, the bound is $`m<3.8`$ eV, and for $`\lambda =4m`$, we get $`m<1.7`$ eV.
We see that the tritium anomaly puts bounds for $`(\lambda ,m)`$ in a very interesting range (a few eV). Of course, a detailed analysis of the experimental data would give finer bounds to this kind of Lorentz invariance violation. Even more interesting would be a possible (future) experimental bound of the type $`R_\mathrm{\Delta }>1+ϵ_\mathrm{\Delta }`$, with $`ϵ_\mathrm{\Delta }>0`$ (that is, the confirmation of the anomaly), which would give a lower bound for $`\lambda `$, showing the presence of Lorentz invariance violation effects in the tritium beta decay.
## 3 Lorentz invariance violations
Special relativity and Lorentz invariance are at the base of our low-energy effective theories. However, it may be possible that these are low-energy symmetries of a larger theory that do not need to be Lorentz invariant. Several attempts have been made to question Lorentz invariance and put bounds on possible violations (see e.g. ).
One way to explore the potentially observable effects of departures from exact Lorentz invariance is to consider the consequences in low-energy processes of possible extensions of the Lorentz-invariant particle energy-momentum relation compatible with translational and rotational invariance in a “preferred frame”. A first possibility is given by the Lorentz-violating class of dispersion relations
$$E^2=𝒑^2+m^2+𝒑^2\left(\frac{|𝒑|}{M}\right)^n,$$
(14)
where $`M`$ is the natural mass scale of the Lorentz non-invariant fundamental theory. An analysis of cosmic ray processes, whose thresholds are drastically changed by the new dispersion relation, puts severe bounds on the scale of the Lorentz violation: $`M`$ has to be several orders of magnitude larger than the Planck mass scale .
Another possible extension of the energy-momentum relation, coming from the introduction of a rotationally invariant two-derivative term in the free Lagrangian, is
$$E^2=𝒑^2+m^2+ϵ𝒑^2,$$
(15)
where $`ϵ`$ is a small coefficient which fixes the maximal attainable velocity of each particle ($`v^2=1+ϵ`$). Differences among maximal attainable velocities of different particles lead to abrupt effects when the dimensionless ratio $`ϵ𝒑^2/m^2`$ is of order unity. The precise tests of special relativity give very strong constraints on this type of extension ($`ϵ<10^{23}`$).
A discussion of departures from exact Lorentz invariance in terms of modifications of the energy-momentum relation leads to consider a third possible extension of the dispersion relation with a term linear in $`|𝒑|`$ ,
$$E^2=𝒑^2+m^2+2\lambda |𝒑|.$$
(16)
The additional term dominates over the standard kinetic term ($`𝒑^2`$) when $`|𝒑|2\lambda `$ and then the nonrelativistic kinematics is drastically changed. Therefore this type of generalized dispersion relation has to be excluded, except just for one case. The neutrino has two characteristic properties: it has a very small mass, and it interacts only weakly. As a result of this combination, we have not any experimental result on its nonrelativistic physics. Therefore, the presence of Lorentz invariance violations affecting the nonrelativistic limit cannot be excluded a priori in the neutrino case.
One possible way to incorporate a departure of Lorentz invariance affecting only to the neutrinos is to assume that the presence of a linear term in the dispersion relation is due to a new interaction which acts as a messenger of the Lorentz non-invariance at high energies. Once again, this is not a weird assumption: the introduction of new interactions is a general practice in the different attempts to explain the smallness of neutrino masses .
In conclusion, a dispersion relation of the form (1) can be considered for the neutrino at low energies. We still assume the existence of further Lorentz noninvariant terms, as those contained in Eqs. (14) and (15), for the neutrino as well as for all other particles. In the rest of the letter, we turn our attention to other implications of Eq. (1).
## 4 Other implications of the new dispersion relation
### 4.1 Neutrino oscillations
Let us first concentrate on the influence on a characteristic low-energy phenomenon for the neutrino: flavour oscillations. In the case of a Lorentz invariance violation independent of flavour, that is,
$$E_i^2=𝒑^2+m_i^2+2\lambda |𝒑|,$$
(17)
one gets the result
$$E_iE_j=\frac{m_i^2m_j^2}{2|𝒑|}+\mathrm{}$$
(18)
so that there is not any footprint of Lorentz noninvariance in neutrino oscillations. If we admit a possible dependence of $`\lambda `$ on the flavour, we get
$$E_iE_j=\lambda _i\lambda _j\frac{(\lambda _i^2\lambda _j^2)(m_i^2m_j^2)}{2|𝒑|}+\mathrm{}$$
(19)
and the oscillation probability becomes
$`P(\nu _i\nu _j)`$ $``$ $`\mathrm{sin}^22\theta \mathrm{sin}^2(L(\mathrm{Km})[(1.27){\displaystyle \frac{(m_i^2m_j^2)(\mathrm{eV}^2)}{E(\mathrm{GeV})}}`$ (20)
$`+(3.54)10^9(\lambda _i\lambda _j)(\mathrm{eV})]),`$
which means an energy independent probability (which goes against the experimental observations in solar neutrino oscillations) unless
$$|\lambda _i\lambda _j|10^{18}(\mathrm{eV}),$$
(21)
where we have used that $`L_{}=10^8`$ Km, and $`\lambda _i^2\lambda _j^2m_i^2m_j^2`$. There will be no footprint of Lorentz invariance violations in the case of atmospheric neutrinos, since $`L_{\mathrm{atm}}10^4L_{}`$, nor in accelerator experiments, with much smaller lengths. The conclusion is that the linear term in $`|𝒑|`$ in the generalized dispersion relation (1) has to be flavour-independent.
### 4.2 Contribution to the energy density of the Universe
The bounds on neutrino masses from their contribution to the energy density of the Universe are not affected by the presence of Lorentz invariance violations, since they only depend on the minimum energy of the neutrino, which is still $`E_{\mathrm{min}}=m`$ (for $`𝒑=0`$), independently of $`\lambda `$.
### 4.3 Neutrinos in astrophysics
The spread of arrival times of the neutrinos from SN1987A, coupled with the measured neutrino energies, provides a simple time-of-flight limit on $`m_{\nu _e}`$.
From Eq. (1) one has for the neutrino velocity
$$\beta =\frac{E}{|𝒑|}=\frac{1}{\sqrt{1\frac{1}{|𝒑|^2}\frac{\lambda ^2m^2}{(1+\lambda /|𝒑|)^2}}}.$$
(22)
Since $`|𝒑|\lambda `$, the limit (23 eV) can be taken as an upper limit for $`\sqrt{\lambda ^2m^2}`$ which is not far away from the range of parameters suggested by the tritium beta-decay anomaly. Then, neutrinos from supernovas can be a good place to look for footprints of Lorentz invariance violations. In fact, we see another implication from Eq. (22) if $`\lambda >m`$: neutrinos travel faster than light, so that those carrying more energy will be the latest to arrive, in contrast with the case of a relativistic dispersion relation.
### 4.4 Consequences in particle and cosmic ray physics
A striking consequence of the presence of the $`\lambda `$-term in the generalized neutrino dispersion relation Eq. (1) is the kinematic prohibition of reactions involving neutrinos at a certain energy. Let us consider ordinary neutron decay: $`np+e^{}+\overline{\nu }`$. At large momentum, the $`\lambda `$-term, which is proportional to $`|𝒑|`$, represents a large contribution to the energy balance, so that the process might be forbidden for neutrons of sufficiently high momentum. This turns out to be the case: it is easy to see that for $`\lambda /|𝒑|1`$ and $`|𝒑|m_n,m_p`$, the total energy of the final state is bounded from below,
$$E_p+E_e+E_{\overline{\nu }}>|𝒑|+\lambda +\frac{(m_p+m_e)^2}{2|𝒑|},$$
(23)
which implies that neutrons with momentum
$$|𝒑|>\frac{m_n^2(m_p+m_e)^2}{2\lambda }=\left[\frac{1.4710^{15}}{2\lambda (\mathrm{eV})}\right](\mathrm{eV})$$
(24)
are stable particles. A similar conclusion is obtained in the case of pions: the desintegration $`\pi ^{}\mu ^{}+\overline{\nu }_\mu `$ is forbidden if the pion has a momentum
$$|𝒑|>\frac{m_\pi ^2m_\mu ^2}{2\lambda }=\left[\frac{8.310^{15}}{2\lambda (\mathrm{eV})}\right](\mathrm{eV}).$$
(25)
This means that pions and neutrons with these energies can form part of cosmic rays, since they are stable particles if the neutrino energy-momentum relation is given by Eq. (1). Therefore, cosmic ray physics might be drastically affected at high energies, of order $`E10^{15}`$$`10^{16}`$ eV. In particular, the presence of stable neutrons and pions in cosmic rays at such energies might contribute to avoiding the well-known GZK cutoff .
## 5 Conclusions
It is surprising that a very simple extension of the Lorentz invariant dispersion relation for the neutrino, consistent with all the constraints from neutrino physics, can affect phenomena of such different energy ranges. It could be behind the tritium beta-decay anomaly, and also lead to a drastic change in the composition of cosmic rays beyond $`10^{16}`$ eV. It seems worthwhile to explore in more detail all the consequences of this violation of Lorentz invariance and possible models of the Lorentz non-invariant physics at high energies incorporating the extended neutrino dispersion relation considered in this letter as a low-energy remnant.
We are grateful to J.L. Alonso for collaboration at early stages of this work, to D.E. Groom for useful comments on the tritium anomaly, to A. Grillo, R. Aloisio and A. Galante for discussions on observable effects of violations of Lorentz invariance, and to J.M. Usón. The work of JMC was supported by EU TMR program ERBFMRX-CT97-0122 and the work of JLC by CICYT (Spain) project AEN-97-1680.
|
warning/0007/gr-qc0007025.html
|
ar5iv
|
text
|
# Generalized Equivalence Principle in Extended New General Relativity
## 1 Introduction
Energy-momentum, angular momentum and electric charge play central roles in modern physics. The conservation properties of the first two are related to the homogeneity and isotropy of space-time, respectively, and charge conservation corresponds to the invariance of the action integral under internal $`U(1)`$ transformations. Also, local quantities such as energy-momentum density, angular momentum density and charge density are well defined if gravitational fields are not present in the system in question.
In general relativity (G. R.), however, well-behaved energy-momentum and angular momentum densities have not yet been defined, although total energy-momen-tum and total angular momentum can be defined for an asymptotically flat space-time surrounding an isolated finite system. The total energy of the system is regarded as the inertial mass multiplied by the square of the velocity of light, and there arises the following question: Is the active gravitational mass of the isolated system equal to its inertial mass? This question regards an aspect of the equivalence principle, and it is usually considered to be affirmatively answered within G. R. However, the equality of the active gravitational mass and the inertial mass is violated for the Schwarzschild metric when it is expressed with using a certain coordinate system.
New general relativity (N. G. R.), which is formulated by gauging coordinate translations and is constructed within the Weitzenböck space-time, is a possible alternative to G. R. The most general gravitational Lagrangian density, which is quadratic in the torsion tensor and is invariant under global Lorentz transformations including also space inversion and general coordinate transformations, has three free parameters, $`c_1,c_2`$ and $`c_3`$. Solar system experiments show that $`c_1`$ and $`c_2`$ are very likely to be equal to $`1/(3\kappa )`$. In Ref., Shirafuji, Nashed and Hayashi give, for the case with $`c_1=1/(3\kappa )=c_2`$, the most general spherically symmetric solution, and they clarified the restriction on the behavior of the vierbeins at spatial infinity imposed by the requirement that the inertial mass is equal to the active gravitational mass. Such analysis has been extended to the case in which $`c_11/(3\kappa )c_2`$ in Ref..
Extended new general relativity (E. N. G. R.), is obtained as a reduction of the $`\overline{\text{Poincaré}}`$ gauge theory (P̄. G. T.) of gravity, which is formulated on the basis of the principal fiber bundle over the space-time, possessing the covering group $`\overline{P}_0`$ of the Poincaré group as the structure group, by following the lines of the standard geometric formulation of Yang-Mills theories as closely as possible. E. N. G. R. is also constructed within the Weitzenböck space-time and has many points in common with N. G. R. The field equations for the vierbeins in E. N. G. R. , for example, are identical to those in N. G. R., if fields with non-vanishing “intrinsic” energy-momentum do not exist.
In this paper, considering charged axi-symmetric solutions in E. N. G. R., we examine the condition imposed on the asymptotic behavior of the field variables by the requirement<sup>3</sup><sup>3</sup>3This is a generalization of the requirement that the inertial mass is equal to the active gravitational mass. that the total energy-momentum, the total angular momentum and the total charge of the system are all equal to the corresponding quantities of the central gravitating body.
This paper is organized as follows. In §2, we give the basic formulation of E. N. G. R. In §3, we calculate conserved quantities for a charged axi-symmetric solution. In §4, we study solutions obtainable from the solution discussed in §3 and examine restrictions on the field variables imposed by the requirement mentioned above. In the final section, we give a summary and discussion.
## 2 Basic formulation
### 2.1 $`\overline{\text{Poincaré}}`$ gauge theory
P̄. G. T. is formulated on the basis of the principal fiber bundle $`𝒫`$ over the space-time $`M`$ possessing the covering group $`\overline{P}_0`$ of the proper orthochronous Poincaré group as the structure group. The space-time is assumed to be a noncompact four- dimensional differentiable manifold having a countable base. The bundle $`𝒫`$ admits a connection $`\mathrm{\Gamma }`$. The translational and rotational parts of the coefficients of $`\mathrm{\Gamma }`$ will be written as $`A^k_\mu `$ and $`A^k_{l\mu }`$, respectively. The fundamental field variables are $`A^k{}_{\mu }{}^{},A^k_{l\mu }`$, the Higgs-type field $`\psi =\{\psi ^k\}`$, and the matter field $`\varphi =\{\varphi ^A|A=1,2,3,\mathrm{},N\}`$.<sup>4</sup><sup>4</sup>4Unless otherwise stated, we use the following conventions for indices: The middle part of the Greek alphabet, $`\mu ,\nu ,\lambda ,\mathrm{}`$, denotes 0, 1, 2 and 3, while the initial part, $`\alpha ,\beta ,\gamma ,\mathrm{}`$, denotes 1, 2 and 3. In a similar way, the middle part of the Latin alphabet, $`i,j,k,\mathrm{}`$, denotes 0, 1, 2 and 3, unless otherwise stated, while the initial part, $`a,b,c,\mathrm{}`$, denotes 1, 2 and 3. The capital letters $`A`$ and $`B`$ are used for indices of components of the field $`\varphi `$, and $`N`$ denotes the dimension of the representation $`\rho `$. These fields transform as<sup>5</sup><sup>5</sup>5For the function $`f`$ on $`M`$, we define $`f_{,\mu }\stackrel{\text{def}}{=}f/x^\mu `$.
$`\psi ^k`$ $`=`$ $`(\mathrm{\Lambda }(a^1))^k{}_{l}{}^{}(\psi ^lt^l),`$
$`A^k_\mu `$ $`=`$ $`(\mathrm{\Lambda }(a^1))^k{}_{l}{}^{}(A^l{}_{\mu }{}^{}+t^l{}_{,\mu }{}^{}+A^l{}_{m\mu }{}^{}t_{}^{m}),`$
$`A^k_{l\mu }`$ $`=`$ $`(\mathrm{\Lambda }(a^1))^k{}_{m}{}^{}A_{}^{m}{}_{n\mu }{}^{}(\mathrm{\Lambda }(a))_{}^{n}{}_{l}{}^{}+(\mathrm{\Lambda }(a^1))^k{}_{m}{}^{}(\mathrm{\Lambda }(a))_{}^{m}{}_{l,\mu }{}^{},`$
$`\varphi ^A`$ $`=`$ $`(\rho ((t,a)^1))^A{}_{B}{}^{}\varphi _{}^{B},`$ (1)
under the $`\overline{\text{Poincaré}}`$ gauge transformation
$$\sigma ^{}(x)=\sigma (x)(t(x),a(x)),t(x)T^4,a(x)SL(2,C).$$
(2)
Here, $`\mathrm{\Lambda }`$ is the covering map from $`SL(2,C)`$ to the proper orthochronous Lorentz group, and $`\rho `$ stands for the representation of the $`\overline{\text{Poincaré}}`$ group to which the field $`\varphi `$ belongs. Also, $`\sigma `$ and $`\sigma ^{}`$ represent the local cross sections of $`𝒫`$. The dual components $`b^k_\mu `$ of the vierbeins $`b^\mu {}_{k}{}^{}/x^\mu `$ are related to the field $`\psi `$ and the gauge potentials $`A^k_\mu `$ and $`A^k_{l\mu }`$ through the relation
$$b^k{}_{\mu }{}^{}=\psi ^k{}_{,\mu }{}^{}+A^k{}_{l\mu }{}^{}\psi _{}^{l}+A^k{}_{\mu }{}^{},$$
(3)
and these transform according to
$$b^k{}_{\mu }{}^{}=(\mathrm{\Lambda }(a^1))^k{}_{l}{}^{}b_{}^{l}_\mu $$
(4)
under the transformation (2). Also, they are related to the metric $`g_{\mu \nu }dx^\mu dx^\nu `$ of $`M`$ through the relation
$$g_{\mu \nu }=\eta _{kl}b^k{}_{\mu }{}^{}b_{}^{l}{}_{\nu }{}^{},$$
(5)
with $`(\eta _{kl})\stackrel{\text{def}}{=}\text{diag}(1,1,1,1)`$.
There is a 2 to 1 bundle homomorphism $`F`$ from $`𝒫`$ to the affine frame bundle $`𝒜(M)`$ over $`M`$, and an extended spinor structure and a spinor structure exist in association with it. The space-time $`M`$ is orientable, which follows from its assumed noncompactness and from the fact that $`M`$ has a spinor structure.
The affine frame bundle $`𝒜(M)`$ admits a connection $`\mathrm{\Gamma }_A`$. The $`T^4`$-part, $`\mathrm{\Gamma }^\mu _\nu `$, and $`GL(4,R)`$-part, $`\mathrm{\Gamma }^\mu _{\nu \lambda }`$, of its connection coefficients are related to $`A^k_{l\mu }`$ and $`b^k_\mu `$ through the relations
$$\mathrm{\Gamma }^\mu {}_{\nu }{}^{}=\delta ^\mu {}_{\nu }{}^{},A^k{}_{l\mu }{}^{}=b^k{}_{\nu }{}^{}b_{}^{\lambda }{}_{l}{}^{}\mathrm{\Gamma }_{}^{\nu }{}_{\lambda \mu }{}^{}+b^k{}_{\nu }{}^{}b_{}^{\nu }_{l,\mu }$$
(6)
by the requirement that $`F`$ maps the connection $`\mathrm{\Gamma }`$ into $`\mathrm{\Gamma }_A`$ , and the space-time $`M`$ is of the Riemann-Cartan type.
The field strengths $`R^k_{l\mu \nu }`$, $`R^k_{\mu \nu }`$ and $`T^k_{\mu \nu }`$ of $`A^k_{l\mu }`$, $`A^k_\mu `$ and of $`b^k_\mu `$ are given by<sup>6</sup><sup>6</sup>6We define $`A_{\mathrm{}[\mu \mathrm{}\nu ]\mathrm{}}`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{1}{2}}(A_{\mathrm{}\mu \mathrm{}\nu \mathrm{}}A_{\mathrm{}\nu \mathrm{}\mu \mathrm{}}),`$ $`A_{\mathrm{}(\mu \mathrm{}\nu )\mathrm{}}`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{1}{2}}(A_{\mathrm{}\mu \mathrm{}\nu \mathrm{}}+A_{\mathrm{}\nu \mathrm{}\mu \mathrm{}}).`$
$`R^k_{l\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`2(A^k{}_{l[\nu ,\mu ]}{}^{}+A^k{}_{m[\mu }{}^{}A_{}^{m}{}_{l\nu ]}{}^{}),`$
$`R^k_{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`2(A^k{}_{[\nu ,\mu ]}{}^{}+A^k{}_{l[\mu }{}^{}A_{}^{l}{}_{\nu ]}{}^{}),`$
$`T^k_{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`2(b^k{}_{[\nu ,\mu ]}{}^{}+A^k{}_{l[\mu }{}^{}b_{}^{l}{}_{\nu ]}{}^{}),`$ (7)
and we have the relation
$$T^k{}_{\mu \nu }{}^{}=R^k{}_{\mu \nu }{}^{}+R^k{}_{l\mu \nu }{}^{}\psi _{}^{l}.$$
(8)
The field strengths $`T^k_{\mu \nu }`$ and $`R^{kl}_{\mu \nu }`$ are both invariant under *internal* translations. The torsion is given by
$$T^\lambda {}_{\mu \nu }{}^{}=2\mathrm{\Gamma }^\lambda {}_{[\nu \mu ]}{}^{},$$
(9)
and the $`T^4`$\- and $`GL(4,R)`$-parts of the curvature are given by
$`R^\lambda _{\mu \nu }`$ $`=`$ $`2(\mathrm{\Gamma }^\lambda {}_{[\nu ,\mu ]}{}^{}+\mathrm{\Gamma }^\lambda {}_{\rho [\mu }{}^{}\mathrm{\Gamma }_{}^{\rho }{}_{\nu ]}{}^{}),`$ (10)
$`R^\lambda _{\rho \mu \nu }`$ $`=`$ $`2(\mathrm{\Gamma }^\lambda {}_{\rho [\nu ,\mu ]}{}^{}+\mathrm{\Gamma }^\lambda {}_{\tau [\mu }{}^{}\mathrm{\Gamma }_{}^{\tau }{}_{\rho \nu ]}{}^{}),`$ (11)
respectively. Also, we have
$`T^k_{\mu \nu }`$ $`=`$ $`b^k{}_{\lambda }{}^{}T_{}^{\lambda }{}_{\mu \nu }{}^{}=b^k{}_{\lambda }{}^{}R_{}^{\lambda }{}_{\mu \nu }{}^{},`$ (12)
$`R^k_{l\mu \nu }`$ $`=`$ $`b^k{}_{\lambda }{}^{}b_{}^{\rho }{}_{l}{}^{}R_{}^{\lambda }{}_{\rho \mu \nu }{}^{},`$ (13)
which follow from Eq. (6).
The covariant derivative of the matter field $`\varphi `$ takes the form
$`D_k\varphi ^A`$ $`=`$ $`b^\mu {}_{k}{}^{}D_{\mu }^{}\varphi ^A,`$
$`D_\mu \varphi ^A`$ $`\stackrel{\text{def}}{=}`$ $`_\mu \varphi ^A+{\displaystyle \frac{i}{2}}A^{lm}{}_{\mu }{}^{}(M_{lm}\varphi )_{}^{A}+iA^l{}_{\mu }{}^{}(P_l\varphi )_{}^{A}.`$ (14)
Here, $`M_{kl}`$ and $`P_k`$ are representation matrices of the standard basis of the group $`\overline{P}_0`$ : $`M_{kl}=i\rho _{}(\overline{M}_{kl}),P_k=i\rho _{}(\overline{P}_k)`$. The matrix $`P_k`$ represents the intrinsic energy-momentum of the field $`\varphi ^A`$, and it is vanishing for all the observed fields.<sup>7</sup><sup>7</sup>7In what follows, the field components $`b^k_\mu `$ and $`b^\mu _k`$ are used to convert Latin and Greek indices, in analogy to the case of $`D_k\varphi ^A`$ and $`D_\mu \varphi ^A`$. Also, raising and lowering the indices $`k,l,m,\mathrm{}`$ are accomplished with the aid of $`(\eta ^{kl})\stackrel{\text{def}}{=}(\eta _{kl})^1`$ and $`(\eta _{kl})`$.
From the requirement of invariance of the action integral under internal $`\overline{P}_0`$ gauge transformations, it follows that the gravitational Lagrangian density is a function of $`T_{klm}`$ and of $`R_{klmn}`$. The gravitational Lagrangian density is identical to that in Poincaré gauge theory, and hence *gravitational field equations take the same forms in these theories*. The field equation for the field $`\psi ^k`$ is automatically satisfied if those for $`A^k_\mu `$ and $`\varphi ^A`$ are both satisfied, and $`\psi ^k`$ is a non-dynamical field in this sense.
### 2.2 Extended new general relativity
#### 2.2.1 Reduction of $`\overline{\text{Poincaré}}`$ gauge theory to an extended new general relativity
In P̄. G. T., we consider the case in which the field strength $`R^{kl}_{\mu \nu }`$ vanishes identically,
$$R^{kl}{}_{\mu \nu }{}^{}0,$$
(15)
and we choose $`SL(2.C)`$-gauge such that
$$A^{kl}{}_{\mu }{}^{}0.$$
(16)
Then, the curvature vanishes, and the space-time reduces to the Weitzenböck space-time, which means that we have a teleparallel theory.
Also, the vierbeins $`b^k_\mu `$, the affine connection coefficients $`\mathrm{\Gamma }^\lambda _{\mu \nu }`$ and the torsion tensor $`T^\lambda _{\mu \nu }`$ are given by
$`b^k_\mu `$ $`=`$ $`\psi ^k{}_{,\mu }{}^{}+A^k{}_{\mu }{}^{},`$ (17)
$`\mathrm{\Gamma }^\lambda _{\mu \nu }`$ $`=`$ $`b^\lambda {}_{l}{}^{}b_{}^{l}{}_{\mu ,\nu }{}^{},`$ (18)
and
$$T^\lambda {}_{\mu \nu }{}^{}=2\mathrm{\Gamma }^\lambda {}_{[\nu \mu ]}{}^{}=2b^\lambda {}_{k}{}^{}b_{}^{k}{}_{[\nu ,\mu ]}{}^{}=b^\lambda {}_{k}{}^{}T_{}^{k}{}_{\mu \nu }{}^{},$$
(19)
respectively. Introducing the volume element $`dv`$ by
$$dv\stackrel{\text{def}}{=}\sqrt{g}dx^0dx^1dx^2dx^3,$$
(20)
with $`g\stackrel{\text{def}}{=}det(g_{\mu \nu })=\{det(b^k{}_{\mu }{}^{})\}^2`$, we consider the action integral
$$𝑰\text{ }=_𝒟L(\psi ^k{}_{,\mu }{}^{},\psi ^k,A^k{}_{\mu ,\nu }{}^{},A^k{}_{\mu }{}^{},A_{\mu ,\nu },A_\mu ,\varphi ^A{}_{,\mu }{}^{},\varphi ^A,\varphi ^A{}_{,\mu }{}^{},\varphi ^A)dv,$$
(21)
where $`A_\mu `$ and $`𝒟`$ denote the electromagnetic vector potential and a compact region in $`M`$, respectively.<sup>8</sup><sup>8</sup>8We consider the case in which the electromagnetic field and the charged field are present. The field $`\varphi ^A`$ is considered to be a field with the electric charge $`q`$. This is preparation for the subsequent sections, in which space-time produced by a charged source is treated. Also, the symbol $``$ represents the operation of complex conjugation, and thus $`\varphi ^A`$ denotes the complex conjugate of $`\varphi ^A`$.
We impose the following requirement:
(R.i) The action $`𝑰`$ is invariant under local internal translations and global $`SL(2,C)`$-transformations. From this, the identities
$`{\displaystyle \frac{\delta 𝑳}{\delta \psi ^k}}+_\mu \left({\displaystyle \frac{\delta 𝑳}{\delta A^k_\mu }}\right)+i{\displaystyle \frac{\delta 𝑳}{\delta \varphi ^A}}(P_k\varphi )^Ai{\displaystyle \frac{\delta 𝑳}{\delta \varphi ^A}}(P_k\varphi )^A0,`$ (22)
$`𝑭_k{}_{}{}^{(\mu \nu )}0,`$ (23)
$`{}_{}{}^{\text{tot}}𝑻_{k}^{}{}_{}{}^{\mu }_\nu 𝑭_k{}_{}{}^{\mu \nu }{\displaystyle \frac{\delta 𝑳}{\delta A^k_\mu }}0,`$ (24)
$`_\mu {}_{}{}^{\text{tot}}𝑺_{kl}^{}{}_{}{}^{\mu }2{\displaystyle \frac{\delta 𝑳}{\delta \psi ^{[k}}}\psi _{l]}2{\displaystyle \frac{\delta 𝑳}{\delta A^{[k}_\mu }}A_{l]\mu }i{\displaystyle \frac{\delta 𝑳}{\delta \varphi ^A}}(M_{kl}\varphi )^A+i{\displaystyle \frac{\delta 𝑳}{\delta \varphi ^A}}(M_{kl}\varphi )^A0`$
follow, where we have defined
$`𝑳`$ $`\stackrel{\text{def}}{=}`$ $`\sqrt{g}L,`$ (26)
$`𝑭_k^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{𝑳}{A^k_{\mu ,\nu }}},`$ (27)
$`{}_{}{}^{\text{tot}}𝑻_{k}^{}^\mu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{𝑳}{\psi ^k_{,\mu }}}+i{\displaystyle \frac{𝑳}{\varphi ^A_{,\mu }}}(P_k\varphi )^Ai{\displaystyle \frac{𝑳}{\varphi ^A_{,\mu }}}(P_k\varphi )^A,`$ (28)
$`{}_{}{}^{\text{tot}}𝑺_{kl}^{}^\mu `$ $`\stackrel{\text{def}}{=}`$ $`2{\displaystyle \frac{𝑳}{\psi ^{[k}_{,\mu }}}\psi _{l]}2𝑭_{[k}{}_{}{}^{\nu \mu }A_{l]\nu }^{}`$ (29)
$`i{\displaystyle \frac{𝑳}{\varphi ^A_{,\mu }}}(M_{kl}\varphi )^A+i{\displaystyle \frac{𝑳}{\varphi ^A_{,\mu }}}(M_{kl}\varphi )^A.`$
By virtue of the identity (24), the density $`{}_{}{}^{\text{tot}}𝑻_{k}^{}^\mu `$ can be expressed in the usual form of the currrent in Yang-Mills theories,
$${}_{}{}^{\text{tot}}𝑻_{k}^{}{}_{}{}^{\mu }=\frac{𝑳}{A^k_\mu }.$$
(30)
When the field equations $`\delta 𝑳/\delta A^k{}_{\mu }{}^{}\stackrel{\text{def}}{=}𝑳/A^k{}_{\mu }{}^{}_\nu (𝑳/A^k{}_{\mu ,\nu }{}^{})=0`$ and $`\delta 𝑳/\delta \varphi ^A=0`$ are both satisfied, we have the following:
(i) The field equation $`\delta 𝑳/\delta \psi ^k=0`$ is automatically satisfied, and hence $`\psi ^k`$ is not an independent dynamical variable.
(ii)
$`_\mu {}_{}{}^{\text{tot}}𝑻_{k}^{}{}_{}{}^{\mu }=0,`$ (31)
$`_\mu {}_{}{}^{\text{tot}}𝑺_{kl}^{}{}_{}{}^{\mu }=0,`$ (32)
which are the differential conservation laws of the dynamical energy-momentum and the “spin” angular momentum, respectively. The assertions (i) and (ii) follow from Eqs. (22) – (2.2.1).
Also, we require the following:
(R.ii) The Lagrangian density $`L`$ is a scalar field on $`M`$.
Then, we have
$$\stackrel{~}{𝑻}_\mu {}_{}{}^{\nu }_\lambda 𝚿_\mu {}_{}{}^{\nu \lambda }\frac{\delta 𝑳}{\delta A^k_\nu }A^k{}_{\mu }{}^{}\frac{\delta 𝑳}{\delta A_\nu }A_\mu 0$$
(33)
and
$$𝚿_\lambda {}_{}{}^{(\mu \nu )}0,$$
(34)
with
$$\stackrel{~}{𝑻}_\mu {}_{}{}^{\nu }\stackrel{\text{def}}{=}\delta _\mu {}_{}{}^{\nu }𝑳𝑭_k{}_{}{}^{\lambda \nu }A_{}^{k}{}_{\lambda ,\mu }{}^{}𝑭^{\lambda \nu }A_{\lambda ,\mu }\frac{𝑳}{\varphi ^A_{,\nu }}\varphi ^A{}_{,\mu }{}^{}\frac{𝑳}{\varphi ^A_{,\nu }}\varphi ^A{}_{,\mu }{}^{}\frac{𝑳}{\psi ^k_{,\nu }}\psi ^k_{,\mu }$$
(35)
and
$`𝚿_\lambda ^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`𝑭_k{}_{}{}^{\mu \nu }A_{}^{k}{}_{\lambda }{}^{}+𝑭^{\mu \nu }A_\lambda ,`$ (36)
$`𝑭^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{𝑳}{A_{\mu ,\nu }}}.`$ (37)
The identities (33) and (34) lead to
$`_\nu \stackrel{~}{𝑻}_\mu ^\nu `$ $`=`$ $`0,`$ (38)
$`_\nu 𝑴_\lambda ^{\mu \nu }`$ $`=`$ $`0`$ (39)
when $`\delta 𝑳/\delta A^k{}_{\mu }{}^{}=0\mathrm{and}\delta 𝑳/\delta A_\mu =0`$, where $`𝑴_\lambda {}_{}{}^{\mu \nu }\stackrel{\text{def}}{=}2(𝚿_\lambda {}_{}{}^{\mu \nu }x^\mu \stackrel{~}{𝑻}_\lambda {}_{}{}^{\nu })`$. Equations (38) and (39) are the differential conservation laws of the canonical energy-momentum and “extended orbital angular momentum”, respectively. We also require invariance of the action under the $`U(1)`$ gauge transformation:
$$A^{}{}_{\mu }{}^{}=A_\mu +\lambda _{,\mu },\varphi ^A=\mathrm{exp}(iq\lambda )\varphi ^A,\psi ^k=\psi ^k,A^k{}_{\mu }{}^{}=A^k{}_{\mu }{}^{},$$
(40)
with $`\lambda `$ being an arbitrary function on $`M`$, from which we can obtain
$`𝑭^{(\mu \nu )}0,`$ (41)
$`𝑱^\mu +iq\left({\displaystyle \frac{𝑳}{\varphi ^A_{,\mu }}}\varphi ^A{\displaystyle \frac{𝑳}{\varphi ^A_{,\mu }}}\varphi ^A\right)0,`$ (42)
$`_\mu \left({\displaystyle \frac{\delta 𝑳}{\delta A_\mu }}𝑱^\mu \right)0,`$ (43)
with
$$𝑱^\mu \stackrel{\text{def}}{=}\frac{𝑳}{A_\mu }.$$
(44)
From the identity (43), the differential conservation law of the electric charge,
$$_\mu 𝑱^\mu =0,$$
(45)
follows when the field equation $`\delta 𝑳/\delta A_\mu =0`$ is satisfied. The functional dependence of $`L`$ is restricted as
$$L=(\psi ^k,T_{klm},F_{\mu \nu },_k\varphi ^A,\varphi ^A,^{}{}_{k}{}^{}\varphi _{}^{A},\varphi ^A),$$
(46)
with $``$ satisfying certain identities, where
$`F_{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`_\mu A_\nu _\nu A_\mu ,`$ (47)
$`_k\varphi ^A`$ $`\stackrel{\text{def}}{=}`$ $`b^\mu {}_{k}{}^{}_{\mu }^{}\varphi ^A\stackrel{\text{def}}{=}b^\mu {}_{k}{}^{}(\varphi ^A{}_{,\mu }{}^{}+iA^k{}_{\mu }{}^{}(P_k\varphi )_{}^{A}iqA_\mu \varphi ^A),`$
$`^{}{}_{k}{}^{}\varphi _{}^{A}`$ $`\stackrel{\text{def}}{=}`$ $`b^\mu {}_{k}{}^{}_{}^{}{}_{\mu }{}^{}\varphi _{}^{A}\stackrel{\text{def}}{=}b^\mu {}_{k}{}^{}(\varphi ^A{}_{,\mu }{}^{}iA^k{}_{\mu }{}^{}(P_k\varphi )_{}^{A}+iqA_\mu \varphi ^A).`$ (48)
The gravitational action<sup>9</sup><sup>9</sup>9This action is identical to the gravitational action in N. G. R.
$$\stackrel{~}{𝑰}_G=_𝒟(c_1t^{klm}t_{klm}+c_2v^kv_k+c_3a^ka_k)𝑑v,$$
(49)
with $`c_i(i=1,2,3)`$ being real constants, satisfies the requirements (R.i) and (R.ii). Here $`t_{klm}`$, $`v_k`$ and $`a_k`$ are irreducible components of the torsion tensor:
$`t_{klm}`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{1}{2}}(T_{klm}+T_{lkm})+{\displaystyle \frac{1}{6}}(\eta _{mk}v_l+\eta _{ml}v_k){\displaystyle \frac{1}{3}}\eta _{kl}v_m,`$ (50)
$`v_k`$ $`\stackrel{\text{def}}{=}`$ $`T^l{}_{lk}{}^{},`$ (51)
$`a_k`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{1}{6}}\epsilon _{klmn}T^{lmn},`$ (52)
where $`\epsilon _{klmn}`$ is a completely anti-symmetric Lorentz tensor with $`\epsilon _{(0)(1)(2)(3)}=1`$.<sup>10</sup><sup>10</sup>10Latin indices are put in parentheses to discriminate them from Greek indices.
The action $`\stackrel{~}{𝑰}_G`$ with<sup>11</sup><sup>11</sup>11We will use natural units in which $`\mathrm{}=c=1`$.
$$c_1=c_2=\frac{1}{3\kappa },$$
(53)
with $`\kappa `$ being the Einstein gravitational constant agrees quite well with experimental results. In what follows, we assume that the condition (53) is satisfied. Thus, our gravitational action is
$$𝑰_G\stackrel{\text{def}}{=}_𝒟L_G𝑑v,$$
(54)
with
$$L_G\stackrel{\text{def}}{=}\frac{1}{3\kappa }(t^{klm}t_{klm}v^kv_k)+c_3a^ka_k.$$
(55)
#### 2.2.2 The gravitational and electromagnetic field equations in vacuum
The electromagnetic Lagrangian density $`L_{em}`$ is given by<sup>12</sup><sup>12</sup>12Here we use Heaviside-Lorentz rationalized unit.
$$L_{em}=\frac{1}{4}g^{\mu \rho }g^{\nu \sigma }F_{\mu \nu }F_{\rho \sigma }.$$
(56)
We consider a system described by the Lagrangian density $`L\stackrel{\text{def}}{=}L_G+L_{em}`$. The gravitational and electromagnetic field equations for this system are the following:
$`G^{\mu \nu }(\{\})+K^{\mu \nu }`$ $`=`$ $`\kappa T_{em}^{\mu \nu },`$ (57)
$`_\mu (\sqrt{g}J^{kl\mu })`$ $`=`$ $`0,`$ (58)
$`_\nu \left(\sqrt{g}F^{\mu \nu }\right)`$ $`=`$ $`0.`$ (59)
Here we have defined
$$G_{\mu \nu }(\{\})\stackrel{\text{def}}{=}R_{\mu \nu }(\{\})\frac{1}{2}g_{\mu \nu }R(\{\}),$$
(60)
and
$$R_{\mu \nu }(\{\})\stackrel{\text{def}}{=}R^\lambda {}_{\mu \lambda \nu }{}^{}(\{\}),R(\{\})\stackrel{\text{def}}{=}g^{\mu \nu }R_{\mu \nu }(\{\}),$$
(61)
with the Riemann-Christoffel curvature tensor
$$R^\lambda {}_{\rho \mu \nu }{}^{}(\{\})\stackrel{\text{def}}{=}2(_{[\mu }\left\{\genfrac{}{}{0pt}{}{\lambda }{\rho \nu ]}\right\}+\left\{\genfrac{}{}{0pt}{}{\lambda }{\sigma [\mu }\right\}\left\{\genfrac{}{}{0pt}{}{\sigma }{\rho \nu ]}\right\}).$$
(62)
Also, $`T_{em}^{\mu \nu }`$ is the energy-momentum tensor of the electromagnetic field,
$$T_{em}^{\mu \nu }\stackrel{\text{def}}{=}F^{\mu \rho }F^{\nu \sigma }g_{\rho \sigma }+g^{\mu \nu }L_{em},$$
(63)
and the tensors $`K^{\mu \nu }`$ and $`J^{kl\mu }`$ are defined by
$$K^{\mu \nu }\stackrel{\text{def}}{=}\frac{\kappa }{\lambda }[\frac{1}{2}\{\epsilon {}_{}{}^{\mu \rho \sigma \lambda }(T^\nu {}_{\rho \sigma }{}^{}T_{\rho \sigma }{}_{}{}^{\nu })+\epsilon {}_{}{}^{\nu \rho \sigma \lambda }(T^\mu {}_{\rho \sigma }{}^{}T_{\rho \sigma }{}_{}{}^{\mu })\}a_\lambda \frac{3}{2}a^\mu a^\nu \frac{3}{4}g^{\mu \nu }a^\lambda a_\lambda ]$$
(64)
and
$$J^{kl\mu }\stackrel{\text{def}}{=}\frac{3}{2}b^k{}_{\rho }{}^{}b_{}^{l}{}_{\sigma }{}^{}\epsilon {}_{}{}^{\rho \sigma \mu \nu }a_{\nu }^{},$$
(65)
respectively, where we have used
$$\lambda \stackrel{\text{def}}{=}\frac{9\kappa }{4\kappa c_3+3}.$$
(66)
#### 2.2.3 An exact solution of gravitational and electromagnetic field equations with a charged rotating source
An exact solution of the field equations (57) – (59), which represents the gravitational and electromagnetic fields surrounding a charged rotating source, is given in Ref.. This will be described below. The vector fields $`b^k_\mu `$ and the electromagnetic potential $`A_\mu `$ have the expressions
$`b^k_\mu `$ $`=`$ $`{}_{}{}^{(0)}b_{}^{k}{}_{\mu }{}^{}+{\displaystyle \frac{a}{2}}l^kl_\mu {\displaystyle \frac{Q^2}{2}}m^km_\mu ,`$ (67)
$`A_\mu `$ $`=`$ $`{\displaystyle \frac{q}{4\pi }}\sqrt{Z}l_\mu ,`$ (68)
where $`{}_{}{}^{(0)}b_{}^{k}_\mu `$ are the dual components of constant vierbeins and are defined by $`{}_{}{}^{(0)}b_{}^{k}{}_{\mu }{}^{}\stackrel{\text{def}}{=}\delta ^k_\mu `$. The functions $`Z,l_\mu ,m_\mu ,l^k`$ and $`m^k`$ are given by
$`Z`$ $`=`$ $`{\displaystyle \frac{N}{D}},`$
$`l_0`$ $`=`$ $`\sqrt{Z},l_\alpha ={\displaystyle \frac{2\sqrt{Z}}{D+r^2+h^2}}\left[Nx_\alpha +{\displaystyle \frac{h^2x^3\delta ^3_\alpha }{N}}\epsilon _{\alpha \beta 3}x^\beta \right],`$
$`m_\mu `$ $`=`$ $`{\displaystyle \frac{l_\mu }{\sqrt{N}}},l^k=\delta ^k{}_{\mu }{}^{}\eta _{}^{\mu \nu }l_\nu ,m^k=\delta ^k{}_{\mu }{}^{}\eta _{}^{\mu \nu }m_\nu ,`$ (69)
where $`D`$ and $`N`$ are given by
$$D=\sqrt{(r^2h^2)^2+4h^2(x^3)^2},N=\frac{\sqrt{r^2h^2+D}}{\sqrt{2}},$$
(70)
with $`r\stackrel{\text{def}}{=}\sqrt{(x^1)^2+(x^2)^2+(x^3)^2}`$. In Eq. (2.2.3), the $`\epsilon _{\alpha \beta \gamma }`$ are three-dimensional totally anti-symmetric tensor with $`\epsilon _{123}=1`$. Also, we have defined
$$a\stackrel{\text{def}}{=}\frac{\kappa m}{4\pi },Q\stackrel{\text{def}}{=}\frac{q}{4\pi }\sqrt{\frac{\kappa }{2}},h\stackrel{\text{def}}{=}\frac{J}{m},$$
(71)
with $`m,J`$ and $`q`$ being the active gravitational mass, the absolute value of the angular momentum, and the electromagnetic charge of the source body, respectively.
For the solution given by Eqs. (67) and (68), the axial vector part $`a_\mu `$ of the torsion tensor vanishes,
$$a_\mu =0,$$
(72)
and the metric is identical to the charged Kerr metric in G. R.
The asymptotic forms of the vierbeins and the electromagnetic vector potential for large $`r`$ are given by
$`b^a_\alpha `$ $`=`$ $`\delta ^a{}_{\alpha }{}^{}+{\displaystyle \frac{1}{2}}(a{\displaystyle \frac{Q^2}{r}}){\displaystyle \frac{n^an_\alpha }{r}}{\displaystyle \frac{ahn^\beta }{2r^2}}(\epsilon _{\alpha \beta 3}n^a+\epsilon ^a{}_{\beta 3}{}^{}n_{\alpha }^{})+O^a{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^3}}\right),`$
$`b^{(0)}_\alpha `$ $`=`$ $`{\displaystyle \frac{n_\alpha }{2r}}\left(a{\displaystyle \frac{Q^2}{r}}\right)+{\displaystyle \frac{ah}{2r^2}}\epsilon _{\alpha \beta 3}n^\beta +O_\alpha \left({\displaystyle \frac{1}{r^3}}\right),`$
$`b^a_0`$ $`=`$ $`{\displaystyle \frac{n^a}{2r}}\left(a{\displaystyle \frac{Q^2}{r}}\right){\displaystyle \frac{ah}{2r^2}}\epsilon ^a{}_{\beta 3}{}^{}n_{}^{\beta }+O^a\left({\displaystyle \frac{1}{r^3}}\right),`$
$`b^{(0)}_0`$ $`=`$ $`1{\displaystyle \frac{1}{2r}}\left(a{\displaystyle \frac{Q^2}{r}}\right)+O\left({\displaystyle \frac{1}{r^3}}\right)`$ (73)
and
$$A_0=\frac{q}{4\pi r}+O\left(\frac{1}{r^3}\right),A_\alpha =\frac{q}{4\pi }\frac{x^\alpha }{r^2}+O_\alpha \left(\frac{1}{r^3}\right),$$
(74)
respectively, which follow from Eqs. (67) and (68). Here, we have defined $`n^\alpha \stackrel{\text{def}}{=}x^\alpha /r`$, and the expression $`O\left(1/r^w\right)`$ with positive $`w`$ denotes a term such that $`lim_r\mathrm{}r^wO\left(1/r^w\right)=`$constant. <sup>13</sup><sup>13</sup>13The symbols $`a`$ and $`\alpha `$ in $`O^a{}_{\alpha }{}^{}(1/r^3),O_\alpha (1/r^3)`$ and $`O^a(1/r^3)`$ are to show that these terms have indices as indicated.
#### 2.2.4 Asymptotic form of $`\psi ^k`$
The space-time in this theory has vanishing curvature tensor. When, additionally, the torsion tensor vanishes identically, the space-time is the Minkowski space-time, for which the translational gauge potentials $`A^k_\mu `$ can be chosen to be zero and Eq. (3) is reduced to
$$b^k{}_{\mu }{}^{}=\psi ^k{}_{,\mu }{}^{}.$$
(75)
For the solution given by Eqs. (67) and (68), we have<sup>14</sup><sup>14</sup>14The square bracket $`[]`$ in the suffix of $`O^k{}_{[\mu \nu ]}{}^{}(1/r^2)`$ indicate that this term is anti-symmetric with respect to $`\mu ,\nu `$.
$$T^k{}_{\mu \nu }{}^{}=O^k{}_{[\mu \nu ]}{}^{}\left(\frac{1}{r^2}\right),$$
(76)
and the space-time asymptotically approaches the Minkowski space-time for large $`r`$.
The above discussion and the consideration given in Ref. to introduce vierbeins suggest that $`\psi ^k`$ can be regarded as a Minkowskian coordinate at spatial infinity, and we are naturally led to employ the following form of $`\psi ^k`$:
$`\psi ^k`$ $`=`$ $`{}_{}{}^{(0)}b_{}^{k}{}_{\mu }{}^{}x_{}^{\mu }+{}_{}{}^{(0)}\psi _{}^{k}+O^k\left({\displaystyle \frac{1}{r^\beta }}\right),(\beta >0)`$
$`\psi ^k_{,\mu }`$ $`=`$ $`{}_{}{}^{(0)}b_{}^{k}{}_{\mu }{}^{}+O^k{}_{\mu }{}^{}\left({\displaystyle \frac{1}{r^{1+\beta }}}\right),`$
$`\psi ^k_{,\mu \nu }`$ $`=`$ $`O^k{}_{(\mu \nu )}{}^{}\left({\displaystyle \frac{1}{r^2}}\right),`$ (77)
where $`{}_{}{}^{(0)}\psi _{}^{k}`$ and $`\beta `$ are constants.
## 3 Equivalence relations for the case of the solution <br>represented by Eqs. (67) and (68)
In this section, on the basis of the discussion in Refs. and , we examine the energy-momentum, the angular momentum and the charge for the solution given in the preceding section.
### 3.1 The case in which $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables
We regard the Lagrangian density $`𝑳\stackrel{\text{def}}{=}\sqrt{g}L`$ as a function of $`\psi ^k,A^k{}_{\mu }{}^{},A_\mu `$ and of their derivatives. For this case, the generator $`M_k`$ of *internal* translations and the generator $`S_{kl}`$ of *internal* Lorentz transformations are the dynamical energy-momentum and the total (=spin $`+`$ orbital) angular momentum, respectively, and they are expressed as
$$M_k\stackrel{\text{def}}{=}_\sigma {}_{}{}^{\text{tot}}𝑻_{k}^{}{}_{}{}^{\mu }d\sigma _\mu $$
(78)
and
$$S_{kl}\stackrel{\text{def}}{=}_\sigma {}_{}{}^{\text{tot}}𝑺_{kl}^{}{}_{}{}^{\mu }d\sigma _\mu .$$
(79)
Here, $`\sigma `$ is a space-like surface, and $`d\sigma _\mu `$ is the surface element on it:
$$d\sigma _\mu =\frac{1}{3!}\epsilon _{\mu \nu \lambda \rho }dx^\nu dx^\lambda dx^\rho .$$
(80)
We have
$$𝑭_k{}_{}{}^{\mu \nu }=𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu }+𝑭^{(2)}{}_{k}{}^{}{}_{}{}^{\mu \nu },$$
(81)
with
$`𝑭^{(1)}{}_{k}{}^{}^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{1}{\kappa \sqrt{g}}}b_{k\rho }{}_{\sigma }{}^{}\left\{(g)g^{\rho [\mu }g^{\nu ]\sigma }\right\}+𝒁_k{}_{}{}^{\mu \nu },`$ (82)
$`𝑭^{(2)}{}_{k}{}^{}^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{3}{2\lambda }}\sqrt{g}\epsilon _k{}_{}{}^{lmn}b_{}^{\mu }{}_{m}{}^{}b_{}^{\nu }{}_{n}{}^{}a_{l}^{},`$ (83)
$`𝒁_k^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{\sqrt{g}}{\kappa }}\{b^{[\mu }{}_{k}{}^{}(b^{\nu ]l}b^\lambda {}_{l.\lambda }{}^{}b^{\lambda l}b^{\nu ]}{}_{l,\lambda }{}^{})+b^\lambda {}_{k}{}^{}b_{}^{[\mu }{}_{l}{}^{}b_{}^{\nu ]l}{}_{,\lambda }{}^{}\},`$ (84)
as is known by using the explicit form of $`L_G`$. From Eqs. (17), (24), (28) and (81), we find that $`{}_{}{}^{\text{tot}}𝑺_{kl}^{}^\mu `$ defined by Eq. (29) can be rewritten as
$`{}_{}{}^{\text{tot}}𝑺_{kl}^{}^\mu `$ $`=`$ $`2_\nu (\psi _{[k}𝑭_{l]}{}_{}{}^{\mu \nu })+{\displaystyle \frac{2}{\kappa }}_\nu (\sqrt{g}b^\mu {}_{[k}{}^{}b_{}^{\nu }{}_{l]}{}^{}{}_{}{}^{(0)}b_{}^{\mu }{}_{[k}{}^{}{}_{}{}^{(0)}b_{}^{\nu }{}_{l]}{}^{})`$ (85)
$`b_{[k\nu }𝑭^{(2)}{}_{l]}{}^{}{}_{}{}^{\mu \nu }+2\psi _{[k}{\displaystyle \frac{\delta 𝑳}{\delta A^{l]}_\mu }},`$
where $`{}_{}{}^{(0)}b_{}^{\mu }_k`$ are the components of the constant vierbeins: $`{}_{}{}^{(0)}b_{}^{\mu }{}_{k}{}^{}\stackrel{\text{def}}{=}\delta ^\mu _k`$. For the vierbeins given by Eq. (67), we have
$$𝒁_k{}_{}{}^{\mu \nu }=0,𝑭^{(2)}{}_{k}{}^{}{}_{}{}^{\mu \nu }=0,$$
(86)
and hence
$$𝑭_k{}_{}{}^{\mu \nu }=𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu }=\frac{1}{\kappa \sqrt{g}}b_{k\rho }_\sigma \left\{(g)g^{\rho [\mu }g^{\nu ]\sigma }\right\}.$$
(87)
#### 3.1.1 Energy-momentum
When the field equation $`\delta 𝑳/\delta A^k{}_{\mu }{}^{}=0`$ is satisfied, Eq. (78) can be rewritten as
$$M_k=_\sigma _\nu 𝑭_k{}_{}{}^{\mu \nu }d\sigma _\mu =_S𝑭_k{}_{}{}^{0\alpha }r_{}^{2}n_\alpha 𝑑\mathrm{\Omega },$$
(88)
by using the identity (24). Here, $`S`$ and $`d\mathrm{\Omega }`$ stand for the two-dimensional surface of $`\sigma `$ and the differential solid angle, respectively. Equations (81), (86) – (88) and (A) give
$$M_{(0)}=m,M_a=0.$$
(89)
The quantity $`M_k`$ is the total energy-momentum vector of the system. The first relation in Eq. (89) expresses the equality of the active gravitational mass and the inertial mass.
#### 3.1.2 Angular momentum
From Eqs. (79) and (85), the toal angular momentum can be expressed as
$$S_{kl}=2_\sigma _\nu \left[\psi _{[k}𝑭_{l]}{}_{}{}^{\mu \nu }+\frac{1}{\kappa }\{\sqrt{g}b^\mu {}_{[k}{}^{}b_{}^{\nu }{}_{l]}{}^{}{}_{}{}^{(0)}b_{}^{\mu }{}_{[k}{}^{}{}_{}{}^{(0)}b_{}^{\nu }{}_{l]}{}^{}\}\right]d\sigma _\mu ,$$
(90)
from which the expression
$`S_{(0)a}`$ $`=`$ $`{}_{}{}^{(0)}\psi _{a}^{}m={}_{}{}^{(0)}\psi _{a}^{}M_{(0)}+{}_{}{}^{(0)}\psi _{(0)}^{}M_a,`$
$`S_{ab}`$ $`=`$ $`J\epsilon _{ab3}=J\epsilon _{ab3}+{}_{}{}^{(0)}\psi _{a}^{}M_b{}_{}{}^{(0)}\psi _{b}^{}M_a`$ (91)
is obtained by the use of Eqs. (2.2.3), (2.2.4), (78), (81), (87) and (A).
In the above, terms of the form $`{}_{}{}^{(0)}\psi _{k}^{}M_l{}_{}{}^{(0)}\psi _{l}^{}M_k`$ are regarded as to represent the conserved orbital angular momentum around the origin of the internal space. Equation (3.1.2) implies that the total angular momentum is equal to the angular momentum of the rotating source.
#### 3.1.3 Canonical energy-momentum and “extended orbital angular momentum”
The generator $`M^c_\mu `$ of coordinate translations and the generator $`L_\mu ^\nu `$ of $`GL(4,𝑹)`$ coordinate transformations are the canonical energy-momentum and the “extended orbital angular momentum”, <sup>15</sup><sup>15</sup>15Note that the anti-symmetric part $`L_{[\mu \nu ]}\stackrel{\text{def}}{=}\eta _{[\nu \lambda }L_{\mu ]}^\lambda `$ is the orbital angular momentum. respectively. They have the expressions
$`M^c_\mu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }\stackrel{~}{𝑻}_\mu {}_{}{}^{\nu }d\sigma _\nu ={\displaystyle _\sigma }_\tau 𝚿_\mu {}_{}{}^{\nu \tau }d\sigma _\nu ,`$ (92)
$`L_\mu ^\nu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }𝑴_\mu {}_{}{}^{\nu \lambda }d\sigma _\lambda =2{\displaystyle _\sigma }_\tau (x^\nu 𝚿_\mu {}_{}{}^{\lambda \tau })d\sigma _\lambda .`$ (93)
The asymptotic behavior of the translational gauge potentials $`A^k_\mu `$ at spatial infinity is given as
$`A^{(0)}_0`$ $`=`$ $`{\displaystyle \frac{a}{2r}}+O\left({\displaystyle \frac{1}{r^{1+\beta }}}\right),A^{(0)}{}_{\alpha }{}^{}={\displaystyle \frac{a}{2r}}n_\alpha +O_\alpha \left({\displaystyle \frac{1}{r^{1+\beta }}}\right),`$
$`A^a_0`$ $`=`$ $`{\displaystyle \frac{a}{2r}}n^a+O^a\left({\displaystyle \frac{1}{r^{1+\beta }}}\right),A^a{}_{\alpha }{}^{}={\displaystyle \frac{a}{2r}}n^an_\alpha +O^a{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^{1+\beta }}}\right),`$ (94)
which are known from Eqs. (17), (2.2.3) and (2.2.4). Then $`M^c_\mu `$ vanishes trivially,
$$M^c{}_{\mu }{}^{}=_S𝑭_k{}_{}{}^{0\alpha }A_{}^{k}{}_{\mu }{}^{}r_{}^{2}n_\alpha d\mathrm{\Omega }=0,$$
(95)
while $`L_\mu ^\nu `$ is expressed as
$$L_\mu {}_{}{}^{\nu }=2_Sx^\nu (𝑭_k{}_{}{}^{0\alpha }A_{}^{k}{}_{\mu }{}^{}+𝑭^{0\alpha }A_\mu )r^2n_\alpha d\mathrm{\Omega },$$
(96)
and the non-zero components are given by
$$L_1{}_{}{}^{1}=L_2{}_{}{}^{2}=L_3{}_{}{}^{3}=\frac{q^2}{6\pi }.$$
(97)
These can be shown by use of Eqs. (36), (74), (81), (3.1.3) and (A). Thus, the orbital angular momentum $`L_{[\mu \nu ]}`$ is vanishing: $`L_{[\mu \nu ]}=0`$.
#### 3.1.4 Charge
The charge is defined as the generator of $`U(1)`$ gauge transformations and is given by
$$C\stackrel{\text{def}}{=}_\sigma _\nu \left(\frac{𝑳}{A_{\mu ,\nu }}\right)d\sigma _\mu =_\sigma _\nu (\sqrt{g}F^{\mu \nu })d\sigma _\mu =_S(^\alpha A^0)n_\alpha r^2𝑑\mathrm{\Omega }=q,$$
(98)
where we have used Eq. (74). This implies the equality of the total charge of the system and the charge of the source.
### 3.2 The case in which $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables
Let us denote $`𝑳`$ and $`L`$, expressed as functions of $`\psi ^k,b^k{}_{\mu }{}^{},A_\mu `$ and their derivatives, by $`\widehat{𝑳}`$ and $`\widehat{L}`$, respectively. The action $`𝑰`$ is now written as
$$\widehat{𝑰}=_𝒟\widehat{L}𝑑v=𝑰.$$
(99)
Various identities can be derived from the requirements (R.i) and (R.ii), among which we have
$`{\displaystyle \frac{\delta \widehat{𝑳}}{\delta \psi ^k}}`$ $``$ $`0,`$ (100)
$`{}_{}{}^{\text{tot}}\widehat{𝑻}_{k}^{}^\mu `$ $``$ $`0,`$ (101)
$`_\mu {}_{}{}^{\text{tot}}\widehat{𝑺}_{kl}^{}{}_{}{}^{\mu }2{\displaystyle \frac{\delta \widehat{𝑳}}{\delta \psi ^{[k}}}\psi _{l]}2{\displaystyle \frac{\delta \widehat{𝑳}}{\delta b^{[k}_\mu }}b_{l]\mu }`$ $``$ $`0,`$ (102)
$`\widehat{\stackrel{~}{𝑻}}_\mu {}_{}{}^{\nu }_\lambda \widehat{𝚿}_\mu {}_{}{}^{\nu \lambda }{\displaystyle \frac{\delta \widehat{𝑳}}{\delta b^k_\nu }}b^k{}_{\mu }{}^{}{\displaystyle \frac{\delta \widehat{𝑳}}{\delta A_\nu }}A_\mu `$ $``$ $`0,`$ (103)
where we have defined
$`\widehat{𝑳}`$ $`\stackrel{\text{def}}{=}`$ $`\sqrt{g}\widehat{L},`$ (104)
$`{}_{}{}^{\text{tot}}\widehat{𝑻}_{k}^{}^\mu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{\widehat{𝑳}}{\psi ^k_{,\mu }}},`$ (105)
$`{}_{}{}^{\text{tot}}\widehat{𝑺}_{kl}^{}^\mu `$ $`\stackrel{\text{def}}{=}`$ $`2{\displaystyle \frac{\widehat{𝑳}}{\psi ^{[k}_{,\mu }}}\psi _{l]}2\widehat{𝑭}_{[k}{}_{}{}^{\nu \mu }b_{l]\nu }^{},`$ (106)
$`\widehat{\stackrel{~}{𝑻}}_\mu ^\nu `$ $`\stackrel{\text{def}}{=}`$ $`\delta _\mu {}_{}{}^{\nu }\widehat{𝑳}\widehat{𝑭}_k{}_{}{}^{\lambda \nu }b_{}^{k}{}_{\lambda ,\mu }{}^{}\widehat{𝑭}^{\lambda \nu }A_{\lambda ,\mu }{\displaystyle \frac{\widehat{𝑳}}{\psi ^k_{,\nu }}}\psi ^k{}_{,\mu }{}^{},`$ (107)
$`\widehat{𝚿}_\lambda ^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`\widehat{𝑭}_k{}_{}{}^{\mu \nu }b_{}^{k}{}_{\lambda }{}^{}+\widehat{𝑭}^{\mu \nu }A_\lambda =\widehat{𝚿}_\lambda {}_{}{}^{\nu \mu },`$ (108)
$`\widehat{𝑭}_k^{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle \frac{\widehat{𝑳}}{b^k_{\mu ,\nu }}}=𝑭_k{}_{}{}^{\mu \nu },\widehat{𝑭}^{\mu \nu }\stackrel{\text{def}}{=}{\displaystyle \frac{\widehat{𝑳}}{A_{\mu ,\nu }}}=𝑭^{\mu \nu }.`$ (109)
From Eqs. (100) and (101), we see that
$$_\mu {}_{}{}^{\text{tot}}\widehat{𝑺}_{kl}^{}{}_{}{}^{\mu }=0$$
(110)
when the field equations for $`b^k_\mu `$ are satisfied. From Eqs. (103) and (108), it follows that
$`_\nu \widehat{\stackrel{~}{𝑻}}_\mu ^\nu `$ $`=`$ $`0,`$ (111)
$`_\nu \widehat{𝑴}_\lambda ^{\mu \nu }`$ $`=`$ $`0`$ (112)
when the field equations $`\delta \widehat{𝑳}/\delta b^k{}_{\mu }{}^{}=0`$ and $`\delta \widehat{𝑳}/\delta A_\mu =0`$ are both satisfied, where $`\widehat{𝑴}_\lambda {}_{}{}^{\mu \nu }\stackrel{\text{def}}{=}2(\widehat{𝚿}_\lambda {}_{}{}^{\mu \nu }x^\mu \widehat{\stackrel{~}{𝑻}}_\lambda {}_{}{}^{\nu })`$. Equations (110) – (112) are the differential conservation laws of the “spin” angular momentum, the canonical angular momentum, and the “extended orbital angular momentum”, respectively.
The density $`{}_{}{}^{\text{tot}}\widehat{𝑺}_{kl}^{}^\mu `$ defined by Eq. (106) can be rewritten as
$`{}_{}{}^{\text{tot}}\widehat{𝑺}_{kl}^{}^\mu `$ $`=`$ $`{\displaystyle \frac{2}{\kappa }}_\nu (\sqrt{g}b^\mu {}_{[k}{}^{}b_{}^{\nu }{}_{l]}{}^{}{}_{}{}^{(0)}b_{}^{\mu }{}_{[k}{}^{}{}_{}{}^{(0)}b_{}^{\nu }{}_{l]}{}^{})b_{[k\nu }𝑭^{(2)}{}_{l]}{}^{}{}_{}{}^{\mu \nu },`$ (113)
by the use of Eqs. (81), (101), (109) and (105).
#### 3.2.1 Energy-momentum
The dynamical energy-momentum $`\widehat{M}_k`$, which is the generator of internal translations, vanishes identically:
$$\widehat{M}_k\stackrel{\text{def}}{=}_\sigma ^{\text{tot}}\widehat{𝑻}_k{}_{}{}^{\mu }d\sigma _\mu 0.$$
(114)
This is evident from Eq. (101).
#### 3.2.2 Spin angular momentum
The generator $`\widehat{S}_{kl}`$ of internal Lorentz transformations is expressed as
$$\widehat{S}_{kl}\stackrel{\text{def}}{=}_\sigma {}_{}{}^{\text{tot}}\widehat{𝑺}{}_{kl}{}^{}{}_{}{}^{\mu }d\sigma _\mu =\frac{2}{\kappa }_S(\sqrt{g}b^0{}_{[k}{}^{}b_{}^{\alpha }{}_{l]}{}^{}{}_{}{}^{(0)}b_{}^{0}{}_{[k}{}^{}{}_{}{}^{(0)}b_{}^{\alpha }{}_{l]}{}^{})n_\alpha r^2d\mathrm{\Omega },$$
(115)
as can be shown by using Eq. (113), and we obtain
$$\widehat{S}_{(0)a}=0,\widehat{S}_{ab}=\frac{1}{3}J\epsilon _{ab3}$$
(116)
by using Eq. (2.2.3).
#### 3.2.3 Canonical energy-momentum and “extended orbital angular momentum”
The generator $`\widehat{M}^c_\mu `$ of coordinate translations and the generator $`\widehat{L}_\mu ^\nu `$ of $`GL(4,𝑹)`$ coordinate transformations are the canonical energy-momentum and the “extended orbital angular momentum”, respectively. They have the expressions
$`\widehat{M}^c_\mu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }\widehat{\stackrel{~}{𝑻}}_\mu {}_{}{}^{\nu }\sigma _{\nu }^{}={\displaystyle _\sigma }_\tau \widehat{𝚿}_\mu {}_{}{}^{\nu \tau }d\sigma _\nu ,`$ (117)
$`\widehat{L}_\mu ^\nu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }\widehat{𝑴}_\mu {}_{}{}^{\nu \lambda }d\sigma _\lambda =2{\displaystyle _\sigma }_\tau (x^\nu \widehat{𝚿}_\mu {}_{}{}^{\lambda \tau })d\sigma _\lambda .`$ (118)
Then we have
$$\widehat{M}^c{}_{0}{}^{}=m,\widehat{M}^c{}_{\alpha }{}^{}=0.$$
(119)
Thus, $`\widehat{M}^c_\mu `$ is the total energy-momentum, and the equality of the active gravitational mass and the inertial mass holds. Also, $`\widehat{L}_\mu ^\nu `$ is given by
$`\widehat{L}_0^0`$ $`=`$ $`2x^0m,\widehat{L}_0{}_{}{}^{\alpha }=0,\widehat{L}_\alpha {}_{}{}^{0}=0,`$
$`\widehat{L}_\alpha ^\beta `$ $`=`$ $`{\displaystyle \frac{2}{3}}J\epsilon _\alpha {}_{}{}^{\beta 3},(\alpha \beta ),\widehat{L}_1{}_{}{}^{1}=\widehat{L}_2{}_{}{}^{2}=\widehat{L}_3{}_{}{}^{3}=\mathrm{}.`$ (120)
Equations (119) and (3.2.3) are obtained by using Eqs. (2.2.3), (74), (87), (108), (109) and (A).
The orbital angular momentum $`\widehat{L}_{[\mu \nu ]}`$ is given by
$$\widehat{L}_{[0\alpha ]}=0,\widehat{L}_{[\alpha \beta ]}=\frac{2}{3}J\epsilon _{\alpha \beta 3}.$$
(121)
If we define the total angular momentum of the system by
$$\widehat{J}_{kl}\stackrel{\text{def}}{=}\widehat{S}_{kl}+{}_{}{}^{(0)}b_{}^{\mu }{}_{k}{}^{}{}_{}{}^{(0)}b_{}^{\nu }{}_{l}{}^{}\widehat{L}{}_{[\mu \nu ]}{}^{},$$
(122)
then we have
$$\widehat{J}_{(0)k}=0,\widehat{J}_{ab}=J\epsilon _{ab3},$$
(123)
and the total angular momentum is equal to the angular momentum of the rotating source.
#### 3.2.4 Charge
The generator $`\widehat{C}`$ of $`U(1)`$ gauge transformations is given by
$$\widehat{C}\stackrel{\text{def}}{=}_\sigma _\nu \left(\frac{\widehat{𝑳}}{A_{\mu ,\nu }}\right)d\sigma _\mu =q,$$
(124)
which implies the equality of the total charge of the system and the charge of the source.
## 4 Restrictions imposed on field variables <br>by the generalized equivalence principle
In the preceding section, we examined the solution given by Eqs. (67) and (68), and the results show that the total energy-momentum, the total angular momentum, and the total charge of the system, which are generators of transformations, are equal to the corresponding active quantities of a central gravitating body. The total mass, which is equal to the total energy divided by the square of the velocity of light, can be regarded as the inertial mass of the system. Thus, the results include the equality of the inertial mass and the active gravitational mass, which implies that the equivalence principle is satisfied by this solution.
In view of the above, we regard, the total momentum, the total angular momentum and the total charge as “inertial” quantities, and we say that a generalized equivalence principle (G. E. P.) is satisfied if the total energy-momentum, total angular momentum and total charge are all equal to the corresponding quantities of the source.
The axial vector part $`a_\mu `$ vanishes for our solution, as stated above, and the field equations (57) – (59) are covariant under general coordinate transformations and under local Lorentz transformations that keep $`a_\mu `$ vanishing. Thus, new solutions can be obtained by applying the general coordinate transformations and restricted local Lorentz transformations $`b^k{}_{\mu }{}^{}=A^k{}_{l}{}^{}(x)b^l_\mu `$ that satisfy the condition <sup>,</sup><sup>16</sup><sup>16</sup>16Note that $`a_\mu `$ is invariant under the local Lorentz transformation $`A^k{}_{l}{}^{}(x)`$ if and only if this condition is satisfied.
$$\epsilon ^{\mu \nu \rho \sigma }b^k{}_{\nu }{}^{}b_{}^{l}{}_{\rho }{}^{}A_{}^{m}{}_{k}{}^{}(x)A_{ml}(x){}_{,\sigma }{}^{}=0$$
(125)
to the solution represented by Eqs. (67) and (68).
In this section, we examine restrictions imposed on solutions by the requirement that this G. E. P. is satisfied. We look for new solutions having suitable asymptotic behavior by considering the following $`\overline{\text{Poincaré}}`$ gauge transformations:
(1) Local $`SL(2.C)`$ transformation
$$H^k{}_{l}{}^{}\stackrel{\text{def}}{=}\left(\mathrm{\Lambda }(a^1)\right)^k{}_{l}{}^{}=\mathrm{\Lambda }^k{}_{l}{}^{}+\mathrm{\Lambda }^k{}_{m}{}^{}\omega _{}^{m}{}_{l}{}^{}(x).$$
(126)
(2) Local internal translation
$$t^k={}_{}{}^{(0)}t_{}^{k}+b^k(x).$$
(127)
Here, $`\mathrm{\Lambda }^k_l`$ and $`{}_{}{}^{(0)}t_{}^{k}`$ denote a constant internal Lorentz transformation and the constant internal translation, respectively, and $`\omega ^k_l`$ and $`b^k`$ are functions satisfying the following conditions:
$`\omega ^k{}_{l}{}^{}(x)`$ $`=`$ $`O^k{}_{l}{}^{}\left({\displaystyle \frac{1}{r^p}}\right),\omega ^k{}_{l,\mu }{}^{}(x)=O^k{}_{l\mu }{}^{}\left({\displaystyle \frac{1}{r^{p+1}}}\right),(p>0)`$ (128)
$`b^k(x)`$ $`=`$ $`O^k\left({\displaystyle \frac{1}{r^\gamma }}\right),b^k{}_{,\mu }{}^{}(x)=O^k{}_{\mu }{}^{}\left({\displaystyle \frac{1}{r^{\gamma +1}}}\right).(\gamma >0)`$ (129)
The transformation $`H^k{}_{l}{}^{}\stackrel{\text{def}}{=}\left(\mathrm{\Lambda }(a^1)\right)^k_l`$ is a Lorentz transformation satisfying Eq. (125), if and only if
$`\omega _{kl}+\omega _{lk}+\omega _{mk}\omega ^m{}_{l}{}^{}=0,`$ (130)
$`\epsilon ^{\mu \nu \lambda \rho }(\omega _{kl,\rho }+\omega _{mk}\omega ^m{}_{l,\rho }{}^{})b^k{}_{\nu }{}^{}b_{}^{l}{}_{\lambda }{}^{}=0`$ (131)
are both satisfied. The conditions (130) and (131) are equivalent to
$$\omega _{\mu \nu }+\omega _{\nu \mu }+\omega _{\lambda \mu }\omega ^\lambda {}_{\nu }{}^{}=0$$
(132)
and
$$X_{\mu \nu \lambda }+X_{\lambda \mu \nu }+X_{\nu \lambda \mu }=0,$$
(133)
respectively, where we have defined
$`\omega _{\mu \nu }`$ $`\stackrel{\text{def}}{=}`$ $`b^k{}_{\mu }{}^{}b_{}^{l}{}_{\nu }{}^{}\omega _{kl}^{},`$ (134)
$`X_{\mu \nu \lambda }`$ $`\stackrel{\text{def}}{=}`$ $`\omega _{\mu \nu ,\lambda }+\omega ^\tau {}_{\mu }{}^{}\omega _{\tau \nu ,\lambda }^{}b^\tau {}_{k}{}^{}b_{}^{k}{}_{\mu ,\lambda }{}^{}\omega _{\tau \nu }^{}b^\tau {}_{k}{}^{}b_{}^{k}{}_{\nu ,\lambda }{}^{}\omega _{\mu \tau }^{}`$ (135)
$`+b^{\tau k}b^\rho {}_{k,\lambda }{}^{}\omega _{\tau \mu }^{}\omega _{\rho \nu }+b^k{}_{\nu }{}^{}b_{}^{\tau }{}_{k,\lambda }{}^{}\omega _{}^{\rho }{}_{\mu }{}^{}\omega _{\rho \tau }^{}.`$
The function $`X_{\mu \nu \lambda }`$ is anti-symmetric with respect to the first two indices:
$$X_{\mu \nu \lambda }=X_{\nu \mu \lambda }.$$
(136)
From Eqs. (128), (132) and (133), $`\omega _{\mu \nu }`$ is known to have the expression
$`\omega _{\mu \nu }`$ $`=`$ $`_\mu \omega _\nu _\nu \omega _\mu +f_{\mu \nu }(x),`$ (137)
with
$`_\mu \omega _\nu _\nu \omega _\mu `$ $`=`$ $`O_{[\mu \nu ]}\left({\displaystyle \frac{1}{r^p}}\right),`$
$`f_{\mu \nu }(x)`$ $`=`$ $`O_{\mu \nu }\left({\displaystyle \frac{1}{r^s}}\right).(p<s)`$ (138)
In addition, we require the leading term of $`\omega _{kl}`$ at spatial infinity to be spherically symmetric. Then we can write
$$\omega ^0=A(r,x^0),\omega ^\alpha =n^\alpha B(r,x^0),$$
(139)
with certain functions $`A`$ and $`B`$ of $`r`$ and $`x^0`$.
Also, we consider the following coordinate transformation:
$`x^\mu `$ $`=`$ $`C^\mu {}_{\nu }{}^{}x_{}^{\nu }+D^\mu (x),`$
$`{\displaystyle \frac{x^\mu }{x^\nu }}`$ $`=`$ $`C^\mu {}_{\nu }{}^{}+a^\mu {}_{\nu }{}^{}(x),`$
$`a^\mu _\nu `$ $`\stackrel{\text{def}}{=}`$ $`D^\mu {}_{,\nu }{}^{},`$
$`a^\mu {}_{\nu }{}^{}(x)`$ $`=`$ $`O^\mu {}_{\nu }{}^{}\left({\displaystyle \frac{1}{r^u}}\right),a^\mu {}_{\nu ,\lambda }{}^{}=O^\mu {}_{\nu \lambda }{}^{}\left({\displaystyle \frac{1}{r^{u+1}}}\right),(u>0)`$ (140)
where $`C^\mu _\nu `$ denotes a constant Lorentz transformation, and $`D^\mu (x)`$ satisfies the condition
$$\underset{r\mathrm{}}{lim}\frac{D^\mu (x)}{r}=0.$$
(141)
We write $`x^\mu /x^\nu `$ as
$$\frac{x^\mu }{x^\nu }=\left(C^1\right)^\mu {}_{\nu }{}^{}+d^\mu {}_{\nu }{}^{}(x),$$
(142)
with $`(C^1)^\mu _\nu `$ being constants satisfying $`(C^1)^\mu {}_{\lambda }{}^{}C_{}^{\lambda }{}_{\nu }{}^{}=\delta ^\mu _\nu `$.
The vierbeins and vector potentials given by
$$b^k{}_{\mu }{}^{}\stackrel{\text{def}}{=}H^k{}_{l}{}^{}\frac{x^\nu }{x^\mu }b^l{}_{\nu }{}^{},A^{}{}_{\mu }{}^{}\stackrel{\text{def}}{=}\frac{x^\nu }{x^\mu }A_\nu ,$$
(143)
with $`b^k_\mu `$ and $`A_\mu `$ given by Eqs. (67) and (68), are solutions of the gravitational and electromagnetic field equations. This is true irrespective of the values of the parameters $`p,s,\beta ,\gamma `$ and $`u`$. The G. E. P. is considered to be satisfied if the energy-momentum, the angular momentum and the charge all have correct transformation properties as their indices indicate. But, this is not necessarily the case for arbitrary values of these parameters; i.e. there are solutions which do not satisfy the G. E. P. We examine restrictions imposed on these parameters by the requirement that the G. E. P. is satisfied.
### 4.1 The case in which $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables
Under the combined transformation of the $`\overline{\text{Poincaré}}`$ gauge transformation given by Eqs. (126) and (127) and satisfying the conditions (130) and (131) and of the coordinate transformation (4), $`𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu },𝑭^{(2)}{}_{k}{}^{}^{\mu \nu }`$ and $`𝑭^{\mu \nu }`$ transform according to
$`𝑭^{(1)}{}_{k}{}^{}^{\mu \nu }`$ $`=`$ $`\mathrm{\Delta }{\displaystyle \frac{x^\mu }{x^\rho }}{\displaystyle \frac{x^\nu }{x^\sigma }}H_k{}_{}{}^{l}𝑭_{}^{(1)}{}_{l}{}^{}^{\rho \sigma }`$ (144)
$`+{\displaystyle \frac{\mathrm{\Delta }}{\kappa }}U_k{}_{}{}^{lmn}{}_{\lambda }{}^{}W_{}^{\rho \sigma \lambda }{}_{lmn}{}^{}{\displaystyle \frac{x^\mu }{x^\rho }}{\displaystyle \frac{x^\nu }{x^\sigma }},`$
$`𝑭^{(2)}{}_{k}{}^{}^{\mu \nu }`$ $`=`$ $`0=\mathrm{\Delta }{\displaystyle \frac{x^\mu }{x^\rho }}{\displaystyle \frac{x^\nu }{x^\sigma }}H_k{}_{}{}^{l}𝑭_{}^{(2)}{}_{l}{}^{}{}_{}{}^{\rho \sigma },`$ (145)
$`𝑭^{\mu \nu }`$ $`=`$ $`\mathrm{\Delta }{\displaystyle \frac{x^\mu }{x^\rho }}{\displaystyle \frac{x^\nu }{x^\sigma }}𝑭^{\rho \sigma },`$ (146)
where we have defined<sup>17</sup><sup>17</sup>17For simplicity, we restrict our consideration to the case in which $`\mathrm{\Delta }>0`$. An extension to the case of arbitrary non-vanishing $`\mathrm{\Delta }`$ can be made without difficulty.
$`U_k{}_{}{}^{lmn}_\lambda `$ $`\stackrel{\text{def}}{=}`$ $`H_k{}_{}{}^{l}V_{}^{mn}{}_{\lambda }{}^{},`$ (147)
$`V^{mn}_\lambda `$ $`\stackrel{\text{def}}{=}`$ $`H^{lm}H_l{}_{}{}^{n}{}_{,\lambda }{}^{}=V^{nm}{}_{\lambda }{}^{},`$ (148)
$`W^{\mu \nu \lambda }_{klm}`$ $`\stackrel{\text{def}}{=}`$ $`b^{[\mu }{}_{k}{}^{}b_{}^{\nu ]}{}_{l}{}^{}b_{}^{\lambda }{}_{m}{}^{}+b^{[\mu }{}_{m}{}^{}b_{}^{\nu ]}{}_{k}{}^{}b_{}^{\lambda }{}_{l}{}^{}+b^{[\mu }{}_{l}{}^{}b_{}^{\nu ]}{}_{m}{}^{}b_{}^{\lambda }{}_{k}{}^{},`$ (149)
$`\mathrm{\Delta }`$ $`\stackrel{\text{def}}{=}`$ $`det\left({\displaystyle \frac{x^\mu }{x^\nu }}\right).`$ (150)
Equations (144) – (146) show that $`𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu },𝑭^{(2)}{}_{k}{}^{}^{\mu \nu }`$ and $`𝑭^{\mu \nu }`$ transform as tensor densities under coordinate transformations. The function $`W^{\mu \nu \lambda }_{klm}`$ is totally anti-symmetric, both in upper indices and in lower indices.
#### 4.1.1 Energy-momentum
From Eqs. (88) and (144), $`M_k`$ is found to transform as<sup>18</sup><sup>18</sup>18$`A{}_{}{}^{[\mu \nu \lambda ]}\stackrel{\text{def}}{=}\frac{1}{3}\{A{}_{}{}^{\mu [\nu \lambda ]}+A{}_{}{}^{\nu [\lambda \mu ]}+A{}_{}{}^{\lambda [\mu \nu ]}\}`$.
$$M^{}{}_{k}{}^{}\stackrel{\text{def}}{=}_\sigma {}_{}{}^{\text{tot}}𝑻_{}^{}{}_{k}{}^{}{}_{}{}^{\mu }d\sigma ^{}{}_{\mu }{}^{}=_\sigma ^{}{}_{\nu }{}^{}𝑭_{}^{}{}_{k}{}^{}{}_{}{}^{\mu \nu }d\sigma ^{}{}_{\mu }{}^{}=\mathrm{\Lambda }_k{}_{}{}^{l}M_{l}^{}+\frac{3}{\kappa }U_k{}_{}{}^{[0\alpha \beta ]}{}_{\beta }{}^{}n_{\alpha }^{}r^2d\mathrm{\Omega },$$
(151)
where we represent $`^{}{}_{\mu }{}^{}=/x^\mu .`$ The energy-momentum $`M_k`$ obeys the correct transformation rule
$$M^{}{}_{k}{}^{}=\mathrm{\Lambda }_k{}_{}{}^{l}M_{l}^{}$$
(152)
if the condition
$$p>\frac{1}{2},s>1$$
(153)
is satisfied.<sup>19</sup><sup>19</sup>19For derivations of the conditions (153), (156) – (4.1.2), (163), (171) and (175), elementary but rather tedious calculations are needed. In Appendix A, we give lists of asymptotic forms of $`𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu },V^{mn}_\lambda `$ and $`W^{\mu \nu \lambda }_{klm}`$ for large $`r`$, which are useful in calculations.
#### 4.1.2 Angular momentum
From Eqs. (90) and (144), we find that $`S_{kl}`$ transforms as
$`S^{}_{kl}`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }{}_{}{}^{\text{tot}}𝑺_{}^{}{}_{kl}{}^{}{}_{}{}^{\mu }d\sigma ^{}_\mu `$
$`=`$ $`2{\displaystyle _\sigma }^{}{}_{\nu }{}^{}\left[\psi ^{}{}_{[k}{}^{}𝑭_{}^{}{}_{l]}{}^{}{}_{}{}^{\mu \nu }+{\displaystyle \frac{1}{\kappa }}(\sqrt{g^{}}b^\mu {}_{[k}{}^{}b_{}^{\nu }{}_{l]}{}^{}b^{(0)\mu }{}_{[k}{}^{}b_{}^{(0)\nu }{}_{l]}{}^{})\right]d\sigma ^{}_\mu `$
$`=`$ $`\mathrm{\Lambda }_k{}_{}{}^{m}\mathrm{\Lambda }_{l}^{}{}_{}{}^{n}(S_{mn}2{}_{}{}^{(0)}t_{[m}^{}M_{n]})`$
$`+2\mathrm{\Lambda }_k{\displaystyle {}_{}{}^{m}}\{\omega _m{}_{}{}^{n}(\psi _nt_n)b_m\}H_l{}_{}{}^{n}𝑭_{n}^{}{}_{}{}^{0\alpha }n_{\alpha }^{}r^2d\mathrm{\Omega }`$
$`+{\displaystyle \frac{2}{\kappa }}{\displaystyle }H_{[k}{}_{}{}^{m}H_{l]}^{}{}_{}{}^{n}(\psi _mt_m)H^{ij}H_i{}_{}{}^{h}{}_{,\beta }{}^{}W_{}^{0\alpha \beta }{}_{njh}{}^{}n_{\alpha }^{}r^2d\mathrm{\Omega }`$
$`+{\displaystyle \frac{2}{\kappa }}{\displaystyle }H_{[k}{}_{}{}^{m}(H_{l]}{}_{}{}^{n}+\mathrm{\Lambda }_{l]}{}_{}{}^{m}\omega _{m}^{}{}_{}{}^{n})(b^0{}_{[m}{}^{}b_{}^{\alpha }{}_{n]}{}^{}{}_{}{}^{(0)}b_{}^{0}{}_{[m}{}^{}{}_{}{}^{(0)}b_{}^{\alpha }{}_{n]}{}^{})n_\alpha r^2d\mathrm{\Omega },`$
where we have defined $`{}_{}{}^{(0)}b_{}^{\mu }{}_{k}{}^{}\stackrel{\text{def}}{=}lim_r\mathrm{}H_k{}_{}{}^{l}(x^\mu /x^\nu )b^\nu {}_{l}{}^{}=\mathrm{\Lambda }_k{}_{}{}^{l}C_{}^{\mu }{}_{\nu }{}^{}{}_{}{}^{(0)}b_{}^{\nu }_l`$. The angular momentum obeys the correct transformation rule<sup>20</sup><sup>20</sup>20The term $`2{}_{}{}^{(0)}t_{[k}^{}M_{l]}`$ comes from the translation (127), and it is an additional orbital angular momentum around the origin of the *internal* space.
$$S^{}{}_{kl}{}^{}=\mathrm{\Lambda }_k{}_{}{}^{m}\mathrm{\Lambda }_{l}^{}{}_{}{}^{n}(S_{mn}2{}_{}{}^{(0)}t_{[m}^{}M_{n]})$$
(155)
if the following conditions are satisfied:
$$p>\frac{1}{2},s>2$$
(156)
and
$$\left[\{p+\beta >1,p+\gamma >1\}\text{or}\left\{O^{(0)}\left(\frac{1}{r^\beta }\right)=f(r,x^0),O^{(0)}\left(\frac{1}{r^\gamma }\right)=g(r,x^0)\right\}\right]$$
(157)
and
$$\left[\{p+\beta >1,p+\gamma >1\}\text{or}\left\{O^a\left(\frac{1}{r^\beta }\right)=n^ah(r,x^0),O^a\left(\frac{1}{r^\gamma }\right)=n^ak(r,x^0)\right\}\right],$$
with $`f,g,h`$ and $`k`$ being some functions of $`r`$ and $`x^0`$, where the terms $`O^k(1/r^\beta )`$ and $`O^k(1/r^\gamma )`$ are those in Eq. (2.2.4) and Eq. (129), respectively.
#### 4.1.3 Canonical energy-momentum and “extended orbital angular momentum”
The transformed canonical energy-momentum $`M^c_\mu `$ and the “extended orbital angular momentum” $`L^{}{}_{\mu }{}^{}^\nu `$ are given by
$`M^c_\mu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }\stackrel{~}{𝑻}^{}{}_{\mu }{}^{}{}_{}{}^{\nu }d\sigma ^{}{}_{\nu }{}^{}={\displaystyle _\sigma }^{}{}_{\lambda }{}^{}𝚿_{}^{}{}_{\mu }{}^{}{}_{}{}^{\nu \lambda }d\sigma ^{}_\nu `$ (158)
$`=`$ $`{\displaystyle 𝚿^{}{}_{\mu }{}^{}{}_{}{}^{\nu \lambda }J\frac{x^0}{x^\nu }\frac{x^\alpha }{x^\lambda }n_\alpha r^2𝑑\mathrm{\Omega }},`$
$`L^{}{}_{\mu }{}^{}^\nu `$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }𝑴^{}{}_{\mu }{}^{}{}_{}{}^{\nu \lambda }d\sigma ^{}{}_{\lambda }{}^{}=2{\displaystyle _\sigma }^{}{}_{\tau }{}^{}(x^\nu 𝚿_\mu ^{}{}_{}{}^{\lambda \tau })d\sigma ^{}_\lambda `$ (159)
$`=`$ $`2{\displaystyle x^\nu 𝚿^{}{}_{\mu }{}^{}{}_{}{}^{\lambda \tau }J\frac{x^0}{x^\lambda }\frac{x^\alpha }{x^\tau }n_\alpha r^2𝑑\mathrm{\Omega }},`$
with $`J\stackrel{\text{def}}{=}det(x^\mu /x^\nu )`$, in which the transformed $`𝚿^{}{}_{\mu }{}^{}^{\nu \lambda }`$ can be obtained from Eqs. (2.1), (36), (144) – (146).
The relation
$$M^c{}_{\mu }{}^{}=0=(C^1)^\nu {}_{\mu }{}^{}M_{}^{c}_\nu $$
(160)
holds without any additional condition on the positive parameters $`p,s,\beta ,\gamma `$ and $`u`$, and $`L_\mu ^\nu `$ and $`L_{[\mu \nu ]}`$ transform according to
$`L^{}{}_{\mu }{}^{}^\nu `$ $`=`$ $`(C^1)^\rho {}_{\mu }{}^{}C_{}^{\nu }{}_{\sigma }{}^{}L_{\rho }^{}{}_{}{}^{\sigma },`$ (161)
$`L^{}_{[\mu \nu ]}`$ $`=`$ $`0=(C^1)^\lambda {}_{\mu }{}^{}(C^1)_{}^{\rho }{}_{\nu }{}^{}L_{[\lambda \rho ]}^{}`$ (162)
under the condition
$$p>1.$$
(163)
Equations (160) – (162) are the transformation rules which we would like $`M^c{}_{\mu }{}^{},L_\mu ^\nu `$ and $`L_{[\mu \nu ]}`$ to obey.
#### 4.1.4 Charge
For the transformed charge
$$C^{}\stackrel{\text{def}}{=}_\sigma ^{}{}_{\nu }{}^{}\left(\frac{𝑳^{\mathbf{}}}{A^{}_{\mu ,\nu }}\right)d\sigma ^{}{}_{\mu }{}^{},$$
(164)
with the Lagrangian density $`𝑳^{\mathbf{}}`$ defined with transformed field variables, we can obtain
$$C^{}=C=q$$
(165)
without imposing any additional condition.
To summarize, the G. E. P. is satisfied if
$$p>1,s>2.$$
(166)
The asymptotic behavior of the transformed dual components of $`b^k{}_{\mu }{}^{}=H^k{}_{l}{}^{}(x^\nu /x^\mu )b^l_\nu `$ and of the transformed vector potential $`A^{}{}_{\mu }{}^{}=(x^\nu /x^\mu )A_\nu `$ can be easily obtained from Eqs. (126), (128), (142) and (166) as
$`b^k_\mu `$ $`=`$ $`\mathrm{\Lambda }^k{}_{l}{}^{}(C^1)_{}^{\nu }{}_{\mu }{}^{}{}_{}{}^{(0)}b_{}^{l}{}_{\nu }{}^{}+O^k{}_{\mu }{}^{}(1/r^u)+O^k{}_{\mu }{}^{}(1/r),(u>0)`$
$`A^{}_\mu `$ $`=`$ $`(C^1)^\nu {}_{\mu }{}^{}A_{\nu }^{}+d^\nu {}_{\mu }{}^{}A_{\nu }^{}=O_\mu (1/r).`$ (167)
### 4.2 The case in which $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables
#### 4.2.1 Energy-momentum and angular momentum
For the generator of $`\overline{\text{Poincaré}}`$ gauge transformations, we have
$`\widehat{M}^{}_k`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }{}_{}{}^{\text{tot}}\widehat{𝑻^{}}{}_{k}{}^{}{}_{}{}^{\mu }d\sigma ^{}{}_{\mu }{}^{}0=\mathrm{\Lambda }_k{}_{}{}^{l}\widehat{M}_{l}^{},`$ (168)
$`\widehat{S}^{}_{kl}`$ $`\stackrel{\text{def}}{=}`$ $`{\displaystyle _\sigma }{}_{}{}^{\text{tot}}\widehat{𝑺^{}}{}_{kl}{}^{}{}_{}{}^{\mu }d\sigma ^{}{}_{\mu }{}^{}=\mathrm{\Lambda }_k{}_{}{}^{m}\mathrm{\Lambda }_{l}^{}{}_{}{}^{n}\widehat{S}_{mn}^{},`$ (169)
which hold without imposing any additional condition.
#### 4.2.2 Canonical energy-momentum and “extended orbital angular momentum”
For $`\widehat{M^c}_\mu `$, we have
$$\widehat{M}^c{}_{\mu }{}^{}\stackrel{\text{def}}{=}_\sigma \widehat{\stackrel{~}{𝑻^{}}}_\mu {}_{}{}^{\nu }d\sigma ^{}{}_{\nu }{}^{}=(C^1)^\nu {}_{\mu }{}^{}\widehat{M}_{}^{c}_\nu $$
(170)
if the conditions
$$p>\frac{1}{2},s>1,p+u>1$$
(171)
are satisfied. The transformed “extended orbital angular momentum”
$$\widehat{L}^{}{}_{\mu }{}^{}{}_{}{}^{\nu }\stackrel{\text{def}}{=}_\sigma \widehat{𝑴^{}}{}_{\mu }{}^{}{}_{}{}^{\nu \lambda }d\sigma ^{}_\lambda $$
(172)
is divergent in general, as is obvious from Eq. (3.2.3). However, the transformed orbital angular momentum $`\widehat{L}^{}_{[\mu \nu ]}`$, and hence the transformed total angular momentum $`\widehat{J}^{}{}_{kl}{}^{}\stackrel{\text{def}}{=}\widehat{S}^{}{}_{kl}{}^{}+{}_{}{}^{(0)}b_{}^{\mu }{}_{k}{}^{}{}_{}{}^{(0)}b_{}^{\nu }{}_{l}{}^{}\widehat{L}_{}^{}_{[\mu \nu ]}`$, are well defined, and they obey the rules
$`\widehat{L}^{}_{[\mu \nu ]}`$ $`=`$ $`(C^1)^\lambda {}_{\mu }{}^{}(C^1)_{}^{\rho }{}_{\nu }{}^{}\widehat{L}_{[\lambda \rho ]}^{},`$ (173)
$`\widehat{J}^{}_{kl}`$ $`=`$ $`\mathrm{\Lambda }_k{}_{}{}^{m}\mathrm{\Lambda }_{l}^{}{}_{}{}^{n}\widehat{J}_{mn}^{}`$ (174)
if the conditions
$$p>\frac{1}{2},s>2,u>1,p+u>2$$
(175)
are satisfied.
#### 4.2.3 Charge
The transformed charge
$$\widehat{C}^{}\stackrel{\text{def}}{=}_\sigma {}_{}{}^{}{}_{\nu }{}^{}\left(\frac{\widehat{𝑳^{}}}{A^{}_{\mu ,\nu }}\right)d\sigma ^{}{}_{\mu }{}^{},$$
(176)
with the Lagrangian density $`\widehat{𝑳^{}}`$ defined with transformed field variables, is evaluated as
$$\widehat{C}^{}=\widehat{C}=q,$$
(177)
without imposing any additional condition.
To summarize, the G. E. P. is satisfied if the conditions in Eq. (175) are satisfied. The asymptotic behavior of the transformed dual components of $`\widehat{b}^k{}_{\mu }{}^{}\stackrel{\text{def}}{=}H^k{}_{l}{}^{}(x^\nu /x^\mu )b^l_\nu `$ and of the transformed vector potential $`\widehat{A}^{}{}_{\mu }{}^{}\stackrel{\text{def}}{=}(x^\nu /x^\mu )A_\nu `$ are easily determined from Eqs. (126), (128), (142) and (175) as
$`\widehat{b}^k_\mu `$ $`=`$ $`\mathrm{\Lambda }^k{}_{l}{}^{}(C^1)_{}^{\nu }{}_{\mu }{}^{}{}_{}{}^{(0)}b_{}^{l}{}_{\nu }{}^{}+O^k{}_{\mu }{}^{}(1/r^p),(p>{\displaystyle \frac{1}{2}})`$
$`\widehat{A}^{}_\mu `$ $`=`$ $`(C^1)^\nu {}_{\mu }{}^{}A_{\nu }^{}+d^\nu {}_{\mu }{}^{}A_{\nu }^{}=O_\mu (1/r).`$ (178)
## 5 Summary and discussion
In extended new general relativity (E. N. G. R.), we have examined exact charged axi-symmetric solutions of the gravitational and electromagnetic field equations in vacuum from the point of view of the equivalence principle.
In this theory, the generators depend on the choice of the set of independent field variables. In §3, we examined the solution represented by Eqs. (67) and (68) for the case in which $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables and for the case in which $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent variables. We have shown the following:
For the case in which $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables, the total energy-momentum, the total angular momentum and the total electric charge of the system are all given by generators of internal transformations. The canonical energy-momentum and the orbital angular momentum vanish trivially.
For the case in which $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables, we have following: (1) The total energy-momentum is given by the generator $`M^c_\mu `$ of the coordinate translations, and the generator $`\widehat{M}_k`$ of the internal translations vanishes identically. (2) The total angular momentum is given by the sum $`\widehat{J}_{kl}\stackrel{\text{def}}{=}\widehat{S}_{kl}+{}_{}{}^{(0)}b_{}^{\mu }{}_{k}{}^{}{}_{}{}^{(0)}b_{}^{\nu }{}_{l}{}^{}\widehat{L}_{[\mu \nu ]}`$ of the generator $`\widehat{S}_{kl}`$ of the internal Lorentz transformations and of the generator $`\widehat{L}_{[\mu \nu ]}`$ of the coordinate Lorentz transformations. (3) The total charge is given by the generators of the internal $`U(1)`$ transformations.
For both cases, the total energy-momentum, the total angular momentum and the total charge of the system are identical to the corresponding active quantities of a central gravitating body.
The total mass, which is equal to the total energy divided by the square of the velocity of light, can be regarded as the inertial mass of the system. Thus, the results mentioned above include the equality of the inertial mass and the active gravitational mass, which implies that the equivalence principle is satisfied by the solution represented by Eqs. (67) and (68). In consideration of this, we have introduced the notion of a generalized equivalence principle (G. E. P.), as stated at the beginning of §4. Solutions obtained by applying {the transformations (126) and (127) satisfying the conditions (125) and (139)$`\}\{`$the coordinate transformation (4)} to the original solution have been examined from the point of view of the G. E. P. The following results have been obtained.
For the case in which $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables, the G. E. P. is satisfied by solutions if the conditions in Eq. (166) are satisfied.
For the case in which $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables, the G. E. P. is satisfied if the conditions in Eq. (175) are satisfied.
We would like to add several comments:
For the internal transformation (126), the condition (166) gives a stronger condition than does the condition (175). For the coordinate transformation (4), no constraint is imposed by the former, while the latter gives the restrictions $`u>1,p+u>2`$.
For the case in which $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables, the G. E. P. is satisfied even by solutions which approach constant values very slowly at spatial infinity, as is seen from Eq. (4.1.4).
This is a direct consequence of the following:
The functions $`𝐅_k^{\mu \nu }`$ and $`𝐅^{\mu \nu }=\sqrt{g}F^{\mu \nu }`$ both describe tensor densities, and hence the total energy-momentum $`M_k`$, total angular momentum $`S_{kl}`$ and the total charge $`C`$ are all independent of the coordinate systems employed.
The canonical energy-momentum $`M^c_\mu `$ and the “extended orbital angular momentum” $`L_\mu ^\nu `$, which are the generators of coordinate transformations, obey the regular transformation rules (160) and (161) under the coordinate transformation (4) with arbitrary positive $`u`$ if the condition (163) is satisfied.
Note that $`b^k_\mu `$ as given by Eq. (4.1.4) approaches a constant value much more slowly than the vierbein components required in the general situation to give reasonable forms of energy-momentum and angular momentum.
In the case in which $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$ is employed as the set of independent field variables, the G.E.P. for the total energy-momentum and for the total angular momentum is established when the vierbeins have asymptotic property as indicated by the first of Eq. (178). This is consistent with the results on the equivalence principle for the energy in new general relativity (N. G. R.).<sup>21</sup><sup>21</sup>21Note that E. N. G. R. is reduced to N. G. R. if fields with nonvanishing $`P_k`$ are not present and if the set $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu ,\varphi ^A,\varphi ^A\}`$ is employed as the set of independent field variables.
The preceding results, together with those in Ref., show that the choice $`\{\psi ^k,A^k{}_{\mu }{}^{},A_\mu \}`$ as the set of independent field variables is preferential to the choice $`\{\psi ^k,b^k{}_{\mu }{}^{},A_\mu \}`$. This is quite natural, because the fields $`\psi ^k,A^k_\mu `$ and $`A_\mu `$ are the fundamental objects and $`b^k_\mu `$ is a composite of $`\psi ^k`$ and $`A^k_\mu `$.
## Appendix A Asymptotic Forms of $`𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu },V^{mn}_\lambda `$ and $`W^{\mu \nu \lambda }_{klm}`$ for Large $`r`$
In this appendix, we give the asymptotic forms of $`𝑭^{(1)}{}_{k}{}^{}{}_{}{}^{\mu \nu },V^{mn}_\lambda `$ and $`W^{\mu \nu \lambda }_{klm}`$ for large $`r`$, which are useful for the calculations in §§3 and 4.
$`𝑭^{(1)}{}_{k}{}^{}^{\mu \nu }`$:
$`𝑭^{(1)}{}_{(0)}{}^{}^{0\alpha }`$ $`=`$ $`{\displaystyle \frac{1}{\kappa }}\left(a{\displaystyle \frac{Q^2}{r}}\right){\displaystyle \frac{n^\alpha }{r^2}}+{\displaystyle \frac{ah}{2\kappa r^3}}\epsilon ^\alpha {}_{\beta 3}{}^{}n_{}^{\beta }+O^\alpha \left({\displaystyle \frac{1}{r^4}}\right),`$
$`𝑭^{(1)}{}_{a}{}^{}^{0\alpha }`$ $`=`$ $`{\displaystyle \frac{1}{\kappa r^2}}(a{\displaystyle \frac{3Q^2}{2r}})n_an^\alpha {\displaystyle \frac{Q^2}{2\kappa r^3}}\delta _a^\alpha `$
$`{\displaystyle \frac{ah}{2\kappa r^3}}(\delta ^{\alpha \beta }3n^\alpha n^\beta )\epsilon _{a\beta 3}+O_a{}_{}{}^{\alpha }\left({\displaystyle \frac{1}{r^4}}\right),`$
$`𝑭^{(1)}{}_{(0)}{}^{}^{\alpha \beta }`$ $`=`$ $`{\displaystyle \frac{3ah}{2\kappa r^3}}(n^\alpha \epsilon ^\beta {}_{\gamma 3}{}^{}n^\beta \epsilon ^\alpha {}_{\gamma 3}{}^{})n^\gamma +{\displaystyle \frac{ah}{\kappa r^3}}\epsilon ^{\alpha \beta }{}_{3}{}^{}+O^{[\alpha \beta ]}\left({\displaystyle \frac{1}{r^4}}\right),`$
$`𝑭^{(1)}{}_{a}{}^{}^{\alpha \beta }`$ $`=`$ $`{\displaystyle \frac{ah}{2\kappa r^3}}(\epsilon _a{}_{}{}^{\alpha }{}_{3}{}^{}n_{}^{\beta }\epsilon _a{}_{}{}^{\beta }{}_{3}{}^{}n_{}^{\alpha })+{\displaystyle \frac{2ah}{\kappa r^3}}(n^\alpha \epsilon ^\beta {}_{\gamma 3}{}^{}n^\beta \epsilon ^\alpha {}_{\gamma 3}{}^{})n_an^\gamma `$ (179)
$`+{\displaystyle \frac{ah}{\kappa r^3}}\epsilon ^{\alpha \beta }{}_{3}{}^{}n_{a}^{}{\displaystyle \frac{Q^2}{2\kappa r^3}}(\delta _a{}_{}{}^{\alpha }n_{}^{\beta }\delta _a{}_{}{}^{\beta }n_{}^{\alpha })+O_a{}_{}{}^{[\alpha \beta ]}\left({\displaystyle \frac{1}{r^4}}\right).`$
$`V^{mn}_\lambda `$:
$`V^{(0)a}_0`$ $`=`$ $`n^a\dot{\mathrm{\Xi }}a{\displaystyle \frac{n^a}{r}}\dot{\mathrm{\Pi }}{\displaystyle \frac{n^a}{2}}\mathrm{\Xi }^2(\dot{\mathrm{\Xi }}{\displaystyle \frac{a}{r}}\dot{\mathrm{\Pi }})+a{\displaystyle \frac{n^a}{r}}\mathrm{\Xi }\mathrm{\Pi }\dot{\mathrm{\Xi }}`$
$`+O^a\left({\displaystyle \frac{1}{r^{s+1}}}\right)+O^a\left({\displaystyle \frac{1}{r^{2p+2}}}\right)+o^a\left({\displaystyle \frac{1}{r^3}}\right),`$
$`V^{(0)a}_\alpha `$ $`=`$ $`{\displaystyle \frac{1}{r}}\left(\delta ^a{}_{\alpha }{}^{}n^an_\alpha \right)\mathrm{\Xi }n^an_\alpha \mathrm{\Xi }^{}{\displaystyle \frac{a}{r^2}}\left(\delta ^a{}_{\alpha }{}^{}2n^an_\alpha \right)\mathrm{\Pi }`$
$`{\displaystyle \frac{an^an_\alpha }{r}}\mathrm{\Pi }^{}+{\displaystyle \frac{n^an_\alpha }{2}}\mathrm{\Xi }^2\mathrm{\Xi }^{}+O^a{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^{s+1}}}\right)`$
$`+O^a{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^{2p+2}}}\right)+o^a{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^3}}\right),`$
$`V^{ab}_0`$ $`=`$ $`{\displaystyle \frac{a}{r}}n^an^b(\mathrm{\Xi }\dot{\mathrm{\Pi }}+\dot{\mathrm{\Xi }}\mathrm{\Pi })a^2{\displaystyle \frac{n^an^b}{r^2}}\mathrm{\Pi }\dot{\mathrm{\Pi }}+{\displaystyle \frac{1}{2}}n^an^b\mathrm{\Xi }^3\dot{\mathrm{\Xi }}`$
$`+O^{ab}\left({\displaystyle \frac{1}{r^{s+1}}}\right)+O^{ab}\left({\displaystyle \frac{1}{r^{2p+2}}}\right)+o^{ab}\left({\displaystyle \frac{1}{r^3}}\right),`$
$`V^{ab}_\alpha `$ $`=`$ $`{\displaystyle \frac{1}{r}}(n^a\delta ^b{}_{\alpha }{}^{}n^b\delta ^a{}_{\alpha }{}^{})(1{\displaystyle \frac{1}{4}}\mathrm{\Xi }^2)\mathrm{\Xi }^2+O^{ab}{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^{s+1}}}\right)`$ (180)
$`+O^{ab}{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^{2p+2}}}\right)+o^{ab}{}_{\alpha }{}^{}\left({\displaystyle \frac{1}{r^3}}\right)`$
with<sup>22</sup><sup>22</sup>22Here, the prime and dot represent derivatives with respect to $`r`$ and $`x^0`$, respectively. For example, $`A^{}\stackrel{\text{def}}{=}A/r`$ and $`\dot{A}\stackrel{\text{def}}{=}A/x^0`$.
$$\mathrm{\Xi }\stackrel{\text{def}}{=}A^{}+\dot{B},\mathrm{\Pi }\stackrel{\text{def}}{=}\dot{A}+\dot{B}A^{}B^{}+\frac{1}{r}(A+B),$$
(181)
and $`o^a{}_{\alpha }{}^{}(1/r^3)`$, for example, denotes a term such that $`lim_r\mathrm{}r^3o^a{}_{\alpha }{}^{}(1/r^3)=0`$.
$`W^{\mu \nu \lambda }_{klm}`$:
$`W^{0\alpha \beta }_{(0)ab}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{b}{}^{}\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{a}{}^{})+{\displaystyle \frac{a}{4r}}\{(\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{b}{}^{}\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{a}{}^{})+(n_a\delta ^\alpha {}_{b}{}^{}n_b\delta ^\alpha {}_{a}{}^{})n^\beta `$
$`+(n_b\delta ^\beta {}_{a}{}^{}n_a\delta ^\beta {}_{b}{}^{})n^\alpha \}+O^{[\alpha \beta ]}_{[ab]}({\displaystyle \frac{1}{r^2}}),`$
$`W^{0\alpha \beta }_{abc}`$ $`=`$ $`{\displaystyle \frac{a}{4r}}\{(\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{b}{}^{}\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{a}{}^{})n_c+(\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{c}{}^{}\delta ^\alpha {}_{c}{}^{}\delta _{}^{\beta }{}_{b}{}^{})n_a`$
$`+(\delta ^\alpha {}_{c}{}^{}\delta _{}^{\beta }{}_{a}{}^{}\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{c}{}^{})n_b\}+O^{[\alpha \beta ]}_{[abc]}({\displaystyle \frac{1}{r^2}}),`$
$`W^{\alpha \beta \gamma }_{(0)ab}`$ $`=`$ $`{\displaystyle \frac{a}{4r}}\{(\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{b}{}^{}\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{a}{}^{})n^\gamma +(\delta ^\beta {}_{a}{}^{}\delta _{}^{\gamma }{}_{b}{}^{}\delta ^\beta {}_{b}{}^{}\delta _{}^{\gamma }{}_{a}{}^{})n^\alpha .`$
$`+(\delta ^\gamma {}_{a}{}^{}\delta _{}^{\alpha }{}_{b}{}^{}\delta ^\gamma {}_{b}{}^{}\delta _{}^{\alpha }{}_{a}{}^{})n^\beta \}+O^{[\alpha \beta \gamma ]}_{[ab]}({\displaystyle \frac{1}{r^2}}),`$
$`W^{\alpha \beta \gamma }_{abc}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\{(\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{b}{}^{}\delta ^\beta {}_{a}{}^{}\delta _{}^{\alpha }{}_{b}{}^{})\delta ^\gamma {}_{c}{}^{}+(\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{c}{}^{}\delta ^\beta {}_{b}{}^{}\delta _{}^{\alpha }{}_{c}{}^{})\delta ^\gamma _a`$ (182)
$`+(\delta ^\alpha {}_{c}{}^{}\delta _{}^{\beta }{}_{a}{}^{}\delta ^\beta {}_{c}{}^{}\delta _{}^{\alpha }{}_{a}{}^{})\delta ^\gamma {}_{b}{}^{}\}`$
$`{\displaystyle \frac{a}{4r}}\{(\delta ^\alpha {}_{a}{}^{}\delta _{}^{\beta }{}_{b}{}^{}\delta ^\beta {}_{a}{}^{}\delta _{}^{\alpha }{}_{b}{}^{})n_c+(\delta ^\alpha {}_{b}{}^{}\delta _{}^{\beta }{}_{c}{}^{}\delta ^\beta {}_{b}{}^{}\delta _{}^{\alpha }{}_{c}{}^{})n_a`$
$`+(\delta ^\alpha {}_{c}{}^{}\delta _{}^{\beta }{}_{a}{}^{}\delta ^\beta {}_{c}{}^{}\delta _{}^{\alpha }{}_{a}{}^{})n_b\}n^\gamma `$
$`{\displaystyle \frac{a}{4r}}[\{(\delta ^\alpha {}_{a}{}^{}n_{b}^{}\delta ^\alpha {}_{b}{}^{}n_{a}^{})n^\beta (\delta ^\beta {}_{a}{}^{}n_{b}^{}\delta ^\beta {}_{b}{}^{}n_{a}^{})n^\alpha \}\delta ^\gamma _c`$
$`+\{(\delta ^\alpha {}_{b}{}^{}n_{c}^{}\delta ^\alpha {}_{c}{}^{}n_{b}^{})n^\beta (\delta ^\beta {}_{b}{}^{}n_{c}^{}\delta ^\beta {}_{c}{}^{}n_{b}^{})n^\alpha \}\delta ^\gamma _a`$
$`+\{(\delta ^\alpha {}_{c}{}^{}n_{a}^{}\delta ^\alpha {}_{a}{}^{}n_{c}^{})n^\beta (\delta ^\beta {}_{c}{}^{}n_{a}^{}\delta ^\beta {}_{a}{}^{}n_{c}^{})n^\alpha \}\delta ^\gamma {}_{b}{}^{}]`$
$`+O^{[\alpha \beta \gamma ]}{}_{[abc]}{}^{}\left({\displaystyle \frac{1}{r^2}}\right).`$
All the other components, which are not listed above, are equal to zero or are obtainable from the listed ones by permutations.
|
warning/0007/hep-ex0007058.html
|
ar5iv
|
text
|
# 1 Motivation
## 1 Motivation
A large number of measurements on the structure of different hadronic objects has been performed in the past decades and elaborate descriptions of them have been developed in terms of parton distribution functions. Frequently, the many data points and multiple parameter descriptions make reading of physics messages from the data not obvious. This contribution attempts an intuitive description of structure functions which has a relation to QCD predictions and is applied to structure function measurements of the proton, the photon, and colour singlet exchange.
## 2 Quark Density and Scaling Violations
The QCD evolution equations predict that measurements of hadronic structures depend on the logarithm of the resolution scale $`Q^2`$ at which the structure is probed. On this basis, the following ansatz to analyse the $`x`$-dependence of structure function data is explored, where $`x`$ denotes the Bjorken fractional momentum of the scattered parton relative to the hadronic object:
$$F_2(x,Q^2)=a(x)\left[\mathrm{ln}\left(\frac{Q^2}{\mathrm{\Lambda }^2}\right)\right]^{\kappa (x)}.$$
(1)
Here $`\mathrm{\Lambda }`$ is a scale parameter, $`a`$ reflects the charge squared weighted quark distributions extrapolated to $`\mathrm{ln}(Q^2/\mathrm{\Lambda }^2)=1`$, and $`\kappa `$ determines the positive and negative scaling violations of $`F_2`$.
In Fig.1, H1 and ZEUS low-$`x`$ data of the proton structure function $`F_2`$ for $`Q^2>2`$ GeV<sup>2</sup> are shown. In each $`x`$-bin, the result of a two-parameter fit according to eq. (1) is shown, using a fixed value of $`\mathrm{\Lambda }=0.35`$ GeV which represents a typical value of the strong interaction scale. Only the total experimental errors have been used, ignoring correlations between individual data points. The same fitting procedure has been applied to BCDMS data which are taken here as a reference sample for the high-$`x`$ region (Fig.2) (after finalising this analysis the author’s attention was called to a study fitting the BCDMS data to a similar expression ). All data turn out to be well described ($`\chi ^2/`$ndf: $`59/107`$ H1, $`112/115`$ ZEUS, $`83/155`$ BCDMS).
The resulting parameters $`a`$ and $`\kappa `$ are summarized in Fig.3 as a function of $`x`$. For $`a`$, the data fits exhibit two distinct regions: around $`x0.3`$ they reflect the valence quark distributions. At low $`x`$, $`a(x)`$ is compatible with converging to a constant value. The resulting scaling violation term $`\kappa `$ appears to rise as $`x`$ decreases, exhibiting the negative and positive scaling violations of $`F_2`$ for $`x`$ above and below $`0.1`$ respectively. The errors in Fig.3 represent the statistical errors of the fits. Both parameters $`a`$ and $`\kappa `$ are anti-correlated as can be seen from neighbouring points. No significant $`Q^2`$-dependence of $`a`$ and $`\kappa `$ has been found when the fits were repeated for two intervals in $`Q^2`$ (above and below $`20`$ GeV<sup>2</sup>).
With $`a`$ being approximately constant below $`x0.004`$, changes of $`F_2`$ at low $`x`$ result from the scaling violation term $`\kappa `$ alone, indicative of the interaction dynamics that drives $`F_2`$ and in support of .
For comparison, the same fits have been applied to recent measurements of the photon structure function $`F_2^\gamma `$ which have been performed at $`e^+e^{}`$ colliders (Fig.4a, $`\chi ^2/`$ndf: $`15/37`$). The photon structure results from fluctuations of a photon into a colour neutral and flavour neutral hadronic state. The values of the parameters $`a`$ and $`\kappa `$ are shown in Fig.3 as the open circles. The photon data exhibit positive scaling violations at all $`x`$ that are distinct from those of a hadronic bound state like the proton. The value of $`\kappa `$ is approximately $`1`$. This is as expected from QCD calculations which predict $`F_2^\gamma `$ for $`0.1<x<1`$ . It is interesting to note that in the low-$`x`$ region around $`x0.1`$ the photon data prefer similar values of $`a`$ to the low-$`x`$ proton data.
In Fig.4b, H1 structure function measurements $`F_2^{D(3)}`$ of colour singlet exchange are compared to the same two-parameter fits as used above ($`\chi ^2/`$ndf: $`24/30`$). Here $`x`$ denotes the fractional momentum of the scattered parton relative to the colour neutral object, which itself carries a fractional momentum $`\xi =0.003`$ relative to the proton. Also these data exhibit scaling violations $`\kappa `$ that are different from the proton measurements at the same values of $`x`$ (Fig.3b). Instead, at $`0.1<x<0.5`$ they are large and similar to the photon data and to the low-$`x`$ proton data. The large rate of events with colour singlet exchange together with the large scaling violations of $`F_2^{D(3)}`$ is suggestive of a gluon dominated exchange. The values of the normalization $`a`$ increase with increasing $`x`$ to about $`a=10`$ as can be seen from extrapolating the curves in Fig.4b towards small $`Q^2`$. These values have large uncertainties of the order of $`100\%`$ and are not shown in Fig.3a.
## 3 Relation of the Parameters with QCD Predictions
In the following only the proton data are considered. The parameter $`a`$ has already been identified as the charge squared weighted quark distributions extrapolated to $`\mathrm{ln}(Q^2/\mathrm{\Lambda }^2)=1`$. An understanding of the parameters $`\mathrm{\Lambda }`$ and $`\kappa `$ can be achieved by comparison with the QCD evolution equation which is written here in the leading order DGLAP approximation:
$$\frac{\mathrm{d}f_i(x,Q^2)}{\mathrm{d}\mathrm{ln}Q^2}=\frac{\alpha _s(Q^2)}{2\pi }\underset{j}{}_x^1\frac{\mathrm{d}y}{y}P_{ij}\left(\frac{x}{y}\right)f_j(y,Q^2).$$
(2)
Here $`f_i,f_j`$ denote the parton densities, $`P_{ij}`$ are the splitting functions, and
$$\alpha _s=\frac{b}{\mathrm{ln}(Q^2/\mathrm{\Lambda }_{QCD}^2)}$$
(3)
is the strong coupling constant.
The derivative of the ansatz chosen here, eq. (1), with respect to $`\mathrm{ln}Q^2`$ gives
$$\frac{\mathrm{d}F_2(x,Q^2)}{\mathrm{d}\mathrm{ln}Q^2}=\frac{1}{\mathrm{ln}(Q^2/\mathrm{\Lambda }^2)}\kappa (x)F_2(x,Q^2),$$
(4)
where relating $`1/\mathrm{ln}(Q^2/\mathrm{\Lambda }^2)`$ with $`\alpha _s`$ in eq. (2) implies association of the scale parameter $`\mathrm{\Lambda }`$ in eq. (1) with the QCD parameter $`\mathrm{\Lambda }_{QCD}`$. The term $`\kappa `$ relates with the sum over the different parton radiation terms in eq. (2) divided by $`F_2`$. $`\kappa `$ increases towards small $`x`$, consistent with larger phase space available for parton radiation.
To match the description chosen here, eq. (1), with the double asymptotic approximation expected from QCD for the gluon-dominated region at small $`x`$, $`F_2\mathrm{exp}\sqrt{\mathrm{ln}x\mathrm{ln}(\mathrm{ln}(Q^2/\mathrm{\Lambda }^2))}`$ , the scaling violation term $`\kappa `$ is required to have a dependence like
$`\kappa \sqrt{\mathrm{ln}x/\mathrm{ln}(\mathrm{ln}(Q^2/\mathrm{\Lambda }^2))}`$. The $`Q^2`$-dependence of $`\kappa `$ is therefore expected to be very small which is in agreement with the experimental observation stated above.
## 4 Concluding Remarks
Based on logarithmic $`Q^2`$-dependence, the presented fits of structure function measurements in bins of $`x`$ give information on the hadronic structures and the parton dynamics.
Published proton data, extrapolated to a low resolution scale $`Q^2`$, show beyond the valence quarks a sea quark density which is compatible with a constant value. Therefore, at small $`x`$, the $`x`$\- and $`Q^2`$-dependencies of $`F_2`$ dominantly originate from the measured scaling violations which are directly related to the parton dynamics of the interaction. More precise data and data reaching smaller values of $`x`$ will determine whether or not the scaling violations further increase towards low-$`x`$ and therefore give valuable information on the parton densities in the proton as $`x`$ approaches $`0`$.
The large scaling violations observed in the structure function data of colour singlet exchange are suggestive of a gluon dominated exchange. More precise measurements will test whether at low $`Q^2`$ one parton carries most of the momentum and will provide a reference for a gluon driven regime.
Judgement on a universal low-$`x`$ behaviour of hadronic structures will result from measurements of the photon structure function. If the photon data show a constant quark density at small $`x`$ similar to the low-$`x`$ proton data, scaling violations of $`F_2^\gamma `$, which deviate from those resulting from the photon splitting into quark-antiquark pairs and approach those observed for the proton, could become visible in the regions around $`x0.03`$ or $`x10^3`$.
## Acknowledgements
For fruitful discussions I wish to thank J. Dainton, J. Gayler, D. Graudenz, T. Naumann, P. Newman, R. Nisius, P. Schleper and F. Zomer. I wish to thank Th. Müller and the IEKP group of the University Karlsruhe for their hospitality, and the Deutsche Forschungsgemeinschaft for the Heisenberg Fellowship.
|
warning/0007/hep-th0007034.html
|
ar5iv
|
text
|
# Magnetic Instability in a Parity Invariant 2D Fermion System
## I Introduction
Interest to the (2+1)-dimensional field theory is heated up by its successful applications to the problems of planar condensed matter systems like the quantum Hall fluid and unconventional superconductors. One of the relevant issues concerns the phenomenon of spontaneous magnetization in the systems of planar fermion matter.
Different authors have given convincing arguments that the perturbative ground state (the one with $`F_{\mu \nu }=0`$) of a 2D dense fermion matter exhibits the magnetic instability. In other terms there does exist the true ground state with the nonzero value of a proper magnetic field.
The peculiar property of planar physics is that spin is not quantized, and spin-up and spin-down states belong to the different representations of the rotation group $`SO(2)`$. Remark, that the existence of magnetized ground state was demonstrated in the systems of fermions with one and the same spin orientation. Hence one can propose the simple mechanism explaining this phenomenon: matter consisting entirely of either up or down spin particles possesses nonvanishing spin density leading to a proper magnetization of the system.
In the present paper we study the double-spin model (also referred to as duplicated) where the matter represents the mixture of opposite spin fermions. Such a consideration can be motivated by the observation<sup>7</sup> that the magnetic properties of duplicated and single-spin (up or down) systems can be quite different.
In the rest part of this section we remind some principal aspects of spontaneous magnetization in the single-spin system. In Section 2 we perform the perturbative study of duplicated system and find out magnetic instability of perturbative ground state. The mechanism of instability in the double-spin model turns out to be different from that in the single-spin systems. Some necessary calculations are placed in the Appendix.
The basic object of our account is the finite temperature effective action $`W[A]`$ which arises after integrating out the fermion fields. In the Matsubara formalism it is given by
$`\mathrm{exp}\left\{W[A]\right\}={\displaystyle D\psi ^{}D\psi \mathrm{exp}\left\{L𝑑\tau d𝐫\right\}},`$ (1)
where $`L`$ is the Euclidean Lagrangian, $`0<\tau <T^1`$, and $`T`$ is the temperature.
Consider the nonrelativistic fermion matter interacting with $`U(1)`$ gauge field. The corresponding Euclidean Lagrangian is given by
$`L={\displaystyle \frac{1}{4}}F_{\mu \nu }F_{\mu \nu }ie\rho A_\tau +\psi ^{}(_\tau +ieA_\tau )\psi +`$ (2)
$`+{\displaystyle \frac{1}{2m}}|(_k+ieA_k)\psi |^2{\displaystyle \frac{\sigma e}{2m}}B\psi ^{}\psi \mu \psi ^{}\psi ,`$ (3)
where $`\rho `$ is the neutralizing charge density, $`\sigma `$ describes the spin degree of freedom, magnetic field is defined as $`B=_2A_1_1A_2`$, and $`\mu `$ is the chemical potential.
The linear part of $`W[A]`$ appears as
$`W_1=e{\displaystyle \{i[\psi ^{}\psi \rho ]A_\tau \frac{\sigma }{2m}\psi ^{}\psi B\}𝑑\tau d𝐫},`$ (4)
where the symbol $`\mathrm{}`$ denotes the thermal average corresponding to $`F_{\mu \nu }=0`$.
From the last expression one can extract the principal conclusion concerning the magnetic properties of the single-spin system. Due to the term, linear in the magnetic field, the extremum of $`W[A]`$ will be realized for some $`B0`$. The common conclusion achieved in the finite temperature analysis of single-spin models, implies that the spontaneous magnetization persists at any temperatures. The reason is quite simple: the finite spin density $`(\sigma /2)\psi ^{}\psi `$ creates the proper magnetization of the system
$`M={\displaystyle \frac{\delta W_1}{\delta B}}={\displaystyle \frac{\sigma e}{2m}}\psi ^{}\psi ,`$ (5)
which survives at any temperatures, provided the particle density is finite.
Concerning the $`A_\tau `$-term, it leads to the electric field corresponding to a local charge distribution $`e\psi ^{}\psi e\rho `$.
## II Double-Spin System
In this section we consider the duplicated model where the matter represents the mixture of opposite spin ($`\pm \sigma `$) fermions.
The system is assumed to be in contact with a particle reservoir which keeps the total particle density fixed and guarantees the chemical equilibrium between spin-up and spin-down subsystems. This can be realized by equalizing the corresponding chemical potentials. Then $`\psi _{}^{}\psi _{}=\psi _{}^{}\psi _{}`$, and the linear part of $`W`$ appears as
$`W_1=ie{\displaystyle [\psi _{}^{}\psi _{}+\psi _{}^{}\psi _{}\rho ]A_\tau 𝑑\tau d𝐫},`$ (6)
where the term linear in the magnetic field is absent since the duplicated model is parity invariant.<sup>*</sup><sup>*</sup>*Magnetic field defined in 2+1 dimensions is a pseudoscalar. Therefore, the spin interaction leads to a parity violation in single-spin systems. Provided spin-up and spin-down fermions are interchanged under the parity transformation the duplicated model is parity invariant. Therefore, the configuration with $`B=0`$ represents the extremum of the corresponding effective action. The prior question which arises in this connection is the one concerning the stability of this extremum. This issue can be studied in terms of the two-point functions calculated within the Gaussian approximation. Considering these functions as integral operators we can introduce the corresponding eigenvalues which characterize the spectrum of gauge field fluctuations. Stability of the perturbative ground state requires the non-negative definiteness of this eigenvalue spectrum. If the spectrum contains at least one negative eigenvalue, then one is faced with an instability leading to the formation of the nontrivial ground state. In what follows we adopt this criterion and search for the negative eigenvalues of the two-point functions determining the Gaussian part. The later is given by
$`W_2={\displaystyle \frac{1}{4}}{\displaystyle F_{\mu \nu }F_{\mu \nu }𝑑x}+{\displaystyle \frac{e^2m}{2}}{\displaystyle A_\mu (x_1)G_{\mu \nu }(x_1x_2)A_\nu (x_2)𝑑x_1𝑑x_2},`$ (7)
where $`dx=d\tau \mathrm{d}𝐫`$, and $`G_{\mu \nu }(x)`$ are the current correlation functions. In the Fourier representation they look as
$`G_{\tau \tau }={\displaystyle \frac{\mathrm{d}𝐤^{}}{2\pi ^2}\frac{f(\beta E_+)f(\beta E_{})}{(\mathrm{𝐤𝐤}^{})^2+m^2\xi ^2}(\mathrm{𝐤𝐤}^{})},`$ (8)
$`G_{\tau i}=G_{i\tau }={\displaystyle \frac{\xi k_i}{𝐤^2}}G_{\tau \tau },`$ (9)
$`G_{ij}=\delta _{ij}{\displaystyle \frac{\xi ^2}{𝐤^2}}G_{\tau \tau }+{\displaystyle \frac{1}{m^2}}(𝐤^2\delta _{ij}k_ik_j)G,`$ (10)
$`G={\displaystyle \frac{\mathrm{d}𝐤^{}}{2\pi ^2}\frac{f(\beta E_+)f(\beta E_{})}{(\mathrm{𝐤𝐤}^{})^2+m^2\xi ^2}(\mathrm{𝐤𝐤}^{})\left\{\frac{𝐤^2}{𝐤^2}2\frac{(\mathrm{𝐤𝐤}^{})^2}{𝐤^4}+\frac{\sigma ^2}{4}\right\}},`$ (11)
where $`\xi =2\pi Tn`$ with $`n=0,\pm 1,\pm 2\mathrm{}`$ , and $`\beta =T^1`$. The quantities $`E_\pm `$ and $`f(z)`$ are given by
$`E_\pm ={\displaystyle \frac{1}{2m}}\left(𝐤^{}\pm {\displaystyle \frac{𝐤}{2}}\right)^2,`$ (12)
$`f(z)={\displaystyle \frac{1}{1+e^{z\mu /T}}}.`$ (13)
The induced Chern-Simons terms corresponding to the spin-up and spin-down fermions are mutually canceled out in accordance with the parity invariance.
From the definition of $`E_\pm `$ it follows that $`E_\pm >E_{}`$ for sgn$`(\mathrm{𝐤𝐤}^{})=\pm 1`$, and since $`f(z)`$ is a decreasing function, we get
$`[f(\beta E_+)f(\beta E_{})](\mathrm{𝐤𝐤}^{})<0.`$ (14)
This implies that $`G_{\tau \tau }>0`$, and the corresponding contributions to $`W_2`$ are positive. The rest part of $`W_2`$ is related with pure magnetic configurations and determines the two-point function
$`\mathrm{\Lambda }(x_1,x_2)={\displaystyle \frac{\delta ^2W_2}{\delta B(x_1)\delta B(x_2)}},`$ (15)
which, up to a constant factor, is the susceptibility of the perturbative ground state.
Because of the translational invariance, it is convenient to analyze the eigenvalues of $`\mathrm{\Lambda }(x_1,x_2)`$ in the Fourier representation. These eigenvalues can be written in the following form
$`\lambda (\xi ,𝐤,T)=1+{\displaystyle \frac{e^2}{m}}G(\xi ,𝐤,T),`$ (16)
where $`G(\xi ,𝐤,T)`$ is given by the Eq. (1).
From these expressions one can detect the possibility of the existence of negative eigenvalues leading to the magnetic instability. In fact, due to the relation (2), the $`\sigma `$-dependent part of $`G`$ is negative, and for sufficiently large values of $`\sigma ^2`$ the quantity $`G(\xi ,𝐤,T)`$ will become negative at least for some $`\xi `$ and $`𝐤`$.
In order, to expose the underlying physics, consider the dense matter of spinless fermions. In the absence of gauge fields fermions are organized in a Fermi sphere. Turning on the homogeneous magnetic field, fermions will be rearranged into the Landau levels, and the orbital diamagnetism will lead to the increase of the energy of the system. Let us now attach the spin to these fermions. Inclusion of the spin with positive (negative) $`\sigma eB`$ decreases (increases) the energy. Due to the chemical equilibrium between the opposite spin subsystems, the partial density of fermions with $`\sigma eB>0`$ exceeds the one corresponding to $`\sigma eB<0`$. Therefore, the spin effects lead to the spin paramagnetism promoting the decrease of the energy. If $`\sigma ^2`$ is large enough, then the paramagnetism dominates over the diamagnetism and causes the overall decrease of the energy of the fermion system.
Further, for sufficiently large values of $`e^2/m`$, the negative values of $`G`$ will dominate in Eq. (3), and for some $`\xi `$ and $`𝐤`$ we shall get $`\lambda <0`$.
In the light of these arguments we can formulate two main points required for the existence of negative modes:
* the magnitude of $`\sigma `$ must be large enough to generate the instability of the fermion matter.
* the value of the fraction $`e^2/m`$ must be large enough to guarantee that the matter instability will not be compensated by the gauge field contributions.
In the following two subsections we present more detailed analysis of the negative modes for static $`(\xi =0)`$ and nonstatic $`(\xi 0)`$ cases, respectively.
### A Static case
Let us first explore the point (i) concerning the negative values of $`G`$. As it is shown in the Appendix, the quantity $`G(0,k,T)`$ can be written in the following form
$`G(0,k,T)={\displaystyle _0^1}{\displaystyle \frac{ds}{4\pi }}{\displaystyle \frac{s^2\sigma ^2}{1+e^{\zeta (1s^2)\mu /T}}},`$ (17)
where $`\zeta =k^2/8mT`$ and $`k=|𝐤|`$.
Assuming that the total particle density is fixed, we express the chemical potential in terms of other parameters. The defining equation is $`\psi _{}^{}\psi _{}+\psi _{}^{}\psi _{}=n_e`$, with
$`\psi _{}^{}\psi _{}=\psi _{}^{}\psi _{}={\displaystyle \frac{mT}{2\pi }}\mathrm{ln}(1+e^{\mu /T}),`$ (18)
and consequently
$`\mu =T\mathrm{ln}\left(e^{1/\mathrm{\Theta }}1\right),`$ (19)
where $`\mathrm{\Theta }=mT/\pi n_e`$.
Consider first the zero temperature limit. In that case the Fermi distribution function becomes steplike, and we get
$`G(0,k,0)={\displaystyle \frac{13\sigma ^2}{12\pi }}{\displaystyle \frac{1}{12\pi }}\sqrt{1{\displaystyle \frac{k_F^2}{k^2}}}\left[13\sigma ^2{\displaystyle \frac{k_F^2}{k^2}}\right]\theta (kk_F),`$ (20)
where the characteristic momentum $`k_F`$ is defined by
$`k_F^2=8m\underset{T0}{lim}\mu =8\pi n_e.`$ (21)
In Fig. 1(a) we present $`G(0,k,0)`$ for the different values of $`\sigma ^2`$. As one can be convinced, the negative eigenvalues appear only for $`\sigma ^2>1/3`$.
Consider now the point (ii) and trace out the condition guaranteeing that the negative values of $`G(0,k,0)`$ will lead to $`\lambda <0`$. For $`\sigma ^2>1/3`$ the minimal value of $`G(0,k,0)`$ is given by
$`G(0,0<k<k_F,0)={\displaystyle \frac{3\sigma ^21}{12\pi }},`$ (22)
and the required condition appears to be
$`{\displaystyle \frac{m}{e^2}}<{\displaystyle \frac{3\sigma ^21}{12\pi }}.`$ (23)
In this case one obtains that $`\lambda (0,k,0)<0`$ for $`0<k<k_c`$, where $`k_c`$ is some critical value which is the solution to $`\lambda (0,k,0)=0`$. Remark, that $`k_c>k_F`$.
The finite temperature behaviour of $`G(0,k,T)`$ is depicted in Fig. 1(b), where the case of $`\sigma ^2=1`$ is represented for different values of the parameter $`\mathrm{\Theta }`$. We see that $`G(0,k,T)`$ tends to zero when $`T`$ increases. Therefore, the negative modes, observed at low temperatures, disappear in the high temperature regime.
Remark, that the relation (5) can be realized only for $`\sigma ^2>1/3`$ and therefore, embraces the points (i) and (ii) simultaneously. Thus, the relation (5) appears to be a sufficient condition for the existence of $`\lambda (0,k,T)<0`$. Moreover, it is also a necessary one, since in the opposite case $`\lambda (0,k,T)`$ is positive for all $`k`$ and $`T`$. In order to check up this assertion, one can carry out some estimates. Note, that in (4) we have $`0<s<1`$. In this interval $`e^{\zeta (1s^2)}>e^{\zeta (1s)}`$ and $`e^{\zeta (1s^2)}<e^{\zeta (1s^3)}`$. Using these inequalities in (4), one gets
$`\lambda (0,k,T)>1+{\displaystyle \frac{e^2}{m}}{\displaystyle \frac{13\sigma ^2}{12\pi }}{\displaystyle \frac{1}{\zeta }}\mathrm{ln}{\displaystyle \frac{1+e^{\mu /T}}{1+e^{\mu /T\zeta }}}.`$ (24)
Further, since $`\zeta >0`$, we have
$`0<{\displaystyle \frac{1}{\zeta }}\mathrm{ln}{\displaystyle \frac{1+e^{\mu /T}}{1+e^{\mu /T\zeta }}}<1`$ (25)
and consequently,
$`\lambda (0,k,T)>1+{\displaystyle \frac{e^2}{m}}{\displaystyle \frac{13\sigma ^2}{12\pi }}.`$ (26)
This is a general relation valid for all values of $`k`$ and $`T`$, and leading to $`\lambda (0,k,T)>0`$ when the condition (5) is not held.
### B Nonstatic case
Integrating out the polar angle in (1), the corresponding expression can be presented in the following form
$`G(\xi ,k,T)={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dz}{4\pi }}\left[{\displaystyle \frac{1\sigma ^2z^2}{2}}x+{\displaystyle \frac{4m^2\xi ^2z^2}{4m^2\xi ^2+k^4(1+z)^2}}{\displaystyle \frac{1}{x}}\right](\epsilon 1)f(\zeta z^2),`$ (27)
$`\epsilon =\sqrt{{\displaystyle \frac{4m^2\xi ^2+k^4(1+z)^2}{4m^2\xi ^2+k^4(1z)^2}}},x=\sqrt{{\displaystyle \frac{1}{2}}{\displaystyle \frac{4m^2\xi ^2+k^4(1z^2)}{4m^2\xi ^2+k^4(1+z)^2}}+{\displaystyle \frac{1}{2\epsilon }}}.`$ (28)
In Fig. 2 we depict the behaviour of $`G(\xi ,k,T)`$ for $`\sigma ^2=0`$. In that case paramagnetism is absent and the eigenvalue spectrum is positively defined.
The case $`\sigma ^2=1`$ is depicted in Fig. 3, where we observe that $`G`$ becomes negative for some values of $`\xi `$ and $`k`$.
From the expression (6) one can extract the asymptotic properties of $`G(\xi ,k,T)`$ for $`k\mathrm{}`$ and $`\xi \mathrm{}`$. In the large-$`k`$ limit we obtain
$`G2{\displaystyle \frac{1\sigma ^2}{2\pi \delta }}{\displaystyle \frac{mT}{k^2}}\mathrm{ln}(1+e^{\mu /T})for\sigma ^21`$ (29)
$`G8{\displaystyle \frac{\delta 2}{\pi \delta ^2}}{\displaystyle \frac{m^2T^2}{k^4}}{\displaystyle _0^{\mathrm{}}}\mathrm{ln}(1+e^{z+\mu /T})𝑑zfor\sigma ^2=1`$ (30)
where $`\delta =1+(2m\xi /k^2)^2`$. These expressions are valid irrespectively whether the fraction $`\xi /k^2`$ vanishes, diverges or stays finite when $`k\mathrm{}`$.
Consider the limit $`\xi \mathrm{}`$. The case when $`k`$ tends to infinity together with $`\xi `$ has been already discussed above. Therefore, we assume that $`k`$ is finite and get
$`G{\displaystyle \frac{1\sigma ^2}{4\pi }}{\displaystyle \frac{Tk^2}{m\xi ^2}}\mathrm{ln}(1+e^{\mu /T})+{\displaystyle \frac{2}{\pi }}{\displaystyle \frac{T^2}{\xi ^2}}{\displaystyle _0^{\mathrm{}}}\mathrm{ln}(1+e^{z+\mu /T})𝑑z.`$ (31)
The asymptotic relations imply that the negative values of $`\lambda `$ can be located in a finite region of the $`(\xi ,k)`$-plane.
Figures 2 and 3 demonstrate that the magnitude of the negative values of $`G`$ tend to zero as $`T`$ increases. Therefore, in the high temperature regime we get $`\lambda >0`$.
## III Acknowledgments
We would like to thank G. Japaridze, L. O’Raifeartaigh, V. Rubakov and P. Sodano for helpful discussions. The work was supported by the grant Intas-Georgia 97-1340.
## IV Appendix
Here we comment on the calculation of $`G(0,𝐤,T)`$. The common details are shown for its $`\sigma `$-dependent part:
$`G_{\tau \tau }(0,𝐤,T)={\displaystyle \frac{\mathrm{d}𝐤^{}}{2\pi ^2}\frac{f(\beta E_+)f(\beta E_{})}{\mathrm{𝐤𝐤}^{}}}.`$ (A.1)
Due to the singularity, the $`E_\pm `$ parts are separately divergent and the integral cannot be decoupled in the corresponding way.
Expanding $`f(\beta E_\pm )`$ in powers of $`\mathrm{𝐤𝐤}^{}`$, we use the polar variables and integrating over $`𝐤^{}`$ get
$`G_{\tau \tau }(0,k,T)={\displaystyle \frac{1}{\pi }}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n\zeta ^n}{n!(2n+1)}}f^{(n)}(\zeta ),`$ (A.2)
where $`\zeta =k^2/8mT`$. Expand $`f^{(n)}(\zeta )`$ and arrange the powers of $`\zeta `$. Besides, we use $`f^{(j)}(0)=(1)^j\varphi ^{(j)}(\mu /T)`$ with $`\varphi (z)=(1+e^z)^1`$ and get
$`G_{\tau \tau }(0,k,T)={\displaystyle \frac{1}{2\sqrt{\pi }}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n\zeta ^n}{\mathrm{\Gamma }(n+3/2)}}\varphi ^{(n)}(\mu /T).`$ (A.3)
Analyze the structure of $`\varphi ^{(n)}(z)`$. From $`\varphi ^{}(z)=\varphi (1\varphi )`$ it follows that $`\varphi ^{(n)}(z)`$ can be presented as an $`(n+1)`$’th order polynomial with respect to $`\varphi `$
$`\varphi ^{(n)}(z)=P_{n+1}[\varphi (z)].`$ (A.4)
These polynomials satisfy the recurrency relation $`P_{n+1}=\varphi (1\varphi )(dP_n/d\varphi )`$ with $`P_1[\varphi ]=\varphi `$. By the direct calculation it can be checked up that the solution to this recurrency chain appears as
$`P_{n+1}[\varphi ]={\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{l=0}{\overset{k}{}}}(1)^l(l+1)^nC_k^l\varphi ^{k+1}.`$ (A.5)
Substituting Eqs. (A.2) and (A.3) into Eq. (A.1), we use
$`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^nx^n}{\mathrm{\Gamma }(n+\nu +1)}}={\displaystyle \frac{2}{\mathrm{\Gamma }(\nu )}}{\displaystyle _0^1}e^{x(1s^2)}s^{2\nu 1}𝑑s,`$ (A.6)
and summing up over $`l`$, get
$`G_{\tau \tau }(0,k,T)={\displaystyle _0^1}{\displaystyle \frac{dt}{\pi }}e^{\zeta (1s^2)}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{[1e^{\zeta (1s^2)}]^k}{[1+e^{\mu /T}]^{k+1}}}.`$ (A.7)
The infinite sum over $`k`$ converges, yielding
$`G_{\tau \tau }(0,k,T)={\displaystyle \frac{1}{\pi }}{\displaystyle _0^1}{\displaystyle \frac{ds}{1+e^{\zeta (1s^2)\mu /T}}}.`$ (A.8)
Performing the similar manipulations for the $`\sigma `$-independent part of $`G`$ and combining with (A.4), we get (4).
|
warning/0007/hep-th0007150.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Describing the laws of physics in terms of underlying symmetries has always been a powerful tool. In this respect, it is interesting to study the kind of symmetries which are allowed in space-time. Within the framework of Quantum Field Theory (unitarity of the $`S`$ matrix etc) it is generally admitted that we cannot go beyond supersymmetry (SUSY). However, the no-go theorem stating that supersymmetry is the only non-trivial extension beyond the Poincaré algebra is valid only if one considers Lie or Super-Lie algebras. Indeed, if one considers Lie algebras, the Coleman and Mandula theorem allows only trivial extensions of the Poincaré symmetry, i.e. extra symmetries must commute with the Poincaré generators. In contrast, if we consider superalgebras, the theorem of Haag, Lopuszanski and Sohnius shows that we can construct a unique (up to the number of supercharges) superalgebra extending the Poincaré Lie algebra non-trivially. It may seem that these two theorems encompass all possible symmetries of space-time. But, if one examines the hypotheses of the above theorems, one sees that it is possible to imagine symmetries which go beyond supersymmetry. Several possibilities have been considered in the literature , the intuitive idea being that the generators of the Poincaré algebra are obtained as an appropriate product of more fundamental additional symmetries. These new generators are in a representation of the Lorentz group which can be neither bosonic nor fermionic (bosonic charges close under commutators and generate a Lie algebra, whilst fermionic charges close under anticommutators and induce super-Lie algebras). In an earlier work we proposed an algebraic structure, called an $`F`$Lie algebra, which makes this idea precise in the context of fractional supersymmetry (FSUSY) of order $`F`$ . Of course, when $`F=1`$ this is a Lie algebra, and when $`F=2`$ this is a Super-Lie algebra. Within the framework of this algebraic structure, we showed that starting from any representation $`𝒟`$ of any Lie algebra $`g`$ it is possible to take in some sense the $`F^{\mathrm{th}}`$root of $`𝒟`$. This means that we were able to consider a representation $`𝒟^{}`$ such that the symmetric tensorial product of order $`F`$ $`𝒮^F\left(𝒟^{}\right)`$ is related to $`𝒟`$ . The representation $`𝒟^{}`$ is in general an infinite dimensional representation of $`g`$, i.e. a Verma module . The purpose of this note is to give an explicit way to realize this construction in terms of monomials, i.e. using appropriate differential realizations of $`g`$.
The content of this paper is as follow. In section 2, we summarize the results of about $`F`$Lie algebras. In section 3 we show how for any Lie algebra we can realize the representations in terms of homogeneous monomials constructing differential realization(s) of the algebra $`g`$. Finally, in section 4 FSUSY is realized for the algebra $`g=su(3)`$. Of course this construction works along the same lines for any Lie algebra.
## 2 Algebraic Structure of Fractional Supersymmetry
In this section, we recall the abstract mathematical structure which generalizes the theory of Lie super-algebras and their (unitary) representations. Let $`F`$ be a positive integer and $`q=\mathrm{exp}(\frac{2i\pi }{F})`$. We consider a complex vector space $`S`$ together with a linear map $`\epsilon `$ from $`S`$ into itself satisfying $`\epsilon ^F=1`$. We set $`A_k=S_{q^k}`$ and $`B=S_1`$ (where $`S_\lambda `$ is the eigenspace corresponding to the eigenvalue $`\lambda `$ of $`\epsilon `$) so that $`S=B_{k=1}^{F1}A_k`$. The map $`\epsilon `$ is called the grading. If $`S`$ is endowed with the following structures we will say that $`S`$ is a fractional super Lie algebra ($`F`$-Lie algebra for short):
1. $`B`$ is a Lie algebra and $`A_k`$ is a representation of $`B`$.
2. There are multilinear, $`B`$equivariant (i.e. which respect the action of $`B`$) maps $`\{,\mathrm{},\}:𝒮^F\left(A_k\right)B`$ from $`𝒮^F\left(A_k\right)`$ into $`B`$. In other words, we assume that some of the elements of the Lie algebra $`B`$ can be expressed as $`F`$th order symmetric products of “more fundamental generators”. Here $`𝒮^F(D)`$ denotes the $`F`$fold symmetric product of $`D`$. It is then easy to see that:
$`\{\epsilon (a_1),\mathrm{},\epsilon (a_F)\}=\epsilon \left(\{a_1,\mathrm{},a_F\}\right),a_iA_k.`$ (2.1)
3. For $`b_iB`$ and $`a_jA_k`$ the following “Jacobi identities” hold:
$`[[b_1,b_2],b_3]+[[b_2,b_3],b_1]+[[b_3,b_1],b_2]=0`$
$`[[b_1,b_2],a_3]+[[b_2,a_3],b_1]+[[a_3,b_1],b_2]=0`$
$`[b,\{a_1,\mathrm{},a_F\}]=\{[b,a_1],\mathrm{},a_F\}+\mathrm{}+\{a_1,\mathrm{},[b,a_F]\}`$ (2.2)
$`{\displaystyle \underset{i=1}{\overset{F+1}{}}}[a_i,\{a_1,\mathrm{},a_{i1},a_{i+1},\mathrm{},a_{F+1}\}]=0.`$
The first identity is the usual Jacobi identity for Lie algebras, the second says that the $`A_k`$ are representation spaces of $`B`$ and the third is just the Leibniz rule (or the equivariance of $`\{,\mathrm{},\}`$). The fourth identity is the analogue of the graded Leibniz rule of Super-Lie algebras for $`F`$Lie algebras
If we want to be able to talk about unitarity, we also require the following additional struc-
ture and in this case, $`S`$ is called an $`F`$Lie algebra with adjoint.
4. A conjugate linear map $``$ from $`S`$ into itself such that:
$`\begin{array}{cc}\mathrm{a})\hfill & (s^{})^{}=s,sS\hfill \\ \mathrm{b})\hfill & [a,b]^{}=[b^{},a^{}]\hfill \\ \mathrm{c})\hfill & \epsilon (s^{})=\epsilon (s)^{}\hfill \\ \mathrm{d})\hfill & \{a_1,\mathrm{},a_F\}^{}=\{\left(a_1\right)^{},\mathrm{},\left(a_F\right)^{}\},aA_k.\hfill \end{array}`$ (2.7)
From a) and c) we see that for $`XB`$ we have $`X^{}B`$, and that for $`XA_k`$, we have $`X^{}A_{Fk}.`$
A unitary representation of an $`F`$Lie algebra with adjoint $`S`$ is a linear map $`\rho :S\mathrm{End}(H)`$, (where $`H`$ is a Hilbert space and $`\mathrm{End}(H)`$ the space of linear operators acting on $`H`$) and a unitary endomorphism $`\widehat{\epsilon }`$ such that $`\widehat{\epsilon }^F=1`$ which satisfy
$`\begin{array}{cc}\mathrm{a})\hfill & \rho \left([x,y]\right)=\rho (x)\rho (y)\rho (y)\rho (x)\hfill \\ \mathrm{b})\hfill & \rho \{a_1.\mathrm{},a_F\}=\frac{1}{F!}\underset{\sigma S_F}{}\rho \left(a_{\sigma (1)}\right)\mathrm{}\rho \left(a_{\sigma (F)}\right)\hfill \\ \mathrm{c})\hfill & \rho (s)^{}=\rho (s^{})\hfill \\ \mathrm{d})\hfill & \widehat{\epsilon }\rho \left(s\right)\widehat{\epsilon }^1=\rho \left(\epsilon \left(s\right)\right)\hfill \end{array}`$ (2.12)
($`S_F`$ being the group of permutations of $`F`$ elements). As a consequence of these properties, since the eigenvalues of $`\widehat{\epsilon }`$ are $`\mathrm{F}^{\mathrm{th}}`$ roots of unity, we have the following decomposition of the Hilbert space
$$H=\underset{k=0}{\overset{F1}{}}H_k,$$
where $`H_k=\left\{\right|hH:\widehat{\epsilon }|h=q^k|h\}`$. The operator $`N\mathrm{End}(H)`$ defined by $`N|h=k|h`$ if $`|hH_k`$ is the “number operator” (obviously $`q^N=\widehat{\epsilon }`$). Since $`\widehat{\epsilon }\rho (b)=\rho (b)\widehat{\epsilon },bB`$ each $`H_k`$ provides a representation of the Lie algebra $`B`$. Furthermore, for $`aA_{\mathrm{}}`$, $`\widehat{\epsilon }\rho (a)=q^{\mathrm{}}\rho (a)\widehat{\epsilon }`$ and so we have $`\rho (a).H_kH_{k+\mathrm{}(\mathrm{mod}F)}`$
Several remarks can be made at that point Firstly, for all $`k=1,\mathrm{},F1`$ it is clear that the subspace $`BA_k`$ of $`S`$ satisfies (2.1-3) and the subspace $`BA_kA_k`$ satisfies (2.1-2.7) (when $`S`$ has an adjoint). Secondly, it is important to notice that bracket $`\{\mathrm{}\}`$ is a priori not defined for elements in different gradings.
The basic idea to define a fractional supersymmetry is the following. Let $`g`$ be a Lie algebra and let $`𝒟,𝒟^{}`$ be representations of $`g`$. The representation $`𝒟^{}`$ is chosen in such a way that $`𝒮^F\left(𝒟^{}\right)`$ is related to $`𝒟`$ ($`𝒮^F\left(𝒟^{}\right)𝒟`$) in a sense specified later on. Then we consider $`B=g𝒟`$, a Lie algebra as the semi-direct product of $`g`$ and $`A_1=𝒟^{}`$. The relation $`𝒮^F\left(𝒟^{}\right)𝒟`$ will be the fundamental one to define an $`F`$Lie algebra.
## 3 Differential realization(s) of Lie algebras
We consider now $`g`$ a semi-simple Lie algebra of rank $`r`$. Let $`h`$ be a Cartan sub-algebra of $`g`$, let $`\mathrm{\Phi }h^{}`$ (the dual of $`h`$) be the corresponding set of roots and let $`f_\alpha `$ be the one dimensional root space associated to $`\alpha \mathrm{\Phi }`$. We chose a basis $`\{H_i,i=1,\mathrm{},r\}`$ of $`h`$ and elements $`E^\alpha f_\alpha `$ such that the commutation relations become
$`[H_i,H_j]`$ $`=`$ $`0`$
$`[H_i,E^\alpha ]`$ $`=`$ $`\alpha ^iE^\alpha `$ (3.1)
$`[E^\alpha ,E^\beta ]`$ $`=`$ $`\{\begin{array}{cc}ϵ\{\alpha ,\beta \}E^{\alpha +\beta }\hfill & \mathrm{if}\alpha +\beta \mathrm{\Phi }\hfill \\ \frac{2\alpha .H}{\alpha .\alpha }\hfill & \mathrm{if}\alpha +\beta =0\hfill \\ 0\hfill & \mathrm{otherwise}\hfill \end{array}`$ (3.5)
We now introduce $`\{\alpha _{(1)},\mathrm{},\alpha _{(r)}\}`$ a basis of simple roots. The weight lattice $`\mathrm{\Lambda }_W(g)h^{}`$ is the set of vectors $`\mu `$ such that $`\frac{2\alpha .\mu }{\alpha .\alpha }\text{ZZ}`$ and, as is well known, there is a basis of the weight lattice consisting of the fundamental weights $`\{\mu _{(1)},\mathrm{},\mu _{(r)}\}`$ defined by $`\frac{2\mu _{(i)}.\alpha _{(j)}}{\alpha _{(j)}.\alpha _{(j)}}=\delta _{ij}`$. A weight $`\mu =\underset{i=1}{\overset{r}{}}n_i\mu _{(i)}`$ is called dominant if all the $`n_i0`$ and it is well known that the set of dominant weights is in one to one correspondence with the set of (equivalence classes of) irreducible finite dimensional representations of $`g`$.
Recall briefly how one can associate a representation of $`g`$ to $`\mu h^{}`$. A given representation of $`g`$ can be defined from a highest weight states $`|\mu >`$ ($`E^\alpha |\mu >=0,\alpha >0,2\frac{\alpha _{(i)}.H}{\alpha ^2}|\mu >=n_i|\mu >),i=1,\mathrm{},r`$). We denote $`h_i=2\frac{\alpha _{(i)}.H}{\alpha ^2}`$. The space obtained from $`|\mu >`$ by the action of the element of $`g`$$`E^{\alpha _{(i_1)}}\mathrm{}E^{\alpha _{(i_k)}}|\mu >`$ clearly define a representation of $`g`$. This construction can be made more precise using the language of Verma module . This representation is denoted $`𝒟_\mu `$. To come back to our original problem, consider a finite dimensional irreducible representation $`𝒟_\mu `$. The basic idea to define an FSUSY associated to $`g𝒟_\mu `$ is to consider the infinite dimensional representation associated to the weight $`\mu /F`$. In , we defined an $`F`$Lie algebra associated to $`g,𝒟_\mu `$ and $`𝒟_{\mu /F}`$. Here, we reproduce these results in an explicit way using the differential realization of $`g`$.
As we have recalled the representations of $`g`$ are just specified by the weight $`\mu =\underset{i=1}{\overset{r}{}}n_i\mu _{(i)}`$. But, among the representations of $`g`$ there are $`r`$ basic representations. These representations are associated to the fundamental weights $`\mu =\mu _{(i)}`$, and all representations can be obtained from (symmetric) tensorial product of these basic representations. Furthermore, all basic representations can be derived from the antisymmetric product of the elementary representations, which are associated to the weight with terminal point in the Dynkin diagram . Consequently, if one is able to obtain a differential realization of the elementary representations of $`g`$ (at maximum $`3`$), one is able to construct all representations easily as we will see. Moreover, for $`su(n+1)=a_n,sp(2n)=c_n`$, we need to consider only the fundamental representation (related to $`\mu =\mu _{(1)}`$). Although for $`so(2n+1)=b_n`$ the vector ($`\mu =\mu _{(1)}`$) and the spinorial ($`\mu =\mu _{(n)}`$) representations and for $`so(2n)=d_n`$ the vector ($`\mu =\mu _{(1)}`$) and the two spinorial representations ($`\mu =\mu _{(n1)},\mu =\mu _{(n)}`$) reproduce all representations. For the exceptional Lie algebras some simplifications may happen from the embedding properties ($`e_8so(16),e_7so(12)su(2),e_6so(10)u(1)`$), and $`f_4,g_2`$ being of small rank calculations can be done easily. In the next subsection we just consider the series $`a_n,b_n,c_n`$ and $`d_n`$ and construct finite and infinite dimensional representations. It is important to emphasize that all the highest weight representations, finite and infinite dimensional, can be obtained in terms of homogeneous monomial of appropriate variables as a consequence of the differential realization of $`g`$. In the next section we apply these realizations to construct explicitly FSUSY.
### 3.1 $`su(n)`$
$`su(n)`$ is a rank $`n1`$ Lie algebra, but it is more convenient for our purpose to define roots as vectors of $`\text{IR }^n`$. Introduce $`e_i,i=1,\mathrm{},n`$ an orthonormal basis of $`\text{IR }^n`$ the simples roots of $`su(n)`$ reads
$$\alpha _{(i)}=e_ie_{i+1},i=1,\mathrm{},n1,$$
(3.6)
and the positive roots
$$\alpha =e_ie_j,1i<jn.$$
(3.7)
Now if we introduce<sup>2</sup><sup>2</sup>2 From now on, all the variables, whatever $`su(n),so(n)`$ or $`sp(2n)`$ are concerned, are positive and different from zero. $`(x_1,\mathrm{},x_n)(\text{IR }^+\{0\})^n`$ we can define the explicit realization of $`su(n)`$:
$`E^\alpha `$ $`=`$ $`E^{e_ie_j}=x_i_{x_j},1i<jn`$ (3.8)
$`h_i`$ $`=`$ $`x_i_{x_i}x_{i+1}_{x_{i+1}},i=1,\mathrm{},n1.`$
Furthermore, all the highest weight representations of $`su(n)`$ can be obtained from (3.8). Indeed, we can check using (3.8) that all the primitive vectors of the basic representations can be realized in terms of antisymmetric products of the $`x`$’s (in this notation we have $`|e_i>=x_i`$)
$`|\mu _{(1)}>`$ $`=`$ $`|e_1>=x_1,`$
$`|\mu _{(2)}>`$ $`=`$ $`|e_1+e_2>=x_1x_2,`$ (3.9)
$`\mathrm{}`$
$`|\mu _{(n1)}>`$ $`=`$ $`|e_1+\mathrm{}+e_{n1}>=x_1\mathrm{}x_{n1}.`$
And correspondingly the representation $`𝒟_\mu `$ associated to the weight $`\mu =\underset{i=1}{\overset{n1}{}}p_i\mu _{(i)}`$ with $`p_i\text{IR }`$ can be obtained from the highest weight $`|\mu >=\left(x_1\right)^{p_1}\mathrm{}\left(x_1\mathrm{}x_{n1}\right)^{p_{n1}}`$ acting with the operators $`E^{\alpha _{(i)}}`$ given in (3.8). When all the $`p_i\text{IN}`$ we obtain a finite dimensional representation of $`su(n)`$ otherwise the representation if infinite dimensional. Of course for $`su(n)`$ only the finite dimensional representations are unitary. Moreover, when $`p_i\text{IR }`$ there is no guaranty that the representation can be exponentiated, namely that they are representation of the Lie group $`SU(n)`$. Finally, notice that the normalizations in (3.8) and (3.1) are not the usual ones, but they are useful for to construct $`F`$Lie algebras. These last properties are also valid for all compact Lie algebras.
### 3.2 $`sp(2n)`$
Using, as for $`su(n)`$, the canonical basis of $`\text{IR }^n`$, $`e_i,i=1,\mathrm{},n`$ the simple roots of $`sp(2n)`$ are
$$\alpha _{(i)}=e_ie_{i+1},i=1,\mathrm{},n1,\alpha _{(n)}=2e_n,$$
(3.10)
$`sp(2n)`$ is not a simply-laced algebra (all roots do not have the same length). The positive roots are given by
$$e_i\pm e_j,1i<jn,2e_i,1in.$$
(3.11)
Constructing the 2n-dimensional representation associated to $`\mu _{(1)}`$ and introducing $`2n`$ variables corresponding to the weight of the representation $`𝒟_{\mu _{(1)}}`$ $`x_i=|e_i>,x_i=|e_i>`$ is is not difficult to define the operators $`E^{\pm \alpha _{(i)}}`$ and $`h_i`$:
$`E^{\alpha _{(i)}}`$ $`=`$ $`E^{e_ie_{i+1}}=x_i_{x_{i+1}}+x_{(i+1)}_{x_i},1in1`$
$`E^{\alpha _{(n)}}`$ $`=`$ $`E^{2e_n}=x_n_{x_n}`$
$`E^{\alpha _{(i)}}`$ $`=`$ $`E^{e_i+e_{i+1}}=x_{i+1}_{x_i}+x_i_{x_{(i+1)}},1in1`$ (3.12)
$`E^{\alpha _{(n)}}`$ $`=`$ $`E^{2e_n}=x_n_{x_n}`$
$`h_i`$ $`=`$ $`x_i_{x_i}x_{i+1}_{x_{i+1}}x_i_{x_i}+x_{(i+1)}_{x_{(i+1)}},i=1,\mathrm{},n1.`$
$`h_n`$ $`=`$ $`x_n_{x_n}x_n_{x_n}.`$
The notations has been chosen in such a way that comparing the weights of the variables $`x>0`$ with the weights of the generators of $`sp(2n)`$, the expression of the $`E^\alpha `$ can directly be read. This holds equally for $`su(n)`$ and $`so(n)`$. Then, the remaining generators can be calculated using the generators associated to the primitives roots. Noticing that $`E^{e_ie_j}`$ maps $`|e_i>|e_j>`$ and $`|e_j>|e_i>`$ we get $`E^{e_ie_j}=ax_i_{x_j}+bx_j_{x_i}`$. The coefficients $`a,b`$ can be determined by multiple commutators. Indeed, writing ($`j>i`$ ) $`j=i+k`$, we get $`e_ie_{i+k}=\alpha _{(i)}+\mathrm{}+\alpha _{(i+k1)}`$ and $`E^{e_ie_j}=[E^{\alpha _{(i+k1)}},\mathrm{}[E^{\alpha _{(i+1)}},E{}_{}{}^{\alpha _{(i)}}]\mathrm{}]`$. And similarly for the other generators. Now as for $`su(n)`$, all representation can be obtained from (3.2):
$`|\mu _{(1)}>`$ $`=`$ $`|e_1>=x_1,`$
$`|\mu _{(2)}>`$ $`=`$ $`|e_1+e_2>=x_1x_2,`$ (3.13)
$`\mathrm{}`$
$`|\mu _{(n1)}>`$ $`=`$ $`|e_1+\mathrm{}+e_{n1}>=x_1\mathrm{}x_{n1},`$
$`|\mu _{(n)}>`$ $`=`$ $`|e_1+\mathrm{}+e_n>=x_1\mathrm{}x_n.`$
To obtain all representations we proceed as for $`su(n)`$. Some remarks can be done at that point. For $`sp(2n)`$, $`\left(𝒟_{\mu _{(1)}}𝒟_{\mu _{(1)}}\right)_{\mathrm{anti}.\mathrm{sym}}=𝒟_{\mu _{(2)}}𝒟_0`$ ($`𝒟_0`$ being the scalar representation) as can be seen directly from (3.2) and (3.2). Similar results hold for $`𝒟_{\mu _{(i)}}`$.
### 3.3 $`so(2n+1)`$
This Lie algebra is the algebra dual of $`sp(2n)`$ and his simple roots are obtained from the simple roots of $`sp(2n)`$ ( $`(\alpha _{(i)})_{so(2n+1)}=2(\alpha _{(i)})_{sp(2n)}/((\alpha _{(i)})_{sp(2n)})^2`$). With the same notations as for $`su(n)`$ and $`sp(2n)`$ we have the simple roots
$$\alpha _{(i)}=e_ie_{i+1},i=1,\mathrm{},n1,\alpha _{(n)}=e_n.$$
(3.14)
The positive roots are given by
$$e_i\pm e_j,1i<jn,e_i,1in.$$
(3.15)
The vectorial representation (of dimension $`2n+1`$) allows to define the set of variables $`x_i=|e_i>,x_0=|0>`$ and $`x_i=|e_i>`$. Constructing explicitly the vectorial representation we obtain
$`E^{\alpha _{(i)}}`$ $`=`$ $`E^{e_ie_{i+1}}=x_i_{x_{i+1}}+x_{(i+1)}_{x_i},1in1`$
$`E^{\alpha _{(n)}}`$ $`=`$ $`E^{e_n}=x_n_{x_0}+x_0_{x_n}`$
$`E^{\alpha _{(i)}}`$ $`=`$ $`E^{e_i+e_{i+1}}=x_{i+1}_{x_i}+x_i_{x_{(i+1)}},1in1`$ (3.16)
$`E^{\alpha _{(n)}}`$ $`=`$ $`E^{e_n}=x_0_{x_n}+x_n_{x_0}`$
$`h_i`$ $`=`$ $`x_i_{x_i}x_{i+1}_{x_{i+1}}x_i_{x_i}+x_{(i+1)}_{x_{(i+1)}},i=1,\mathrm{},n1.`$
$`h_n`$ $`=`$ $`x_n_{x_n}x_n_{x_n}.`$
If consider the $`p`$froms as for $`su(n)`$ and $`sp(2n)`$ we observe that
$`|\mu _{(1)}>`$ $`=`$ $`|e_1>=x_1,`$
$`|\mu _{(2)}>`$ $`=`$ $`|e_1+e_2>=x_1x_2,`$ (3.17)
$`\mathrm{}`$
$`|\mu _{(n1)}>`$ $`=`$ $`|e_1+\mathrm{}+e_{n1}>=x_1\mathrm{}x_{n1},`$
$`|2\mu _{(n)}>`$ $`=`$ $`|e_1+\mathrm{}+e_n>=x_1\mathrm{}x_n,`$
and all representations, except the spinorial one $`D_{\mu _{(n)}}`$ can be obtained form (3.3) and (3.3). Then if we want to obtain all representations in terms of appropriate monomials another differential realization, associated to $`D_{\mu _{(n)}}`$ has to be constructed along the same lines (constructing explicitly the spinorial representation). However, if one consider the highest weight states $`|\mu _n>=(x_1\mathrm{}x_n)^{1/2}`$ one is able to define a primitive vector having all properties of the primitive vector of the spinorial representation. But, in such a representation the operators $`E^\alpha `$ are not nilpotent. This means that this representation is precisely a Verma module $`𝒱_{\mu _{(n)}}`$. Then $`𝒱_{\mu _{(n)}}`$ has a unique maximal proper sub-representation $`M_{\mu _{(n)}}`$ and the quotient $`𝒱_{\mu _{(n)}}/M_{\mu _{(n)}}`$ is $`𝒟_{\mu _{(n)}}`$ (see e.g. ) <sup>3</sup><sup>3</sup>3If we use (3.3) for $`n=1`$ i.e. for $`so(3)`$ and we consider $`(x_1)^{1/2}`$ as a primitive vector of the spinorial representation we can easily see that for any $`p>0`$ $`\left(E^{\alpha _{(1)}}\right)^p(x_1)^{1/2}0`$ but $`E^{\alpha _{(1)}}\left(E^{\alpha _{(1)}}\right)^2(x_1)^{1/2}=0`$. So the representation $`𝒱=\left\{\left(E^{\alpha _{(1)}}\right)^p(x_1)^{1/2},p0\right\}`$ is precisely a Verma module. But it can be easily seen that $`M=\left\{\left(E^{\alpha _{(1)}}\right)^p(x_1)^{1/2},p2\right\}`$ is the maximal sub-representation of $`𝒱`$ ($`mM,E^{\alpha _{(1)}}(m),E^{\alpha _{(1)}}(m),h_1(m)M`$) and then $`𝒟=𝒱/M`$ is the two-dimensional spinorial representation.. But, constructing explicitly the representation $`𝒱_{\mu _{(n)}}`$ and introducing at each step a new variable, $`y_1,\mathrm{},y_{2^n}`$, one is able to construct the differential realization of the spinorial representation. Indeed, for any representation $`𝒟_\mu `$ such a process can be equally applied and the differential realization of $`𝒟_\mu `$ can be reached straightforwardly. Of course this is also possible for $`su(n),sp(2n)`$ and $`so(2n)`$.
### 3.4 $`so(2n)`$
The case of $`so(2n)`$ is similar to $`so(2n+1)`$ but in this case we have two spinorial representations. With the same notations as before we introduce the set of primitive roots
$$\alpha _{(i)}=e_ie_{i+1},i=1,\mathrm{},n1,\alpha _{(n)}=e_{n1}+e_n,$$
(3.18)
and the positive roots are given by
$$e_i\pm e_j,1i<jn.$$
(3.19)
From the 2n-dimensional vector representation we obtain ($`x_i=|e_i>,x_i=|e_i>`$)
$`E^{\alpha _{(i)}}`$ $`=`$ $`E^{e_ie_{i+1}}=x_i_{x_{i+1}}+x_{(i+1)}_{x_i},1in1`$
$`E^{\alpha _{(n)}}`$ $`=`$ $`E^{e_{n1}+e_n}=x_{n1}_{x_n}+x_n_{x_{(n1)}}`$
$`E^{\alpha _{(i)}}`$ $`=`$ $`E^{e_i+e_{i+1}}=x_{i+1}_{x_i}+x_i_{x_{(i+1)}},1in1`$ (3.20)
$`E^{\alpha _{(n)}}`$ $`=`$ $`E^{e_{n1}e_n}=x_n_{x_{n1}}+x_{(n1)}_{x_n}`$
$`h_i`$ $`=`$ $`x_i_{x_i}x_{i+1}_{x_{i+1}}x_i_{x_i}+x_{(i+1)}_{x_{(i+1)}},i=1,\mathrm{},n1.`$
$`h_n`$ $`=`$ $`x_{n1}_{x_{n1}}+x_n_{x_n}x_n_{x_n}x_{(n+1)}_{x_{(n+1)}}.`$
And, as for $`so(2n+1)`$ we observe that the $`p`$forms are
$`|\mu _{(1)}>`$ $`=`$ $`|e_1>=x_1,`$
$`|\mu _{(2)}>`$ $`=`$ $`|e_1+e_2>=x_1x_2,`$ (3.21)
$`\mathrm{}`$
$`|\mu _{(n1)}+\mu _{(n)}>`$ $`=`$ $`|e_1+\mathrm{}+e_{n1}>=x_1\mathrm{}x_{n1},`$
$`|2\mu _{(n)}>`$ $`=`$ $`|e_1+\mathrm{}+e_n>=x_1\mathrm{}x_n.`$
Then two more differential realizations (the two spinorial) allow to construct all representations of $`so(2n)`$. However, if we use (3.4) for all representations of $`so(2n)`$, the spinorial representations will be infinite dimensional Verma modules.
To conclude this section recall once again that on the one hand the differential realization(s) considered for $`su(n),sp(2n),so(2n+1)`$ and $`so(2n)`$ are just a convenience and simply reproduce known results on Lie algebras. Of course similar expressions (in terms of bosonic or fermionic oscillators) are given in standard text-books (see e.g. ) but to my knowledge they were not used to construct representations in a systematic way from the fundamental one $`𝒟_{\mu _{(1)}}`$, except for $`su(2),sl(2,\text{IR })`$ and $`sl(2,\text{lC})`$ (see ). Indeed, the advantage of the differential realization (3.8),(3.2),(3.3) and (3.4) is to permit constructing explicitly the representations in terms of appropriate monomials. For such representations, finite or infinite dimensional, one, two or at maximum three realizations of the Lie algebra $`g`$ are just needed. In addition, acting with (3.8), (3.2), (3.3) or (3.4), on the states of the corresponding representation $`𝒟_\mu `$ (of dimension $`d`$), and introducing new variables $`y_1,\mathrm{},y_d`$ at each step, permits the construction of the differential realization of $`𝒟_\mu `$ directly form $`𝒟_{\mu _{(1)}}`$. These differential realizations are also convenient to obtain representations neither bounded from below nor from above (for instance, in the case of $`su(n)`$, the representation obtained with $`x_1^{p_1}\mathrm{}x_n^{p_n},p_i\text{IN}`$ has no primitive vector). On the other hand, they are very useful to construct FSUSY as shown in the next section.
## 4 Fractional supersymmetry and Lie algebras
For any Lie algebra $`g`$ and any representation $`𝒟_\mu `$, one is able to construct an associated FSUSY as it has been established in . One possible solution is to consider the infinite dimensional representation $`𝒟_{\mu /F}`$. At that point, the results of the previous section apply and allow to define an $`F`$Lie algebra associated to $`g,𝒟_\mu `$ and $`𝒟_{\mu /F}`$. Indeed, this explicit procedure has only be done for $`so(1,2)`$ . For higher rank Lie algebras this construction was obtained abstractly in terms of Verma modules . But now, we observe that the differential realization of $`so(1,2)`$ extends along the same lines for any Lie algebra. Then, it becomes possible to realize, as it has been done for $`so(1,2)`$, an $`F`$Lie algebra for any Lie algebra. This explicit construction being analogous for any $`g`$ we just reproduce it for the rank two Lie algebra $`su(3)`$ (weight diagrams and representations can be represented graphically) but we have to keep in mind that this procedure, to construct an $`F`$Lie algebra, is equally valid for any Lie algebra $`g`$.
Denote $`\alpha ,\beta `$ the simple roots of $`su(3)`$ and $`\gamma =\alpha +\beta `$ the third positive root. Introduce, as in section 3, $`x_1,x_2,x_3>0`$ of weight $`|\mu _{(1)}>=x_1,|\mu _{(1)}\alpha >=x_2,|\mu _{(1)}\alpha \beta >=x_3`$. The generators takes the form
$`\begin{array}{cc}E^\alpha =x_1_{x_2}\hfill & E^\alpha =x_2_{x_1}\hfill \\ E^\beta =x_2_{x_3}\hfill & E^\beta =x_3_{x_2}\hfill \\ E^\gamma =x_1_{x_3}\hfill & E^\gamma =x_3_{x_1}\hfill \end{array}`$ (4.4)
$`h_1=x_1_{x_1}x_2_{x_2}`$
$`h_2=x_2_{x_2}x_3_{x_3}.`$
We would like to construct an $`F`$Lie algebra associated to the three dimensional representation $`𝒟_{\mu _{(1)}}`$ (this could have been done for any representation of $`su(3)`$). In the realization (4.4) the vectorial representation writes
$$𝒟_{\mu _{(1)}}=\{\begin{array}{ccc}x_1\hfill & =|\mu _{(1)}>,\hfill & \\ x_2\hfill & =|\mu _{(1)}\alpha >\hfill & =E^\alpha |\mu _{(1)}>,\hfill \\ x_3\hfill & =|\mu _{(1)}\alpha \beta >\hfill & =E^\beta E^\alpha |\mu _{(1)}>.\hfill \end{array}$$
(4.5)
So, we consider
$$B=su(3)𝒟_{\mu _{(1)}}$$
(4.6)
for the bosonic (graded zero part) of the $`F`$Lie algebra. The natural representation to define the “$`\mathrm{F}^{\mathrm{th}}`$root” of $`𝒟_{\mu _{(1)}}`$ is $`𝒟_{\mu _{(1)}/F}`$. So, we take for the graded one part
$$A_1=𝒟_{{\scriptscriptstyle \frac{\mu _{(1)}}{F}}}$$
(4.7)
In the realization (4.4) this representation writes
$`𝒟_{{\scriptscriptstyle \frac{\mu _{(1)}}{F}}}=\{`$ $`|\frac{\mu _{(1)}}{F}n\alpha p\beta >=(x_1)^{1/F}\left(\frac{x_2}{x_1}\right)^n\left(\frac{x_3}{x_2}\right)^p`$
$`\left(E^\beta \right)^p\left(E^\alpha \right)^n|\frac{\mu _{(1)}}{F}>,n\text{IN},0pn\},`$
leading to the following weight diagram
To define an $`F`$Lie algebra associated to $`su(3)𝒟_{\mu _{(1)}}𝒟_{\mu _{(1)}/F}`$ we consider the representation (reducible)
$$𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)=\{_{i=1}^F(x_1)^{1/F}\left(\frac{x_2}{x_1}\right)^{n_i}\left(\frac{x_3}{x_2}\right)^{p_i},n_i\text{IN},0p_in_i\},$$
(4.9)
with $``$ the symmetric tensorial product.
In we compare $`x_1`$, the primitive vector of $`𝒟_{\mu _{(1)}}`$ with $`^F\left(x_1\right)^{1/F}`$ we observe that these two vectors, as primitive vectors, satisfy the same properties
$$\begin{array}{cc}h_1(x_1)=x_1,\hfill & h_1\left(^F\left(x_1\right)^{1/F}\right)=^F\left(x_1\right)^{1/F},\hfill \\ h_2(x_1)=0,\hfill & h_2\left(^F\left(x_1\right)^{1/F}\right)=0,\hfill \\ E^{\alpha ,\beta ,\gamma }(x_1)=0,\hfill & E^{\alpha ,\beta ,\gamma }\left(^F\left(x_1\right)^{1/F}\right)=0.\hfill \end{array}$$
(4.10)
But now, if we construct the representation from these primitives vectors, on the one hand we get $`𝒟_{\mu _{(1)}}`$ and on the other hand the infinite dimensional representation
$`^F\left(x_1\right)^{1/F}=\{`$ $`|\mu _{(1)}n\alpha p\beta >=`$
$`\left(E^\beta \right)^p\left(E^\alpha \right)^n\left(^F\left(x_1\right)^{1/F}\right),n\text{IN},0pn\}.`$
But a direct calculation show that the following relations hold
$`E^\alpha |\mu _{(1)}2\alpha >=0`$
$`E^\alpha |\mu _{(1)}2\alpha \beta >=0`$ (4.12)
$`E^\gamma |\mu _{(1)}2\alpha 2\beta >=0.`$
This means that $`𝒱_{\mu _{(1)}}=^F\left(x_1\right)^{1/F}`$ is a Verma module (the operator $`E^\alpha `$ is not nilpotent) and $`M=\{\left(E^\beta \right)^p\left(E^\alpha \right)^n\left(^F\left(x_1\right)^{1/F}\right),n\text{IN},0pn,(n,p)(0,0),(1,0),(1,1)\}`$ is the maximal proper sub-representation of $`𝒱_{\mu _{(1)}}`$ ($`Tsu(3),mM,T(m)M`$). Then $`𝒱_{\mu _{(1)}}/M`$ and $`𝒟_{\mu _{(1)}}`$ are isomorphic. Then, from $`𝒱_{\mu _{(1)}}𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)`$ one can find an injection $`i:𝒱_{\mu _{(1)}}𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)`$. Conversely, from $`𝒟_{\mu _{(1)}}𝒱_{\mu _{(1)}}/M`$ we can define a surjection $`\pi :𝒱_{\mu _{(1)}}𝒟_{\mu _{(1)}}`$. Consequently the following diagram
$$\begin{array}{ccccc}𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)& \stackrel{i}{}& 𝒱_{\mu _{(1)}}& \stackrel{\pi }{}& 𝒟_{\mu _{(1)}},\end{array}$$
shows that we cannot define an $`F`$Lie algebra in such a way (because we cannot find a mapping from $`𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)`$ into $`𝒟_{\mu _{(1)}}`$ as stated in property 2 of $`F`$Lie algebras, see section 2). Indeed, in FSUSY in $`1+2`$ dimensions was constructed along these lines but we did not have the structure of $`F`$Lie algebra.
To obtain an $`F`$Lie algebra some constraints have to be introduced. Following , we define $``$ the vector space of functions on $`x_1,x_2,x_3>0`$. The multiplication map $`m_n:\times \mathrm{}\times `$ given by $`m_n(f_1,\mathrm{},f_n)=f_1\mathrm{}f_n`$ is multilinear and totally symmetric. Hence, it induces a map $`\mu _F`$ from $`𝒮^F\left(\right)`$ into $``$. Restricting to $`𝒮^F\left(𝒟_{\mu _{(1)}}\right)`$ one gets
$`\mu _F:𝒮^F\left(𝒟_{\mu _{(1)}}\right)`$ $``$ $`𝒮_{\mathrm{red}}^F\left(𝒟_{\mu _{(1)}}\right)`$
$`_{i=1}^F(x_1)^{1/F}\left(\frac{x_2}{x_1}\right)^{n_i}\left(\frac{x_3}{x_2}\right)^{p_i}`$ $``$ $`x_1\left(\frac{x_2}{x_1}\right)^{\underset{i=1}{\overset{F}{}}n_i}\left(\frac{x_3}{x_2}\right)^{\underset{i=1}{\overset{F}{}}p_i}.`$ (4.13)
We observe that $`𝒮_{\mathrm{red}}^F\left(𝒟_{\mu _{(1)}}\right)=\{x_1\left(\frac{x_2}{x_1}\right)^n\left(\frac{x_3}{x_2}\right)^p,n\text{IN},0pn\}𝒟_{\mu _{(1)}}`$, meaning that we can find an injection $`i^{}`$ from $`𝒟_{\mu _{(1)}}`$ into $`𝒮_{\mathrm{red}}^F\left(𝒟_{\mu _{(1)}}\right)`$. This representation is reducible but indecomposable. Namely, we cannot find a complement of $`𝒟_{\mu _{(1)}}`$ in $`𝒮_{\mathrm{red}}^F\left(𝒟_{\mu _{(1)}}\right)`$ stable under $`su(3)`$. For instance $`x_1\left(\frac{x_2}{x_1}\right)^2`$ is such that $`E^\alpha \left(x_1\left(\frac{x_2}{x_1}\right)^2\right)=2x_2`$, but $`E^\alpha (x_2)=0`$. As before we observe that the diagram
$$\begin{array}{ccccc}𝒟_{\mu _{(1)}}& \stackrel{i^{}}{}& 𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)_{\mathrm{red}}& \stackrel{\mu _F}{}& 𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)\end{array}$$
leads to the same conclusion on the structure of $`F`$Lie algebra. With these simple observations we can conclude as in that we cannot define an $`F`$Lie algebra with $`B=su(3)𝒟_{\mu _{(1)}}`$. To obtain such a structure, one possible solution is to extend $`𝒟_{\mu _{(1)}}`$ into an infinite dimensional representation. For instance,
$$\left(su(3)𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)_{\mathrm{red}}\right)𝒟_{\mu _{(1)}/F}$$
(4.14)
has a structure of $`F`$Lie algebra (a similar structure could have been defined with $`(su(3)`$ $`𝒱_{\mu _{(1)}})𝒟_{\mu _{(1)}/F}`$).
The problem to construct an $`F`$Lie algebra associated to $`su(3),𝒟_{\mu _{(1)}},𝒟_{\mu _{(1)}/F}`$ is basically related to the fact that we would like to relate a finite dimensional representation $`𝒟_{\mu _{(1)}}`$ with an infinite dimensional one $`𝒟_{\mu _{(1)}/F}`$. One possible solution, as we just have seen, is to extend $`𝒟_{\mu _{(1)}}`$ into a infinite dimensional (reducible but indecomposable) representation $`𝒮^F\left(𝒟_{\mu _{(1)}/F}\right)_{\mathrm{red}}`$. This procedure being quite general, works similarly for any Lie algebra $`g`$.
## 5 Conclusion
Under some assumptions symmetries beyond supersymmetry can be constructed. Fractional supersymmetry and $`F`$Lie algebras are a possible solutions. In this note, we have shown that differential realization(s) of Lie algebras is(are) an useful tool(s) to construct explicitly a structure of $`F`$Lie algebra associated to any representation $`𝒟_\mu `$ of any Lie algebra $`g`$. The basic point is to consider the infinite dimensional representation $`𝒟_{\mu /F}`$. We have shown, that considering an infinite dimensional (reducible but indecomposable) representation extending $`𝒟_\mu `$ enables us to construct an $`F`$Lie algebra (this solves the problem related to the fact that, in general, $`𝒟_\mu `$ is finite dimensional, although $`𝒟_{\mu /F}`$ is infinite dimensional). Another possible solution, if to extend the Lie algebra $`g`$ into a infinite dimensional Lie algebra. When $`g=so(1,2)`$ this algebra reduces to the centerless Virasoro algebra . Another advantage of the differential realization of section 3 is the possibility to construct explicitly, in a differential way, this infinite dimensional algebra (this will be done elsewhere, at least for $`su(n)`$).
Finally, we would like to conclude that for $`g=so(1,2)`$ unitary representations have been constructed . It has been observed that it is a symmetry which acts on relativistic anyons . The question of the interpretation of FSUSY in higher dimensional space-time is still open.
Acknowledgements: J’exprime mes plus vifs remerciements aux organisateurs du Réseau de Physique Théorique du Maroc, et en particulier à A. El Hassouni et E. H. Saidi, pour leur invitation à participer au workshop Non Commutative Geometry and Superstring Theory, 16 et 17 Juin 2000, Rabat.
|
warning/0007/math0007164.html
|
ar5iv
|
text
|
# Dimensions of Prym Varieties
## 1. Introduction
The most familiar Prym variety arises from a (possibly branched) double cover $`\pi :XY`$ of curves. In this situation, there is a surjective norm map $`\mathrm{Nm}:\mathrm{Jac}(X)\mathrm{Jac}(Y)`$, and the Prym (another abelian variety) is a connected component of its kernel. Another way to think of this is that the involution $`\sigma `$ of the double cover induces an action of $`/2`$ on the vector space $`H^0(X,\omega _X)`$, which can then be decomposed as a representation of $`/2`$. The Jacobian of the base curve Y and the Prym correspond to the trivial and sign representations, respectively. The Prym variety can be defined as the component containing the identity of $`(\mathrm{Jac}(X)_{}\epsilon )^\sigma `$, where $`\epsilon `$ denotes the sign representation of $`/2`$.
The generalization of this construction that we will study in this paper is as follows. Let $`G`$ be a finite group, and $`\pi :XY`$ be a tame Galois branched cover, with Galois group $`G`$, of smooth projective curves over an algebraically closed field. The action of $`G`$ on $`X`$ induces an action on the vector space of differentials $`H^0(X,\omega _X)`$, and on the Jacobian $`\mathrm{Jac}(X)`$. For any representation $`\rho `$ of $`G`$, we define $`\mathrm{Prym}_\rho (X)`$ to be the connected component containing the identity of $`(\mathrm{Jac}(X)_{}\rho ^{})^G`$. The vector space $`H^0(X,\omega _X)`$ will decompose as a $`[G]`$-module into a direct sum of isotypic pieces
(1)
$$H^0(X,\omega _X)=\underset{j=1}{\overset{N}{}}\rho _jV_j$$
where $`\rho _1,\mathrm{},\rho _N`$ are the irreducible representations of $`G`$. If $`G`$ is such that all of its representations are rational, then the Jacobian will also decompose, up to isogeny, into a direct sum of Pryms \[D2\]:
(2)
$$\mathrm{Jac}(X)\underset{j=1}{\overset{N}{}}\rho _j\mathrm{Prym}_{\rho _j}(X).$$
In particular, if $`G`$ is the Weyl group of a semisimple Lie algebra, then it will satisfy this property.
The goal of this paper is to compute the dimension of such a Prym variety. This formula is given in section 2, with a proof that uses only the Riemann-Hurwitz theorem and some character theory. Special cases of this formula relevant to integrable systems have appeared previously \[A, Me, S, MS\].
One motivation for this work comes from the study of algebraically integrable systems. An algebraically integrable system is a Hamiltonian system of ordinary differential equations, where the phase space is an algebraic variety with an algebraic (holomorphic, over $``$) symplectic structure. The complete integrability of the system means that there are $`n`$ commuting Hamiltonian functions on the $`2n`$-dimensional phase space. For an algebraically integrable system, these functions should be algebraic, in which case they define a morphism to an $`n`$-dimensional space of states for the system. The flow of the system will be linearized on the fibers of this morphism, which, if they are compact, will be $`n`$-dimensional abelian varieties.
Many such systems can be solved by expressing the system as a Lax pair depending on a parameter $`z`$. The equations can be written in the form $`\frac{d}{dt}A=[A,B]`$, where $`A`$ and $`B`$ are elements of a Lie algebra $`𝔤`$, and depend both on time $`t`$ and on a parameter $`z`$, which is thought of as a coordinate on a curve $`Y`$. In this case, the flow of the system is linearized on a subtorus of the Jacobian of a Galois cover of $`Y`$. If it can be shown that this subtorus is isogenous to a Prym of the correct dimension, then the system is completely integrable.
In section 3, we will briefly discuss two examples of such systems, the periodic Toda lattice and Hitchin systems. Both of these are important in Seiberg-Witten theory, providing solutions to $`𝒩=2`$ supersymmetric Yang-Mills gauge theory in four dimensions.
This work appeared as part of a Ph.D. thesis at the University of Pennsylvania. The author would like to thank her thesis advisor, Ron Donagi, for suggesting this project and for many helpful discussions. Thanks are also due to David Harbater, Eyal Markman, and Leon Takhtajan.
## 2. Dimensions
We can start by using the Riemann-Hurwitz formula to find the genus $`g_X`$ of $`X`$, which will be the dimension of the whole space $`H^0(X,\omega _X)`$ and of $`\mathrm{Jac}(X)`$. Since $`\pi :XY`$ is a cover of degree $`|G|`$, we get
(3)
$$g_X=1+|G|(g1)+\frac{\mathrm{deg}R}{2}$$
where $`g`$ is the genus of the base curve $`Y`$ and $`R`$ is the ramification divisor.
The first isotypic piece we can find the dimension of is $`V_1`$, corresponding to the trivial representation. The subspace where $`G`$ acts trivially is the subspace of differentials which are pullbacks by $`\pi `$ of differentials on $`Y`$. This tells us that $`dimV_1=dimH^0(Y,\omega _Y)=g`$.
In the case of classical Pryms, where $`G=/2`$, there is only one other isotypic piece, $`V_\epsilon `$ corresponding to the sign representation $`\epsilon `$. Thus we have
(4)
$$dimV_\epsilon =g_Xg=g1+\frac{\mathrm{deg}R}{2}.$$
For larger groups $`G`$, there are more isotypic pieces, but we also have more information: we can look at intermediate curves, i.e. quotients of $`X`$ by subgroups $`H`$ of $`G`$. Differentials on $`X/H`$ pull back to differentials on $`X`$ where $`H`$ acts trivially. Thus
(5)
$$H^0(X/H,\omega _{X/H})=\underset{j=1}{\overset{N}{}}(\rho _j)^HV_j.$$
The map $`\pi _H:X/HY`$ will be a cover of degree $`\frac{|G|}{|H|}`$, so Riemann-Hurwitz gives us the following formula for the genus $`g_H`$ of $`X/H`$, which is the dimension of $`H^0(X/H,\omega _{X/H})`$:
(6)
$$g_H=1+\frac{|G|}{|H|}(g1)+\frac{\mathrm{deg}R_H}{2}.$$
where again $`R_H`$ is the ramification divisor.
We can further analyze the ramification divisor, by classifying the branch points according to their inertial groups. Since $`\pi :XY`$ is a Galois cover of curves over $``$, all of the inertial groups must be cyclic.
###### Lemma 1.
Let $`G`$ be a finite group all of whose characters are defined over $``$. If two elements $`x,yG`$ generate conjugate cyclic subgroups, then they are conjugate.
Proof (adapted from \[BZ\]): We want to show that for any character $`\chi `$ of $`G`$, $`\chi (x)=\chi (y)`$. Then the properties of characters will tell us that $`x`$ and $`y`$ must be in the same conjugacy class.
We may assume that $`x`$ and $`y`$ generate the same subgroup, $`H`$. Then $`y=x^k`$ for some integer $`k`$ relatively prime to $`|H|`$. Let $`\chi `$ be a character of $`G`$, and $`\rho :GGL(n,)`$ a representation with character $`\chi `$. Then $`\rho (x)`$ will be a matrix with eigenvalues $`\lambda _1,\mathrm{},\lambda _n`$, and $`\rho (y)`$ will have eigenvalues $`\lambda _1^k,\mathrm{},\lambda _n^k`$. Since $`x^{|H|}=1`$, we have $`\lambda _1^{|H|}=\mathrm{}=\lambda _n^{|H|}=1`$. Let $`\xi `$ be a primitive $`|H|`$th root of unity. Then we can write $`\lambda _1=\xi ^{\nu _1},\mathrm{},\lambda _n=\xi ^{\nu _n}`$ for some integers $`\nu _i`$. Now $`\chi (x)`$ = Trace($`\rho (x)`$) = $`\lambda _1+\mathrm{}+\lambda _n`$, and $`\chi (y)`$ = $`\chi (x^k)`$ = $`\lambda _1^k+\mathrm{}+\lambda _n^k`$. Thus $`\chi (y)`$ will be the image of $`\chi (x)`$ under the element of Gal($`(\xi )/`$) which sends $`\xi \xi ^k`$. Since the values of $`\chi `$ are rational, this element will act trivially, so $`\chi (y)=\chi (x)`$. $`\mathrm{}`$
From now on, we will suppose that $`G`$ is such that all of its characters are rational. (This will be true, for instance, if $`G`$ is a Weyl group). Pick representative elements $`h_1\mathrm{}h_N`$ for each conjugacy class in $`G`$, and let $`H_1\mathrm{}H_N`$ be the cyclic groups that each of them generates. By Lemma 1, this will be the whole set (up to conjugacy) of cyclic subgroups of $`G`$. We can partially order this set of cyclic subgroups by their size, so that $`H_1`$ is the trivial subgroup. Now we can classify the branch points: let $`R_k,k=2\mathrm{}N`$ be the degree of the branch locus with inertial group conjugate to $`H_k`$ (ignoring the trivial group). Over each point of the branch locus where the inertial group is conjugate to $`H_k`$, there will be $`|G|/|H_k|`$ points in the fiber. Thus the degree of the ramification divisor $`R`$ of $`\pi :XY`$ will be
(7)
$$\mathrm{deg}R=\underset{k=1}{\overset{N}{}}(|G|\frac{|G|}{|H_k|})R_k$$
For each quotient curve $`X/H`$, each point in the fiber of $`\pi _H:X/HY`$ over a point with inertial group $`H_k`$ will correspond to a double coset $`H_k\backslash G/H`$. Thus the degree of the ramification divisor $`R_H`$ will be
(8)
$$\mathrm{deg}R_H=\underset{k=1}{\overset{N}{}}(\frac{|G|}{|H|}\mathrm{\#}(H_k\backslash G/H))R_k.$$
Combining these formulas with the earlier Riemann-Hurwitz computations, we get:
(9)
$$g_X=1+|G|(g1)+\underset{k}{}(|G|\frac{|G|}{|H|})\frac{R_k}{2}$$
(10)
$$g_H=1+\frac{|G|}{|H|}(g1)+\underset{k}{}(\frac{|G|}{|H|}\mathrm{\#}(H_k\backslash G/H))\frac{R_k}{2}$$
Since the genera $`g_H`$ are exactly the dimensions $`dimH^0(X/H,\omega _{X/H})`$, we also have
(11)
$$g_H=\underset{j=1}{\overset{N}{}}dim\rho _j^HdimV_j.$$
For each subgroup $`H`$, this is a linear equation for the unknown dimensions $`dimV_j`$ in terms of the genus $`g_H`$. Thus by taking quotients by the set of all cyclic subgroups $`H_1\mathrm{}H_N`$, we get a system of $`N`$ equations. We wish to invert the matrix $`dim\rho _j^{H_i}`$ and find the $`N`$ unknowns $`dimV_j`$.
###### Lemma 2.
The matrix $`dim\rho _j^{H_i}`$ is invertible.
Proof: We show that the rows of the matrix are linearly independent, using the fact that rows of the character table are linearly independent. First, note that $`dim\rho _j^{H_i}`$, the dimension of the subspace of $`\rho _j`$ invariant under $`H_i`$, is equal to the inner product of characters $`\mathrm{Res}_{H_i}^G\rho _j,\mathrm{𝟏}`$, which we can read off from the character table of $`G`$ as
(12)
$$dim\rho _j^{H_i}=\frac{1}{|H_i|}\underset{a_iH_i}{}\chi _{\rho _j}(a_i).$$
Compare this matrix to the matrix of the character table $`\chi _{\rho _j}(a_i)`$. From (12) we see that each row is a sum of multiples of rows of the character table. Since each element of a subgroup has order less than or equal to the order of the subgroup, the rows of the character table being added to get row $`i`$ appear at or below row $`i`$ in the character table. Thus if we write the matrix $`dim\rho _j^{H_i}`$ in terms of the basis of the character table, we will get a lower triangular matrix with non-zero entries on the diagonal. By row reduction, we see that the linear independence of the rows of $`dim\rho _j^{H_i}`$ is equivalent to the linear independence of the rows of the character table. $`\mathrm{}`$
###### Theorem 1.
For each nontrivial irreducible representation $`\rho _j`$ of $`G`$, $`V_j`$ has dimension
(13)
$$(dim\rho _j)(g1)+\underset{k=1}{\overset{N}{}}\left((dim\rho _j)(dim\rho _j^{H_k})\right)\frac{R_{H_k}}{2}$$
Proof: Since the matrix $`dim\rho _j^{H_i}`$ is invertible, there is a unique solution to the system of equations (11), so we only need to show that this is a solution. Namely, given this formula for $`dimV_j`$, and combining (10) and (11), we wish to show that for each cyclic subgroup $`H_i`$,
(14)
$$\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}dimV_j=1+\frac{|G|}{|H_i|}(g1)+\underset{k}{}(\frac{|G|}{|H_i|}\mathrm{\#}(H_k\backslash G/H_i))\frac{R_k}{2}.$$
Note that on the left side we are summing over all representations, not just the nontrivial ones, so our notation will be simpler if we write $`dimV_1=g`$ in a similar form to (11). For the trivial representation $`\rho _1`$, $`(dim\rho _1)(dim\rho _1^{H_k})=0`$ (since $`\rho _1`$ is fixed by any subgroup $`H_k`$), so
(15)
$$dimV_1=1+(dim\rho _1)(g1)+\underset{k=1}{\overset{N}{}}\left((dim\rho _1)(dim\rho _1^{H_k})\right)\frac{R_{H_k}}{2}.$$
The sum on the left hand side of (14) will be
(16)
$$1+\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}\left((dim\rho _j)(g1)+\underset{k=1}{\overset{N}{}}\left((dim\rho _j)(dim\rho _j^{H_k})\right)\frac{R_{H_k}}{2}\right).$$
Let us look at the $`(g1)`$ term and the $`R_{H_k}`$ terms separately. For the $`(g1)`$ coefficient, we can write both $`dim\rho _j^{H_i}`$ and $`dim\rho _j`$ in terms of characters of $`G`$ (as in (12)) and exchange the order of summation to get
(17)
$$\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}dim\rho _j=\frac{1}{|H_i|}\underset{a_iH_i}{}\underset{j=1}{\overset{N}{}}\chi _{\rho _j}(a_i)\chi _{\rho _j}(e)$$
where $`e`$ is the identity element of $`G`$. The inner sum amounts to taking the inner product of two columns of the character table of $`G`$. The orthogonality of characters tells us that this inner product will be zero unless the two columns are the same, in this case if $`a_i=e`$. Thus the sum over elements in $`H_i`$ disappears, and we get the sum of the squares of the dimensions of the characters:
(18)
$$\frac{1}{|H_i|}\underset{j=1}{\overset{N}{}}\chi _{\rho _j}(e)^2=\frac{|G|}{|H_i|}.$$
which is what we want.
The $`R_{H_k}`$ term looks like
(19)
$$\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}\underset{k=1}{\overset{N}{}}\left((dim\rho _j)(dim\rho _j^{H_k})\right)\frac{R_{H_k}}{2}.$$
We can distribute and rearrange the sums to get:
(20)
$$\underset{k=1}{\overset{N}{}}\left(\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}dim\rho _j\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}dim\rho _j^{H_k}\right)\frac{R_{H_k}}{2}.$$
As in (17) and (18), the first term becomes $`\frac{|G|}{|H_i|}`$. The second term is also the inner product of columns of the character table:
(21)
$$\underset{j=1}{\overset{N}{}}dim\rho _j^{H_i}dim\rho _j^{H_k}=\frac{1}{|H_i|}\frac{1}{|H_k|}\underset{a_iH_i}{}\underset{a_kH_k}{}\underset{j=1}{\overset{N}{}}\chi _{\rho _j}(a_i)\chi _{\rho _j}(a_k).$$
This will be zero unless $`a_i`$ and $`a_k`$ are conjugate, in which case $`\chi _{\rho _j}(a_i)=\chi _{\rho _j}(a_k)`$ and character theory tells us (see for example \[FH\], p. 18) that
(22)
$$\underset{j=1}{\overset{N}{}}\chi _{\rho _j}(a_i)^2=\frac{|G|}{c(a_i)},$$
where $`c(a_i)`$ is the number of elements in the conjugacy class of $`a_i`$. Now the second term has become
(23)
$$\frac{|G|}{|H_i||H_k|}\underset{\{a_i,a_k\}}{}\frac{1}{c(a_i)}$$
where the sum is taken over pairs of elements $`a_iH_i,a_kH_k`$ such that $`a_i`$ and $`a_k`$ are conjugate. This is exactly the number of double cosets $`\mathrm{\#}(H_k\backslash G/H_i)`$.
Adding up all of the terms, the sum on the left hand side becomes
(24)
$$1+\frac{|G|}{|H_i|}(g1)+(\frac{|G|}{|H_i|}\mathrm{\#}(H_k\backslash G/H_i))\frac{R_{H_k}}{2}$$
which is exactly the right hand side. $`\mathrm{}`$
###### Corollary 1.
For each nontrivial irreducible representation $`\rho _j`$ of $`G`$, $`\mathrm{Prym}_{\rho _j}(X)`$ has dimension
(25)
$$(dim\rho _j)(g1)+\underset{k=1}{\overset{N}{}}\left((dim\rho _j)(dim\rho _j^{H_k})\right)\frac{R_{H_k}}{2}.$$
$`\mathrm{}`$
## 3. Integrable Systems.
Periodic Toda lattice.
The periodic Toda system is a Hamiltonian system of differential equations with Hamiltonian
$$H(p,q)=\frac{|p|^2}{2}+\underset{\alpha }{}e^{\alpha (q)}$$
where $`p`$ and $`q`$ are elements of the Cartan subalgebra $`𝔱`$ of a semisimple Lie algebra $`𝔤`$, and the sum is over the simple roots of $`𝔤`$ plus the highest root. This system can be expressed in Lax form \[AvM\] $`\frac{d}{dt}A=[A,B]`$, where $`A`$ and $`B`$ are elements of the loop algebra $`𝔤^{(1)}`$. and can be thought of as elements of $`𝔤`$ which depend on a parameter $`z^1`$. For $`𝔰𝔩(n)`$, $`A`$ is of the form
$$\left(\begin{array}{cccc}y_1& 1& & x_0z\\ x_1& y_2& \mathrm{}& \\ & \mathrm{}& \mathrm{}& 1\\ z& & x_{n1}& y_n\end{array}\right)$$
For any representation $`\varrho `$ of $`𝔤`$, the spectral curve $`S_\varrho `$ defined by the equation $`det(\varrho (A(z)\lambda I)=0`$ is independent of time (i.e. is a conserved quantity of the system). The spectral curve is a finite cover of $`Y`$ which for generic $`z`$ parametrizes the eigenvalues of $`\varrho (A(z))`$. While the eigenvalues are conserved by the system, the eigenvectors are not. The eigenvectors of $`\varrho (A)`$ determine a line bundle on the spectral cover, so an element of $`\mathrm{Jac}(S_\varrho )`$. The flow of the system is linearized on this Jacobean. Since the original system of equations didn’t depend on a choice of representation $`\varrho `$, the flow is actually linearized on an abelian variety which is a subvariety of $`\mathrm{Jac}(S_\varrho )`$ for every $`\varrho `$.
In fact, instead of considering each spectral cover we can look at the cameral cover $`X^1`$. This is constructed as a pullback to $`^1`$ of the cover $`𝔱𝔱/G`$, where $`G`$ is the Weyl group of $`𝔤`$. This cover is pulled back by the rational map $`^1𝔱/G`$ defined by the class of $`A(z)`$ under the adjoint action of the corresponding Lie group. (For $`A(z)`$ a regular semisimple element of $`𝔰𝔩(n)`$, this map sends $`z`$ to the unordered set of eigenvalues of $`A(z)`$.) Thus the cameral cover is a finite Galois cover of $`^1`$ whose Galois group $`G`$ is the Weyl group of $`𝔤`$. The flow of the Toda system is linearized on the Prym of this cover corresponding to the representation of $`G`$ on $`𝔱^{}`$. This is an $`r`$-dimensional representation, where $`r`$ is the rank, so the dimension of this Prym is
$$r(1)+\underset{k=1}{\overset{N}{}}\left(r(dim𝔱^{H_k})\right)\frac{R_{H_k}}{2}.$$
The ramification of this cover has been analyzed in \[D1\] and \[MS\]. There are $`2r`$ branch points where the inertial group $`H`$ is $`/2`$ generated by one reflection, so for each of these $`dim𝔱^H`$ is $`r1`$. There are also two points ($`z=0`$ and $`\mathrm{}`$) where the inertial group $`H`$ is generated by the Coxeter element, the product of the reflections corresponding to the simple roots. This element of $`G`$ doesn’t fix any element of $`𝔱`$, so for these two points $`dim𝔱^H=0`$. Thus the dimension of the Prym is
$`r+(r(r1)){\displaystyle \frac{2r}{2}}+(r0){\displaystyle \frac{2}{2}}`$
$`=r.`$
Since the original system of equations had a $`2r`$-dimensional phase space, this is the answer that we want.
Hitchin systems. Hitchin showed \[H\] that the cotangent bundle to the moduli space of semistable vector bundles on a curve $`Y`$ has the structure of an algebraically completely integrable system. His proof, later extended to principal $`𝒢`$ bundles with any reductive Lie group $`𝒢`$ \[F,S\], uses the fact that this moduli space is equivalent (by deformation theory) to the space of *Higgs pairs*, pairs $`(P,\varphi )`$ of a principal bundle and an endomorphism $`\varphi H^0(Y,ad(P)\omega _Y)`$. As in the case of the Toda system, the key construction is of a cameral cover of $`Y`$. The eigenvalues of $`\varphi `$, which are sections of the line bundle $`\omega _Y`$, determine a spectral cover of $`Y`$ in the total space bundle. The eigenvectors determine a line bundle on this spectral cover. The Hitchin map sends a Higgs pair $`(P,\varphi )`$ to the set of coefficients of the characteristic polynomial. Each coefficient is a section of a power of $`\omega _Y`$, so the image of the Hitchin map is $`B:=_{i=1}^rH^0(Y,\omega _Y^{d_i})`$, where the $`d_i`$ are the degrees of the basic invariant polynomials of the Lie algebra $`𝔤`$.
Again, we can consider instead the cameral cover $`X_bY`$, which is obtained as a pullback to $`Y`$ vi $`\varphi `$ of $`𝔱\omega _Y𝔱\omega _Y/G`$. The generic fiber of the Hitchin map is isogenous to $`\mathrm{Prym}_𝕥(X)`$, which has dimension
$$r(g1)+\underset{k=1}{\overset{N}{}}\left(r(dim𝔱^{H_k})\right)\frac{R_{H_k}}{2}$$
By looking at the generic fiber, we can restrict our attention to cameral covers where the only ramification is of order two, with inertial group $`H`$ generated by one reflection. The last piece of information we need to compute the dimension is the degree of the branch divisor of $`XY`$.
The cover $`𝔱\omega _Y𝔱\omega _Y/G`$ is ramified where any of the roots, or their product, is equal to zero. There are $`(dim𝒢r)`$ roots, so this defines a hypersurface of degree $`(dim𝒢r)`$ in the total space of $`\omega _Y`$. The ramification divisor of $`XY`$ is the intersection of this hypersurface with the section $`\varphi `$, which is the divisor corresponding to the line bundle $`\omega _Y^{(dim𝒢r)}.`$ Thus the degree of the branch divisor will be $`(dim𝒢r)(2g2)`$.
Combining all of this information, we see that the dimension of the Prym is
$`dim\mathrm{Prym}_𝕥(X)`$ $`=`$ $`r(g1)+(r(r1)){\displaystyle \frac{(dim𝒢r)(2g2)}{2}}`$
$`=`$ $`r(g1)+(dim𝒢r)(g1)`$
$`=`$ $`dim𝒢(g1).`$
By comparison, the dimension of the base space is
$$\mathrm{\Sigma }_{i=1}^rh^0(Y,\omega _Y^{d_i})$$
The sum of the degrees $`d_i`$ of the basic invariant polynomials of $`𝔤`$ is the dimension of a Borel subalgebra, $`(dim𝒢+r)/2`$. For $`g>1`$, Riemann-Roch gives
$`\mathrm{\Sigma }_{i=1}^rh^0(Y,\omega _Y^{d_i})`$ $`=`$ $`\mathrm{\Sigma }_{i=1}^r(2d_i1)(g1)`$
$`=`$ $`(dim𝒢+rr)(g1)`$
$`=`$ $`dim𝒢(g1).`$
Which, as Hitchin said, “somewhat miraculously” turns out to be the same thing.
Markman \[Ma\] and Bottacin \[B\] generalized the Hitchin system by twisting the line bundle $`\omega _Y`$ by an effective divisor $`D`$. The effect of this is to create a family of integrable systems, parametrized by the residue of the Higgs field $`\varphi `$ at $`D`$. The base space of each system is a fiber of the map
$$B:=\underset{i=1}{\overset{r}{}}H^0(Y,\omega _Y(D)^{d_i})$$
$$$$
$$\overline{B}:=\text{the space of possible residues at }D$$
which sends the set of $`r`$ sections in $`B`$ to its set of residues at $`D`$. At each point of $`D`$, there are $`r`$ independent coefficients, so the dimension of $`\overline{B}`$ is $`r(\mathrm{deg}D)`$. Thus the base space of each system has dimension
$`dimBdim\overline{B}`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{r}{}}}h^0(Y,\omega _Y(D)^{d_i})r(\mathrm{deg}D)`$
$`=`$ $`{\displaystyle \underset{i=1}{\overset{r}{}}}(d_i(2g2+\mathrm{deg}D)(g1))r(\mathrm{deg}D)`$
$`=`$ $`(1/2)(dim𝒢+r)(2g2+\mathrm{deg}D)r(g1)r(\mathrm{deg}D)`$
$`=`$ $`(dim𝒢)(g1)+{\displaystyle \frac{dim𝒢r}{2}}\mathrm{deg}D`$
Markman showed that the generic fiber of this system is again isogenous to $`\mathrm{Prym}_𝔱(X)`$, where $`X`$ is a cameral cover of the base curve $`Y`$. The construction of the cameral cover is similar to the case of the Hitchin system, except that $`\varphi `$ is a section of $`ad(P)\omega _Y(D)`$. Thus the ramification divisor is $`(\omega _Y(D))^{(dim𝒢r)}`$, and the dimension is
$`dim\mathrm{Prym}_𝔱(X)`$ $`=`$ $`r(g1)+{\displaystyle \frac{(dim𝒢r)(2g2+\mathrm{deg}D)}{2}}`$
$`=`$ $`dim𝒢(g1)+{\displaystyle \frac{(dim𝒢r)}{2}}\mathrm{deg}D.`$
Again, this is the same dimension as the base of the system.
|
warning/0007/math0007185.html
|
ar5iv
|
text
|
# Finiteness of simple homotopy type up to 𝑠-cobordism of aspherical 4-manifolds
## 0. Introduction
The Borel conjecture says that closed aspherical manifolds are determined by their fundamental groups, i.e., an isomorphism between the fundamental groups of two closed aspherical manifolds is induced by a homeomorphism of the manifolds. In dimension greater than $`4`$ this question is answered in positive for the class of manifolds with nonpositively curved Riemannian metric: this is the largest class of manifolds for which the answer is known so far. The advantage in higher dimension is the availability of the $`s`$-cobordism theorem and the surgery theory. In dimension $`3`$ the answer is known for a vast class of manifolds: namely, Haken manifolds and hyperbolic manifolds. The answer will be complete in dimension $`3`$ provided Thurston’s Geometrization conjecture is true. The dimension $`4`$ case is also not yet settled. For example the $`s`$-cobordism theorem and the exactness of the surgery sequence is known only for $`4`$-manifolds with elementary amenable fundamental groups.
In this paper we show that for a class of aspherical $`4`$-manifolds; up to $`s`$-cobordism there are only finitely many $`4`$-manifolds simple homotopy equivalent to a given member of this class. Due to the unavailability of the $`4`$-dimensional $`s`$-cobordism theorem we cannot quite say that there are only finitely many simple homotopy type up to homeomorphism of a manifold from this class.
## 1. Finiteness of simple homotopy type
Let $`𝒞`$ be the class of compact orientable aspherical $`4`$-manifolds so that for each $`M𝒞`$ there is a fiber bundle projection $`M𝕊^1`$ with irreducible fiber $`N`$ and either $`M`$ has nonempty boundary or it is closed and satisfies one of the following properties:
Here recall that a subgroup $`H`$ of a group $`G`$ is called square root closed if for any $`xG`$, $`x^2H`$ implies $`xH`$. And a $`3`$-manifold is called irreducible if any embedded $`2`$-sphere in it bounds an embedded $`3`$-disc. An irreducible $`3`$-manifold with nonempty boundary has nonvanishing first Betti number. Also any irreducible $`3`$-manifold with nonvanishing first Betti number is Haken.
The dual graph of the $`JSJ`$-decomposition has vertices the pieces in the decomposition and edges are tori which are common boundary component of two pieces. By fundamental group of a vertex or an edge we mean the fundamental group of the associated spaces.
Before we state our main theorem we recall the definition of homotopy-cobordant structure sets: Let $`M`$ be a compact manifold. Define $`𝒮_{TOP}^s(M,M)=\{(N,f)|f:NM,`$ where $`N`$ a compact manifold, $`f`$ a simple homotopy equivalence, $`f|_N`$ is a homeomorphism onto $`M`$ $`\}/`$, where $`(N_1,f_1)(N_2,f_2)`$ if there is a map $`F:WM`$ with domain $`W`$ a $`s`$-cobordism with $`W=N_1N_2`$ and $`F|_{N_i}=f_i`$. If the Whitehead group of $`\pi _1(M)`$ vanishes then this is the usual homotopy-topological structure set of $`M`$ provided the $`s`$-cobordism theorem is true in dim$`M`$. The dimension $`4`$ $`s`$-cobordism theorem is known to be true only for $`4`$-manifolds with elementary amenable fundamental group (see \[FQ\]).
In this paper we prove the following theorem:
###### Theorem 1.1
Let $`M𝒞`$. Then the set $`𝒮_{TOP}^s(M)`$ is finite when $`M`$ is closed and for $`n1`$ $`𝒮_{TOP}^s(M\times 𝔻^n,(M\times 𝔻^n))`$ has only one element.
###### Corollary 1.2
Let $`M𝒞`$ and $`N`$ be any other $`4`$-manifold homotopy equivalent to $`M`$. Then there are integers $`r`$ and $`s`$ so that $`M\mathrm{\#}r(𝕊^2\times 𝕊^2)`$ is diffeomorphic to $`N\mathrm{\#}s(𝕊^2\times 𝕊^2)`$. Here $`\mathrm{\#}`$ denotes connected sum. In such a case $`M`$ and $`N`$ are called stably diffeomorphic.
###### Demonstration Proof of Corollary 1.2
Follows from the main theorem in \[D\].
###### Remark Remark 1.3
As it is not yet known if a $`s`$-cobordism between two $`4`$-manifolds is trivial we cannot quite conclude that the manifolds in the class $`𝒞`$ has finitely many homotopy type up to homeomorphism.
Here we deduce an interesting corollary:
###### Corollary 1.4
Let $`M`$ be a nonsingular complex affine surface (i.e., a nonsingular complex algebraic surface in the complex space $`^n`$) which is a fiber bundle over the circle with irreducible (in $`3`$-manifold sense) fiber. Then for $`n1`$ $`M\times 𝔻^n`$ has only one homotopy type (with homotopy which are homeomorphism outside a compact set) up to homeomorphism.
###### Demonstration Proof
Using a suitable Morse function on $`M`$ (for example consider the polynomial function $`xx_0^2`$ for a fixed $`x_0M`$) it is easily deduced (by Morse theory) that $`M`$ is diffeomorphic to the interior of a compact aspherical $`4`$-manifold. This follows because the restriction to $`M`$ of any polynomial function has only finitely many critical value. (see corollary 2.8 in \[M\]). The Corollary follows.∎
## 2. Proof of the theorem 1.1
At first we check that the fundamental group of any of the $`4`$-manifolds in the class $`𝒞`$ has vanishing Whitehead group.
If $`N`$ is nonempty and has $`𝕊^2`$ as a boundary component then by irreducibility $`N`$ is homeomorphic to $`𝔻^3`$ and hence $`M`$ is homeomorphic to $`𝔻^3\times 𝕊^1`$. In this particular case the theorem is known. So we assume that if $`N\mathrm{}`$ then genus of any component of $`N`$ is $`1`$.
Note that for any $`M𝒞`$ the fiber $`N`$ of the fiber bundle $`M𝕊^1`$ is a Haken $`3`$-manifold in the cases $`(1)(4)`$. This implies $`\pi _1(N)𝒞l`$ from the notation of \[W\] and hence it has vanishing Whitehead group. Also $`\pi _1(N)`$ is regular coherent.
Now we can use the Mayer-Vietoris exact sequence (for $`K`$-theory) from \[W\] (Sections 17.1.3 and 17.2.3) to deduce that $`\pi _1(M)`$ has vanishing Whitehead group.
A general version of this fact is proved in (lemma V.3, \[H1\]).
If $`N`$ is hyperbolic then by the Mostow rigidity theorem the monodromy diffeomorphism of the fiber bundle $`M𝕊^1`$ is homotopic to an isometry of finite order and hence $`\pi _1(M)`$ is isomorphic to the fundamental group of a $`4`$-manifold $`M^{}`$ which has a $`^3\times `$ structure. Since $`M^{}`$ is nonpositively curved $`Wh(\pi _1(M))=Wh(\pi _1(M^{}))=0`$ by \[FJ\]. This conclusion also can be made by noting that in fact the monodromy diffeomorphism is (topologically) isotopic (see \[G\] and \[GMT\]) to a (finite order) isometry and hence the fiber bundle $`M`$ itself has $`^3\times `$ structure.
In \[Ro1\] and \[Ro2\] we proved the following theorem:
###### Theorem 2.1
(Theorem 1.2 in \[Ro1\] and Theorem 1.1 and 1.3 in \[Ro2\]) Let $`N`$ be a compact orientable irreducible $`3`$-manifold so that one of the following properties is satisfied:
Then for $`n2`$ $`N\times 𝔻^n`$ has only one homotopy type up to homeomorphism.
Here note that the case when $`N`$ has nonempty boundary is included in case $`(1)`$. In \[Ro2\] a large class of examples of $`3`$-manifolds is given satisfying the property $`(2)`$ and $`(3)`$ in the above theorem. In fact it was shown there that if we consider the Jaco-Shalen and Johannson ($`JSJ`$) decomposition of the Haken manifold with $`T`$ as one of the decomposing torus then the square root closed condition depends only on the pieces which abut the torus $`T`$. Also a large class of examples of $`3`$-manifolds are given which has a square root closed incompressible torus boundary component.
We recall the Wall-Novikov surgery exact sequence here:
Let $`𝒮(X,X)`$ denote the topological structure set of $`X`$ of the group of homotopy type of $`X`$ up to homeomorphism. For precise definition see any reference on surgery theory or in \[Ro1\]. (Here note that the differentiable structure set is not a group.) In terms of this group Theorem 2.1 say that $`𝒮(N\times 𝔻^n,(𝒮(N\times 𝔻^n))`$ is trivial for $`n2`$. We always assume that $`Wh(\pi _1(X))=0`$. Then these groups fit into a long exact sequence of groups:
$$\mathrm{}𝒮_{n1}(X)H_n(X,𝕃_0)L_n(\pi _1(X))𝒮_n(X)\mathrm{}$$
Here $`𝒮_n(X)`$ are the total surgery obstruction group of Ranicki (\[R1\]) and they are in bijection with $`𝒮(X\times 𝔻^n,(X\times 𝔻^n))`$ with a different indexing.
Note that the fundamental group of any $`M𝒞`$ is of the form $`\pi _1(M)=\pi _1(N)`$. From Theorem 2.1, the main theorem in \[S\] for the case $`(3)`$ and by Farrell-Jones Topological Rigidity theorem for nonpositively curved Riemannian manifold (in the hyperbolic case) (\[FJ\]) it follows that the (assembly) map $`H_n(N,𝕃_0)L_n(\pi _1(N))`$ is an isomorphism for large $`n`$.
Now the proof of the theorem follows from the following facts:
Fact 1: the Whitehead group of $`\pi _1(M)`$ vanishes.
Fact 2: the Ranicki Mayer-Vietoris exact sequence of surgery groups for groups which are semidirect product of a group with the infinite cyclic group (see \[R2\]).
Fact 3: the Mayer-Vietoris exact sequence of the generalized homology theory for $`K(\pi ,1)`$ spaces with coefficient in the surgery spectrum $`𝕃_0`$.
Fact 4: naturality of the assembly map and an application of five-lemma together with Theorem 2.1 and Siebenmann’s periodicity theorem (\[KS\]).
Fact 5: the corollary to the theorem V.12 in \[H1\] which says that if the assembly map $`H_5(M,𝕃_0)L_5(\pi _1(M))`$ is an epimorphism then the set $`𝒮_{TOP}^s(M)`$ is finite. ∎
###### Remark Remark 2.2
Here we remark that the Theorem 1.1 will be true for all compact $`4`$-manifolds which fiber over the circle if Thurston’s conjecture is true: i.e., if any aspherical closed $`3`$-manifold is either Haken, hyperbolic or Seifert fibered space, and if Theorem 2.1 is true for any Haken $`3`$-manifold.
###### Remark Remark 2.3
Hillman informed the author that he thinks it follows from \[H3\] that if $`M^{}`$ is homotopically equivalent to $`M𝒞`$ then in fact $`M^{}`$ and $`M`$ are $`s`$-cobordant.
###### Remark Remark 2.4
In \[H1\] Hillman proved Theorem 1.1 for the case when $`M`$ also supports a complex structure. Also note that in the theorem the case of certain complex surfaces is included in case $`(3)`$ because in (\[H2\]) it was proved if a complex surface fibers over the circle then the fiber is a Seifert fibered space. Hillman proved the Theorem 1.1 when the fiber is an arbitrary Seifert fibered space $`N`$ assuming that $`N`$ supports a nonpositively curved Riemannian metric (page 81 in \[H1\]). However, the unit tangent bundle of a closed oriented surface of genus $`2`$ is a Seifert fibered space which has $`\stackrel{~}{𝕊𝕃(2,)}`$ structure but does not support any nonpositively curved Riemannian metric (see \[Ro1\]).
Acknowledgement: The author would like to thank J.A. Hillman for sending his reprints. He also thanks Mehta Research Institute in Allahabad where this work was done.
|
warning/0007/hep-th0007024.html
|
ar5iv
|
text
|
# Untitled Document
HU–EP–00/27 hep-th/0007024
Noncommutative Compactifications of Type I Strings
on Tori with Magnetic Background Flux
Ralph Blumenhagen<sup>1</sup> e-mail: blumenha@physik.hu-berlin.de, Lars Görlich<sup>2</sup> e-mail: goerlich@physik.hu-berlin.de, Boris Körs<sup>3</sup> e-mail: koers@physik.hu-berlin.de and Dieter Lüst<sup>4</sup> e-mail: luest@physik.hu-berlin.de
Humboldt-Universität zu Berlin, Institut für Physik,
Invalidenstrasse 110, 10115 Berlin, Germany
Abstract
We construct six- and four-dimensional toroidal compactifications of the Type I string with magnetic flux on the D-branes. The open strings in this background probe a noncommutative internal geometry. Phenomenologically appealing features such as chiral fermions and supersymmetry breaking in the gauge sector are naturally realized by these vacua. We investigate the spectra of such noncommutative string compactifications and in a bottom-up approach discuss the possibility to obtain the standard or some GUT like model.
06/2000
1. Introduction
The search for realistic string vacua is one of the burning open problems within superstring theory. A phenomenologically viable string compactification should contain at least three chiral fermion generations, the standard model gauge group and broken space-time supersymmetry. In the context of ‘conventional’ string compactifications the requirement of getting chiral fermions is usually achieved by considering compact, internal background spaces with nontrivial topology rather than simple tori. In particular, when analyzing the Kaluza-Klein fermion spectra a net-fermion generation number arises if the internal Dirac operator has zero modes. For example, considering heterotic string compactifications on Calabi-Yau threefolds , the net-generation number is equal to $`|\chi |/2`$, where $`\chi `$ is the Euler number of the Calabi-Yau space. Chiral fermions are also present in a large class of heterotic orbifold compactifications , as well as in free bosonic and fermionic constructions. Type II string models with chiral fermions can be constructed by locating D-branes at transversal orbifold or conifold singularities , or by considering intersections of D-branes and NS-branes ; chiral type I models were first proposed in . Moreover orbifold compactificactions of eleven-dimensional M-theory can lead to chiral fermions, as discussed e.g. in .
The phenomenological requirement of breaking space-time supersymmetry can be met in various ways. In the context of heterotic string compactifications gaugino condensation or the Scherk-Schwarz mechanism \[11,,12\] lead to potentially interesting models with supersymmetry broken at low energies. In addition, as it was realized more recently, type II models on nontrivial background spaces with certain D-brane configurations possess broken space-time supersymmetry. Especially, when changing the GSO-projections tachyon free type 0 orientifolds in four dimensions can be constructed . Alternatively, orientifolds on six-dimensional orbifolds with brane-antibrane configurations provide interesting scenarios , where supersymmetry is left unbroken in the gravity bulk, but broken in the open string sector living on the brane-antibrane system.
Finally the quest for a realistic gauge group with sufficiently low rank is met in heterotic strings by choosing appropriate gauge vector bundles on the Calabi-Yau spaces , which can be alternatively described by turning on Wilson lines in Calabi-Yau or also in orbifold compactifications . On the type II side the rank of the gauge group can be also lowered by Wilson lines or, in the T-dual picture, by placing the branes at different positions inside the internal space.
As it should have become clear from the previous discussion, ‘standard’ heterotic, type I or type II compactifications on simple 6-tori do not meet any of the three above requirements. However, as we will discuss in this paper, turning on magnetic fluxes in the internal directions of the D-branes, thereby inducing mixed Neumann-Dirichlet boundary conditions for open strings equivalent to a noncommutative internal geometry on the branes, all three goals can be achieved in one single stroke.<sup>1</sup> In a previous paper we have discussed type I string compactifications on noncommutative asymmetric orbifold spaces. Specifically, we will discuss type I string compactifications on a product of $`d`$ noncommutative two-tori to $`102d`$ non-compact Minkowski dimensions ($`d=2,3`$), i.e. the ten-dimensional background spaces $`M^{10}`$ we are considering have the following form
$$M^{10}=R^{1,92d}\times \underset{j=1}{\overset{d}{}}T_{(j)}^2.$$
The coordinates in the internal space possess the following commutation relations
$$[X_{102j},X_{112j}]=i\theta ^{(j)},j=1,\mathrm{},d.$$
These commutation relations will be realized by D9-branes with constant background magnetic fluxes $`F^{(j)}`$ turned on in the directions of the 2-tori \[19--22\], corresponding to the following noncommutative deformation parameter $`\theta ^{(j)}`$ in eq.(1.1):
$$\theta ^{(j)}=\frac{2\pi \alpha ^{}F^{(j)}}{1+(F^{(j)})^2}.$$
The entire internal noncommutative torus will actually consist out of different sectors with different noncommutative deformation parameters, because we will introduce several D9-branes with different magnetic fluxes. We will show that the spectrum of open strings, with mixed boundary conditions in the internal directions is generically chiral, breaks space-time supersymmetry and leads to gauge groups of lower rank. It is however important to stress that the effective gauge theories in the uncompactified part of space-time are still commutative, and therefore are Lorentz invariant and local field theories.
This construction is the D-brane extended version of , where it was already observed that turning on magnetic flux in a toroidal type I compactification leads to supersymmetry breaking and chiral massless spectra in four space-time dimensions. However, the consistency conditions for such models were derived in the effective non-supersymmetric gauge theories, leaving the actual string theoretic conditions an open issue. We will show that, with all the insights gained in the description of D-branes with magnetic flux, we are now able to achieve a complete string theoretic understanding, giving rise to certain extensions and modifications of the purely field theoretical analysis. As a solution to the tadpole cancellation conditions we can get different sectors of D-branes with different magnetic fluxes, corresponding to different noncommutative boundary conditions. Chirality then arises in sectors of open strings which have ends on branes with different gauge flux, while the presence of any solitary flux is not sufficient. The gauge groups that act on the D-branes with non-vanishing flux are unitary instead of orthogonal or symplectic in accord with the general statement that only these are compatible with a noncommutative deformation of the coordinate algebra.
For this kind of models, it is sometimes very helpful to employ an equivalent T-dual description, where the background fields vanish and the torus is entirely commutative, but the D-branes intersect at various different angles . This description allows to present a more intuitive picture of the open string sector involved in such models. Chiral fermions then arise due to the nontrivial geometric boundary conditions of the intersecting D-branes, which at the same time also break space-time supersymmetry and lower the rank of the gauge group.
The paper is organized as follows. In the next section we analyze the one-loop amplitudes and the resulting tadpole cancellation conditions for D9-branes with mixed Neumann-Dirichlet boundary conditions moving in the background of $`d`$ two-dimensional tori ($`d=2,3`$). In section 3 we discuss specific six-dimensional models ($`d=2`$) working out the non-supersymmetric, chiral spectrum. We also point out some subtleties involving the mechanisms of supersymmetry breaking in ‘nearly’ supersymmetric brane configurations. In section 4 we move on to chiral, non-supersymmetric four-dimensional models ($`d=3`$), reconsider in particular the model presented in with GUT-like gauge group $`G=U(5)\times U(3)\times U(4)\times U(4)`$ and display another 4 generation model with ‘standard model’ gauge group $`G=U(3)\times U(2)\times U(1)^r`$.<sup>2</sup> For other recent bottom up attempts to obtain GUT’s and the standard model from branes see . Some phenomenological problems of this model are stressed at the end.
2. One loop amplitudes
In it was observed that turning on magnetic flux in a toroidal type I compactification leads to supersymmetry breaking and in general to chiral massless spectra in four space-time dimensions. The consistency conditions for such models were derived in the effective non-supersymmetric gauge theory but not in the full string theory. In this section we will show that, with the inclusion of D-branes with magnetic flux, respectively D-branes at angles, we are now able to derive the string theoretic tadpole cancellation conditions.
2.1. D9-branes with magnetic fluxes
As our starting point we consider the orientifold
$$\frac{\mathrm{Type}\mathrm{IIB}\mathrm{on}T^{2d}}{\mathrm{\Omega }},$$
where $`\mathrm{\Omega }`$ denotes the word-sheet parity transformation. In the following we will assume that $`T^{2d}`$ splits into a direct product of $`d`$ two-dimensional tori $`T_{(j)}^2`$ with coordinates $`X_1^{(j)}`$, $`X_2^{(j)}`$ and radii $`R_1^{(j)}`$, $`R_2^{(j)},j=1,\mathrm{},d`$. Their complex structures will always be taken to be purely imaginary, and the antisymmetric NSNS tensorfield will be set to zero. Turning on magnetic flux, $`F_{12}^{(j)}=F^{(j)}`$, on a D9-brane changes the pure Neumann boundary conditions into mixed Neuman-Dirichlet conditions
$$\begin{array}{cc}& _\sigma X_1^{(j)}+F^{(j)}_\tau X_2^{(j)}=0,\hfill \\ & _\sigma X_2^{(j)}F^{(j)}_\tau X_1^{(j)}=0,\hfill \end{array}$$
Let us consider different kinds of D9<sub>μ</sub>-branes, labelled by $`\mu \{1,\mathrm{},K\}`$, distinguished by different magnetic fluxes on at least one torus. Specifically we are only considering branes which are characterized by $`d`$ sets of two integers $`(n_\mu ^{(j)},m_\mu ^{(j)})`$ corresponding to the electric respectively magnetic charges<sup>3</sup> All models considered in correspond to the subset of branes with $`n_\mu ^{(j)}=1`$ for all $`j`$ and $`\mu `$..
The magnetic fluxes of such a brane are given by
$$F_\mu ^{(j)}=\frac{m_\mu ^{(j)}}{n_\mu ^{(j)}R_1^{(j)}R_2^{(j)}}.$$
Following the boundary conditions eq.(2.1) define an open string metric on $`T_{(j)}^2`$ in the following way:
$$G_\mu ^{(j)kl}=\frac{1}{1+(F_\mu ^{(j)})^2}\delta ^{kl}.$$
The deformation parameter of the noncommutative torus is given by
$$\theta _\mu ^{(j)kl}=\frac{2\pi \alpha ^{}F_\mu ^{(j)}}{1+(F_\mu ^{(j)})^2}ϵ^{kl}.$$
Let us now discuss the Kalazu-Klein mass spectrum of open strings on the noncommutative torus. Since the translations on the torus are not deformed by eq.(1.1), the center of mass Kaluza-Klein momenta on $`T_{(j)}^2`$ are unchanged and are given by
$$p_{\mu 1}^{(j)}=\frac{r_\mu ^{(j)}}{n_\mu ^{(j)}R_1^{(j)}},p_{\mu 2}^{(j)}=\frac{s_\mu ^{(j)}}{n_\mu ^{(j)}R_2^{(j)}}.$$
Here the electric charge of the brane enters as the number of times it wraps the torus while the integers $`r_\mu ^{(j)}`$ and $`s_\mu ^{(j)}`$ are the two Kaluza-Klein momentum numbers along $`T_{(j)}^2`$. Howover the masses of the open string Kalazu-Klein states do depend on the magnetic fluxes. Using the boundary state formalism one can compute the mass spectrum of the open string Kalazu-Klein states on $`T_{(j)}^2`$ with mixed boundary conditions specified by the two integers $`n_\mu ^{(j)}`$ and $`m_\mu ^{(j)}`$ with the following result ,:
$$(M_\mu ^{(j)})^2=\frac{(s_\mu ^{(j)}R_1^{(j)})^2+(r_\mu ^{(j)}R_2^{(j)})^2}{(m_\mu ^{(j)})^2+(n_\mu ^{(j)}R_1^{(j)}R_2^{(j)})^2}.$$
This allows use to extract the ‘noncommutative’ Kaluza-Klein momenta and winding contributions
$$\begin{array}{cc}\hfill \stackrel{~}{p}_\mu ^{(j)}& =\frac{n_\mu ^{(j)}R_1^{(j)}R_2^{(j)}}{(m_\mu ^{(j)})^2+(n_\mu ^{(j)}R_1^{(j)}R_2^{(j)})^2}\left(r_\mu ^{(j)}R_2^{(j)}+is_\mu ^{(j)}R_1^{(j)}\right)=\frac{1}{1+(F_\mu ^{(j)})^2}\left(p_{\mu 1}^{(j)}+ip_{\mu 2}^{(j)}\right),\hfill \\ \hfill w_\mu ^{(j)}& =\frac{m_\mu ^{(j)}}{(m_\mu ^{(j)})^2+(n_\mu ^{(j)}R_1^{(j)}R_2^{(j)})^2}\left(s_\mu ^{(j)}R_1^{(j)}+ir_\mu ^{(j)}R_2^{(j)}\right)=\frac{F_\mu ^{(j)}}{1+(F_\mu ^{(j)})^2}\left(p_{\mu 2}^{(j)}+ip_{\mu 1}^{(j)}\right),\hfill \end{array}$$
(here written in complex notation) where the mass formula eq.(2.1) is given as $`(M_\mu ^{(j)})^2=|\stackrel{~}{p}_\mu ^{(j)}|^2+|w_\mu ^{(j)}|^2`$. Of course, for $`F_\mu ^{(j)}=0`$ the momenta $`p_\mu ^{(j)}`$ and $`\stackrel{~}{p}_\mu ^{(j)}`$ agree, and $`w_\mu ^{(j)}=0`$. Note that the KK masses in eq.(2.1) can be also expressed by using the standard KK momenta eq.(2.1) togother with the open string metric eq.(2.1) as
$$(M_\mu ^{(j)})^2=p_{\mu k}^{(j)}G_\mu ^{(j)kl}p_{\mu l}^{(j)}.$$
To convince ourselves that we are indeed dealing with a noncommutative torus for open strings which end on the $`D9_\mu `$-brane with magnetic flux $`F_\mu ^{(j)}`$, let us compute the OPE of the vertex operators $`𝒪(z)=e^{ip_\mu ^{(j)}X^{(j)}}(z)`$ ,:
$$\begin{array}{cc}\hfill e^{ip_\mu ^{(j)}X^{(j)}}(\tau )e^{iq_\mu ^{(j)}X^{(j)}}(\tau ^{})& =(\tau \tau ^{})^{\frac{2\alpha ^{}p_\mu ^{(j)}q_\mu ^{(j)}}{1+(F_\mu ^{(j)})^2}}\hfill \\ & \mathrm{exp}\left(i\pi \alpha ^{}\frac{F_\mu ^{(j)}}{1+(F_\mu ^{(j)})^2}ϵ_{kl}p_{\mu k}^{(j)}q_{\mu l}^{(j)}\right)e^{i(p_\mu ^{(j)}+q_\mu ^{(j)})X^{(j)}}(\tau ^{})+\mathrm{}.\hfill \end{array}$$
For generic, internal momenta $`p_\mu ^{(j)}`$, $`q_\mu ^{(j)}`$ in eq.(2.1), the phase factor is indeed nontrivial, and hence the torus is noncommutative. The commutator of the coordinates of the open string endpoints on the $`D9_\mu `$-brane is given by ,
$$[X_1^{(j)}(\tau ),X_2^{(j)}(\tau )]=\frac{2\pi i\alpha ^{}F_\mu ^{(j)}}{1+(F_\mu ^{(j)})^2},$$
in agreement with eqs.(1.1),(2.1).
2.2. D$`(9d)`$-branes at angles
Instead of working with D9-branes with various magnetic fluxes, we will now use the T-dual description in terms of D-branes at angles , which allows to present a more intuitive picture of the open string sector involved in such models. Applying a T-duality in all $`X_2^{(j)}`$ directions
$$R_2^{(j)}R_2^{(j)^{}}=1/R_2^{(j)},$$
leads to boundary conditions for D$`(9d)`$-branes intersecting at angles, where the angle of the D$`(9d)`$-brane relative to the $`X_1^{(j)}`$ axes is given by
$$\mathrm{tan}\varphi ^{(j)}=F^{(j)}.$$
(In the following we will omit the prime on the dual radii.) This T-duality also maps $`\mathrm{\Omega }`$ onto $`\mathrm{\Omega }`$, where $``$ acts as complex conjugation on all the $`d`$ complex coordinates along the $`T_{(j)}^2`$ tori. Thus, instead of (2.1) we are considering the orientifold
$$\frac{\mathrm{Type}\mathrm{II}\mathrm{on}T^{2d}}{\mathrm{\Omega }}.$$
For $`d`$ even we have to take type IIB and for $`d`$ odd type IIA. Note that after performing this T-duality transformation the internal coordinates are completely commutative.
Let $`j\{1,\mathrm{},d\}`$ again label the $`d`$ different two-dimensional tori and $`\mu \{1,\mathrm{},K\}`$ the different kinds of D$`(9d)`$-branes, which are distinguished by different angles on at least one torus. Moreover, we are only considering branes which do not densely cover any of the two-dimensional tori. Thus, the position of a D$`(9d)`$-brane is described by two sets of integers $`(n_\mu ^{(j)},m_\mu ^{(j)})`$, labelling how often the D-branes are wound around the two fundamental cycles of each $`T_{(j)}^2`$. The angles of such a brane with the axes $`X_1^{(j)}`$ are given by
$$\mathrm{tan}\varphi _\mu ^{(j)}=\frac{m_\mu ^{(j)}R_2^{(j)}}{n_\mu ^{(j)}R_1^{(j)}}.$$
All these conventions are shown in figure 1.
Figure 1
Since $`\mathrm{\Omega }`$ reflects the D-branes at the axis $`X_1^{(j)}`$, for each brane labelled by $`(n_\mu ^{(j)},m_\mu ^{(j)})`$ we must also introduce the mirror brane with $`(n_\mu ^{}^{(j)},m_\mu ^{}^{(j)})=(n_\mu ^{(j)},m_\mu ^{(j)})`$. The values $`m_\mu ^{(j)}=0n_\mu ^{(j)}`$ and $`n_\mu ^{(j)}=0m_\mu ^{(j)}`$ correspond to branes located along one of the axis. The horizontal D-branes translate via T-duality into D9-branes with vanishing flux and the vertical ones into branes of lower dimension with pure Dirichlet boundary conditions.
The questions we are going to deal with in the following are: Is it possible to cancel all or at least the RR tadpoles originating from the Klein bottle amplitude by D9-branes with non-vanishing magnetic fluxes $`F^{(j)}`$, or equivalently by D$`(9d)`$-branes at nontrivial angles $`\varphi ^{(j)}`$? Taken that supersymmetry is broken generically by such a background, are there configuration which still preserve some amount of supersymmetry? This would provide a string scenario with partial supersymmetry breaking. Finally, what are the phenomenological properties of such compactifications? Concerning the first question we find a positive answer in the sense that the RR tadpole can be cancelled, while supersymmetry is always broken entirely, albeit sometimes in a rather subtle manner. This comes along with a non-vanishing NSNS tadpole and the presence of tachyons. Interestingly, these models generically contain chiral fermions motivating us to study how far one can get in deriving the standard model in this setting. However, later we will mention an obstacle to construct phenomenologically realistic models in this simple approach.
Technically we first have to compute all contributions to the massless RR tadpole. The cancellation conditions will then imply relations for the number of D9-branes and their respective background fluxes. This computation will be performed in the T-dual picture, where D9-branes with background fields are mapped to D$`(9d)`$-branes, and the background fields translate into relative angles. This picture allows to visualize the D-branes easily and gives a much better intuition than dealing with sets of D9-branes, all filling the same space but differing by background fields.
2.3. Klein bottle amplitude
The loop channel Klein bottle amplitude for (2.1) can be computed straightforwardly
$$𝒦=2^{(5d)}c(11)_0^{\mathrm{}}\frac{dt}{t^{(6d)}}\frac{1}{4}\frac{\vartheta \left[\genfrac{}{}{0pt}{}{0}{1/2}\right]^4}{\eta ^{12}}\underset{j=1}{\overset{d}{}}\left[\underset{r,s\text{ZZ}}{}e^{\pi t\left(r^2/\left(R_1^{(j)}\right)^2+s^2\left(R_2^{(j)}\right)^2\right)}\right],$$
with $`c=V_{102d}/\left(8\pi ^2\alpha ^{}\right)^{5d}`$. Transforming (2.1) into tree channel, one obtains the following massless RR tadpole
$$_0^{\mathrm{}}𝑑l\mathrm{\hspace{0.17em}2}^{(13d)}\underset{j=1}{\overset{d}{}}\left(\frac{R_1^{(j)}}{R_2^{(j)}}\right).$$
The tree channel Klein bottle amplitude allows to determine the normalization of the corresponding crosscap states
$$|C=2^{(d/24)}\left(\underset{j=1}{\overset{d}{}}\sqrt{\frac{R_1^{(j)}}{R_2^{(j)}}}\right)\left(|C_{\mathrm{NS}}+|C_\mathrm{R}\right).$$
2.4. Annulus amplitude
Next we calculate all contributions of open strings stretching between the various D$`(9d)`$-branes, generically located at nontrivial relative angles. We will both include the case, where the relative angle is vanishing, i.e. the background gauge flux is equal on both branes, and the case, where the angle is $`\pi /2`$ and the field gets infinitely large on, say, $`p`$ of the tori.
We start with the contributions of strings with both ends on the same brane. The T-dual of the Kaluza-Klein and winding spectrum in eq.(2.1) reads
$$M_\mu ^2=\underset{j=1}{\overset{d}{}}\left(\left(\frac{r_\mu ^{(j)}}{V_\mu ^{(j)}}\right)^2+\left(s_\mu ^{(j)}\right)^2\left(\frac{R_1^{(j)}R_2^{(j)}}{V_\mu ^{(j)}}\right)^2\right)$$
with
$$V_\mu ^{(j)}=\sqrt{\left(R_1^{(j)}n_\mu ^{(j)}\right)^2+\left(R_2^{(j)}m_\mu ^{(j)}\right)^2}$$
denoting the volume of the brane on $`T_{(j)}^2`$. It is now straightforward to compute the loop channel annulus amplitude for open strings starting and ending on the same brane and transform it to the tree channel
$$\stackrel{~}{A}_{\mu \mu }=cN_\mu ^2(11)_0^{\mathrm{}}𝑑l\frac{1}{2^{(d+1)}}\underset{j=1}{\overset{d}{}}\frac{\left(V_\mu ^{(j)}\right)^2}{R_1^{(j)}R_2^{(j)}}\frac{\vartheta \left[\genfrac{}{}{0pt}{}{1/2}{0}\right]^4}{\eta ^{12}}\underset{r,s}{}e^{\pi l\stackrel{~}{M}_\mu ^2}$$
with
$$\stackrel{~}{M}_\mu ^2=\underset{j=1}{\overset{d}{}}\left(\left(r_\mu ^{(j)}\right)^2\left(V_\mu ^{(j)}\right)^2+\left(s_\mu ^{(j)}\right)^2\left(\frac{V_\mu ^{(j)}}{R_1^{(j)}R_2^{(j)}}\right)^2\right).$$
$`N_\mu `$ counts the numbers of different kinds of branes. Using (2.1) one can determine the normalization of the boundary state, which has the schematic form
$$|D_\mu =2^{(d/2+1)}\left(\underset{j=1}{\overset{d}{}}\frac{\left(V_\mu ^{(j)}\right)^2}{R_1^{(j)}R_2^{(j)}}\right)\left(|D_{\mu ,\mathrm{NS}}+|D_{\mu ,\mathrm{R}}\right).$$
Reflecting the brane on a single $`T_{(j)}^2`$ by a $`\pi `$ rotation onto itself corresponds to $`(n_\mu ^{(j)},m_\mu ^{(j)})(n_\mu ^{(j)},m_\mu ^{(j)})`$ and, as can be determined in the boundary state approach, changes the sign of the RR charge, thus exchanging branes and anti-branes. Using the boundary state (2.1) we can compute the tree channel annulus amplitude for an open string stretched between two different D-branes
$$\begin{array}{cc}\hfill \stackrel{~}{A}_{\mu \nu }& =_0^{\mathrm{}}𝑑lD_\mu |e^{lH_{cl}}|D_\nu =\hfill \\ & \frac{1}{2}cN_\mu N_\nu I_{\mu \nu }_0^{\mathrm{}}𝑑l(1)^d\underset{\alpha ,\beta \{0,1/2\}}{}(1)^{2(\alpha +\beta )}\frac{\vartheta \left[\genfrac{}{}{0pt}{}{\alpha }{\beta }\right]^{4d}\underset{j=1}{\overset{d}{}}\vartheta \left[\genfrac{}{}{0pt}{}{\alpha }{(\varphi _\nu ^{(j)}\varphi _\mu ^{(j)})/\pi +\beta }\right]}{\eta ^{123d}_{j=1}^d\vartheta \left[\genfrac{}{}{0pt}{}{1/2}{(\varphi _\nu ^{(j)}\varphi _\mu ^{(j)})/\pi +1/2}\right]},\hfill \end{array}$$
where the coefficient
$$I_{\mu \nu }=\underset{j=1}{\overset{d}{}}\left(n_\mu ^{(j)}m_\nu ^{(j)}m_\mu ^{(j)}n_\nu ^{(j)}\right)$$
is the (oriented) intersection number of the two branes. It gives rise to an extra multiplicity in the annulus loop channel, which we have to take into account, when we compute the massless spectrum. In order to properly include the case where some $`\varphi _\mu ^{(j)}=\varphi _\nu ^{(j)}`$, one needs to employ the relation
$$\underset{\psi 0}{lim}\frac{2\mathrm{sin}(\pi \psi )}{\vartheta \left[\genfrac{}{}{0pt}{}{1/2}{1/2+\psi }\right]}=\frac{1}{\eta ^3}$$
and include a sum over KK momenta and windings as in (2.1). The contribution to the massless RR tadpole due to (2.1) and (2.1) is
$$_0^{\mathrm{}}𝑑lN_\mu N_\nu \mathrm{\hspace{0.17em}2}^{(3d)}\underset{j=1}{\overset{d}{}}\left(\frac{\left(R_1^{(j)}\right)^2n_\mu ^{(j)}n_\nu ^{(j)}+\left(R_2^{(j)}\right)^2m_\mu ^{(j)}m_\nu ^{(j)}}{R_1^{(j)}R_2^{(j)}}\right).$$
The loop channel annulus can be obtained by a modular transformation
$$\begin{array}{cc}\hfill A_{\mu \nu }=cN_\mu N_\nu I_{\mu \nu }_0^{\mathrm{}}\frac{dt}{t^{(6d)}}\frac{1}{4}\underset{\alpha ,\beta \{0,1/2\}}{}& (1)^{2(\alpha +\beta )}e^{2i\alpha _j(\varphi _\nu ^{(j)}\varphi _\mu ^{(j)})}e^{i\pi d/2}\hfill \\ & \times \frac{\vartheta \left[\genfrac{}{}{0pt}{}{\beta }{\alpha }\right]^{4d}\underset{j=1}{\overset{d}{}}\vartheta \left[\genfrac{}{}{0pt}{}{(\varphi _\nu ^{(j)}\varphi _\mu ^{(j)})/\pi \beta }{\alpha }\right]}{\eta ^{123d}_{j=1}^d\vartheta \left[\genfrac{}{}{0pt}{}{(\varphi _\nu ^{(j)}\varphi _\mu ^{(j)})/\pi 1/2}{1/2}\right]}.\hfill \end{array}$$
Of course, one can alternatively start from the loop channel, putting in the intersection numbers as an extra multiplicity by hand. The loop channel annulus amplitude looks like a twisted open string sector and considering for instance the NS sector, $`\alpha =\beta =0`$, of (2.1) one can expand the $`\vartheta `$-functions in (2.1) as
$$\frac{\vartheta \left[\genfrac{}{}{0pt}{}{0}{0}\right]^{4d}}{\eta ^{12d}}\underset{j=1}{\overset{d}{}}\left[\left(\underset{l^{(j)}1}{}q^{ϵ^{(j)}l^{(j)}}\right)\frac{\underset{n^{(j)}1}{}(1+q^{n^{(j)}ϵ^{(j)}\frac{1}{2}})(1+q^{n^{(j)}+ϵ^{(j)}\frac{1}{2}})}{_{n^{(j)}1}(1q^{n^{(j)}ϵ^{(j)}})(1q^{n^{(j)}+ϵ^{(j)}})}\right],$$
with $`ϵ^{(j)}=(\varphi _\mu ^{(j)}\varphi _\nu ^{(j)})/\pi `$. The non-negative integers $`l^{(j)}`$ correspond to the Landau-levels in .
2.5. Möbius amplitude
Computing the overlap between the crosscap state (2.1) and a boundary state (2.1) yields the contribution of the brane $`D(9p)_\mu `$ to the Möbius amplitude
$$\begin{array}{cc}\hfill \stackrel{~}{M}_\mu =cN_\mu \mathrm{\hspace{0.17em}2}^5(1)^d& _0^{\mathrm{}}𝑑l\underset{j=1}{\overset{d}{}}m_\mu ^{(j)}\hfill \\ & \times \underset{\alpha ,\beta \{0,1/2\}}{}(1)^{2(\alpha +\beta )}\frac{\vartheta \left[\genfrac{}{}{0pt}{}{\alpha }{\beta }\right]^{4d}\underset{j=1}{\overset{d}{}}\vartheta \left[\genfrac{}{}{0pt}{}{\alpha }{\varphi _\mu ^{(j)}/\pi +\beta }\right]}{\eta ^{123d}_{j=1}^d\vartheta \left[\genfrac{}{}{0pt}{}{1/2}{\varphi _\mu ^{(j)}/\pi +1/2}\right]},\hfill \end{array}$$
with argument $`q=\mathrm{exp}(4\pi l)`$. Therefore the contribution to the RR tadpole is
$$_0^{\mathrm{}}𝑑lN_\mu \mathrm{\hspace{0.17em}2}^{(9d)}\underset{j=1}{\overset{d}{}}\left(\frac{R_1^{(j)}}{R_2^{(j)}}n_\mu ^{(j)}\right).$$
The overall sign in (2.1) and (2.1) is fixed by the tadpole cancellation condition. In the loop channel the contribution of the Möbius strip results from strings starting on one brane and ending on its mirror partner. The extra multiplicity given by the numbers $`m_\mu ^{(j)}`$ of intersection points invariant under $``$ needs to be regarded as before. Now we have all the ingredients to study the relations which derive from the cancellation of massless RR tadpoles.
3. Compactifications to six dimensions
We are compactifying type I strings on a four-dimensional torus and cancel the tadpoles by introducing stacks of D9-branes with magnetic fluxes. The T-dual arrangement of D7-branes at angles looks like the situation depicted in figure 2, where we have drawn only two types of D7-branes labelled by $`\mu `$ and $`\nu `$ and their mirror partners $`\mu ^{}`$ and $`\nu ^{}`$, the angles being chosen arbitrary.
Figure 2
3.1. Six-dimensional models
The complete annulus amplitude is a sum over all open strings stretched between the various D7-branes
$$\begin{array}{cc}\hfill \stackrel{~}{A}_{tot}=& \underset{\mu =1}{\overset{K}{}}\left(\stackrel{~}{A}_{\mu \mu }+\stackrel{~}{A}_{\mu ^{}\mu ^{}}+\stackrel{~}{A}_{\mu \mu ^{}}+\stackrel{~}{A}_{\mu ^{}\mu }\right)+\hfill \\ & \underset{\mu <\nu }{}\left(\stackrel{~}{A}_{\mu \nu }+\stackrel{~}{A}_{\nu \mu }+\stackrel{~}{A}_{\mu ^{}\nu ^{}}+\stackrel{~}{A}_{\nu ^{}\mu ^{}}+\stackrel{~}{A}_{\mu \nu ^{}}+\stackrel{~}{A}_{\nu \mu ^{}}+\stackrel{~}{A}_{\mu ^{}\nu }+\stackrel{~}{A}_{\nu ^{}\mu }\right).\hfill \end{array}$$
Using (2.1) and (2.1) and adding up all these various contributions yields the following two RR tadpoles
$$\begin{array}{cc}& _0^{\mathrm{}}𝑑l\mathrm{\hspace{0.17em}8}\left(\frac{R_1^{(1)}R_1^{(2)}}{R_2^{(1)}R_2^{(2)}}\right)\left(\underset{\mu =1}{\overset{K}{}}N_\mu n_\mu ^{(1)}n_\mu ^{(2)}\right)^2+_0^{\mathrm{}}𝑑l\mathrm{\hspace{0.17em}8}\left(\frac{R_2^{(1)}R_2^{(2)}}{R_1^{(1)}R_1^{(2)}}\right)\left(\underset{\mu =1}{\overset{K}{}}N_\mu m_\mu ^{(1)}m_\mu ^{(2)}\right)^2.\hfill \end{array}$$
For the total Möbius amplitude we obtain the RR tadpole
$$_0^{\mathrm{}}𝑑l\mathrm{\hspace{0.17em}2}^8\frac{R_1^{(1)}R_1^{(2)}}{R_2^{(1)}R_2^{(2)}}\underset{\mu =1}{\overset{K}{}}N_\mu n_\mu ^{(1)}n_\mu ^{(2)}.$$
Note, the two special cases of $`N_9`$ horizontal and $`N_5`$ vertical D7-branes are contained in (3.1) and (3.1) by setting $`N_\mu =N_9/2`$ respectively $`N_\mu =N_5/2`$. Choosing the minus sign in (3.1) we get the two RR tadpole cancellation conditions
$$\begin{array}{cc}& \frac{R_1^{(1)}R_1^{(2)}}{R_2^{(1)}R_2^{(2)}}:\underset{\mu =1}{\overset{K}{}}N_\mu n_\mu ^{(1)}n_\mu ^{(2)}=16,\hfill \\ & \frac{R_2^{(1)}R_2^{(2)}}{R_1^{(1)}R_1^{(2)}}:\underset{\mu =1}{\overset{K}{}}N_\mu m_\mu ^{(1)}m_\mu ^{(2)}=0.\hfill \end{array}$$
As one might have expected, pure D9-branes with $`m_\mu ^{(j)}=0`$ only contribute to the tadpole proportional to the volume of the torus, while the D5-branes with $`n_\mu ^{(j)}=0`$ to the one proportional to the inverse volume. Remarkably, by choosing multiple winding numbers, $`n_\mu ^{(j)}>1`$, one can reduce the rank of the gauge group. As usual in non-supersymmetric models, there remains an uncancelled NSNS tadpole, which needs to be cancelled by a Fischler-Susskind mechanism.
In the next section we shall show that except for the trivial case, when $`m_\mu ^{(1)}=m_\mu ^{(2)}=0`$ for all $`\mu `$, i.e. vanishing gauge flux on all the D9-branes, supersymmetry is broken and tachyons develop for open strings stretched between different branes. In contrast to the breaking of supersymmetry in a brane-antibrane system these tachyons cannot be removed by turning on Wilson-lines, which is related via T-duality to shifting the position of the branes by some constant vector. At any non trivial angle there always remains an intersection point of two D7-branes where the tachyons can localize. Also the lowest lying bosonic spectrum depends on the radii of the torus, which determine the relative angles. The zero point energy in the NS sector of a string stretching between two different branes is shifted by
$$\mathrm{\Delta }E_{0,\mathrm{NS}}=\frac{1}{2}\underset{j=1}{\overset{d}{}}\frac{\varphi _\mu ^{(j)}\varphi _\nu ^{(j)}}{\pi }$$
using the convention $`\varphi _\mu ^{(j)}\varphi _\nu ^{(j)}(0,\pi /2]`$. Even assuming a standard GSO projection, the lightest physical state can easily be seen to be tachyonic except for the supersymmetric situation with $`\varphi _\mu ^{(1)}\varphi _\nu ^{(1)}=\varphi _\mu ^{(2)}\varphi _\nu ^{(2)}`$. We shall find in the next section that tadpole cancellation prohibits this solution, except when all fluxes vanish.
On the contrary, the chiral fermionic massless spectrum is independent of the moduli and we display it in table 1.
spin rep. number $`(1,2)`$ $`𝐀_\mu +\overline{𝐀}_\mu `$ $`4m_\mu ^{(1)}m_\mu ^{(2)}`$ $`(1,2)`$ $`𝐀_\mu +\overline{𝐀}_\mu +𝐒_\mu +\overline{𝐒}_\mu `$ $`2m_\mu ^{(1)}m_\mu ^{(2)}(n_\mu ^{(1)}n_\mu ^{(2)}1)`$ $`(1,2)`$ $`(𝐍_\mu ,𝐍_\nu )+(\overline{𝐍}_\mu ,\overline{𝐍}_\nu )`$ $`(n_\mu ^{(1)}m_\nu ^{(1)}+m_\mu ^{(1)}n_\nu ^{(1)})(n_\mu ^{(2)}m_\nu ^{(2)}+m_\mu ^{(2)}n_\nu ^{(2)})`$ $`(1,2)`$ $`(𝐍_\mu ,\overline{𝐍}_\nu )+(\overline{𝐍}_\mu ,𝐍_\nu )`$ $`(n_\mu ^{(1)}m_\nu ^{(1)}m_\mu ^{(1)}n_\nu ^{(1)})(n_\mu ^{(2)}m_\nu ^{(2)}m_\mu ^{(2)}n_\nu ^{(2)})`$
Table 1: Chiral 6D massless open string spectrum.
($`𝐀_\mu `$ and $`𝐒_\mu `$ denote the antisymmetric resp. symmetric tensor representations with respect to $`U(N_\mu )`$, $`SO(N_\mu )`$ or $`Sp(N_\mu )`$.) Since $`\mathrm{\Omega }`$ exchanges a brane with its mirror brane, the Chan-Paton indices of strings ending on a stack of branes with non-vanishing gauge flux have no $`\mathrm{\Omega }`$ projection and the gauge group is $`U(N_\mu )`$. If $`\mathrm{\Omega }`$ leaves branes invariant, i.e. the flux vanishes or is infinite, corresponding to pure D9- or D5-branes, the gauge factor is $`SO(N_\mu )`$ or $`Sp(N_\mu )`$, respectively.
The degeneracy of states stated in the third column of table 1 is essentially given by the intersection numbers of the D7-branes. Whenever it is formally negative, one has to pick the $`(2,1)`$ spinor of opposite chirality taking into account the opposite orientation of the branes at the intersection. As was pointed out earlier, a change of the orientation switches the RR charge in the tree channel translating into the opposite GSO projection in the loop channel. Therefore the other chirality survives the GSO projection in the R sector. If the multiplicity is zero, this does not mean that there are no massless open string states in this sector, it only means that the spectrum is not chiral. This happens precisely when two branes lie on top of each other in one of the two $`T_{(j)}^2`$ tori. Then the extra zero modes give rise to an extra spinor state of opposite chirality. The chiral spectrum shown in table 1 does indeed cancel the irreducible $`R^4`$ and $`F^4`$ anomalies.
We have also considered a $`\text{ZZ}_2`$ orbifold background, together with nonvanishing magnetic flux, which changes the transversality condition to
$$\frac{R_2^{(1)}R_2^{(2)}}{R_1^{(1)}R_1^{(2)}}:\underset{\mu =1}{\overset{K}{}}N_\mu m_\mu ^{(1)}m_\mu ^{(2)}=16$$
and leads to a projection $`SO(N_\mu ),Sp(N_\mu )U(N_\mu /2)`$ on pure D9- and D5-branes but no further changes on D9-branes with nonvanishing flux. In this background it appears to be possible to construct also supersymmetric models .
3.2. Supersymmetry Breaking
One might suspect that there exist nontrivial configurations of D-branes cancelling all tadpoles while preserving maybe a reduced number of supersymmetries. Now, we would like to show that no such nontrivial supersymmetric configurations exist. We assume the absence of anti-branes from the beginning as their presence breaks supersymmetry anyway.
Going to the T-dual picture with D-branes at angles, we can apply the results of for the supersymmetry preserved by two D-branes intersecting at some relative angles $`\varphi ^{(1)}`$ and $`\varphi ^{(2)}`$ on the two tori. Whenever the square of the operation $`\mathrm{\Theta }`$ which rotates one brane onto the other has eigenvalues 1 when acting on the supercharges
$$\mathrm{\Theta }^2Q_\alpha =Q_\alpha ,$$
some supersymmetry is preserved. Here $`Q_\alpha `$ denote the sixteen tendimensional spinor states with internal quantum numbers $`(\pm 1/2,\pm 1/2,\mathrm{})`$ and some definite chirality. Hence, $`\mathrm{\Theta }^2`$ has eigenvalues $`\mathrm{exp}(\pm i\varphi ^{(1)}\pm i\varphi ^{(2)})`$ being equal to 1 exactly if $`\varphi ^{(1)}=\pm \varphi ^{(2)}\mathrm{mod}2\pi `$, which preserves the supercharges $`\pm (1/2,1/2,\mathrm{})`$ or $`\pm (1/2,1/2,\mathrm{})`$ respectively. An important point to notice is, that the angles must not be measured modulo $`\pi `$, but instead the relative orientations of the respective branes need to be regarded and a consistent convention in measuring angles only modulo $`2\pi `$ in $`(\pi ,\pi ]`$ must be adopted. For this purpose, on each $`T_{(j)}^2`$ we associate the direction of the vector $`(n_\mu ^{(j)}R_1^{(j)},m_\mu ^{(j)}R_2^{(j)})`$ to the respective D7-brane. To preserve supersymmetry, one of the two conditions, $`\varphi ^{(1)}=\pm \varphi ^{(2)}\mathrm{mod}2\pi `$, must hold in any open string sector. Therefore all branes need to be at equal or opposite angles on the two tori.
Now let us first consider brane configurations without D5-branes, i.e. without vertical D7-branes, and concentrate on the second condition in (3.1), which we call the transversality condition. We are free to choose $`m_\mu ^{(1)}0`$. In order to have any chance to satisfy the transversality condition, there is at least one pair of branes, D7<sub>1</sub> and D7<sub>2</sub>, with $`m_1^{(2)}0`$ and $`m_2^{(2)}0`$. One can easily convince oneself, that whenever D7<sub>1</sub> and D7<sub>2</sub> have relative angles $`\varphi ^{(1)}=\pm \varphi ^{(2)}`$ this cannot be the true for either D7<sub>1</sub> and D7$`_2^{}`$ or D7$`_1^{}`$ and D7<sub>2</sub>.
We still need to include vertical D7-branes to complete the proof. They have a positive contribution to the transversality condition in (3.1). In order to get a net negative contribution from D7-branes at angles in $`(0,\pi /2)`$ relative to the $`X_1^{(j)}`$ axes, we need $`\pm m_\mu ^{(1)}0`$ and $`m_\mu ^{(2)}0`$ for some $`\mu `$. These branes have $`\varphi ^{(1)}=\varphi ^{(2)}`$ relative to their mirrors, preserving $`\pm (1/2,1/2,\mathrm{})`$, but at best may only have $`\varphi ^{(1)}=\pi \varphi ^{(2)}`$ relative to the vertical D7-branes. Thus, such configurations are not supersymmetric, either.
Figure 3
Such a situation has been illustrated in figure 3. If one flips the orientation of the vertical D7-branes on the second torus, the configuration turns supersymmetric, as all brane sectors now satisfy $`\varphi ^{(1)}=\varphi ^{(2)}`$. But one finds, that in order to cancel the non-abelian anomaly then, one is required to take a different chirality for the D5-D9 strings as compared to the rest of the spectrum and therefore the sector behaves as in the presence of an anti-D5-brane. This refers to the change of the GSO projection in the loop channel induced by flipping the orientation.
4. Four dimensional models
The completely analogous computation as in six dimensions can be performed for the compactification of type I strings on a 6-torus in the presence of additional gauge fields. Now we cancel the tadpoles by D9-branes with magnetic fluxes on all three 2-tori respectively, in the T-dual picture, by D6-branes at angles. One obtains four independent tadpole cancellation conditions
$$\begin{array}{cc}& \frac{R_1^{(1)}R_1^{(2)}R_1^{(3)}}{R_2^{(1)}R_2^{(2)}R_2^{(3)}}:\underset{\mu =1}{\overset{K}{}}N_\mu n_\mu ^{(1)}n_\mu ^{(2)}n_\mu ^{(3)}=16\hfill \\ & \frac{R_1^{(1)}R_2^{(2)}R_2^{(3)}}{R_2^{(1)}R_1^{(2)}R_1^{(3)}}:\underset{\mu =1}{\overset{K}{}}N_\mu n_\mu ^{(1)}m_\mu ^{(2)}m_\mu ^{(3)}=0\hfill \\ & \frac{R_2^{(1)}R_1^{(2)}R_2^{(3)}}{R_1^{(1)}R_2^{(2)}R_1^{(3)}}:\underset{\mu =1}{\overset{K}{}}N_\mu m_\mu ^{(1)}n_\mu ^{(2)}m_\mu ^{(3)}=0\hfill \\ & \frac{R_2^{(1)}R_2^{(2)}R_1^{(3)}}{R_1^{(1)}R_1^{(2)}R_2^{(3)}}:\underset{\mu =1}{\overset{K}{}}N_\mu m_\mu ^{(1)}m_\mu ^{(2)}n_\mu ^{(3)}=0.\hfill \end{array}$$
(For convenience they are given in the picture with D6-branes at angles.) Again the gauge group contains a $`U(N_\mu )`$ factor for each stack of D9-branes with non-vanishing flux, an $`SO(N_\mu )`$ gauge factor for a stack with vanishing flux and an $`Sp(N_\mu )`$ factor for a stack of D5-branes. The general spectrum of chiral fermions with respect to the gauge group factors is presented in table 2.
rep. number $`(𝐀_\mu )_L`$ $`8m_\mu ^{(1)}m_\mu ^{(2)}m_\mu ^{(3)}`$ $`(𝐀_\mu )_L+(𝐒_\mu )_L`$ $`4m_\mu ^{(1)}m_\mu ^{(2)}m_\mu ^{(3)}(n_\mu ^{(1)}n_\mu ^{(2)}n_\mu ^{(3)}1)`$ $`(𝐍_\mu ,𝐍_\nu )_L`$ $`(n_\mu ^{(1)}m_\nu ^{(1)}+m_\mu ^{(1)}n_\nu ^{(1)})(n_\mu ^{(2)}m_\nu ^{(2)}+m_\mu ^{(2)}n_\nu ^{(2)})(n_\mu ^{(3)}m_\nu ^{(3)}+m_\mu ^{(3)}n_\nu ^{(3)})`$ $`(\overline{𝐍}_\mu ,𝐍_\nu )_L`$ $`(n_\mu ^{(1)}m_\nu ^{(1)}m_\mu ^{(1)}n_\nu ^{(1)})(n_\mu ^{(2)}m_\nu ^{(2)}m_\mu ^{(2)}n_\nu ^{(2)})(n_\mu ^{(3)}m_\nu ^{(3)}m_\mu ^{(3)}n_\nu ^{(3)})`$
Table 2: Left-handed 4D massless open string spectrum.
Whenever the intersection number in the second column is formally negative, one again has to take the conjugate representation. The spectrum in table 2 is free of non-abelian gauge anomalies. The remarks concerning the bosonic NS part of the spectrum made above for six dimensions also apply here. Masses depend on the radii and we have not been able to produce an otherwise consistent model free of tachyons or even preserving supersymmetry.
In the next subsections we discuss some examples and point out some phenomenological issues for these models.
4.1. A 24 generation $`SU(5)`$ model
Having found a way to break supersymmetry, to reduce the rank of the gauge group and to produce chiral spectra in four space-time dimensions, it is tempting to search in a compact bottom-up approach for brane configurations producing massless spectra close to the standard model. The tachyons are not that dangerous from the effective field theory point of view, as they simply may serve as Higgs-bosons for spontaneous gauge symmetry breaking, anticipating a mechanism to generate a suitable potential keeping their vacuum expectation values finite. In a three generation GUT model was presented, which we shall revisit in the following. The gauge group of the model is $`G=U(5)\times U(3)\times U(4)\times U(4)`$ with maximal rank, so that we have to choose all $`n_\mu ^{(j)}=1`$. The following choice of $`m_\mu ^{(j)}`$ then satisfies all tadpole cancellation conditions (4.1):
$$m_\mu ^{(j)}=\left(\begin{array}{cccc}3& 5& 1& 1\\ 1& 1& 1& 1\\ 1& 1& 1& 1\end{array}\right).$$
This configuration of D6-branes is displayed in figure 4, where the mirror branes have been omitted. The chiral part of the fermionic massless spectrum is shown in table 3.
Figure 4
rep. number $`(\mathrm{𝟏𝟎},\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏})`$ $`24`$ $`(\mathrm{𝟏},\mathrm{𝟑},\mathrm{𝟏},\mathrm{𝟏})`$ $`40`$ $`(\overline{\mathrm{𝟓}},\overline{\mathrm{𝟑}},\mathrm{𝟏},\mathrm{𝟏})`$ $`8`$ $`(\mathrm{𝟏},\mathrm{𝟏},\overline{\mathrm{𝟔}},\mathrm{𝟏})`$ $`8`$ $`(\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟏},\mathrm{𝟔})`$ $`8`$
Table 3: Chiral left-handed fermions for the 24 generation model.
No chiral fermions transform under both the $`U(5)\times U(3)`$ gauge group and the $`U(4)\times U(4)`$ gauge group, but there will of course also be non-chiral bifundamentals. If we think of the $`SU(5)`$ factor as a GUT gauge group, then this model has 24 generations<sup>4</sup> In this model was advocated as a three generation model. We can formally reproduce the model in by dividing the matrix (4.1) by a factor of two. However, this is inconsistent as it would violate the condition that the $`m_\mu ^{(j)}`$’s have to be integers. Thus, we conclude that in string theory only the choice (4.1) is correct and the model is actually a 24 generation model.. We shall see in the following that it is actually impossible to get a model with three or any odd number of generations in this framework.
4.2. A four generation model
The tadpole cancellation condition
$$\underset{\mu =1}{\overset{K}{}}N_\mu n_\mu ^{(1)}n_\mu ^{(2)}n_\mu ^{(3)}=16$$
tells us that we can reduce the rank of the gauge group right from the beginning by choosing some $`n_\mu ^{(j)}>1`$. Therefore, we can envision a model where we start with the gauge group $`U(3)\times U(2)\times U(1)^r`$ at the string scale. In order to have three quark generations in the $`(\mathrm{𝟑},\mathrm{𝟐})`$ representation of $`SU(3)\times SU(2)`$, we necessarily need $`I_{12}=3`$ and $`I_{12^{}}=0`$. However, this is not possible, as $`I_{\mu \nu }I_{\mu \nu ^{}}`$ is always an even integer. The model we found closest to the 4 generation standard model is presented in the following. We choose the gauge group $`U(3)\times U(2)\times U(1)^2`$ and the following configuration of four stacks of D-branes:
$$n_\mu ^{(j)}=\left(\begin{array}{cccc}1& 1& 1& 1\\ 1& 1& 1& 1\\ 1& 1& 1& 10\end{array}\right),m_\mu ^{(j)}=\left(\begin{array}{cccc}0& 2& 2& 0\\ 0& 1& 2& 0\\ 1& 0& 0& 1\end{array}\right).$$
The configuration has been illustrated in figure 5.
Figure 5
The resulting chiral massless spectrum is shown in table 4.
$`SU(3)\times SU(2)\times U(1)^4`$ number $`(\mathrm{𝟑},\mathrm{𝟐})_{(1,1,0,0)}`$ $`2`$ $`(\mathrm{𝟑},\mathrm{𝟐})_{(1,1,0,0)}`$ $`2`$ $`(\overline{\mathrm{𝟑}},\mathrm{𝟏})_{(1,0,1,0)}`$ $`4`$ $`(\overline{\mathrm{𝟑}},\mathrm{𝟏})_{(1,0,1,0)}`$ $`4`$ $`(\mathrm{𝟏},\mathrm{𝟐})_{(0,1,0,1)}`$ $`2`$ $`(\mathrm{𝟏},\mathrm{𝟐})_{(0,1,0,1)}`$ $`2`$ $`(\mathrm{𝟏},\mathrm{𝟏})_{(0,0,1,1)}`$ $`4`$ $`(\mathrm{𝟏},\mathrm{𝟏})_{(0,0,1,1)}`$ $`4`$
Table 4: Chiral left-handed fermions for the 4 generation model.
Computing the mixed $`G^2U(1)`$ anomalies one realizes that one of the abelian gauge factors is anomalous, which needs to be cured by the Green-Schwarz mechanism. The other three anomaly-free abelian gauge groups include a suitable hypercharge $`U(1)`$
$$U(1)_Y=\frac{1}{3}U(1)_1+U(1)_3U(1)_4,$$
so that the spectrum finally looks like the one in table 5.
$`SU(3)\times SU(2)\times U(1)_Y\times U(1)^2`$ number $`(\mathrm{𝟑},\mathrm{𝟐})_{(\frac{1}{3},1,0)}`$ $`2`$ $`(\mathrm{𝟑},\mathrm{𝟐})_{(\frac{1}{3},1,0)}`$ $`2`$ $`(\overline{\mathrm{𝟑}},\mathrm{𝟏})_{(\frac{4}{3},0,1)}`$ $`4`$ $`(\overline{\mathrm{𝟑}},\mathrm{𝟏})_{(\frac{2}{3},0,1)}`$ $`4`$ $`(\mathrm{𝟏},\mathrm{𝟐})_{(1,1,0)}`$ $`2`$ $`(\mathrm{𝟏},\mathrm{𝟐})_{(1,1,0,)}`$ $`2`$ $`(\mathrm{𝟏},\mathrm{𝟏})_{(0,0,1)}`$ $`4`$ $`(\mathrm{𝟏},\mathrm{𝟏})_{(2,0,1)}`$ $`4`$
Table 5: Chiral left-handed fermions for the 4 generation model.
We found a semi-realistic, non-supersymmetric, four generation standard-model like spectrum with two gauged flavour symmetries and right-handed neutrinos. In order to determine the Higgs sector, we would have to investigate the bosonic part of the spectrum. However, this is not universal but depends on the radii of the six-dimensional torus. We will not elaborate this further but instead discuss another important issue concerning the possible phenomenological relevance of these models.
Since we break supersymmetry already at the string scale $`M_s`$, in order to solve the gauge hierarchy problem we must choose $`M_s`$ in the TeV region. Let us employ the T-dual picture of D6-branes at angles again to analyze the situation in more detail. Using the relations
$$M_{pl}^2\frac{M_s^8V_6}{g_s^2},\frac{1}{(g_{\mathrm{YM}}^{(\mu )})^2}\frac{M_s^3V_\mu }{g_s},$$
where $`V_\mu `$ denotes the volume of some D6-brane in the internal directions
$$V_\mu =\underset{j=1}{\overset{3}{}}V_\mu ^{(j)}$$
and $`g_{\mathrm{YM}}^{(\mu )}`$ the gauge coupling on this brane. They imply
$$M_s\alpha _{\mathrm{YM}}^{(\mu )}M_{pl}\frac{V_\mu }{\sqrt{V_6}},$$
Therefore, for the TeV scenario to work one needs
$$\frac{V_\mu }{\sqrt{V_6}}<<1$$
for all D6-branes. However, chirality for the fermionic spectrum of an open string stretched between any two D6-branes implied that the two branes in question do not lie on top of each other on any of the three $`T_{(j)}^2`$ tori. In other words the two branes already span the entire torus and the condition (4.1) cannot be realized.
5. Conclusions
In this paper we have investigated type I string compactifications on noncommutative tori, which are due to constant magnetic fields along the world volumes of the D9-branes being wrapped around the internal space. The key features of these models in six and four dimensions are the presence of chiral fermions, low rank gauge groups and broken space-time supersymmetry. In fact, we found a four-dimensional model with standard model gauge group $`SU(3)\times SU(2)\times U(1)_Y`$ (times some Abelian flavour gauge groups) with four generations of standard model fermions and also a four-dimensional 24 generation $`SU(5)`$ GUT-like model. On the other hand, while displaying promising features like getting chiral spectra, these simple models suffer from other phenomenological difficulties, like even numbers of generations and the problem of splitting the string and the Planck scale sufficiently. It remains to be seen whether more complicated backgrounds with magnetic flux can improve this situation.
Since the deformation parameters of the noncommutative tori, we have considered, correspond to rational (up to some trivial volume dependence) magnetic fluxes, it was always possible to choose a T-dual description where the internal space is commutative, but instead the various D-branes intersect at particular rational angles. It would be interesting to find also noncommutative compactifications which do not allow for an equivalent, T-dual commutative description. It would be also interesting to discuss the properties of the effective quantum field theories which originate from noncommutative compactifications along the lines of ref..
Acknowledgements
The group is supported in part by the EEC contract ERBFMRXCT96-0045. B. K. also wants to thank the Studienstiftung des deutschen Volkes for support.
References
relax E. Witten, Search for realistic Kaluza-Klein theory, Nucl. Phys. B186 (1981) 412. relax P. Candelas, G. Horowitz, A. Strominger and E. Witten, Vacuum configurations for superstrings, Nucl. Phys. B258 (1985) 46. relax L. Dixon, J.A. Harvey, C. Vafa and E. Witten, Strings on orbifolds, Nucl. Phys. B261 (1985) 678. relax W. Lerche, D. Lüst and A.N. Schellekens, Chiral, four-dimensional heterotic strings from self-dual lattices, Nucl. Phys. B287 (1987) 477. relax H. Kawai, D. Lewellen and S.H. Tye, Construction of fermionic string models in four dimensions, Nucl. Phys. B288 (1987) 1; I. Antoniadis, C. Bachas and C. Kounnas, Four-dimensional superstrings, Nucl. Phys. B289 (1987) 87. relax M. Douglas and G. Moore, D-branes, quivers and ALE instantons, hep-th 9603167; I.R. Klebanov and E. Witten, Superconformal field theory on three-branes at a Calabi-Yau singularity, Nucl. Phys. B536 (1998) 199, hep-th/9807080. relax K. Landsteiner, E. Lopez and D. Lowe, Duality of chiral $`N=1`$ supersymmetric gauge theories via branes, JHEP 9802 (1998) 007, hep-th/9801002; I. Brunner, A. Hanany, A. Karch and D. Lüst, Brane dynamics and chiral non-chiral transitions, Nucl. Phys. B528 (1998) 197, hep-th/9801017; A. Hanany and A. Zaffaroni, On the realization of chiral four-dimensional gauge theories using branes, JHEP 9805 (1998) 001, hep-th/9801134. relax C. Angelantonj, M. Bianchi, G. Pradisi, A. Sagnotti and Y. Stanev, Chiral asymmetry in four-dimensional open string vacua, Phys. Lett. B385 (1996) 96, hep-th/9606169. relax P. Horava and E. Witten, Heterotic and type I string dynamics from eleven dimensions, Nucl. Phys. B460 (1996) 506, hep-th/9510209; E. Witten, Strong coupling expansion of Calabi-Yau compactification, Nucl. Phys. B471 (1996) 135, hep-th/9602070; M. Faux, D. Lüst and B.A. Ovrut, Intersecting orbifold planes and local anomaly cancellation in M-theory, Nucl. Phys. B554 (1999) 437, hep-th/9903028; V. Kaplunovsky, J. Sonnenschein, S. Theisen and S. Yankielowicz, On the duality between perturbative orbifolds and M-theory on $`T^4/Z_N`$, hep-th/9912144; M. Faux, D. Lüst and B.A. Ovrut, Local Anomaly Cancellation, M-theory orbifolds and phase transitions, hep-th/0005251. relax J.P. Derendinger, L.E. Ibanez and H.P. Nilles, On the low-energy $`D=4`$, $`N=1`$ supergravity theory extracted from the $`D=10`$, $`N=1`$ superstring, Phys. Lett. B155 (1985) 65; M. Dine, R. Rohm, N. Seiberg and E. Witten, Gluino condensation in superstring models, Phys. Lett. B156 (1985) 55. relax J. Scherk, J.H. Schwarz, Spontaneous breaking of supersymmetry through dimensional reduction, Phys. Lett. B82 (1979) 60; How to get masses from extra dimensions, Nucl. Phys. B153 (1979) 61; E. Cremmer, J. Scherk, J.H. Schwarz, Spontaneously broken $`𝒩=8`$ supergravity, Phys. Lett. B84 (1979) 83; S. Ferrara, C. Kounnas, M. Porrati, F. Zwirner, Superstrings with spontaneously broken supersymmetry and their effective theories, Nucl. Phys. B318 (1989) 75. relax I. Antoniadis, E. Dudas, A. Sagnotti, Supersymmetry breaking, open strings and M-theory, Nucl. Phys. B544 (1999) 469, hep-th/9807011; I. Antoniadis, G. D’Appollonio, E. Dudas, A. Sagnotti, Partial breaking of supersymmetry, open strings and M-theory, Nucl. Phys. B553 (1999) 133, hep-th/9812118; Open descendants of $`\mathrm{𝖹𝖹}_2\times \mathrm{𝖹𝖹}_2`$ freely acting orbifolds, Nucl. Phys. B565 (2000) 123, hep-th/9907184. relax C. Angelantonj, Non-tachyonic open descendants of the 0B string theory, Phys. Lett. B444 (1998) 309, hep-th/9810214; R. Blumenhagen, A. Font and D. Lüst, Tachyon free orientifolds of type 0B strings in various dimensions, Nucl. Phys. B558 (1999) 159, hep-th/9904069; R. Blumenhagen, A. Font, A. Kumar and D. Lüst, Aspects of type 0 string theory, Class. Quant. Grav. 17 (2000) 989, hep-th/9908155; R. Blumenhagen and A. Kumar, A Note on Orientifolds and Dualities of Type 0B String Theory, Phys.Lett. B464 (1999) 46, hep-th/9906234. relax I. Antoniadis, E. Dudas and A. Sagnotti, Brane supersymmetry breaking, Phys. Lett. B464 (1999) 38, hep-th/9908023; G. Aldazabal and A.M. Uranga, Tachyon-free Non-supersymmetric Type IIB Orientifolds via Brane-Antibrane Systems, JHEP 9910 (1999) 024, hep-th/9908072; G. Aldazabal, L.E. Ibanez and F. Quevedo, Standard-like Models with Broken Supersymmetry from Type I String Vacua, JHEP 0001 (2000) 031, hep-th/9909172; A D-Brane Alternative to the MSSM , hep-ph/0001083; C. Angelantonj, I. Antoniadis, G. D’Appollonio, E. Dudas, A. Sagnotti, Type I vacua with brane supersymmetry breaking, hep-th/9911081; C. Angelantonj, R. Blumenhagen and M.R. Gaberdiel, Asymmetric orientifolds, brane supersymmetry breaking and non BPS branes, hep-th/0006033. relax E. Witten; New issues in manifolds of $`SU(3)`$ holonomy, Nucl. Phys. B268 (1986) 79. relax L.E. Ibanez, H.P. Nilles and F. Quevedo, Reducing the rank of the gauge group in orbifold compactifications of the heterotic string, Phys. Lett. B192 (1987) 332. relax A. Connes, M.R. Douglas and A. Schwarz, Noncommutative Geometry and Matrix Theory: Compactification on Tori, JHEP 9802 (1998) 003, hep-th/9711162; M.R. Douglas and C. Hull, D-branes and the Noncommutative Torus, JHEP 9802 (1998) 008, hep-th/9711165. relax R. Blumenhagen, L. Görlich, B. Körs and D. Lüst, Asymmetric Orbifolds, Noncommutative Geometry and Type I Vacua, hep-th/0003024. relax C.G. Callan, C. Lovelace, C.R. Nappi and S.A. Yost, String Loop Corrections To Beta Functions, Nucl. Phys. B288 (1987) 525; A. Abouelsaood, C.G. Callan, C.R. Nappi and S.A. Yost, Open Strings in Background Gauge Fields, Nucl. Phys. B280 (1987) 599; Y.-K.E. Cheung and M. Krogh, Noncommutative Geometry from D0-branes in a Background B-field, Nucl.Phys. B528 (1998) 185, hep-th/9803031; J. Fröhlich, O. Grandjean and A. Recknagel, Supersymmetric Quantum Theory and Noncommutative Geometry, Commun. Math. Phys. 203 (1999) 119, math-ph/9807006; F. Ardalan, H. Arfaei and M. M. Sheikh-Jabbari, Dirac Quantization of Open Strings and Noncommutativity in Branes, hep-th/9906161; C.-S. Chu and P.-M. Ho, Noncommutative Open String and D-brane, Nucl.Phys. B550 (1999) 151, hep-th/9812219; Constrained Quantization of Open String in Background B Field and Noncommutative D-brane, hep-th/9906192; J. Madore, S. Schraml, P. Schupp and J. Wess, Gauge theory on noncommutative spaces, hep-th/0001203; B. Jurco, P. Schupp and J. Wess, Noncommutative gauge theory for Poisson manifolds, hep-th/0005005. relax V. Schomerus, D-branes and Deformation Quantization, JHEP 9906 (1999) 030, hep-th/9903205. relax N. Seiberg and E. Witten, String Theory and Noncommutative Geometry, JHEP 9909 (1999) 032, hep-th/9908142. relax F. Ardalan, H. Arfaei and M. M. Sheikh-Jabbari, Noncommutative Geometry From Strings and Branes, JHEP 9902 (1999) 016, hep-th/9810072. relax C. Bachas, A way to break supersymmetry, hep-th/9503030. relax R. Blumenhagen, L. Görlich and B. Körs, Supersymmetric Orientifolds in 6D with D-Branes at Angles, Nucl. Phys. B569 (2000) 209, hep-th/9908130; A New Class of Supersymmetric Orientifolds with D-Branes at Angles, hep-th/0002146; Supersymmetric 4D Orientifolds of Type IIA with D6-branes at Angles, JHEP 0001 (2000) 040, hep-th/9912204. relax I. Antoniadis, E. Kiritsis and T.N. Tomaras, A D-brane alternative to unification, hep-th/0004214; G. Aldazabal, L. E. Ibanez, F. Quevedo and A. M. Uranga, D-Branes at Singularities : A Bottom-Up Approach to the String Embedding of the Standard Model, hep-th/0005067; A. Krause, A small cosmological constant, grand unification and warped geometry, hep-th/0006226. relax B. Chen, H. Itoyama, T. Matsuo and K. Murakami, p-p’ System with B-field, Branes at Angles and Noncommutative Geometry, hep-th/9910263. relax C.-S. Chu, Noncommutative Open String: Neutral and Charged, hep-th/0001144. relax C. Angelantonj, I. Antoniadis, E. Dudas and A. Sagnotti, in preparation. relax M. Berkooz, M.R. Douglas and R.G. Leigh, Branes Intersecting at Angles, Nucl. Phys. B480 (1996) 265, hep-th/9606139. relax J. Gomis, T. Mehen and M.B. Wise, Quantum field theories with compact noncommutative extra dimensions, hep-th/0006160.
|
warning/0007/math-ph0007011.html
|
ar5iv
|
text
|
# Fusion rules for the continuum sectors of the Virasoro algebra with 𝑐=1
## 1 Introduction
“Fusion rules” describe the product of two superselection charges and the decomposition of the product into irreducible charges. They thus constitute an important characteristics for the charge structure of a quantum field theory.
The general definition of the composition of charges (“DHR product”) was first given in . In two-dimensional conformal quantum field theory, other notions of fusion became more popular, but every evidence shows that these describe the same abstract charge structure.
The actual computation of the fusion rules in concrete models is in general a difficult task, and almost always relies on some specific apriori knowledge. If the QFT at hand is the fixpoint subalgebra of another QFT with respect to a compact gauge group, then harmonic analysis determines the composition law for those sectors which appear in the decomposition of the vacuum sector of the larger algebra . The fusion rules then follow the composition of the representations of the gauge group. In low-dimensional theories, a gauge group is in general not present, but in favorable cases, modular transformation properties or “null vectors” can be exploited.
In the present letter we treat a model where the standard strategies are not applicable: the chiral stress-energy tensor of a 1+1-dimensional conformal quantum field theory with $`c=1`$. (A chiral field can be treated like a “one-dimensional QFT”.) Its algebra $`A`$ is the fixpoint algebra of the chiral $`SU(2)`$, level $`k=1`$, current algebra $`B`$ with respect to its global $`SU(2)`$ symmetry , and the positive-energy representations of the current algebra contain a discrete series of superselection sectors of $`A`$. But besides the discrete series there is a continuum of further sectors which do not arise by restriction from $`B`$. These sectors have no “null vectors” and hence infinite asymptotic dimension , so that the Verlinde formula or Nahm’s prescription are not applicable.
We adopt a method due to Fredenhagen for the computation of the fusion rules: A charged state $`\omega `$ is described by a positive map $`\chi `$ of the algebra into itself such that
$$\omega =\omega _0\chi $$
where $`\omega _0`$ is the vacuum state. The correspondence between states and positive maps is 1:1 provided the charge is strictly localized. This yields a product of states defined by
$$\omega _1\times \omega _2:=\omega _0\chi _1\chi _2.$$
The GNS representation $`\pi _{\omega _1\times \omega _2}`$ is always a subrepresentation of the DHR product of GNS representations $`\pi _{\omega _1}\times \pi _{\omega _2}`$ , and is expected to exhaust it as the positive maps vary within their equivalence class.
For two states $`\omega _1`$ and $`\omega _2`$ belonging to the discrete and continuous sectors, respectively, we shall determine (by a new method) the sectors to which the product states belong.
## 2 Fusion rules for $`c=1`$
The superselection sectors of the stress-energy tensor with $`c=1`$ are uniquely determined by their ground state energy $`h0`$ for the conformal Hamiltonian $`L_0`$. The sectors $`[h=s^2]`$ with $`s_0`$, arise as subrepresentations of the vacuum representation of the $`SU(2)`$ current algebra $`B`$, and those with $`s_0+\frac{1}{2}`$ arise in the spin-$`\frac{1}{2}`$ representation of $`B`$. Those with $`h(\frac{1}{2})^2`$ constitute the continuum. For each of these representations, the partition function is well known :
$$\text{Tr}\mathrm{exp}(\beta \pi _h(L_0))=\{\begin{array}{ccc}t^hp(t)\hfill & \text{if}& h(\frac{1}{2})^2,\hfill \\ (t^{s^2}t^{(s+1)^2})p(t)\hfill & \text{if}& h=s^2,s\frac{1}{2}_0,\hfill \end{array}$$
where $`t=\text{e}^\beta `$ and $`p(t)=_n(1t^n)^1`$.
The positive maps describing the charged states are of the form (cf. Lemma 2.1)
$$\chi =\mu \alpha _g|_A.$$
Here $`g`$ is a smooth $`SU(2)`$ valued function, and $`\alpha _g`$ the automorphism of the current algebra $`B`$ induced from the local gauge transformation (Bogolyubov automorphism) of the underlying chiral fermion doublet,
$$\psi _i(x)\underset{j}{}\psi _j(x)g_{ji}(x).$$
$`\mu =𝑑\mu (k)\gamma _k`$ is the average over the global gauge group $`SU(2)`$ acting by automorphisms $`\gamma _k`$. Since $`\mu `$ is a positive map of $`B`$ onto $`A`$, $`\chi _g`$ is a positive map of $`A`$ onto $`A`$.
The induced action of $`\alpha _g`$ on the currents $`j(f)j^a(f_a)=:\psi (x)f(x)\psi (x)^{}:dx`$ (with an $`su(2)`$ valued test function $`f(x)=f_a(x)T^a`$) is explicitly computed as
$$\alpha _g(j(f))=j(gfg^1)\frac{i}{2\pi }\text{Tr}(fg^1g)\mathrm{𝟏},$$
and its restriction to the Sugawara stress-energy tensor $`T=\frac{\pi }{3}g_{ab}:j^aj^b:`$ is
$$\alpha _g(T(f))=T(f)ij(fgg^1)\frac{1}{4\pi }f\text{Tr}(gg^1gg^1)\mathrm{𝟏}.$$
The central terms arise, of course, from normal ordering. To be specific, we choose the functions
$$g_q(x)=\left(\begin{array}{cc}\mathrm{exp}(iq\lambda (x))& 0\\ 0& \mathrm{exp}(iq\lambda (x))\end{array}\right)$$
where $`\lambda (x)=i\mathrm{log}\frac{1+ix}{1ix}`$ interpolates between $`\lambda (\mathrm{})=\pi `$ and $`\lambda (+\mathrm{})=+\pi `$, and $`q`$ is a real parameter whose role as a charge will be exhibited in Lemma 2.1. <sup>1</sup><sup>1</sup>1It appears that one could also use the embedding $`T=\pi :jj:`$ of $`A`$ into a $`U(1)`$ current algebra $`C`$. The problem would be that the conditional expectation $`\mu `$ which takes the homomorphisms $`\alpha _g:AC`$ back onto $`A`$ in order to obtain $`\chi =\mu \alpha _q|_A`$ is not explicitly known in that case.
At this point, we have to distinguish the quasilocal algebras $`A_{\mathrm{local}}`$ and $`B_{\mathrm{local}}`$ generated by field operators smeared with test functions, and the global algebras $`A_{\mathrm{global}}`$ and $`B_{\mathrm{global}}`$ generated by field operators smeared with “admissible” functions which are test functions up to polynomials of order $`2(d1)`$ where $`d`$ is the scaling dimension. It is well known that the fields as distributions extend to these enlarged test function spaces, so that
$$L_n=\frac{1}{2}(1ix)^{1n}(1+ix)^{1+n}T(x)𝑑x\text{and}Q_n^a=(1ix)^n(1+ix)^nj^a(x)𝑑x$$
are defined as closed unbounded operators. The specific automorphisms $`\alpha _q\alpha _{g_q}`$ extend to the operators $`Q_n^3B_{\mathrm{global}}`$ and $`L_nA_{\mathrm{global}}`$:
$$\alpha _q(Q_n^3)=Q_n^3+q\delta _{n,0},\alpha _q(L_n)=L_n+2qQ_n^3+q^2\delta _{n,0}(q),$$
but they extend to $`Q_n^\pm B_{\mathrm{global}}`$ only if $`q\frac{1}{2}`$:
$$\alpha _q(Q_n^\pm )=Q_{n\pm 2q}^\pm (q\frac{1}{2}).$$
(Our basis of $`SU(2)`$ and hence of the fields $`j^a`$ is such that $`[Q_n^+,Q_m^{}]=2Q_{n+m}^3+n\delta _{n+m,0}`$, $`[Q_n^3,Q_m^\pm ]=Q_{n+m}^\pm `$, $`[Q_n^3,Q_m^3]=\frac{1}{2}n\delta _{n+m,0}`$.)
Our first Lemma establishes the relation between the parameters $`q`$ and $`h`$:
2.1. Lemma: The state $`\omega _q\omega _0\chi _q\omega _0\mu \alpha _q|_A=\omega _0\alpha _q|_A`$ is a ground state in the irreducible sector $`[h=q^2]`$.
Proof: Since the operators $`L_n`$ and $`Q_n^3`$ ($`n0`$) annihilate the vacuum, $`\alpha _q(L_n)`$ annihilate the vacuum for $`n>0`$ and $`\alpha _q(L_0)`$ has eigenvalue $`q^2`$. It follows that $`\omega _q`$ is a ground state for $`L_0`$ with ground state energy $`q^2`$. Q.E.D.
Thus, in order to compute the fusion rules $`[h_1]\times [h_2]`$ (where $`h_i=q_i^2`$) one has to determine the GNS representation for the product state
$$\omega _{q_1}\times \omega _{q_2}=\omega _0\alpha _{q_1}\mu \alpha _{q_2}|_A=_{SU(2)}𝑑\mu (k)\omega _0\alpha _{q_1}\gamma _k\alpha _{q_2}|_A.$$
This state is a continuous mixture of states $`\omega _k\omega _0\alpha _k`$ induced by the homomorphisms
$$\alpha _k\alpha _{q_1}\gamma _k\alpha _{q_2}|_A$$
of $`A_{\mathrm{local}}`$ into $`B_{\mathrm{local}}`$. (We suppress the explicit reference to the involved charges $`q_1`$ and $`q_2`$.) These homomorphisms extend to $`A_{\mathrm{global}}`$ for generic $`kSU(2)`$ only if $`q_1\frac{1}{2}`$, as can be seen from the above transformation formulae. The following argument is more physical: If one evaluates $`\omega _k(T(f)^2)`$ for test functions $`f`$, then one finds that the contributions from the current two-point functions diverge for generic $`q`$ as $`f`$ is replaced by the function $`\frac{1}{2}(1+x^2)`$. Hence the operator $`L_0=\frac{1}{2}(1+x^2)T(x)𝑑x`$ has a finite expectation value but infinite variance in these states.
This is why we shall restrict ourselves to the case $`q_1\frac{1}{2}`$. Since $`q`$ and $`q`$ give rise to the same sector $`[h=q^2]`$, we shall even assume $`q_1\frac{1}{2}_0`$.
Now we exploit the fact that $`\gamma _k`$ is implemented by a unitary operator in $`B_{\mathrm{global}}`$ of the form $`U(k)=\mathrm{exp}(i\kappa _aQ_0^a)`$ on which $`\alpha _{q_1}`$ is well defined. Hence
$$\alpha _k=\text{Ad}(V(k))\alpha _{q_1}\alpha _{q_2}|_A=\text{Ad}(V(k))\alpha _{q_1+q_2}|_A$$
with $`V(k)=\alpha _{q_1}(U(k))=\mathrm{exp}(i\kappa _a\alpha _{q_1}(Q_0^a))`$. It is more convenient to express $`U(k)`$ in the form
$$U(k)=\mathrm{exp}(i\frac{k_2^{}}{k_1}Q_0^{})k_1^{2Q_0^3}\mathrm{exp}(i\frac{k_2}{k_1}Q_0^+)\text{for}k=\left(\begin{array}{cc}k_1& ik_2\\ ik_2^{}& k_1^{}\end{array}\right).$$
($`k_1^{2Q_0^3}`$ is well defined since $`2Q_0^3`$ has integer spectrum.) Application of $`\alpha _{q_1}`$ yields
$$V(k)=k_1^{2q_1}\mathrm{exp}(i\frac{k_2^{}}{k_1}Q_{2q_1}^{})k_1^{2Q_0^3}\mathrm{exp}(i\frac{k_2}{k_1}Q_{+2q_1}^+).$$
2.2. Lemma: The product state $`\omega _{q_1}\times \omega _{q_2}`$ is a convex integral over states $`\omega _0\alpha _k`$, $`kSU(2)`$. Each state $`\omega _0\alpha _k`$ on $`A`$ is a finite convex sum
$$\omega _0\alpha _k=\underset{\nu =0}{\overset{2q_1}{}}\left(\begin{array}{c}2q_1\\ \nu \end{array}\right)|k_1|^{2(2q_1\nu )}|k_2|^{2\nu }\omega _{q_1,q_2}^{(\nu )}$$
of states
$$\omega _{q_1,q_2}^{(\nu )}()=\frac{(2q_1\nu )!}{(2q_1)!\nu !}((Q_{2q_1}^{})^\nu \mathrm{\Omega },\alpha _{q_1+q_2}()(Q_{2q_1}^{})^\nu \mathrm{\Omega }).$$
Since only the weights depend on the group element $`kSU(2)`$, the product state $`\omega _{q_1}\times \omega _{q_2}`$ is a finite convex sum of the same states $`\omega _{q_1,q_2}^{(\nu )}`$.
Proof: The first statement just summarizes the precedent discussion. We have $`\omega _0\alpha _k=(V(k)^{}\mathrm{\Omega },\alpha _{q_1+q_2}()V(k)^{}\mathrm{\Omega })`$, and $`V(k)^{}\mathrm{\Omega }=(k_1^{})^{2q_1}\mathrm{exp}(i\frac{k_2^{}}{k_1^{}}Q_{2q_1}^{})\mathrm{\Omega }`$ because $`Q_n^a`$ annihilate the vacuum for $`n0`$ (remember our choice $`q_1\frac{1}{2}_0`$). The power series expansion of the exponential yields vectors $`(Q_{2q_1}^{})^\nu \mathrm{\Omega }`$ with energy $`2q_1\nu `$ and Cartan charge (the eigenvalue of $`Q_0^3`$) $`C=\nu `$. These vectors vanish for $`\nu >2q_1`$ because the vacuum Hilbert space $`H`$ of $`B`$ does not contain vectors with energy less than $`C^2`$. This fact is read off the following expression for the partition function for the vacuum representation:
$$\text{Tr}\mathrm{exp}(\beta L_0\eta Q_0^3)=\underset{j_0}{}\underset{m=j}{\overset{j}{}}z^m(t^{j^2}t^{(j+1)^2})p(t)(z=\text{e}^\eta ,t=\text{e}^\beta )$$
in which the power of $`t`$ is always at least the square of the power of $`z`$. Since $`\alpha _{q_1+q_2}(L_n)`$ does not change the Cartan charge $`C`$, the vectors $`(Q_{2q_1}^{})^\nu \mathrm{\Omega }`$ have only diagonal matrix elements for $`\alpha _{q_1+q_2}(A)`$, showing the convex decomposition. The proper normalization of the states $`\omega _{q_1,q_2}^{(\nu )}(1)=1`$ can be checked recursively in $`\nu `$. Q.E.D.
The problem has thus been reduced to the determination of the GNS representations $`\pi _{q_1,q_2}^{(\nu )}`$ for the states $`\omega _{q_1,q_2}^{(\nu )}`$. One can easily compute that these states are eigenstates of $`L_0`$ with energy $`(q_1+q_2)^22\nu q_2`$, but they are not ground states in general. It is therefore not possible to determine the sectors directly via their ground state energies. Instead, it turns out to be possible to compute the partition function for the representations induced by these states. This is our main result.
2.3. Proposition: Let $`q_1\frac{1}{2}_0`$. If $`q_2\frac{1}{2}`$, then $`\pi _{q_1,q_2}^{(\nu )}`$ is irreducible and belongs to the sector $`[h=(q_1+q_2\nu )^2]`$. If $`q_2\frac{1}{2}`$, then $`\pi _{q_1,q_2}^{(\nu )}`$ is a direct sum of sectors from the set $`\{[h=s^2]:s|q_1+q_2\nu |+_0\}`$.
Proof: The vector $`(Q_{2q_1}^{})^\nu \mathrm{\Omega }`$ has Cartan charge $`C=\nu `$. This value is not changed by application of $`\alpha _{q_1+q_2}(L_n)`$, hence $`\pi _{q_1,q_2}^{(\nu )}`$ is a subrepresentation of the representation $`\alpha _{q_1+q_2}`$ on the subspace $`H_{C=\nu }=P_\nu H`$ of Cartan charge $`\nu `$ in the vacuum representation of $`B`$. The partition function for the latter representation is
$$\text{Tr}P_\nu \mathrm{exp}(\beta \alpha _{q_1+q_2}(L_0))=\text{e}^{(q_1+q_2)^2\beta }\text{Tr}P_\nu \mathrm{exp}(\beta L_02(q_1+q_2)\beta Q_0^3).$$
From the previous expression for the vacuum partition function, we obtain
$$\text{Tr}P_\nu \mathrm{exp}(\beta L_0\eta Q_0^3)=z^\nu t^{\nu ^2}p(t)(z=\text{e}^\eta ,t=\text{e}^\beta )$$
by collecting the terms $`z^\nu `$, and hence
$$\text{Tr}P_\nu \mathrm{exp}(\beta \alpha _{q_1+q_2}(L_0))=t^{(q_1+q_2\nu )^2}p(t).$$
If $`q_1+q_2\nu \frac{1}{2}`$, then this is the partition function of the irreducible sector $`[h=(q_1+q_2\nu )^2]`$. Hence $`\alpha _{q_1+q_2}(A)`$ acts irreducibly on $`H_{C=\nu }`$, and must coincide with its subrepresentation $`\pi _{q_1,q_2}^{(\nu )}`$. If on the other hand $`q_1+q_2\nu \frac{1}{2}`$, then the above equals the sum of the partition functions $`(t^{s^2}t^{(s+1)^2})p(t)`$ of the sectors $`[h=s^2]`$ with $`s|q_1+q_2\nu |+_0`$. Thus $`\pi _{q_1,q_2}^{(\nu )}`$ is the direct sum of a subset of these sectors. Q.E.D.
As mentioned in the introduction, the product of states, computed here, might accidentally not exhaust the DHR product. But this degeneracy disappears if the positive map $`\chi _2`$ is perturbed by the adjoint action of some isometry $`aA`$. We note that the argument leading to Prop. 2.3 is in fact stable if $`\chi _{q_2}`$ is replaced by $`\text{Ad}(a^{})\chi _{q_2}`$. Namely, because $`a`$ is $`SU(2)`$-invariant, one has $`\text{Ad}(a^{})\gamma _k\alpha _{q_2}=\text{Ad}(U(k)a^{})\alpha _{q_2}`$, so it is sufficient to replace in the above argument the vectors $`(Q_{2q_1}^{})^\nu \mathrm{\Omega }`$ by the perturbed vectors $`\alpha _{q_1}(a)(Q_{2q_1}^{})^\nu \mathrm{\Omega }`$ which still belong to $`H_{C=\nu }`$. In the case $`q_2\frac{1}{2}`$, the perturbed GNS representation $`\pi _{q_1,q_2}^{(\nu )}`$ will still belong to the irreducible sector $`[h=(q_1+q_2\nu )^2]`$.
Thus, combining Lemma 2.2 with the Proposition, we obtain
2.4. Corollary: Let $`q_1\frac{1}{2}_0`$ and $`q_2\frac{1}{2}`$. The fusion rules for the sectors $`[h_i=q_i^2]`$ are
$$[h_1]\times [h_2]=\underset{\nu =0}{\overset{2q_1}{}}[h^{(\nu )}]\text{with}h^{(\nu )}=(q_1+q_2\nu )^2.$$
## 3 Comments
We have studied the decomposition into irreducibles of the product of sectors (“fusion rules”) for the chiral stress-energy tensor with $`c=1`$. We succeeded to compute the fusion rules for two sectors with ground state energies $`h_i`$ where $`[h_1]`$ is a special sector, $`h_1(\frac{1}{2}_0)^2`$, and $`[h_2]`$ belongs to the continuum of sectors, $`h_2_+(\frac{1}{2}_0)^2`$, Cor. 2.4. This result was not accessible by the prevailing methods for the computation of fusion rules. The case where both sectors belong to the continuum should in principle also be studied with the present method, but becomes technically very intricate.
When both sectors $`[h_i]`$ are special, we would have expected $`SU(2)`$-like fusion rules since the special sectors $`[h=s^2]`$, $`s\frac{1}{2}_0`$, arise by restriction of the vacuum and spin-$`\frac{1}{2}`$ representations of $`B`$ to the fixpoint algebra $`A`$ on the subspaces of $`SU(2)`$ charge $`s`$. This is, however, not reproduced by Prop. 2.3 and Lemma 2.2: Although the unperturbed states $`\omega _{q_1,q_2}^{(\nu )}`$ have finite energy and hence only finitely many of the possible sectors according to Prop. 2.3 really contribute to them, this limitation will disappear if $`\chi _{q_2}`$ is perturbed as described above. Moreover, if $`h_i=s_i^2`$ with $`0<s_2<s_1`$, the sectors $`[h=s^2]`$ with $`0s<|s_1s_2|`$ should not occur according to $`SU(2)`$, while they are not excluded by Prop. 2.3, and are really found to be present by more explicit computations.
This state of affairs has a simple explanation: For $`q\frac{1}{2}`$, the positive maps $`\chi _q`$ transfer not only the $`SU(2)`$ charge $`s=|q|`$ but in fact, as explained below, a mixture of all charges $`s|q|+_0`$. These admixtures are not seen if evaluated in the vacuum state (Lemma 2.1), but become visible if evaluated in a generic state of $`A`$, e.g., upon perturbation of $`\chi _q`$. The product states $`\omega _0\chi _{q_1}\chi _{q_2}`$, too, are sensitive to admixtures to $`\chi _{q_2}`$, which accounts for the presence of “too many” sectors contributing to the fusion rules as inferred from Lemma 2.2 and Prop. 2.3.
Let us explain why $`\chi _q`$ is capable of transferring the “wrong” charges if $`q\frac{1}{2}`$, but not if $`q\frac{1}{2}`$, and why this is not in conflict with the statement in that the correspondence between states and positive maps is 1:1. The argument is very similar to the one in the proof of Prop. 2.3. If $`\chi _q`$ is evaluated in some perturbed state $`\omega =(a\mathrm{\Omega },a\mathrm{\Omega })`$ with $`aA`$, we have $`\omega \chi _q=\omega \alpha _q`$ since $`a`$ and $`\omega _0`$ are $`SU(2)`$ invariant. Thus the GNS representation $`\pi _\omega `$ for $`\omega `$ is a subrepresentation of the representation $`\alpha _q`$ on the subspace $`H_{C=0}=P_0H`$ of Cartan charge $`C=0`$ in the vacuum representation of $`B`$ (to which $`a\mathrm{\Omega }`$ belongs). The partition function for this representation has been computed above (putting $`q_1=0,\nu =0,q_2=q`$):
$$\text{Tr}P_0\mathrm{exp}(\beta \alpha _q(L_0))=t^{q^2}p(t).$$
This is the character of the irreducible representation $`[h=q^2]`$ if $`q\frac{1}{2}`$, but is the sum of infinitely many irreducible characters for $`[h=s^2]`$, $`s|q|+_0`$, if $`q\frac{1}{2}`$.
By testing with suitable operators $`aA_{\mathrm{global}}`$, one finds that the “wrong” sectors are indeed present. Remember that the 1:1 correspondence between states and positive maps requires that the charge is strictly localized, while the automorphisms $`\alpha _q`$ in our analysis are only asymptotically localized (the derivative $`g_q(x)`$ vanishes asymptotically). Of course our choice for $`\alpha _q`$ was dictated by the simplicity of the transformation formulae for $`L_n`$ and $`Q_n^a`$. The unpleasant feature of the wrong sectors is the price for that simplification.
The fusion rules in Cor. 2.4 are not affected by this complication.
This work is based on the Diploma Thesis of the second author .
|
warning/0007/cond-mat0007431.html
|
ar5iv
|
text
|
# Impurity in a 𝑑-wave superconductor: Kondo effect and STM spectra
\[
## Abstract
We present a theory for recent STM studies of Zn impurities in the superconductor $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_{8+\delta }`$, using insights from NMR experiments which show that there is a net $`S=1/2`$ moment on the Cu ions near the Zn. We argue that the Kondo spin dynamics of this moment is the origin of the low bias peak in the differential conductance, rather than a resonance in a purely potential scattering model. The spatial and energy dependence of the STM spectra of our model can also fit the experiments.
\]
Recent progress in scanning tunneling microscopy (STM) of surfaces of the high temperature superconductor $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_{8+\delta }`$ has given us a high resolution probe of its electronic correlations. Especially notable have been studies in the vicinity of Zn impurities on the Cu sites, and the site-specific information on the variation in the pairing correlations. The $`\mathrm{Zn}^{++}`$ ion has spin $`S=0`$, and so it appears natural to interpret the experiments using a theoretical model in which each Zn site acts a potential scatterer of Cu $`3d`$ electrons forming a $`d`$-wave BCS state. However, several key experimental features do not fit easily in this simple picture. There is a large peak in the differential conductance close to zero bias near the Zn site: it is tempting to identify this peak with a quasi-bound state in the potential scattering model, but such a state appears at low energies only for a range of very large potential values depending upon microscopic details . Furthermore, the spatial dependence of the zero bias peak is unexpected : it is largest on the Zn site, and with a local maximum on the second neighbor Cu sites (see Fig 3c in Ref ), and this differs from the general expectations of the potential scattering model . Finally, the observed spatially integrated spectrum is asymmetric between positive and negative bias , while the potential model predicts approximate symmetry .
We shall address these questions here using a rather different theoretical model which was originally motivated by other experiments on Zn or Li impurities (the $`\mathrm{Li}^+`$ ion is also $`S=0`$). A series of beautiful NMR experiments have clearly shown that each impurity, despite having no on-site spin, induces a local, unpaired $`S=1/2`$ moment on the Cu ions in its vicinity at intermediate energy scales. In the underdoped regime, this can be understood by the confining property of the host antiferromagnet, in which the impurity is a localized “holon” which binds the moment of a $`S=1/2`$ “spinon”. At larger doping, a related picture can be developed by analogy with the theory for moment formation in the disordered metallic state of Si:P—small variations in the potential combine with strong local correlations to induce very localized spin excitations. Other theoretical perspectives on local moment formation in the cuprates have also been given. Further evidence for $`S=1/2`$ local moments near Zn sites appears from neutron scattering experiments: it has been argued that these are required to explain the strong effects of a small concentration of Zn impurities on a “resonance peak” in the spin dynamic structure factor.
This paper will study a $`S=1/2`$ local moment near the Zn/Li site coupled by exchange interactions to the fermionic $`S=1/2`$ excitations of a $`d`$-wave superconductor. We shall describe the effects of the Kondo screening of the local moment, by the fermionic excitations, on the STM spectra: we shall show that our model leads naturally to a peak in the differential conductance at a low bias of order the Kondo temperature, with a spatial distribution which can fit the experiments. Our spatially integrated spectrum also has a bias asymmetry which increases with decreasing doping.
We begin by describing our model Hamiltonian, $`H=H_{\mathrm{BCS}}+H_{\mathrm{imp}}`$, for a single Zn impurity . The first term describes the host superconductor, which we model by a simple BCS Hamiltonian
$$H_{\mathrm{BCS}}=\underset{𝐤}{}\mathrm{\Psi }_𝐤^{}\left[(\epsilon _𝐤\mu )\tau ^z+\mathrm{\Delta }_𝐤\tau ^x\right]\mathrm{\Psi }_𝐤.$$
Here $`\mathrm{\Psi }_𝐤=(c_𝐤,c_𝐤^{})`$ is a Nambu spinor at momentum $`𝐤=(k_x,k_y)`$ ($`c_{𝐤\alpha }`$ annihilates an electron with spin $`\alpha `$ on a 3d orbital), $`\tau ^{x,y,z}`$ are Pauli matrices in particle-hole space, and $`\mu `$ is a chemical potential. For the kinetic energy, $`\epsilon _𝐤`$ we have first ($`t`$), second ($`t^{}`$), and third ($`t^{\prime \prime })`$ neighbor hopping, while we assume a $`d`$-wave form for the BCS pairing function $`\mathrm{\Delta }_𝐤=(\mathrm{\Delta }_0/2)(\mathrm{cos}k_x\mathrm{cos}k_y)`$. The Zn impurity is at $`𝐫=𝐫_0`$, and is described by
$$H_{\mathrm{imp}}=\underset{𝐫𝒩}{}K(𝐫)\stackrel{}{S}c_\alpha ^{}(𝐫)\frac{\stackrel{}{\sigma }_{\alpha \beta }}{2}c_\beta (𝐫)+Uc_\alpha ^{}(𝐫_0)c_\alpha (𝐫_0),$$
where $`\sqrt{N_s}c_\alpha (𝐫)=_kc_{𝐤\alpha }e^{i𝐤𝐫}`$ ($`N_s`$ is the number of sites in the lattice) annihilates an electron at site $`𝐫`$, $`\stackrel{}{\sigma }`$ are the Pauli matrices in spin space, and $`𝒩`$ is a set of sites in the neighborhood of $`𝐫_0`$. The spin degree of freedom induced by the impurity is represented by the $`S=1/2`$ operator $`\stackrel{}{S}`$. The on-site potential scattering of the Zn ion is represented by $`U`$: previous analyses of STM spectra of Zn ions included only this second term in $`H_{\mathrm{imp}}`$ and omitted the degrees of freedom represented by $`\stackrel{}{S}`$. We assume that effects due to spatial variations in the self-consistent pairing amplitude near the impurity, along with those due to Hartree renormalizations from the Coulomb interactions, have been absorbed into the effective parameters $`K(𝐫)`$ and $`U`$ (as in ); therefore, as the dopant charge carriers are moving in a background of ions that are both doubly charged ($`\mathrm{Cu}^{++}`$ and $`\mathrm{Zn}^{++}`$), we expect $`U`$ to be negative and measuring mainly the shift in the $`d`$ level energy on the Zn (along the same lines, $`U`$ should be larger and positive for $`\mathrm{Li}^+`$ impurities).
An important ingredient in our computation, which influences the spatial form of the STM spectrum, is the $`𝐫`$ dependence of the spin-dependent interaction $`K(𝐫)`$. This is quite difficult to determine from first principles, and we will instead use the spatial information obtained by analyses of Knight shifts in NMR experiments . The $`S=1/2`$ moment is found to be concentrated on the 4 Cu nearest neighbors of the Zn ion (see Fig 1g in and Fig 19 in ), and the moment on the Zn site itself, $`𝐫=𝐫_0`$, is negligible. We expect that the dopant holes will have a strong preference to reside on a Cu site with a moment (as the Cu spins on the other sites have paired with each other and gained exchange energy); this attraction is realized by $`K(𝐫)`$, and we take $`K(𝐫)=K_1`$ for the 4 sites $`𝐫𝐫_0=(\pm 1,0),(0,\pm 1)`$, while $`K(𝐫_0)=0`$. In principle, it is not difficult to also include the smaller $`K(𝐫)`$ values at larger values of $`|𝐫𝐫_0|`$, but we will neglect them here for simplicity. We also note that while the above form for $`K(𝐫)`$ is reasonable, there is no justification for a corresponding form for the potential scattering term, which is surely largest at $`𝐫=𝐫_0`$, as in $`H_{\mathrm{imp}}`$.
The Kondo effect in $`H`$ is related to that in a class of models which have been much studied recently : these models have a single spin coupled to a fermionic bath whose local density of states, $`\rho _{\mathrm{}}(ϵ)`$, vanishes as a power law near the Fermi energy, $`\rho _{\mathrm{}}(ϵ)|ϵ|^r`$, where $`ϵ`$ is measured from the Fermi level. The usual Kondo effect in a Fermi liquid corresponds to $`r=0`$, while the present $`d`$-wave superconductor has $`r=1`$. The value of $`r`$ and the presence or absence of particle-hole symmetry play a key role in determining the low energy physics, and a comprehensive phase diagram has been presented by Ingersent and collaborators . For $`r=1`$ and with perfect particle-hole symmetry (for $`H`$ this corresponds to $`t^{}=0`$, $`\mu =0`$, and $`U=0`$) there is in fact no Kondo effect: the spin is free at low energies for all values of the exchange, $`K`$. However, there is a quantum phase transition at a finite magnitude of particle-hole symmetry breaking and at a finite $`K`$, to a phase in which the spin is Kondo screened below an energy scale $`T_K`$. The universal critical theory for this quantum critical point is not known and we shall not discuss it here.
Here, we shall describe the dynamics of $`H`$ using a “large $`N`$” approach . While this method has numerous artifacts near the quantum-critical point just mentioned, and is not quantitatively accurate, it does capture the qualitative physics of the Kondo screened phase in an effective manner, and this is our primary interest. We represent the spin $`\stackrel{}{S}`$ by a fermion, $`f_\alpha `$, which obeys the constraint $`f_\alpha ^{}f_\alpha =1`$. We impose this by a Lagrange multiplier $`\lambda `$, and decouple the exchange interactions by complex Hubbard-Stratonovich fields $`\phi _𝐫`$; $`\lambda `$ and $`\phi _𝐫`$ are approximated by their static, real saddle point values. The physical quantities are expressed in terms of the Green’s function of the Nambu spinor $`F=(f_{},f_{}^{})`$, which in Matsubara frequency, $`\omega _n`$, is $`𝒯(\omega _n)=F(\omega _n)F^{}(\omega _n)`$. We have
$$𝒯^1(\omega _n)=i\omega _n+\lambda \tau ^z\underset{𝐫,𝐫^{}𝒩}{}\phi _𝐫\phi _𝐫^{}\tau ^zG(𝐫,𝐫^{},\omega _n)\tau ^z$$
where $`G`$ is the $`\mathrm{\Psi }`$ Green’s function with potential scattering alone:
$`G(𝐫`$ $`,𝐫^{},\omega _n)=G^0(𝐫𝐫^{},\omega _n)UG^0(𝐫𝐫_0,\omega _n)`$ (2)
$`\times \tau ^z[1+UG^0((0,0),\omega _n)\tau ^z]^1G^0(𝐫_0𝐫^{},\omega _n);`$
$`G^0`$ is the Green’s function of the host $`N_sG^0(𝐫,\omega _n)=_𝐤e^{i𝐤𝐫}[i\omega _n+(\epsilon _k\mu )\tau ^z+\mathrm{\Delta }_k\tau ^x]^1`$. The values of $`\lambda `$ and $`\phi _𝐫`$ were obtained by numerically solving the constraint equations $`T_{\omega _n}\text{Tr}[\tau ^z𝒯(\omega _n)]=0`$ and $`\phi _𝐫=(TK(𝐫)/2)_{\omega _n}_{𝐫^{}𝒩}\phi _𝐫^{}\text{Tr}[\tau ^z𝒯(\omega _n)\tau ^zG(𝐫^{},𝐫,\omega _n)]`$ ($`T`$ is the temperature, and Boltzmann’s constant $`k_B=1`$). Finally, the tunneling density of states (DOS) was obtained as $`\text{Im Tr}(\stackrel{~}{G}(1+\tau ^z))/2`$, where $`\stackrel{~}{G}`$ is the full $`\mathrm{\Psi }`$ Green’s function
$`\stackrel{~}{G}(𝐫,𝐫^{},\omega _n)=G(𝐫,𝐫^{},\omega _n)`$ (3)
$`+`$ $`{\displaystyle \underset{𝐬,𝐬^{}𝒩}{}}\phi _𝐬\phi _𝐬^{}G(𝐫,𝐬,\omega _n)\tau ^z𝒯(\omega _n)\tau ^zG(𝐬^{},𝐫^{},\omega _n).`$ (4)
We now describe our numerical results. For small $`K_1`$, the constraint equations have only the solution $`\phi _𝐫=0`$. In the present large $`N`$ approach, $`\stackrel{}{S}`$ is completely decoupled from the fermions at such a saddle point. This solution corresponds to the free spin phase, and the STM spectra of the large $`N`$ theory are identical to that of the purely potential scattering model. To include the spin dynamics, we take $`K_1>K_{1c}`$ below so that $`\phi _𝐫0`$; we found that the lowest free energy saddle points had a $`d`$-wave pattern with $`\phi _𝐫=+[]\phi `$ for $`𝐫𝐫_0=(\pm 1,0)[(0,\pm 1)]`$. Provided particle-hole symmetry is absent, the onset of a non-zero $`\phi `$ corresponds to a phase transition to the Kondo screened phase at $`K_{1c}`$ (for the particle-hole symmetric case, there is no Kondo screening , and the large $`N`$ equations show that this is so, even for $`K_1>K_{1c}`$). The assumption of eventual low energy Kondo screening is also in accord with indications in NMR and thermodynamic measurements . The large $`N`$ value for $`K_{1c}`$ is believed to be an overestimate: this are other shortcomings will be addressed in a forthcoming work employing an alternative “non-crossing approximation” .
For $`K_1>K_{1c}`$, the energy dependence of the tunneling spectrum is dominated by the form of the scattering matrix $`𝒯(\omega )`$; the imaginary part of $`𝒯(\omega )`$ shows a pronounced maximum at an energy $`\omega _0`$ (which becomes a very sharp peak near the Kondo transition), so the location of possible peaks in the local DOS is given by $`\pm \omega _0`$. On the other hand, $`|\omega _0|`$ can be also be identified with the Kondo temperature, $`T_K`$. To see this, consider the local impurity susceptiblity, which is given by $`\chi _{\text{l}oc}=(T/4)_{\omega _n}\text{Tr}[𝒯(\omega _n)^2]`$. Assuming the simple pole structure $`𝒯(\omega _n)=(i\omega _n+\omega _0\tau ^z)^1`$ it is evident that $`\chi _{\text{l}oc}`$ changes its character at a $`T`$ of order $`|\omega _0|T_K`$.
We show in Fig 1 the energy and spatial dependence of the local DOS of $`H`$ for a typical set of parameters at 14% hole doping. Most significant is the pronounced peak at $`𝐫_0`$ (on the Zn site) at a bias, $`\omega =\omega _0`$. This was a robust feature of our results: the peak retained a small negative bias and width for a wide range of $`K_1>K_{1c}`$ and $`U<0`$. Away from the central site, the peaks were much weaker, and varied slightly depending upon $`U`$ and doping: for $`0.4`$ eV $`<U<0`$ at optimal doping (and more robustly at lower doping) we obtained the alternating intensity pattern in Fig 1b . The spatial integral of the spectrum at negative bias is larger than that at positive bias, and the ratio is quoted in Fig 1. For comparison we show in Fig 2 the analogous results for a purely potential scattering model ($`K_1=0`$ and $`\stackrel{}{S}`$ absent), where we see dramatically different features. Now the largest peak is on the first neighbors; notice also the change in the color scale in Fig 2, and that this peak is not as pronounced as the Zn site peak in Fig 1. Further, we had to choose a very large value of $`|U|`$ to make the peak sharp and at a small negative bias. Both these features are sensitive to variations in the values of the doping , the band structure ($`t^{}/t`$, $`t^{\prime \prime }/t`$) and $`U`$: it is not difficult to find broader peaks at higher energies, and to switch the largest peak to positive bias.
We make a few remarks on the spatial dependencies in Fig 1. The dominant contribution at peak energy comes from the second term on the right hand side of (4), and both the normal ($`\tau ^z`$) and anomalous ($`\tau ^x`$) components of $`G`$ are largest for sites on opposite sublattices. Using the location of the four sites in $`𝒩`$ and the $`d`$-wave structure of both the superconducting gap function $`\mathrm{\Delta }_𝐤`$ and the $`\phi _𝐫`$ fields, we see that at $`𝐫_0`$ the normal (anomalous) Green’s functions destructively (constructively) interfere leading to a large peak at $`\omega _0`$. This interference is also responsible for the asymmetry of the spatially integrated spectrum. (We note that for an $`s`$-wave pattern in the $`\phi _𝐫`$ fields the interference mechanism is similar, with the difference that the peak at $`𝐫_0`$ appears at $`+\omega _0`$.) An analogous discussion can be applied to Fig 2, where interference of scattering from different $`𝐫`$ is absent. We also note that the coherence peaks, appearing in the DOS at the gap energy, are much suppressed at the impurity site and its neighbors, although we did not include a spatial variation of the pairing amplitude. Furthermore, upon moving to the metallic state by setting $`\mathrm{\Delta }_0=0`$, our approach is closely related to that of , and then the large DOS around the Fermi level leads to a Fano lineshape.
To summarize, this paper has introduced a new model ($`H=H_{\mathrm{BCS}}+H_{\mathrm{imp}}`$) for STM spectra of Zn/Li impurities in the high temperature superconductors. The model appears to be required to explain NMR and neutron scattering measurements on the same system. It is therefore satisfying that our model also leads to a low bias peak in the STM; the measured bias of this peak (1.5 meV) suggests that the Kondo screening of the moment (as can be measured directly in NMR) occurs at a temperature of order 15 K.
We conclude by noting related issues: (i) Our model admits a simple generalization to Ni impurities. The Ni<sup>++</sup> ion has $`S=1`$, and so we add an additional $`S=1`$ degree of freedom $`\stackrel{}{S}_{\mathrm{Ni}}`$. A simple choice for a Hamiltonian is $`H+K^{}\stackrel{}{S}_{\mathrm{Ni}}\stackrel{}{S}`$. Depending upon the values of $`K^{}`$ and $`K(𝐫)`$, a rich variety of behaviors appear possible. A likely possibility, suggested by the NMR experiments , is that $`\stackrel{}{S}`$ and $`\stackrel{}{S}_{\mathrm{Ni}}`$ combine to form a $`S=1/2`$ moment; the Kondo coupling of this effective moment is expected to be ferromagnetic —so a $`S=1/2`$ moment remains unscreened. The STM spectra will interpolate between the results described here, and those in the presence of a static local magnetic field , and this appears to be the case. (ii) In a recent study of the broadening of the collective spin resonance mode by Zn impurities, the Kondo screening of $`\stackrel{}{S}`$ by the fermionic quasiparticles was neglected. This is valid as long as the energy of the resonance mode, $`\mathrm{\Delta }_{\mathrm{res}}40`$ meV, is larger than $`T_K`$; this condition is obeyed by the $`T_K`$ values discussed above.
We thank H. Alloul, W. Atkinson, A. Balatsky, S. Davis, G. Khaliullin, A. MacDonald, and S. Pan for valuable discussions, and US NSF Grant No DMR 96–23181 and the DFG (VO 794/1-1) for support.
|
warning/0007/hep-ph0007289.html
|
ar5iv
|
text
|
# BNL-HET-00/29, hep-ph/0007289 — July 2000 Virtual corrections to 𝑔𝑔→𝐻 to two loops in the heavy top limit
## 1 Introduction
The Standard Model of Elementary Particle Physics (SM) has been verified in many of its details with enormous precision over the last 20 years. However, several fundamental questions remain unanswered. Perhaps the most important unresolved problem concerns the origin of the particle masses.
In the SM, the underlying mechanism for mass generation is spontaneous breakdown of the electro-weak gauge symmetry. In its minimal version (which we will assume throughout this paper) it predicts a single, as yet undetected physical particle, the Higgs boson. It is determined to be electrically neutral and of spin zero. Its mass, however, is a free parameter of the theory.
The extensive searches for the Higgs boson at particle colliders have set a lower bound on its mass, the latest update yielding a limit of around 108 GeV . On the other hand, theoretical predictions for physical observables depend on the Higgs mass through radiative corrections. Comparison of the existing experimental precision data with their theoretically predicted values allows to derive a most probable range for the Higgs mass which turns out to be roughly between 100 and 200 GeV .
With LEP being close to its maximum possible energy, the attention concerning Higgs search is turning towards future experiments at hadron colliders, in particular LHC, scheduled for the year 2007, or already Tevatron’s Run II, starting in 2001.
The dominant production mechanism for a Higgs boson with a mass below 1 TeV at the LHC will be through gluon-gluon fusion (for a review see ). The coupling of the gluons to the Higgs boson is mediated through a quark loop, Fig. 1 (a). In the heavy quark limit, the corresponding form factor becomes independent of the quark mass. Thus, this process can be used, for example, to count the number of heavy quarks that may exist beyond the third generation.
The current theoretical prediction for this reaction carries an uncertainty of about a factor of 1.5 to 2. It is therefore important to improve on the theoretical accuracy. In this paper we provide a gauge invariant ingredient to the complete next-to-next-to-leading order prediction, namely the virtual corrections up to order $`\alpha _s^4`$. The calculation is, to our knowledge, the first application of a recently introduced method that allows to relate the relevant set of vertex diagrams to the more familiar class of three-loop two-point functions.
## 2 Effective Lagrangian
As it was mentioned before, the coupling of the gluons to the Higgs boson is mediated through a quark loop, Fig. 1 (a). Since all quarks except for the top are much lighter than the current lower limit on the Higgs mass, we will neglect their masses in the following. In this case, the top quark is the only one that couples directly to the Higgs boson, because the Higgs-fermion vertex is proportional to the fermion mass.
The leading order result has been known for quite a while . At the parton level it reads:
$$\begin{array}{cc}\hfill \sigma _{\text{LO}}(ggH)& =\frac{G_\text{F}\alpha _s^2(\mu ^2)}{128\sqrt{2}\pi }\tau ^2\delta (1z)\left|1+(1\tau )f(\tau )\right|^2,\hfill \\ \hfill f(\tau )& =\{\begin{array}{cc}\mathrm{arcsin}^2\frac{1}{\sqrt{\tau }},\hfill & \tau 1,\hfill \\ \frac{1}{4}\left[\mathrm{log}\frac{1+\sqrt{1\tau }}{1\sqrt{1\tau }}i\pi \right]^2,\hfill & \tau <1,\hfill \end{array}\hfill \\ \hfill \tau & =4M_\mathrm{t}^2/M_\mathrm{H}^2,z=M_\mathrm{H}^2/s,\hfill \end{array}$$
(1)
where $`s`$ is the partonic cms energy and $`G_\text{F}`$ is the Fermi coupling constant. $`\alpha _s`$ is the strong coupling constant which depends on the renormalization scale $`\mu `$. $`M_\mathrm{t}`$ is the pole mass of the top quark, and $`M_\mathrm{H}`$ is the Higgs mass. In order to arrive at the cross section for hadron collisions, $`\sigma _{\text{LO}}`$ has to be folded with the gluon distribution functions.
The currently favored values for the Higgs mass appear to be not too different from the top quark mass. However, since the threshold for open top production is at $`M_\mathrm{H}=2M_\mathrm{t}`$, an expansion in terms of small Higgs mass is expected to work well for $`M_\mathrm{H}<2M_\mathrm{t}`$. In fact, for the most interesting mass range of 100 GeV $`M_\mathrm{H}`$ 200 GeV it appears that at next-to-leading order in $`\alpha _s`$ the complete result is excellently approximated by the leading term in an expansion in $`M_\mathrm{H}^2/M_\mathrm{t}^2`$ . Therefore we think it is reasonable to adopt this limit also at next-to-next-to-leading order (NNLO).
All calculations in this paper have been performed using dimensional regularization in $`D=42ϵ`$ space time dimensions. Unless stated otherwise, renormalized expressions are to be understood in the $`\overline{\text{MS}}`$ scheme, bare ones will be marked by the superscript “B”. Furthermore, all the quantities used in the following refer to the five flavor effective theory. For example, by $`\alpha _s`$ we mean the running coupling constant with five active flavors, $`\alpha _s^{(5)}(\mu ^2)`$.
The most convenient way to obtain the leading term in $`M_\mathrm{H}^2/M_\mathrm{t}^2`$ is to use the following effective Lagrangian for the Higgs-gluon interaction, where the top quark has been integrated out:
$$_{\mathrm{eff}}=\frac{H}{v}C_1^\text{B}\frac{1}{4}(G_{\mu \nu }^\text{B})^2=\frac{H}{v}C_1\frac{1}{4}(G_{\mu \nu })^2.$$
(2)
Here, $`G_{\mu \nu }`$ is the gluonic field strength tensor in the effective five flavor theory, and $`v`$ is the vacuum expectation value of the Higgs field, related to the Fermi constant by $`v=(\sqrt{2}G_\text{F})^{1/2}`$. The coefficient function $`C_1`$ has been computed in up to $`𝒪(\alpha _s^4)`$. In order to obtain the cross section for $`ggH`$ with NNLO accuracy, however, it will only be needed up to $`𝒪(\alpha _s^3)`$ :
$$\begin{array}{cc}\hfill C_1=& \frac{1}{3}\frac{\alpha _s}{\pi }\left\{1+\frac{11}{4}\frac{\alpha _s}{\pi }+\left(\frac{\alpha _s}{\pi }\right)^2\left[\frac{2777}{288}+\frac{19}{16}l_t+n_l\left(\frac{67}{96}+\frac{1}{3}l_t\right)\right]\right\},\hfill \end{array}$$
(3)
where $`l_t=\mathrm{ln}(\mu ^2/M_t^2)`$, with $`M_t`$ the on-shell top quark mass. Here and in the following, the number of (light) flavors $`n_l`$ will eventually be set to five.
The renormalized operator in Eq. (2) is related to the bare one through
$$(G_{\mu \nu })^2=\frac{1}{1\beta (𝒩\alpha _s)/ϵ}(G_{\mu \nu }^\text{B})^2,$$
(4)
where
$$𝒩=\mathrm{exp}\left[ϵ\left(\gamma _\text{E}+\mathrm{ln}4\pi \right)\right],$$
(5)
and $`\beta (\alpha _s)`$ governs the running of the strong coupling constant:
$$\mu ^2\frac{\mathrm{d}}{\mathrm{d}\mu ^2}\alpha _s=\alpha _s\beta (\alpha _s).$$
(6)
Its perturbative expansion is known up to $`𝒪(\alpha _s^4)`$ , but for the present purpose the terms up to $`𝒪(\alpha _s^2)`$ are sufficient:
$$\begin{array}{cc}\hfill \beta (\alpha _s)=& \frac{\alpha _s}{\pi }\left[\frac{1}{4}\left(11\frac{2}{3}n_l\right)+\frac{\alpha _s}{\pi }\frac{1}{16}\left(102\frac{38}{3}n_l\right)\right].\hfill \end{array}$$
(7)
Using the Lagrangian of Eq. (2) for the Higgs-gluon interaction instead of the full SM, the number of loops reduces by one. For example, in leading order one obtains the tree diagram shown in Fig. 1 (b). The corresponding expression for the partonic cross section is
$$\sigma _{\text{LO}}(ggH)=\frac{G_\text{F}M_\mathrm{H}^2\alpha _s^2}{288\sqrt{2}\pi }\delta (1z)$$
(8)
and coincides with the limit $`\tau \mathrm{}`$ of Eq. (1), of course.
Going to higher order, one has to compute virtual corrections to diagram Fig. 1 (b). However, these will contain infra-red and collinear divergences. The infra-red divergences will be canceled when adding the corrections from real radiation of quarks and gluons, but the sum will still have collinear divergences. The latter will disappear only when the Altarelli-Parisi splitting functions are taken into account up to the appropriate order . At the next-to-leading order, a full calculation has been carried out in . The correction terms increase the cross section by about a factor of 1.5 to 2 in the relevant Higgs mass range.
At NNLO, only a few ingredients for the full answer are available: In the one-loop amplitude for the radiation of a single quark or gluon in gluon-gluon, gluon-quark, and quark-quark fusion was obtained, and contains the tree-level amplitude for the double-emission of gluons and quarks in these reactions. Both of these results still have to be integrated over the corresponding two- and three-particle phase space, which certainly is a rather non-trivial task by itself.
In this paper, we want to add the virtual two-loop corrections to the list of available knowledge at NNLO. Together with the phase-space integrated expressions for the real radiation and the Altarelli-Parisi splitting functions, one will then be able to arrive at a finite result for the partonic cross section. In order to obtain a physically accessible quantity, one needs to fold this partonic result with the corresponding parton distribution functions. However, they have yet to be evaluated to the appropriate order in $`\alpha _s`$ (see, e.g. ).
## 3 Calculation of the two-loop diagrams
Examples of diagrams contributing to the virtual two-loop corrections to the process $`ggH`$ are shown in Fig. 2. The Higgs boson couples to the effective vertex resulting from the Lagrangian of Eq. (2). The two gluons are on-shell ($`p_1^2=p_2^2=0`$), which is why — after extraction of the tensor structure — the diagrams depend only on one kinematic variable, $`(p_1+p_2)^2=q^2=M_\mathrm{H}^2`$.
In addition to their topology and the power of the denominators, the resulting integrals can be classified by the power of irreducible numerators, i.e. invariants of momenta that cannot be expressed in terms of denominators.<sup>1</sup><sup>1</sup>1There is a freedom in choosing the specific invariants, of course. In , the integrals with unit (or zero) power of denominators and low powers of irreducible numerators were evaluated using Feynman parameterization and dispersion techniques.
In recurrence relations based on the integration-by-parts algorithm for the planar two-loop integrals were derived, reducing them to convolutions of one-loop integrals. These relations allow to compute any such planar diagram with arbitrary powers of the denominators and irreducible numerators.
In our case, however, we also need to compute non-planar diagrams, e.g. Fig. 2 (b). For this reason, we follow an algorithm that has recently been published by Baikov and Smirnov . It relates the recurrence relations for $`l`$-loop integrals with $`n+1`$ external legs to the ones for ($`l+1`$)-loop integrals with $`n`$ external legs. Here we have $`n=l=2`$, and thus the massless two-loop vertex diagrams of Fig. 2 are mapped onto massless three-loop two-point functions. The algorithm to compute the latter ones is known and implemented in the computer program MINCER , written in FORM . Following the recipe of , we modified the MINCER routines such that they are applicable to the class of two-loop three-point functions at hand. For the generation of the diagrams we used QGRAF as integrated in the program package GEFICOM <sup>2</sup><sup>2</sup>2I acknowledge the kind permission by the authors of GEFICOM to use this program..
The only integral that can not be reduced to convolutions of one-loop integrals in this approach is the non-planar one with all propagators appearing in single power, and with numerator equal to one. However, the result for this integral is known as an expansion in $`ϵ`$ up to its finite part .
As a check of our setup we re-did the calculation of the electro-magnetic quark form factor in QCD to two loops and found full agreement with . We performed this calculation in a general $`R_\xi `$ gauge and explicitly checked its gauge parameter independence in this way. We also computed the two-loop three-gluon vertex in $`R_\xi `$ gauge with two gluons on-shell and found agreement with . Finally, the calculation of the present paper was also performed in $`R_\xi `$ gauge and we verified that the gauge parameter dependence disappears in the sum of all diagrams.
## 4 Results
The virtual cross section for the process $`ggH`$ can be written as
$$\sigma _{\mathrm{virt}}=4\pi \frac{M_\mathrm{H}^2}{v^2}\delta \left(1z\right)\left(\frac{C_1}{1\beta (𝒩\alpha _s)/ϵ}\right)^2\frac{1}{256(1ϵ)^2}\underset{\mathrm{pol}}{}\left|\right|^2,$$
(9)
where
$$=\epsilon ^{a,\mu }(p_1,\lambda _1)\epsilon ^{b,\nu }(p_2,\lambda _2)𝒜_{\mu \nu }^{ab}(p_1,p_2),$$
(10)
and $`C_1`$ is the coefficient function given in Eq. (3). The factor $`1/256/(1ϵ)^2`$ comes from the average over initial state polarizations and color. The kinematical constraint on the momenta is
$$p_1+p_2=q,q^2=M_\mathrm{H}^2,p_1^2=p_2^2=0.$$
(11)
$`\epsilon ^{a,\mu }(p,\lambda )`$ is the polarization vector of a gluon with momentum $`p`$ and polarization $`\lambda `$, and $`a,b`$, respectively $`\mu ,\nu `$ are the SU(3)-color and the Lorentz indices. Adopting a covariant $`R_\xi `$ gauge, the polarization vectors obey the relation
$$\underset{\lambda }{}\epsilon _\mu ^a(p,\lambda )\epsilon _\nu ^b(p,\lambda )=g_{\mu \nu }\delta ^{ab}.$$
(12)
The tensor $`𝒜_{\mu \nu }^{ab}`$ may be written as
$$𝒜_{\mu \nu }^{ab}=\frac{\delta ^{ab}}{2M_\mathrm{H}^2}\left[a(p_1p_2)g_{\mu \nu }+bp_{1\nu }p_{2\mu }+cp_{1\mu }p_{2\nu }+dp_{1\mu }p_{1\nu }+ep_{2\mu }p_{2\nu }\right].$$
(13)
The last two terms seem to violate gauge invariance. However, to lowest order they vanish, and starting from next-to-leading order, their contribution to the squared matrix element gets canceled by the diagrams with ghosts in the initial state. We explicitly checked this cancelation by computing these ghost diagrams to the corresponding order. Further, we have
$$a=b,$$
(14)
and thus the $`c`$ term in (13) does not contribute to the total rate. Therefore we will only quote the result for $`a`$ in the following. Its loop expansion can be written as
$$a=1+\frac{\alpha _s^\text{B}}{\pi }a^{(1)}+\left(\frac{\alpha _s^\text{B}}{\pi }\right)^2a^{(2)}+\mathrm{}.$$
(15)
The first order correction has been computed in up to the finite part. Since it involves a second order pole in $`ϵ`$ and we are also interested in the finite part of its square, we need its expansion up to $`ϵ^2`$:
$$\begin{array}{c}\hfill a^{(1)}=𝒩\left(\frac{\mu ^2}{M_\mathrm{H}^2}\right)^ϵ\left\{\frac{3}{2ϵ^2}+\frac{3}{4}\zeta _2+ϵ\left(\frac{3}{2}+\frac{7}{2}\zeta _3\right)+ϵ^2\left(\frac{9}{2}+\frac{141}{32}\zeta _4\right)\right\}+𝒪(ϵ^3),\end{array}$$
(16)
where $`\mu `$ is the ’t Hooft mass, and $`\zeta _n\zeta (n)`$ is Riemann’s zeta function ($`\zeta _2=\pi ^2/6`$; $`\zeta _31.20206`$; $`\zeta _4=\pi ^4/90`$). Furthermore, it is understood that $`(1)^ϵ\mathrm{exp}(+iϵ\pi )`$.
To this we add the second order correction:
$$\begin{array}{cc}\hfill a^{(2)}& =𝒩^2(\frac{\mu ^2}{M_\mathrm{H}^2})^{2ϵ}\{\frac{9}{8ϵ^4}+\frac{1}{ϵ^3}[\frac{33}{32}+\frac{1}{16}n_l]+\frac{1}{ϵ^2}[\frac{67}{32}\frac{9}{16}\zeta _2+\frac{5}{48}n_l]\hfill \\ & +\frac{1}{ϵ}\left[\frac{17}{12}+\frac{99}{32}\zeta _2\frac{75}{16}\zeta _3+n_l\left(\frac{19}{72}\frac{3}{16}\zeta _2\right)\right]\hfill \\ & +\frac{5861}{288}+\frac{201}{32}\zeta _2+\frac{11}{16}\zeta _3\frac{189}{32}\zeta _4+n_l[\frac{605}{216}\frac{5}{16}\zeta _2\frac{7}{8}\zeta _3]\},\hfill \end{array}$$
(17)
where, as before, $`n_l`$ is the number of light quark flavors, $`n_l=5`$.
Ultra-violet renormalization of the strong coupling constant is given by $`\alpha _s^\text{B}=Z_\alpha (𝒩\alpha _s)\alpha _s`$, where $`Z_\alpha `$ is related to the $`\beta `$ function of Eq. (7) through
$$\alpha _s\frac{}{\alpha _s}\mathrm{ln}Z_\alpha (\alpha _s)=\frac{\beta (\alpha _s)}{ϵ\beta (\alpha _s)}.$$
(18)
## 5 Ratio of time-like to space-like form factor
Both as a check and as an estimate on the magnitude of the corrections, we may consider the ratio of the time-like to the space-like form factor. It is free from infra-red singularities and contains the presumably most significant contributions stemming from the analytic continuation of the factor $`(\mu ^2/(q^2))^ϵ`$ from space-like to time-like values of $`q^2`$.
As was shown in for the quark form factor in QCD , this ratio can be derived up to $`𝒪(\alpha _s^2)`$ from the one-loop terms of the form factor by combining them with a known two-loop anomalous dimension. In the case of $`ggH`$ this anomalous dimension is given by $`9/4`$ times the one given in . Following the derivation of and setting $`\mu ^2=M_\mathrm{H}^2`$, we arrive at the following expression:
$$\begin{array}{cc}\hfill \left|\frac{a(M_\mathrm{H}^2)}{a(M_\mathrm{H}^2)}\right|^2& =1+\frac{3}{2}\pi ^2\frac{\alpha _s(M_\mathrm{H}^2)}{\pi }+\left(\frac{\alpha _s(M_\mathrm{H}^2)}{\pi }\right)^2\left(\frac{3}{4}\pi ^4+\frac{67}{8}\pi ^2\frac{5}{12}\pi ^2n_l\right)+𝒪(\alpha _s^3)\hfill \\ & 1+14.8\frac{\alpha _s(M_\mathrm{H}^2)}{\pi }+153.2\left(\frac{\alpha _s(M_\mathrm{H}^2)}{\pi }\right)^2+𝒪(\alpha _s^3)\hfill \\ & 1+0.528+0.1721.700,\hfill \end{array}$$
(19)
where we inserted $`n_l=5`$ in the second line. The third line displays separately the LO, NLO and the NNLO contribution, as well as their sum, for $`\alpha _s(M_\mathrm{H}^2)=0.112`$. Eq. (19) fully agrees with a direct evaluation of the ratio using the expressions (15), (16), and (17) for $`a`$. This provides a check on the terms $`\alpha _s^n/ϵ^{2nk}`$ ($`k=0,1,2`$) of our result. The numerical value of the NLO correction in Eq. (19) reflects the largeness of the full NLO terms as obtained in . The number for the NNLO corrections gives some hope towards a certain degree of convergence for the perturbative series of the full result for $`\sigma (ggH)`$. Concluding this section, let us note that an interesting extension of this discussion could be the resummation of the leading terms along the lines of .
## 6 Conclusions and Outlook
We used the recently introduced method of in order to calculate the NNLO virtual corrections to the production cross section of Higgs bosons in gluon fusion. The result is a gauge invariant component of the full cross section. The next step towards a complete answer for the NNLO rate is to integrate the squared amplitudes for the real radiation processes over the phase space. This is work in progress . Finally, one has to convolute the full partonic cross section with the parton distribution functions. Their evaluation to the relevant order is therefore certainly a very important task.
## Acknowledgments
I would like to thank the High Energy/Nuclear Theory group at BNL, in particular A. Czarnecki, S. Dawson, W. Kilgore, and W. Vogelsang, for encouragement and valuable discussions. Furthermore, I acknowledge discussions and comments by K. Melnikov and T. Seidensticker. This work was supported by the Deutsche Forschungsgemeinschaft.
|
warning/0007/hep-ph0007139.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The Standard Model (SM) of high energy physics, despite its eminent success, suffers from certain drawbacks, the hierarchy problem being a major one. Supersymmetry (SUSY) provides an elegant solution and, consequently, has been a cornerstone in attempts to build models going beyond the SM. It is manifestly clear, though, that SUSY must be broken, at least at low energies. This forces us onto a new problem. Since SUSY cannot be broken in the observable sector in a phenomenologically consistent way , one is forced to introduce a hidden sector wherein the breaking takes place. The question as to how this breaking is conveyed to the observable sector is yet to be settled. The idea that gravity plays the primary mediating role has, historically, been the most popular one. In such supergravity (SUGRA)-inspired models, SUSY breaking occurs at a very high scale (typically well above the grand unification scale) and is communicated to the visible sector through gravitational interactions, the only one common to both the sectors. The gravitino turns out be heavy (at or above the electroweak scale) and, generically, the lightest of the neutralinos is the lightest supersymmetric particle (LSP). Such scenarios suffer from a potential drawback though: interactions of heavy fields above the Grand Unification scale can induce large flavour-changing neutral currents at low energies . Questions like these as well as the fact that we still do not have a complete theory of SUSY breaking through gravitation have, in recent years, prompted research into alternative mechanisms for SUSY breaking.
One such mechanism postulates a set of particles (the “messenger sector” MS) that both transform non-trivially under the SM gauge group, as well as interact with the hidden sector. The latter interaction, which communicates SUSY breaking from the hidden sector to the MS superfield(s), could have a characteristic scale as low as $`𝒪(10^{23})\mathrm{TeV}`$ . The SM gauge interactions can then serve to communicate the breaking to the observable sector. This assures, for example, that the MSSM sfermions with the same quantum numbers are degenerate at the scale. Furthermore, given the limited range of renormalization group running, they continue to be approximately degenerate at the electroweak scale thereby avoiding the flavour problem. Even more interestingly, the gravitino in these gauge mediated SUSY breaking (GMSB) models turns out to be superlight, in contrast to the case of the supergravity models. Consequently, the lightest of the usual superpartners (now the next to lightest supersymmetric particle or NLSP) can now decay into its SM counterpart and the gravitino.
We see thus, that, apart from its purely theoretical aspects, the dynamics of SUSY breaking is likely to leave its imprint on low energy phenomenology as well . With the spectrum changing significantly, search strategies need to be modified. Furthermore, there could be cases where, even after SUSY signals have been established, an understanding of the mode of SUSY breaking remains elusive . Such “inadequacies” of the simplest strategies thus call for new ones to be developed. We shall attempt to do this in the context of $`e^{}e^{}`$ coliders.
We structure the rest of this article as follows. In Section 2, we present a very brief review of the GMSB models. The following section deals with selectron pair production at $`e^{}e^{}`$ colliders. In sections 4 and 5, we examine the signal and background for cases with selectron NLSP and neutralino NLSP respectively. Section 6 examines the possibility of identifying between GMSB and SUGRA-inspired models. Finally, we conclude in Section 7.
## 2 The spectrum in GMSB models
Renormalizability of the theory, coupled with economy of field content, dictates that the messenger sector be comprised of chiral superfields such that their SM gauge couplings are vectorial in nature. Most GMSB models actually consider these fields to be in ($`5+\overline{5}`$) or ($`10+\overline{10}`$) representations of $`SU(5)`$. This construct, while not mandatory, helps preserve the successful SUSY-GUT prediction of the weak mixing angle. The maximum number of messenger families is constrained by the twin requirements of low energy supersymmetry breaking and perturbativity upto the grand unification scale to five of ($`5+\overline{5}`$)s or to one ($`10+\overline{10}`$) in addition to two pairs of ($`5+\overline{5}`$).
Restricting ourselves, for the time being, to a single pair of MS supermultiplets ($`\mathrm{\Psi }+\overline{\mathrm{\Psi }}`$), consider a term in the superpotential of the form $`\lambda 𝒮\overline{\mathrm{\Psi }}\mathrm{\Psi }`$, where $`𝒮`$ is an SM singlet. The scalar ($`S`$) and auxilliary ($`F_S`$) components of $`𝒮`$ may acquire vacuum expectation values (vevs) through their interactions with the hidden sector fields. SUSY breaking is thus communicated to the MS, with the fermions and sfermions acquiring different masses. This, in turn, is communicated to the SM fields resulting in the gauginos and sfermions acquiring masses at the one-loop and two-loop levels respectively. The expressions, in the general case of multiple messenger pairs and/or gauge singlets $`𝒮_i`$, is a somewhat complicated function of $`MS`$ and $`\mathrm{\Lambda }F_S/S`$. However if there be just one such singlet, the expressions for masses at the messenger scale $`M`$ simplify to
$$\begin{array}{ccc}\hfill \stackrel{~}{M}_i(M)& =& N_m\frac{\alpha _i(M)}{4\pi }\mathrm{\Lambda }f_1\left(\frac{\mathrm{\Lambda }}{M}\right)\hfill \\ \hfill \stackrel{~}{m}_{\stackrel{~}{f}}^2(M)& =& 2N_m\mathrm{\Lambda }^2f_2\left(\frac{\mathrm{\Lambda }}{M}\right)\underset{i=1}{\overset{3}{}}\kappa _iC_i^{\stackrel{~}{f}}\left(\frac{\alpha _i(M)}{4\pi }\right)^2.\hfill \end{array}$$
(1)
where $`N_m`$ is the number of messenger generations. In eq.(1), $`C_i^{\stackrel{~}{f}}`$ are the quadratic Casimirs for the sfermion in question. The factors $`\kappa _i`$ equal 1, 1 and $`5/3`$ for $`SU(3)`$, $`SU(2)`$ and $`U(1)`$ respectively with the gauge couplings so normalized that $`\kappa _i\alpha _i`$ are equal at the messenger scale. The threshold functions are given by
$`f_1(x)`$ $`=`$ $`{\displaystyle \frac{1+x}{x^2}}\mathrm{log}(1+x)+\left(xx\right)`$ (2)
$`f_2(x)`$ $`=`$ $`{\displaystyle \frac{(1+x)}{x^2}}\left[\mathrm{log}(1+x)+2Li_2\left({\displaystyle \frac{x}{1+x}}\right){\displaystyle \frac{1}{2}}Li_2\left({\displaystyle \frac{2x}{1+x}}\right)\right]+\left(xx\right).`$ (3)
The superparticle masses at the electroweak scale are obtained from those in eq.(1) by evolving the appropriate renormalization group equations. For the scalar masses, the $`D`$-terms need to be added too.
## 3 Selectron production in polarized $`e^{}e^{}`$ colliders
At an $`e^{}e^{}`$ collider, the dominant production mode for supersymmetric particles is that of a pair of selectrons . The relevant term in the Lagrangian reads
$$=e_{\mathrm{em}}\overline{\chi }_a^0(l_{ia}P_L+r_{ia}P_R)e\stackrel{~}{ϵ}_i,$$
where $`\chi _a^0`$ ($`a=1\mathrm{}4`$) represent the neutralino fields and $`e_i`$ refer to $`\stackrel{~}{e}_{L,R}`$ as the case may be. Consider the polarized electron scattering
$$e^{}(\lambda _1)+e^{}(\lambda _2)\stackrel{~}{e}_1^{}(m_1)+\stackrel{~}{e}_2^{}(m_2)$$
(4)
which proceeds through the $`t`$\- and $`u`$-channel exchanges of each of the four neutralinos. The corresponding differential cross sections are given by
$$\begin{array}{ccc}\hfill \frac{\mathrm{d}\sigma }{\mathrm{d}t}& =& \frac{\pi \alpha ^2}{4s}\underset{a,b}{}[sM_aM_bD_{ab}(\frac{1}{tM_a^2}+\frac{1}{uM_b^2})^22I_{ab}\frac{(utm_1^2m_2^2)}{(tM_a^2)(uM_b^2)}\hfill \\ & & +(utm_1^2m_2^2)C_{ab}(\frac{1}{(tM_a^2)(tM_b^2)}+\frac{1}{(uM_a^2)(uM_b^2)})]\hfill \\ \hfill D_{ab}& & l_{1a}l_{1b}l_{2a}l_{2b}(1\lambda _1)(1\lambda _2)+r_{1a}r_{1b}r_{2a}r_{2b}(1+\lambda _1)(1+\lambda _2)\hfill \\ \hfill C_{ab}& & l_{1a}l_{1b}r_{2a}r_{2b}(1+\lambda _1)(1\lambda _2)+r_{1a}r_{1b}l_{2a}l_{2b}(1\lambda _1)(1+\lambda _2)\hfill \\ \hfill I_{ab}& & l_{1a}r_{1b}r_{2a}l_{2b}(1+\lambda _1)(1\lambda _2)+r_{1a}l_{1b}l_{2a}r_{2b}(1\lambda _1)(1+\lambda _2),\hfill \end{array}$$
(5)
where $`M_a`$ are the masses of the neutralinos. The masses, as well as the couplings $`l_{ia}`$ and $`r_{ia}`$, are, of course, determined by $`N_m`$, $`\mathrm{\Lambda }`$, $`M`$ as well as $`\mathrm{tan}\beta `$ and $`\mu `$. Since the coupling of a fermion-sfermion pair to a higgsino is proportional to the fermion mass, clearly the higgsino components of the neutralinos play only a small part in selectron production. In other words, the dependence of the cross section on $`\mu `$ (and $`\mathrm{tan}\beta `$) is of a minor nature (see Fig. 1). As far as decays of the selectron are concerned, these parameters do play a more significant role though. A small $`\mu `$, for example, results in the some of the neutralinos being light, thus affording a decay channel which might not be otherwise available to the selectron. Inspite of the small coupling of the electron-selectron pair to the higgsinos, such decays might be competitive with the gravitino mode. However, since we would be explicitly considering such cascading decays of the selectron, we are justified in neglecting the dependence on $`\mu `$ and $`\mathrm{tan}\beta `$.
For a given ratio $`\mathrm{\Lambda }/M`$, both the scalar and the gaugino masses grow with $`\mathrm{\Lambda }`$ (eq. 1). Consequently, the production cross section falls steeply with $`\mathrm{\Lambda }`$ (see Fig. 2$`a`$). The fall is understandably steeper for larger $`N_m`$ as the selectron mass grows as $`\mathrm{\Lambda }\sqrt{N_m}`$. The behaviour for small $`\mathrm{\Lambda }`$ is more subtle. The total cross section is a complicated function of $`M_a/\sqrt{s}`$ and $`m_{\stackrel{~}{e}}/\sqrt{s}`$. Combined with the fact that the couplings with the $`\stackrel{~}{B}`$ and $`\stackrel{~}{W}_3`$ are different, this can lead to a situation where the cross sections do not actually fall with $`N_m`$ (Fig. 2$`a`$). On the other hand, as the mass parameter $`M`$ enters eq.(1) only logarithmically, one may expect that the cross-sections would change only fractionally as this parameter is varied (for a fixed $`\mathrm{\Lambda }`$). This is borne out by Fig. 2$`b`$. Since such deviations are almost of the order of statistical fluctuations in the signal itself, for the rest of our study, we will not consider any explicit dependence on the mass ratio $`M/\mathrm{\Lambda }`$.
As we have already pointed out, in GMSB models, the $`\stackrel{~}{e}_L`$ is distinctly heavier than the $`\stackrel{~}{e}_R`$. Hence, we shall concentrate on the pair-production process $`e^{}e^{}\stackrel{~}{e}_R\stackrel{~}{e}_R`$. Once produced, the selectron may decay to either $`e^{}+\stackrel{~}{G}`$ or $`e^{}+\chi _1^0`$ (if kinematically allowed). In the first case the final state comprises of two electrons and missing momentum, while the second case has two photons in addition. As the backgrounds are quite different, we shall now examine each case individually.
## 4 Selectron as the NLSP
With the selectrons decaying into an electron and a gravitino each, the SM background comprises of $`e^{}e^{}\nu _i\overline{\nu }_i`$. The main contributions to the latter clearly arise from the “resonant” processes $`e^{}e^{}e^{}\nu _eW^{}`$ and $`e^{}e^{}e^{}e^{}Z`$ and have been discussed at some length in refs. . The $`W`$ contribution is dominant but can be suppressed by right-polarizing the electron beams. This also serves to enhance the selectron production rate. Of course, the ideal state of a fully polarized beam is virtually unattainable and hereon we shall assume the electron beams to be 90% right-polarized.
It is obvious that the phasespace distribution of the signal events would depend crucially on the mass of the selectron and, to a lesser extent, on the neutralino masses. For purposes of comparison between the signal and the background, we will concentrate on two specific choices in the parameter space marked in Table 1.
Whereas the electrons from an $`\stackrel{~}{e}_R`$ would be produced isotropically, the background events would prefer to have at least one of them to be close to the beam pipe . We thus demand that both the electrons must satisfy
$$|\eta _e|<3,$$
(6)
a requirement in consistency with the angular coverage of proposed detectors. In addition, the leptons must have sufficient momentum to be detectable, viz
$$p_T(e)>5\mathrm{GeV},$$
(7)
and be separated enough to be individually resolved:
$$\mathrm{\Delta }R\sqrt{(\mathrm{\Delta }\eta )^2+(\mathrm{\Delta }\varphi )^2}>0.2,$$
(8)
where $`\mathrm{\Delta }\varphi `$ refers to their azimuthal separation. In addition, we demand that the missing momentum be large enough:
$$p_T/>20\mathrm{GeV}.$$
(9)
With these cuts in place, the total SM background is very flat over the region $`\sqrt{s}=500\mathrm{GeV}`$$`1\mathrm{TeV}`$ and amounts to approximately $`19.5\mathrm{fb}`$ over the entire range. In Table 1, we display the signal cross-section for some representative values of GMSB parameters. One can get $`10^4`$ events per year assuming integrated luminosity of $`50`$ $`\mathrm{fb}^1`$ for $`\sqrt{s}=500`$ GeV and $`10^3`$ events for $`\sqrt{s}=1`$ TeV machine (for the same luminosity). In Fig. 3, we present the phasespace distribution for the background events. Also shown in the figure are the corresponding signal profiles for the two particular points in the parameter space. Let us begin by discussing these distributions.
At first sight, it might seem surprising that the transverse momentum distribution (see Fig. 3$`a`$) for the signal events do not show the characteristic Jacobian peaks. This is not surprising though, as, for such behaviour to be exhibited, an electron should only appear as a decay product of a particular particle. However, in the case at hand, the two electrons cannot be distinguished from each other and hence have to be ordered in some fashion, whether energy or magnitude of transverse momentum or rapidity. We choose the first option, namely energy ordering. Any such ordering will tend to destroy features indicative of individual decay product kinematics, and this is particularly true of set ($`A`$). For set ($`B`$), on the other hand, the selectron mass is much closer to $`\sqrt{s}/2`$ and hence they are produced with very little momenta. Consequently, the effect of ordering is relatively smaller and the remnant of the Jacobian peak more pronounced.
The rapidity distribution (Fig. 3$`b`$) for the SM background shows clearly that the softer electron prefers to lie closer to the beam pipe while the harder one is much more central. This is reflective of the singularities in the photon-mediated contributions. This also makes itself felt in the $`p_T/`$ distributions (see Fig. 3$`c`$) where the SM cross sections fall off much more steeply than those for the signal.
Although the SM background is not too big, it might be desirable to reduce it further without sacrificing a large fraction of the signal. This becomes particularly important when the signal size reduces either on account of $`m_{\stackrel{~}{e}_R}`$ being close to the kinematic limit or other (nonstandard) decay modes becoming available to the selectron. A look at Figs. 3 tells clearly that this goal is unlikely to be achieved by imposing harder cuts on either the individual electron $`p_T`$s or on the missing momentum. Removing events with $`\eta _e>2`$ is an option, but even then, the improvement is merely quantitative and not a qualitative one.
In actuality, such a goal is much better realized by examining the double differential cross section rather than making the individual cuts of eqs.(69) any stronger. As the $`W`$-mediated diagrams have been suppressed by right-polarizing the electron beams, the bulk of the background owes its origin to the $`e^{}e^{}Z`$ final state with the $`Z`$ decaying invisibly. It is easy to see that for such a process, the energies of the two electrons satisfy the conditions
$$\begin{array}{ccc}\hfill E_1+E_2& & \frac{sm_Z^2}{2\sqrt{s}}\hfill \\ \hfill (2E_1\sqrt{s})(2E_2\sqrt{s})& & m_Z^2.\hfill \end{array}$$
(10)
In Fig. 4$`a`$, we present a scatter plot of the SM background for an accumulated luminosity of $`50\mathrm{fb}^1`$. Superimposed on it are the two curves of eq.(10). Eliminating the part of the phase space bounded by the two curves reduces the SM background from $`19\mathrm{fb}`$ to approximately $`1fb`$. The preponderance of points just below the straight line can be attributed to the contributions from a slightly off-shell $`Z`$. A very large fraction of these could be eliminated by modifying the straight line curve by replacing $`m_Z`$ with, say, $`m_Z2\mathrm{\Gamma }_Z`$. Points well outside this region, on the other hand, owe their origin to the $`W`$-mediated diagrams and would disappear in the limit of fully right-polarized electron beams.
In contrast to the above, the electrons in the signal events are distributed evenly within the square region defined by
$$\begin{array}{ccc}\hfill E_{\mathrm{min}}& & E(e_i)E_{\mathrm{max}},\hfill \\ \hfill E_{\mathrm{min},\mathrm{max}}& =& \frac{\sqrt{s}}{4}\left(1\frac{m_{\stackrel{~}{G}}^2}{m_{\stackrel{~}{e}}^2}\right)(1\beta ),\hfill \\ \hfill \beta & =& \left(1\frac{4m_{\stackrel{~}{e}}^2}{s}\right)^{1/2}.\hfill \end{array}$$
(11)
This is illustrated in Fig. 4$`b`$. While a significant fraction of the signal cross section could be lost on imposing the cuts corresponding to eq.(10), the signal to noise ratio shows an enormous improvement.
## 5 Neutralino as the NLSP
As we have already pointed out, for $`N_m=1`$, the lightest neutralino is always the NLSP. Even for $`N_m>1`$, this may continue to be the case especially if the higgsino mass parameter $`\mu `$ is small. However, in all such cases the gaugino mass parameter $`M_2`$ is larger than the mass of the $`\stackrel{~}{e}_R`$. Consequently, the selectron decays mainly into an electron and the lightest neutralino with the latter cascading into a photon and the gravitino. The final state thus comprises of a pair each of electrons and photons and missing energy due to the gravitinos. The energy and angular distributions would obviously depend on the masses of the selectron and the lightest neutralino. We will again concentrate on two representative points (see Table. 2) in our analysis of the signal and comparison with the background.
An exact calculation of the (6-body) SM background is an onerous task. It can be easily seen though that the bulk of the background arises from the two resonant processes:
1. $`e^{}e^{}e^{}e^{}\gamma \gamma Ze^{}e^{}\gamma \gamma \nu _i\overline{\nu }_i`$ and
2. $`e^{}e^{}e^{}\overline{\nu }_e\gamma \gamma W^{}e^{}e^{}\gamma \gamma \nu _e\overline{\nu }_e`$ .
Once soft and collinear singularities have been removed by an appropriate set of cuts, one expects these cross sections to be smaller than those in the previous section by a factor $`𝒪(\alpha ^2)`$. Thus the kinematical cuts required over and above those of eqs.(69) are dictated not by the need to minimize background, but by detector acceptances. To be specific, we demand that
$$\begin{array}{ccccccc}\hfill |\eta _\gamma |& <& 3\hfill & & \hfill p_T(\gamma )& >& 10\mathrm{GeV}\hfill \\ \hfill \mathrm{\Delta }R_{\gamma \gamma }& >& 0.2\hfill & & \hfill \mathrm{\Delta }R_{e\gamma }& >& 0.2.\hfill \end{array}$$
(12)
With these set of cuts, the surviving cross section, for $`\sqrt{s}=500\mathrm{GeV}`$, from process $`(i)`$ above is approximately $`0.004\mathrm{fb}`$ while that from the second one is approximately $`0.001\mathrm{fb}`$. Thus, for all practical purposes, we have a background free situation. The surviving size of the signal is quite similar to that in the previous section as the cuts of eq.(12) do not take away much of the signal.
In Fig. 5, we present the signal event distributions for the two particular parameter choices indicated in Table. 2. Comparing Figs. 5$`a`$ and 3$`a`$, one is struck by the similarity between the curves for parameter sets ($`A`$) and ($`C`$) on the one hand and those for ($`B`$) and ($`D`$) on the other. This can be understood by realizing that the shape of electron transverse momenta distribution is determined by the masses of the selectron and the particles it is decaying into. Since cases ($`A`$) and ($`C`$) correspond to very similar $`m_{\stackrel{~}{e}}`$, it is only natural that the $`p_T`$ spectrum would look similar. The larger mass of the neutralino (as compared to that for the gravitino) is reflected in smaller value of the maximal $`p_T`$ allowed to the electrons. Analogous statements apply to points ($`B`$) and ($`D`$) as well.
We turn now to the photon spectra. As the neutralinos are produced in (isotropic) scalar decays and as they themselves decay into a photon and gravitino, there are no nontrivial angular correlations. Consequently, the spectrum is determined by kinematics alone. This being identical to that for squark decay into massless particles, the decay distributions (Fig. 5$`c,d`$) are very similar to those for the other set. A similar story obtains for the missing transverse momentum as well.
## 6 Distinguishing from non-GMSB models
In the last two sections we have seen that, for $`\mathrm{\Lambda }\stackrel{<}{}\mathrm{\hspace{0.25em}200}\mathrm{TeV}`$, the signal from a GMSB model stands well and truly above the SM background. This is particularly true for the neutralino NLSP case where one expects less than one event from SM processes. This brings us to the more important question, namely how to recognize if supersymmetry breaking is driven by gauge mediation. It goes without saying, though, that short of determining the entire spectrum, one can only draw (strong) inferences and this is what we shall aim to do in this section.
Considering the selectron NLSP case first, it is clear that eq.(11) can be used to determine the masses of both the selectron and the supersymmetric particle $`X`$ (gravitino or neutralino) that it is decaying into. Of course, measurement of the edge of phasespace is always beset with inaccuracies. However, at this stage we do not need to know the mass of $`X`$ very accurately. In fact, as long as the experimentally deduced value $`m_X\stackrel{<}{}\mathrm{\hspace{0.25em}30}\mathrm{GeV}`$ or so, the rest of argument follows. LEP data already tells us that such a light neutralino can only be the bino (primarily)<sup>1</sup><sup>1</sup>1Unless there exist neutralinos, and hence gauge symmetries, going beyond the MSSM. We do not consider such exotic models. Now, the $`\stackrel{~}{e}_R`$ does not couple with the $`\stackrel{~}{W}_3`$ and its coupling with the higgsinos is suppressed by the electron mass. Consequently, for a $`\stackrel{~}{e}_R`$ of fixed mass, the production cross section (eq.5) is determined essentially by the bino mass $`M_1`$. Working in the limit $`M_2,\mu M_1`$, the chirality structure of the amplitude ensures that, for small values of $`M_1^2/s`$, the cross section grows as this ratio (see Fig. 6). For large values of the ratio, though, the cross section would fall off. Realistic values for $`M_2`$ and $`\mu `$ would alter our simplistic arguments to a degree, but such effects are too small to be noticeable in the graph that we present.
That we have produced a pair of $`\stackrel{~}{e}_R`$s and not $`\stackrel{~}{e}_L`$s we can deduce from the polarization of the initial state. Its mass, as we have already seen, can be determined from the energy distribution, and if necessary, refined by a threshold scan. At this stage, Fig. 6 can be used to “determine” $`M_1`$ from the experimentally measured cross section and compare it with the direct, if inaccurate, measurement from the endpoint analysis. Clearly, the consistency between the two values would be much higher for the GMSB hypothsis than for the non-GMSB case. Thus, rate counting helps us to to distinguish between the light gravitino and light bino cases.
The neutralino NLSP case presents us with an additional complication. Presumably we could have had $`\stackrel{~}{e}_Re^{}+\chi _2^0`$ in a non-GMSB scenario followed by $`\chi _2^0\chi _1^0+\gamma `$. We can again measure both $`m_{\stackrel{~}{e}_R}`$ and $`m_\chi `$ by determining the phasespace boundaries of the electrons and employing a relation analogous to eq.(11). What about the mass of the gravitino (neutralino) in the second stage of the decay? In fact, a corresponding relation can be derived for the energy of the photons:
$$\begin{array}{ccc}\hfill E_\gamma ^{\mathrm{max},\mathrm{min}}& =& A(1\pm \beta )(1\pm \zeta )\hfill \\ \hfill A& & \frac{\sqrt{s}}{8}\left(1+\frac{m_\chi ^2}{m_{\stackrel{~}{e}}^2}\right)\left(1\frac{m_{\stackrel{~}{G}}^2}{m_\chi ^2}\right)\hfill \\ \hfill \zeta & & \frac{m_\chi ^2m_{\stackrel{~}{G}}^2}{m_\chi ^2+m_{\stackrel{~}{G}}^2},\hfill \end{array}$$
(13)
with $`\beta `$ defined as in eq.(11).
In Figs. 7($`a,b`$), we exhibit the distributions in the scalar sums of electron and photon energies respectively. The first, namely $`E_{ee}^{\mathrm{sum}}E_{e_1}+E_{e_2}`$, leads to a symmetric distribution as in the case of a selectron NLSP. On the other hand, the distribution in $`E_{\gamma \gamma }^{\mathrm{sum}}`$ shows a high energy tail. The tail is purely a kinematic feature and can be derived from a generalization of the Dalitz plot. A better understanding of the same can be obtained from the scatter plots of Figs. 7($`c,d`$).
As in the previous case, the endpoints of the electron spectrum can be used to deduce both $`m_{\stackrel{~}{e}_R}`$ and $`m_\chi `$. Once $`m_{\stackrel{~}{e}_R}`$ is measured, eq.(13) can be used to determine both $`m_\chi `$ and $`m_{\stackrel{~}{G}}`$. This, thus, also serves as a consistency check. At this stage we can again take recourse to Fig. 6 to argue that the existence of such a light bino (as in a non-GMSB model) would have implied a small cross section. Moreover, if such a bino were to exist, the selectron would have a substantial branching into it. Hence, nonobservance of an excess in the ($`e^{}e^{}+`$ missing energy) final state is yet another argument against a spectrum with a heavy gravitino but a light bino.
## 7 Summary
The $`e^{}e^{}`$ option of the Next Generation Linear collider can be a very effective tool in the search for physics beyond the SM. In this paper, we have studied the feasibility of using such a machine to probe Gauge Mediated Supersymmetry Breaking. The process of choice is the pair-production of right-handed selectrons, not in the least because of their being significantly lighter than their left-handed counterparts.
If the selectron be the NLSP, the signal comprises two electrons accompanied by a missing momentum. Right-polarizing (90%) the electron beams helps eliminate the bulk of the SM background apart from increasing the signal strength as well. The already very good signal to noise ratio can be enhanced even further by imposing correlated cuts on the electron energies. The neutralino NLSP case, on the other hand results in a spectacular final state comprising a pair each of electrons and photons accompanied by missing momentum. The SM background is virtually nonexistent. In either case, the rates are high enough for the selectron to be detectible almost upto the kinematic limit.
The energy correlations (electrons for the selectron NLSP case and both electrons and photons in the neutralino NLSP case) are characteristic and can be used to determine the masses of both the produced particle and its decay products. Furthermore, such information gleaned from the differential distributions, used in conjunction with rate counts, can be used to distingguish GMSB from alternate scenarios of supersymmetry breaking (including, but not limited to, the case of supergravity-inspired models without gaugino mass unification).
Acknowledgement
DC acknowledges the Department of Science and Technology, India for the Swarnajayanti Fellowship grant. DKG acknowledges the hospitality of the Theory Divison, CERN, Geneva and the Laboratorie de Physique Particles (LAPP), Annecy, where a part of this work was done.
|
warning/0007/astro-ph0007122.html
|
ar5iv
|
text
|
# Comment on “On physical interpretation of the Poynting-Robertson effect”
## 1 Introduction
Poynting-Robertson (P-R) effect causes small bodies in circum-solar orbit, such as dust, assumed to totally absorb intercepted radiation and re-emit isotropically in the bodies’ rest-frame (“rest-isotropically”), to inspiral with orbits of decreasing eccentricity (Harwit 1985; Burns 1979). The paradoxical, and hence interesting, element of the effect is that outward-directed raditation from a primary (the Sun or a star) can cause an otherwise stably orbiting dust to infall. Although the effect is familiar enough through treatment in standard textbooks, a physical understanding of its origin has been unclear (for a historical introduction, see Robertson 1937; Klačka 1993; Srikanth 1999).
More often than not the drag has been attributed to the front-back asymmetry of the dust re-emission as seen in the heliocentric frame. The valid objection to this is that rest-symmetric emission should not alter the intertial state of the center-of-mass, as is evident in the instantaneous rest-frame of the dust. In response to this, the aberration of sunlight is invoked to explain the drag in the dust frame. But to this we raise the objection that it is inconsistent to invoke two different causes (dust-emission asymmetry / sunlight aberration) in the two different frames (heliocentric frame / instantaneous dust rest-frame).
The fact of a given cause engendering an observed effect is absolute, and not relative to the choice of reference frame. The covariance of the force equations automatically enforces this: the status of a given tensor as vanishing or non-vanishing is itself absolute, even though its components transform covariantly. The belief that the cause of the P-R effect is also frame-dependent is a popular fundamantal misconception surrounding the effect. A subtlety to be aware of is frame-dependent manifestations of a single cause (eg., the magnetic force between two current carrying wires transforms into an electric in the rest frame of the valence electrons; this happens basically because these two fields are part of a single Maxwell tensor field). However, in the P-R effect, absorption and re-emission are in principle distinct processes and transform independently of each other. Strangely, this curious state of impasse doesn’t seem to have attracted sufficient mainstream attention.
In Srikanth (Icarus 1999), the P-R “paradox” was resolved by carefully distinguishing the contributions to the dust dynamics from the absorption and re-emission. We showed that solar radiation absorption plays an explicit role in the drag (i.e, azimuthal slow-down) of the dust, while re-emission, assumed rest-isotropic, does not. The slow-down may be visualized as a consequence of dust absorption and angular momentum conservation.
The present paper is intended to further clarify this result, and in specific to respond to comments made in Klačka (astro-ph/0006426: hereafter Klačka). This is worth the effort, since history affirms that the paradoxical nature of this effect can cause confusion.
## 2 The equations of motion
It is sufficient for our purpose to consider a simple model of a spherical, completely absorbing dust, with the added generalization that the absorption and re-emission parameters of the dust are mutually distinct. The generally covariant equations governing the motion of the dust can be written as:-
$$\frac{Dp^\mu }{D\tau }\frac{dm}{d\tau }u^\mu +m\frac{Du^\mu }{D\tau }=f_{ext}^\mu ,$$
(1)
where $`p^\mu =mu^\mu `$ is four-momentum of the dust, $`\tau `$ proper time, $`f_{ext}^\mu `$ is the external four-force, and the operator $`D/D\tau `$ is the “total” covariant derivative in the General Relativistic sense and includes gravitational effects. Thus gravitation is not part of $`f_{ext}^\mu `$. The mass change term accounts for possible change in the dust’s internal energy due to heating or cooling (Srikanth 1999),
The primary radiation is considered as a plane-parallel beam flowing radially outward, represented by the force $`f_{rad}^\mu =ϵl^\mu `$, where $`l^\mu `$ is a dimensionless null-vector, with its spatial part being purely radial (Robertson 1937). The scalar $`ϵ`$ $`\left(>0\right)`$ is the rest-rate of absorption of solar radiation by the dust, with dimension momentum over time. We assume that the dust absorbs all the incident radiation and re-emits rest-isotropically. The relativistic four-force associated with a rest-isotropic emission is the time-like four-vector $`f_{emit}^\mu =\left(\xi /c^2\right)u^\mu `$, where $`u^\mu `$ is four-velocity, and $`\xi `$ $`\left(>0\right)`$ is the rest-frame energy emission rate. Accordingly, Eq. (1) becomes:-
$`{\displaystyle \frac{dm}{d\tau }}u^\mu +m{\displaystyle \frac{Du^\mu }{D\tau }}`$ $`=`$ $`f_{rad}^\mu +f_{emit}^\mu ,`$ (2)
$`=`$ $`ϵl^\mu \left(\xi /c^2\right)u^\mu .`$
This can be split into scalar and spacelike componants as follows.
Contracting Eq. (2) by $`u_\mu `$, we get the scalar equation:
$$c^2\frac{dm}{d\tau }=ϵl^\alpha u_\alpha \xi .$$
(3)
Substituting this back into Eq. (2), we get the true equations of motion:-
$$m\left(\tau \right)\frac{Du^\mu }{D\tau }=ϵ\left(l^\mu \frac{l^\alpha u_\alpha }{c^2}u^\mu \right),$$
(4)
where the bracketed term is just the part of radiation orthogonal to $`u^\mu `$. The significance of this split is that Eq. (3) describes the internal energy change of the dust but not its motion, while Eq. (4) describes its motion. Both equations are in general coupled because of the time-dependence of mass. Implementing the metric for a weak static gravitational field, and letting $`vc`$, Eq. (4) can be shown to lead to the usual non-relativistic equations of motion for the P-R effect (Srikanth 1999).
## 3 Identifying the P-R drag
The first term on the right-hand side of Eq. (4) is the radiation pressure term. The second term, which is formally the relativistic generalization of a friction force, is responsible for the azimuthal deceleration of the dust. Hence it is called the drag term. Switching it off, as it were, leads to a stably orbiting dust with the gravitational field modified by radiation pressure. To check that this is the case, in Eq. (2), we put $`ϵ=0`$, which leads to:
$`c^2{\displaystyle \frac{dm}{d\tau }}`$ $`=`$ $`\xi .`$ (5)
$`m\left(\tau \right){\displaystyle \frac{Du^\mu }{D\tau }}`$ $`=`$ $`0.`$ (6)
Eq. (6) shows that in the absence of absorption, the dust moves along a geodesic. The mass of the dust is time-dependent, but does not affect its motion. Hence absorption is a necessary condition for dust slow-down, while re-emission is not a sufficient condition for slow-down.
On the other hand, associating the four-momentum carried away from the dust, which is the second term in the r.h.s of Eq. (2), with the drag, as Klačka recommends, is misleading. In Eq. (3), putting $`\xi =0`$ leads to:
$$c^2\frac{dm}{d\tau }=ϵl^\alpha u_\alpha ,$$
(7)
while Eq. (4) remains unchanged. Thus, we still have the slow-down of the dust in the absence of re-emission, though the trajectory is modified by the mass change. This shows that re-emission is not a necessary condition for slow-down, while absorption is a sufficient condition (assuming all the usual non-P-R aspects of problem as given). If $`dm/d\tau 0`$ then re-emission contributes implicitly via the time-dependent mass. However, this is not a friction or drag-like contribution. Depending on whether the dust heats up or cools down, this process renders the dust less or more susceptible to drag.
In summary, re-emission is neither a necessary nor sufficient condition for the drag, while absorption is necessary. The association of re-emission with the drag is erroneous except in the case of an isothermal ($`dm/d\tau =0`$) dust, as seen via Eq. (3). Because historically the P-R effect had been studied with an implicit assumption of isothermality, this equality led to a general confusion of the relative roles played by these two processes in the drag, as discussed in the Introduction. Indeed, the two-parameter model of the P-R effect that we have adopted was motivated by this observation.
## 4 Isothermality
As seen from its rest frame, an isothermal dust emits as much as it absorbs (Srikanth 1999). However, in any other frame, as seen from Eq. (2), the excess of absorbed energy over re-emission is balanced by kinetic energy. Klačka argues that this distinction between the reference frames is obvious, and no more necessary to make explicit than to state that the laws of reflection must be referred to the rest frame of the reflecting surface.
However, in fact, this analogy is not apt. Here, the rest-frame of the reflector is a natural choice for reference frame. The simplest statement of the law of reflection also refers to this frame. But, in the P-R effect, while the convenient choice of frame is the heliocentric frame, the simplest possible satisfaction of isothermality (by the equality of emitted to absorbed radiation) holds good in the dust rest-frame. This has lead in some extant literature to the erroneous association of the simplest version of isothermality with the heliocentric reference frame. The purpose of mentioning this point in Srikanth (1999) was of pedagogical and historical interest.
## 5 Conclusions
Some clarifications on an earlier result reached by us in Srikanth (1999) are given. We find that in the P-R effect dust absorption is a necessary condition for drag, whereas re-emission, assumed to be rest-isotropic, is neither necessary nor sufficient. This result is independent of the reference frame used for the description of the effect.
## Acknowledgements
I am thankful to Dr. S. P. K. Rajaguru and Dr. Gajendra Pandey.
|
warning/0007/cond-mat0007280.html
|
ar5iv
|
text
|
# On-site magnetization in open antiferromagnetic chains: a classical analysis versus NMR experiments in a spin-1 compound
\[
## Abstract
The response of an open spin chain with isotropic antiferromagnetic interactions to a uniform magnetic field is studied by classical Monte Carlo simulations. It is observed how the induced on-site magnetization is non uniform, due to the occurrence of edge staggered terms which decay exponentially over a distance equal to the zero field correlation length of the infinite chain. The total magnetic moment associated to each staggered term is found to be about half of the original single-spin magnitude and to decrease as the inverse of temperature (i.e. to behave as a Curie-like moment). The numerical results are compared to recent NMR findings in spinless-doped Y<sub>2</sub>BaNiO<sub>5</sub>; the remarkable agreement found shows that, for temperatures above the Haldane gap, the classical approach gives a correct picture of the boundary effects observed in the Heisenberg $`S`$=1 chain.
\]
Because of the richness of their phase diagrams, Heisenberg spin systems with low-dimensional antiferromagnetic interactions (HAF’s) are presently attracting strong interest. Magnetic correlations in spatially-homogeneus HAF’s have been probed dynamically, in Fourier-space, by neutron scattering, but can hardly be visualized experimentally in a static real-space picture. Translational invariance, infact, makes each site equivalent to the others, thus preventing any spatial oscillation of the spin direction. In recent experiments, the correlation properties of low-dimensional HAF’s have been investigated through NMR imaging of the spin polarization induced by a uniform field around non-magnetic defects that break the translational invariance. Relevant results have been obtained in copper oxide two-dimensional compounds, in spin ladders, in half-integer and integer spin chains.
In $`S=1`$ chains in particular, the local magnetization $`S_i^z`$ has been resolved site by site, through <sup>89</sup>Y NMR in Mg-doped Y<sub>2</sub>BaNiO<sub>5</sub> , for temperatures ranging from $`T0.35J/k_B`$ to $`TJ/k_B`$, being $`J`$ the AF exchange constant. $`S_i^z`$ shows an alternate component which is maximum around impurities (i.e. close to chain boundaries) and vanishes exponentially over a distance equal to the zero-field correlation length of the bulk. Boundary staggered defects with total spin $`S=1/2`$ and a size of the order of the bulk correlation length are actually expected, for $`S=1`$ HAF chains, in the limit $`T0`$. At very low temperatures in fact, the magnetic properties of these systems are controlled by edge-induced triplet states, in which $`S_i^z`$ shows the profile described above. This $`T0`$ argument however, can hardly be used by itself in order to explain the experimental evidences in Ref. 5, since - on increasing temperature - the excited states above the Haldane gap $`\mathrm{\Delta }_H0.4J`$ should be taken into account (in open $`S=1`$ chains $`\mathrm{\Delta }_H`$ corresponds to the separation between the second and the third lowest lying energy levels). A finite temperature analysis, in which all the excited states are correctly treated, was recently carried out by Alet and Sørensen, using Quantum Monte Carlo techniques. Excellent agreement with the experimental data derived from NMR imaging has been obtained.
The purpose of the present work is to study whether or not the alternating magnetization observed in finite $`S=1`$ chains is a pure quantum mechanical effect somewhat reminiscent of the liquid-like ground state. The response of open chains with various lengths $`L`$ to a uniform magnetic field is thus studied in the limit of infinite-$`S`$, proving that edge staggered defects with spin-$`S/2`$ originate from the translational invariance breaking, even in the framework of a classical model. It is shown that the characteristic spatial extention of these defects corresponds to the zero-field correlation length of the thermodynamic chain ($`L\mathrm{}`$) and that, at high temperature, the behavior of $`S_i^z`$ calculated for infinite-$`S`$ tracks the experimental findings in the $`S=1`$ chain.
In the classical limit ($`S\mathrm{}`$), the Hamiltonian for a nearest-neighbor Heisenberg chain of $`L`$ sites in an external field $`H`$ takes the form
$$=J\underset{i=1}{\overset{L}{}}\stackrel{}{S}_i\stackrel{}{S}_{i+1}g\mu _BH\underset{i=1}{\overset{L}{}}S_i^z,$$
(1)
where $`\stackrel{}{S}_i`$’s are classical vectors, whose magnitude is taken to be $`\left|\stackrel{}{S}_i\right|=\sqrt{S\left(S+1\right)}`$ (i.e $`\left|\stackrel{}{S}_i\right|=\sqrt{2}`$ in this case), and $`J`$ is positive for antiferromagnets.
The thermal expectation value of the on-site magnetic moment has been calculated by the Wolff cluster algorithm in the temperature range $`50KT285K`$, for $`L`$ up to 31 spins. In order to compare numerical data with NMR results in doped Y<sub>2</sub>BaNiO<sub>5</sub>, $`J/k_B`$ has been set equal to 285 K . The effect of the magnetic field $`H`$ has been included using a standard Metropolis algorithms, which associates a flip probability to the whole cluster. As the field intensity is low, the flip acceptance remains close to one.
Typical results of the simulations for different values of $`H`$ are shown in Fig. 1, proving that the response of a finite classical chain to a homogeneous field is non uniform. In Fig. 2, magnetization profiles at characteristic temperatures and $`H`$=14.1 Tesla, are shown together with <sup>89</sup>Y NMR spectra in Mg-doped Y<sub>2</sub>BaNiO<sub>5</sub> (at the same temperatures) from Ref. 5. The spectra were obtained at fixed frequency ($`\nu _{rf}`$ =29.4 MHz) by sweeping the magnetic field $`H_0`$ in a narrow range around 14.1 Tesla. The intensity of the NMR signal is proportional to the number of <sup>89</sup>Y nuclei which obey the resonance condition
$$H_0=2\pi \frac{\nu _{rf}k}{\gamma }AS_i^z,$$
(2)
where $`\gamma `$ is the <sup>89</sup>Y-gyromagnetic factor, $`A=`$1.3 Tesla is the <sup>89</sup>Y - Ni<sup>2+</sup> hyperfine coupling constant in Y<sub>2</sub>BaNiO<sub>5</sub> and $`k`$ is an $`i`$-independent factor that accounts for the chemical and orbital shifts. Using Eq. (2), the local magnetization of the open classical chain and the position of the peaks in the NMR spectra can be directly compared. Since $`k`$ is not known precisely a priori, we fix it by matching the spin-polarization at the centre of the chain with the maximum of the NMR central line. Then, as sketched in Fig. 2, a satellite peak is found in correspondence of each value of magnetization taken by the edge spins.
The magnetization profiles can be analyzed in detail, by observing that $`S_i^z`$ consists in a uniform part and of staggered contributions decaying away from each boundary. The following phenomenological function
$$S_i^z=S_b^z+S_1^z\left[(1)^{(i1)}e^{\frac{(i1)}{\xi }}+(1)^{(Li)}e^{\frac{(Li)}{\xi }}\right]$$
(3)
is thus used to fit the numerical data. Expression (3) describes very well the behavior of $`S_i^z`$ in all the investigated temperature range (see for instance solid lines in Fig. 2) and the extracted fitting parameters - $`S_b^z`$, $`S_1^z`$ and $`\xi `$ \- turn out to be independent of the chain length $`L`$. The two alternating terms in Eq. (3) give an in phase (out of phase) contribution in chains with odd (even) number of spins. In real systems, consisting in an ensamble of segments with different length, such effect induces a distribution of $`S_i^z`$ and thus a broadening of the NMR satellite peaks, as actually observed in Fig. 2. Here we do not investigate this subject quantitatively, being mainly interested to discuss the behavior of the parameters extracted from the classical Monte Carlo data in light of the values for the same quantities obtained by NMR in the $`S`$=1 chain.
In Fig. 3a), the temperature behavior of the uniform term $`S_b^z`$ is reported. $`S_b^z`$ depends linearly on the field up to 40 Tesla (inset) and follows strictly the magnetization $`\chi H`$ that would be obtained by using for $`\chi `$ the zero-field susceptibility of the infinite classical chain (solid line in Fig. 3a)). This result should be intuitively expected, since the staggered terms in Eq. (3) involve only boundary spins and thus, in the thermodynamic limit, the ”bulk” magnetization is due only to $`S_b^z`$. The first-site magnitude $`S_1^z`$ and the decay-length $`\xi `$ of the staggered contributions, extracted from the simulation, are reported respectively in Fig. 3b) and 4 and compared with the experimental values of $`S_1^z`$ and $`\xi `$ from <sup>89</sup>Y NMR in doped Y<sub>2</sub>BaNiO<sub>5</sub> . The agreement between experimental results in the HAF $`S`$=1 one-dimensional compound and predictions of the classical model is remarkable at high temperature. Small quantitative deviations, which likely prelude to a more sensible departure at lower temperatures, are observed only around $`T=`$100K$`\mathrm{\Delta }_H/k_B`$. At the moment, the lack of experimental data for $`S_1^z`$ and $`\xi `$ below 100 K (due to the broadening of the NMR lines) prevents a precise definition of the temperature range in which the classical model accounts for the magnetic response of the open $`S=1`$ chain. Fig. 4 also shows how the decay length $`\xi `$ of the staggered magnetization in a weak external field ($`H`$=14.1 Tesla) tracks strictly the behavior of the zero-field spin-spin correlation length (solid line), calculated by Fisher for the infinite-volume chain.
In the inset of Fig. 4, the total magnetic moment associated to each staggered contribution,
$$S_T^z=\underset{i=1}{\overset{L}{}}S_1^z(1)^{(i1)}e^{\frac{(i1)}{\xi }},$$
(4)
is plotted as a function of temperature. $`S_T^z`$ is well reproduced by a Curie-like law
$$S_T^z=\frac{g\mu _BS_f^2H}{3k_BT},$$
(5)
characteristic of non-interacting classical moments of magnitude $`S_f`$. The value of $`S_f`$ that optimizes the fitting (solid line in the inset in Fig. 4) is 0.72, about half of the site-spin magnitude $`\sqrt{S\left(S+1\right)}=\sqrt{2}`$. This result extends to a temperature region in which the system shows classical behavior the low-$`T`$ picture for gapped spin chains, in which $`S`$/2 degrees of freedom develop at the chain edges. Moreover, since the temperature dependence of $`S_T^z`$ is mainly controlled by $`S_1^z`$ for $`\mathrm{\Delta }_H/k_BTJ/k_B`$, even the first-site staggered moment displays approximatively a Curie-like behavior.
In summary, our numerical results prove that the alternate boundary magnetization observed experimentally in finite-length $`S=1`$ HAF chains is correctly predicted, for $`\mathrm{\Delta }_H/k_BTJ/k_B`$, by a strictly classical analysis of the Heisenbeg model. In light of this evidence we conclude that the occurrence of Curie-like fractional-spin defects is a general feature in HAF one-dimensional systems with open boundaries, not related to the occurrence of a spin gap. On the other hand, the size of these staggered defects, which is fixed by the bulk spin-spin correlation length, is affected by quantum fluctuations and is thus expected to remain finite when the ground state is spin-liquid like.
The authors are grateful to A. Rigamonti, E. Sørensen and M. Fabrizio for stimulating discussions.
|
warning/0007/hep-th0007189.html
|
ar5iv
|
text
|
# Fractional and Integer Charges from Levinson’s Theorem
## I Introduction
Many field theory solitons have especially interesting properties when they are coupled to fermions, because they act as strong background fields that can drastically alter the Dirac spectrum. Solitons that break $`C`$ and $`CP`$ invariance can introduce asymmetries in the Dirac spectrum, causing the soliton to carry fermion number. By adiabatically turning on the soliton from a trivial background, one can observe this fermion number as a level crossing in the Dirac spectrum. In addition, solitons with nontrivial topological boundary conditions can carry fractional fermion number .
Blankenbecler and Boyanovsky showed how to use a phase shift representation of the density of states to calculate these fermion quantum numbers. There is no need to consider any interpolation of the background field. Together with the bound states, the behavior of the phase shift at threshold gives the integer charge. Fractional charges naturally appear in the behavior of the phase shift at large $`k`$.
In recent work on variational computations of soliton energies , it has been essential to know these fermion numbers, since one wants to compare configurations with the same fermion charge. This work has led us to extend the method of Blankenbecler and Boyanovsky in two directions. First, we have shown how to extend dimensional regularization to a phase shift formalism. We are then able to reconsider the anomaly calculation of Ref. starting with an explicitly gauge-invariant regulator, rather than enforcing gauge invariance by hand on the result. We can thus explicitly see the difference between the naive regulator and the gauge-invariant regulator. Second, we demonstrate an extension of this work to chiral bag models in three spatial dimensions, giving a practical example of a fractional fermion number.
Section II describes the Green’s function formalism that relates the phase shifts to the Fock space expansion for the charge. This result can then be strengthened by the use of Levinson’s theorem. In Section III we demonstrate the effects of the choice of regulator in a one-dimensional gauge theory. We show that by choosing a gauge-invariant regulator we automatically ensure that there are no anomalies in the vector current. In Section IV we perform explicit calculations in chiral bag models. In Appendix A, we develop the formalism for computing fermion phase shifts in arbitrary dimensions, and demonstrate that it leads to results that agree with the standard result from Feynman perturbation theory for the contribution of the tadpole graph to the effective energy. In Appendix B, we derive results used in Section IV for the three-dimensional bag.
## II Spectral Analysis in Soliton Backgrounds
### A Density of states
Our primary tool will be the scattering phase shifts. The phase shifts are useful because they unambiguously and quantitatively track the changes of the spectrum of small oscillations around the background fields, even though that spectrum is continuous and infinite, and integrals over it of physical quantities may be divergent.
To derive our results, we consider a single Dirac fermion in $`n`$ space dimensions. We will assume that the soliton background has spherical symmetry, so that the spectrum decomposes into a sum over eigenchannels $`\alpha `$. In one dimension there are just two channels, for even and odd parity, while in three dimensions the sum will run over parity and over either total spin or grand spin, also including the degeneracy within each channel. We begin by writing the fermion charge density as
$$j^0(x)=\frac{1}{2}\mathrm{\Omega }|[\mathrm{\Psi }^{}(x),\mathrm{\Psi }(x)]|\mathrm{\Omega }$$
(1)
where $`x`$ is an $`n`$-dimensional vector. We have been careful to order the anticommuting fermion fields to maintain charge conjugation invariance: $`[\overline{\mathrm{\Psi }},\mathrm{\Psi }]`$ is even under $`𝒞`$ and $`[\overline{\mathrm{\Psi }},\gamma _5\mathrm{\Psi }]`$ and $`[\overline{\mathrm{\Psi }},\gamma ^\mu \mathrm{\Psi }]`$ are odd under $`𝒞`$.
We make the usual Fock decomposition in terms of the eigenstates $`\psi _\alpha ^\omega `$ of the single-particle Dirac equation,
$$\mathrm{\Psi }(x,t=0)=\underset{\alpha }{}\left(_0^{\mathrm{}}b_\alpha ^\omega \psi _\alpha ^\omega (x)𝑑\omega +_{\mathrm{}}^0d_\alpha ^\omega {}_{}{}^{}\psi _{\alpha }^{\omega }(x)𝑑\omega \right)$$
(2)
where the integral over $`\omega `$ includes both continuum and bound states. We have normalized the wavefunctions by
$$\psi _\alpha ^\omega (x)^{}\psi _\alpha ^{}^\omega ^{}(x)𝑑x=\delta (\omega \omega ^{})\delta (\alpha \alpha ^{})$$
(3)
where the delta function of $`\omega `$ on the right hand side is interpreted as a Dirac delta function for continuum states and a Kronecker delta for bound states. The creation and annihilation operators satisfy
$$\{b_\alpha ^\omega ,b_\alpha ^{}^\omega ^{}{}_{}{}^{}\}=\{d_\alpha ^\omega ,d_\alpha ^{}^\omega ^{}{}_{}{}^{}\}=\delta (\omega \omega ^{})\delta (\alpha \alpha ^{})$$
(4)
with all other anticommutators vanishing. Thus the charge density becomes
$$j^0(x)=\frac{1}{2}\underset{\alpha }{}\left(_{\mathrm{}}^{\mathrm{}}\mathrm{sgn}(\omega )|\psi _\alpha ^\omega (x)|^2𝑑\omega \right).$$
(5)
We next consider the Green’s function for the fermion field
$`G(x,y,t)`$ $`=`$ $`i\mathrm{\Omega }|T(\mathrm{\Psi }(x,t)\mathrm{\Psi }(y,0)^{})|\mathrm{\Omega }`$ (6)
$`=`$ $`i{\displaystyle \underset{\alpha }{}}\left({\displaystyle _{\mathrm{}}^0}𝑑\omega e^{i\omega t}\psi _\alpha ^\omega (x)\psi _\alpha ^\omega (y)^{}\mathrm{\Theta }(t){\displaystyle _0^{\mathrm{}}}𝑑\omega e^{i\omega t}\psi _\alpha ^\omega (x)\psi _\alpha ^\omega (y)^{}\mathrm{\Theta }(t)\right)`$ (7)
and its Fourier transform
$$G(x,y,E)=\underset{\alpha }{}_{\mathrm{}}^{\mathrm{}}\frac{d\omega }{2\pi }\frac{\psi _\alpha ^\omega (x)\psi _\alpha ^\omega (y)^{}}{E\omega +i\mathrm{sgn}(\omega )ϵ}$$
(8)
whose trace gives the density of states according to
$$\rho (\omega )=\mathrm{sgn}(\omega )\frac{dN(\omega )}{d\omega }=\mathrm{Im}\mathrm{Tr}\frac{1}{\pi }G(x,x,\omega )d^nx$$
(9)
giving as a result
$$\rho (\omega )=\frac{1}{\pi }\underset{\alpha }{}d^nx|\psi _\alpha ^\omega (x)|^2.$$
(10)
for $`|\omega |>m`$. (For $`|\omega |<m`$, the density of states consists of a sum of delta functions for each bound state.) A more useful form of this equation is obtained by subtracting the free density of states $`\rho ^0(\omega )`$, giving
$$\rho (\omega )\rho ^0(\omega )=\frac{1}{\pi }\underset{\alpha }{}d^nx(|\psi _\alpha ^\omega (x)|^21).$$
(11)
We can also express the density of states in each channel in terms of the S-matrix:
$$\rho _\alpha (\omega )\rho _\alpha ^0(\omega )=\mathrm{sgn}(\omega )\frac{1}{2\pi i}\frac{d}{d\omega }\mathrm{log}S_\alpha (\omega )=\mathrm{sgn}(\omega )\frac{1}{\pi }\frac{d\delta _\alpha (\omega )}{d\omega }.$$
(12)
The total density of states is then
$$\rho (\omega )=\underset{\alpha }{}\rho _\alpha (\omega ).$$
(13)
Putting these results into eq. (5) we thus obtain an integral over the continuum and a sum over bound states
$`Q`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}\left({\displaystyle _m^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}{\displaystyle \frac{d\delta _\alpha (\omega )}{d\omega }}+{\displaystyle \underset{\omega _\alpha ^j>0}{}}{\displaystyle \frac{1}{2}}{\displaystyle _m^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}{\displaystyle \frac{d\delta _\alpha (\omega )}{d\omega }}{\displaystyle \underset{\omega _\alpha ^j<0}{}}{\displaystyle \frac{1}{2}}\right)`$ (14)
$`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{\alpha }{}}\left(\delta _\alpha (m)\delta _\alpha (\mathrm{})\pi n_\alpha ^>+\pi n_\alpha ^<\delta _\alpha (m)+\delta _\alpha (\mathrm{})\right)`$ (15)
where $`n_\alpha ^>`$ and $`n_\alpha ^<`$ give the number of bound states with positive and negative energy respectively in each channel. This is our main result. We will see that it generalizes unchanged to cases with fractional charges, which appear in the phase shift at $`\omega =\pm \mathrm{}`$.
### B Levinson’s Theorem
We have seen that the phase shifts keep track of the rearrangement of the fermion spectrum in a quantitative way, by telling us the density of states. Levinson’s theorem shows that the phase shifts also track how many states have left the continuum at each threshold. This property will give us a physical motivation for eq. (15), and will also enable us to simplify it.
Levinson’s theorem in three dimensions and in the negative parity channel in one dimension gives the number of states $`N`$ that have left the continuum by passing through the threshold at $`m`$ as
$$\delta (m)\delta (\mathrm{})=N\pi .$$
(16)
These states typically appear as bound states (which give delta functions in the density of states), though it is possible that in cases where the spectrum is not charge conjugation invariant, they can reenter the continuum of states with opposite energy. In the positive parity channel in one dimension, Levinson’s theorem is modified to
$$\delta (m)\delta (\mathrm{})=(N\frac{1}{2})\pi .$$
(17)
In one dimension and in the $`\mathrm{}=0`$ channel in three dimensions, there is the possibility of a state whose wavefunction goes to a constant at $`r=\mathrm{}`$, rather than decaying exponentially like a bound state or oscillating like a continuum state. Such “half-bound” states are on the verge of becoming bound and count as $`\frac{1}{2}`$ in Levinson’s theorem. In the free case, for example, $`\delta (\omega )=0`$ for all $`\omega `$, but eq. (17) still holds because there is a half-bound state with $`\psi =\mathrm{constant}`$. For fuller discussion of these results, see Ref. .
Computing the fermion number of a field configuration now becomes a matter of simple counting. We consider each channel separately. If a bound state leaves the positive continuum but appears as a positive energy bound state, it has not changed the fermion number of the configuration. However, if it crosses $`\omega =0`$ and becomes a negative energy bound state, it is now filled in the vacuum and gives a fermion number of one. Thus the fermion number of a field configuration is given by
$$Q=\frac{1}{2\pi }\underset{\alpha }{}\left(\delta _\alpha (m)\delta _\alpha (\mathrm{})\pi n_\alpha ^>+\pi n_\alpha ^<\delta _\alpha (m)+\delta _\alpha (\mathrm{})\right)$$
(18)
which is exactly (15).
This interpretation suggests a further simplification: since Levinson’s theorem tracks all states that enter and leave the two continua, even in the presence of CP-violation, we have the restriction that
$$\delta _\alpha (m)\delta _\alpha (\mathrm{})+\delta _\alpha (m)\delta _\alpha (\mathrm{})\pi n_\alpha ^<\pi n_\alpha ^>=0,$$
(19)
so that
$`Q`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{\pi }}\left(\delta _\alpha (m)\delta _\alpha (\mathrm{})\pi n_\alpha ^>\right)`$ (20)
$`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{\pi }}\left(\pi n_\alpha ^<\delta _\alpha (m)+\delta _\alpha (\mathrm{})\right).`$ (21)
In the positive parity channel in one space dimension, we must subtract 1 from the left-hand side of eq. (19) and $`\frac{1}{2}`$ from the subsequent expressions for $`Q`$ because of the modification to Levinson’s theorem in eq. (17).
## III Electrostatics and the need for regularization
The conserved charges we consider are not renormalized. That is, they do not receive any contributions from the counterterms of the theory. Nonetheless, it is essential to include the effects of the regularization procedure used to define the theory. The example of QED in 1+1 dimensions provides a clear illustration of this subtlety. Although the theory is finite, the regularization process is nontrivial.
The Lagrangian is
$$=\frac{1}{4e^2}F_{\mu \nu }F^{\mu \nu }+\frac{1}{2}[\overline{\mathrm{\Psi }},\left(\gamma ^\mu (i_\mu A_\mu )m\right)\mathrm{\Psi }]$$
(22)
where we again have used the commutator to ensure that the free theory is $`C`$ and $`CP`$ invariant. In ordinary perturbation theory, the vacuum polarization diagram computed in $`d`$ space-time dimensions is
$$\mathrm{\Pi }_{\mu \nu }(p)=2ie^2N_d_0^1𝑑\xi \frac{d^dk}{(2\pi )^d}\frac{2\xi (1\xi )(g_{\mu \nu }p^2p_\mu p_\nu )+g_{\mu \nu }\left(m^2p^2\xi (1\xi )+k^2(\frac{2}{d}1)\right)}{(k^2+p^2\xi (1\xi )m^2)^2}$$
(23)
where $`2N_d`$ is the dimension of the Dirac algebra. If we had not regulated the theory by analytically continuing the space-time dimension, we would not have found the last term, which vanishes if we set $`d=2`$ from the outset. Keeping $`d2`$ shows that this term exactly cancels the two terms that precede it, leaving the transverse form of the vacuum polarization that is required by gauge invariance. Thus we must include in our definition of the field theory the additional information that the theory is regulated in order to preserve gauge invariance at the quantum level, and dimensional regularization provides a convenient way to implement this requirement.
The vacuum polarization diagram reflects the effect of the anomaly. The anomaly is obtained from the leading correction to the vector current, which is related to the polarization tensor by
$$j_\mu (x)=d^dy\mathrm{\Pi }_{\mu \nu }(xy)A^\nu (y)$$
(24)
where $`\mathrm{\Pi }_{\mu \nu }(x)`$ denotes the Fourier transform of eq. (23). Thus a completely transverse polarization tensor corresponds to a conserved vector current. Setting $`d=2`$ from the outset gives an anomalous vector current
$`_\mu j^\mu `$ $`=`$ $`_\mu \overline{\mathrm{\Psi }}\gamma ^\mu \mathrm{\Psi }={\displaystyle \frac{eN_d}{\pi }}_\mu A^\mu `$ (25)
$`_\mu j_5^\mu `$ $`=`$ $`_\mu \overline{\mathrm{\Psi }}\gamma ^\mu \gamma _5\mathrm{\Psi }=0.`$ (26)
Including the contribution proportional to $`\frac{d}{2}1`$ in eq. (23) transfers this anomaly to the axial current
$`_\mu j^\mu `$ $`=`$ $`_\mu \overline{\mathrm{\Psi }}\gamma ^\mu \mathrm{\Psi }=0`$ (27)
$`_\mu j_5^\mu `$ $`=`$ $`_\mu \overline{\mathrm{\Psi }}\gamma ^\mu \gamma _5\mathrm{\Psi }={\displaystyle \frac{eN_d}{2\pi }}ϵ^{\mu \nu }F_{\mu \nu }.`$ (28)
If we choose a configuration with $`A_1=0`$ and adiabatically turn on a configuration $`A_0(x)`$ between $`t=\mathrm{}`$ and $`t=0`$, then integrating eq. (26) we obtain
$$Q=j^0(x)𝑑x=\frac{eN_d}{2\pi }A_0(x)𝑑x$$
(29)
for the naive regulator while $`Q=0`$ in dimensional regularization.
The phase shift approach shows exactly the same behavior. We consider the example of an electrostatic square well potential with depth $`\phi `$ and width $`2L`$. First we ignore subtleties of regularization and compute directly in $`d=2`$. The phase shift in the negative parity channel $`\delta ^{}(\omega )`$ is determined by
$$\frac{m+\omega }{k}\mathrm{tan}(kL+\delta ^{}(\omega ))=\frac{m+\omega +e\phi }{q}\mathrm{tan}qL$$
(30)
with $`k=\sqrt{\omega ^2m^2}`$ and $`q=\sqrt{(\omega +e\phi )^2m^2}`$. Similarly, the phase shift in the positive parity channel $`\delta ^+(\omega )`$ is determined by
$$\frac{k}{m+\omega }\mathrm{tan}(kL+\delta ^+(\omega ))=\frac{q}{m+\omega +e\phi }\mathrm{tan}qL.$$
(31)
As $`\omega \pm \mathrm{}`$, the total phase shift approaches $`\pm 2e\phi L`$, giving a fractional contribution to the total fermion charge in agreement with eq. (29). Although such fractional charges are possible (and we will see examples of them in the next section), in this case the result indicates that the method of calculation has not preserved gauge invariance.
To preserve gauge invariance, we regularize the computation by computing the phase shifts as analytic functions of the space dimension $`n`$. Since we are only concerned with the contribution from $`\omega \pm \mathrm{}`$, we can consider just the leading Born approximation. In Appendix A we have extended the method of dimensional regularization of the phase shifts in Ref. to fermions. In arbitrary dimensions, the total phase shift is a sum over channels labeled by total spin $`j=\frac{1}{2},\frac{3}{2}\mathrm{}`$ and by parity. Summing over parity, the leading Born approximation to total phase shift in each $`j`$ channel is given by
$$\delta _{n,j}^{(1)}=\omega e\pi _0^{\mathrm{}}A_0(r)\left(J_{\frac{n}{2}+j\frac{3}{2}}(kr)^2+J_{\frac{n}{2}+j\frac{1}{2}}(kr)^2\right)r𝑑r$$
(32)
which has degeneracy $`d(j)`$ given by eq. (A20). Summing over $`j`$ using eq. (A17) and eq. (A27) yields the leading Born approximation to the total phase shift in $`n`$ space dimensions
$`\delta _n^{(1)}(\omega )`$ $`=`$ $`\omega e\pi N_d{\displaystyle \underset{\mathrm{}=0}{\overset{\mathrm{}}{}}}D(\mathrm{}){\displaystyle _0^{\mathrm{}}}A_0(r)J_{\frac{n}{2}+\mathrm{}1}(kr)^2r𝑑r`$ (33)
$`=`$ $`\omega k^{n2}{\displaystyle \frac{N_de\pi }{2^{n2}\mathrm{\Gamma }(\frac{n}{2})^2}}{\displaystyle _0^{\mathrm{}}}A_0(r)r^{n1}𝑑r`$ (34)
$`=`$ $`\omega k^{n2}{\displaystyle \frac{N_dL^ne\phi \pi }{2^{n2}n\mathrm{\Gamma }(\frac{n}{2})^2}}`$ (35)
which reduces to $`N_de\phi L`$ if we send $`n1`$ and take the limit as $`\omega \pm \mathrm{}`$. But the order of these limits is essential: if we first regulate the theory by holding the dimension fixed at $`n<1`$, we then see that the contribution as $`\omega \pm \mathrm{}`$ vanishes. Only after we have taken the $`\omega \pm \mathrm{}`$ limits do we send $`n1`$. This procedure, dictated by dimensional regularization, preserves gauge invariance and gives no fractional charge.
We note that other regularization methods commonly used in phase shift calculations, such as zeta-function regularization, would not preserve gauge invariance and would thus lead to the same spurious fractional result.
## IV Fractional Charges
### A Chiral bag model in one dimension
Chiral bag models provide simple illustrations of fractional fermion numbers. We begin with a Dirac fermion in one dimension on the half-line $`x>0`$, subject to the boundary condition
$$ie^{i\gamma _5\theta }\mathrm{\Psi }=\gamma ^1\mathrm{\Psi }$$
(36)
at $`x=0`$, with $`\frac{\pi }{2}\theta \frac{\pi }{2}`$. We consider the Dirac Hamiltonian
$$\gamma ^0(i\gamma ^1_x+m)\mathrm{\Psi }=\omega \mathrm{\Psi }$$
(37)
in the basis $`\gamma ^0=\sigma _3`$ and $`\gamma ^1=i\sigma _2`$. The free solutions are
$$\varphi ^\pm (x)=\left(\begin{array}{c}\pm k\\ \omega m\end{array}\right)e^{\pm ikx}$$
(38)
where $`k=\sqrt{\omega ^2m^2}`$. The full solutions are then given in terms of the phase shifts by
$$\psi ^\pm (x)=\varphi ^{}(x)+e^{2i\delta ^\pm (\omega )}\varphi ^+(x).$$
(39)
We can then solve for the phase shifts using the boundary condition, eq. (36). We obtain
$`\mathrm{cot}\delta ^+(\omega )`$ $`=`$ $`{\displaystyle \frac{k}{\omega m}}\mathrm{tan}\beta `$ (40)
$`\mathrm{tan}\delta ^{}(\omega )`$ $`=`$ $`{\displaystyle \frac{k}{\omega m}}\mathrm{tan}\beta `$ (41)
where $`\beta =\frac{\pi }{4}\frac{\theta }{2}`$. To find the bound states, we look for solutions of the form
$$\varphi (x)=\left(\begin{array}{c}i\kappa \\ \omega m\end{array}\right)e^{\kappa x}$$
(42)
with $`\kappa =\sqrt{m^2\omega ^2}`$. Imposing the boundary condition gives
$$\kappa =(m+\omega )\mathrm{tan}\beta $$
(43)
so there is always exactly one bound states for $`\frac{\pi }{2}\theta \frac{\pi }{2}`$. Plugging these results into eq. (15), we find that the fermion number is $`\frac{\theta }{\pi }`$.
### B Chiral bag model in 3 dimension
This simple model generalizes naturally to 3 dimensions. We consider an isodoublet of Dirac fermions subject to the boundary condition
$$ie^{i\theta \stackrel{}{\tau }\widehat{n}\gamma _5}\mathrm{\Psi }=\stackrel{}{\gamma }\widehat{n}\mathrm{\Psi }$$
(44)
imposed on a sphere of radius $`R`$, where $`\stackrel{}{\tau }`$ are the isospin Pauli matrices. This condition is not invariant under space and isospin rotations individually, but it is invariant under combined space and isospin rotations, and under parity. Thus we can decompose the scattering problem into eigenchannels labelled by grand spin $`G=0,1,2\mathrm{}`$ and parity. This calculation is outlined in Appendix B.
In the $`G=0`$ channel, the phase shifts in the two parity channels are
$$\delta _0^\pm (\omega )=\mathrm{Arg}\left(ih_1(kR)\frac{k}{\pm \omega m}\mathrm{cos}\theta +ih_0(kR)(1\mathrm{sin}\theta )\right)$$
(45)
where $`h_n(x)`$ are spherical Hankel functions of the first kind and $`k=\sqrt{\omega ^2m^2}`$. Extracting the contribution from the $`G=0`$ channel is straightforward. Using eq. (15), we find a contribution of $`\frac{\theta }{\pi }`$ to the fermion number.
For $`G>0`$ there are two states for each choice of $`G`$ and parity. The total phase shift in each channel is given by
$`\delta _G^\pm (\omega )=\mathrm{Arg}`$ $`(\mathrm{sin}\theta {\displaystyle \frac{h_G(kR)^2}{(\omega m)R}}+{\displaystyle \frac{k}{\pm \omega m}}h_G(kR)(h_{G+1}(kR)h_{G1}(kR))`$ (47)
$`\mathrm{cos}^2\theta [h_{G+1}(kR)h_{G1}(kR)\left({\displaystyle \frac{k}{\pm \omega m}}\right)^2h_G(kR)^2]).`$
The contribution from $`G>0`$ needs to be treated with care. For large $`k`$, the dominant contribution comes from $`G+\frac{1}{2}kR`$. We must follow a consistent regularization procedure in order to obtain the correct order of limits. In this model a simple cutoff suffices. We first compute the total phase shift at $`k=\mathrm{\Lambda }`$, summed over all partial waves. This sum then has a smooth limit as $`\mathrm{\Lambda }\mathrm{}`$. If we had taken the limit in the other order by considering the $`k\mathrm{}`$ limit in each partial wave separately, we would incorrectly conclude that the contribution from $`G>0`$ was identically zero.
For numerical computations, it is convenient to first consider $`\frac{dQ}{d\theta }`$ and then integrate to obtain $`Q`$ as a function of $`\theta `$. Figure 1 shows the $`\frac{dQ}{d\theta }`$ computation for one value of $`\theta `$. Fixing a large value of the cutoff $`\mathrm{\Lambda }`$, we then sum over both parities from $`G=0`$ up to $`G_{\mathrm{max}}\mathrm{\Lambda }R`$, so that the contribution from the higher values of $`G`$ is negligible. We then compute the $`\frac{dQ}{d\theta }`$ for each $`\theta `$ from eq. (15) using $`k=\mathrm{\Lambda }`$ in place of $`k=\mathrm{}`$.
Carrying out this computation numerically for each $`\theta `$, we find agreement with the established result
$$\frac{dQ}{d\theta }=\frac{1}{\pi }(1\mathrm{cos}2\theta ).$$
(48)
so that
$$Q=\frac{1}{\pi }(\theta \mathrm{sin}\theta \mathrm{cos}\theta ).$$
(49)
## V Conclusions
By explicitly introducing appropriate regulators, we have been able to unambiguously compute fermion quantum numbers in a variety of models. We have seen that by introducing the regulator as part of the definition of the theory, we have no need (and no freedom) to adjust the results of our computations after the fact. This determinism leads to a computation that is both theoretically elegant and computationally practical.
### Acknowledgments
We would like to thank S. Forte, J. Goldstone, and V. P. Nair for helpful conversations, suggestions, and references. This work is supported in part by funds provided by the U.S. Department of Energy (D.O.E.) under cooperative research agreement #DF-FC02-94ER40818 and the Deutsche Forschungsgemeinschaft (DFG) under contract We 1254/3-1.
## A Dirac Equation in Arbitrary Dimensions
In this Appendix we discuss the Dirac equation for a radially symmetric potential $`V=V(r)`$ for an arbitrary number of spatial dimensions $`n`$. We calculate the leading Born approximation to the phase shift in $`n`$ dimensions and find its contribution to the Casimir energy. We verify that this result agrees with what we obtain from the tadpole graph in standard Feynman perturbation theory using dimensional regularization.
Our starting point is the free Dirac equation in $`n`$ spatial dimensions,
$$\left(i\stackrel{}{\alpha }\stackrel{}{}+\beta m\right)\mathrm{\Psi }=H\mathrm{\Psi }=\omega \mathrm{\Psi }.$$
(A1)
The spinor $`\mathrm{\Psi }`$ has $`2N_d`$ components and accordingly the Dirac matrices $`\alpha _j,j=1,\mathrm{},n`$ and $`\beta `$ have $`2N_d\times 2N_d`$ elements, where $`N_d=2^{(n1)/2}`$ for $`n`$ odd and $`N_d=2^{n/2}`$ for $`n`$ even. We will generalize the case of $`n`$ odd, though our results will not depend on this choice. We choose the basis
$$\beta =\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)\mathrm{and}\alpha _j=\left(\begin{array}{cc}0& \mathrm{\Lambda }_j\\ \mathrm{\Lambda }_j& 0\end{array}\right)j=1,\mathrm{},n.$$
(A2)
The Clifford algebra is obtained by demanding the anti–commutator $`\{\mathrm{\Lambda }_i,\mathrm{\Lambda }_j\}=2\delta _{ij}`$. In analogy to the spin generators in $`n=3`$ we define the commutator
$$[\mathrm{\Lambda }_i,\mathrm{\Lambda }_j]=2i\mathrm{\Sigma }_{ij}$$
(A3)
which obeys the $`SO(n)`$ commutation relation
$$[\mathrm{\Sigma }_{ij},\mathrm{\Sigma }_{kl}]=i\left(\delta _{ik}\mathrm{\Sigma }_{jl}+\delta _{jl}\mathrm{\Sigma }_{ik}\delta _{il}\mathrm{\Sigma }_{jk}\delta _{jk}\mathrm{\Sigma }_{il}\right).$$
(A4)
We define the orbital angular momentum operator
$$L_{ij}=i\left(x_i_jx_j_i\right)$$
(A5)
which also satisfies the $`SO(n)`$ algebra
$$[L_{ij},L_{kl}]=i\left(\delta _{ik}L_{jl}+\delta _{jl}L_{ik}\delta _{il}L_{jk}\delta _{jk}L_{il}\right).$$
(A6)
We can then put these together to form the total spin operator
$$J_{ij}=L_{ij}+\frac{1}{2}\mathrm{\Sigma }_{ij}$$
(A7)
which commutes with the Hamiltonian: $`[H,J_{ij}]=0`$. Having obtained the algebra, we next need to find the Casimir eigenvalues of
$$L^2=\frac{1}{2}\underset{i,j}{}L_{ij}^2\mathrm{\Sigma }^2=\frac{1}{2}\underset{i,j}{}\mathrm{\Sigma }_{ij}^2\mathrm{and}J^2=\frac{1}{2}\underset{i,j}{}J_{ij}^2.$$
(A8)
The eigenvalues of $`L^2`$ are those of $`SO(n)`$, $`\mathrm{}(\mathrm{}+n2)`$. To find $`\mathrm{\Sigma }^2`$, we consider its trace, which is just the number of independent matrices $`\mathrm{\Sigma }_{ij}`$, $`\frac{1}{2}n(n1)`$. Then to obtain the Casimir eigenvalue of $`J^2`$, all we need to find is $`L\mathrm{\Sigma }=\frac{1}{2}_{i,j}L_{ij}\mathrm{\Sigma }_{ij}`$. We use the second order equations obtained from eq. (A1), which are generalized Bessel equations. We first remark that $`\widehat{r}\stackrel{}{\mathrm{\Lambda }}`$ has zero total spin ($`[\widehat{r}\stackrel{}{\mathrm{\Lambda }},J_{ij}]=0`$) and eigenvalue $`\mathrm{}=1`$ with respect to $`L^2`$. Therefore the appropriate spinor with definite parity can be parameterized as
$$\mathrm{\Psi }=\left(\begin{array}{c}if(r)Y_{l,s,j}\\ g(r)\left(\widehat{r}\stackrel{}{\mathrm{\Lambda }}\right)Y_{l,s,j}\end{array}\right)$$
(A9)
where $`Y_{l,s,j}`$ denote generalized spinor spherical harmonics. The radial functions obey the coupled first order equations
$`\left[_r+{\displaystyle \frac{n1+R}{2r}}\right]g(r)`$ $`=`$ $`(m\omega )f(r)`$ (A10)
$`\left[_r+{\displaystyle \frac{n1R}{2r}}\right]f(r)`$ $`=`$ $`(m+\omega )g(r)`$ (A11)
where $`R=L\mathrm{\Sigma }+n1`$ contains the desired eigenvalue. We can decouple these equations to obtain second order equations for $`f(r)`$ and $`g(r)`$. By demanding that $`f(r)`$ and $`g(r)`$ obey generalized Bessel equations with orbital angular momentum $`\mathrm{}`$ and $`\mathrm{}^{}`$ respectively, we find
$$R=1\pm (n22\mathrm{})\mathrm{and}R=1\pm (n2+2\mathrm{}^{}).$$
(A12)
In view of the above mentioned properties of $`\widehat{r}\stackrel{}{\mathrm{\Lambda }}`$ we have $`\mathrm{}^{}=\mathrm{}\pm 1`$. Hence the two relations in eq. (A12) are consistent if
$$R=n+2\mathrm{}1\mathrm{for}\mathrm{}^{}=\mathrm{}+1\mathrm{and}R=3n2\mathrm{}\mathrm{for}\mathrm{}^{}=\mathrm{}1.$$
(A13)
Putting these results together we find the Casimir eigenvalue
$$J^2=\mathrm{}(\mathrm{}+n2)+\frac{n(n1)}{8}+\{\begin{array}{cc}\mathrm{},\hfill & \mathrm{}^{}=\mathrm{}+1\hfill \\ 2n\mathrm{},\hfill & \mathrm{}^{}=\mathrm{}1\hfill \end{array}.$$
(A14)
Defining $`j=\frac{1}{2}(\mathrm{}+\mathrm{}^{})`$ yields
$$J^2=\left(j\frac{1}{2}\right)\left(j+n\frac{3}{2}\right)+\frac{n(n1)}{8}$$
(A15)
for both cases. The above definition of $`j`$ also ensures that (as for $`n=3`$) there are two independent solutions for a given $`j`$: (i) $`\mathrm{}=j+\frac{1}{2},\mathrm{}^{}=j\frac{1}{2}`$ and (ii) $`\mathrm{}=j\frac{1}{2},\mathrm{}^{}=j+\frac{1}{2}`$. These solutions have opposite parity.
Finally, we have to find the degeneracy factor, $`d(j)`$. We use two trace relations that are simple in the basis appropriate for $`\mathrm{\Sigma }_{ij}`$ and $`L_{ij}`$. Written in the basis for $`J_{ij}`$ they connect different representations and provide information about the degeneracy factors. We have
$$\underset{j=\mathrm{}\frac{1}{2}}{\overset{j=\mathrm{}+\frac{1}{2}}{}}d(j)(R(j)n+1)=0$$
(A16)
and
$$\underset{j=\mathrm{}\frac{1}{2}}{\overset{j=\mathrm{}+\frac{1}{2}}{}}d(j)=N_dD(\mathrm{}).$$
(A17)
where we have defined $`d(\frac{1}{2})=0`$ independent of $`n`$. The first condition is nothing but the tracelessness of $`L\mathrm{\Sigma }`$ while the second gives the number of states for a given orbital angular momentum $`\mathrm{}`$,
$$D(\mathrm{})=\frac{\mathrm{\Gamma }(n+\mathrm{}2)}{\mathrm{\Gamma }(n1)\mathrm{\Gamma }(\mathrm{}+1)}(n+2\mathrm{}2).$$
(A18)
Eq. (A16) can be re–expressed as
$$d(\mathrm{}+\frac{1}{2})=\frac{n+\mathrm{}2}{\mathrm{}}d(\mathrm{}\frac{1}{2})$$
(A19)
which after substitution into eq. (A17) yields the final result
$$d(j)=N_d(j+\frac{1}{2})\frac{\mathrm{\Gamma }(n+j\frac{3}{2})}{\mathrm{\Gamma }(n1)\mathrm{\Gamma }(j+\frac{3}{2})}.$$
(A20)
Equation (A19) represents a recursion relation between $`d(j)`$ and $`d(j+1)`$ that can straightforwardly be shown to be satisfied by the degeneracy factor of eq. (A20). We note that as $`n1`$, eq. (A20) gives zero in all channels except $`j=\frac{1}{2}`$, where it is one. As in the bosonic case, this limit gives the reduction to the positive and negative parity channels in one dimension.
To show how the dimensional regularization of the phase shifts corresponds to ordinary dimensional regularization of Feynman perturbation theory, we compute the contribution to the Casimir energy from the leading Born approximation to the phase shifts using the method of , and compare it to the energy of the tadpole graph, with both quantities evaluated in $`n`$ dimensions.
The leading Born approximation to the phase shift is, summing over parity channels,
$$\delta _{n,j}^{(1)}(k)=\frac{\pi }{2}_0^{\mathrm{}}𝑑rV(r)r\left(J_{\frac{n}{2}+j\frac{3}{2}}(kr)^2+J_{\frac{n}{2}+j\frac{1}{2}}(kr)^2\right)$$
(A21)
for $`k=\sqrt{\omega ^2m^2}`$. The leading Born approximation to the Casimir energy is then given by
$$\mathrm{\Delta }E_n^{(1)}[\stackrel{}{\varphi }]=\frac{1}{\pi }_0^{\mathrm{}}𝑑k(\sqrt{k^2+m^2}m)\underset{j}{}d(j)\frac{d\delta _{n,j}^{(1)}}{dk}.$$
(A22)
Summing over all channels we find the total phase shift
$`\delta _n^{(1)}(k)`$ $`=`$ $`{\displaystyle \underset{j=\frac{1}{2},\frac{3}{2}\mathrm{}}{}}d(j)\delta _{n,j}^{(1)}(k)`$ (A23)
$`=`$ $`{\displaystyle \frac{\pi }{2}}{\displaystyle _0^{\mathrm{}}}𝑑rV(r)r{\displaystyle \underset{j=\frac{1}{2},\frac{3}{2}\mathrm{}}{}}d(j)\left(J_{\frac{n}{2}+j\frac{3}{2}}(kr)^2+J_{\frac{n}{2}+j\frac{1}{2}}(kr)^2\right)`$ (A24)
$`=`$ $`{\displaystyle \frac{\pi }{2}}{\displaystyle _0^{\mathrm{}}}𝑑rV(r)r{\displaystyle \underset{\mathrm{}=0,1,2\mathrm{}}{}}\left(d(\mathrm{}+{\displaystyle \frac{1}{2}})J_{\frac{n}{2}+\mathrm{}1}(kr)^2+d(\mathrm{}{\displaystyle \frac{1}{2}})J_{\frac{n}{2}+\mathrm{}1}(kr)^2\right)`$ (A25)
$`=`$ $`{\displaystyle \frac{\pi }{2}}{\displaystyle _0^{\mathrm{}}}𝑑rV(r)r{\displaystyle \underset{\mathrm{}=0,1,2\mathrm{}}{}}N_dD(\mathrm{})J_{\frac{n}{2}+\mathrm{}1}(kr)^2.`$ (A26)
Using the Bessel function identity
$$\underset{\mathrm{}=0}{\overset{\mathrm{}}{}}\frac{(2q+2\mathrm{})\mathrm{\Gamma }(2q+\mathrm{})}{\mathrm{\Gamma }(\mathrm{}+1)}J_{q+\mathrm{}}(z)^2=\frac{\mathrm{\Gamma }(2q+1)}{\mathrm{\Gamma }(q+1)^2}\left(\frac{z}{2}\right)^{2q}$$
(A27)
and setting $`q=\frac{n}{2}1`$, we sum over $`\mathrm{}`$ and obtain for the Casimir energy
$$\mathrm{\Delta }E^{(1)}[\stackrel{}{\varphi }]=2N_d\frac{V}{(4\pi )^{\frac{n}{2}}\mathrm{\Gamma }\left(\frac{n}{2}\right)}(n2)_0^{\mathrm{}}(\omega m)k^{n3}𝑑k$$
(A28)
where
$$V=V(x)d^nx=\frac{2\pi ^{\frac{n}{2}}}{\mathrm{\Gamma }\left(\frac{n}{2}\right)}_0^{\mathrm{}}V(r)r^{n1}𝑑r.$$
(A29)
The $`k`$ integral can be calculated in the vicinity of $`n=\frac{1}{2}`$ and then analytically continued, yielding
$$_0^{\mathrm{}}(\omega m)k^{n3}𝑑k=\frac{m^{n1}}{4\sqrt{\pi }}\mathrm{\Gamma }\left(\frac{1n}{2}\right)\mathrm{\Gamma }\left(\frac{n2}{2}\right).$$
(A30)
Hence we find
$$\mathrm{\Delta }E^{(1)}[\stackrel{}{\varphi }]=2N_d\frac{V}{(4\pi )^{\frac{n+1}{2}}}\mathrm{\Gamma }\left(\frac{1n}{2}\right)m^{n1}$$
(A31)
which is exactly what we obtain using standard dimensional regularization for the tadpole diagram in $`n+1`$ space–time dimensions.
## B Chiral Bag Model S-matrix
In this Appendix we outline the derivation of eq. (45) and eq. (47). We begin with spinors that are eigenstates of parity and total grand spin , where grand spin $`\stackrel{}{G}`$ is the sum of total spin $`\stackrel{}{j}=\stackrel{}{l}+\frac{1}{2}\stackrel{}{\sigma }`$ and isospin $`\frac{1}{2}\stackrel{}{\tau }`$. For a given grand spin $`G`$ with $`z`$-component $`M`$, we will find the scattering wavefunctions in terms of the spherical harmonic functions $`𝒴_{j,\mathrm{}}`$ with $`j=G\pm \frac{1}{2}`$ and $`\mathrm{}=j\pm \frac{1}{2}`$. These are two-component spinors in both spin and isospin space. While grand spin is conserved, the boundary condition in eq. (44) mixes states with different ordinary spin $`j`$. For the channels with parity $`()^G`$ we have two spinors that solve the Dirac equation away from the boundary:
$$\left(\begin{array}{c}ig_1(r)𝒴_{G+\frac{1}{2},G}\\ f_1(r)𝒴_{G+\frac{1}{2},G+1}\end{array}\right)\mathrm{and}\left(\begin{array}{c}ig_2(r)𝒴_{G\frac{1}{2},G}\\ f_2(r)𝒴_{G\frac{1}{2},G1}\end{array}\right).$$
(B1)
For zero grand spin the second spinor is absent. We now introduce linear combinations that define the S-matrix:
$`\mathrm{\Psi }_1`$ $`=`$ $`\left(\begin{array}{c}iw^+h_G(kr)𝒴_{G+\frac{1}{2},G}\\ w^{}h_{G+1}(kr)𝒴_{G+\frac{1}{2},G+1}\end{array}\right)+S_{11}^{+,G}\left(\begin{array}{c}iw^+h_G^{}(kr)𝒴_{G+\frac{1}{2},G}\\ w^{}h_{G+1}^{}(kr)𝒴_{G+\frac{1}{2},G+1}\end{array}\right)`$ (B4)
$`+S_{21}^{+,G}\left(\begin{array}{c}iw^+h_G^{}(kr)𝒴_{G\frac{1}{2},G}\\ w^{}h_{G1}^{}(kr)𝒴_{G\frac{1}{2},G1}\end{array}\right)`$
$`\mathrm{\Psi }_2`$ $`=`$ $`\left(\begin{array}{c}iw^+h_G(kr)𝒴_{G\frac{1}{2},G}\\ w^{}h_{G1}(kr)𝒴_{G\frac{1}{2},G1}\end{array}\right)+S_{22}^{+,G}\left(\begin{array}{c}iw^+h_G^{}(kr)𝒴_{G\frac{1}{2},G}\\ w^{}h_{G1}^{}(kr)𝒴_{G\frac{1}{2},G1}\end{array}\right)`$ (B8)
$`+S_{12}^{+,G}\left(\begin{array}{c}iw^+h_G^{}(kr)𝒴_{G+\frac{1}{2},G}\\ w^{}h_{G+1}^{}(kr)𝒴_{G+\frac{1}{2},G+1}\end{array}\right)`$
where $`h_{\mathrm{}}(kr)`$ refers to the spherical Hankel functions suitable to parameterize an incoming spherical wave. We have introduced the kinematic factors $`w^+=\sqrt{1+\frac{m}{\omega }}`$ and $`w^{}=\mathrm{sgn}(\omega )\sqrt{1\frac{m}{\omega }}`$, where $`\omega =\pm \sqrt{k^2+m^2}`$ and $`m`$ are the energy and mass of the Dirac particle respectively. In the case $`G=0`$, the components with $`j=G\frac{1}{2}`$ are absent and the S-matrix has only a single component $`S_{11}^{+,0}=\mathrm{exp}(2i\delta _0^+)`$.
Imposing the boundary condition, eq. (44), on these wavefunctions gives
$`\left(\begin{array}{cc}\mathrm{cos}\theta & i\widehat{r}\stackrel{}{\tau }\mathrm{sin}\theta i\widehat{r}\stackrel{}{\sigma }\\ i\widehat{r}\stackrel{}{\tau }\mathrm{sin}\theta +i\widehat{r}\stackrel{}{\sigma }& \mathrm{cos}\theta \end{array}\right)\mathrm{\Psi }_n|_{r=R}=0\mathrm{for}j=1,2.`$ (B9)
For each $`n=1,2`$ the projection onto grand spin spherical harmonics yields two equations, which allows us to extract all four components of the S-matrix. It is convenient to express the result in the form of a matrix equation:
$`\left(\begin{array}{cc}X& Y\\ \overline{X}& \overline{Y}\end{array}\right)\left(\begin{array}{cc}S_{11}^{+,G}& S_{12}^{+,G}\\ S_{21}^{+,G}& S_{22}^{+,G}\end{array}\right)=\left(\begin{array}{cc}X^{}& Y^{}\\ \overline{X}^{}& \overline{Y}^{}\end{array}\right)M^{}`$ (B10)
where the star denotes complex conjugation. The components of the matrix $`M`$ are given by
$`X`$ $`=`$ $`h_G^{}(kR)\mathrm{cos}\theta +{\displaystyle \frac{k}{\omega +m}}\left(1+{\displaystyle \frac{1}{2G+1}}\mathrm{sin}\theta \right)h_{G+1}^{}(kR)`$ (B11)
$`\overline{X}`$ $`=`$ $`h_G^{}(kR)\mathrm{cos}\theta +{\displaystyle \frac{k}{\omega +m}}\left(1+{\displaystyle \frac{1}{2G1}}\mathrm{sin}\theta \right)h_{G1}^{}(kR)`$ (B12)
$`Y`$ $`=`$ $`{\displaystyle \frac{2\sqrt{G(G+1)}}{2G+1}}{\displaystyle \frac{k}{\omega +m}}h_{G1}^{}(kR)\mathrm{sin}\theta `$ (B13)
$`\overline{Y}`$ $`=`$ $`{\displaystyle \frac{2\sqrt{G(G+1)}}{2G+1}}{\displaystyle \frac{k}{\omega +m}}h_{G+1}^{}(kR)\mathrm{sin}\theta .`$ (B14)
The total phase shift in this channel is then given by
$$\delta _G^+=\mathrm{Arg}\left(\mathrm{det}S^{+,G}\right)=\mathrm{Arg}\left(\frac{\mathrm{det}M^{}}{\mathrm{det}M}\right)$$
(B15)
leading to eq (47). For the channels with parity $`()^{G+1}`$ the computation proceeds analogously from the definition of the corresponding S-matrix:
$`\mathrm{\Psi }_1`$ $`=`$ $`\left(\begin{array}{c}iw^+h_{G+1}(kr)𝒴_{G+\frac{1}{2},G+1}\\ w^{}h_G(kr)𝒴_{G+\frac{1}{2},G}\end{array}\right)+S_{11}^{,G}\left(\begin{array}{c}iw^+h_{G+1}^{}(kr)𝒴_{G+\frac{1}{2},G+1}\\ w^{}h_G^{}(kr)𝒴_{G+\frac{1}{2},G}\end{array}\right)`$ (B18)
$`+S_{21}^{,G}\left(\begin{array}{c}iw^+h_{G1}^{}(kr)𝒴_{G\frac{1}{2},G1}\\ w^{}h_G^{}(kr)𝒴_{G\frac{1}{2},G}\end{array}\right)`$
$`\mathrm{\Psi }_2`$ $`=`$ $`\left(\begin{array}{c}iw^+h_{G1}(kr)𝒴_{G\frac{1}{2},G1}\\ w^{}h_G𝒴_{G\frac{1}{2},G}\end{array}\right)+S_{22}^{,G}\left(\begin{array}{c}iw^+h_{G1}^{}(kr)𝒴_{G\frac{1}{2},G1}\\ w^{}h_G^{}𝒴_{G\frac{1}{2},G}\end{array}\right)`$ (B22)
$`+S_{12}^{,G}\left(\begin{array}{c}iw^+h_{G+1}^{}(kr)𝒴_{G+\frac{1}{2},G+1}\\ w^{}h_G^{}(kr)𝒴_{G+\frac{1}{2},G}\end{array}\right).`$
|
warning/0007/quant-ph0007096.html
|
ar5iv
|
text
|
# Finite resolution of time in continuous measurements: phenomenology and the modelPublished in Physics Letters A231, 1-8 (1997)
## 1 Introduction
During last decades the theory of continuous quantum measurements has been under thorough investigation both with the help of models and in the framework of different phenomenological approaches The interest to this field significantly increased in connection with the quantum Zeno effect predicted in and experimentally verified in . Phenomenological approaches have an advantage of being model-independent. One of the first approaches of this type applied to continuous quantum measurements was one based on restricted path integrals (RPI). It was proposed by the author (see also ) following an idea of R.Feynman . A Lindblad-type master equation for the density matrix of an open measured system can be derived from this approach . Analogous equations follow from concrete models of continuous quantum measurements .
To describe monitoring a quantum observable in the framework of the RPI approach, one has to define quantum corridors, corresponding to different readouts of the measurement. In the preceding works the quantum corridors were used which corresponded to the assumption that monitoring is performed with the ideal resolution of time. In the present paper we shall consider a more general definition of quantum corridors including the effect of a finite resolution of time. The evolution of a measured system is presented in the resulting theory by an influence functional, but it cannot be described by a differential equation (for example a master equation).
Finally the results of the phenomenological consideration will be compared with conclusions based on a model. For this goal a modification of the model will be presented which allows one to take into account a finite resolution of time. The description of the measured system following from the modified model will be shown to agree with the conclusions of the RPI approach.
## 2 Quantum corridors
A measured system is considered in the RPI theory of continuous measurements as an open system. The back influence of a measuring device (environment) onto the measured system is taken into account by restricting the Feynman path integral presenting the propagator. The restriction is determined by the information about the measured system supplied by the measurement. Let us outline this approach (see for details).
The evolution of a closed quantum system during a time interval $`T`$ is described by the evolution operator $`U_T`$. A matrix element of this operator between the states with definite positions (in the configuration space) is called the propagator and may be expressed in the form of the Feynman path integral (the variables $`q`$ and $`p`$ may be multidimensional)
$$U_T(q^{\prime \prime },q^{})=q^{\prime \prime }|U_T|q^{}=_q^{}^{q^{\prime \prime }}\mathrm{d}[p,q]\mathrm{exp}\left[\frac{i}{\mathrm{}}_0^T(p\dot{q}H(p,q,t))\right].$$
(1)
If the system undergoes a continuous (prolonged in time) measurement and therefore is considered as being open, its evolution may be described (in the RPI approach) by the set of partial evolution operators $`U_T^\alpha `$ depending on outputs (readouts) $`\alpha `$ of the measurement:
$$|\psi _T^\alpha =U_T^\alpha |\psi _0,\rho _T^\alpha =U_T^\alpha \rho _0\left(U_T^\alpha \right)^{}.$$
(2)
The partial propagators are expressed by restricted path integrals. This means that the path integral for $`U_T^\alpha `$ must be of the form (1) but with the integration restricted according to the information given by the measurement readout $`\alpha `$. The information given by $`\alpha `$ may be presented by a weight functional $`w_\alpha [p,q]`$ (positive, with values between 0 and 1) so that the partial propagator has to be written as a weighted path integral
$$U_T^\alpha (q^{\prime \prime },q^{})=q^{\prime \prime }|U_T^\alpha |q^{}=_q^{}^{q^{\prime \prime }}\mathrm{d}[p,q]w_\alpha [p,q]\mathrm{exp}\left[\frac{i}{\mathrm{}}_0^T(p\dot{q}H(p,q,t))\right].$$
(3)
The probability density for $`\alpha `$ to arise as a measurement readout is given by the trace of the density matrix $`\rho _T^\alpha `$ so that the probability for $`\alpha `$ to belong to some set $`𝒜`$ of readouts is
$$\mathrm{Prob}(\alpha 𝒜)=_𝒜d\alpha \mathrm{Tr}\rho _T^\alpha $$
(4)
with an appropriate measure $`\mathrm{d}\alpha `$ on the set of readouts.
The preceding consideration concerns the situation when the measurement readout $`\alpha `$ is known (a selective description of the measurement). If the readout is unknown (a non-selective description), the evolution of the measured system may be presented by the complete density matrix
$$\rho _T=d\alpha \rho _T^\alpha =d\alpha U_T^\alpha \rho _0\left(U_T^\alpha \right)^{}.$$
(5)
The generalized unitarity condition
$$d\alpha \left(U_T^\alpha \right)^{}U_T^\alpha =\mathrm{𝟏}$$
(6)
provides conservation of probabilities.
In the special case, when monitoring an observable $`A=A(p,q,t)`$ is considered as a continuous measurement, the measurement readout is given by the curve
$$[a]=\{a(t)|0tT\}$$
characterizing values of this observable in different time moments. If the square average deflection is taken as a measure of the deviation of the observable $`A(t)=A(p(t),q(t),t)`$ from the readout $`a(t)`$, then the weight functional describing the measurement may be taken<sup>1</sup><sup>1</sup>1The choice of the weight functional depends on the class of measurements under consideration. in the Gaussian form:
$$w_{[a]}[p,q]=\mathrm{exp}\left[\kappa _0^T(A(t)a(t))^2dt\right].$$
(7)
The constant $`\kappa `$ characterizes the resolution of the measurement and may be expressed in terms of the “measurement error” $`\mathrm{\Delta }a_T`$ achieved during the period $`T`$ of the measurement, $`\kappa =1/T\mathrm{\Delta }a_T^2`$. The error $`\mathrm{\Delta }a_T`$ decreases with the duration $`T`$ of the measurement increased, $`\mathrm{\Delta }a_T1/\sqrt{T}`$.
The resulting path integral
$$U_T^{[a]}(q^{\prime \prime },q^{})=_q^{}^{q^{\prime \prime }}\mathrm{d}[p,q]\mathrm{exp}\left\{\frac{i}{\mathrm{}}_0^T\left(p\dot{q}H\right)dt\kappa _0^T\left(A(t)a(t)\right)^2dt\right\}$$
(8)
has the form of a conventional (non-restricted) Feynman path integral (1) but with the non-Hermitian effective Hamiltonian
$$H_{[a]}(p,q,t)=H(p,q,t)i\kappa \mathrm{}\left(A(p,q,t)a(t)\right)^2$$
(9)
instead of the original Hamiltonian $`H`$. Therefore, the partial propagator (8) satisfies a Schrödinger equation with the effective Hamiltonian.
This allows one to describe a continuous measurement (monitoring) without calculating a restricted path integral. Instead, one may solve the Schrödinger equation (with the effective Hamiltonian) for a wave function of the system:
$$\frac{}{t}|\psi _t=\frac{i}{\mathrm{}}H_{[a]}=\left(\frac{i}{\mathrm{}}H\kappa \left(Aa(t)\right)^2\right)|\psi _t.$$
(10)
If the initial wave function $`\psi _0`$ corresponds to the initial state of the measured system, then the solution $`\psi _T`$ in the final time moment presents the state of the system after the measurement, under the condition that the measurement readout is $`[a]`$.
The wave function $`\psi _T`$ obtained in this way has a non-unit norm. If the initial wave function is normalized, then the norm of the final wave function, according to Eq. (4), determines the probability density of the measurement output: $`P[a]=\psi _T^2`$. Solving the Schrödinger equation for the same initial condition but for different readouts $`[a]`$, one has a probability distribution over all possible scenarios of the measurement with the corresponding final states of the measured system.
The non-selective description of the measurement (if the readout is unknown) is given by the density matrix $`\rho _t`$ defined by (5) and satisfying the equation
$$\dot{\rho }=\frac{i}{\mathrm{}}[H,\rho ]\frac{\kappa }{2}[A,[A,\rho ]].$$
(11)
The influence of the measuring device (measuring medium) on the measured system may be described by an influence functional $`W[p,q|p^{},q^{}]`$ (see ) in the sense that the ‘superpropagator’ describing the evolution of the density matrix
$$\rho _T(q,q^{})=dq_0dq_0^{}U(q,q^{}|q_0,q_0^{})\rho _0(q_0,q_0^{})$$
(12)
may be presented in the form of a double path integral:
$`U(q,q^{}|q_0,q_0^{})={\displaystyle _q^{q_0}}\mathrm{d}[p,q]{\displaystyle _q^{}^{q_0^{}}}\mathrm{d}[p^{},q^{}]W[p,q|p^{},q^{}]`$ (13)
$`\times `$ $`\mathrm{exp}\left[{\displaystyle \frac{i}{\mathrm{}}}{\displaystyle _0^T}[(p\dot{q}H(p,q,t))(p^{}\dot{q}^{}H(p^{},q^{},t))]\right]`$
In the present case the influence functional may easily be derived from Eq. (5) and has the form
$$W[p,q|p^{},q^{}]=\mathrm{d}[a]w_{[a]}[p,q]w_{[a]}[p^{},q^{}].$$
(14)
With the weight functional (7) this gives, up to an inessential number factor,
$$W[p,q|p^{},q^{}]=\mathrm{exp}\left[\frac{\kappa }{2}_0^Tdt(A(p,q,t)A(p^{},q^{},t))^2\right].$$
(15)
## 3 Quantum corridors for a finite time resolution
It has been assumed in the preceding arguments that the time is measured precisely. This means that the number $`a(t)`$ is an estimate, supplied by the measurement, of a value $`A(t)`$ of the observable $`A`$ at the precisely known instant $`t`$. This assumption is not always realistic. A real device gives an estimate of the observable $`A`$ over some finite time interval of the duration, say, $`\tau `$. We shall refer to this situation as “measuring time with the resolution $`\tau `$”.
To be more concrete, let $`a(t)`$ be an estimate, due to the measurement, of the entity
$$\overline{A}(t)=A_t=dt^{}\mathrm{\Pi }_t(t^{})A(t^{})$$
(16)
defined by an appropriate ‘form-factor’ $`\mathrm{\Pi }_t`$ (depending on the type of the measuring device). Then, instead of Eq. (7), the weight functional for restricting the path integral should be defined as follows:
$$\stackrel{~}{w}_{[a]}[p,q]=\mathrm{exp}\left[\kappa _0^T(a(t)\overline{A}(t))^2dt\right].$$
(17)
We arrive therefore, instead of Eq. (8), to the following expression for the partial propagator:
$$U_T^{[a]}(q^{\prime \prime },q^{})=_q^{}^{q^{\prime \prime }}\mathrm{d}[p,q]\mathrm{exp}\left\{\frac{i}{\mathrm{}}_0^T\left(p\dot{q}H\right)dt\kappa _0^T\left(\overline{A}(t)a(t)\right)^2dt\right\}.$$
(18)
It differs radically in that $`\overline{A}(t)`$ depends on $`A(t^{})`$ for different time moments $`t^{}`$. Therefore, the description of the evolution is ‘not local in time’ and cannot be reduced to an effective Hamiltonian in analogy with Eqs. (910).
The partial propagators (18) describe an evolution of the measured system selectively, i.e. with the measurement output $`[a]`$ taken into account. A non-selective description is given by the general formula (5) resulting in the present case in an influence functional of the form
$`\stackrel{~}{W}[p,q|p^{},q^{}]`$ $`=`$ $`{\displaystyle \mathrm{d}[a]\stackrel{~}{w}_{[a]}[p,q]\stackrel{~}{w}_{[a]}[p^{},q^{}]}`$ (19)
$`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{\kappa }{2}}{\displaystyle _0^T}dt(\overline{A}(p,q,t)\overline{A}(p^{},q^{},t))^2\right].`$
###### Remark 1
The measure $`\mathrm{d}[a]`$ of integration over measurement readouts has to be chosen in such a way that the generalized unitarity Eq. (6) be valid. Eq. (19) is valid with the conventional functional measure $`\mathrm{d}[a]_t\mathrm{d}a(t)`$ which provides the generalized unitarity either for a linear measured system or for a system with a not too large nonlinearity and a measurement with a not too high resolution. I a general case a weight has to be included in the measure $`\mathrm{d}[a]`$. Eq. (19) should be modified in this case.
The influence functional (19) enables one to describe the evolution of the measured system by Eqs. (1213) (with $`\stackrel{~}{W}`$ instead of $`W`$), but no differential equation in time (analogous to Eq. (11)) exists for the resulting density matrix. This is a consequence of non-locality in time.
In the rest of the paper we shall consider a concrete model of the monitoring the position of a particle to verify that it actually leads to the influence functional of the form Eq. (19).
## 4 Finite resolution in the model of a continuous measurement
The well-known model of a quantum diffusion proposed by Caldeira and Leggett may be considered as a model for a continuous measurement, namely, for monitoring the position of a particle by a measuring medium. In this model the decoherence (measurement) is caused by the interaction of the particle with modes of the crystal.
One more model of this type has been proposed in . This model also consists of a particle in some medium. However the “atoms” of the medium are modelled as oscillators not interacting with each other. Decoherence is caused in this case by interaction of the particle with internal degrees of freedom of the atoms (presented as degrees of freedom of the oscillators). Let us remark that in most cases such a model is more realistic than the Caldeira-Leggett model because the decoherence due to internal structure of atoms is more efficient (more fast) than the decoherence by modes of a crystal. Now we shall modify this model to take into account a finite resolution of time.
The measuring medium in the model consists of atoms in the nods of a cubic lattice. Each atom is presented as an oscillator. We shall present each atom as a family of oscillators of different frequencies. However let us begin by the case of a single oscillator with the frequency $`\omega `$ as in . The Hamiltonian of the interaction between the oscillator and the particle is taken in as
$$H_{\mathrm{int}}=\underset{k}{}H_k=\underset{k}{}\gamma _\omega q_k\mathrm{exp}\left[\frac{(𝐫𝐜_k)^2}{l^2}\right].$$
(20)
Here $`𝐫`$ is a position of the measured particle while $`q_k`$ is a canonical coordinate of the $`k`$th oscillator, $`𝐜_k`$ its location, $`\gamma _\omega `$ the interaction constant, and the length $`l`$ characterizes the range of the interaction. The interaction of the particle with each of the atoms (oscillators) may be considered as a measurement of its location with the precision $`l`$.
To solve this model, the discrete lattice of atoms was replaced in by a continuous distribution of them with the constant density $`n`$. This enables one to calculate the influence functional describing the influence of the medium on the particle:
$`W[𝐫_1,𝐫_2]`$ $`=`$ $`\mathrm{exp}\{{\displaystyle _0^T}\mathrm{d}t{\displaystyle _0^T}\mathrm{d}t^{}\nu (\omega )\mathrm{cos}\omega (tt^{})[\mathrm{exp}({\displaystyle \frac{(𝐫_1(t)𝐫_1(t^{}))^2}{2l^2}})`$ (21)
$`+`$ $`\mathrm{exp}({\displaystyle \frac{(𝐫_2(t)𝐫_2(t^{}))^2}{2l^2}})2\mathrm{exp}({\displaystyle \frac{(𝐫_2(t)𝐫_1(t^{}))^2}{2l^2}})]\}`$
where
$$\nu (\omega )=n\left(\frac{\pi l^2}{2}\right)^{3/2}\frac{\gamma _\omega ^2}{4\mathrm{}m\omega }.$$
(22)
This simple model is not enough realistic because it does not describe measuring of time: the state of an oscillator after the interaction with the particle contains no information about the time of the interaction. To describe the measurement of time, it was assumed in that the interaction of each oscillator is turned on for a short time. Let us consider now a more realistic model for the measurement of time.
For this goal, assume that each atom has a more complicated internal structure so that the information about the time of the interaction is recorded in its state. To give a model of the internal structure of the atom, let us present it as a family of oscillators with different frequencies. Now the information about the time of interaction with the particle is recorded in phase relations between different oscillators of the same ‘atom’.
Going over to the calculation, we have now integrate the exponent in Eq. (21) over the frequencies $`\omega `$ of the oscillators forming the model of an atom. The influence functional takes the form
$`W[𝐫_1,𝐫_2]=\mathrm{exp}\{{\displaystyle \frac{\kappa l^2}{2}}{\displaystyle _0^T}\mathrm{d}t{\displaystyle _0^T}\mathrm{d}t^{}\mathrm{\Pi }(tt^{})[\mathrm{exp}({\displaystyle \frac{(𝐫_1(t)𝐫_1(t^{}))^2}{2l^2}})`$ (23)
$`+\mathrm{exp}({\displaystyle \frac{(𝐫_2(t)𝐫_2(t^{}))^2}{2l^2}})2\mathrm{exp}({\displaystyle \frac{(𝐫_2(t)𝐫_1(t^{}))^2}{2l^2}})]\}`$
with
$$\mathrm{\Pi }(t)=\frac{2}{\kappa l^2}\nu (\omega )\mathrm{cos}\omega t\mathrm{d}\omega ,\mathrm{\Pi }(t)dt=1.$$
(24)
Let the width of the ‘form-factor’ $`\mathrm{\Pi }(t)`$ be of the order of $`\tau `$. Then for not too large initial energy of the particle, $`E_0Ml^2/2\tau ^2`$, the exponentials in the square brackets may be evaluated up to the first order to give
$`W[𝐫_1,𝐫_2]=\mathrm{exp}\{{\displaystyle \frac{\kappa }{4}}{\displaystyle _0^T}\mathrm{d}t^{}{\displaystyle _0^T}\mathrm{d}t^{\prime \prime }\mathrm{\Pi }(t^{}t^{\prime \prime })`$ (25)
$`[2(𝐫_2(t^{})𝐫_1(t^{\prime \prime }))^2(𝐫_1(t^{})𝐫_1(t^{\prime \prime }))^2(𝐫_2(t^{})𝐫_2(t^{\prime \prime }))^2]\}`$
Let the function $`\mathrm{\Pi }(t)`$ be symmetrical. Then it may be expressed through the function of two arguments $`\mathrm{\Pi }_t(t^{})`$ (depending only on the difference $`|tt^{}|`$):
$$\mathrm{\Pi }(t^{}t^{\prime \prime })=dt\mathrm{\Pi }_t(t^{})\mathrm{\Pi }_t(t^{\prime \prime }),\mathrm{\Pi }_t(t^{})dt^{}=1.$$
(26)
Let us substitute this expression in Eq. (25) and make use of the relation
$`{\displaystyle dt^{}dt^{\prime \prime }\mathrm{\Pi }_t(t^{})\mathrm{\Pi }_t(t^{\prime \prime })}`$ (27)
$`\times `$ $`[2(𝐫_2(t^{})𝐫_1(t^{\prime \prime }))^2(𝐫_1(t^{})𝐫_1(t^{\prime \prime }))^2(𝐫_2(t^{})𝐫_2(t^{\prime \prime }))^2]`$
$`=`$ $`2{\displaystyle dt[𝐫_2_t𝐫_1_t]^2}`$
where the notation (16) is exploited. Then we immediately see that the influence functional has the form
$$W[𝐫,𝐫^{}]=\mathrm{exp}\left[\frac{\kappa }{2}_0^Tdt(𝐫_t𝐫^{}_t)^2\right]$$
(28)
in accord with Eq. (19). Thus, the prediction of the phenomenological RPI approach coincides with what follows from the concrete model of the measurement even in the case when a finite resolution of time is taken into account.
## 5 Conclusion
We have shown, both in the RPI phenomenological approach and in the framework of the concrete model of a continuous measurement, that the finiteness of the resolution in measuring time moments leads to the replacement of the value of an observable $`A(t)`$ by its ‘time-coarse-graining’ $`\overline{A}(t)=A_t`$ in the expression for the influence functional. In a special case when the position $`𝐫(t)`$ of a particle is continuously measured, $`𝐫(t)`$ has to be replaced by $`𝐫_t`$. As a result, the time evolution of the measured system cannot be presented by a differential equation in time. Neither the master equation Eq. (11) nor the Schrödinger equation with a complex Hamiltonian Eq. (10) is correct in this case.
Time coarse-graining is a characteristic of the ‘measuring device’ or ‘measuring medium’. It characterizes inertial properties of the measuring setup. It is evident that the effect of time coarse-graining is negligible if all physically relevant frequencies are lower than the inverse period of coarse-graining: $`\mathrm{\Omega }\tau ^1`$. In this case the master equation and the Schrödinger equation with a complex Hamiltonian are correct. Physically this means that inertial properties of the measuring medium are negligible.
ACKNOWLEDGEMENT
The author is indebted to V.Namiot for stimulating discussions. This work was supported in part by the Deutsche Forschungsgemeinschaft.
|
warning/0007/hep-ph0007040.html
|
ar5iv
|
text
|
# 1 The standard model of elementary particles
## 1 The standard model of elementary particles
In this section we define the standard model of elementary particles (SM) by introducing the particle content and the parameters of this model. We aim at giving a short introduction, at illustrating how this model is used, and at discussing (some of) its merits and limitations (to go deeper in the subjects touched, one could make reference to the seminal papers on the SM , to books and reviews , and to the Particle Data Group biannual report ).
### 1.1 The basic blocks
There are two groups of spin 1/2 particles, quark and leptons, divided in 4 groups with different values of the electric charge $`Q`$
$$\text{MATTER FERMIONS}\{\begin{array}{ccccc}\text{leptons}\hfill & Q=0,\hfill & \nu _\mathrm{e}\nu _\mu \nu _\tau ;\hfill & Q=1,\hfill & \mathrm{e}\mu \tau .\hfill \\ \text{quarks}\hfill & Q=2/3,\hfill & uct;\hfill & Q=1/3,\hfill & dsb.\hfill \end{array}$$
(respectively named: e-, mu- and tau-neutrinos; electron, muon, tau; up, charm and top quarks; down, strange and bottom quarks; recall the existence of their anti-particles, with opposite charge). A different mass distinguishes among the three particles in each group<sup>2</sup><sup>2</sup>2Incidentally, mass and spin are fundamental entities (=they identify certain irreducible representations) of the underlying space-time group of symmetry (which has as subgroups the 4-dimensional translations and rotations–Lorentz subgroup). Note however that such an axiomatic definition of “mass” requires that the system can be considered isolated, that is usually true only in particular conditions. (neutrino masses are to a certain extent special, and we will discuss them in a few pages). Thence, the fundamental particles can be arranged in three “families”, with increasingly heavier members–but otherwise identical. Free quarks have never been observed. It is thought that they can manifest as such only in very energetic processes; and that they necessarily bind to form the “hadrons”
$$\text{HADRONS}\{\begin{array}{cc}\text{mesons}\text{ (}q\overline{q}\text{ states–bosons)}\hfill & \pi ^\pm \pi ^0K\rho \mathrm{}\varphi \mathrm{}J/\mathrm{\Psi }\mathrm{}\hfill \\ \text{baryons}\text{ (}qqq\text{ states–fermions)}\hfill & pn\mathrm{\Delta }^\pm \mathrm{}\mathrm{\Lambda }\mathrm{}\mathrm{\Lambda }_b\mathrm{}\hfill \end{array}$$
the binding is provided by “strong” interactions, a prerogative of the quarks; the forces between nuclei are regarded as residual strong interactions.
Actually, the list of stable spin 1/2 particles is even shorten than the above ones; the electron, the proton $`p`$ (perhaps), the neutrinos (perhaps).
Then come the integer spin fundamental particles, that have the special role of mediating the interactions:
* The graviton (spin 2) related to gravitational forces (that, strictly speaking are not part of the SM);
* the gluon $`g`$ (spin 1) that carries strong interactions;
* the photon $`\gamma `$ (spin 1) that carries electromagnetic interactions;
* the $`W^\pm `$ and $`Z^0`$ bosons (spin 1) that originate charged and neutral (current) weak interactions (responsible for instance of the nuclear $`\beta `$ decays–with emission of electrons);
* and finally the Higgs boson (a scalar, with spin zero) that mediates Yukawa interactions (see below).
The last one is a hypothetical particle; however it has an outstanding importance in the SM.
### 1.2 How interactions are understood
There is a concept that underlies the theory of spin 1 interactions: the gauge principle (which can be given an essential role in formulating the theory of gravity, too; and perhaps, the one of Yukawa interactions—but this is already a speculation). We will introduce it, starting to venture in the formalism. Let us consider the equation that describes the free propagation of a relativistic spin 1/2 particle:
$$[i\underset{\mu =0,1,2,3}{}\gamma ^\mu \frac{}{x^\mu }m]\psi (x)=0.\text{(}\text{Dirac}\text{ equation)}$$
(1)
Here, $`m`$ is the mass of the particle; $`\gamma ^\mu `$ are $`4\times 4`$ matrices that guarantee the covariance of the equation; $`\psi `$ has 4 components, necessary to describe the spin of a massive particle (thenceforth named: 4-spinor). The spinorial index and $`x`$ describe the “space-time” transformation properties of $`\psi ;`$ if we add more indices, we can construct equations that are covariant under other symmetries<sup>3</sup><sup>3</sup>3It is well known that symmetries have a central role in elementary physics, being related to conservation laws; the gauge principle further augments their importance.. To be specific, let us consider the transformation $`\psi _aU_{ab}\psi _b,`$ where we parameterise $`U`$ by introducing the “generators” $`T_{ab}^B`$: $`U=\mathrm{exp}(i_B\alpha ^BT^B)`$ (the imaginary unity $`i`$ in the equations above are just due to the tradition; it is important instead to recall that the number of generators is a characteristic of the group of symmetry; 1 for U(1), 3 for SU(2), 8 for SU(3), etc—SU($`n`$) being the group of unitary matrices of dimension $`n`$ with unit determinant). The symmetry consist in the fact that the trasformed spinor still obeys the Dirac equation (=the equation is “covariant” under the transformation); and this is easy to prove, since neither the mass nor the $`\gamma `$-matrices are transformed.
However, in consideration of the local character of the space-time, one is lead to wonder whether local symmetries $`U(x)=\mathrm{exp}(i_B\alpha ^B(x)T^B)`$ also hold true. One readily verifies that the Dirac equation is not covariant in this enlarged context, due to the fact that the partial derivative $`/x^\mu `$ (that describes how the particle propagates in the space-time) do act on the parameters of the transformation. The “gauge principle” consists in insisting on covariance, by modifying the Dirac equation in the following manner:
$$\begin{array}{c}\left[i\underset{\mu =0,1,2,3}{}\gamma ^\mu \left(\frac{}{x^\mu }+igA_\mu ^BT^B\right)m\right]\psi (x)=0\hfill \\ A_\mu ^BT^BUA_\mu ^BT^BU^1\frac{i}{g}U\frac{U^1}{x^\mu }\hfill \end{array}$$
(2)
we introduced a parameter (“coupling”) $`g,`$ and a set of fields labelled with space-time index $`\mu `$ (that transforms as a 4-vector=that describes a spin 1 particle); the indices $`a,b\mathrm{}`$ are not indicated explicitly and a sum over $`B`$ is understood. Different group of symmetries, different particles. Note that in the special case of the U(1) symmetry group, only one particle $`A_\mu `$ must be introduced, and due to the commutativity in the group, it has simple transformation properties since $`UA_\mu TU^1=A_\mu T.`$ Identifying $`g`$ with the electron charge e, $`T`$ with value of the charge in unities of e, it becomes evident that the U(1) theory is just the electromagnetism<sup>4</sup><sup>4</sup>4Actually, this is the reason why one speaks of “gauge principle”, since we rather directly generalise the well-known gauge invariance to induce the existence of new interactions related to new symmetries.. For non-commutative groups (say, SU($`n`$), where $`UVVU`$ in general), instead, one notes that these spin-1 particles transform in a non trivial manner also under global (non-local) transformations, exactly due to the $`UA_\mu ^BT^BU^1`$ term. Then, it comes without surprise the fact that these particle interact not only with the matter fermions (quarks and leptons) but also among themselves; however, this fact is of great significance, since it means that the “superposition principle” of ordinary electromagnetic interactions (“the light does not interact with the light”) is not of general validity<sup>5</sup><sup>5</sup>5This is among the most prominent peculiarity of the strong interactions, and a crucial ingredient to account for their behaviour. Note incidentally that the existence of bound states of gluons (“glueballs”) has been postulated, and there is theoretical support and circumstantial evidence of the correctness of this hypothesis, see e.g. ..
### 1.3 First foundation of the standard model
Now we can appreciate the first foundation of the standard model, namely, its gauge group
$$G_{321}=\text{SU(3)}_c\times \text{SU(2)}_L\times \text{U(1)}_Y$$
the sign $`\times `$ means that the three subgroups commute (=they do not communicate among them—neither they do, incidentally, with the space-time symmetries). The SU(3)<sub>c</sub> part is related to the existence of gluons, the rest (“electroweak” group) to the photon and the spin 1 bosons related to weak interactions.
How matter fermions behave under this group? The assignment is done in a visibly asymmetric manner (see table 1), since the “left” and “right” spinors have different behaviours (even worser: only the “left” neutrino is assumed to exist–an ontological asymmetry). With the adjectives “left” and “right”, we refer to the most elementary objects that describe particles with spin 1/2, called (Weyl) 2-spinors<sup>6</sup><sup>6</sup>6In fact, the Lorentz group SO(1,3) can be seen as a “complexification” of the group SO(4)$``$SU(2)$`{}_{L}{}^{}\times `$SU(2)$`_R,`$ (the symbol “$``$” means that the two algebræ are the same, and the indices “$`L`$” and “$`R`$” distinguish the two copies of SU(2)). We perceive then the close analogy with the usual group of rotations, SO(3)$``$SU(2) (relation familiar from quantum mechanics). The representations of the Lorentz group can be then labelled by a pair of integers $`|n,m|;`$ the smallest non-trivial ones, $`|1,0|`$ and $`|0,1|,`$ denote left and right 2-spinors respectively. Both of them are needed to construct a Dirac spinor.. The “left-right” asymmetric assignment of the elementary particles in the SM accounts for a peculiarity of the weak interactions, the violation of the parity reflection symmetry. The generator of electric charge is the sum of the third SU(2)<sub>L</sub> generator and of the U(1)<sub>Y</sub> (hypercharge) generator $`Y`$:
$$Q=T_{3L}+Y$$
By using table 1 and previous equation, one can check that left and right states couple with the same charge $`Q`$ to the photon, as it should be for parity to be conserved in electromagnetic interactions (the same is true for the gluons). At this stage, we introduced 3 gauge couplings $`g_3,`$ $`g_2`$ and $`g_1;`$ one parameter each gauge subgroup.
The elementary interactions can be conveniently represented as shown in fig. 1. Each of these plots corresponds to a certain “Feynman rule” (elementary diagram), that are used to calculate the amplitudes of probabilities of the admitted physical processes. For instance, the Feynman rule of the first plot (quark-quark-gluon) tells us the presence of a $`g_3`$ coupling constant, and other factors related to the spin structure, too; it corresponds actually to the term of interaction between $`\psi _a`$ and $`A_\mu ^B`$ in equation<sup>7</sup><sup>7</sup>7Usually, a theory of the interactions is not formulated in terms of the equations of motions, but in terms of a Lagrangian, invariant under the symmetry. In this example, the invariant terms that is represented by the first Feynman rule is $`gT_{ab}^B\times \overline{\psi }_a\gamma ^\mu A_\mu ^B\psi _b.`$ Trying to resemble the notation of section 1.2 (and what is shown in the Feynman rule) we can illustrate this as an “elementary reaction” $`\psi _b\psi _aA_\mu ^B;`$ we warn however that this cannot be considered as an actual reaction (due to energy non-conservation) but only as an element of a reaction. 2. Joining the elementary diagrams in the admitted manners (matching the type of lines) one obtains the list of the possible processes in terms of “Feynman diagrams”. Feynman diagrams are in fact a very convenient manner to organise the actual computation; in the following, however, we will use them mostly to illustrate the content of the theory, and not to perform computations.
All this story of symmetries and interactions is beautiful, but there is an high price to pay for that:
> No mass is allowed, exactly due to $`G_{321}`$ (standard model) gauge invariance!
### 1.4 Origin of the masses
Here we come to the second foundation of the SM. The masses arise in a completely particular manner: through a symmetry breaking ascribed to the vacuum . More specifically, it is postulated that a scalar SU(2)<sub>L</sub>-doublet exists (the Higgs boson), with hypercharge $`1/2,`$ and it obtains a vacuum expectation value:
$$H=\left|\begin{array}{c}H^0\\ H^{}\end{array}\right|\text{ such that }\mathrm{Re}[H^0]0$$
This hypothesis is not as strange as it might sound, since it is a rather common situation in physics of condensed matter, e.g. a spontaneous magnetisation breaks the rotational symmetry<sup>8</sup><sup>8</sup>8A similar effect exists for strong interactions: The vacuum is responsible of the fact that the quarks in the proton have an effective (“constituent”) mass, and this is in strict correspondence with a “dynamical” reduction of symmetry .. In consequence of this assumption, the electroweak invariance is lost (only the electric charge symmetry remains untouched), and the $`W^\pm `$ and $`Z^0`$ bosons acquire mass. Also, by introducing 9 new parameters between the Higgs boson and the quarks and leptons (Yukawa couplings), one can account for their masses. See figure 2, and note the left-right structure of the couplings of quarks and leptons (Dirac type mass, like the one in section 1.2). The only communication between different families in the SM arises at this point, and involves only the charged weak currents, To explain this well, it is necessary to go into some subtleties. (1) The particles that have interactions with the $`W`$ are only the doublets, $`q_L`$ and $`l_L,`$ so we have interaction structures like $`u_L^iW^+d_L^i;`$ however, (2) (as we see from fig. 2) the particles with definite mass require to match left and right quark (lepton) states, namely, $`d_L^i`$ should match with $`d_R^i,`$ and similarly for $`u`$ quarks and charged leptons. So, these two are different prescriptions, or in other terms there is a clash between the “interaction eigenstates” and “mass eigenstates”. Normally, people refer to mass eigenstates (e.g. the top; the electron; etc.); thence charged currents must be non-diagonal (see fig. 3). For leptons, the absence of neutrino masses implies that the non-diagonality is just formal (of no physical significance): e.g. $`\nu _\mathrm{e}`$ is by definition the state associated to the electron by charged weak currents.
We are close to the end of the list of parameters, and we must introduce now the self-interaction of the Higgs particle (scalar potential):
$$U=\frac{\lambda }{4}\left(|H|^2v^2\right)^2$$
(3)
with two parameters, $`\lambda `$ (the Higgs boson self-coupling) and $`v`$ (the vacuum expectation value, as it should be clear). Notice that we omitted to list an additive constant; this has no dynamical meaning if gravity is ignored, otherwise, it should be identified as (a contribution to) the vacuum energy, and we know by sure that this is quite small in comparison with the “natural” scale $`v^410^{50}`$ GeV/cm<sup>3</sup>!!! Actually, there are still some more parameters, related to the non-abelian gauge bosons (one for SU(3)<sub>c</sub> and one for SU(2)<sub>L</sub>) that formally can be written as interaction terms like $`\theta \times G_{ab}G_{cd}ϵ^{abcd}.`$ While the one related to SU(2)<sub>L</sub> is considered harmless, the one related to SU(3)<sub>c</sub> poses serious problems for phenomenology (“strong CP” problem ), and thence has to be small. We will not list these in the following, but we have to stress that these are very delicate and mysterious points of the SM; surely they indicate (some of) its frontiers.
### 1.5 Successes and troubles
Let us summarise the content of the model by some counting. The number of matter fermions for family is 15, 3 leptons, and 4 quarks which come in three types (“colour”). Three families, 45 matter particles. Then there are 9 massless spin 1 bosons (photon and gluons), and three that are massive instead. The number of parameters is
$$18(=3+9+4+2).$$
These are all the fundamental parameters<sup>9</sup><sup>9</sup>9We put aside those parameters like those in form factors, partonic distributions, fragmentation functions, masses of hadrons… that cannot be calculated and have to be measured at present, but that are believed not to have a fundamental nature. Progresses are expected from computer simulations of the SU(3)<sub>c</sub> theory on a discretized space-time (lattice). that are compatible with the requisite of “renormalizability”. In short, this means “calculability” of the theory; in more diffuse terms, a theory of quantised fields (that represent particles) is termed renormalizable when the quantum fluctuations (referred as “loops”, in connection with their representations by Feynman diagrams) produce relatively harmless infinities, namely only those that can be re-absorbed in the parameters of the theory. So a non-renormalizable theory must be modified to become consistent, for instance:
(1) adding new parameters (no need in the SM–well, apart from three more of them, the $`\theta `$ parameters and the cosmological constant, that we brutally set to zero);
(2) or new fields (e.g. if only left electrons were to exist, the electromagnetic theory would be inconsistent due to so-called quantum anomalies; a right handed electron cures this problem. More complex is the case of electroweak interactions, but due to a conspiracy of all the particles in a family, no such trouble exists);
(3) or sometimes, it is necessary to reinterpret the theory as a non-fundamental one (e.g. effective 4 fermion interactions result when a virtual $`W`$ boson is exchanged between two pairs of them; but a 4 fermion fundamental interaction, in itself, would be not renormalizable).
The proof that the SM is renormalizable is not simple, however, gauge invariance turns out to be a crucial ingredient for that .
The list of successes of the SM is impressivly long; broadly, can be divided in those related to strong interactions (with few parameters) and those of electroweak interactions (with several parameters). We will limit here to two of them, that are of rather non-trivial nature (for a full account, books and reviews should be consulted):
(1) Electroweak precision measurements
Due to the assumption that the Higgs particle is a doublet, the scalar potential has a special structure, that ensures a relation between the $`W`$ and $`Z`$ masses and the gauge couplings of the SU(2)<sub>L</sub> and U(1)<sub>Y</sub> groups ($`g_2`$ and $`g_1`$) at the lowest order:
$$\frac{M_W^2}{M_Z^2}=\frac{1}{1+(g_1/g_2)^2}$$
(4)
this relation would not hold, for instance, if the Higgs boson were in the “triplet” representations of SU(2)$`_L.`$ This argument suggests that the hypothesis of “doublets” is correct, at least in first approximation. Actually, there are calculable corrections to this relation due to “loop” effects, which increase the left-hand side by:
$$\rho =1+\frac{3}{4}\frac{g_2^2}{(4\pi )^2}\times \frac{m_{top}^2}{M_W^2}+\text{smaller terms;}$$
(5)
$`(4\pi )^2`$ is the typical loop factor. These calculations have been first made by Veltman ; the predictions have been verified at LEP and other experiments (see for instance the reviews in ). What we want to emphasise is that a new parameter, the top mass, enters into the game, so that precise measurements of $`M_W,`$ $`M_Z`$ and of the coupling constants give informations on the top mass, even if we were to ignore that the top quark existed <sup>10</sup><sup>10</sup>10An analogy can be drawn with ordinary quantum mechanics: the second-order perturbative corrections to the $`i`$-th level, $`\delta E_i^{(2)}=_j|V_{ij}|^2/(E_iE_j)`$ depends on the intermediate $`j`$-th levels, even if we don’t see them directly.. Similarly, by this method it is possible to get some information on the Higgs boson mass. The sensitivity is only logarithmic, much weaker than the one to the top mass, however present data are precise enough to show an indication for a relatively light Higgs boson (assuming that this is the only new particle that contributes).
(2) Neutral currents
As we saw, in the SM it is possible to change the family only at the price of emitting a charged spin 1 boson (fig. 3). However, by a combination of two charged currents, it is possible to have an effective neutral current transition with change of family. The diagram shown in fig. 4 gives in fact an amplitude for electromagnetic dipole transition, with a dipole $`D`$ of the size
$$D(d_id_j\gamma )=V_{ki}^{}V_{kj}\times \frac{eg_2^2}{(4\pi )^2}\times \frac{m_{d_i}}{M_W^2}\times f\left(\frac{m_{u_k}^2}{M_W^2}\right)\text{where }ij$$
(6)
($`f`$ is an adimensional loop function ). Beside the crucial electroweak ingredient, to obtain the correct quantitative estimate it is necessary to consider gluonic corrections ; so, it is rather remarkable such a transition ($`bs\gamma `$) has been observed, and with a rate that agrees with the one predicted in the SM<sup>11</sup><sup>11</sup>11There is nothing similar in the standard model for leptonic transition. So, a positive observation of, say, $`\mu \mathrm{e}\gamma `$ transition would be of enormous importance. The existence of this transition is actually predicted in extensions of the SM, like in simple supersymmetric models..
The SM has also some weak points; we saw the large number of parameters, the undetected Higgs particle (that will be discussed further later on), and the replica of the fermions in three families should be added to the list. But the fatal failure seems to be the following one: Neutrinos are predicted to be massless, while experiments operated in underground sites suggest that they are massive . This rigidity of the SM is due to the global symmetries it has<sup>12</sup><sup>12</sup>12To be fair, one must remind that “quantum anomalies” exist, implying that $`B+L`$ symmetry is broken in the SM ($`B+L`$ is not a gauge symmetry, however!). This breaking is manifest when the symmetry SU(2)<sub>L</sub> gets restored, and this has interesting cosmological consequences (e.g., a pre-existing leptonic asymmetry can be converted into the observed baryonic asymmetry). A more detailed discussion is in the contribution of Auriemma.: A strict conservation of the total number of quarks (equivalent to the “baryon number” $`B`$); and also of the number of leptons of each family (conservation of the “family lepton numbers” $`L_{\mathrm{e},\mu ,\tau },`$ the sum being called the lepton number $`L`$). In the language of Feynman diagrams, this means that quark lines are not turned into lepton lines (or viceversa); and that the arrows in the fermionic lines never clash. The problem of massive neutrinos is very urgent in our view, and will be discussed further in the following.
However, another very evident (and, perhaps, deeper) limitation of the standard model of elementary particles is that gravity is not included. This has to do with the fact that quantising gravity turned out to be a very difficult problem. Anyhow, it is rather reasonable to think that the model of elementary particle would look rather different at the Planck scale $`(\mathrm{}c/G_{Newton})^{1/2}1.2\times 10^{19}`$ GeV. There are ideas on how to search for manifestations of quantum gravity by using ultra-high energy cosmic rays<sup>13</sup><sup>13</sup>13This was discussed by Grillo in this Conference, in connection with the absence of a cutoff in cosmic ray proton spectrum above $`E>5\times 10^{19}`$ eV, which could be related with the $`\gamma `$’s with $`E>10^{13}`$ eV coming from cosmological distances. See also the contributions of Blasi, Petrukhin and Stanev. Note that the center of mass energies ($`\sqrt{s}=(2M_pE)^{1/2}`$) for cosmic ray proton-proton collisions are 2 TeV at the “knee”, 70 TeV at the “ankle” and 200 TeV for UHECR’s: All three exceed the energies of present accelerators.; also, one imagines that a proper understanding of black holes would require to make further steps toward a full theory of gravity. The concerns regarding the status of the cosmological constant stem from similar considerations.
## 2 Modifying the standard model
In the following, we focus on massive neutrinos , in consideration of the urgent character of the problem they pose for the SM, and also because certain reasonable answers can be offered. After a first harsh (“phenomenological”) approach, we will consider more refined and satisfying (“theoretical”) answers; ideas like existence of “right-handed” neutrinos, “quark-lepton symmetry” and “grand unification” will be introduced. At the end, we will come back to Higgs particle and to related arguments for supersymmetry at the weak scale.
### 2.1 A harsh way to massive neutrinos
It is not possible to introduce masses of neutrinos in the SM. However, it is possible to write neutrino masses with 2-spinors only; instead than the usual structure $`q_Rq_L,`$ one can use $`\overline{\nu _L}\nu _L,`$ in the sense that the role of the right state can be played by the anti-neutrino: This is what is called Majorana mass. (For reappraisal, one sees that this step is forbidden, unless one admits that the symmetry U(1)<sub>Y</sub> is violated, since particles and anti-particles have opposite charge—with Feynman diagrams, one says that this mass term produces clashing arrows). Here, we postpone the problem of theoretical justification, and simply assume the existence of Majorana neutrino mass terms. These can be arranged in a mass matrix $`,`$ which is symmetric, and thence can be decomposed as follows:
$$_{\mathrm{}\mathrm{}^{}}=\underset{j=1,2,3}{}U_\mathrm{}j^{}m_j\mathrm{exp}(i\xi _j)U_\mathrm{}^{}j^{}\mathrm{},\mathrm{}^{}=\mathrm{e},\mu ,\tau $$
(7)
$`m_j`$ are the three neutrino masses; $`U_\mathrm{}j`$ is the MNS (after Maki, Nakagawa and Sakata ) unitary mixing matrix, analogous to the CKM matrix (and with the same number of physical parameters); and $`\xi _j`$ are the so-called “Majorana” phases (two of them having physical meaning). All in all, we have 9 new physical parameters.
#### 2.1.1 What do we know on massive neutrinos?
Now we have to face the question; What is actually known from neutrino experiments? As reviewed by Kajita and Stanev in this conference, there is evidence that neutrinos oscillate, as suggested by Pontecorvo . The relevant experiments can be conceptualised in three main steps; the production of a neutrino state, its propagation, and finally its detection. The states produced and detected are the “interaction eigenstates”, that however need not coincide with the “mass”, or more in general with the “propagation eigenstates”. In more precise terms, a neutrino generated by weak interactions (say, a $`\nu _\mu ,`$ tagged by a $`\mu `$ produced in association) is supposed to be a superposition of propagation eigenstates, $`\nu _{1,2,3}`$; if the masses of neutrinos are different, these states propagate differently; thence, a subsequent detector could reveal a disappearance of the original neutrino $`\nu _\mu `$ (or the appearance of a new type neutrino, say $`\nu _\tau `$). This is described by probabilities of survival (or of conversion) whose entities depend on the neutrino mixing and the differences of masses squared. So: (1) atmospheric neutrinos inform us on one squared mass difference and one mixing angle (these results can be regarded as indication of $`\nu _\mu `$ disappearance; Super-Kamiokande gives also indications of $`\nu _\tau `$ appearance via neutral currents events); (2) solar neutrinos inform us on another (squared) mass difference and another mixing angle<sup>14</sup><sup>14</sup>14In fact, more than one region of the parameter space is compatible with present information. This indication of massive neutrinos will largely deserve further studies and attentions. (this results can be regarded as $`\nu _\mathrm{e}`$ disappearance; neutral current signals can be seen by Super-Kamiokande and SNO detectors together). So, there are 5 parameters that are simply unknown at present. They are<sup>15</sup><sup>15</sup>15In this context, the LSND indication cannot be explained. So, this might open further space for surprises.
1. the mixing between $`\nu _\mathrm{e}`$ and the state responsible of atmospheric oscillations, $`|U_{\mathrm{e3}}|;`$
2. a phase that can cause CP violating oscillations (included in $`U_\mathrm{}i`$);
3. the mass of the lightest neutrino;
4. the Majorana phases $`\xi _j.`$
We have plenty of choice of what to study next! However, difficulties are not lacking, too…
#### 2.1.2 Can we measure something more?
Let us discuss the parameters of the previous list.
(1) The first parameter is bound by the CHOOZ experiment to be small, $`|U_{\mathrm{e3}}|^2<2.5\times 10^2`$ (actually, the bound depends to a certain extent on the mass differences). In fact, there are some theoretical arguments<sup>16</sup><sup>16</sup>16Theoretical expectations are not misleading, if considered in the proper manner; however, we feel necessary to stress at this point the provisional character of these expectations. suggesting that this parameter is even smaller, at the level of $`|U_{\mathrm{e3}}|^210^3;`$ these will be discussed in section 2.2.3. So, the effects on oscillation probabilities could be quite tiny and would be not possibly found even after refining cosmic ray neutrino experiments. Actually, it seems rather difficult that present generation of long-baseline experiments will achieve sensitivity to such a small value (compare with the report of Scapparone for this Conference). A completely different possibilities is offered by the detection of a type-II supernova, since the neutrino burst are substantially affected by the presence of even such a small value of $`|U_{\mathrm{e3}}|`$ .
(2) As for the CP violating phase, it is rather difficult to foresee the perspectives of measurement at the present stage. It seems to us that success could be possible only if the solar neutrino problem has relatively large values of the mass difference (“large angle Mikheyev-Smirnov-Wolfenstein” solution –see contribution of Kajita); this possibility could be tested by the KamLAND experiment which aims at detecting neutrinos from distant reactors.
(3) The most sensitive tool to the lightest neutrino mass is the search of anomalies in the endpoint spectra of $`\beta `$ (=electron) emitters with low $`Q`$-value, since in that kinematical regime the mass of the neutrino emitted in association becomes relevant. The experimental searches (Troitsk, Mainz) bound the mass of this neutrino to be below the few eV range (theoretically, it is not excluded that the actual value is close to this one; in that case, however, oscillation experiments would force the conclusion that the spectrum is almost degenerate, which does not sound particularly appealing). Note that close type-II supernovas could give some information on this parameter, too. Also, cosmological considerations give valuable information on massive neutrinos as sub-dominant components of the Universe (assuming of course that the main components and the dynamics are well understood), since the big-bang picture suggests the existence of a sea of relic neutrinos, a close relative of the cosmic microwave (photon) background. In passing, we have to say that it is very hard to conceive other (non-gravitational) tests of the relic neutrinos hypothesis.
(4) The Majorana phases are very-very difficult to measure directly. Actually, the only handle of which we are aware is a rather indirect one, namely a measurement of these parameters in combination with other ones, which could be possible observing the neutrinoless-$`2\beta `$ decay (see fig. 5). In this transition, a nucleus increases its charge by two unities, by emitting two electrons simultaneously. If detected, this transition could inform us on the size of $`|_{\mathrm{ee}}|^2,`$ which depends also on the Majorana phases. The uncertainties in the description of nuclear effects could limit quantitatively the precision of our inferences; however, a detection of this transition would be of great significance, since it would strongly suggest that neutrino have Majorana type mass: oscillation experiments cannot help do this. The present limit on $`|_{\mathrm{ee}}|`$ has been obtained by the Heidelberg-Moscow Collaboration and is in the $`0.11`$ eV range.
However, even in the optimistic assumption all the unknown quantities we discussed will be measured, it is evident that one of the parameters of the neutrino mass matrix will remain unmeasured ($`54=1`$).
### 2.2 A kind way to massive neutrinos
Now, we discuss some (theoretically) respectable extensions of the standard model, with massive neutrinos and in general with new physics.
#### 2.2.1 Adding new particles
We start considering the Higgs boson $`H`$ and the left-leptons $`l_L`$, that have the same gauge numbers, $`|1,2,1/2|`$ (notation as in table 1). An admitted interaction must be invariant, so the product $`l_LH^{}`$ seems to be OK. But it is the space-time structure that is not OK, since combining a spinor and a scalar we do not obtain an invariant. One simple solution is to introduce a new fermion $`\nu _R,`$ with no gauge numbers, and construct $`l_LH^{}\overline{\nu _R}.`$ The answer we get is rather obvious<sup>17</sup><sup>17</sup>17This procedure, however, is of wide validity . If for instance we wonder which particle couples a lepton and a quark, we have just to multiply two such fermions, and deduce from gauge invariance the properties of this hypothetical particle. Also, we can list the effective interactions that are gauge invariant, which do not satisfy the requisite of renormalizability (“higher dimensional” operators).: we just introduced something that must be called “right-handed” neutrino, and we obtained a Yukawa interaction completely analogous with quark or charged leptons Yukawa terms (this term alone would originate what is called a Dirac mass for neutrinos). But we learned that $`\nu _R`$ has no gauge interactions, so for consistence we must consider the invariant $`\nu _R\nu _R,`$ too (=Majorana mass for the right-handed neutrino). These two elements give rise to the famous “see-saw” model for the masses of the left-neutrinos (which is Majorana in type; the name comes from the fact that, the more the right-handed neutrino masses go up, the more the left-neutrino masses go down): see fig. 6.
It should be noted that, even if the right-handed neutrino has super-heavy mass and cannot be produced at accelerators, it still leaves its footprints in the masses of left-neutrinos. The generated mass can be regarded effectively as a $`(l_LH^{})^2`$ term; it reveals its non-fundamental nature from the fact that it is not renormalizable (one can draw here a very close analogy with the 4 fermion weak interactions). Note that the number of parameters is still enlarged! However, some of these parameters might have some interesting manifestations, for instance they can generate a leptonic asymmetry in the early Universe .
There is at least one alternative possibility that we feel should be mentioned: if, in order to obtain Majorana masses for neutrinos, we directly multiply two leptonic doublets, we find the quantum numbers $`|1,3+1,1|.`$ Thence, by postulating a triplet scalar $`\mathrm{\Delta }`$ with U(1)<sub>Y</sub> charge $`1,`$ we can construct the invariant term $`l_Ll_L\mathrm{\Delta };`$ and if we give a tiny (see eq. 5 and discussion therein) vacuum expectation value to the neutral component of the triplet, we find again Majorana masses. So, the question arises, which of these (or other) mechanism for massive neutrinos is the correct one? This is very difficult to answer convincingly, and it could be considered a frontier of the “theory beyond the SM”!
#### 2.2.2 Larger gauge groups
We proceed in the presentation of promising ideas, and pass to the concept of grand unification . The existence of right-handed neutrinos–whose interest has been already discussed–can be argued on the basis of quark-lepton symmetric spectrum (even if the Majorana mass of the right-handed neutrinos is by itself a point of asymmetry). This hypothesis becomes compulsory if we assume that the SM is a group of residual symmetry of certain “unification groups.” We just present few examples and limit ourselves to illustrate how the matter fermions (and the right-handed neutrinos) fit in the representations of larger groups:
SU(5). (Georgi-Glashow) This group of $`5\times 5`$ matrices obviously includes SU(3)<sub>c</sub> and SU(2)<sub>L</sub> as the upper $`3\times 3`$ and lower $`2\times 2`$ blocks, while U(1)<sub>Y</sub> is in the diagonal. The representations where the matter fermions sit are:
$$\mathrm{𝟏𝟎}=\left|\begin{array}{ccccc}0& \overline{u_{R3}}& \overline{u_{R2}}& u_{L1}& d_{L1}\\ & 0& \overline{u_{R1}}& u_{L2}& d_{L2}\\ & & 0& u_{L3}& d_{L3}\\ & & & 0& \overline{\mathrm{e}_R}\\ & & & & 0\end{array}\right|\overline{\mathrm{𝟓}}=\left|\begin{array}{c}\overline{d_{R1}}\\ \overline{d_{R2}}\\ \overline{d_{R3}}\\ \nu _L\\ \mathrm{e}_L\end{array}\right|\mathbf{1}=\nu _R$$
(8)
(1,2,3 are the index of SU(3)$`_c;`$ the 10 matrix is antisymmetric). In this context, the right-handed neutrino has the same status as in the SM.
SU(4)$`\times `$SU(2)$`{}_{L}{}^{}\times `$SU(2)<sub>R</sub> (Pati-Salam). This group is very satisfying as for the quark-lepton symmetry. The fermions are assigned to:
$$|\mathrm{𝟒},\mathrm{𝟐},\mathrm{𝟏}|=\left|\begin{array}{cccc}u_{L1}& u_{L2}& u_{L3}& \nu _L\\ d_{L1}& d_{L2}& d_{L3}& \mathrm{e}_L\end{array}\right||\mathrm{𝟒},\mathrm{𝟏},\mathrm{𝟐}|=\left|\begin{array}{cccc}u_{R1}& u_{R2}& u_{R3}& \nu _R\\ d_{R1}& d_{R2}& d_{R3}& \mathrm{e}_R\end{array}\right|$$
(9)
(note the presence of right-handed neutrinos). Leptons, in a sense, are just the fourth type of quarks. For us is difficult to see this, and maintain the opinion that it has no meaning. One could consider a weak point of this assumption the fact that there are three different (non-unified) gauge subgroups (however, by imposing a parity that relates $`L`$\- and $`R`$-subgroups this improves).
SO(10) (Georgi, Fritzsch-Minkowski). This group includes SU(5) (since a complex 5 vector can be obviously mapped in a real 10 vector) but also the Pati-Salam group (since SO(6)$``$SU(4)). Each family of matter fermions sits in a single representation:
$$\mathrm{𝟏𝟔}=\mathrm{𝟏𝟎}+\overline{\mathrm{𝟓}}+\mathrm{𝟏}=|\mathrm{𝟒},\mathrm{𝟐},\mathrm{𝟏}|+|\mathrm{𝟒},\mathrm{𝟏},\mathrm{𝟐}|$$
(10)
Like SU(5), this group has an unique coupling. However, the breaking of SO(10) down to the SM might take place in several steps (with several “intermediate” scales).
Whatever the gauge group, it is needed that the large gauge symmetries are broken (and also that the fermion masses are generated). This calls for scalars, and opens up many possibilities (and troubles, see section 2.3). We mention only one specific point here . As we saw, the SO(10) or Pati-Salam group naturally includes right-handed neutrinos, which is certainly a good thing. However, the reduction of SU(2)<sub>R</sub> symmetry typically requires the existence of scalar “right” triplets; the symmetry forces the existence of “left” triplets; so that, there are two competing sources for massive neutrinos: see-saw and “left” triplet.
Apart from neutrino masses, the unification groups have (some) predictivity on the unification of gauge couplings, of fermion masses, and also on proton decay. From the experimental point of view, the last aspect is surely the most interesting. Proton decay could be due to the new gauge bosons of the grand unified group, that permit communication between quarks and leptons; however there are also other possibilities, e.g. new scalar particles might have an important role (this happens commonly in supersymmetric models). Trying to summarise in a few words the present situation: proton decay is not experimentally found (strongest limits come from the Super-Kamiokande experiment) but there are some theoretical models that offer hopes of detection for future detectors (new detectors like ICARUS have superior properties, but its mass might limit its discovery potential … should a Mega-Kamiokande be built?). Experimental and theoretical proposals, however, are still at a stage of discussion; the “future” seems not close.
#### 2.2.3 Use of quark-lepton symmetry
Here, we present an ansatz for massive neutrinos, that we consider quite reasonable since it is based on a principle: quark-lepton symmetry (or more precisely, quark-lepton correspondence). One starts to note that the masses of up-type quarks increase strongly changing family, in comparison with what happens for down type quarks, or also for charged leptons. This difference of hierarchies motivate the assumption that neutrinos have still weaker differences among them. An actual implementation of this idea, due to Sato and Yanagida<sup>18</sup><sup>18</sup>18This model has been inspired by SU(5), since the weaker hierarchy was explained by saying that the $`\overline{\mathrm{𝟓}}`$-plets of second and third families (that contain $`\mu `$ and $`\tau `$ neutrinos) do not pay-off hierarchy factors; while this always happen for $`\mathrm{𝟏𝟎}`$-plets. , is the assumption that the neutrino mass has the structure:
$$\left|\begin{array}{ccc}\epsilon ^2& \epsilon & \epsilon \\ & 1& 1\\ & & 1\end{array}\right|\epsilon =1/20\mathrm{sin}^2\theta _Cm_\mu /m_\tau $$
(11)
(where the Cabibbo angle $`\theta _C`$ is a parameter of the CKM matrix). To make the statement sufficiently vague (or equivalently, sufficiently precise) one postulates that the elements of the matrix include also coefficients order unity (these are in fact are essential in order to generate three different neutrino masses).
The weak points of this assumption are that:
(1) the overall scale is not predicted;
(2) the hierarchy of masses between the two heavier neutrinos tends to be rather weak.
The advantages (after the weak points are made up, by adjusting the unknown “coefficients” and the overall scale) are that:
(1) a large mixing for atmospheric neutrinos is automatic;
(2) there is a prediction of “large angle Mikheyev-Smirnov-Wolfenstein” solution (with the correct type of mass hierarchy);
(3) the value of $`U_{\mathrm{e3}}`$ is predicted to be small.
Note, incidentally, that also neutrinoless $`2\beta `$ transitions are predicted to be suppressed, since $`|_{\mathrm{ee}}|^2\mathrm{\Delta }m_{atm}^2\times ϵ^4.`$ This is essentially a manifestation of the fact that neutrinos are supposed to obey a sort of family hierarchy, and is tightly related to the suppression of $`U_{\mathrm{e3}},`$ since $`U_{\mathrm{e3}}ϵ.`$
### 2.3 Troubles with fundamental scalars, and supersymmetry as solution
Here we come to one trouble of the SM (even more severe in the unified theories). This is, in essence, a problem of hierarchy of scales . One can say that, due to quantum fluctuations, heavy masses creep in the Higgs boson ($``$ weak boson) scale and want to destabilise it. For instance, the diagram in the fig. 7 changes the coefficient of the bilinear term $`\mu ^2\times |H|^2`$ by an amount of the order of
$$\delta \mu ^2\frac{Y^2}{(4\pi )^2}\times M_{\nu _R}^2\text{ where }\{\begin{array}{c}Y=\text{Yukawa}\text{ coupling of neutrinos,}\hfill \\ M_{\nu _R}=\text{Mass of right-neutrinos.}\hfill \end{array}$$
(12)
which (apart for the loop pre-factor) seems to produce a huge scale, comparable to $`M_{\nu _R}`$ (not of the order of the electroweak scale as we need). To be picky, one can say that the individual contributions to $`\mu ^2`$ are not separately measured; that only their sum has physical meaning; and that we should not ask the theory to say more than it can. However, in this manner we would most probably give up any chance of predicting these fundamental parameters, since a hypothetical (more complete) theory would be forced to explain a very precise fine-tuning<sup>19</sup><sup>19</sup>19Note the purely theoretical character of this problem; in this sense, one could say that there is a disturbing situation with fundamental scalars, but not an untolerable one.. This situation motivated several extensions of the SM, and all of them have new physics (close) at the electroweak scale; for instance, it was postulated that the Higgs particle is not fundamental—but instead a pion-like object (technicolor).
We will concentrate the discussion on supersymmetric models . Supersymmetry is an extension of the space-time group, which relates fermions and bosons by a symmetry transformation. In the models that could be possibly relevant for electroweak scale physics, supersymmetry commutes (=is unrelated) with the gauge group; so that any ordinary particle<sup>20</sup><sup>20</sup>20In principle, the lepton doublet and one of the Higgs bosons could be “partners”; in practise, this type of model would have too large violations of the leptonic number (and other troubles) and for this reason is not pursued. obtains a “partner”, and we have scalar electrons (sleptons), fermionic gluon (gluino), etc. Actually, it is necessary that the number of Higgs doublets is at least two<sup>21</sup><sup>21</sup>21A problem that we cannot discuss here is: How the additional scalars of the model are prevented from obtaining a vacuum expectation value? We have to limit ourselves to say that for some value of the parameters it can be done.. We recommend to make reference to the contributions of Ganis and Denegri for a more detailed description of these models, and discussion of the perspectives of confirmation. The connection with the “hierarchy problem” is due to an amazing property of supersymmetry as a quantum field theory: that the “loops” involving bosons and fermions compensate each other, and contributions like those in eq. 12 do not arise. At this point, however, we have to recall that an extension of the SM should have broken supersymmetry in order to be realistic (otherwise, for instance, the “partners” would have the same mass of the ordinary particles). The actual mechanism for supersymmetry breaking is an open question at present, which surely is not a nice feature, even if there are reasonable proposals. However, if one assumes that the breaking scale is order of the TeV, and restricts the allowed breaking terms to so-called “soft-breaking”, the quantum properties are maintained and the hierarchy problem is under control<sup>22</sup><sup>22</sup>22Despite the desire to mantain as much predictivity as we can, we are forced to introduce new parameters in order to do this (for a precise counting, see for instance the review on supersymmetry in ).. In this paragraph, we presented an instant summary of the idea of “supersymmetry at the weak scale”. Much more could be said, however, the key question that we have to answer is: Do these considerations have any relevance to the description of Nature? Let us list three considerations that suggest (but not “imply”) an affirmative answer:
* Electroweak precision measurements suggest the existence of a light Higgs particle as predicted in (minimal) supersymmetric extensions of the standard model.
The prediction is mainly due to the fact that the Higgs boson self-coupling $`\lambda `$ in eq. 3 turns out to be a combination of the (measured) electroweak gauge constants (the scalar potential is quite constrained in these models). This prediction can be tested at future colliders (… or if we are really lucky, even before; see Denegri, these Proceedings).
* Gauge couplings unify in the context of low energy supersymmetric model.
Here we mean that: the extrapolation of the couplings is compatible with the hypothesis of grand unified dynamics (broken below a certain scale $`M_{GU},`$ see fig. 8) . This does not happen in the ordinary SM (not within a model with a single scale of breaking). It is rather remarkable that the large unification scale, $`M_{GU}2\times 10^{16}`$ GeV, is comparable with what is suggested by a see-saw mechanism for massive neutrinos. Can proton decay provide the crucial confirmation of this indication? Specific channels exist , like
$$pK^+\overline{\nu };$$
(13)
the presence of a kaon (the characteristic aspect) results from the fact that proton decay is expected to be due to Yukawa type interactions, and to observe for this reason the family hierarchy (see fig. 9). However, the dependence on the unknown aspects of the model is strong (and no doubt that the usual Yukawa sector already challenges our understanding), and, theoretically, it is not excluded that the proton decay process is rather suppressed.
* A cold dark matter candidate exists.
The lightest supersymmetric particle (LSP) is stable due a discrete symmetry that can be incorporated in the model (which, for the record, is called R parity<sup>23</sup><sup>23</sup>23For completeness, we must add that theoretical models can be constructed in which this symmetry is broken; the cold dark matter candidate disappears, but such a breaking could account for massive neutrinos. At present, however, this possibility is not considered of particular appeal, due to the rather ad-hoc values of the parameters that are required to account for the masses of the neutrinos.). Is the galactic “LSP cloud” responsible of the modulated signal seen in the DAMA experiment shown by Incichitti at this Conference? If this were true, this result would be of enormous importance, not only for the direct detection of dark matter but also as a first signal of “supersymmetry at the electroweak scale”; and it would also open quite interesting perspectives for future collider searches, as discussed by Ganis. Further studies and confirmations are of essential importance (Incichitti).
We would like to add a comment on the Higgs boson mass. A value on the large side (say, $`>120`$ GeV) would indicate in a supersymmetric context a rather strong hierarchy between the vacuum expectation values of the two Higgs doublets, $`H_{up}^0H_{down}^0.`$ This would suggest an entire series of theoretical and phenomenological questions; for example, the (Yukawa type) proton decay is expected to be enhanced in this regime. Instead, if the Higgs boson mass turns out to be really large (say, $`>150`$ GeV) it seems not easy to avoid the conclusion that “supersymmetry at the electroweak scale” is in trouble; this will be the crucial test of the model. Finally, we remind that in the SM there is a limit on the Higgs boson mass suggested by the consideration that the self-coupling $`\lambda `$ should be not driven negative as an effect of the quantum fluctuations, (“vacuum stability” ) at least up to the Planck scale where new effects most probably appear. It is rather funny, but this lower limit almost coincides with the upper limit in the (minimal) supersymmetric extension of the standard model. So (if a joke is permitted) we present a prediction for LHC:
$$m_H=135\pm 5\text{ GeV}$$
(14)
the reason is that this value will increase the entropy in the minds of several theorists. Note, however, that the decay of a standard and supersymmetric Higgs particles with the same mass (or also the production rate–“cross-sections”) could be rather different; thence, these measurements would offer a possibility to distinguish between the SM and its supersymmetric extension even in this tricky case.
## 3 (Not quite a) conclusion
We would like to close this pages by spending few words of caution, to remind that failures of the standard model have been often claimed in the past years (today, several of them are considered dubious or simply wrong tracks). Here is an arbitrary selection:
$$\begin{array}{ccc}\text{THEORETICAL}\hfill & & \text{EXPERIMENTAL}\hfill \\ \text{INTERPRETATION}\hfill & & \text{ANOMALY}\hfill \\ & & \\ \text{leptoquark}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \text{High }x\text{ and }Q^2\text{ events at }\text{HERA}\hfill \\ \text{compositeness}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \text{Excess of 4-jet events at }\text{ALEPH}\hfill \\ \text{light gravitino}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \mathrm{ee}\gamma \gamma E/\text{ event at }\text{CDF}\hfill \\ \text{17 keV neutrino}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \text{bump in }\beta \text{ spectra (}\text{Simpson, …}\text{)}\hfill \\ \text{monopole}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \text{induced currents (}\text{Cabrera}\text{)}\hfill \\ \text{proton decay}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \text{contained multitrack events at }\text{KGF}\hfill \\ \text{}\hfill & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..& \text{}\hfill \end{array}$$
Is there any moral behind these stories? Maybe not; however:
1) they suggest to go slowly and carefully from data to theories and back (because of possible pitfalls of interpretation, of suggestion, etc.);
2) they witness how hard is to reach the frontiers of standard model; and, also, how strong is the desire of particle physicists to find them!
I am grateful to the Organizers and Partecipants (in particular to B. Alessandro, R. Bandiera, P. Blasi, S. Colafrancesco, F. Giovannelli, A. Grillo, G. Mannocchi, R. Ramelli, P.G. Rancoita, E. Scapparone, A. Stamerra, A. Surdo) for the most pleasant and informative discussions, and to F. Cavanna for a careful reading of the manuscript. I would like to take this occasion to thank R. Barbieri, V. Berezinksy, S. Bertolini, W. Buchmüller, A. Di Giacomo, A. Masiero, N. Paver, S. Petcov, E. Roulet, G. Senjanović, A.Yu. Smirnov, M. Veltman and T. Yanagida to whom I owe what I know on the SM and its extensions, and who largely deserve the credit for niceties in the presentation; errors and misinterpretations of course are mine.
|
warning/0007/hep-th0007044.html
|
ar5iv
|
text
|
# Contents
## 1 Introduction
Usually, supersymmetry is associated with non-negative energy. In asymptotically flat spaces this is related to the time translation operator which is a square of fermionic operators $`E=Q^20`$. In curved space the positivity of energy theorem is proved via Nestor–Israel–Witten construction where it is usually assumed that the space is non-singular. In case of black holes with the singularity covered by the horizon, the argument about the positivity of energy was also extended .
Recently there were number of reasons to reconsider the issues of supersymmetry, in general. In the brane world scenarios with fine tuning where singular branes are introduced , the status of the supersymmetric embedding was not clear. The first attempts to relax the fine tuning was to find smooth solutions of supergravity where domain walls are build from some scalar fields . With respect to the ‘alternative to compactification scenario’ , no-go theorems were established for the BPS smooth solutions of a certain class of supergravity theories ($`N=2`$ with vector and tensor multiplets ) on the basis of the negative definiteness of the derivative of the $`\beta `$-function in the relevant renormalization group behavior (UV behavior).
More recently, the complete $`N=2`$ supergravity theory was constructed in where also the hypermultiplets are included. The status of hypermultiplets in the adS brane world is not yet settled<sup>1</sup><sup>1</sup>1J. Louis, talk at SUSY2K, July 2000., more work is required. Even if the solutions with the IR fixed point will be found, there will still be a problem to find a configuration in a smooth supergravity which will connect two such IR fixed points. This may be also explained following a suggestive argument<sup>2</sup><sup>2</sup>2K. Stelle, talks at Fradkin and Gürsey conferences, June 2000.: BPS solutions are expected to be given by harmonic functions, in codimension 1 they must have a kink. A kink can not appear as a smooth solution. The complimentary argument was given in : the fermion mass does not change sign when passing through the wall, which may explain the absence of smooth solutions with the required properties. The version of a no-go theorem for smooth brane worlds from compactifications, under some assumptions about the potential, was recently proposed in . None of these arguments seems to give an unconditional final proof that there are no smooth domain walls for a brane world scenario. However, they suggest to find out whether the clear framework is available for the non-smooth supersymmetric solutions.
Thus the purpose of this paper is to generalize supersymmetry for singular spaces, where the curvature may have some $`\delta `$-function singularities. In singular spaces there was no consistent and complete definition of supersymmetry so far<sup>3</sup><sup>3</sup>3We will comment on the available information below. and the related issue of non-negative energy was not clearly addressed.
An important reason to reformulate supersymmetry in singular spaces is related to supersymmetric domain walls: objects of codimension $`d=1`$, i.e. $`(D2)`$-branes in $`D`$-dimensional space. They may be associated with $`D`$-form fluxes which are dual to a scalar and piecewise constant. Such forms do not fall off with distance. In an infinite volume, they may lead to an infinite energy and would be unphysical. Therefore, objects like the $`8`$-brane in $`D=10`$ and 3-branes in $`D=5`$ may not exist as independent objects. They may need some planes that serve as sinks for the fluxes for supersymmetric configurations of codimension 1 . Note that the fluxes in the branes of higher codimension do vanish at infinity and this problem can be avoided.
Usually local supersymmetry is realized in supergravity when the Lagrangian is integrated over a continuous space. Under supersymmetry, the Lagrangian transforms as a total derivative. The parameters of local supersymmetry are assumed to fall off at infinity and therefore the action is invariant. If in addition to the supergravity action one considers the $`\kappa `$-symmetric worldvolume action of the positive tension brane of codimension $`d3`$, it provides the $`\delta `$-functional sources for the harmonic functions that describe the configuration, $`_i_iH=T\delta ^d(\stackrel{}{y})`$.
In the case of branes of codimension $`d=1`$ defined on an orbifold $`S^1/_2`$, the harmonic functions satisfy the equations $`_y_yH=T\delta (y)T\delta (y\stackrel{~}{y})`$. This can be seen e.g. in the Hor̆ava–Witten (HW) construction developed for 5-dimensional 3-branes in . The metric of such solutions depends on $`H=c+T|y|`$ and has two kinks at the orbifold fixed points: one at $`y=0`$ and the other at $`y=\stackrel{~}{y}`$ where $`\stackrel{~}{y}`$ is identified with $`\stackrel{~}{y}`$. The supersymmetry in HW construction includes the contribution from anomalies and requires quantum consistency. It has some problematic features related to higher order corrections and quartic fermionic terms.
To set up a new general point of view on supersymmetry in singular spaces, and to clarify the issue of energy of brane worlds, we introduce here a set of rules defining consistent supersymmetry on singular spaces.
The most important new steps include: i) some constants (masses or gauge couplings) have to become “$`_2`$-odd” to make supersymmetry commuting with $`_2`$-symmetry. This can be achieved by promoting such a constant to the status of a supersymmetry singlet field, as suggested in ; ii) moreover, one has to add to the theory the $`(D1)`$-form potential<sup>4</sup><sup>4</sup>4The importance of the $`(D1)`$-form field was realized by Duff and van Nieuwenhuizen , who pointed out 20 years ago the quantum inequivalence of the theories with and without such field in the context of trace anomalies and 1-loop counterterms in topologically non-trivial backgrounds. At about the same time Aurilia, Nicolai and Townsend have found that the $`(D1)`$-form, which propagates no physical particles, carries a surprising physics. They have looked at $`\theta `$ parameter in QCD and at the cosmological constant in supergravity. to compensate the variation of the Lagrangian proportional to the derivative of the new field; iii) one can find afterwards the new bulk and the brane actions, which are separately invariant under supersymmetry.
The unusual features of the new supersymmetry are the presence of the supersymmetry singlet field and $`(D1)`$-form field in the bulk, and the fact that the purely bosonic action on the brane is supersymmetric due to the fact that its fermionic partner is a $`_2`$-odd fermion which vanishes on the brane.
In absence of brane actions, the new fields become irrelevant: on shell for the 4-form and supersymmetry singlet, the bulk action reduces to the standard supergravity action supersymmetric under the standard rules.
In general, when the brane actions are added, the new fields play an important role in understanding the energy issue. We will find that the total energy of supersymmetric configurations vanishes locally at each brane. The positive (negative) energy of the brane tensions are compensated separately by the terms with the derivative of the supersymmetry singlet field. The energy of any static $`y`$-dependent bosonic configuration vanishes locally, $`E=0`$, in analogy with the vanishing of the Hamiltonian in a closed universe<sup>5</sup><sup>5</sup>5We are grateful to A. Linde who suggested this analogy..
The strategy is applied in detail in the particular case of a 3-brane in $`D=5`$ on the basis of $`U(1)`$ gauged $`N=2`$ supergravity interacting with abelian vector multiplets. We expect that it will work in other cases as well: in $`D=5`$ one can try to include more general gauging and tensor and hypermultiplets. The gauged $`N=4`$ and $`N=8`$ supergravities in $`D=5`$ are also natural candidates for an analogous extension. We will give a brief discussion of the particularly interesting case of the 8-brane in $`D=10`$.
The paper has two main parts.
In Part I, The Supersymmetric Theory: Bosons And Fermions, we construct the supersymmetric actions in the bulk and in the brane. To do so, we first identify the $`_2`$ operator in section 2. Then, we construct the supersymmetric theory with three steps in section 3. The final result is written down in section 4
In Part II, The Background: Vanishing Fermions, Bosons Solve Equation of Motion, we study bosonic solutions of the theory from part I and their unbroken supersymmetries. Section 5 discusses the vanishing of the energy. The BPS equations and preserved supersymmetries in these singular spaces are discussed in section 6. We consider two cases, the fixed scalars, with doubling of supersymmetry, and the general stabilization equations in very special geometry, with 1/2 preserved supersymmetry. We show the resulting formulas for the example of the STU wall, and for a particular Calabi–Yau wall in section 7.
Finally, in Part III, we discuss the similar mechanism for the 8-brane in 10 dimensions.
The notations and use of indices are presented in appendix A. For those unfamiliar with the 5-dimensional supergravity, and for establishing the related notations, we present its structure in appendix B. In appendix C we discuss the previous attempts to define supersymmetry on orbifolds.
PART I. The Supersymmetric Theory: Bosons And Fermions
## 2 Local supersymmetry and $`_2`$
Our setup has the following basic features.
1. We start with the standard action $`S_{bulk}`$ of five-dimensional gauged supergravity with a symplectic Majorana supersymmetry with parameters $`ϵ^i`$, $`i=1,2`$, thus having 8 real components, coupled with an arbitrary number of vector multiplets .
2. *Orbifold construction*. Fields live on a circle in the direction $`x^5`$, with an orbifold condition. The circle implies that $`\mathrm{\Phi }(x^5)=\mathrm{\Phi }(x^5+2\stackrel{~}{x}^5)`$, where $`\stackrel{~}{x}^5`$ is some arbitrary parameter setting the length of the circle. We use a concept of parity to split the fields in even and odd under a $`_2`$:
$$\mathrm{\Phi }_{even}(x^5)=\mathrm{\Phi }_{even}(x^5),\mathrm{\Phi }_{odd}(x^5)=\mathrm{\Phi }_{odd}(x^5).$$
(2.1)
This implies that the odd fields vanish at $`x^5=0`$ and at $`x^5=\stackrel{~}{x}^5`$, where we will put the branes. The supersymmetries are also split in half even ($`ϵ_+`$) and half are odd ($`ϵ_{}`$). Both $`ϵ_\pm `$ have 4 real components. The bulk action is even, and all transformation rules are consistent with the assignments.
3. *The brane action* $`S_{brane}`$ is introduced. We place two branes, one at $`x^5=0`$ and another one at $`x_5=\stackrel{~}{x}_5`$. The actions depend on the values of the bulk fields at $`x^5=0`$ and $`x_5=\stackrel{~}{x}_5`$. As the odd fields are zero on the brane, the brane action only depends on the even fields. Only the supersymmetries $`ϵ_+`$ act on these fields.
We analyse all possibilities to make parity assignments consistent with supersymmetry commuting with $`_2`$-symmetry. We conclude that the consistent supersymmetry on orbifolds without brane actions does not fix the $`_2`$-properties of the gauge coupling. However, after adding the brane action we must require the gauge coupling to be $`_2`$-odd<sup>6</sup><sup>6</sup>6The second possibility, $`_2`$-even gauge coupling, as chosen in , will be discussed in Appendix C..
We have to treat the fact that $`g`$ is only piecewise constant. This approach is inspired by . It consists of the following steps
1. Replace in the action the constant $`g`$ by a scalar function $`G(x)`$ and keep this field a supersymmetry singlet. The supersymmetry singlet field was introduced in the context of $`\kappa `$-symmetric brane actions in . This field has some peculiar properties: its shift under translation vanishes on shell but not off shell.
2. Add a Lagrange multiplier $`A_{\mu \nu \rho \sigma }`$ that imposes the constraint $`_\mu G=0`$. At this point this reproduces the known ’bulk’ actions of 5 dimensions . The $`G`$-field dependent action is invariant by appropriately choosing supersymmetry transformations as well as local $`U(1)`$ gauge transformations for $`A_{\mu \nu \rho \sigma }`$.
3. Add source terms that are separately invariant, but may contain $`\delta (x^5)`$ and $`\delta (x^5\stackrel{~}{x}^5)`$ and are dependent on $`A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}`$, where these are the components in the 4 dimensions of the brane. After these additions the $`A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}`$ field equation says that $`_5G(x)\delta (x^5)\delta (x^5\stackrel{~}{x}^5)`$, and thus $`G(x)\epsilon (x^5)`$. This provides a supersymmetric mechanism to change the sign of the coupling constant (and of the fermion mass) when passing through the wall.
We will first consider the possibilities for a $`_2`$ in the 5-dimensional action. Define the operator $`\mathrm{\Pi }`$ for any field, being $`+1`$ or $`1`$ whether it is even or odd under $`x^5x^5`$. For the symplectic Majorana spinors, the splitting might involve some projection operators. Let us look for a $`_2`$ of the form (for any symplectic Majorana spinor $`\lambda ^i`$)
$$\lambda ^i(x^5)=\mathrm{\Pi }(\lambda )\gamma _5M^i{}_{j}{}^{}\lambda _{}^{j}(x^5),$$
(2.2)
where $`\mathrm{\Pi }(\lambda )`$ is thus a number $`+1`$ or $`1`$, while $`M^i_j`$ is so far an undetermined $`2\times 2`$ matrix. If this has to be a $`_2`$, the operation should square to $`\text{ }\text{}`$. Therefore $`M`$ should square to $`\text{ }\text{}`$. Notice that this is independent of whether we included $`\gamma _5`$ in (2.2) or not. Next, this has to be consistent with the reality condition (symplectic Majorana condition, see appendix A). This implies that $`M`$ should satisfy $`M^C\sigma _2M^{}\sigma _2=M`$, or thus, $`M=\mathrm{i}m_0\text{ }\text{}+m_1\sigma _1+m_2\sigma _2+m_3\sigma _3`$ with $`m_0`$, $`m_1`$, $`m_2`$ and $`m_3`$ real. The $`M^2=\text{ }\text{}`$ condition implies that
$$M^i{}_{j}{}^{}=m_1(\sigma _1)^i{}_{j}{}^{}+m_2(\sigma _2)^i{}_{j}{}^{}+m_3(\sigma _3)^i{}_{j}{}^{},\text{with}m_1,m_2,m_3,$$
(2.3)
which means that $`M`$ is hermitian and traceless. If $`\gamma _5`$ were not included in (2.2), the numbers $`m`$ would have to be pure imaginary, and the $`M^2=\text{ }\text{}`$ condition would imply $`M=\text{ }\text{}`$, but then we would have no projection at all. Thus we conclude that (2.2) with (2.3) is the only one that is possible.
Lowering the two indices on $`M`$, it becomes a symmetric matrix:
$$M_{ij}=m_1(\sigma _3)_{ij}\mathrm{i}m_2(\text{ }\text{})_{ij}+m_3(\sigma _1)_{ij}.$$
(2.4)
The parity will be related to the fifth direction. We will therefore assign a negative parity to the $`x^5`$ coordinate, or $`\mathrm{\Pi }(_5)=1`$. Consider now the supersymmetry transformation laws in the bulk, see (B.12). For one of the fermions we may arbitrary assign even parity, e.g. for the components of the gravitino in the directions excluding ‘5’, i.e. $`\psi _{\underset{¯}{\mu }}`$. Consider first the supersymmetry transformation laws that are independent of the gauge coupling $`g`$. The consistency of the parity assignments determines
$`\mathrm{\Pi }(e_{\underset{¯}{\mu }}{}_{}{}^{m})=1,\mathrm{\Pi }(e_5{}_{}{}^{5})=1,\mathrm{\Pi }(A_5^I)=1,`$
$`\mathrm{\Pi }(e_5{}_{}{}^{m})=1,\mathrm{\Pi }(e_{\underset{¯}{\mu }}{}_{}{}^{5})=1,\mathrm{\Pi }(A_{\underset{¯}{\mu }}^I)=\mathrm{\Pi }(\mathrm{\Lambda }^I)=1,`$
$`\mathrm{\Pi }(\psi _{\underset{¯}{\mu }})=\mathrm{\Pi }(ϵ)=1,\mathrm{\Pi }(\psi _5)=\mathrm{\Pi }(\lambda )=1.`$ (2.5)
Note that also the supersymmetry parameters got a parity projection and that the parameters of the gauge transformations, $`\mathrm{\Lambda }^I`$, have to be odd.
Now we consider the terms with $`g`$ in the gravitino transformation law (B.12). They depend on the constant matrix $`Q_{ij}`$, see (B.6). Taking the parity transformation of both sides, we find that they are proportional to
$$M^j{}_{i}{}^{}Q_{}^{k}{}_{j}{}^{}ϵ_{k}^{}=\mathrm{\Pi }(g)Q^j{}_{i}{}^{}M_{}^{k}{}_{j}{}^{}ϵ_{k}^{},$$
(2.6)
where we allowed a parity transformation of $`g`$. We find that if $`Q`$ and $`M`$ commute (which means that they are proportional) one needs $`\mathrm{\Pi }(g)=1`$, while if they anticommute (thus are taken in orthogonal directions in the $`SU(2)`$-space), $`\mathrm{\Pi }(g)=1`$.
Taking an anticommuting $`M`$ and $`Q`$ brings us to the setup of . We will see below, that the addition of brane actions forces us to take a matrix that commutes with $`Q`$. If $`Q`$ and $`M`$ commute, one has two possibilities to implement the parity assignment $`\mathrm{\Pi }(g)=1`$. In the approach of , who take $`M^i{}_{j}{}^{}=(\sigma _3)^i_j`$, this assignment is realized by replacing $`g`$ by $`g\epsilon (x^5)`$. On the other hand, we will be able to make this assignment by promoting $`g`$ to a field.
To continue in the direction of , one has to provide the consistent definition of supersymmetry with step functions and delta functions present in supersymmetry rules and in the action. The construction of the higher order fermionic terms and the structure of the algebra of supersymmetry in singular spaces in these approaches are difficult as the HW theory shows. Instead of this, a new way to introduce supersymmetry in singular spaces will be developed below.
## 3 Supersymmetry in a singular space
### 3.1 Step 1: the bulk action
We consider 5-dimensional supergravity coupled to $`n`$ vector multiplets. The coupling is determined by a 3-index real symmetric constant tensor $`C_{IJK}`$, where the indices run over $`n+1`$ values. We consider a $`U(1)`$ gauging of the $`R`$-symmetry group determined by real constants $`V_I`$. The direction in $`SU(2)`$ space is determined by a matrix $`Q_i^j`$. Obviously, the choice of that direction has no physical consequences. The full action, to which we will refer as the GST action , and transformation rules, are given in appendix B. Replacing the coupling constant by a scalar $`G(x)`$, the action is not any more invariant supersymmetry,
$$e^1\delta (ϵ)_{GST}=\left[\mathrm{i}\frac{3}{2}\overline{\psi }{}_{\mu }{}^{i}\gamma _{}^{\mu \nu }ϵ^jW\overline{\psi }{}_{\mu }{}^{i}\gamma _{}^{\mu \nu \rho }ϵ^jA_\rho ^{(R)}+\frac{3}{2}\overline{\lambda }{}_{x}{}^{i}W_{}^{,x}\gamma ^\nu ϵ^j\right]Q_{ij}_\nu G,$$
(3.1)
and neither under the $`U(1)`$ gauged symmetry,
$$e^1\delta _R_{GST}=\frac{1}{2}\left[\overline{\psi }_\mu ^i\gamma ^{\mu \nu \rho }\psi _\rho ^j+\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\nu }\lambda ^{jx}\right]Q_{ij}\mathrm{\Lambda }_R_\nu G,$$
(3.2)
where $`\mathrm{\Lambda }_R`$ is the parameter of the $`R`$-symmetry
$$\mathrm{\Lambda }_R=V_I\mathrm{\Lambda }^I.$$
(3.3)
### 3.2 Step 2: the four-form
In the second step, we add the following Lagrange multiplier term:
$$S_A=\frac{1}{4!}\mathrm{d}^5x\epsilon ^{\mu \nu \rho \sigma \tau }A_{\mu \nu \rho \sigma }_\tau G.$$
(3.4)
Now we can make the action invariant under supersymmetry. This is obtained 1) by taking $`G`$ invariant under supersymmetry, and 2) by defining the variation of $`A`$ such that all $`_\mu G`$ terms in the transformation of the rest of the action are cancelled. The fact that $`G`$ is invariant is consistent with the algebra because the translation of $`G`$ is a field equation ($`G`$ is a supersymmetry singlet ). Thus the algebra is realized on-shell. The resulting variation of $`A_{\mu \nu \rho \sigma }`$ under supersymmetry is
$$\delta (ϵ)A_{\mu \nu \rho \sigma }=\left[6\overline{ϵ}^i\gamma _{[\mu \nu \rho }\psi _{\sigma ]}^jW+\mathrm{i}\overline{ϵ}{}_{}{}^{i}\gamma _{[\mu \nu }^{}A_\rho ^{(R)}\psi _{\sigma ]}^j\mathrm{i}\frac{3}{2}\overline{ϵ}{}_{}{}^{i}\gamma _{\mu \nu \rho \sigma }^{}\lambda _x^jW^{,x}\right]Q_{ij}.$$
(3.5)
Under gauge transformations the 4-form transforms as follows
$$\frac{1}{4!}\epsilon ^{\mu \nu \rho \sigma \tau }\delta _RA_{\mu \nu \rho \sigma }=\frac{1}{2}e\left[\overline{\psi }_\mu ^i\gamma ^{\mu \tau \rho }\psi _\rho ^j+\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\tau }\lambda ^{jx}\right]Q_{ij}\mathrm{\Lambda }_R.$$
(3.6)
We define the covariant flux $`\widehat{F}`$ as follows:
$`\widehat{F}`$ $``$ $`\frac{1}{4!}e^1\epsilon ^{\mu \nu \rho \sigma \tau }_\mu A_{\nu \rho \sigma \tau }+\frac{1}{2}\overline{\psi }_\mu ^i\gamma ^{\mu \nu \rho }A_\nu ^{(R)}Q_{ij}\psi _\rho ^j+\frac{1}{2}\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\mu }A_\mu ^{(R)}Q_{ij}\lambda ^{jx}`$ (3.7)
$`+\frac{3}{2}\left[\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\mu }\psi _\mu ^jW^{,x}\mathrm{i}\frac{1}{2}\overline{\psi }{}_{\mu }{}^{i}\gamma _{}^{\mu \nu }\psi _\nu ^jW\right]Q_{ij}.`$
The closure of the algebra on shell is due to a $`G`$ field equation
$$\widehat{F}=12G\left(W^2\frac{3}{4}(\frac{W}{\phi ^x})^2\right)+\mathrm{i}\overline{\lambda }{}_{}{}^{ix}\lambda _{}^{jy}\left(\frac{1}{4}g_{xy}W+\sqrt{\frac{3}{2}}T_{xyz}W^{,z}\right)Q_{ij},$$
(3.8)
where the bosonic part is related to the potential, given by
$$V=6G^2\left[W^2\frac{3}{4}(\frac{W}{\phi ^x})^2\right].$$
(3.9)
This equation relates the flux to the potential via the singlet field $`G`$ (up to terms with fermions). We will see later that on shell the flux will change sign when passing through the wall. This will explain the role of the wall as a sink for the flux.
Remarks:
* the action is invariant under 8 supersymmetries.
* the Lagrangian is invariant up to a total derivative. This is sufficient if the fields either drop off at infinity (as it is supposed to be in the 4-dimensional spacetime), or the space is cyclic and the fields are continuous (as it is supposed to be in the $`x^5`$ direction). To have at the end $`G(x)`$ piecewise constant, but not everywhere the same, it is clear that we need at least 2 branes where it can jump.
* The procedure outlined in step 2 is very general and does not depend on the details of the configuration.
We have explained in section 2 that the GST-action allows two different parity assignments for $`G`$. The action (3.4) respects both choices with the assignments
$$\{\begin{array}{ccc}\mathrm{\Pi }(G)=1,& \mathrm{\Pi }(A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }})=+1,& \mathrm{\Pi }(A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }5})=1,\\ \mathrm{\Pi }(G)=+1,& \mathrm{\Pi }(A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }})=1,& \mathrm{\Pi }(A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }5})=+1.\end{array}$$
(3.10)
### 3.3 Step 3: the brane as a sink for the flux
If we have only the bulk action, the field equation still implies that $`G`$ is constant everywhere. The flux is proportional to the potential and on shell for $`A_{\mu \nu \rho \sigma \tau }`$ and $`G`$ we recover the standard supergravity.
We thus need sources to modify the field equation on $`G`$. These we can choose according to a physical situation. For reproducing the scenario described above, we take two branes positioned at $`x_5=0`$ and at $`x_5=\stackrel{~}{x}_5`$. For both branes we introduce a worldvolume action, which basically is the Dirac–Born–Infeld action in a curved background with all excitations of the worldvolume fields set equal to zero.
The brane action includes the pullback of the metric, the scalars and the 4-form of the bulk action. The coefficient of the determinant of the metric is taken to be the function $`W`$. It is related to the central charge, similar to what was obtained for black holes in $`N=2`$, $`d=4`$ in . The Wess–Zumino term describes the charge of the domain wall.
Remember that, as explained after (2.1), odd fields vanish on the branes. Therefore, if we want to use the pullback of the components $`A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}`$ on the brane, we need that their parity is even. This forces us to take the first choice in (3.10). This is consistent with the scenario of , but it is problematic to incorporate the approach of in our framework of ‘consistent supersymmetry on singular spaces’.
The brane action is
$$S_{brane}=2gd^5x\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right)\left(e_{(4)}3\alpha W+\frac{1}{4!}\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}\right),$$
(3.11)
where $`e_{(4)}`$ is the determinant of the 4 by 4 vierbein $`e_{\underset{¯}{\mu }}^m`$, and $`\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}=\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }5}`$ is in the same way a 4-density. The factor $`\alpha =\pm 1`$ is a sign to be chosen later. The new field equation for $`A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}`$ is
$$_5G(x^5)=2g\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right),$$
(3.12)
which has as solution
$$G(x)=g\epsilon (x^5),$$
(3.13)
for $`\stackrel{~}{x}^5<x^5<\stackrel{~}{x}^5`$. One should understand $`\epsilon (x^5)`$ as the function that is $`+1`$ for $`0<x^5<\stackrel{~}{x}^5`$, and $`1`$ for $`\stackrel{~}{x}^5<x^5<0`$. Thus it has also a jump at $`x^5=\stackrel{~}{x}^5\stackrel{~}{x}^5`$, and
$$\frac{1}{2}_5\epsilon (x^5)=\delta (x^5)\delta (x^5\stackrel{~}{x}^5).$$
(3.14)
Now we may look at the equations of motion for the flux with account of the value of the on-shell $`G`$-field. The bosonic part of the flux (3.8) is on-shell
$`\widehat{F}_{onshell,bos}=\epsilon (x^5)12g\left[W^2{\displaystyle \frac{3}{4}}({\displaystyle \frac{W}{\phi ^x}})^2\right].`$ (3.15)
Clearly the flux is changing the sign when passing through the brane, which justifies the title of this section. This may be contrasted with the properties of the potential when the $`G`$-field is on shell:
$$V_{onshell}=6g^2\left[W^2\frac{3}{4}(\frac{W}{\phi ^x})^2\right],$$
(3.16)
where the standard assumption is made that $`\epsilon ^2(x^5)=1`$. The potential does not care about the existence of the brane.
The fermion mass terms on shell for the $`G`$-field also change the sign across the wall, as it follows from the terms in the action that are quadratic in fermions and linear in $`G`$.
We now consider the invariance of the brane action. The fünfbein satisfies $`e_5^m=e_{\underset{¯}{\mu }}^5=0`$. This is due to the parity assignment and orbifold condition, which implies that only even parity fields are non-zero on the brane. The variation of the brane action is ($`A_{\underset{¯}{\mu }}^{(R)}`$ is zero on the brane, and therefore those contributions can be neglected)
$`\delta S_{brane}`$ $`=`$ $`3g{\displaystyle \mathrm{d}^5x\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right)}`$ (3.17)
$`\times \left[We_{(4)}\overline{ϵ}^i\gamma ^me_m^{\underset{¯}{\mu }}\left(\alpha \psi _{\underset{¯}{\mu }i}\mathrm{i}\gamma _5Q_{ij}\psi _{\underset{¯}{\mu }}^j\right)+W_{,x}\overline{ϵ}^i\left(\mathrm{i}\alpha \lambda _i^x\gamma _5Q_{ij}\lambda ^{xj}\right)\right].`$
This vanishes if we apply the $`_2`$ projections of section 2 with
$$M_{ij}=\mathrm{i}\alpha Q_{ij},\alpha =\pm 1.$$
(3.18)
It thus implies that
$`\psi _{\underset{¯}{\mu }i}(x^5)`$ $`=`$ $`\mathrm{i}\alpha Q_{ij}\gamma _5\psi _{\underset{¯}{\mu }}^j(x^5),`$
$`\lambda _i(x^5)`$ $`=`$ $`\mathrm{i}\alpha Q_{ij}\gamma _5\lambda ^j(x^5),`$
$`ϵ_i(x^5)`$ $`=`$ $`\mathrm{i}\alpha Q_{ij}\gamma _5ϵ^j(x^5),`$ (3.19)
such that (3.17) vanishes.
## 4 Summary of $`d=5`$ supersymmetry on $`S^1/_2`$
In summary, the new Lagrangian in a singular space with the new supersymmetry is given by
$$S_{new}(x^5)=S_{bulk}+S_{brane},S_{bulk}=S_{GST}(gG(x))\mathrm{d}^5xeF(x)G(x).$$
(4.1)
Here $`F`$ is a curl of the 4-form $`F=\frac{1}{4!}\epsilon ^{\mu \nu \rho \sigma \tau }_\tau A_{\mu \nu \rho \sigma }`$ and $`S_{GST}(g)=S_0+gS_1+g^2S_2`$ is the standard $`U(1)`$ gauged supergravity of , see appendix B. The standard supergravity at $`g=0`$ is called ungauged supergravity. We will use the notation $`_0`$ for its Lagrangian. Our new bulk theory has Lagrangian $`_0`$, there are terms linear in $`G(x)`$, which are proportional to the flux $`\widehat{F}`$ defined in (3.7), and terms quadratic in the field $`G(x)`$, see (3.9),
$`_{bulk}`$ $`=`$ $`_0+6eG^2(x)\left(W^2{\displaystyle \frac{3}{4}}W_{,x}^2\right)eG(x)\widehat{F}`$ (4.2)
$`+e\mathrm{i}G(x)Q_{ij}\overline{\lambda }{}_{}{}^{ix}\lambda _{}^{jy}\left(\frac{1}{4}g_{xy}W+\sqrt{{\displaystyle \frac{3}{2}}}T_{xyz}W^{,z}\right).`$
In addition, we will denote by $`\delta _0`$ the part of supersymmetry that acts at $`g=0`$ and which forms the supersymmetry transformations of ungauged supergravity. For completeness we repeat here the brane actions.
$$_{brane}=2g\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right)\left(e_{(4)}3W+\frac{1}{4!}\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}\right),$$
(4.3)
The new supersymmetry rules are
$`\delta (ϵ)\{e_\mu ^m,A_\mu ^I,\phi ^x\}`$ $`=`$ $`\delta _0\{e_\mu ^m,A_\mu ^I,\phi ^x\},`$
$`\delta (ϵ)\psi _{\mu i}`$ $`=`$ $`\delta _0(ϵ)\psi _{\mu i}+G(x)A_\mu ^{(R)}Q_{ij}ϵ^j+\mathrm{i}\frac{1}{2}G(x)\gamma _\mu ϵ^jWQ_{ij},`$
$`\delta (ϵ)\lambda _i^x`$ $`=`$ $`\delta _0(ϵ)\lambda _i^x\frac{3}{2}G(x)ϵ^jW^{,x}Q_{ij},`$
$`\frac{1}{4!}\delta (ϵ)A_{\mu \nu \rho \sigma }`$ $`=`$ $`\left[6\overline{ϵ}^i\gamma _{[\mu \nu \rho }\psi _{\sigma ]}^jW+\mathrm{i}\overline{ϵ}{}_{}{}^{i}\gamma _{[\mu \nu }^{}A_\rho ^{(R)}\psi _{\sigma ]}^j\mathrm{i}\frac{3}{2}\overline{ϵ}{}_{}{}^{i}\gamma _{\mu \nu \rho \sigma }^{}\lambda _x^jW^{,x}\right]Q_{ij},`$
$`\delta (ϵ)G`$ $`=`$ $`0.`$ (4.4)
The new gauge $`R`$-symmetry transformations are
$`\delta _RA_\mu ^{(R)}`$ $`=`$ $`_\mu \mathrm{\Lambda }_R,`$
$`\delta _R\psi _\mu ^i`$ $`=`$ $`G(x)\mathrm{\Lambda }_RQ^{ij}\psi _{\mu j},`$
$`\delta _R\lambda ^{xi}`$ $`=`$ $`G(x)\mathrm{\Lambda }_RQ^{ij}\lambda _{\mu j}^x,`$
$`\epsilon ^{\mu \nu \rho \sigma \tau }\delta _RA_{\mu \nu \rho \sigma }`$ $`=`$ $`\frac{1}{2}\overline{\psi }_\mu ^i\gamma ^{\mu \tau \rho }\mathrm{\Lambda }_RQ_{ij}\psi _\rho ^j\frac{1}{2}\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\tau }\mathrm{\Lambda }_RQ_{ij}\lambda ^{jx}.`$ (4.5)
The action is defined by an integral over a product space $`𝐌`$ of the 4-dimensional manifold and an orbifold, $`𝐌=𝐌_4\times \frac{S^1}{_2}`$.
$$S^{new}=_𝐌d^4x𝑑x^5^{new}.$$
(4.6)
The 4d manifold is non-compact and the 5th dimension is compact but has no boundaries in $`S^1/_2`$. Therefore, all surface terms in the variation of the action vanish assuming as usual that the parameters of supersymmetry decrease at infinities of the 4d space. The bulk and the brane actions are separately invariant. The supersymmetry transformations form an on-shell closed algebra.
PART II. The Background: Vanishing Fermions, Bosons Solve Equation of Motion
## 5 Vanishing energy
It is known that the Hamiltonian of the spatially closed universe vanishes since in absence of boundaries it is given by a diffeomorphism constraint . The basic argument goes as follows. In 4d space when the ansatz for the metric is taken in the form $`ds^2=(N^2N_iN^i)dt^2+2N_idx^idt+h_{ij}dx^idx^j`$ one finds that the Hamiltonian of constraint is $`H\frac{S}{N}=0`$. Here $`H`$ has the contribution both from the gravity and from matter. Still one has to keep in mind that the definition of the energy of the closed universe is rather subtle.
In our new supersymmetric theory we also face the problem of how to define the energy, in general. In our case the space is not an asymptotically flat or an anti-de Sitter space. Moreover, since our space is singular, we can not easily apply the Nestor–Israel–Witten construction, which for asymptotically flat or anti-de Sitter spaces would predict a non-negative energy.
Therefore, we would perform here only a partial analysis of the energy issue for supersymmetric theories in singular spaces, which has a clear conceptual basis. We hope, however, that a more general treatment of this problem is possible.
Here we limit ourselves to configurations which depend only on $`x^5`$. For such configurations the natural definition of the energy functional was suggested and studied in . We are using the warped metric in the form
$$ds^2=a^2(x^5)dx^{\underset{¯}{\mu }}dx^{\underset{¯}{\nu }}\eta _{\underset{¯}{\mu }\underset{¯}{\nu }}+(dx^5)^2.$$
(5.1)
Starting with the new Lagrangian (4.1), we may present the energy functional for static $`x^5`$-dependent bosonic configurations as follows
$`E(x^5)`$ $`=`$ $`6a^2a^2+\frac{1}{2}a^4(\phi ^x)^2+a^4V\frac{1}{4!}\epsilon ^{\mu \nu \rho \sigma 5}A_{\mu \nu \rho \sigma }G^{}`$ (5.2)
$`+2g\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right)\left(3a^4\alpha W+\frac{1}{4!}\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}\right).`$
Here means $`\frac{}{x^5}`$. This expression in turn can be given in the BPS-type form closely related to but still different due to i) the presence of the 4-form and the supersymmetry singlet off shell, ii) the presence of $`\alpha =\pm `$, which comes from the choice of the $`_2`$ action on the fermions and is introduced in (3.18).
$`E(x^5)_{BPS}`$ $`=`$ $`{\displaystyle \frac{1}{2}}a^4\left\{\left[\phi ^x3\alpha GW^{,x}\right]^212[{\displaystyle \frac{a^{}}{a}}+\alpha GW]^2\right\}+3\alpha [a^4GW]^{}`$ (5.3)
$`+\left[2g\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right)G^{}\right]\left(3a^4\alpha W+\frac{1}{4!}\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}\right).`$
The tension of the first brane, $`T_{x^5=0}^1`$, is equal to $`6g\alpha W(x^5=0)`$. The tension of the second brane, $`T_{x^5=\stackrel{~}{x}^5}^2`$, is given by $`6g\alpha W(x^5=\stackrel{~}{x}^5)`$. Even if any of them is negative, this causes no problem since we have a compensating contribution to the energy on each brane. In presence of the supersymmetry singlet field, there is an additional contribution to the energy at each brane due to the gradient of the supersymmetry singlet field $`G`$. The term with $`G^{}`$ cancels the tension contributions at each brane separately since $`G^{}=2g\left(\delta (x^5)\delta (x^5\stackrel{~}{x}^5)\right)`$ due to the field equations for the 4-form. With account of the $`A`$ and $`G`$ equations and no boundary condition, which allows to ignore $`+3\alpha [a^4GW]^{}`$, the energy functional takes the form
$`E(x^5)_{BPS}`$ $`=`$ $`{\displaystyle \frac{1}{2}}a^4\left\{\left[\phi ^x3\alpha GW^{,x}\right]^212[{\displaystyle \frac{a^{}}{a}}+\alpha GW]^2\right\}.`$ (5.4)
Note that the energy functional is still not positive definite: one perfect square with the kinetic energy of the normal scalar $`\phi `$ is positive, however that with the kinetic energy of the conformal factor of the metric is negative, as it should be. One might have a concern about some configurations where the negative contribution will dominate over the positive one, which will lead to the instability of the theory. However, using the equations of motion for $`g_{00}`$ and $`A_{\mu \nu \rho \sigma }`$ (not the BPS equations), one finds that the energy vanishes for any $`x^5`$-dependent solution of the equations of motion. This can also be reinterpreted as derived using the canonical formulation of the gravitational theory where one starts from
$$_M\frac{1}{2}\sqrt{g}R+_M𝒦=\stackrel{~}{},$$
(5.5)
where $`𝒦`$ is such that the resulting action has no second derivatives on the metric. The action of matter fields can be added and as in one finds that an on-shell energy $`E`$, at least for time-independent configurations, vanishes
$$E=0$$
(5.6)
for any solution of the equations of motion.
One very important property of the vanishing of the energy of the closed universe is that it vanishes locally and not due to the compensation of the total energy, positive and negative, in different parts of the universe. As we explained this happens also here: the cancellation of energy on each brane takes case separately: the tension term is cancelled by the energy of the supersymmetry singlet field $`G(x)`$.
## 6 BPS construction in singular spaces
The squared terms in (5.3) suggest the BPS conditions:
$$g_{xy}(\phi ^y)^{}=+3\alpha G(x^5)W_{,x},\frac{a^{}}{a}=\alpha G(x^5)W,$$
(6.1)
where we can use the on shell value of $`G(x^5)`$, which is equal to $`g\epsilon (x^5)`$, so that we obtain
$$g_{xy}(\phi ^y)^{}=+3\alpha g\epsilon (x^5)W_{,x},\frac{a^{}}{a}=\alpha g\epsilon (x^5)W.$$
(6.2)
From now on we will make the choice $`\alpha =1`$, i.e. pick up a particular property of fermions under parity. Physics depends on the sign of $`\alpha g`$. Therefore, the change in the sign of $`\alpha `$ can always be compensated by a change of the sign of $`g`$.
One can verify that the jump conditions on the branes<sup>7</sup><sup>7</sup>7It was observed in that in $`N=8`$ gauged supergravity the jump conditions on the branes may be satisfied if the tensions are related to the superpotential. However, putting the step functions in supersymmetry transformations by hand may in general cause problems with higher order corrections. Our work makes it plausible that $`N=8`$ gauged supergravity with addition of the flux and supersymmetry singlet may be constructed with the complete and consistent supersymmetry. This will generate the step functions in supersymmetry rules in presence of branes. , derived starting with the second order differential equations, are satisfied automatically due to the new supersymmetry (4.4), the on-shell condition (3.12) and the presence of the 5-form flux, changing the sign when passing through the wall.
Let us also consider the Killing spinors. In the background with only non-vanishing scalars and the warped metric (5.1), which has as only non-zero components of the spin connection $`\omega _{\underset{¯}{\mu }}^{m5}=a^{}\delta _{\underset{¯}{\mu }}^m`$, the transformations of the spinors are
$`\delta (ϵ)\lambda _i^x`$ $`=`$ $`\mathrm{i}\frac{1}{2}\gamma _5\phi ^xϵ_i\frac{3}{2}GW^xQ_{ij}ϵ^j,`$
$`\delta (ϵ)\psi _{\underset{¯}{\mu }i}`$ $`=`$ $`_{\underset{¯}{\mu }}ϵ_i+\frac{1}{2}\delta _{\underset{¯}{\mu }}^m\gamma _m\left(a^{}\gamma _5ϵ_i+\mathrm{i}aGWQ_{ij}ϵ^j\right),`$
$`\delta (ϵ)\psi _{5i}`$ $`=`$ $`ϵ_i^{}+\frac{1}{2}\mathrm{i}GW\gamma _5Q_{ij}ϵ^j.`$ (6.3)
To solve these, we split
$`ϵ_i`$ $`=`$ $`ϵ_i^++ϵ_i^{},`$
$`ϵ_i^\pm `$ $`=`$ $`\frac{1}{2}\left(ϵ_i\pm \mathrm{i}\gamma _5Q_{ij}ϵ^j\right)=\pm \mathrm{i}\gamma _5Q_{ij}ϵ^{\pm j}.`$ (6.4)
Using the conditions (6.1), the last equation of (6.3) gives the dependence of the supersymmetries on $`x_5`$. We obtain
$$ϵ_i^\pm =a^{\pm 1/2}ϵ_i^\pm (x^{\underset{¯}{\mu }}).$$
(6.5)
The second equation gives
$$_{\underset{¯}{\mu }}ϵ_i^++_{\underset{¯}{\mu }}ϵ_i^{}+\delta _{\underset{¯}{\mu }}^m\gamma _ma^{}\gamma _5ϵ_i^{}=0.$$
(6.6)
The solutions are thus
$$ϵ_i=a^{1/2}ϵ_i^{+(0)}+a^{1/2}\left(1\frac{a^{}}{a}x^{\underset{¯}{\mu }}\gamma _{\underset{¯}{\mu }}\gamma _5\right)ϵ_i^{(0)},$$
(6.7)
where $`ϵ_i^{\pm (0)}`$ are constant spinors with each only 4 real components due to the projection (6.4). Remains the first Killing equation, which implies
$$\phi ^xϵ_i^{(0)}=0.$$
(6.8)
There are thus two possibilities to solve the Killing equations.
* Maximal unbroken supersymmetry in the bulk ($`N=2`$).
Here we require that the scalars are strictly constant, $`\phi ^x=0`$, and the superpotential is independent of the scalars, $`\frac{W}{\phi ^x}=0`$, at the solution. No constraints on Killing spinors arise from the gaugino.
From the gravitino transformation, we have shown in (6.7) that in the warped geometry a result similar to takes place, with doubling of supersymmetries: the constant spinors $`ϵ_i^{\pm (0)}`$ give together 8 unbroken supersymmetries.
* 1/2 of the maximal unbroken supersymmetry in the bulk ($`N=1`$).
When the scalars are not constant, then there is the extra condition (6.8), leaving just one projected supersymmetry with 4 real components,
$$ϵ_i\mathrm{i}\gamma _5Q_{ij}ϵ^j=0,ϵ_i=a^{1/2}ϵ_i^{(0)},$$
(6.9)
where $`ϵ_i^{(0)}`$ is constant. Note that this projection of the supersymmetries is on the brane consistent with (3.19). Thus, we remain in the bulk as well as on the brane with $`1/2`$ of the original supersymmetries. Vice versa, imposing the projection (6.9) one derives from the vanishing of (6.3) that the conditions (6.1) should be satisfied.
### 6.1 Fixed scalars, doubling of supersymmetries <br>and an alternative to compactification world brane
The field equation for the 4-form and $`G`$-field Killing equations are solved if (6.1) are solved. In this section we look for the very particular solutions of these equations with maximal unbroken supersymmetry that have everywhere constant scalars<sup>8</sup><sup>8</sup>8These solutions remind the so called double-extreme black holes , which have fixed scalars in their solutions ..
The ‘fixed scalar domain wall solution’ is given by
$$(\phi ^y)^{}=0,\left(\frac{W}{\phi ^x}\right)_{crit}=0,\frac{a^{}}{a}=g\epsilon (x^5)W_{crit}.$$
(6.10)
The solution is given by the supersymmetric attractor equation in the form
$$C_{IJK}\overline{h}^J\overline{h}^K=q_I,$$
(6.11)
where
$$\overline{h}^K\sqrt{W_{crit}}h^K,$$
(6.12)
and we have used charges normalized as
$$q_I\sqrt{\frac{2}{3}}V_I,\text{such that}W=h^Iq_I,W_{,x}=\sqrt{\frac{2}{3}}h_x^Iq_I.$$
(6.13)
Consistency implies
$$W_{crit}(C_{IJK},q_I)=h_{crit}^Iq_I=(\overline{h}^Iq_I)^{2/3}=(C_{IJK}\overline{h}^I\overline{h}^J\overline{h}^K)^{2/3}.$$
(6.14)
In some cases the explicit solution of the attractor equation is known in the form
$$\overline{h}^I(C_{IJK},q_I)$$
(6.15)
(see e.g. where many examples are given). The metric is
$$ds^2=e^{2gW_{crit}|x^5|}dx^{\underset{¯}{\mu }}dx^{\underset{¯}{\nu }}\eta _{\underset{¯}{\mu }\underset{¯}{\nu }}+(dx^5)^2.$$
(6.16)
If we choose $`gW_{crit}`$ to be positive (which means that at $`x^5=0`$ we have a positive tension brane), the metric is that introduced in , where two branes are present at some finite distance from each other. We will refer to this scenario as RSI. The second brane has a negative tension and can be sent to infinity, in principle, which leads to an alternative to compactification, discussed in . We will refer to this as RSII. Whether this limit is totally consistent is an independent issue, however the warp factor in the metric can be chosen to exponentially decrease away from the positive tension brane. This is not in a contradiction with the no-go theorem . For constant scalars at the critical point, the metric behaves differently from the case when scalars are not constant but approaching the critical point. This can also be explained by the doubling of unbroken supersymmetries in the bulk for these solutions. Indeed, the gaugino transformations are vanishing without any constraint on the Killing spinors and the gravitino transformations also have an 8-dimensional zero mode as shown in (6.7). Note that the curvature is everywhere constant, except on the branes where the metric has a cusp. For example, the scalar curvature for this solution, is equal to
$$\frac{1}{4}R=5g^2W_{crit}^2g(\delta (x^5)\delta (x^5\stackrel{~}{x}^5))W_{crit}.$$
(6.17)
The simplest example of fixed scalar domain wall (related to double-extreme $`STU`$ black holes) comes out from the M5-brane compactified on $`T^6`$ so that $`C_{IJK}h^Ih^Jh^K=STU=1`$ and $`W_{crit}=3(q_Sq_Tq_U)^{1/3}`$ (see sec. 4.3 in ). Here $`q_S,q_T,q_U`$ are FI terms.
### 6.2 BPS equations in very special geometry (with vector multiplets)
In the context of our present work in the more general case with vector multiplets present and non-constant scalars, we will first change coordinates, write down the energy functional in the new coordinate system and proceed by solving the BPS conditions following from the energy functional. We take
$$ds^2=a^2(y)dx^{\underset{¯}{\mu }}dx^{\underset{¯}{\nu }}\eta _{\underset{¯}{\mu }\underset{¯}{\nu }}+a^4(y)dy^2,$$
(6.18)
so that $`\frac{}{x^5}=a^2\frac{}{y}`$ and we will use $`\dot{}`$ for $`\frac{}{y}`$. The BPS-type energy functional for static $`y`$-dependent bosonic configurations following from the new Lagrangian (4.1) is
$`E(y)`$ $`=`$ $`{\displaystyle \frac{1}{2}}a^2\left\{[a^2\dot{\phi ^x}3GW^{,x}]^212[a\dot{a}+GW]^2\right\}+3{\displaystyle \frac{d}{dy}}[a^4GW]`$
$`+\left[2g\left(\delta (y^5)\delta (y^5\stackrel{~}{y}^5)\right)\dot{G}\right]\left(3a^4\alpha W+\frac{1}{4!}\epsilon ^{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}A_{\underset{¯}{\mu }\underset{¯}{\nu }\underset{¯}{\rho }\underset{¯}{\sigma }}\right).`$
The stabilization equations for the energy are the BPS conditions:
$$a^2g_{xy}\dot{\phi ^y}=+3G(y)W_{,x},a\dot{a}=G(y)W,$$
(6.19)
where on shell for the 4-form field
$$G(y)=g\epsilon (y),\dot{G}=2g\left(\delta (y)\delta (y\stackrel{~}{y})\right).$$
(6.20)
The solution of these closely related equations is known both in the context of black holes and domain walls in smooth supergravities. The difference is that we have now derived the BPS equations from supersymmetry with the step functions, and the signs are correlated with the choice of the parity assignments for fermions. This takes care of the jump conditions on the branes where our solutions have kinks.
The equations (6.19) can be combined to one $`(n+1)`$-component equation with a free index $`I`$ in very special geometry. The equations (B.3) imply for any derivative
$$g_{xy}\dot{\phi }^y=\sqrt{\frac{3}{2}}h_{Ix}\dot{h}^I=\sqrt{\frac{3}{2}}h_x^I\dot{h}_I.$$
(6.21)
The BPS equations are then
$`h_x^I\left(a^2\dot{h}_I+2q_IG(y)\right)=0,`$ $`2a\dot{a}+2h^Iq_IG(y)=0,`$
$`h_x^I\left({\displaystyle \frac{d}{dy}}(a^2h_I)+2q_IG(y)\right)=0,`$ $`h^I\left({\displaystyle \frac{d}{dy}}(a^2h_I)+2q_IG(y)\right)=0.`$ (6.22)
These $`n+1`$ equations are equivalent to
$$\frac{d}{dy}(a^2h_I)+2q_IG(y)=0.$$
(6.23)
We can rewrite it using the expression for the dual coordinate $`h_I=C_{IJK}h^Jh^K`$
$$\frac{d}{dy}(C_{IJK}\stackrel{~}{h}^J\stackrel{~}{h}^K)=2G(y)q_I\text{where}\stackrel{~}{h}^Ia(y)h^I.$$
(6.24)
### 6.3 Supersymmetric Domain Walls with non-constant scalars
Using the on-shell expression for $`G(y)`$ we may solve this and get that
$$C_{IJK}\stackrel{~}{h}^J\stackrel{~}{h}^K=H_I(y),$$
(6.25)
where $`H_I`$ is a harmonic function,
$$H_I=c_I2gq_I|y|,$$
(6.26)
satisfying the equation
$$\frac{d}{dy}\frac{d}{dy}H_I=4gq_I[\delta (y)\delta (y\stackrel{~}{y})].$$
(6.27)
This is an explicit answer for a given $`C_{IJK}`$ and $`H_I`$ under the condition that we know how to solve the algebraic attractor equation (6.11). If we know the solution of the algebraic attractor equation (6.11) in the form (6.15), we simply replace the $`q_I`$ in this expression by the harmonic functions $`H_I(y)`$.
$$\stackrel{~}{h}^I(y)=\stackrel{~}{h}^I(C_{IJK},H_I(y)),$$
(6.28)
and we can find the scalars and the metric. Contracting the attractor equation with $`\stackrel{~}{h}^I`$ we find that
$$C_{IJK}\stackrel{~}{h}^I\stackrel{~}{h}^J\stackrel{~}{h}^K=a^3(y)C_{IJK}h^Ih^Jh^K=a^3(y)\text{or}a^2(y)=h^IH_I.$$
(6.29)
Our solution for the domain wall metric is therefore given by (6.18) with $`a(y)`$ given by this expression.
The choice of coordinates for the scalars $`\phi ^x`$ has been left unspecified so far. One possible choice is that similar to the ‘special coordinates’ in $`d=4`$, $`N=2`$:
$$\varphi ^I=\frac{\stackrel{~}{h}^I(y)}{\stackrel{~}{h}^0(y)}=\{1,\phi ^x(y)\}.$$
(6.30)
The general class of domain wall solutions as described in the previous section has many particular realizations, depending on the choice of the intersection numbers $`C_{IJK}`$ as well as on the choice of the initial conditions at $`y=0`$ defined by the first terms in the harmonic functions, $`c_I`$. Only if at least some of the $`c_I`$ are not vanishing, would the scalars depend on $`y`$. Otherwise all ratios of harmonic functions will be $`y`$-independent and therefore all scalars will be constants and the solutions will be those from the previous section.
The properties of the solutions with non-constant scalars depend on the choice of the model, i.e. the choice of the intersection numbers $`C_{IJK}`$, and also on various choices of the signs between the first and the second term in the harmonic functions. The options are:
* All $`c_I`$ have the opposite sign to all $`gq_I`$, i.e. none of the harmonic functions vanishes at some value of $`y`$, $`|H_I|=|c_I|+|2gq_I||y|>0`$. In such case one can directly use most of the solutions of the stabilization equations from and present the relevant domain walls. Particularly interesting ones require the second wall to cut off the part of the 5D space where the relevant CY cycles collapse and/or the metric of the moduli space vanishes.
* At least some of the $`c_I`$ have the same sign as $`gq_I`$, e.g. for positive $`c_1`$ and positive $`gq_1`$, $`H_1=c_12gq_1|y|`$ vanishes at $`|y|_{sing}=\frac{c_1}{2gq_1}`$. Such cases definitely require the second brane at $`|\stackrel{~}{y}|`$ to be placed before $`|y|_{sing}`$ since at $`|y|_{sing}`$ the metric of the moduli space and the space-time metric may vanish.
## 7 Examples
Here we briefly present a couple of examples. It would be interesting to undertake a more careful study of these solutions.
### 7.1 STU Wall
We consider again $`C_{IJK}h^Ih^Jh^K=STU=1`$ and the harmonic functions are given by
$$H_S=c_S2gq_S|y|,H_T=c_T2gq_T|y|,H_U=c_U2gq_U|y|.$$
(7.1)
If all $`c`$ are positive and all $`gq`$ are negative, all harmonic functions are strictly positive. The metric is
$$ds^2=3(H_SH_TH_U)^{1/3}(dx)^2+3^2(H_SH_TH_U)^{2/3}dy^2,$$
(7.2)
and the moduli are
$$S=\left(\frac{H_TH_U}{H_S^2}\right)^{1/3},T=\left(\frac{H_SH_U}{H_T^2}\right)^{1/3},U=\left(\frac{H_TH_S}{H_U^2}\right)^{1/3}.$$
(7.3)
We use the normalization $`3(c_Sc_Tc_U)^{1/3}=1`$ so that near $`|y|=0`$ the domain wall metric (6.18) tends to a flat Minkowski metric.
$$ds_{y0}^2dx^{\underset{¯}{\mu }}dx^{\underset{¯}{\nu }}\eta _{\underset{¯}{\mu }\underset{¯}{\nu }}+dy^2.$$
(7.4)
If near the second wall at $`|y|=|\stackrel{~}{y}|`$ the values of the second terms in the harmonic functions are much larger than the first ones, the metric near this brane approaches the boundary of the $`adS`$ space with
$$ds^2\rho ^2dx^{\underset{¯}{\mu }}dx^{\underset{¯}{\nu }}\eta _{\underset{¯}{\mu }\underset{¯}{\nu }}+R_{AdS}^2\left(\frac{d\rho }{\rho }\right)^2,$$
(7.5)
where $`y=\frac{1}{2}R_{AdS}\rho ^2`$ and $`R_{AdS}^2=9g^2(q_Sq_Tq_U)^{2/3}`$.
If in any of the harmonic functions the sign between the terms is opposite, and the harmonic function vanishes at $`|y|_{sing}`$, the second brane has to be at the position $`|\stackrel{~}{y}|<|y|_{sing}`$.
### 7.2 Calabi-Yau Wall (Base $`𝐏_2`$ Vacuum)
Here we take an example of a large complex structure limit of a particular Calabi-Yau threefold which defines a cubic form for 5D supergravity.
$$𝒱(h^1,h^2)=9(h^1)^3+9(h^1)^2h^2+3h^1(h^2)^2=1.$$
(7.6)
The attractor equation for this theory was solved in . We will use this solution to present one for the CY domain wall. We perform a coordinate redefinition to variables $`U`$ and $`T`$, related to the basic cycles $`h^1`$ and $`h^2`$ by $`U=6^{\frac{1}{3}}h^1`$, $`T=6^{\frac{1}{3}}(h^2+\frac{3}{2}h^1)`$ and
$$𝒱(U,T)=\frac{3}{8}U^3+\frac{1}{2}UT^2=1.$$
(7.7)
The relevant harmonic functions are given by
$$H_U=c_U2gq_U|y|,H_T=c_T2gq_T|y|,$$
(7.8)
and we choose $`H_U>0`$ and $`H_T>0`$. The domain wall metric takes the form (6.18), where
$$a^2(|y|)=\left[H_U\frac{2}{\sqrt{3}}\sqrt{H_U\sqrt{H_U^2\frac{9}{4}H_T^2}}+H_T\sqrt{3}\sqrt{H_U+\sqrt{H_U^2\frac{9}{4}H_T^2}}\right]^{2/3}.$$
(7.9)
The moduli are given by
$`U(|y|)={\displaystyle \frac{2}{\sqrt{3}a(|y|)}}\sqrt{H_U\sqrt{H_U^2{\displaystyle \frac{9}{4}}H_T^2}},T(|y|)={\displaystyle \frac{\sqrt{3}}{a(|y|)}}\sqrt{H_U+\sqrt{H_U^2{\displaystyle \frac{9}{4}}H_T^2}}.`$ (7.10)
The range of the variables is determined by the positivity of the basic cycles $`h^1`$ and $`h^2`$, which translates for our solution into
$$2H_U3H_T>0,H_U>0,H_T>0.$$
(7.11)
To satisfy this condition at $`|y|=0`$ we have to require that $`c_U>0`$, $`c_T>0`$ and $`2c_U3c_T>0`$. Note that if $`H_T`$ would vanish, which may happen for $`gq_T>0`$ and $`c_T>0`$ at $`|y|_{sing}=\frac{c_T}{2gq_T}`$, this would be the point where the cycle $`h^1`$ would collapse. Indeed at $`H_T=0`$, the ratio $`\frac{U}{T}=\frac{h^1}{h^2+(3/2)h^1}=0`$. At this point also $`a(|y|_{sing})=0`$ and the spacetime metric is singular. To avoid such a singularity we may require that the second brane is at
$$|\stackrel{~}{y}|<|y|_{sing}=\frac{c_T}{2gq_T}.$$
(7.12)
Even if none of the harmonic functions $`H_U`$ and $`H_T`$ vanishes, i.e. $`gq_T`$ and $`gq_U`$ are both negative, we still have an inequality
$$2c_U3c_T|y|2g(2q_U3q_T)>0.$$
(7.13)
If the condition $`g(2q_U3q_T)>0`$ is satisfied, then there is also an upper limit for the distance between the branes:
$$|\stackrel{~}{y}|<\frac{2c_U3c_T}{2g(2q_U3q_T)}.$$
(7.14)
At this point $`h^2=6^{1/3}(T\frac{3}{2}U)=0`$, i.e. the second cycle is collapsing. Thus to prevent this from happening, we have to put the second wall before this value of $`|y|`$.
As we already mentioned, a more careful and detailed analysis of domain walls of section 6.3 in application to other CY spaces, as it was done for CY black holes, may lead to more interesting and diverse configurations.
PART III. Discussion:10D supersymmetry and 8-branes on the orbifold.
The massive 10D supergravity theory of Romans describing the 8-brane is expected to have the same basic features as 5D supergravity describing the 3-brane. Thus we expect using and the results of the present paper that the 10D supersymmetry on $`S^1/_2`$ can be realized as follows:
$$S_{new}=S_{bulk}+S_{brane},S_{bulk}=S_{Romans}(mM(x))+S_A=S_{BGPT}+\mathrm{fermions}.$$
(8.1)
Here $`S_{Romans}(mM(x))`$ is the Romans action where the constant mass $`m`$ is replaced by a field $`M(x)`$ and the 9-form field is added in $`S_A=ϵ^{(10)}𝑑A_9M`$ so that the total bosonic bulk action is that given by Bergshoeff, Green, Papadopoulos and Townsend in (we have not given the fermion terms which can be taken from the Romans action in ). Note that a slightly different version of the bosonic bulk plus brane action, including a kinetic term for the 9-form potential and a brane kinetic term for the BI vector fields, has been given in <sup>9</sup><sup>9</sup>9We obtain a kinetic term for the 9-form potential if we eliminate $`M`$ as auxiliary field, rather than using the field equation for the 9-form..
The brane action has to be added to the bulk action so that the on-shell value of the $`M`$-field is piecewise constant. In particular, for an $`S^1/_2`$ orbifold, the value of $`M`$ should change sign across the wall. The parity assignments of the fields follow straightforwardly by reducing the $`D=11`$ parity assignments under $`yy`$, where $`y`$ refers to the direction between the “end of the world” branes of Hor̆ava and Witten, over a direction different from $`y`$, i.e. a direction inside the branes.
Without the brane action, the on-shell field $`M`$ is the same constant everywhere and this prevents us from getting the harmonic functions depending on moduli $`|y|`$. The brane actions for the 8-brane are expected to contain the following terms
$$_{brane}2m\left(\delta (y)\delta (y\stackrel{~}{y})\right)\left(e_{(9)}e^{\frac{5\sigma }{4}}+\frac{1}{9!}\epsilon ^{\underset{¯}{\mu }_1\mathrm{}\underset{¯}{\mu }_9}A_{\underset{¯}{\mu }_1\mathrm{}\underset{¯}{\mu }_9}\right),$$
(8.2)
where $`\sigma `$ is the Romans scalar which is related to the standard dilaton $`\varphi `$ via $`\sigma =4/5\varphi `$ such that the brane tension is proportional to $`1/g_s=e^{<\varphi >}`$. The coordinate $`y`$ refers to the direction connecting the two branes. The modified supersymmetry transformations in the presence of the supersymmetry singlet field $`M`$ and of the 9-form are expected to exist and the bulk and the brane actions will be supersymmetry invariant. Note that the brane action only contains bosonic terms. It is supersymmetric due to the orbifold condition $`\psi _{\underset{¯}{\mu }}(y)=\gamma _y\psi _{\underset{¯}{\mu }}(y)`$.
The new physics due to the presence of the brane action will show up via the $`M`$-field equation. The bosonic part of this equation following from the bulk action is given in and reads
$`\widehat{F}ϵ\widehat{F}_{10}=M(x)eK(B).`$ (8.3)
Here $`K`$ is a polynomial in terms of the massive 2-form field $`B`$ and $`\widehat{F}_{10}`$ is the covariant curvature of $`A_9`$. With account of the brane action (8.2) added according to our rule of consistent supersymmetry in the singular spaces, the on-shell value of the $`M`$-field will be proportional to the step function:
$`\widehat{F}=m\epsilon (y)eK(B).`$ (8.4)
Clearly the flux is changing sign when passing through the brane, according to the field equations of the theory.
One would hope that the improved supersymmetric formulation of the 8-brane of string theory in the framework of 10D supergravity in a singular space will shed some light on the M-theory and its 9-brane. Since 11D supergravity does not admit a cosmological constant, one may in this way uplift from 10D to 11D some important information on extended objects of string theory, which is difficult to understand from the 11D massless supergravity point of view.
A related general problem to which the construction of this paper may be useful is to clarify the appearance of chiral fermions on the brane in the context of supersymmetric theory with extended supersymmetry in the bulk.
## Acknowledgments.
We are grateful to A. Ceresole, G. Gibbons, A. Giveon, S. Gubser, S. Ferrara, N. Lambert, W. Lerche, A. Linde, J. Louis, A. Lukas, D. Marolf, P. Mayr, K. Stelle, B. Ovrut, W. Unrhu and D. Waldram for interesting and useful discussions and to the Theory Division of CERN where this work was initiated and to large extent performed. This work was supported by the European Commission TMR programme ERBFMRX-CT96-0045, in which E. Bergshoeff is associated with Utrecht university. The work of R.K was supported by NSF grant PHY-9870115.
## Appendix A Notations
The metric is $`(++++)`$, and $`[ab]`$ denotes antisymmetrization with total weight 1, thus $`\frac{1}{2}(abba)`$. We use the following indices
$`\mu `$ $`0,\mathrm{},3,5`$ local spacetime
$`\underset{¯}{\mu }`$ $`0,\mathrm{},3`$ 4-d local spacetime
$`a`$ $`0,\mathrm{},3,5`$ tangent spacetime
$`m`$ $`0,\mathrm{},3`$ tangent spacetime in 4 dimensions
$`i`$ $`1,2`$ $`SU(2)`$
$`I`$ $`0,\mathrm{},n`$ vectors
$`x`$ $`1,\mathrm{},n`$ scalars in vector multiplets. (A.1)
For the curvatures and connections, we use the conventions:
$`\omega _\mu ^{ab}`$ $`=`$ $`2e^{\nu [a}_{[\mu }e_{\nu ]}{}_{}{}^{b]}e^{\nu [a}e^{b]\sigma }e_{\mu c}_\nu e_\sigma {}_{}{}^{c},`$
$`R_{\mu \nu }^{ab}`$ $`=`$ $`2_{[\mu }\omega _{\nu ]}^{ab}+2\omega _{[\mu }^{ac}\omega _{\nu ]}{}_{c}{}^{}{}_{}{}^{b},`$
$`R_{\mu \nu }`$ $`=`$ $`R_{\mu \rho }{}_{}{}^{ba}e_{b}^{}{}_{}{}^{\rho }e_{\nu a}^{},R=g^{\mu \nu }R_{\mu \nu },`$
$`R^\mu _{\nu \rho \sigma }`$ $`=`$ $`R_{\rho \sigma }{}_{}{}^{ab}e_{a}^{\mu }e_{\nu b}=2_{[\rho }\mathrm{\Gamma }_{\sigma ]\nu }^\mu +2\mathrm{\Gamma }_{\tau [\rho }^\mu \mathrm{\Gamma }_{\sigma ]\nu }^\tau .`$ (A.2)
For the metric (5.1), the only non-zero components of the spin connection are
$$\omega _{\underset{¯}{\mu }}^{m5}=a^{}\delta _{\underset{¯}{\mu }}^m.$$
(A.3)
Here means $`\frac{}{x^5}`$. The non-zero curvature components, Ricci tensor and scalar curvature are
$`R_{5\underset{¯}{\mu }}{}_{}{}^{m5}=a^{\prime \prime }\delta _{\underset{¯}{\mu }}^m,R_{\underset{¯}{\mu }\underset{¯}{\nu }}{}_{}{}^{mn}=2a^2\delta _{\underset{¯}{\mu }}^{[m}\delta _{\underset{¯}{\nu }}^{n]},`$
$`R_{55}=4a^1a^{\prime \prime },R_{\underset{¯}{\mu }\underset{¯}{\nu }}=(aa^{\prime \prime }+3a^2)\eta _{\underset{¯}{\mu }\underset{¯}{\nu }},R=8a^1a^{\prime \prime }+12a^2a^2,`$
$`eR=12a^2a^2+8\left(a^{}a^3\right)^{}.`$ (A.4)
The Levi–Civita tensor is real, and
$$\epsilon _{abcde}\epsilon ^{abcde}=5!,\epsilon ^{\mu \nu \rho \sigma \tau }=ee_a^\mu e_b^\nu \mathrm{}e_e^\tau \epsilon ^{abcde}.$$
(A.5)
The gamma matrices are related by
$$\gamma ^{abcde}=\mathrm{i}\epsilon ^{abcde}.$$
(A.6)
$`SU(2)`$ indices are raised and lowered with $`\epsilon _{ij}`$, where $`\epsilon _{12}=\epsilon ^{12}=1`$, in NW–SE convention:
$$X^i=\epsilon ^{ij}X_j,X_i=X^j\epsilon _{ji}.$$
(A.7)
Spinor indices are omitted. The charge conjugation $`𝒞`$ and $`𝒞\gamma _a`$ are antisymmetric. $`𝒞`$ is unitary and $`\gamma _a`$ is hermitian apart from the timelike one, that is antihermitian. The bar is the Majorana bar:
$$\overline{\lambda }^i=(\lambda ^i)^T𝒞.$$
(A.8)
Define the charge conjugation operation on spinors as
$$(\lambda ^i)^C\alpha ^1B^1\epsilon ^{ij}(\lambda ^j)^{},\overline{\lambda }^{iC}\overline{(\lambda ^i)^C}=\alpha ^1(\overline{\lambda }{}_{}{}^{k})^{}B\epsilon ^{ki},$$
(A.9)
where $`B=𝒞\gamma _0`$, and $`\alpha `$ is arbitrary $`\pm 1`$ when you use the convention that complex conjugation does not interchange the order of spinors, or $`\pm \mathrm{i}`$ when complex conjugation does interchange the order of spinors. Symplectic Majorana spinors satisfy $`\lambda =\lambda ^C`$. Charge conjugation acts on gamma matrices as $`(\gamma _a)^C=\gamma _a`$, does not change the order of matrices, and works on matrices in $`SU(2)`$ space as $`M^C=\sigma _2M^{}\sigma _2`$. Complex conjugation can then be replaced by charge conjugation, if for every bispinor one inserts a factor $`1`$. Then e.g. the expression
$$\overline{\lambda }^i\gamma ^\mu _\mu \lambda _i$$
(A.10)
is real for symplectic Majorana spinors. For more details, see e.g. .
## Appendix B Supergravity in 5 dimensions with vector multiplets
We present here 5-dimensional supergravity with vector multiplets, which we will use in the main text. To put it in perspective, we repeat that the pure supergravity was constructed in . The coupling with vector multiplets was obtained in , and with gauging of the vectors in . This was extended to tensor multiplets in , and an action coupled to an arbitrary quaternionic manifold is obtained in .
We consider the coupling of abelian vector multiplets to supergravity. The fields are the fünfbein, the gravitino, the vectors, scalars and gauginos:
$$e_\mu ^a,\psi _\mu ^i,A_\mu ^I,\phi ^x,\lambda ^{ix}.$$
(B.1)
The vector multiplets couplings are determined by the symmetric real constant tensor $`C_{IJK}`$. The scalars appear in the functions<sup>10</sup><sup>10</sup>10These functions are arbitrary up to non-degeneracy conditions that the $`(n+1)\times (n+1)`$ matrix $`(h^I,h_x^I)`$ (see definition below) should be invertible. $`h^I(\phi )`$, satisfying
$$h^I(\phi )h^J(\phi )h^K(\phi )C_{IJK}=1,$$
(B.2)
and define the quantities
$`h_I(\phi )C_{IJK}h^J(\phi )h^K(\phi )=a_{IJ}h^J,a_{IJ}2C_{IJK}h^K+3h_Ih_J,`$
$`h_x^I\sqrt{\frac{3}{2}}h_{,x}^I(\phi ),h_{Ix}a_{IJ}h_x^J=\sqrt{\frac{3}{2}}h_{I,x}(\phi ),`$
$`g_{xy}h_x^Ih_y^Ja_{IJ}=2h_x^Ih_y^JC_{IJK}h^K,T_{xyz}C_{IJK}h_x^Ih_y^Jh_z^K,`$ (B.3)
with <sub>,x</sub> an ordinary derivative with respect to $`\phi ^x`$. The $`I`$-type indices are lowered or raised with $`a_{IJ}`$ or its inverse, which we assume to exist, and the same holds for the metric $`g_{xy}`$ used for the indices on scalars and gauginos.
Many identities have been derived in previous papers. The most useful ones for us are
$`h_Ih_x^I=0,h_Ih_J+h_I^xh_{Jx}=a_{IJ}.`$
$`h_{Ix;y}h_{Ix,y}\mathrm{\Gamma }_{xy}^zh_{Iz}=\sqrt{\frac{2}{3}}\left(h_Ig_{xy}+T_{xyz}h_I^z\right),`$ (B.4)
with the connection $`\mathrm{\Gamma }_{xy}^z`$ as usual defined such that $`g_{xy;z}=0`$.
We consider just one gauged $`U(1)`$, whose coupling constant is called $`g`$, and whose gauge vector is a linear combination of the vectors,
$$A_\mu ^{(R)}V_IA_\mu ^I,$$
(B.5)
where $`V_I`$ are real constants. It gauges a direction in $`SU(2)`$ space determined by a matrix (with real $`q_1`$, $`q_2`$ and $`q_3`$):
$`Q_i{}_{}{}^{j}=\mathrm{i}(q_1\sigma _1+q_2\sigma _2+q_3\sigma _3),(q_1)^2+(q_2)^2+(q_3)^2=1,`$
$`Q_{ij}=\mathrm{i}q_1\sigma _3q_2\text{ }\text{}+\mathrm{i}\sigma _1q_3,Q^{ij}=\mathrm{i}q_1\sigma _3q_2\text{ }\text{}\mathrm{i}\sigma _1q_3,`$
$`Q_{ij}Q^{jk}=\delta _i^k,Q_i{}_{}{}^{j}Q_{j}^{}{}_{}{}^{k}=\delta _i^k.`$ (B.6)
The equations of are obtained by choosing $`q_1=q_3=0`$, $`q_2=1`$. They, and , introduce also other functions that are useful when one considers non-abelian gauging:
$$P_{Iij}=V_IQ_{ij},P_{ij}=h^IV_IQ_{ij},P_{ij}^x=h^{xI}V_IQ_{ij}.$$
(B.7)
We will use the quantities
$$W=\sqrt{\frac{2}{3}}h^IV_I,$$
(B.8)
with derivative
$$W_{,x}\frac{}{\phi ^x}W=\frac{2}{3}h_x^IV_I.$$
(B.9)
The action is then $`S_{GST}=\mathrm{d}^5x_{GST}`$, with
$`e^1_{GST}`$ $`=`$ $`\frac{1}{2}R(\omega )\frac{1}{2}\overline{\psi }_\mu ^i\gamma ^{\mu \nu \rho }\left(_\nu (\omega )\psi _{\rho i}+gA_\nu ^{(R)}Q_{ij}\psi _\rho ^j\right)\frac{1}{4}a_{IJ}F_{\mu \nu }^IF^{J\mu \nu }`$ (B.10)
$`+\frac{1}{6\sqrt{6}}e^1\epsilon ^{\mu \nu \rho \sigma \lambda }C_{IJK}F_{\mu \nu }^IF_{\rho \sigma }^JA_\lambda ^K`$
$`\frac{1}{2}g_{xy}_\mu \phi ^x^\mu \phi ^yV(x)`$
$`\frac{1}{2}\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\mu }\left(_\mu (\omega )\lambda _i^x+\mathrm{\Gamma }_{yz}^x(_\mu \phi ^y)\lambda _i^z+gA_\mu ^{(R)}Q_{ij}\lambda ^{jx}\right)`$
$`\mathrm{i}\frac{\sqrt{6}}{16}\left[\overline{\psi }_\mu ^i\gamma ^{\mu \nu \rho \sigma }\psi _{\nu i}F_{\rho \sigma }^Ih_I+2\overline{\psi }^{\mu i}\psi _i^\nu F_{\mu \nu }^Ih_I\right]`$
$`+\frac{1}{4}\overline{\lambda }{}_{}{}^{ix}\gamma _{}^{\mu }\gamma ^{\nu \rho }\psi _{i\mu }F_{\nu \rho }^Ih_{Ix}\frac{\mathrm{i}}{2}\overline{\lambda }{}_{x}{}^{i}\gamma _{}^{\mu }\gamma ^\nu \psi _{i\mu }_\nu \phi ^x`$
$`+\frac{\mathrm{i}}{4}\sqrt{\frac{2}{3}}(\frac{1}{4}g_{xy}h_I+T_{xyz}h_I^z)\overline{\lambda }{}_{}{}^{ix}\gamma _{}^{\mu \nu }\lambda _i^yF_{\mu \nu }^I`$
$`+gQ_{ij}\left[\mathrm{i}\overline{\lambda }{}_{}{}^{ix}\lambda _{}^{jy}\left(\frac{1}{4}g_{xy}W+\sqrt{\frac{3}{2}}T_{xyz}W^{,z}\right)\frac{3}{2}\overline{\lambda }{}_{x}{}^{i}W_{}^{,x}\gamma ^\mu \psi _\mu ^j+\mathrm{i}\frac{3}{4}\overline{\psi }{}_{\mu }{}^{i}\gamma _{}^{\mu \nu }\psi _\nu ^jW\right]`$
$`+\text{4-fermion terms.}`$
where ($`W^{,x}g^{xy}W_{,y}`$)
$`_\mu `$ $`=`$ $`_\mu +\frac{1}{4}\omega _{\mu ,ab}\gamma ^{ab},F_{\mu \nu }^I=2_{[\mu }A_{\nu ]}^I,`$
$`V(x)`$ $`=`$ $`4g^2V_IV_JC^{IJK}h_K=g^2\left(6W^2+\frac{9}{2}W^{,x}W_{,x}\right).`$ (B.11)
The action is invariant under the supersymmetry transformations
$`\delta (ϵ)e_\mu ^a`$ $`=`$ $`\frac{1}{2}\overline{ϵ}{}_{}{}^{i}\gamma _{}^{a}\psi _{\mu i},`$
$`\delta (ϵ)\psi _{\mu i}`$ $`=`$ $`_\mu ϵ_i+\frac{1}{4}\gamma ^{cd}\widehat{\omega }_{\mu ,cd}ϵ_i+gA_\mu ^{(R)}Q_{ij}ϵ^j+{\displaystyle \frac{\mathrm{i}}{4\sqrt{6}}}(\gamma _{\mu \nu \rho }4g_{\mu \nu }\gamma _\rho )ϵ_ih_I\widehat{F}^{\nu \rho I}`$
$`\frac{1}{12}\gamma _{\mu \nu }ϵ^j\overline{\lambda }{}_{i}{}^{x}\gamma _{}^{\nu }\lambda _{jx}+\frac{1}{48}\gamma _{\mu \nu \rho }ϵ^j\overline{\lambda }{}_{i}{}^{x}\gamma _{}^{\nu \rho }\lambda _{jx}+\frac{1}{6}ϵ^j\overline{\lambda }{}_{i}{}^{x}\gamma _{\mu }^{}\lambda _{jx}\frac{1}{12}\gamma ^\nu ϵ^j\overline{\lambda }{}_{i}{}^{x}\gamma _{\mu \nu }^{}\lambda _{jx}`$
$`+\mathrm{i}\frac{1}{2}g\gamma _\mu ϵ^jWQ_{ij},`$
$`\delta (ϵ)A_\mu ^I`$ $`=`$ $`{\displaystyle \frac{1}{2}}\overline{ϵ}^i\gamma _\mu \lambda _i^xh_x^I+{\displaystyle \frac{\mathrm{i}\sqrt{6}}{4}}h^I\overline{\psi }_\mu ^iϵ_i,`$
$`\delta (ϵ)\phi ^x`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2}}\overline{ϵ}^i\lambda _i^x,`$
$`\delta (ϵ)\lambda _i^x`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}}{2}}\widehat{\overline{)}}\phi ^xϵ_i+{\displaystyle \frac{1}{4}}h_I^x\gamma ^{\mu \nu }ϵ_i\widehat{F}_{\mu \nu }^Ig\frac{3}{2}ϵ^jW^{,x}Q_{ij}`$ (B.12)
$`{\displaystyle \frac{\mathrm{i}}{2}}h_I^xh_{Iy,z}\lambda _i^y\overline{ϵ}^j\lambda _j^z+{\displaystyle \frac{\mathrm{i}}{\sqrt{6}}}ϵ^j\overline{\lambda }{}_{i}{}^{y}\lambda _{j}^{z}T_{xyz},`$
where
$`\widehat{\omega }_{\mu ab}`$ $`=`$ $`\omega _{\mu ab}\frac{1}{4}\left(\overline{\psi }_b^i\gamma _\mu \psi _{ai}+2\overline{\psi }_\mu ^i\gamma _{[b}\psi _{a]i}\right),`$
$`\widehat{F}_{\mu \nu }^I`$ $`=`$ $`2_{[\mu }A_{\nu ]}^I+h_x^I\overline{\psi }{}_{[\mu }{}^{i}\gamma _{\nu ]}^{}\lambda _i^x+\frac{1}{4}\mathrm{i}\sqrt{6}h^I\overline{\psi }_{[\mu }^i\psi _{\nu ]i},`$
$`\widehat{}_\mu \phi ^x`$ $`=`$ $`_\mu \phi ^x\frac{\mathrm{i}}{2}\overline{\psi }{}_{\mu }{}^{j}\lambda _{j}^{x}.`$ (B.13)
## Appendix C Previous work on supersymmetry in singular spaces
The first idea to have two supersymmetric domain walls relating the 11-dimensional physics with the 10-dimensional one, was suggested in HW . They made the connection between the bulk and the branes at the quantum level, i.e. 1-loop anomalies of the bulk action are cancelled by a classical non-invariance of the action on the brane. The non-trivial flux that introduces the step functions into the theory, is related to an anomaly. The realization of the supersymmetry of this combined bulk and brane actions was successful at the order $`\kappa ^{2/3}`$, which was established by rather involved calculations. However, at the order $`\kappa ^{4/3}`$ things go out of control. Some $`\delta (0)`$ terms occur. The hope was expressed that a proper quantum M-theory treatment will lead to quantum consistency. The important aspect of this theory was the appearance of chiral fermions on the brane.
The next step in this development was undertaken by , where HW theory was compactified on a CY space and the “5D supergravity” with the step functions in the supersymmetry rules, was recovered. The supersymmetry in this approach relies on the supersymmetry of HW theory and therefore would be difficult to complete.
More recently, two new approaches to realize the supersymmetric RS scenario were suggested in the framework of 5D supergravity: one in and an alternative one in . The basic difference is that in the gauge coupling constant is even across the wall and in it is taken to be odd as in HW theory. We have shown in section 2, using a standard basis for symplectic spinors, that there are indeed two possibilities to make the $`_2`$ symmetry on spinors commuting with the supersymmetry: one with even gauge coupling, when the projector in parity rules anticommutes with the projector in supersymmetry rules, and the second one with the odd gauge coupling, when the projector in parity rules commutes with the projector in supersymmetry rules. So far both versions were fine.
In the case of even $`g`$ , the bulk action is simply the action of supergravity, without step functions: therefore the standard supersymmetry rules are valid for the bulk action and there is no compelling reason to add the brane actions. The motivation for adding the brane actions comes from the properties of the singular background: if one would consider configurations with unbroken supersymmetries for domain walls, one would not solve Killing equations everywhere. This enforces some particular changes in supersymmetry rules at the singularities only. The brane actions are added to compensate the supersymmetry of the bulk action, which is not supersymmetric after the $`\delta `$-functional changes in the supersymmetry rules. We have checked that if one would start not from pure supergravity without matter, as in , but with additional multiplets, the change of supersymmetry rules would depend on the scalars in the background and the issue of the closure of the algebra would become obscure as the scalars had to satisfy equations of motion. The conclusion of this analysis was that it will be difficult to generalize the approach of in pure supergravity to more general theories with matter multiplets.
We therefore decided to continue along the line of , where the gauge coupling $`g`$ was taken to be odd. This enforced us to introduce step functions in the supersymmetry rules and as a consequence the bulk supergravity action was not supersymmetric anymore: the brane action of was designed to compensate the problem of the bulk action. This was conceptually and technically correct from our perspective. The problem was how to make the total system complete. Later when the supersymmetry singlet and the 4-form potential were introduced, we were able to construct a consistent and complete supersymmetric theory where the bulk and the brane actions are supersymmetric. Remarkably, if we would not care about the conceptual issues and would only try to accomplish the complete supersymmetric theory, we would have to use the odd gauge coupling to have a consistent implementation of the 4-form field, see section 3.3, where this is explained.
Here we also would like to comment on a very interesting paper . It does not deal with fermions, therefore it does not address directly the issues discussed in our paper. However, it brings some deep insights from the perspective of the brane actions: in particular the role of the $`(D1)`$-form is stressed as required for the Wess–Zumino terms of codimension 1 branes, as well as the fact that the bulk action has to be supplemented by the term that is quadratic in the $`D`$-form flux. In our supersymmetric theory, we have not eliminated the field $`G`$ as an auxiliary field by its equation of motion. If we would have done it using the action in the form of (4.2) and (4.3), we would find out that the bulk action has a term quadratic in the flux $`\widehat{F}`$. We expect that our supersymmetric action with the brane action where the worldvolume degrees of freedom are not excited, may be further developed in the spirit of and lead to a deeper understanding of the total dynamical system.
|
warning/0007/astro-ph0007250.html
|
ar5iv
|
text
|
# RAPID STATE TRANSITIONS IN THE GALACTIC BLACK HOLE CANDIDATE SOURCE GRS 1915+105
## 1. INTRODUCTION
The X-ray transient GRS 1915+105 was discovered with the WATCH all sky X-ray monitor (Castro-Tirado et al. 1992) and it has been exhibiting a wide variety of temporal variability in its X-ray emission. It has been identified with a superluminal radio source (Mirabel & Rodriguez 1994) and attempts have been made to connect the jet production with the hard X-ray emission from the accretion disk (Harmon et al. 1997). A simultaneous X-ray and infrared observation of the source established a close link between the non-thermal infrared emission (presumably coming from a “baby-jet” analog of the superluminal ejection events) and the X-ray emission from the accretion disk (Eikenberry et al. 1998). A diverse variability in the X-ray emission, episodes of superluminal jet emission and evidences of disk-jet symbiosis has made GRS 1915+105 an ideal source to examine several concepts of accretion physics, black-hole environment, disk-jet connection etc.
Greiner et al. (1996) presented the X-ray variability characteristics of GRS 1915+105 using the ROSSI XTE data which included repeated patterns of brightness sputters, large amplitude oscillations, fast oscillations and several incidences of prolonged lulls. Narrow quasi-periodic oscillations (QPO) in the X-ray emission were discovered from the source using the Indian X-ray Astronomy Experiment (Agrawal et al. 1996) and the RXTE (Morgan & Remillard 1996). Chen, Swank & Taam (1997) found that the narrow QPO emission is a characteristic feature of the hard branch and it is absent in the soft branch which corresponds to the very high state similar to those of other black hole candidates. Morgan, Remillard, & Greiner (1997) made a systematic study of the different QPOs seen in GRS 1915+105 which ranged from 3 mHz to 67 Hz and based on the power spectra, light curves and energy spectra classified the X-ray emission into four different emission states, none of which appear identical to any of the canonical states of black hole binaries. Paul et al. (1998b), however, found that the 0.5 – 10 Hz QPO traces the change of state from a “flaring state” to a low-hard state quite smoothly along with other X-ray characteristics like the low frequency variability. Trudolyubov, Churazov, & Gilfanov (1999) studied the 1996/1997 low luminosity state and state transitions using the RXTE data and concluded that the QPO centroid frequency is correlated with the spectral and timing parameters and these properties are similar to other Galactic black hole candidates in the intermediate state. Muno, Morgan, & Remillard (1999) sampled the RXTE data over a wide range of properties and found that the 0.5 – 10 Hz QPO can be used as a tracer of the spectral state of the source and the source mainly stays in two states: the spectrally hard state when the QPO is present and the soft state when the QPO is absent.
The X-ray intensity variations seen in GRS 1915+105 were classified as lull, flare and splutter in Greiner et al. (1996), and each of these states last for a few hundred seconds. Similar X-ray light curves were analyzed in detail by Belloni et al. (1997a; 1997b) who classified them as “outbursts”; they found that the source shows distinct and different spectral and temporal characteristics during the “quiescence” and outburst. Taam, Chen, and Swank (1997) detected a wide range of transient activity including regular bursts with a recurrence time of about one minute and irregular bursts. Belloni et al. (1997b) made a detailed spectral analysis during a sequence of “bursts” and discovered a strong correlation between the quiescent phase and burst duration. Paul et al. (1998a) detected several types of bursts using the IXAE data and found evidence for matter disappearing into the event horizon of the black hole. Yadav et al. (1999) made a systematic analysis of these bursts and classified them based on the recurrence time.
Yadav et al. (1999) presented a comprehensive picture for the origin of these bursts in the light of the recent theories of advective accretion disk. It was suggested that the peculiar bursts are characteristic of the change of state of the source. The source can switch back and forth between the low-hard state and the high-soft state near critical accretion rates in a very short time scale, giving rise to the irregular and quasi-regular bursts. The fast time scale for the transition of the state is explained by invoking the appearance and disappearance of the advective disk in its viscous time scale. The periodicity of the regular bursts is explained by matching the viscous time scale with the cooling time scale of region beyond a shock front or a centrifugal barrier (Chakrabarti & Titarchuk 1995).
In this paper we present the results of a study of the time variability and spectral characteristics of the source GRS 1915+105 during the irregular bursts. We show that during these bursts the source makes distinct transitions between the two spectral states in a few seconds. In section 2 we present results obtained from an analysis of the data obtained from the Proportional Counter Array (PCA) of the Rossi X-ray Timing Explorer (RXTE). In section 3 we discuss the importance of our results and in the last section a summary of the results is given.
## 2. DATA SELECTION AND ANALYSIS
GRS 1915+105 was in a low-hard state during 1996 December to 1997 March when the hard X-ray spectral index ($``$2.0) and the soft X-ray flux (300 - 500 mCrab) were low (Greiner et al. (1998); trud:98). The source started a new outburst around 1997 April-May when the soft X-ray flux started increasing and the X-ray spectrum became soft (spectral index increased to 3 $``$ 4). It reached the high-soft state in August 1997. The 1.3 to 12.2 keV X-ray light curve of the source obtained from the RXTE ASM archives is shown in Figure 1, from 1997 January to 1997 September. Individual dwell data (see Levine et al. 1996) are plotted against the ASM day numbers (which is equal to MJD - 49353). A few dates (in 1997) are marked on the top of the figure. There are about 20 dwells on the source per day, lasting for about 90 s. Data are in ASM counts s<sup>-1</sup> (1 Crab = 75 ASM counts s<sup>-1</sup>).
The RXTE public archive contains several observations on GRS 1915+105 using the RXTE PCA (Jahoda et al. 1996). These observations typically last for a few thousand seconds and the start times of these observations are marked as circles in Figure 1. Several of these observations are analyzed and reported in the literature and these are marked with vertical arrows in the figure. The observation times of the regular bursts reported as ‘rings’ in the PCA color-color diagrams by Vilhu & Nevalainen (1998) are marked with a ‘v’ in the figure and the ‘irregular bursts’ reported by Belloni et al. (1997b) is marked as ‘B’. The peculiar repeated observation of the ‘kink’ in the light curve followed by a ‘lull’ and then ringing flares reported by Eikenberry et al. (1998) is marked by an ‘E’. Similar episodes of flares have seen by Markwardt et al. (1999) and it is marked with an ‘M’. The remaining observations which are reported in the literature are from Trudolyubov, Churazov, & Gilfanov (1999) and Muno et al. (1999). The observation times of GRS 1915+105 by the PPCs on board IXAE (Paul el al. 1998a; Yadav et al. 1999) are marked as stars in Figure 1.
It can be seen from the figure that the source was in a stable low-hard state up to 1997 April 25 (day number 1210) when the average ASM count was 20 s<sup>-1</sup>. Barring three episodes of dipping behaviors (which could be absorption dips seen in other black hole candidate sources like 4U 1630-47 - see Kuulkers et al. 1998), the source flux was stable with a rms deviation of 15%. It should be noted here that the ASM data, with a dwell time of 90 s and about 20 dwells per day, is insensitive to short term variabilities. The spectral and temporal behavior during the low-hard state was stable characterized by a hard spectrum (with the power-law index of $``$2, and the total flux in the power-law component being $``$80%) and 0.5 – 10 Hz QPOs (Trudolyubov et al. 1999; Muno et al. 1999). The fact that the canonical low-hard states of black hole candidate sources have a negligible thermal component (Chitnis et al. 1998) prompted Trudolyubov et al. (1999) to characterize this state as an “intermediate state” and they conclude that with the lowering of the accretion rate the source should go to the canonical hard state. Since the source was in similar state on several occasions (1996 July-August; 1997 October; 1998 September-October), we treat this state as the “low-hard state” of GRS-1915+105.
After 1997 April 25 the source started a steady increase in its X-ray emission with an average increase in the ASM count rate of 0.65 s<sup>-1</sup> day<sup>-1</sup> and reaching a count rate of 76 s<sup>-1</sup> in the middle of July (day number 1290). The variability, as can be deduced from the ASM count rates and measured as the fraction of rms to mean, steadily increased from 15% to 40%. It should be noted that during the low-hard state of 1996 July-August the source showed similar variability behavior: the rms variation was low (5 - 10%) during the low-hard state and it was $``$40% just before and after this state (Paul et al. 1998b). The variability, however, decreased to $`<`$15% around 1997 June (Day number 1260). During this period the source showed evidences of continuous “ringing” flares (Vilhu and Nevalainen 1998; Yadav et al. 1999), with time scales of 30 – 60 s and these will not be evident in the ASM data. The variability increased again from June end (Day number 1275) and the flaring state continued for about 20 more days with the ASM variability being 30 - 40%. The source reached a steady state with low variability ($``$10%). The ringing flares started again in the beginning of 1997 August (Yadav et al. 1999) and towards the end of this state the peculiar outbursts accompanied by infrared flares were observed on 1997 August 14-15 (Eikenberry et al. 1998). The source was relatively quiet thereafter for about 10 days and then went into a flaring state. On September 9 the peculiar “outburst” was seen again (Markwardt et al. 1999).
For a given accreting system, it is reasonable to assume that the long term changes in the X-ray emission is due to changes in the accretion rate. Unlike Cygnus X-1 where the source transition typically takes place in a few days, the state transition in GRS 1915+105 took about three months. If we attribute the state transition as due to a change in the accretion rate near some critical accretion rate (as has been done in Chakrabarti & Titarchuk 1995), the slow state transition can be used to study the system in detail at this critical rate. In this scenario the variety of bursts and variability seen in GRS 1915+105 can be attributed to phenomena near this critical state. Here we address the question of the fast state transition that has occurred in 1997 June 18. From Figure 1 we can see that the bursting properties systematically changed during the transition. In Yadav et al. (1999) (see their Figure 2) it was shown that for about a month the source was in a burst mode with a slow transition from regular bursts to irregular bursts and then again to a regular burst of shorter duration. In Yadav et al. (1999) it was postulated that the irregular bursts are manifestations of rapid state changes. To quantify the properties of the spectral states we have taken two RXTE observations each in the low-hard and high-soft states. In the low-hard state the observation carried out on 1997 March 26 (towards the end of the low-hard state) and another on 1997 May 8 (in the beginning of state transition) were selected. In the high-soft state observations carried out on 1997 July 20 and 1997 August 19 were chosen. The 1997 May 8 and 1997 August 19 observations are the same that are used in Yadav et al. (1999). The X-ray emission properties of the source in the spectral states as deduced from these observations are compared with that obtained during the irregular bursts observed on 1997 June 18. Some salient features of the selected observations are given in Table 1. We use the standard 2 mode data which gives 128 channel energy spectra every 16 s (used for the spectral analysis) and 2 channel binned data below 13 keV every 0.125 ms (used for the timing analysis).
### 2.1. The low-hard and the high-soft states of GRS 1915+105
The light curves of the selected low-hard and high-soft states of the source are shown in Figure 2, with a time resolution of 1 s. The data is for all the five Proportional Counter Units (PCUs) and the energy range is 2 $``$ 13 keV. The average count rate during the low-hard state of March 26 is 3484 s<sup>-1</sup> and it increased to 5600 s<sup>-1</sup> for the May 8 data. During the high-soft states of July 20 and August 19, the observed count rates were 18330 s<sup>-1</sup> and 21669 s<sup>-1</sup>, respectively.
Several features of the state changes in black hole candidates are discernible from the light curve itself. On March 26 (when the count rate was low) there was higher variability in the 1 $``$ 10 s time scale but the source was stable at longer time scales. The situation was opposite in the high state of August 19: there is large variability at longer time scale. On May 8 the count rate increased by about 50%, but the variability characteristics were essentially the same. On July 20, though the count rate was similar to that of August 19, the light curve appears similar to that seen during the low state, with several additional spikes at a few tens of seconds. The behavior of the source is similar to the Galactic black hole candidate source Cygnus X-1 (see Rao et al 1998) when the source showed essentially similar pattern of changes.
To quantify the timing characteristics, we have obtained the power density spectrum (PDS) for these four observations and these are displayed in Figure 3. The low-hard state PDS displays all the characteristics of PDS generally observed for other Galactic black hole sources during the low-hard state (see Rao et al. 1998 and references therein): a flat spectrum up to a break frequency in the vicinity of a few Hz and spectrum falling steeply thereafter. The narrow quasi-periodic oscillation (QPO) along with its first harmonic, seen during the other low-hard states of GRS 1915+105 (Paul et al. 1997; Morgan et al. 1997), is evident in the PDS. In the high state the PDS is a featureless power-law above 0.2 Hz. In the high state of July 20, however, there is a low frequency QPO at 0.13 Hz. The salient features of the PDS in the two states are given in Table 2. To highlight the differences in the PDS at the low and high frequencies, we have fitted the PDS with a power-law and a Lorentzian (whenever a QPO is present) in the frequency ranges 0.1 $``$ 1 Hz and 1 $``$ 10 Hz. For the high state data a single power-law was fitted in the entire frequency range. The distinguishing features of the spectral states are 1) a flat spectrum below 1 Hz for the low-hard state with the power-law index ranging from -0.5 to -0.2 compared to an index $``$ -1.3 in the high-soft state, 2) presence of a narrow 0.5 – 10 Hz QPO feature in the low-hard state and 3) higher variability ($``$15%) in the 1 $``$ 10 Hz range in the low-hard state compared to the low variability ($`<`$5%) in the same frequency range for the high-soft state.
The other distinguishing characteristics of the intensity states of Galactic black hole candidate sources is the energy spectra. The low state is dominated by a thermal-Compton spectrum (which can be approximated to a power-law at lower energies) along with a blackbody emission component, which generally is modeled as a disk blackbody emission. The disk blackbody emission increases in intensity and dominates the spectrum in the high state.
We have generated the 128 channel energy spectra from the Standard 2 mode of the PCA for each of the above observations. Standard procedures for data selection, background estimation and response matrix generation have been applied. To avoid the extra systematic errors in the response matrix of PCA, we have restricted our analysis to the energy range of 3 $``$ 26 keV. Data from all the PCUs are added together. We have fitted the energy spectrum of the source using a model consisting of disk-blackbody and power-law with absorption by intervening cold material parameterized as equivalent Hydrogen column density, N<sub>H</sub>. The value of N<sub>H</sub> has been kept fixed at 6 $`\times `$ 10<sup>22</sup> cm<sup>-2</sup>. We have included a Gaussian line near the expected K<sub>α</sub> emission from iron and absorption edge due to iron. These features help to mimic the reflection spectrum usually found in other Galactic black hole candidate sources like Cygnus X-1 (Gierlinski et al. 1997). Systematic errors of 1% have been added to the data. We obtain reduced $`\chi ^2`$ values in the range of 1 – 3.
The unfolded spectrum for the two spectral states are shown in Figure 4. The individual model components are shown separately. Scales for the axes are kept the same for all the plots. The derived parameters are given in Table 3. The quoted errors are nominal 90% confidence levels obtained by the condition of $`\chi _{min}^2`$ \+ 2.7. The inner disk radius is calculated using an inclination angle of 70 for the accretion disk and a distance of 10 kpc to the source (see Muno et al. 1999). The iron line was found to have an equivalent width of $``$100 eV.
We can identify the following distinguishing features in the energy spectrum in the two spectral states of the source: 1) in the low-hard state the disk blackbody component has lower temperature ($`<`$1 keV) and larger inner disk radius ($`>`$50 km) compared to the high-soft state, which has the inner disk temperature of $``$2 keV and inner disk radius of $``$20 km. 2) the disk blackbody component has a 3 $``$ 26 keV flux of $``$10<sup>-9</sup> erg cm<sup>-2</sup> s<sup>-1</sup> ($`<`$10% of the total flux) whereas the disk blackbody flux increases by a factor of more than 30 in th high-soft state. In fact it becomes the predominant component with $`>`$65% of the observed flux being in this component. 3) The power-law index becomes noticeably steeper in the high-soft state.
### 2.2. Rapid state transitions
Now we show that during the irregular bursts observed on 1997 June 18 the source GRS 1915+105 made rapid intensity transitions between two levels and the X-ray emission properties in these two levels are identical to the spectral states of the source. Belloni et al. (1997b) have shown that during the irregular bursts observed on 1997 June 18 the source exhibited a repeating pattern of intensity states characterized by well defined spectral states. We have selected one irregular burst from this observation for the timing and spectral analysis. The power density spectrum (PDS) were separately obtained for the burst and quiescent time. The results are shown in Figure 5, right panel, separately for quiescent data and the burst data. The results are also given in Table 2. The results of the spectral analysis is shown in the left panels of the same figure and the derived parameters are given in Table 3. The scales for figures for the PDS and the energy spectra are identical to Figure 3 and Figure 4, respectively.
The remarkable similarity between the low-hard state and burst quiescence is evident from the figure. They show a power-law spectral index (between 0.1 and 1 Hz) of -0.3 and -0.5 respectively. A QPO feature at $``$4 Hz is seen during the low-hard state and there is an indication of a narrow QPO feature at 3.2 Hz during the burst quiescence. The observed PDS characteristics are very similar during the high-soft state of 1997 August 19 and the burst time observations of 1997 June 18. The energy spectra, too, highlights the similarity between the spectral state changes and the bursts (see Table 3 and Figure 5). All the identifying features of the spectral states like the changes in the disk blackbody temperature, radius and flux are evident in the burst time too.
These observations strongly support the suggestion, made in Yadav et al. (1999), that the source makes rapid transition between the low-hard state and high-soft state in very short time scales. The difference in the PDS is predominantly above 1 Hz and hence to dramatize our findings we have plotted the observed sub-second variability in these two states. The sub-second variability is characterized as the rms (root mean square) normalized to mean for 10 bins of 0.1 s duration. The contribution from counting statistics has been subtracted. In Figure 6, the observed count rates and the variability are plotted. The irregular burst described earlier is shown in Figure 6a (upper-left panel) and the corresponding variability is shown in Figure 6b (lower-left panel). The rms variability makes a transition when the source count rate changes. The observed count rates during the low-hard state (1997 May 8 observations) and the high-soft state (1997 August 19 observations, shown as stars) are shown in Figure 6c. The variability values for these two states are plotted in Figure 6d, with the high-soft state shown as stars. The similarity between the burst quiescence and the low-hard state on the one hand and the burst and the high-soft state is very striking. The total count rate (in 2 $``$ 13 keV range) is 3000 s<sup>-1</sup> and the variability is 9.3% during the burst quiescence, which compares well with the values of 5400 s<sup>-1</sup> and 12.9%, respectively, observed during the low-hard state of 1997 May. The source had a count rate of 19000 s<sup>-1</sup> and variability of 3.4% during the high-soft state of 1997 August, which is distinctly different from the values found during the low-hard state but very similar to the values of 25000 s<sup>-1</sup> and 2.9%, respectively, found during the burst.
To highlight the spectral similarities we have shown the spectral ratios in Figure 7. Four spectral files are generated. Quiescent data for the irregular burst were accumulated in a spectral file, referred to as B<sub>Q</sub>, for burst quiescence. Similarly burst data for the same burst is accumulated in the file B<sub>H</sub> (burst-high). Spectral data obtained on 1997 May 8 (during the low-hard state of the source) are accumulated in the file L<sub>H</sub> and the high-soft state data (1997 August 19) was accumulated in the file H<sub>S</sub>.
The ratio of B<sub>Q</sub> to L<sub>H</sub>, as a function of energy, is plotted in Figure 7a, and B<sub>H</sub> to H<sub>S</sub> ratio is shown in Figure 7b. All the spectra are normalized to the observed values at 10 keV. It can be seen that the spectra during the burst quiescence is very similar to that during the low-hard state of the source and burst spectrum is very similar to that seen during the high-soft state of the source, within a factor of two. To emphasize the spectral change during the burst the spectral ratio seen during the burst (B<sub>Q</sub> to B<sub>H</sub>) is plotted in Figure 7c and, for comparison, the ratio of the low-hard to high-soft spectra are plotted in Figure 7d. It can be seen that the burst quiescence spectrum is very different from the burst spectrum characterized by a hard component above 10 keV. This type of a hard component is evident in the low-hard state spectrum as compared to the high-soft state (Figure 7d). Hence we conclude that during an irregular burst the source shows a spectral change which is quite similar to the spectral change seen during the transition from a low-hard state to a high-soft state.
## 3. DISCUSSION
Though GRS 1915+105 defies a neat classification between low-hard and high-soft state as seen in other Galactic black hole candidate sources, it does exhibit extended low-hard states (Greiner et al. (1998); Morgan et al. (1997)). Outside such extended low-hard states the source can be further sub-divided into several states namely bright state (high-soft state), chaotic state and flaring state (Morgan et al. (1997)). We have shown here that in at least one type of flaring state when the source was showing irregular bursts, the source traverses between low-hard state and the high-soft state.
The PDS presented here during the burst quiescence and the low-hard state are similar to that seen during the extended low-hard state of 1996 July-August (Morgan et al. (1997); Paul et al. (1997)). The PDS is flat up to a few Hz with evidence of QPO in the range of 0.5 to 5 Hz. The average power in the range of 0.1 to 1 Hz presented here is 0.01 (rms/mean)<sup>2</sup> Hz<sup>-1</sup> which is consistent with the earlier observations. The PDS during the burst and the high-soft state are similar to that seen during the bright state of 1996 April (Morgan et al. (1997)) with the PDS being very steep between 0.1 and 5 Hz.
We emphasize here that the high-soft and low-hard states of the Galactic black hole candidate sources are generally identified only from the strength and shape of the spectra. Since we see spectral changes in a very short time we can conclude that the canonical states too change in such time scales. The timing characteristics like the shape of PDS, QPO, and rms variations give additional support to this. This reinforces the conclusion drawn earlier (Yadav et al. 1999) that the source can make very rapid state transitions.
There have been attempts to explain the rapid variability seen in GRS 1915+105 using disk instability models. Taam et al. (1997) have used scaling laws for viscosity and found that the instability time scales of the inner accretion disk can explain the time scales of the regular bursts seen in GRS 1915+105. Belloni et al. (1997b) have tried to explain the repeated patterns as due to the appearance and disappearance of inner accretion disk. Nayakshin, Rappaport, & Melia (2000) have investigated the different accretion models and viscosity prescriptions and attempted to explain the temporal behavior of GRS 1915+105. In particular, they have shown that the accretion instability in a slim disk, as invoked by Taam et al. (1997), is not likely to adequately account for the behavior of GRS 1915+105 (because of the difficulty in maintaining the high state for long) and the appearance of inner accretion disk, as postulated by Belloni et al. (1997), will take a time scale comparable to the burst time scale. Though Nayakshin et al. (2000) were able to reproduce many characteristics in the X-ray variability for GRS 1915+105, they were unable to explain the rise/fall times or the $`<`$ 10 s oscillations.
The results presented in this paper, namely, the occurrence of fast changes in the luminosity state, put additional constraints to any disk instability model. The fact that the spectral states are observed for long durations show that these states are stable solutions for an accretion disk rather than changes in any stability parameters. Such stable solutions changing in a short time scale is difficult to explain in terms of the disappearance of slim disks. Hence in Yadav et al. (1999) it was postulated that these changes are due to appearance and disappearance of sub-Keplerian disks above the slim disks.
Recently, Chakrabarti (1999) have presented a possible solution to the state transition based on a Two Component Accretion Flow (TCAF) model given by Chakrabarti & Titarchuk (1995). They have invoked the initiation of mass outflow (see also Das & Chakrabarti 1999) as the cause of the start of the state change and the catastrophic Compton cooling of the material in a spherical volume causing the reverting back to the high state. This model essentially reinforces the suggestion of Yadav et al. (1999) but gives a physical basis for the start of the event. In a later work, Chakrabarti & Manickam (2000) have expanded this model further and derived a correlation between the burst time and the QPO frequency. It would be interesting to make a detailed time resolved X-ray spectroscopy during a regular burst to find the slow evolution of the optical depth of the Compton cloud.
Since the canonical spectral states of black hole binaries remain for considerably long duration (up to years) in the same state with similar properties, it is usually assumed that they are some stable solutions for the accretion disk. It is normally believed that for a given source the total accretion rate is the governing parameter for the spectral changes. Our finding that the source makes very fast repeated state transitions, implies that the two solutions for the accretion disk (corresponding to the two spectral states) must exist for roughly the same total mass accretion rate because the time scale of the state change ($``$10s) observed during the irregular bursts is much smaller than the time scale for the readjustment of the accretion disk for any global mass accretion rate changes. The time scale for the latter must be of the order of thousand seconds because it is normally believed that far away from the compact object the accretion disk can be described as the classical thin disk which has a very large viscous time scale. Hence two solutions for the accretion disk must exist for the same global accretion rate.
Such behavior of fast spectral changes must be occurring in other black hole binaries where a wide range of X-ray luminosities are observed, like GRO J1655-40. Remillard et al. (1999) have pointed out the similarity between the temporal properties of GRS 1915+105 and GRO J1655-40 when each source teeters between relative stability and a state of intense oscillations. It will be very interesting to find long duration irregular bursts in GRO J1655-40 because in this source the binary period and the mass of the black hole are known and hence it will prove an ideal ground to further refine the TCAF model.
## 4. CONCLUSION
We have presented the X-ray spectral and time variability characteristics of GRS 1915+105 when the source was exhibiting irregular bursts. The spectral characteristics strongly suggest the view that the source makes a state transition in a very fast time scale (a few seconds). We have found that the temporal variabilities also support this conclusion. We conclude that the source can make spectral state transition only if the advective and standard thin disk co-exist, as suggested by Chakrabarti & Titarchuk (1995).
This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center Online Service, provided by the NASA/Goddard Space Flight Center.
|
warning/0007/hep-ph0007235.html
|
ar5iv
|
text
|
# Running of Axial Coupling Constant
## Abstract
We illustrate that both of massless pseudoscalar meson fields and constitutent quark fields are dynamical field degrees of freedom in energy region between chiral symmetry spontaneously broken scale and quark confinement scale. This physical configuration yields a renormalization running axial coupling constant $`g__A(\mu )`$. The one-loop renormalization prediction $`g__A(\mu =m__N)=0.75`$ and $`g__A(\mu =m_\pi )=0.96`$ agree with $`\beta `$ decay of neutron and $`\pi ^0\gamma \gamma `$ respectively. We also calculate $`\rho ^\pm \gamma \pi `$ and $`\omega \gamma \pi `$ decays up to the next to leading order of $`N_c^1`$ expansion and including all order effects of vector meson momentum expansion. The results strongly support prediction of renormalization running, $`g__A(\mu =m_\rho )=0.77`$.
14.40.-n,12.39.-x,11.40.Ha,13.20.Jf
The asymptotic freedom, chiral symmetry spontanously broken(CSSB) and quark confinement(QC) are essential features of QCD, which are related to the non-Abelian guage symmetry structure of QCD. They divide QCD-description of strong interaction into three different energy regions. It is well-known that the dynamical field degrees of freedom in the energy region above CSSB scale are current quarks and gluons, and in the energy region blow QC scale are pseudoscalar meson fields purely. In the energy region between CSSB scale and QC scale, the CSSB implies that the dynamical field degrees of freedom should be constituent quarks(quasi-particle of quarks), gluons and Goldstone bosons(massless pseudoscalar meson fields) associated with CSSB. Nevertheless, it is still debated whether pseudoscalar mesons are double counted when they are regarded as fundamental fields, as well as light bound states of quarks. Theoretically, it is clear that there is not double counting problem, since fundamental pseudoscalar mesons are light bound states of light current quarks instead of heavy constituent quarks. Practically, however, due to lack of essential evidences to confirm this point, the pseudoscalar mesons are still regared as composited fields of constituent quarks in many works of low energy QCD in the literatures. The purpose of this present paper is to provide an evidence to confirm that pseudoscalar mesons are fundamental dynamical degrees of freedom in the energy region below CSSB scale indeed.
Our analysis is based on to solve a puzzle which also lies low energy strong interaction. In QCD below CSSB scale, the axial weak current with light flavors can be defined with a non-unit coupling constant,
$`j_{5\mu }^a=g__A\overline{\psi }\gamma _\mu \gamma _5{\displaystyle \frac{\tau ^a}{2}}\psi ,(a=1,2,3),`$
where $`\overline{\psi }=(\overline{u},\overline{d})`$ and $`\tau ^a`$ are Pauli matrices. If we recognize that the wavefunction of nucleon has $`SU(3)_{\mathrm{color}}\times SU(2)_{\mathrm{flavor}}`$ symmetry, the beta decay of neutron requires $`g__A=0.75`$. Meanwhile, the PCAC(partially conserved axial current) hypothesis and Adler-Bell-Jackiw anomaly imply axial current divergence
$`^\mu j_{5\mu }^{(3)}={\displaystyle \frac{f_\pi }{2}}m_\pi ^2\pi ^0+{\displaystyle \frac{N_c}{24\pi }}g__A\alpha _{\mathrm{e}.\mathrm{m}.}ϵ^{\mu \nu \alpha \beta }F_{\mu \nu }F_{\alpha \beta },`$ (1)
where $`f_\pi =185.2`$MeV is the pion decay constant and $`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu `$ is photon field strength. Thus in chiral limit, the transition matrix element for the $`\pi ^0\gamma \gamma `$ reduces to
$`<\gamma (ϵ_1,k_1),\gamma (ϵ_2,k_2)|\pi ^0(q)>=(2\pi )^4\delta ^4(qk_1k_2)(i){\displaystyle \frac{2N_c}{3\pi f_\pi }}g__A\alpha _{\mathrm{e}.\mathrm{m}.}ϵ^{\mu \nu \alpha \beta }ϵ_{1\mu }k_{1\nu }ϵ_{2\alpha }k_{2\beta }.`$ (2)
The experimental data for $`\mathrm{\Gamma }(\pi ^0\gamma \gamma )`$ yields axial coupling constant $`g__A=1`$ for $`N_c=3`$. This value is very different from one determined by $`\beta `$ decay of neutron. Then how to resolve this puzzle is a open question. In this paper, we will provide a possible explanation on this problem in the framework of constituent quark model. In this interpretation, the psudoscalar mesons must be regarded as independent dynamical field degrees of freedom in this energy region. The renormalization of pseudoscalar-constituent quarks loop effects will lead to aixal coupling constant $`g__A`$ running with renormalization scale. Our results for one-loop renormalization are that taking $`g__A(\mu =m_N)=0.75`$ as input, we predict $`g__A(\mu =m_\pi )=0.96`$ which matchs with requirement of $`\pi ^0\gamma \gamma `$ decay well. This result is an evdience that pseudoscalar mesons are fundamental fields in energy scale below CSSB. In order to confirm this conclusion further, we will calculate the anomal decays $`\rho ^\pm \gamma \pi ^\pm `$ and $`\omega \gamma \pi ^0`$. In this type of decays, effects of $`g__A`$ are the leading order. Our calculations will be up to the next to leading order of $`N_c^1`$ expansion and including all order effects of vector meson momentum expansion. The results strongly support prediction of one-loop renormalization for axial coupling constant, $`g__A(\mu =m_\rho )=0.77`$.
The simplest version of chiral quark model was originated by Weinberg, and developed by Manohar and Georgi provides a QCD-inspired description on the constituent quark model. At chiral limit, it is parameterized by the following $`SU(3)__V`$ invariant chiral constituent quark lagrangian
$`_\chi `$ $`=`$ $`i\overline{\psi }(/+/\mathrm{\Gamma }+g__A/\mathrm{\Delta }\gamma _5)\psi m\overline{\psi }\psi +{\displaystyle \frac{F^2}{16}}Tr_f\{_\mu U^\mu U^{}\}.`$ (3)
Here $`Tr_f`$ denotes trace in SU(3) flavour space, $`\overline{\psi }=(\overline{u},\overline{d},\overline{s})`$ are constituent quark fields. The $`\mathrm{\Delta }_\mu `$ and $`\mathrm{\Gamma }_\mu `$ are defined as follows,
$`\mathrm{\Delta }_\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}\{\xi ^{}(_\mu ir_\mu )\xi \xi (_\mu il_\mu )\xi ^{}\},`$ (4)
$`\mathrm{\Gamma }_\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}\{\xi ^{}(_\mu ir_\mu )\xi +\xi (_\mu il_\mu )\xi ^{}\},`$ (5)
and covariant derivative are defined as follows
$`_\mu U`$ $`=`$ $`_\mu Uir_\mu U+iUl_\mu =2\xi \mathrm{\Delta }_\mu \xi ,`$ (6)
$`_\mu U^{}`$ $`=`$ $`_\mu U^{}il_\mu U^{}+iU^{}r_\mu =2\xi ^{}\mathrm{\Delta }\xi ^{},`$ (7)
where $`l_\mu =v_\mu +a_\mu `$ and $`r_\mu =v_\mu a_\mu `$ are linear combinations of external vector field $`v_\mu `$ and axial-vector field $`a_\mu `$, $`\xi `$ associates with non-linear realization of spontanoeusly broken global chiral symmetry $`G=SU(3)_L\times SU(3)_R`$ introduced by Weinberg,
$$\xi (\mathrm{\Phi })g_R\xi (\mathrm{\Phi })h^{}(\mathrm{\Phi })=h(\mathrm{\Phi })\xi (\mathrm{\Phi })g_L^{},g_L,g_RG,h(\mathrm{\Phi })H=SU(3)__V.$$
(8)
Explicit form of $`\xi (\mathrm{\Phi })`$ is usual taken
$$\xi (\mathrm{\Phi })=\mathrm{exp}\{i\lambda ^a\mathrm{\Phi }^a(x)/2\},U(\mathrm{\Phi })=\xi ^2(\mathrm{\Phi }),$$
(9)
where the Goldstone boson $`\mathrm{\Phi }^a`$ are treated as pseudoscalar meson octet. The constituent quark fields transform as matter fields of SU(3)$`__V`$,
$$\psi h(\mathrm{\Phi })\psi ,\overline{\psi }\overline{\psi }h^{}(\mathrm{\Phi }).$$
(10)
$`\mathrm{\Delta }_\mu `$ is SU(3)$`__V`$ invariant field gradients and $`\mathrm{\Gamma }_\mu `$ transforms as field connection of SU(3)$`__V`$
$$\mathrm{\Delta }_\mu h(\mathrm{\Phi })\mathrm{\Delta }_\mu h^{}(\mathrm{\Phi }),\mathrm{\Gamma }_\mu h(\mathrm{\Phi })\mathrm{\Gamma }_\mu h^{}(\mathrm{\Phi })+h(\mathrm{\Phi })_\mu h^{}(\mathrm{\Phi }).$$
(11)
Thus the lagrangian( 3) is invariant under $`G_{\mathrm{global}}\times G_{\mathrm{local}}`$.
Let us consider one-loop effects renormalization of lagrangian (3) It should be pointed out that, this model as low energy effective model of QCD is not completely renormalizable. The reason is that more and more new divergent terms will apear when loop effects are included. Fortunately, the terms in lagrangian (3) are still renormalizable.. In terms of defining the following “renormalization” quantities,
$$\begin{array}{c}\psi \psi __R=Z_\psi ^{1/2}\psi ,g__Ag_{_{RA}}=Z_gg__A,\\ mm__R=m+\delta m,FF__R=Z__F^{1/2}F,\\ \xi (\mathrm{\Phi })\xi __R(\mathrm{\Phi })=Z_\xi ^{1/2}\xi (\mathrm{\Phi })\{\begin{array}{c}\mathrm{\Gamma }_{\mu R}=Z_\xi ^1\mathrm{\Gamma }_\mu ,\\ \mathrm{\Delta }_{\mu R}=Z_\xi ^1\mathrm{\Delta }_\mu ,\\ U__R=Z_\xi ^1U,\end{array}\end{array}$$
(12)
the divergences yielded by loop-effects of constituent quarks and massless Goldstone fields can be cancelled by divergent constants $`Z_\psi ,Z_g,\delta m,Z__F`$ and $`Z_\xi `$. In this paper, we focus our attention on renormalization of axial constant $`g__A`$. The diagrams in fig. 1 concern our purpose. In $`\overline{\mathrm{MS}}`$ scheme, explicit calculation gives
$`Z_\psi `$ $`=`$ $`1{\displaystyle \frac{2m__R^2}{\mathrm{\Lambda }_\chi ^2}}g_{_{RA}}^2N_ϵ,Z_\psi \delta m={\displaystyle \frac{2m__R^2}{3\mathrm{\Lambda }_\chi ^2}}g_{_{RA}}^2N_ϵ,`$ (13)
$`Z_\psi Z_\xi `$ $`=`$ $`1{\displaystyle \frac{m__R^2}{4\mathrm{\Lambda }_\chi ^2}}g_{_{RA}}^2N_ϵ,Z_\psi Z_\xi Z_g^1=1{\displaystyle \frac{m__R^2}{\mathrm{\Lambda }_\chi ^2}}({\displaystyle \frac{3}{2}}+{\displaystyle \frac{5}{12}}g_{_{RA}}^2)N_ϵ,`$ (14)
where $`\mathrm{\Lambda }_\chi =2\pi f_\pi 1.2`$GeV is CSSB scale and
$`N_ϵ=\underset{ϵ0}{lim}{\displaystyle \frac{2}{ϵ}}\gamma __E+\mathrm{ln}4\pi .`$ (15)
Then renomalization lagrangian is written as follow
$`__R`$ $`=`$ $`i\overline{\psi }__R(/+/\mathrm{\Gamma }__R)\psi __R\{1{\displaystyle \frac{2m__R^2}{3\mathrm{\Lambda }_\chi ^2}}g_{_{RA}}^2\mathrm{ln}{\displaystyle \frac{\mu ^2}{m^2}}+O({\displaystyle \frac{m__R^4}{\mathrm{\Lambda }_\chi ^4}})\}m__R\overline{\psi }__R\psi __R`$ (17)
$`+\{1+{\displaystyle \frac{m__R^2}{\mathrm{\Lambda }_\chi ^2}}({\displaystyle \frac{3}{2}}{\displaystyle \frac{4}{3}}g_{_{RA}}^2)\mathrm{ln}{\displaystyle \frac{\mu ^2}{m^2}}+O({\displaystyle \frac{m__R^4}{\mathrm{\Lambda }_\chi ^4}})\}g_{_{RA}}\overline{\psi }__Ri/\mathrm{\Delta }\gamma _5\psi __R+{\displaystyle \frac{F__R^2}{16}}Tr_f\{_\mu U__R^\mu U__R^{}\},`$
where $`\mu `$ is renomalization scale. The $`\mu `$-independence of renormalization requires the constituent quark mass $`m__R`$ and axial constant $`g_{_{RA}}`$ should be $`\mu `$-dependent, and should satisfy(for sake of convenience, we omit the subscript “$`R`$” in $`m__R`$ and $`g_{_{RA}}`$ hereafter)
$`{\displaystyle \frac{m(\mu ^{})}{m(\mu )}}`$ $`=`$ $`1{\displaystyle \frac{2\overline{m}^2}{3\mathrm{\Lambda }_\chi ^2}}\overline{g}__A^2\mathrm{ln}{\displaystyle \frac{\mu ^2}{\mu ^2}}+O({\displaystyle \frac{m^4}{\mathrm{\Lambda }_\chi ^4}}),`$ (18)
$`{\displaystyle \frac{g__A(\mu ^{})}{g__A(\mu )}}`$ $`=`$ $`1+{\displaystyle \frac{\overline{m}^2}{\mathrm{\Lambda }_\chi ^2}}({\displaystyle \frac{3}{2}}{\displaystyle \frac{4}{3}}\overline{g}__A^2)\mathrm{ln}{\displaystyle \frac{\mu ^2}{\mu ^2}}+O({\displaystyle \frac{m^4}{\mathrm{\Lambda }_\chi ^4}}),`$ (19)
where
$`\overline{m}={\displaystyle \frac{m(\mu )+m(\mu ^{})}{2}},\overline{g}__A={\displaystyle \frac{g__A(\mu )+g__A(\mu ^{})}{2}}.`$ (20)
Then inputting $`m(\mu =m_\rho )=480`$MeV <sup>§</sup><sup>§</sup>§The $`O(p^4)`$ chiral coupling constant $`L_5=(\frac{3}{8}g^2\frac{N_c}{16\pi ^2})\frac{m}{2B_0}`$ is predicted in this model. Here $`g`$ and $`B_0`$ are two constants absorbing divergence from quark loops. In particular, $`g=\pi ^1`$ and $`B_0=m_\pi ^2/(m_u+m_d)2`$GeV is fitted at $`\mu =m_\rho `$. Thus $`m=480`$MeV is fitted by $`L_5(\mu =m_\rho )=(1.4\pm 0.5)\times 10^3`$. and $`g__A(\mu =m_N)=0.75`$, we have
$`m(\mu =m_N)`$ $`=`$ $`494\mathrm{M}\mathrm{e}\mathrm{V},m(\mu =m_\pi )=368\mathrm{M}\mathrm{e}\mathrm{V}`$ (21)
$`g__A(\mu =m_\rho )`$ $`=`$ $`0.77,g__A(\mu =m_\pi )=0.96.`$ (22)
In particular, $`g__A(\mu =m_\pi )=0.96`$ is agree with $`\pi ^0\gamma \gamma `$ requirement.
Several remarks are necessary here. 1) The renormalization running of $`m`$ and $`g__A`$ is rather larger. This point can be easily understood, since there is no a rather small parameter to suppress loop effects in this energy region. This is a feature of low energy QCD. For example, a low energy coupling constant of chiral perturbative theory at $`O(p^4)`$, $`L_5`$, is equal to $`(2.2\pm 0.5)\times 10^3`$ at energy scale $`\mu =m_\eta `$ but $`(1.4\pm 0.5)\times 10^3`$ at energy scale $`\mu =m_\rho `$. 2) If pseudoscalar mesons are not independent dynamical field degrees of freedom, all diagrams in fig.1 will be absent. Then we can not understand why the value axial constant $`g__A`$ should be equal to 0.75 in $`\beta `$ decay of neutron but be equal to 1 in $`\pi ^0\gamma \gamma `$. Therefore, our results in eqs. (18) and (17) is a direct support that massless pseudoscalar mesons as Goldstone bosons associating CSSB are fundamental dynamical degrees of freedom. 3) The loop effects also generate some high derivative terms contributing to axial constant, e.g., $`\overline{\psi }^2/\mathrm{\Delta }\gamma _5\psi `$. However, we expect that this type terms are suppressed by $`p^2/\mathrm{\Lambda }_\chi ^2`$ expansion(where $`p`$ denote four-momentum of axial current). For $`\beta `$ decay of neutron and $`\pi ^0\gamma \gamma `$, the transition momentum are both very small comparing with $`\mathrm{\Lambda }_\chi `$. So the contributions from high derivative terms are very small that it can not explain the difference of axial constant in two reactions. 4) If we inserted Adler-Bell-Jackiw anomal term with unit axial constant into original lagrangian (3), unambiguously, it should cause double counting, since triangle diagrams of constituent quarks also generate Adler-Bell-Jackiw anomally. 5) In lagrangian (3) pseudoscalar mesons are massless. It is consistent with Goldstone theorem for CSSB.
In the following, for checking the result of eq. (18), we will provide a complete calculation on decays $`\rho ^\pm \gamma \pi ^\pm `$ and $`\omega \gamma \pi ^0`$. Since the chiral expansion converges slowly at energy scale of vector meson masses, we have to include high order contribution of chiral expansion into our results. In this paper, our calculation will be up to the next to leading order of $`N_c^1`$ expansion, and including all important effects of momentum expansion.
Due to WCCWZ realization for vector meson resonances, the lagrangian can be easily extended to included the lowest vector meson resonances
$`_\chi `$ $`=`$ $`i\overline{\psi }(/+/\mathrm{\Gamma }+g__A/\mathrm{\Delta }\gamma _5i/V)\psi m\overline{\psi }\psi +{\displaystyle \frac{F^2}{16}}Tr_f\{_\mu U^\mu U^{}\}+{\displaystyle \frac{1}{4}}m_0^2Tr_f\{V_\mu V^\mu \}.`$ (23)
Here $`V_\mu `$ denotes vector meson octet and singlet, or more convenience, due to OZI rule, they are combined into a singlet “nonet” matrix
$$V_\mu (x)=\lambda 𝐕_\mu =\sqrt{2}\left(\begin{array}{ccc}\frac{\rho _\mu ^0}{\sqrt{2}}+\frac{\omega _\mu }{\sqrt{2}}& \rho _\mu ^+& K_\mu ^+\\ \rho _\mu ^{}& \frac{\rho _\mu ^0}{\sqrt{2}}+\frac{\omega _\mu }{\sqrt{2}}& K_\mu ^0\\ K_\mu ^{}& \overline{K}_\mu ^0& \varphi _\mu \end{array}\right).$$
(24)
It transforms homogeneously under SU(3)$`__V`$
$$V_\mu h(\mathrm{\Phi })V_\mu h^{}(\mathrm{\Phi }),$$
(25)
Thus the lagrangian (23) is still invariant under $`G_{\mathrm{global}}\times G_{\mathrm{local}}`$. The effective action describing meson interaction can be obtained via integrating over degrees of freedom of fermions in lagrangian (23)
$$e^{iS_{\mathrm{eff}}}𝒟\overline{q}𝒟qe^{i{\scriptscriptstyle d^4x_\chi (x)}}=<vac,out|in,vac>_{V,\mathrm{\Delta },\mathrm{\Gamma }},$$
(26)
where $`<vac,out|in,vac>_{V,\mathrm{\Delta },\mathrm{\Gamma }}`$ is vacuum expectation value in presence external sources. The above path integral can be performed explicitly, and heat kernal method has been used to regulate the result. However, this method is extremely difficult to compute very high order contributions in practice. This diffculty can be overcomed via calculating one-loop diagrams of constituent quarks directly. This method can capture all high order contributions of the chiral expansion.
In interaction picture, the equation( 26) is rewritten as follow
$`e^{iS_{\mathrm{eff}}}`$ $`=`$ $`<0|𝒯_qe^{i{\scriptscriptstyle d^4x_\chi ^\mathrm{I}(x)}}|0>`$ (27)
$`=`$ $`{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}i{\displaystyle d^4p_1\frac{d^4p_2}{(2\pi )^4}\mathrm{}\frac{d^4p_n}{(2\pi )^4}\stackrel{~}{\mathrm{\Pi }}_n(p_1,\mathrm{},p_n)\delta ^4(p_1p_2\mathrm{}p_n)}`$ (28)
$``$ $`i\mathrm{\Pi }_1(0)+{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}i{\displaystyle \frac{d^4p_1}{(2\pi )^4}\mathrm{}\frac{d^4p_{n1}}{(2\pi )^4}\mathrm{\Pi }_n(p_1,\mathrm{},p_{n1})},`$ (29)
where $`𝒯_q`$ is time-order product of constituent quark fields, $`_\chi ^\mathrm{I}`$ is interaction part of lagrangian( 23), $`\stackrel{~}{\mathrm{\Pi }}_n(p_1,\mathrm{},p_n)`$ is one-loop effects of constituent quarks with $`n`$ external sources, $`p_1,p_2,\mathrm{},p_n`$ are four-momentas of $`n`$ external sources respectively and
$$\mathrm{\Pi }_n(p_1,\mathrm{},p_{n1})=d^4p_n\stackrel{~}{\mathrm{\Pi }}_n(p_1,\mathrm{},p_n)\delta ^4(p_1p_2\mathrm{}p_n).$$
(30)
To get rid of all disconnected diagrams, we have
$`S_{\mathrm{eff}}`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}S_n,`$ (31)
$`S_1`$ $`=`$ $`\mathrm{\Pi }_1(0),`$ (32)
$`S_n`$ $`=`$ $`{\displaystyle \frac{d^4p_1}{(2\pi )^4}\mathrm{}\frac{d^4p_{n1}}{(2\pi )^4}\mathrm{\Pi }_n(p_1,\mathrm{},p_{n1})},(n2).`$ (33)
Hereafter we will call $`S_n`$ as $`n`$-point effective action.
At the leading order of $`N_c^1`$ expansion, the 3-point $`\rho \gamma \pi `$ and $`\omega \gamma \pi `$ effective action is generated by triangle diagram of constituent quarks
$`\mathrm{\Pi }_3(q,k)`$ $`=`$ $`{\displaystyle \frac{N_c}{3\pi ^2gf_\pi }}eg__A{\displaystyle _0^1}𝑑x_1{\displaystyle _0^{1x_1}}𝑑x_2\{1(x_1+x_2)(1x_1x_2){\displaystyle \frac{q^2}{2m^2}}\}^1`$ (35)
$`\times ϵ^{\mu \nu \alpha \beta }q_\alpha k_\beta \{\rho _\mu ^i(q)A_\nu (k)\pi _i(qk)+3\omega _\mu (q)A_\nu (k)\pi ^0(qk)\},`$
where $`g`$ is an universal coupling constant which absorbs nonrenormalizable logarithmic divergence from constituent quark loops. In the above equation, $`g`$ is from vector meson field normalization
$`\rho _\mu ^i{\displaystyle \frac{1}{g}}\rho _\mu ^i,\omega _\mu {\displaystyle \frac{1}{g}}\omega _\mu .`$
In ref., $`g=\pi ^1`$ has been fitted by the first KSRF sum rule.
It is rather complicate to calculate contribution from meson one-loop, i.e., the next to leading order contribution of $`N_c^1`$ expansion. The diagrams in fig.2 concern to our calculation. Here it is requirement of unitarity of $`S`$-matrix to sum over chain-like approximation in fig.2-b).
For obtain the tadpole correction in fig.2-a), at leading order of $`N_c^1`$ expansion, vertices of $`\rho \gamma \pi KK(\eta \eta )`$ and $`\omega \gamma \pi KK(\eta \eta )`$ are needed if we treat pion as massless particle. In scheme of dimensional regularization, only the vertices generated by triangle diagram of constituent quarks yield non-zero contribution,
$`\delta \mathrm{\Pi }_3(q,k)`$ $`=`$ $`{\displaystyle \frac{N_c}{16\pi ^2gf_\pi }}eg__A{\displaystyle _0^1}𝑑x_1{\displaystyle _0^{1x_1}}𝑑x_2\{1(x_1+x_2)(1x_1x_2){\displaystyle \frac{q^2}{2m^2}}\}^1ϵ^{\mu \nu \alpha \beta }q_\alpha k_\beta \pi ^l\phi ^a\phi ^b`$ (39)
$`Tr\{{\displaystyle \frac{1}{3}}(\rho _\mu ^i(q)\lambda ^i+3\omega _\mu (q)+eQA_\mu (q))(\rho _\nu ^j(k)\lambda ^j+3\omega _\nu (k)+eQA_\nu (k))([\lambda ^a,\lambda ^l]\lambda ^b+\lambda ^a[\lambda ^l,\lambda ^b])`$
$`+eA_\mu (q)([\lambda ^a,Q]\lambda ^b+\lambda ^a[Q,\lambda ^b])(\rho _\mu ^i(k)\lambda ^i+3\omega _\mu (k))\pi ^l`$
$`+e(\rho _\mu ^i(q)\lambda ^i+3\omega _\mu (q))A_\nu (k)([\lambda ^a,Q]\lambda ^b+\lambda ^a[Q,\lambda ^b])\pi ^l\},(i,j,l=1,2,3;a,b=4,\mathrm{},8)`$
where $`Q=\mathrm{diag}\{2/3,1/3,1/3\}`$ is charge operator of quark fields, $`\phi ^a`$ denotes $`K`$ or $`\eta `$-meson fields in internal line. For sake of convenience, we can assume that the masses of $`K`$ and $`\eta `$ are degenerated. Then integral over $`\phi `$ fields, we have tadpole correction to $`\rho \gamma \pi `$ and $`\omega \gamma \pi `$ vertices
$`\mathrm{\Pi }_3^{tad}(q,k)={\displaystyle \frac{4}{3}}\zeta \mathrm{\Pi }_3(q,k),\zeta ={\displaystyle \frac{m__K^2}{8\pi ^2f_\pi ^2}}({\displaystyle \frac{4\pi \mu ^2}{m__K^2}})^{ϵ/2}\mathrm{\Gamma }(1{\displaystyle \frac{D}{2}}){\displaystyle \frac{m__K^2}{8\pi ^2f_\pi ^2}}\lambda ,`$ (40)
where $`\mathrm{\Pi }_3(q,k)`$ is defined in eq.( 35), $`\lambda `$ absorb the quadratic divengence from meson loops. $`\lambda 2/3`$ has been fitted by Zweig rule.
Next we will calculate the contribution in fig.2-b). It is equivalent to calculate the diagram in fig.3-a). In other words, we need to an effective VPP(where “V” denotes $`\rho `$ or $`\omega `$ meson, “P” denotes pseudoscalar meson octet) vertex and $`\gamma \pi `$-PP vertex. In fig.3-b) we have shown that the effective VPP vertex includes not only tree level vertex but also meson loop contribution.
At leading order of $`N_c^1`$ expansion, the anomal $`\gamma \pi `$-PP vertex is generated by both of triangle and box diagrams of constituent quarks,
$`\mathrm{\Gamma }_{\gamma 3P}(q_2,k_1,k_2,k_3)={\displaystyle \frac{3i}{8f_\pi ^2}}eB(q^2)ϵ^{\mu \nu \alpha \beta }A_\mu (q_2)k_{1\nu }k_{2\alpha }k_{3\beta }Tr_f\{QP(k_1)P(k_2)P(k_3)\},`$ (41)
$`B(q^2)={\displaystyle \frac{N_c}{3\pi ^2f_\pi }}g__A({\displaystyle _0^1}𝑑x_1{\displaystyle _0^{1x_1}}𝑑x_2\{1{\displaystyle \frac{q^2}{2m^2}}(x_1+x_2)(1x_1x_2)\}^1{\displaystyle \frac{g__A^2}{6}}),`$ (42)
where $`q^2=(k_1+k_2)^2`$. The effective $`\rho `$-PP vertex and $`\omega KK`$ vertex(here we omit those vertices suppressed by isospin conversation) have been derived in ref.
$`\mathrm{\Gamma }_{\rho PP}(q,k_1,k_2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}g_{\rho PP}(q^2)(q^2k_{2\mu }q_\mu qk_2)Tr_f\{\rho ^\mu (q)P(k_1)P(k_2)\},`$ (43)
$`\mathrm{\Gamma }_{\omega KK}(q,k_1,k_2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}g_{\omega KK}(q^2)(q^2k_{2\mu }q_\mu qk_2)\omega ^\mu \{(K^+(k_1)K^{}(k_2)+K^0(k_1)\overline{K}^0(k_2))+(k_1k_2)\},`$ (44)
with
$`g_{\rho PP}(q^2)`$ $`=`$ $`{\displaystyle \frac{A_1(q^2)+g__A^2A_2(q^2)}{gf_\pi ^2(1+2\zeta )(1+\mathrm{\Sigma }(q^2))}},`$ (45)
$`g_{\omega KK}(q^2)`$ $`=`$ $`{\displaystyle \frac{A_1(q^2)+g__A^2A_2(q^2)}{gf_\pi ^2(1+2\zeta )(12\mathrm{\Sigma }_K(q^2))}},`$ (46)
where
$`A_1(q^2)`$ $`=`$ $`g^2{\displaystyle \frac{N_c}{\pi ^2}}{\displaystyle _0^1}𝑑xx(1x)\mathrm{ln}(1{\displaystyle \frac{x(1x)q^2}{m^2}}),`$ (47)
$`A_2(q^2)`$ $`=`$ $`g^2+{\displaystyle \frac{N_c}{2\pi ^2}}{\displaystyle _0^1}dx_1{\displaystyle _0^1}dx_2(x_1x_1^2x_2)\{1+{\displaystyle \frac{m^2}{m^2x_1(1x_1)(1x_2)q^2}}`$ (49)
$`+\mathrm{ln}(1{\displaystyle \frac{x_1(1x_1)(1x_2)q^2}{m^2}})\},`$
$`\mathrm{\Sigma }(q^2)`$ $`=`$ $`\{1+{\displaystyle \frac{q^2(A_1(q^2)+2g__A^2A_2(q^2))}{2f_\pi ^2(1+11\zeta /3)}}\}(4\mathrm{\Sigma }_\pi (q^2)2\mathrm{\Sigma }_K(q^2)),`$ (50)
$`\mathrm{\Sigma }_\pi (q^2)`$ $`=`$ $`{\displaystyle \frac{q^2}{16\pi ^2f_\pi ^2}}\{{\displaystyle \frac{\lambda }{6}}+{\displaystyle _0^1}𝑑tt(1t)\mathrm{ln}{\displaystyle \frac{t(1t)q^2}{m__K^2}}+{\displaystyle \frac{i}{6}}Arg(1)\theta (q^24m_\pi ^2)\},`$ (51)
$`\mathrm{\Sigma }_K(q^2)`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2f_\pi ^2}}\{\lambda (m__K^2{\displaystyle \frac{q^2}{6}})+{\displaystyle _0^1}𝑑t[m__K^2t(1t)q^2]\mathrm{ln}(1{\displaystyle \frac{t(1t)q^2}{m__K^2}})\}.`$ (52)
Then due to soft-pion theorem, the pseudoscalar meson loops in fig.2-b) contribution to $`\rho \gamma \pi `$ and $`\omega \gamma \pi `$ vertices as follow
$`\mathrm{\Pi }_3^{1loop}(q,k)`$ $`=`$ $`eq^2g_{\rho PP}(q^2)[A(q^2)+2A(0)][\mathrm{\Sigma }_\pi (q^2){\displaystyle \frac{1}{2}}\mathrm{\Sigma }_K(q^2)]ϵ^{\mu \nu \alpha \beta }q_\alpha k_\beta \rho _\mu ^i(q)A_\nu (k)\pi _i(qk)`$ (54)
$`+{\displaystyle \frac{1}{2}}eq^2g_{\omega KK}(q^2)[A(q^2)+2A(0)]\mathrm{\Sigma }_K(q^2)ϵ^{\mu \nu \alpha \beta }q_\alpha k_\beta \omega _\mu (q)A_\nu (k)\pi ^0(qk).`$
Here $`\rho \gamma \pi `$ coupling receives contributions from both of pion-loop and $`K`$-loop, but $`\omega \gamma \pi `$ coupling receives dominant contribution from $`K`$-loop only.
Eq.(54) tegother with eqs.(35) and (40) give “complete” $`\rho \gamma \pi `$ and $`\omega \gamma \pi `$ coupling up to the next to leading of $`N_c^1`$ expansion at least,
$`\mathrm{\Pi }_3^c(q,k)`$ $`=`$ $`eϵ^{\mu \nu \alpha \beta }q_\alpha k_\beta \{g_{\rho \gamma \pi }(q^2)\rho _\mu ^i(q)A_\nu (k)\pi _i(qk)+g_{\omega \gamma \pi }(q^2)\omega _\mu (q)A_\nu (k)\pi ^0(qk)\},`$ (55)
$`g_{\rho \gamma \pi }(q^2)`$ $`=`$ $`{\displaystyle \frac{N_c}{3\pi ^2gf_\pi }}g__A(1{\displaystyle \frac{4}{3}}\zeta ){\displaystyle _0^1}𝑑x_1{\displaystyle _0^{1x_1}}𝑑x_2\{1(x_1+x_2)(1x_1x_2){\displaystyle \frac{q^2}{2m^2}}\}^1`$ (57)
$`q^2g_{\rho PP}(q^2)[A(q^2)+2A(0)][\mathrm{\Sigma }_\pi (q^2){\displaystyle \frac{1}{2}}\mathrm{\Sigma }_K(q^2)],`$
$`g_{\omega \gamma \pi }(q^2)`$ $`=`$ $`{\displaystyle \frac{N_c}{\pi ^2gf_\pi }}g__A(1{\displaystyle \frac{4}{3}}\zeta ){\displaystyle _0^1}𝑑x_1{\displaystyle _0^{1x_1}}𝑑x_2\{1(x_1+x_2)(1x_1x_2){\displaystyle \frac{q^2}{2m^2}}\}^1`$ (59)
$`+{\displaystyle \frac{1}{2}}q^2g_{\omega KK}(q^2)[A(q^2)+2A(0)]\mathrm{\Sigma }_K(q^2).`$
For $`g__A(\mu =m_\rho )=0.77`$, the above results yield
$`B(\rho ^\pm \pi ^\pm \gamma )=4.88\times 10^4,B(\omega \pi ^0\gamma )=8.9\%.`$ (60)
These results agree with data $`B(\rho ^\pm \pi ^\pm \gamma )=(4.5\pm 0.5)\times 10^4`$ and $`B(\omega \pi ^0\gamma )=(8.5\pm 0.5)\%`$ well. If we take $`g__A=1`$ as a comparison, the theoretical predictions are
$`B(\rho ^\pm \pi ^\pm \gamma )=8.33\times 10^4,B(\omega \pi ^0\gamma )=15.0\%.`$ (61)
These results obviously disagree with data. Therefore, the theoretical prediction (60) provides an evidence to confirm the result in eq. (18).
To conclude, we discuss the one-loop renormalization in Manohar-Georgi model. This renormalization leads to running of axial coupling constant and constituent quark mass. Inputting $`g__A(\mu =m_N)=0.75`$ which agree with $`\beta `$ decay of neutron, the prediction $`g__A(\mu =m_\pi )=0.96`$ agree with reqiurement of $`\pi ^0\gamma \gamma `$ decay. We also notice that, if massless pseudoscalar meson fields were not independent degrees of freedom in energy between CSSB scale and QC scale, we could not interpret $`\beta `$ decay of neutron and $`\pi ^0\gamma \gamma `$ decay simultaneously. For example, because axial constant is not unit in Extend Nambu-Jona-Lasinio model, the model can not yield right prediction for $`\pi ^0\gamma \gamma `$ decay. The similar problem also exists in the models of ref.. Therefore, our result in this paper can be reagrd as an evidence that pseudoscalar meson fields are independent degrees of freedom in energy between CSSB scale and QC scale.
The renormalization of axial constant also predicts $`g__A(\mu =m_\rho )=0.77`$. For confirming this prediction, we calculate anomal vector meson decays, $`\rho ^\pm \gamma \pi `$ and $`\omega \gamma \pi `$. In this type of decays, effects of $`g__A`$ is leading order. Our results are up to the next to leading order of $`N_c^1`$ expansion and include all order effects of vector meson momentum expansion. The theoretical predictions for branch ratios of these two decays agree with results of renormalization and is against $`g__A=1`$.
|
warning/0007/astro-ph0007059.html
|
ar5iv
|
text
|
# The Initial Mass Function of Low-Mass Stars and Brown Dwarfs in TaurusVisiting Astronomer, Kitt Peak National Observatory, National Optical Astronomy Observatories, which is operated by the Association of Universities for Research in Astronomy, Inc. (AURA) under cooperative agreement with the National Science Foundation.
## 1 Introduction
A great deal of insight into the process of star formation can be gained by measuring the initial mass function (IMF) of low-mass stars and brown dwarfs. The presence of a flattening or turnover in the IMF, the value of such a characteristic mass, the shape of the IMF into the substellar regime, the minimum mass at which objects form in isolation, and the variation of these properties with environmental conditions can provide discriminating tests of the wide range of theories for how the masses of stars are determined.
In an effort to measure accurately and precisely the low-mass IMF under a wide range of conditions, the populations within the nearby star forming clusters IC 348 (Luhman et al. 1998b), $`\rho `$ Oph (Luhman & Rieke 1999), and the Trapezium Cluster (Luhman et al. 2000) were studied in detail. As summarized by Luhman et al. (2000), the stellar IMFs in these clusters were found to be similar over a wide range of stellar densities ($`\rho `$ Oph, $`n=0.2`$-$`1\times 10^3`$ pc<sup>-3</sup>; IC 348, $`n=1\times 10^3`$ pc<sup>-3</sup>; Trapezium, $`n=1`$-$`5\times 10^4`$ pc<sup>-3</sup>), while modest variations in the substellar IMFs were not ruled out. The data for star forming clusters (see also Hillenbrand & Carpenter 2000), young open clusters (Bouvier et al. 1998; Barrado y Navascués et al. 2000), and the field (Reid et al. 1999) were shown to be consistent with the same IMF, one that is flat or rises slowly from the substellar regime to about 0.6 $`M_{}`$, and then rolls over into a power law that continues from about 1 $`M_{}`$ to higher masses, consistent the notion that the majority of field stars were born in clusters (Lada, Strom, & Myers 1993). Meanwhile, a different IMF characterizes Galactic globular clusters (Paresce & De Marchi 2000), one that can be described by a log-normal form with a characteristic mass of 0.33 $`M_{}`$. These studies of star forming and open clusters, the field, and globular clusters represent the most reliable measurements of the low-mass IMF to date. The observed variation in the IMF from the field and Galactic clusters to globular clusters motivates further studies of the low-mass IMF under a greater variety of environments.
Whereas the young stellar populations from the previous studies represent the clustered mode of star formation ($`n=10^2`$-$`10^4`$ pc<sup>-3</sup>), the stars in the Taurus-Auriga molecular cloud are forming in relative isolation (1-10 pc<sup>-3</sup>). Because of the compact nature of star forming clusters, candidate members can be efficiently identified through optical and infrared (IR) imaging and confirmed through multi-object spectroscopy. The completeness limits in mass and reddening of such samples are easily defined with IR photometry, which is essential for a measurement of the IMF (see, e.g., Luhman et al. 2000). In contrast, the population in the Taurus star forming region is spread across a much larger area of sky, dramatically increasing the field star contamination and reducing the observing efficiency. Consequently, unlike the unbiased samples defined for clusters, young stars in Taurus are usually identified through specific selection criteria (e.g., emission in H$`\alpha `$, X-rays, or the far-IR). Such samples are not satisfactory for measurements of the IMF because the biases and mass completeness limits are not readily quantified.
In a few recent studies, representative measurements of the low-mass population in Taurus have been attempted. Strom & Strom (1994) and Luhman & Rieke (1998) (hereafter LR98) used optical and IR imaging and spectroscopy to estimate the IMF in L1495, one of the more well-defined stellar aggregates in Taurus. To search for objects at lower masses over a larger area, Briceño et al. (1998) (hereafter BHSM) obtained $`RI`$ imaging of the L1495, B209, L1529, and L1551 dark clouds. Because of the proximity and youth of Taurus, the locus of members in the BHSM optical color-magnitude diagram was well-separated from most of the background population. BHSM obtained spectroscopy of candidate low-mass members to search for signatures of youth and membership, such as Li absorption, strong H$`\alpha `$ emission, and pre-main-sequence spectral features. Although their observations were sensitive to unreddened young brown dwarfs with masses of $`0.03`$ $`M_{}`$ (M8-M9), they found objects down to only $`0.08`$ $`M_{}`$ (M6-M6.5). However, because the observations were conducted in the optical and the members become redder with later types, the survey was complete to progressively lower levels of extinction with decreasing mass. Consequently, it was unclear whether the absence of brown dwarfs in the BHSM study was statistically significant.
To measure an IMF for Taurus that is representative at masses of $`0.02`$ $`M_{}`$, I have obtained spectroscopy of low-mass candidates identified by LR98 and have searched for brown dwarfs at low masses and moderately high reddenings by combining deep $`I`$ and $`z\mathrm{}`$ imaging of the BHSM fields with IR photometry from the Two-Micron All-Sky Survey (2MASS). The techniques employed by Luhman et al. (2000) and references therein are used to calculate the IMF for this area. The Taurus IMF is found to have the same shape as the Trapezium IMF from 0.1-1 $`M_{}`$, but with moderately fewer stars above 1 $`M_{}`$ and a significant deficit of brown dwarfs. The implications of these results for theories of the IMF are discussed.
## 2 Observations and Data Analysis
### 2.1 Spectroscopy
I selected for IR spectroscopy 16 of the 17 low-mass candidates identified by LR98 in $`JK_\mathrm{s}`$ imaging of a $`10\mathrm{}\times 10\mathrm{}`$ field towards the area of highest extinction in the L1495 dark cloud. Additional spectra were obtained for the dwarf spectral type standards GL 382 (M2V), GL 381 (M2.5V), GL 51 (M5V), LHS 2243 (M8V), LHS 2065 (M9V), BRI 0021-0214 ($``$M9.5), and 2MASP J0345432$`+`$254023 (L0V) and the optically-classified pre-main-sequence sources V410 X-ray 3 (M6), 5a (M5.5), and 6 (M5.5). The observations were performed with the near-IR long-slit spectrometer CRSP at the Kitt Peak 4 meter telescope on the nights of 1998 November 30 and December 1. The 75 l mm<sup>-1</sup> grating was used with the $`1\stackrel{}{\mathrm{.}}9`$ slit, providing coverage of the entire $`K`$-band with a two-pixel resolution of $`R=\lambda /\mathrm{\Delta }\lambda =300`$. Wavelength calibration was performed with HeNeAr lamp spectra. LR 1 and LR 15/V410 Anon 14 were also observed at a spectral resolution of $`R=1200`$ with the IR spectrometer FSpec (Williams et al. 1993) at the Multiple Mirror Telescope (MMT) on Mount Hopkins on 1997 October 14. The remaining observing and data reduction procedures for both instruments were similar to those described by Luhman et al. (1998b). I also obtained optical spectra of LR 10/V410 Anon 3, LR 13/V410 Anon 1, LR 14/V410 Anon 2, LR 15/V410 Anon 14, V410 X-ray 8b, 8d, and 8e, where the latter three stars are candidate counterparts to X-ray sources in the study of L1495 by Strom & Strom (1994). These measurements were made with the Red Channel Spectrograph at the MMT on 1997 November 27-29 and are identical to observations of stars in IC 348 by Luhman et al. (1998b) during the same nights.
### 2.2 Photometry
For optical imaging, I selected the regions observed by BHSM towards the L1495, B209, L1529, and L1551 dark clouds (see their Figure 1). Images of these fields were obtained with the four shooter camera at the Fred Lawrence Whipple Observatory 1.2 m telescope on 1999 October 12 and 13 under photometric conditions. The instrument contained four $`2048\times 2048`$ CCDs separated by $`45\mathrm{}`$ and arranged in a $`2\times 2`$ grid. After binning $`2\times 2`$ during readout, the plate scale was $`0\stackrel{}{\mathrm{.}}67`$ pixel<sup>-1</sup>. Two positions separated by $`40\mathrm{}`$ in right ascension and declination were observed towards L1495, B209, L1529, and contiguous north and south sections of L1551. At each position, images were obtained at $`I`$ ($`\lambda _{eff}8100`$ Å) and $`z\mathrm{}`$ ($`\lambda _{eff}9100`$ Å) with exposure times of 1 and 20 min for each filter. The images were bias subtracted, divided by dome flats, registered, and combined into one image at each band. Image coordinates and photometry were measured with DAOFIND and PHOT under the IRAF package APPHOT. Aperture photometry was extracted with a radius of four pixels. The background level was measured in an annulus around each source and subtracted from the photometry, where the inner radius of the annulus was five pixels and the width was one pixel. The photometry was calibrated in the Cousins $`I`$ system through observations of standards across a range of colors (Landolt 1992). Because the $`I`$ filter for these observations was similar to Cousins $`I`$, the color transformation was small. The $`z\mathrm{}`$ filter is similar to that of the Sloan Digital Sky Survey (Fukugita et al. 1996). The $`z\mathrm{}`$ data were calibrated by the standard star with the most neutral colors ($`VI=0.2`$) and assuming $`Iz\mathrm{}=0`$. This crude calibration is sufficient for this study since the $`z\mathrm{}`$ data are used only in identifying likely background stars and young low-mass candidates through the relative $`Iz\mathrm{}`$ colors. Saturation occurred near $`I12.5`$. The completeness limits were $`I21`$ and $`z\mathrm{}19.5`$ as inferred from a comparison of the data to the number of stars as a function of magnitude in the Galactic models of Bahcall & Soneira (1981). Typical photometric uncertainties are 0.04 mag at $`I=20`$ and $`z\mathrm{}=18.5`$ and 0.1 mag at $`I=21`$ and $`z\mathrm{}=19.5`$. The plate solution was derived from coordinates of sources observed in the 2MASS Spring 1999 Release Point Source Catalog that appeared in the optical images and were not saturated.
The centers and dimensions of the fields from which photometry is extracted are $`(\alpha ,\delta )(2000)=(4^\mathrm{h}18^\mathrm{m}36\stackrel{\mathrm{s}}{\mathrm{.}}65`$, $`28\mathrm{°}23\mathrm{}30\stackrel{}{\mathrm{.}}5`$) and $`23\stackrel{}{\mathrm{.}}42\times 23\stackrel{}{\mathrm{.}}42`$ for L1495, $`(\alpha ,\delta )(2000)=(4^\mathrm{h}14^\mathrm{m}12\stackrel{\mathrm{s}}{\mathrm{.}}15`$, $`28\mathrm{°}10\mathrm{}51\stackrel{}{\mathrm{.}}5`$) and $`23\stackrel{}{\mathrm{.}}42\times 23\stackrel{}{\mathrm{.}}42`$ for B209, $`(\alpha ,\delta )(2000)=(4^\mathrm{h}32^\mathrm{m}31\stackrel{\mathrm{s}}{\mathrm{.}}1`$, $`24\mathrm{°}23\mathrm{}37\stackrel{}{\mathrm{.}}5`$) and $`23\stackrel{}{\mathrm{.}}1\times 22\stackrel{}{\mathrm{.}}5`$ for L1529, and $`(\alpha ,\delta )(2000)=(4^\mathrm{h}31^\mathrm{m}47\stackrel{\mathrm{s}}{\mathrm{.}}8`$, $`18\mathrm{°}09\mathrm{}35\stackrel{}{\mathrm{.}}0`$) and $`23\stackrel{}{\mathrm{.}}18\times 46\stackrel{}{\mathrm{.}}83`$ for L1551. 2MASS photometry is currently unavailable for the northern 70% of the L1551 field (north of $`\delta =18\mathrm{°}00\mathrm{}06\stackrel{}{\mathrm{.}}0`$). Out of nearly 2000 2MASS sources within these four fields, 66 objects were not measured in the optical data because they fell within gaps between the detectors, were contaminated by bad pixels, or were not point sources (e.g., galaxies). These 2MASS sources are excluded from the study. In addition, a known asteroid that was flagged in the 2MASS database was rejected. In the areas covered by 2MASS, stars saturated in the optical data were included in the compilation of sources through their 2MASS photometry and coordinates. But in the portion of L1551 not available in 2MASS, no photometry is available for saturated optical sources, and thus they are omitted. The exceptions are V826 Tau, XZ Tau, HL Tau, V827 Tau, V710 Tau A and B, which are young stars with previously measured photometry. Two additional young stars that fall in the L1551 field and lack 2MASS data are L1551/IRS5 and LkH$`\alpha `$ 358. Because the former is broad and nebulous in the optical data, photometry and coordinates were not measured for it in this work. For these eight stars, IR photometry is taken from Kenyon & Hartmann (1995). The recently discovered low-mass objects MHO 4, MHO 5, MHO 6, MHO 7, and MHO 9 (BHSM) lack 2MASS photometry or other previous IR data. Photometry and coordinates for spectroscopically confirmed young members within the imaged fields of Taurus are listed in Table 1. When available, the coordinates in Table 1 are those measured from the optical data. For the other stars – 2MASS sources that were saturated or too red to be detected in the optical images – the 2MASS coordinates are given.
## 3 Individual Source Characteristics
### 3.1 Spectral Types
The targets in the spectroscopic sample can be foreground stars, background stars, or young low-mass members of Taurus. Foreground M dwarfs are identified by the lack of reddening in their spectra and colors and by the strong absorption in the optical Na and K transitions (Luhman et al. 1998a, 1998b). The star LR 10/V410 Anon 3 exhibits these properties with a spectral type of M5V, which is consistent with the classification as a foreground M4.5V star by Strom & Strom (1994). Other previously known foreground M stars in L1495 include V410 Anon 12 and 17 (Strom & Strom 1994; BHSM). Note that V410 Anon 3 was mistakenly omitted as an alternate designation for LR 10 in Table 2 of LR98. In addition, LR 11 and LR 17/V410 Anon 27 are in fact the same star. As discussed by Strom & Strom (1994), V410 Anon 9 may be a field star close to and behind the L1495 cloud. It is taken as a background star in the remaining analysis. The question of membership of this star has no bearing on the analysis of the IMF because it has a reddening higher than the limit used in defining the IMF sample.
Young late-type members of Taurus are expected to have strong absorption in TiO, VO, and H<sub>2</sub>O at optical and near-IR wavelengths. Thus, they should be easily identified through low-resolution spectroscopy from $`I`$ through $`K`$. On the other hand, background stars are predominantly giants and early-type stars. These stars exhibit featureless spectra at low resolution in the optical and IR, except for giants which have strong CO absorption at 2.3 $`\mu `$m. The following targets from the spectroscopic sample are classified as background stars: V410x8b (RC, C), 8d (RC), 8e (RC, LR98), LR 2 (C), LR 3/V410 Anon 19 (C), LR 4/V410 Anon 21 (C), LR 5-8 (C), LR 9/V410 Anon 26 (C), LR 12 (C), LR 13/V410 Anon 1 (RC), LR 14/V410 Anon 2 (RC), LR 15/V410 Anon14 (BHSM, RC, F), and LR 17/V410 Anon 27 (C). The classifications are from the new data with Red Channel (RC), CRSP (C), and FSpec (F) and from the previous work of BHSM and LR98.
A spectral type of K4-K5 is measured for LR 1 from the FSpec data in the manner described by LR98. This star cannot be a foreground star because of its reddened colors. It is too bright for its spectral type to be a background star (i.e., above the main sequence), therefore it is taken as a young member of Taurus. Spectral types for LR 1 and all other known young sources within the four Taurus fields are listed in Table 1.
### 3.2 Extinctions
Extinctions for the known young stars in the four Taurus fields are now estimated.
In the following analysis, standard dwarf colors are taken from the compilation of Kenyon & Hartmann (1995) for types earlier than M0 and from the young disk populations described by Leggett (1992) for types of M0 and later. The IR colors from Leggett (1992) are on the CIT photometric system. Most of the IR data in Table 1 are from the 2MASS survey, where the photometric system is similar to CIT. The IR colors of the standards and the eight young stars taken from Kenyon & Hartmann (1995) are transformed from Johnson-Glass to CIT (Bessell & Brett 1988) during the analysis, although the colors of those young stars remain in Johnson-Glass in Table 1. Reddenings are calculated with the extinction law of Rieke & Lebofsky (1985).
In the common method for estimating the reddening towards a young star, a color excess is calculated by assuming the intrinsic color, typically that of a standard dwarf at the spectral type in question. To ensure that the color excess reflects only the effect of reddening, contamination from short or long wavelength excess emission is minimized by selecting colors between the $`R`$ and $`H`$ bands. Because $`RI`$ is less susceptible to excess emission than $`JH`$, the former is used for measuring reddening in this study when it is available. The $`RI`$ colors of 35 stars in Table 1 are provided by BHSM, Strom & Strom (1994), and Kenyon & Hartmann (1995). The reddenings computed by BHSM for their late M objects are slightly lower than the values reported here because of the differing references for standard dwarf $`RI`$ colors between the two studies. Although dwarf colors have been assumed here, the intrinsic $`RI`$ colors of young stars do appear to depart from dwarf values for M4-M5 types (Luhman 1999).
Meyer, Calvet, & Hillenbrand (1997) estimated the locus of intrinsic $`JH`$ and $`HK`$ colors for a sample of classical T Tauri stars (CTTS) by dereddening the observed colors with the extinctions measured from $`RI`$. The locus had an origin that fell near the intrinsic colors of an M0 dwarf and it extended to redder $`JH`$ and $`HK`$ with a well-defined slope, a behavior that was reproduced by models of star-disk systems. When the same experiment is repeated for the sources with $`RI`$ in this study, the stars are distributed around the CTTS locus of Meyer et al. (1997) in a similar fashion as in Figure 4 of LR98 for L1495. However, the scatter is larger than that found by Meyer et al. (1997), probably because the sample here is not as homogeneous in spectral type or as well-studied photometrically.
To calculate reddenings for each of the 14 stars in Table 1 that has a spectral type but lacks an $`RI`$ measurement, the $`JH`$ and $`HK`$ colors are dereddened to a CTTS locus where the origin coincides with the dwarf colors for the star’s spectral type. The locus is given the slope measured by Meyer et al. (1997) for M0 stars, which they predicted to remain relatively constant with spectral type. This method assumes that the central stars of young mid- and late-M stars have dwarf-like $`JH`$ and $`HK`$ colors, which is suggested by the work of Luhman (1999). Most of the stars dereddened in this fashion have extinctions that are nearly the same as those derived by simply assuming dwarf colors without including the CTTS locus, i.e., they have little IR excess emission. An exception is PSC04154+2823, which exhibits an excess in $`HK`$ that is larger than expected from a reddened CTTS. To estimate the extinction for this object, the $`JH`$ color is dereddened to a value of 1.1, which is roughly the maximum intrinsic color of the CTTS locus. In the compilation of Kenyon & Hartman (1995), the binary V710 Tau ($`3\stackrel{}{\mathrm{.}}24`$, Leinert et al. 1993) is resolved in the IR photometry but not in the optical data. Because the intrinsic optical colors of the components probably differ significantly given their spectral types, the IR colors are used in measuring the extinctions towards each component. Reddenings and luminosities are not calculated for the three sources lacking spectral types and the class I object L1551/IRS5.
### 3.3 Effective Temperatures and Bolometric Luminosities
For reasons described in previous studies (e.g., Luhman 1999), $`I`$ and $`J`$ are the preferred bands for measuring bolometric luminosities of young low-mass stars. When $`J`$ is available, the luminosities in Table 1 are computed from standard dwarf bolometric corrections (see Luhman 1999), the dereddened $`J`$-band measurements, and a distance modulus of 5.76 (Wichmann et al. 1998). For the five members that lack IR data, luminosities are estimated from the $`I`$-band photometry obtained in this study. Because HL Tau, PSC04154+2823, and MHO 1 have strong excess emission in the near-IR, the luminosities inferred from $`J`$ have large uncertainties. In addition, such sources are often highly variable, as found for PSC04154+2823 when the photometry of this work and BHSM are compared ($`\mathrm{\Delta }I=1.66`$).
Spectral types of M0 and earlier are converted to effective temperatures with the dwarf temperature scale of Schmidt-Kaler (1982). For spectral types later than M0, the adopted temperature scale is that developed by Luhman (1999) for use with the evolutionary models of Baraffe et al. (1998). Detailed discussions of the temperature scales and evolutionary models for young low-mass stars are found in LR98, Luhman (1999), and Luhman et al. (2000).
## 4 The Taurus Stellar Population
### 4.1 Membership and Completeness
#### 4.1.1 Separation of Field Stars and Candidate Low-Mass Members
To distinguish cool, low-mass cluster members from background stars in a color-magnitude diagram, the photometric bands should be selected such that the directions of decreasing mass and increasing extinction differ as much as possible. For example, in a plot of $`JH`$ versus $`H`$, members of a population will move vertically down the diagram with lower mass and cooler temperatures because the near-IR colors of most spectral types fall within a small range of values. Meanwhile, extinction moves stars down and to the right at a relatively shallow angle in this diagram. As a result, cluster members and reddened background stars are mixed together and cannot be separated with IR colors alone. However, in a color-magnitude diagram that includes at least one optical band (e.g., $`RI`$, $`Iz\mathrm{}`$, $`IJ`$), the color increases rapidly with later spectral types and the reddening vector has a steep slope downward. In such a diagram, the intersection of reddened background stars and low-mass members is minimized, allowing for the efficient rejection of most of the field stars.
For the above reasons, BHSM used an optical color-magnitude diagram to search for new low-mass members of Taurus in fields towards L1495, B209, L1529, and L1551. At stellar masses, spectroscopy was obtained for nearly all candidate members; only a few objects near the locus of Taurus members were not observed spectroscopically. BHSM obtained spectra for all but one of the brown dwarf candidates at $`I18`$ in the diagram of $`RI`$ versus $`I`$ (their Figure 2). The object lacking a spectrum is a likely brown dwarf (see § 4.1.3) and is shown as an open triangle in the various color-color and color-magnitude diagrams. Thus, in the $`Iz\mathrm{}`$ versus $`I`$ and $`JH`$ versus $`IK_s`$ diagrams in Figs. 1 and 2, after plotting this brown dwarf candidate and the spectroscopically confirmed members, all other stars at $`I18`$ can be labeled as field stars. Because the spectroscopy is not 100% complete for the candidates at stellar masses, a few of the objects taken as field stars may be young stars.
I can now use the color-color and color-magnitude diagrams presented here to separate background stars and candidate low-mass members that are beyond the sensitivity of the BHSM study ($`I=18`$-21). To determine the regions in Figs. 1 and 2 where low-mass Taurus members should appear, I must plot the intrinsic colors and magnitudes expected for young brown dwarfs at masses of 0.02 and 0.08 $`M_{}`$. The $`Iz\mathrm{}`$ colors for these masses are estimated in the following manner. The combination of the models of Baraffe et al. (1998) and the compatible temperature scale of Luhman (1999) indicate that brown dwarfs at masses of 0.02 and 0.08 $`M_{}`$ and an age of 1 Myr should have spectral types near M9 and M6.5, respectively. From the locus of reddened background stars in Figure 2, I measure a reddening slope of $`E(IK_s)/E(JH)=4.0`$. By combining this slope with the dwarf colors of Leggett (1992) for spectral types of M9 and M6.5, I plot the reddening vectors for 0.02 and 0.08 $`M_{}`$ from $`A_H=0`$-1.4 in Figure 2. The $`E(IK_s)/E(JH)`$ slope is also combined with $`E(JH)=0.107`$ $`A_V`$ (Rieke & Lebofsky 1985) and $`A_{K_s}=0.116`$ $`A_V`$ (extrapolated from $`A_K`$ of Rieke & Lebofsky 1985) to derive $`A_I=0.544`$ $`A_V`$. After measuring a slope of $`E(Iz\mathrm{})/E(JH)=0.75`$ from the distribution of field stars in $`Iz\mathrm{}`$ versus $`JH`$, I arrive at $`E(Iz\mathrm{})=0.08`$ $`A_V`$. From this extinction relation and the reddenings in Table 1, the intrinsic $`Iz\mathrm{}`$ colors of the M6-M6.25 Taurus members are calculated. These dereddened $`Iz\mathrm{}`$ colors are then extrapolated to M6.5 and M9 ($`Iz\mathrm{}=1.6`$ and 1.75) by using the variation in $`IZ`$ from M6 to M9 for objects in the field and the Pleiades open cluster (Steele & Howells 2000; Zapatero Osorio et al. 1999). The magnitudes at $`I`$ for 0.02 and 0.08 $`M_{}`$ at an age of 1 Myr are calculated by combining the bolometric corrections and distance modulus from § 3.3 with the luminosities predicted by Baraffe et al. (1998). The resulting positions for these two masses are shown in Figure 1. Most stars in these Taurus fields have ages of $`3`$ Myr, with a few as old as 10 Myr (§ 4.2.1). Therefore, to obtain a census that is complete down to 0.02 $`M_{}`$, the boundary in Figure 1 used in selecting low-mass candidates must include objects at 0.02 $`M_{}`$ with ages as old as 10 Myr. From 1 to 10 Myr, the evolutionary models of Baraffe et al. (1998) and Burrows et al. (1997) predict a decrease in luminosity of 0.5 and 0.9 mag, respectively. Thus, to be conservative in the rejection of field stars, a reddening vector is placed at 1 mag below the position of 0.02 $`M_{}`$ in Figure 1; any low-mass members with masses above 0.02 $`M_{}`$ should fall above this line. The nature of stars above the reddening vector will be examined in more detail in § 4.1.3. Stars below this vector are rejected as field stars for the purposes of calculating the IMF down to 0.02 $`M_{}`$. However, substellar members of Taurus with masses of $`0.01`$ $`M_{}`$ ($``$L0) could fall below the reddening vector because the $`Iz\mathrm{}`$ color saturates at early L types (Steele & Howells 2000). As demonstrated in Figure 1, most of the targets identified as foreground and background stars through spectroscopy could have been rejected by deep $`I`$ and $`z\mathrm{}`$ photometry alone had it been available.
#### 4.1.2 Selection of the Reddening Limit of the IMF Sample
Because the mass and reddening vectors are roughly perpendicular in a near-IR color-magnitude diagram and because all of the late-type members should have similar intrinsic IR colors, completeness in mass and reddening are readily evaluated with near-IR data. Thus, by combining the 2MASS data with the membership information from the spectroscopy and the optical color-magnitude diagrams of BHSM and this work, I can determine the optimum reddening limit for the sample used in calculating the IMF. This reddening limit will be high enough to include a large number of cluster members while low enough to achieve completeness to very low masses. In Figure 3, the near-IR color-magnitude diagram is generated from the 2MASS data for the four Taurus fields, with the exception of the portion of L1551 that was unavailable in 2MASS (see § 2.2). The membership status is established for most of the faint sources at $`JH1.5`$. Because this color corresponds to a reddening of $`A_H=1.4`$ for young brown dwarfs ($`JH0.65`$; Luhman 1999), I select a reddening limit of $`A_H1.4`$ for the sample from which the IMF is computed. If a young brown dwarf with $`A_H1.4`$ exhibited an excess at $`JH`$ arising from emitting circumstellar material, it could fall in the region of high $`JH`$ in Figure 3 where membership status is uncertain for most objects. However, the fraction of stellar members that have such large excesses at $`JH`$ is very small. Furthermore, the young brown dwarfs studied by Luhman (1999) in IC 348 appear to have negligible excess emission in $`JH`$.
#### 4.1.3 Any Brown Dwarfs within the Reddening Limit?
I now determine whether there are any new low-mass candidates within the reddening limit that is used for the IMF measurement. In Figs. 1-4, stars that have not been rejected as field stars in § 4.1.1 and are not known members are divided into three categories: stars detected only in the 2MASS data (no $`I`$ and $`z\mathrm{}`$), optical sources with high reddening (high $`JH`$), and optical sources with low reddening (low $`JH`$). For the former sources, the membership status cannot be unambiguously determined because only IR data is available. However, this is not important for the IMF calculation because these objects are either fainter than the $`H`$ magnitude of the desired completeness limit of 0.02 $`M_{}`$ or are redder than the extinction limit of $`A_H1.4`$, as shown in Figure 3 and discussed further in § 4.1.5. The high reddening category consists of sources that fall to the right of the $`A_H1.4`$ reddening vectors in Figure 3; these stars are beyond the extinction limit for the IMF and therefore are not considered further. Finally, the low reddening objects are defined as those between the $`A_H1.4`$ reddening vectors for 0.02 and 0.08 $`M_{}`$ or below the vector for 0.02 $`M_{}`$ in Figure 3. Because the three objects in this category are within the reddening limit of the IMF sample, their membership status must be examined to determine whether they should be added to the IMF. The brightest of these three sources is 2MASSs J0418511+281433, the brown dwarf candidate of BHSM mentioned in § 4.1.1. The 2MASS and optical photometry for this object are $`H=13.19`$, $`JH=0.71`$, $`HK_s=0.44`$, $`I=16.77`$, and $`Iz\mathrm{}=1.67`$. As shown in Figure 2 of BHSM and in Figs. 1 and 2, 2MASSs J0418511+281433 has the colors and magnitudes expected for a young brown dwarf. The second brightest source is 2MASSs J0432086+242213, which has $`H=15.11`$, $`JH=1.47`$, $`HK_s=0.22`$, $`I=20.61`$, and $`Iz\mathrm{}=2.18`$. This object does not have a high enough ratio of $`IK_s`$ to $`JH`$ to be a late-type member of Taurus, as demonstrated in Figure 2. The third object, 2MASSs J0417504+281440, has $`H=15.33`$, $`JH=1.45`$, $`HK_s=0.92`$, $`I=22.31`$, and $`Iz\mathrm{}=2.34`$. This source has $`JH`$ and $`IK_s`$ colors that are consistent with a late M type (reddest of the three open triangles in Figure 2). However, because it falls at the detection limit in both the optical and IR data, its positions in the color-color and color-magnitude diagrams are uncertain. This source is taken to be a field star for the remainder of this work; the membership status of this one object is not important for the conclusions of this study. In summary, there is one compelling brown dwarf candidate lacking spectroscopy and within the reddening limit of $`A_H1.4`$ used for the IMF in L1495, B209, L1529, and the portion of L1551 currently available through 2MASS.
#### 4.1.4 Special Considerations for L1551
For the portion of the L1551 field that is not included in the 2MASS photometry, the identification of low-mass candidates and estimation of completeness must be performed with the optical data alone. The objects in this region that are not rejected as field stars by the $`Iz\mathrm{}`$ versus $`I`$ diagram are labeled with open boxes in Figure 1. Without IR data, the completeness is not assured for high reddenings at substellar masses. Therefore, rather than use a reddening limit of $`A_H1.4`$ to calculate the IMF as done in the other fields, a smaller range of extinctions must be considered for this part of L1551. Because 12 of the 14 known young stars in L1551 have $`A_H=0`$-0.5, $`A_H0.5`$ is taken as the reddening limit for the calculation of the IMF in this part of L1551. For the IMF to be representative for masses of $`0.02`$ $`M_{}`$, low-mass candidates must be identified down to the magnitude of a 0.02 $`M_{}`$ brown dwarf at an age of 10 Myr with a reddening of $`A_H=0.5`$, which corresponds to $`I=20.6`$. This magnitude was computed by combining the luminosities and temperatures predicted by Baraffe et al. (1998) with the temperature scale and bolometric corrections discussed in § 3.3. There is only one L1551 object above this brightness that is not rejected as a field star. In the other fields covered by 2MASS, all but one of the stars falling above the reddening vector in Figure 1 were found to be either highly reddened or a field star when the optical and IR data were combined. Thus, because it is likely that this one object in L1551 would also be rejected in the same manner if IR photometry were available, it is not considered a likely brown dwarf.
#### 4.1.5 The Mass Completeness of the Reddening-Limited Sample
The reddening-limited sample described here would not be an accurate reflection of the Taurus population if the average extinctions were a function of mass, e.g., young brown dwarfs were more highly reddened than stars. In previous magnitude-limited searches for low-mass stars and brown dwarfs in star forming clusters (e.g., Luhman 1999; Wilking, Greene, & Meyer 1999) no difference in reddening characteristics has been found between the stellar and substellar members. The data in this study can be used as an additional test of whether brown dwarfs are hidden by higher amounts of extinction than the known stellar population. Most of the targets of the IR spectroscopy were selected from the imaging of LR98 towards a $`10\mathrm{}\times 10\mathrm{}`$ region of the L1495 dark cloud. As shown in the IR color-magnitude diagram for this area in Figure 4, no low-mass members are found at a dereddened magnitude of $`H>10.5`$-13 in a sample where the membership status is complete to large reddenings ($`A_V20`$). Although this test suffers from poor number statistics, it is consistent with the result from studies of other young regions that a reddening-limited measurement produces a representative IMF down to substellar masses.
The sample selected for measuring the IMF is defined by $`A_H0.5`$ for a portion of L1551 and $`A_H1.4`$ elsewhere. As discussed in the previous two sections, out of the objects with optical photometry, there is one likely brown dwarf within these reddening limits and lacking spectroscopy; it will be added to the IMF in § 4.2.1. For stars with only IR photometry, membership cannot be determined. However, none of these stars are within the reddening ($`A_H1.4`$) and mass ($`0.02`$ $`M_{}`$) limits for ages of $`1`$ Myr, as shown in Figure 3. Because the 1 and 3 Myr isochrones are nearly coincident at 0.02 $`M_{}`$, this statement holds for an age limit of 3 Myr as well. The sample is incomplete only for objects that are simultaneously old (10 Myr) and at the reddening and mass limits. Because such objects occupy a tiny fraction of the mass-reddening-age phase space and because most members of these Taurus fields have ages of $`3`$ Myr (see Figure 5), this incompleteness is negligible. Because there remain $`5`$ candidate members at stellar masses that lack spectra in the color-magnitude diagram of BHSM, the completeness for brown dwarfs above 0.02 $`M_{}`$ is as good or better than that for stars. Thus, although the census of members may not be 100% complete in any mass interval, it is not biased against brown dwarfs and is representative down to a mass of 0.02 $`M_{}`$.
### 4.2 The Initial Mass Function
#### 4.2.1 Taurus
In the Hertzsprung-Russell (H-R) diagram in Figure 5, the models of Baraffe et al. (1998) are plotted with the Taurus members from Table 1, with the exception of L1551/IRS5, the three objects with unknown spectral types, and the two earliest stars. These models are used to infer masses for individual sources from their estimated temperatures and luminosities. For stars in the reddening-limited sample that are above the solar mass track of Baraffe et al. (1998), the prescription of Luhman et al. (2000) is followed; the two stars are placed in one mass bin from log $`M=0.05`$ to 0.35 (0.89-2.2 $`M_{}`$). 2MASSs J0418511+281433 is the one likely Taurus member that lacks a measured spectral type and is within the reddening limit of the IMF sample. If an age of 1 Myr is assumed for this object, the optical and IR photometry (§ 4.1.3) imply a mass of $`0.03`$ $`M_{}`$ with the models of Baraffe et al. (1998). Because the masses of many of the low-mass sources in the Trapezium IMF of Luhman et al. (2000) were derived through photometry alone, they used the bin size of $`\mathrm{\Delta }\mathrm{log}`$ $`M=0.4`$ at all masses. For all but one of the sources in the Taurus IMF, on the other hand, more precise masses are inferred through spectroscopy and placement on the H-R diagram, thus, smaller mass bins of $`\mathrm{\Delta }\mathrm{log}`$ $`M=0.2`$ are used below log $`M=0.05`$. The IMF from the reddening-limited sample for the four Taurus fields contains 40 sources and is presented in Figure 6.
#### 4.2.2 Comparison of Taurus and Clusters
Luhman et al. (2000) found that the stellar IMFs in the young clusters IC 348, $`\rho `$ Oph, and the Trapezium are consistent with the same mass function, while modest variations in the substellar IMFs are possible. Because of the superior number statistics in the Trapezium sample, that IMF is taken to represent star forming clusters in the following comparison between Taurus and clusters.
For a meaningful comparison of the IMFs in Taurus and the Trapezium, it is important that the IMFs have been constructed from samples that are physically equivalent and relatively primordial. Thus, the minimum separations of objects in the two samples should correspond to the same physical scale and should be less then the distance at which disruption through dynamical interactions is significant. These criteria have been satisfied fairly well; the IMFs for the Trapezium and Taurus consist of sources with separations greater than $`0\stackrel{}{\mathrm{.}}7`$ and $`2\mathrm{}`$, respectively, which are equal physical distances. Furthermore, these separations are less than the length scales in the Trapezium ($`1\mathrm{}`$) and Taurus ($`90\mathrm{}`$) that mark the transitions from the binary regime to large-scale clustering (Gomez et al. 1993; Larson 1995; Simon 1997; Nakajima et al. 1998; Bate, Clarke, & McCaughrean 1998), which in turn should be less than the maximum binary separations (Bate et al. 1998).
Because the samples in Taurus and the Trapezium have been designed to be comparable in their binarity and because the same techniques were employed in estimating the masses for each population, the IMFs for the two regions can be reliably compared. In addition, the optimum reddening limit for each sample is $`A_H=1.4`$ (except for L1551), making the two IMFs particularly suitable for comparison. The methods of these studies should also produce reasonably accurate measurements of the IMFs in each region (see Luhman et al. 2000).
The IMFs for Taurus and the Trapezium are plotted together in Figure 6, where both mass functions are representative down to 0.02 $`M_{}`$. The Trapezium IMF has been normalized to the Taurus data between 0.1 and 1 $`M_{}`$. Over this mass range, the two mass functions are consistent with the same shape; a peak at 0.6-1 $`M_{}`$ followed by slight decline and flattening down to 0.1 $`M_{}`$. However, below 0.1 $`M_{}`$ the two IMFs differ greatly. If the IMF in Taurus were the same as that in the Trapezium, $`12.8\pm 1.8`$ brown dwarfs ($`>0.02`$ $`M_{}`$) are expected in the former, but only one is found. Relative to the Trapezium, Taurus exhibits a deficit of higher mass stars as well. At 1-2 $`M_{}`$ the renormalized Trapezium IMF contains $`7.8\pm 1.4`$ stars while the Taurus sample has only 2 stars. Neither IMF is characterized by a log-normal distribution.
The deficit of brown dwarfs in Taurus relative to the Trapezium motivates searches for variations in substellar mass functions in regions of intermediate stellar densities. For instance, the number statistics of the substellar census in IC 348 (Luhman et al. 1998b; Luhman 1999) and $`\rho `$ Oph (Luhman & Rieke 1999; Wilking et al. 1999) can be improved to determine whether they differ from the Trapezium at a significant level.
The significant number of brown dwarfs in the field (Reid et al. 1999) and the scarcity of them in Taurus combined with the similarity of the IMFs in star forming clusters, young open clusters, and the field (Luhman et al. 2000) are convincing evidence that the field is populated by stars born in clusters rather than isolated regions (Lada, Strom, & Myers 1993).
#### 4.2.3 Implications for Theories of the IMF
Luhman et al. (2000) compared predictions of theories of the IMF to the shape of the low-mass IMF, its approximate invariance at stellar masses among IC 348, $`\rho `$ Oph, and the Trapezium, and the minimum mass observed for isolated objects ($`0.01`$ $`M_{}`$), possibly below the deuterium burning mass limit (0.013-0.015 $`M_{}`$, Burrows et al. 1997). In addition, while similarities between the mass functions of pre-stellar clumps and the stellar IMF have recently been reported and suggested as evidence that the process of fragmentation determines the masses of stars (Motte, André, & Neri 1998; Testi & Sargent 1998), Luhman et al. (2000) found that a closer comparison of these pre-stellar mass functions to accurate IMFs recently derived for star forming regions is not yet conclusive. I now include the implications of the new IMF for Taurus and its behavior relative to the mass function for the Trapezium.
The high-mass IMF in Taurus is first considered. Elmegreen (1997, 1999, 2000) contended that differences in measurements of IMF slopes may be statistical fluctuations rather than true variations in the IMF. Regions with fewer high mass stars simply do not have enough members to populate the mass function to high masses. He points to the agreement between cluster IMFs and the integrated galaxy IMF as supporting evidence. However, it is clear from the work presented here and by Luhman et al. (2000) that star forming clusters rather than sparse regions like Taurus are responsible for the Galactic field population. Thus, a true deficit of high-mass stars in Taurus would not be reflected in the Galactic IMF. Indeed, a comparison of Taurus and the Trapezium suggests that the lack of high-mass stars is not because of poor sampling of the IMF, although the statistical significance of this result should be improved through an expansion of this work to include a larger fraction of the Taurus population. Elmegreen does refer to possible exceptions to his model, such as the extreme field in the LMC and Milky Way (Massey et al. 1995), where the high-mass IMFs are much steeper than those in the Galactic and cluster IMFs. In fact, Elmegreen (1999) briefly refers to a scenario in which a low-density star forming region like Taurus could produce fewer high-mass stars.
The predictions of individual theories are now compared to the low-mass IMFs in Taurus and the Trapezium.
While wind-limited accretion models for the IMF (Adams & Fatuzzo 1996) can account for the value of the low-mass cutoff in Taurus, they do not explain the change in the minimum mass from Taurus to the Trapezium, the very low value of the cutoff in the Trapezium, or the deficit of high-mass stars in Taurus. Furthermore, if the IMF is determined by a large number of variables, a log-normal form is predicted for the IMF, which is not observed in either region.
Elmegreen (1997, 1999) presented a model where the IMF is controlled by fragmentation from hierarchical clouds modulated by the Jeans condition. The shape of the IMF at intermediate masses in Taurus and the Trapezium – a power law slope followed by a turnover and flattening – is reproduced by this model (and many others; Scalo 1999). Both the turnover mass and the minimum mass should be reflections of the Jeans mass, thus they should change together from one region to another. However, this is not the case in the Taurus and the Trapezium populations, which exhibit the same turnover masses of $`0.8`$ $`M_{}`$ but very different low-mass cutoffs of $`0.01`$ $`M_{}`$ and $`0.08`$ $`M_{}`$, respectively.
Theories of the IMF that rely on interactions of cores or protostars (Lejeune & Bastien 1986; Murray & Lin 1996; Price & Podsiadlowski 1995; Bonnell et al. 1997) have difficulty in explaining the similarity of the turnover masses in the Trapezium and Taurus, where the stellar densities differ by three orders of magnitude. Dynamical interactions and competitive accretion between protostellar cores were also important in shaping the IMF in numerical modeling of the gravitational collapse and fragmentation of a dense cloud core by Klessen, Burkert, & Matthew (1998). The predicted IMF exhibited a log-normal form that was centered at a mass described by the product of the Jeans mass of the system and the star forming efficiency. As demonstrated here and by Luhman et al. (2000), the IMF is not characterized by a log-normal function, with the possible exception of that in globular clusters.
Any theory of the IMF must explain the differences in the frequencies of brown dwarfs and (more tentatively) of high-mass stars between Taurus and young clusters. One property that varies dramatically from Taurus to star forming clusters is the level of turbulence observed in cloud cores (Myers 1998), which may be the ultimate origin of these IMF variations. For instance, it is possible that both the growth and the fragmentation of pre-stellar cores are enhanced in more turbulent environments, which could broaden the IMF in a dense cluster relative to that in an isolated region (Myers 2000b). Further modeling of the formation of stellar clusters within turbulent dense cores (Klessen et al. 1998; Myers 1998, 2000a) may help explain the observed behavior of the low-mass IMF.
## 5 Conclusion
An IMF that is representative for masses down to 0.02 $`M_{}`$ has been measured for fields near the L1495, B209, L1529, and L1551 dark clouds in the Taurus star forming region. These data and similar studies of star forming clusters provide powerful constraints on theories of star formation.
The observations for star forming clusters as summarized by Luhman et al. (2000) are first reviewed:
1. Above 0.1 $`M_{}`$, the IMFs in the core of IC 348, the cloud core of $`\rho `$ Oph, and the Trapezium Cluster are indistinguishable within the counting uncertainties.
2. Below 0.1 $`M_{}`$, the IMFs in the three clusters are roughly similar, although the statistical uncertainties in the data for $`\rho `$ Oph and IC 348 allow for modest variations.
3. Data for the Trapezium, young open clusters, and the field are consistent with the same IMF, which differs from the IMF that characterizes globular clusters (Paresce & De Marchi 2000).
4. The IMF for young clusters and the field is flat or slowly rising from the substellar regime to $`0.6`$ $`M_{}`$ and then rolls over into a power law that continues from $`1`$ $`M_{}`$ to higher masses with a slope similar to or somewhat larger than the Salpeter value of 1.35; it cannot be described by a log-normal function.
5. In the Trapezium, the IMF is flat down to 0.02 $`M_{}`$ or lower and contains a population of $`50`$ likely brown dwarfs.
6. The least massive objects observed in the Trapezium appear to have masses of $`0.01`$ $`M_{}`$.
The new conclusions provided by this study of Taurus are as follows:
1. Above 1 $`M_{}`$, the fields of Taurus in this study have proportionately fewer stars than the Trapezium at a modest level of significance.
2. From 0.1-1 $`M_{}`$, the IMF in Taurus matches that of the Trapezium; both regions have a turnover mass near $`0.8`$ $`M_{}`$ and a slow decline and flattening to lower masses.
3. Below 0.1 $`M_{}`$, there is a significant deficit of objects in Taurus relative to the Trapezium. If the Trapezium IMF is normalized to the Taurus IMF by the numbers of objects between 0.1-1 $`M_{}`$, then $`12.8\pm 1.8`$ brown dwarfs at $`>0.02`$ $`M_{}`$ are present in the Trapezium where one is found in the Taurus fields.
4. In summary, if the Trapezium IMF is steepened above 1 $`M_{}`$ and suppressed below 0.1 $`M_{}`$, then the result is the IMF for Taurus; it cannot be described by a log-normal function.
I am grateful to F. Allard, I. Baraffe, and F. D’Antona for access to their most recent calculations. Discussions with L. Hartmann, G. Rieke, J. Stauffer and, in particular, P. Myers are appreciated. I also thank C. Briceño for providing details concerning his published work. K. L. was funded by a postdoctoral fellowship at the Harvard-Smithsonian Center for Astrophysics. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center, funded by the National Aeronautics and Space Administration and the National Science Foundation.
|
warning/0007/math0007055.html
|
ar5iv
|
text
|
# On the Stability of the Standard Riemann Semigroup
## 1. Introduction
Under suitable assumptions on the function $`f:\mathrm{\Omega }^n`$ (with $`\mathrm{\Omega }^n`$), the system
(1.1)
$$u_t+\left[f(u)\right]_x=0$$
generates a Standard Riemann Semigroup (SRS) $`S:[0,+\mathrm{}[\times 𝒟𝒟`$, see . Aim of this paper is to investigate the dependence of $`S`$ upon the flow function $`f`$.
Several papers in the current literature are concerned with the existence of an SRS, see for example and the references in . On the contrary, in the present paper the existence of an SRS is assumed as a starting point and the focus is on the correspondence $`fS`$. In fact, the results in this paper imply that the SRS $`S`$ is a Lipschitzean function of the flow $`f`$, with respect to the $`C^0`$ norm of $`Df`$. An immediate consequence is the following. Assume that $`f`$ depends on the parameters $`(p_1,\mathrm{},p_m)`$ that may vary in a compact subset of $`^m`$. Given a continuous functional $`J`$ defined on the solution $`u`$ at time $`t`$ to the Cauchy problem for (1.1), the present result ensures the continuity of the map $`(p_1,\mathrm{},p_m)J(u(t))`$ hence, by Weierstrass Theorem, the optimization problem admits a solution.
For the sake of completeness, we only recall here that the existence of the SRS for the $`n\times n`$ system (1.1) was first proved in . The main assumptions there are that $`Df`$ is strictly hyperbolic with every characteristic field either linearly degenerate or genuinely nonlinear and that the initial data has sufficiently small total variation. More recently, the existence of the SRS was extended also to the non genuinely nonlinear setting in the $`2\times 2`$ case, see .
Below we shall restrict our attention to standard solutions to Riemann problems and, hence, to general Cauchy problems for Conservation Laws. Here, by standard solutions we refer to those introduced by Lax and then generalized by Liu . Various extensions of the present work to other types of solvers are straightforward.
The present paper is organized as follows. In the next section we state the main results. The following Sections 4 and 5 are devoted to two applications: the classical – relativistic limit of Euler equations and scalar conservation laws with $`L^{\mathrm{}}`$ initial data. The proofs are given in Section 3.
## 2. Notation and Main Results
Consider the following hyperbolic system of conservation laws in one space dimension
(2.1)
$$u_t+\left[f(u)\right]_x=0$$
where $`f:\mathrm{\Omega }^n`$ is in $`\text{Hyp}(\mathrm{\Omega })`$, i.e. $`f`$ is a sufficiently smooth function that generate a SRS $`S^f:[0,+\mathrm{}[\times 𝒟^f𝒟^f`$. Recall that by SRS generated by $`f`$ (see ) we mean a map $`S^f:[0,+\mathrm{}[\times 𝒟^f𝒟^f`$ with the following properties:
* $`S^f`$ is a semigroup: $`S_0^f=\mathrm{Id}`$ and $`S_t^fS_s^f=S_{t+s}^f`$;
* $`S^f`$ is Lipschitz continuous: there exists a positive $`L_f`$ such that for all positive $`t,s`$ and for all $`u,w𝒟^f`$, $`S_t^fuS_s^fw_{L^1}L_f\left(|ts|+uw_{L^1}\right)`$;
* if $`u`$ is piecewise constant, then for $`t`$ small, $`S_t^fu`$ coincides with the glueing of standard solutions to Riemann problems.
For all $`u𝒟^f`$, it is well known (see ) that the map $`tS_t^fu`$ is a weak entropic solution to (2.1).
Given $`f\text{Hyp}(\mathrm{\Omega })`$, let $`(𝒟^f)`$ be the set of all piecewise constant functions in $`𝒟^f`$ having a single jump at the origin. In other words, $`(𝒟^f)`$ is the set of initial data to the Riemann problems
(2.2)
$$\{\begin{array}{c}u_t+f(u)_x=0\\ u(0,x)=\{\begin{array}{cc}u^{}\hfill & \text{if }x<0\hfill \\ u^+\hfill & \text{if }x>0\hfill \end{array}\end{array}$$
Below, by solution to (2.2) we always refer to the standard Lax (see ) self–similar entropic solutions.
Let $`f,g\text{Hyp}(\mathrm{\Omega })`$ with
(2.3)
$$𝒟^g𝒟^f$$
and define the “distance” between $`f`$ and $`g`$ as (cfr. )
(2.4)
$$\widehat{d}(f,g)=\underset{u(𝒟^g)}{sup}\frac{1}{|u^+u^{}|}S_1^fuS_1^gu_{L^1}$$
The distance $`\widehat{d}(f,g)`$ is well defined due to (2.3). The main result of the present paper is the following theorem.
###### Theorem 2.1.
Let $`f\text{Hyp}(\mathrm{\Omega })`$. Then, for all $`g\text{Hyp}(\mathrm{\Omega })`$ with $`𝒟^g𝒟^f`$ and for all $`u𝒟^g`$
(2.5)
$$S_t^fuS_t^gu_{L^1}L_f\widehat{d}(f,g)_0^t\text{Tot.Var.}\left(S_t^gu\right)𝑑t.$$
Recall that $`L_f`$ is the Lipschitz constant of the semigroup $`S^f`$, see (ii) above. The proof of Theorem 2.1 is deferred to Section 3.
Remark that $`\widehat{d}`$ generalizes the analogous quantity $`\widehat{d}_{\text{lin}}`$ defined in with reference to the linear case. Let $`𝕄_d^{n\times n}`$ denote the set of $`n\times n`$ diagonalizable matrices with real eigenvalues. Note that $`𝕄_d^{n\times n}\text{Hyp}(^n)`$. Fix a $`v^n`$, $`v0`$. Denote by $`A^tv`$ the solution evaluated at time $`t`$ of the linear system
$$\{\begin{array}{c}u_t+Au_x=0\\ u(0,x)=v\chi _{[0,+\mathrm{})}(x)\end{array}$$
(here, $`\chi _I`$ is the characteristic function of the interval $`I`$). Theorem 2.3 in shows that
(2.6)
$$\widehat{d}_{\text{lin}}(A,B)\underset{v:|v|=1}{sup}A^1vB^1v_{L^1},$$
is a distance on $`𝕄_d^{n\times n}`$ such that for all $`A,B𝕄^{n\times n}`$
(2.7)
$$BA\widehat{d}_{\text{lin}}(A,B),$$
$``$ being the usual operator norm. Moreover, $`(𝕄_d^{n\times n},\widehat{d}_{\text{lin}})`$ is a complete metric space. Clearly, if $`f`$ and $`g`$ are linear, then $`\widehat{d}_{\text{lin}}(f,g)=\widehat{d}(f,g)`$.
Furthermore, $`\widehat{d}`$ is related to $`\widehat{d}_{\text{lin}}`$ computed on the derivatives of the flow functions, as shown by the following Proposition.
###### Proposition 2.2.
Let $`f,g\text{Hyp}(\mathrm{\Omega })`$ with $`𝒟^g𝒟^f`$. Then
(2.8)
$$\widehat{d}(f,g)\underset{u\mathrm{\Omega }}{sup}\widehat{d}_{\text{lin}}(Df(u),Dg(u)).$$
Thus, $`\widehat{d}(f,g)`$ seems stronger than the $`C^0`$ distance between $`Df`$ and $`Dg`$, in the sense of (2.7). Nonetheless, Corollary 2.5 below shows that once the flow functions and the domains $`𝒟^f`$, $`𝒟^g`$ are fixed, i.e. the total variation of the solutions $`S_t^fu`$, $`S_t^gu`$ are uniformly bounded, we can estimate the r.h.s. in (2.5) by means of $`DfDg_{C^0}`$.
Theorem 2.1 shows that the key point in the stability of the SRS w.r.t. the flow function lies in the dependence of only the solution to Riemann problems upon the flow function. From the more abstract point of view of quasidifferential equations in metric spaces (see ) this is equivalent to relate the distance between semigroups to the distance between the vector fields generated by the semigroups.
As in (see also ), in a metric space $`(E,d)`$ define an equivalence relation on all the Lipschitz curves $`\gamma :[0,1]E`$ exiting a fixed point $`u`$ as
(2.9)
$$\gamma \gamma ^{}\text{ iff }\underset{\theta 0^+}{lim}\frac{d(\gamma (\theta ),\gamma ^{}(\theta ))}{\theta }=0.$$
The quotient space $`T_u`$ so obtained is naturally equipped with the metric
(2.10)
$$\widehat{d}(v_1,v_2)\underset{\theta 0^+}{lim\; sup}\frac{d(\gamma _1(\theta ),\gamma _2(\theta ))}{\theta },$$
where $`\gamma _i`$ is a representative of the equivalence class $`v_i`$. By (2.9), $`\widehat{d}`$ does not depend on the particular representatives chosen. A map $`v:E_{uE}T_u`$ is a vector field, provided $`v(u)T_u`$ for all $`u`$.
Let $`S:E\times [0,+\mathrm{}[E`$ be a Lipschitzean semigroup, i.e. $`S`$ satisfies
$$\begin{array}{c}S_0=𝕀_E\\ S_sS_t=S_{s+t}\end{array}\text{ and }\begin{array}{c}L>0\text{ such that}\\ t,s>0\text{ and }u,w𝒟\\ d(S_tu,S_sw)L\left(|ts|+d(u,w)\right).\end{array}$$
Then, $`S`$ naturally defines a vector field $`v_S`$ on $`E`$ by
(2.11)
$$v_S(u)\text{ is the equivalence class of the orbit }\theta S_\theta u.$$
Theorem 2.1 has a natural abstract counterpart, namely
###### Proposition 2.3.
Let $`S`$, $`S^{}`$ be two Lipschitz semigroups on $`E`$ generating the vector fields $`v`$ and $`v^{}`$, respectively. Denote with $`L`$ the Lipschitz constant of, say, $`S`$. Then
(2.12)
$$d(S_tu,S_t^{}u)L_0^t\widehat{d}(v(S_t^{}u),v^{}(S_t^{}u))𝑑t.$$
The above proposition is an immediate corollary of the following widely used (see and the references in ) error estimate:
###### Lemma 2.4.
Given a Lipschitz semigroup $`S:E\times [0,+\mathrm{}[E`$ with Lipschitz constant $`L`$, for every Lipschitz continuous map $`w:[0,T]E`$ one has
(2.13)
$$d(w(T),S_tw(0))L_0^T\underset{h0^+}{lim\; inf}\frac{d(w(t+h),S_hw(t))}{h}dt.$$
The proof of Theorem 2.1 consists of the following steps.
* Find an explicit definition of the vector field $`v_S`$ generated by the SRS $`S`$.
* Compute the r.h.s. in (2.12).
* Apply Lemma 2.4.
Note that this procedure requires the mere existence of the SRS. In several cases (see for instance ) the existence of the SRS is achieved through the construction of a sequence $`S^n`$ of uniformly Lipschitzean approximate semigroups defined on piecewise constant functions. In all these cases, the vector field $`v_n`$ generated by $`S^n`$ on the set of piecewise constant functions simply consists in the gluing of solutions to Riemann problems. Thus, under the further assumption that such an approximating sequence $`S^n`$ exists, the step (1) above could be avoided.
The quantity $`\widehat{d}`$ in (2.4) is thus the natural tool to estimate the dependence of the SRS upon the flow function (note also that no constant is involved in (2.5)). However, in view of possible applications of Theorem 2.1, we provide an estimate of the r.h.s. in (2.5) in terms of handier quantities.
###### Corollary 2.5.
Let $`f\text{Hyp}(\mathrm{\Omega })`$ and assume that
$$𝒟^f\{uL^1(,K):\text{Tot.Var.}(u)M\}$$
for suitable positive $`M`$ and compact $`K^n`$. Then, there exists a positive constant $`C`$ such that for all $`g\text{Hyp}(\mathrm{\Omega })`$ with $`𝒟^g𝒟^f`$ and for all $`u𝒟^g`$
(2.14)
$$S_t^fuS_t^gu_{L^1}CDfDg_{C^0(\mathrm{\Omega })}t.$$
The above is the counterpart of the well known estimate for solutions of bounded variation in the scalar case given in . Note that, differently from the linear case, the “distance” $`\widehat{d}(f,g)`$ is equivalent to $`DfDg_{C^0}`$ because the total variation of both solutions $`S^fu`$ and $`S^gu`$ is fixed by the domain $`𝒟^f`$: thus the case of example of Remark 3.3 in is not valid here.
For scalar equations and assuming that the flow functions $`f`$ and $`g`$ are strictly convex, we are able to extend the estimate in to $`L^{\mathrm{}}`$ initial data.
###### Theorem 2.6.
Assume that the scalar flow functions $`f`$ and $`g`$ are uniformly strictly convex in a compact interval $`K`$: i.e. $`f^{\prime \prime }(u),g^{\prime \prime }(u)\kappa >0`$ for all $`uK`$. Let $`u`$ (resp. $`w`$) denote the solution to
$$\{\begin{array}{c}u_t+\left[f(u)\right]_x=0\\ u(0,x)=u_o(x),\end{array}\text{ resp.}\{\begin{array}{c}w_t+\left[g(w)\right]_x=0\\ w(0,x)=u_o(x)\end{array}$$
with the same initial data $`u_oL^{\mathrm{}}(,K)`$. Denote by $`\widehat{\lambda }`$ an upper bound for the characteristic speeds, i.e. $`\widehat{\lambda }\mathrm{max}_{uK}\{|f^{}(u)|,|g^{}(u)|\}`$. Then
(2.15)
$$_a^b|u(t,x)w(t,x)|dx2\text{diam}(K)t\frac{ba+4\widehat{\lambda }t}{\kappa t}\underset{uK}{\mathrm{max}}|f^{}(u)g^{}(u)|.$$
## 3. Proof of the main results
This section is devoted to the proofs of Theorem 2.1, Proposition 2.2 and Corollary 2.5. The first step consists in the explicit computation of the vector field $`v(u)`$ generated by a SRS in the sense of (2.11). This procedure follows . Remark that, given an $`f\text{Hyp}(\mathrm{\Omega })`$, such construction is accomplished for all $`u𝒟^f`$.
Fix $`f\text{Hyp}(\mathrm{\Omega })`$ and $`u𝒟^f`$. $`Du`$ stands for the total variation of the weak derivative of $`u`$. Let $`\widehat{\lambda }`$ be a constant strictly greater than all the characteristic speeds induced by $`f`$ on $`𝒟^f`$. Let $`\xi `$ be given. We denote by $`\omega `$ the self similar solution of the Riemann problem
$$\{\begin{array}{c}\omega _t+f(\omega )_x=0\\ \omega (0,x)=\{\begin{array}{cc}u(\xi )\hfill & \text{if }x<0\hfill \\ u(\xi +)\hfill & \text{if }x>0\hfill \end{array}\end{array}$$
and let $`U_\xi ^{\mathrm{}}`$ be the function
(3.1)
$$U_\xi ^{\mathrm{}}(\theta ,x)\{\begin{array}{cc}\omega (\theta ,x\xi )\hfill & \text{if }|x\xi |\widehat{\lambda }\theta \hfill \\ u(x)\hfill & \text{if }|x\xi |>\widehat{\lambda }\theta \hfill \end{array}$$
Moreover, define $`U_\xi ^{\mathrm{}}`$ as the solution to the linear hyperbolic problem with constant coefficients
(3.2)
$$\{\begin{array}{c}\omega _\theta +Df(\xi )\omega _x=0\\ \omega (0,x)=u(x).\end{array}$$
Given $`ϵ>0`$ by an $`ϵ`$–covering of the real line we mean a family
(3.3)
$$\{I_1,\mathrm{},I_N,I_1^{},\mathrm{},I_M^{}\}$$
of open intervals which cover $``$ such that:
* the intervals $`I_\alpha `$ are mutually disjoint and no point $`x`$ lies inside more than two distinct intervals $`I_\beta ^{}`$;
* for every $`\alpha `$ there exists $`\xi _\alpha I_\alpha `$ such that $`Du(I_\alpha \{\xi _\alpha \})<ϵ/N`$;
* for every $`\beta `$, $`Du(I_\beta ^{})<ϵ`$.
Now, let $`ϵ_n`$ be a sequence strictly decreasing to $`0`$. If $`_n=\{I_1,\mathrm{},I_{N(n)},I_1^{},\mathrm{},I_{M(n)}^{}\}`$ is an $`ϵ_n`$–covering of $``$, then also $`_{n,\theta }=\{I_{1,\theta },\mathrm{},I_{N(n),\theta },I_{1,\theta }^{},\mathrm{},I_{M(n),\theta }^{}\}`$ is an $`ϵ_n`$–covering, where $`I_{j,\theta }I_j+\lambda (\theta ,\theta )`$ and $`\theta [0,\theta _n]`$, for $`\theta _n`$ sufficiently small and such that the sequence $`\theta _n`$ strictly decreases to $`0`$. Define now
$$u_n^\theta (x)\{\begin{array}{cc}U_{\xi _\alpha }^{\mathrm{}}(\theta ,x)\hfill & \text{if }xI_{\alpha ,\theta }\hfill \\ U_{\xi _\beta }^{\mathrm{}}(\theta ,x)\hfill & \text{if }xI_{\beta ,\theta }^{},x_\alpha I_{\alpha ,\theta },x_{\beta ^{}<\beta }I_{\beta ^{},\theta }^{}\hfill \end{array}$$
We finally obtain the vector field $`v_{S^f}`$ letting $`v_{S^f}(u)`$ be the equivalence class (w.r.t. (2.9)) of the curve $`\theta u^\theta `$, where
(3.4)
$$u^\theta su_n^{\theta _n}+(1s)u_{n1}^{\theta _{n1}}\text{ if }\theta =s\theta _n+(1s)\theta _{n1},s[0,1].$$
In it is shown that the trajectories of SRS $`S^f`$ generated by the hyperbolic system of conservation laws $`u_t+f(u)_x=0`$ are the solution of the quasilinear equation $`\dot{u}=v_{S^f}(u)`$, where the vector field $`v_{S^f}(u)`$ is generated by the curve (3.4).
Let $`f`$, $`g`$ be as in Theorem 2.1, and denote with $`u^\theta `$, $`w^\theta `$ the two curves generating the vector fields $`v_{S^f}`$ and $`v_{S^g}`$ induced by the SRSs $`S^f`$ and $`S^g`$. In view of (2.12), we now pass to compute the distance (2.10) between $`v_{S^f}(u)`$ and $`v_{S^g}(u)`$. A simple computation shows that by Proposition 2.2
(3.5) $`{\displaystyle \frac{d(u^{\theta _n},w^{\theta _n})}{\theta _n}}`$ $`={\displaystyle \frac{1}{\theta _n}}w^{\theta _n}u^{\theta _n}_{L^1}`$
$`{\displaystyle \underset{\alpha }{}}Du(\xi _\alpha )\widehat{d}_{\text{lin}}(f,g)+{\displaystyle \underset{\beta }{}}\text{Tot.Var.}(u,I_\beta ^{})\widehat{d}(Df(\xi _\beta ),Dg(\xi _\beta ))`$
$`\text{Tot.Var.}(u)\widehat{d}(f,g).`$
As a consequence, if $`s`$ is as in (3.4)
(3.6) $`{\displaystyle \frac{d(u^\theta ,w^\theta )}{\theta }}`$ $`={\displaystyle \frac{s(u_n^{\theta _n}w_n^{\theta _n})+(1s)(u_{n1}^{\theta _{n1}}w_{n1}^{\theta _{n1}})_{L^1}}{\theta }}`$
$`{\displaystyle \frac{s\theta _n+(1s)\theta _{n1}}{\theta }}\text{Tot.Var.}(u)\widehat{d}(f,g)`$
$`=\text{Tot.Var.}(u)\widehat{d}(f,g)`$
hence
(3.7)
$$\widehat{d}(v_{S^f}(u),v_{S^g}(u))=\underset{n+\mathrm{}}{lim\; sup}\frac{d(u^\theta ,w^\theta )}{\theta }\text{Tot.Var.}(u)\widehat{d}(f,g).$$
By Proposition 2.3, applying (3.7) to $`S_t^gu`$, the proof of Theorem 2.1 is completed.
In the next part of this section we give a proof of Corollary 2.5. We only need to prove that there exists a constant $`C`$ such that $`\widehat{d}(f,g)CDfDg_{C^0}`$: this means that for all Riemann problems (2.2) we have
(3.8)
$$S_1^fuS_1^gu_{L^1}CDfDg_{C^0}\left|u^+u^{}\right|.$$
We recall that $`S_t^gu`$ is a self similar solution, obtained by piecing together centered rarefaction waves and jump discontinuities. By the $`L_{\text{loc}}^1`$ dependence of $`S^f`$, formula (3.8) is proved if we can verify it for Riemann problems generating a single wave in the solution $`S_t^gu`$. We then have to consider two cases.
If $`S_t^gu`$ is a centered rarefaction wave, then by the Lipschitz continuity of the solution for $`t>0`$, the functions $`S_{t+h}^gu`$ and $`S_h^fS_t^gu`$ solves in the broad sense the quasilinear versions of (2.1):
(3.9)
$$\left[S_h^fS_t^gu\right]_h+Df\left[S_h^fS_t^gu\right]_x=0\text{and}\left[S_{t+h}^gu\right]_h+Dg\left[S_{t+h}^gu\right]_x=0.$$
Applying Lemma 2.4, we obtain
(3.10) $`S_1^fuS_1^gu_{L^1}`$ $`L_f{\displaystyle _0^1}\underset{h0}{lim\; inf}{\displaystyle \frac{S_h^fS_t^guS_{t+h}^gu_{L^1}}{h}}dt`$
(3.11) $`=`$ $`L_f{\displaystyle _0^1}Df\left[S_t^gu\right]_xDg\left[S_t^gu\right]_x_{L^1}`$
$``$ $`CDfDg_{C^0}|u^+u^{}|,`$
since in this case the $`L^1`$ limits as $`h0`$ of $`(S_h^fS_t^gu)/h`$ and $`(S_{t+h}^gu)/h`$ exist by (3.9) and are equal to $`(DfDg)[S_t^gu]_x`$.
Now we consider the case in which the jump $`u^{},u^+`$ is solved by a shock travelling with speed $`\sigma `$, where $`\sigma `$ is given by the Rankine–Hugoniot condition
(3.12)
$$g(u^+)g(u^{})=\sigma \left(u^+u^{}\right).$$
To prove (3.8), we approximate the solution $`S_t^gu`$ with the Lipschitz continuous function $`\stackrel{~}{u}(t)`$ defined as
$$\stackrel{~}{u}(t,x)=u^{}\chi _{(\mathrm{},0]}\left(x\sigma t\right)+\left(u^+u^{}\right)\mathrm{min}(\frac{x\sigma t}{\delta },1)\chi _{(0,+\mathrm{})}\left(x\sigma t\right).$$
Roughly speaking, $`\stackrel{~}{u}`$ is obtained from $`S_t^gu`$ by substituting the jump $`u^{},u^+`$ at $`x\sigma t`$ with a linear function. Using the Lipschitzeanity of $`S^g`$ we can write
(3.13) $`S_h^fS_t^guS_{t+h}^gu_{L^1}`$ $`S_h^fS_t^guS_h^f\stackrel{~}{u}(t)_{L^1}+S_h^f\stackrel{~}{u}(t)\stackrel{~}{u}(t+h)_{L^1}`$
$`+\stackrel{~}{u}(t+h)S_{t+h}^gu_{L^1}`$
$`{\displaystyle \frac{\delta }{2}}(1+L^f)|u^+u^{}|+\stackrel{~}{u}(t+h)S_h^f\stackrel{~}{u}(t)_{L^1}.`$
The last term above can be evaluated again by means of Lemma 2.4:
(3.14) $`\stackrel{~}{u}(t+h)S_h^f\stackrel{~}{u}(t)_{L^1}L^f{\displaystyle _t^{t+h}}\underset{\xi 0^+}{lim\; inf}{\displaystyle \frac{\stackrel{~}{u}(\eta +\xi )S_\xi ^f\stackrel{~}{u}(\eta )_{L^1}}{\xi }}d\eta `$
$`=hL^f{\displaystyle _0^\delta }\left|(\dot{\sigma }_\alpha 𝕀Dg(u_\alpha ^{}+(u_\alpha ^+u_\alpha ^{}){\displaystyle \frac{y}{\delta }}))(u_\alpha ^{}+(u_\alpha ^+u_\alpha ^{}){\displaystyle \frac{y}{\delta }})\right|dy`$
$`hL^f\underset{xK}{sup}Df(x)Dg(x)\left|u_\alpha ^+u_\alpha ^{}\right|,`$
where we use $`\dot{x}_\alpha [\sigma ϵ,\sigma +ϵ]`$ and the relation
$$\sigma (u_\alpha ^+u_\alpha ^{})=f(u_\alpha ^+)f(u_\alpha ^{})=_0^1Df\left((1s)u_\alpha ^{}+su_\alpha ^+\right)𝑑s(u_\alpha ^+u_\alpha ^{}).$$
In fact, since $`u_{\alpha ,\delta }`$ is Lipschitz continuous, we can use Lebesgue’s dominated convergence theorem as in the previous case. Letting $`\delta `$ tend to $`0`$, we obtain finally
(3.15)
$$\frac{1}{h}S_h^fS_t^guS_{t+h}^g_{L^1}L_f\underset{xK}{sup}Df(x)Dg(x)\left|u_\alpha ^+u_\alpha ^{}\right|.$$
This concludes the proof of Corollary 2.5: in fact an application of Lemma 2.4 gives immediately (3.8).
To end this section, we prove Proposition 2.2. Fix $`u_o\mathrm{\Omega }`$ and $`v^n`$ with $`|v|=1`$. Assume that for all positive and sufficiently small $`h`$ the function $`u_h=u_o+h\chi _{[0,+\mathrm{})}v`$ is in $`𝒟^g`$. Let
$$\stackrel{~}{f}_h(u)=\frac{1}{h}\left(f(u)f(u_o)\right),\stackrel{~}{g}_h(u)=\frac{1}{h}\left(g(u)g(u_o)\right)$$
and note that using a simple rescaling, we can write
$$S_1^{\stackrel{~}{f}_h}\left(\frac{1}{h}(u_hu_o)\right)S_1^{\stackrel{~}{g}_h}\left(\frac{1}{h}(u_hu_o)\right)_{L^1}=\frac{1}{h}S_1^fu_hS_1^gu_h_{L^1}$$
(recall that $`(1/h)(u_hu_o)=\chi _{[0,+\mathrm{})}v`$ is independent from $`h`$). Hence, passing to the limit $`h0`$
(3.16) $`Df(u_o)^1vDg(u_o)^1v_{L^1}`$ $`=\underset{h0}{lim}{\displaystyle \frac{1}{h}}S_1^fu_hS_1^gu_h_{L^1}`$
$`\widehat{d}_{\text{lin}}(Df(u_o),Dg(u_o))`$ $`=\underset{v}{sup}\underset{h0}{lim}{\displaystyle \frac{1}{h}}S_1^fu_hS_1^gu_h_{L^1}`$
and the proof is completed.
## 4. The Classical Limit of the Relativistic Euler Equations
In this section we apply Corollary 2.5 to the classical limit of the relativistic Euler equations, generalizing what was obtained in .
The relativistic $`p`$–system, (see ) is
(4.1)
$$\{\begin{array}{c}\left[\rho +\left(\rho +\frac{1}{c^2}p\right)\frac{(v/c)^2}{1(v/c)^2}\right]_t+\left[\left(\rho +\frac{1}{c^2}p\right)\frac{v}{1(v/c)^2}\right]_x=0\\ \left[\left(\rho +\frac{1}{c^2}p\right)\frac{v}{1(v/c)^2}\right]_t+\left[\left(\rho +\frac{1}{c^2}p\right)\frac{v^2}{1(v/c)^2}+p\right]_x=0\end{array}$$
Above, $`\rho `$ is the mass–energy density of the fluid, $`v`$ the classical coordinate velocity, $`p`$ the pressure and $`c`$ the light speed. We show below that as $`c+\mathrm{}`$, the problem (4.1) approaches to its classical counterpart
(4.2)
$$\{\begin{array}{c}\left[\rho \right]_t+\left[\rho v\right]_x=0\\ \left[\rho v\right]_t+\left[\rho v^2+p\right]=0\end{array}$$
in the sense that the SRS $`S^c`$ generated by (4.1) converges to the SRS $`S`$ generated by (4.2) on a domain containing all physically reasonable data. In particular, the total variation of the data need not be small.
In (4.1) a standard choice (see and the references therein) for the pressure law is
(4.3)
$$p=\sigma ^2\rho ,$$
$`\sigma `$ being the sound speed.
Fix a positive lower bound for the density $`\rho _{\mathrm{min}}`$ and for the light speed $`c_o`$. Without any loss in generality, we may assume $`\sigma <c_o`$.
Let $`𝒱_M=\{(\rho ,\rho v)\text{BV}(,(\rho _{\mathrm{min}},+\mathrm{})\times ):\text{Tot.Var.}(\rho )+\text{Tot.Var.}(\rho v)M\}`$. In it is proved that for any $`M>0`$, (4.1) generates a SRS $`S^{c,M}`$ defined on a domain $`𝒟^{c,M}`$ containing $`𝒱_M`$ and consisting of functions of total variation bounded by, say, $`M^3`$ (provided $`M`$ is sufficiently large). Similarly, (4.2) generates a SRS $`S^M`$ on a domain $`𝒟^M`$ containing $`𝒱_{M^3}`$ and contained, say, in $`𝒱_{M^9}`$.
Thus, for all sufficiently large $`c`$ and $`M`$, there exist domains $`𝒟^{c,M}`$, $`𝒟^M`$ such that
(4.4)
$$𝒱_M𝒟^{c,M}𝒱_{M^3}𝒟^M𝒱_{M^9}$$
and, moreover, the problems (4.1) and (4.2) generate the SRSs
$$S^{c,M}:𝒟^{c,M}\times (0,+\mathrm{})𝒟^{c,M}\text{ and }S^M:𝒟^M\times (0,+\mathrm{})𝒟^M.$$
We are now ready to state and prove the following application of Corollary 2.5.
###### Theorem 4.1.
Fix a positive $`\rho _{\mathrm{min}}`$ and sufficiently large $`M`$, $`c_o`$. Let $`𝒟^{c,M}`$ and $`𝒟^M`$ satisfy to (4.4) for $`c>c_o`$. Then there exists a constant $`C`$ such that for all $`cc_0`$ and for all $`u𝒟^{M,c}`$
(4.5)
$$S_t^{c,M}uS_t^Mu_{L^1}C\frac{1}{c^2}t.$$
In particular, by (4.4), the bound (4.5) holds for all initial data $`u`$ with $`\text{Tot.Var.}(u)<M`$.
###### Proof.
Note that (4.1) and (4.2) in conservation form become, respectively,
(4.6)
$$\{\begin{array}{c}\rho _t+q_x=0\\ q_t+\left(\varphi _c(\rho ,q)\frac{q^2}{\rho }+p\right)_x=0\end{array}\text{ and }\{\begin{array}{c}\rho _t+q_x=0\\ q_t+\left(\frac{q^2}{\rho }+p\right)_x=0\end{array}$$
where
$$\varphi _c(\rho ,q)=1+\frac{1}{c^2}\left(1\frac{v^2(\rho ,q)}{c^2}\right)\frac{p}{\rho +{\displaystyle \frac{v^2(\rho ,q)}{c^2}}{\displaystyle \frac{p}{c^2}}}.$$
Call $`f_c`$ and, respectively, $`f`$ the fluxes in the two systems (4.6). Then, the estimate $`Df_cDf_{C^0}C(1/c^2)`$ follows from straightforward computations and completes the proof. ∎
We remark that the particular pressure law (4.3) is necessary only to ensure the existence of the SRS $`S^{c,M}`$ in the large. The above procedure remains true under much milder assumptions on the equation of state. Note moreover that the rate of convergence $`𝒪(1/c^2)`$ is exactly the one expected by the convergence of relativistic to classical mechanics.
It is of interest to mention that (4.5) proves also the uniform Lipschitz continuity of all semigroups $`S^{c,M}`$ for all sufficiently large $`c`$.
###### Remark 4.2.
As is well known, linearizing (4.2) around $`\rho =\rho _o`$, $`v=0`$ at constant entropy leads to the wave equation. Corollary 2.5 ensures that the solutions to (4.2) converge to the linearized equation in $`L^1`$ over finite time intervals.
## 5. Stability of a scalar equation w.r.t. flux
In this section we prove Theorem 2.6. We consider two scalar equation,
(5.1)
$$u_t+f(u)_x=0,$$
(5.2)
$$v_t+g(u)_x=0,$$
with the same initial condition: $`u(0,x)=v(0,x)=u_0`$. We assume that $`f`$, $`g`$ are strictly convex $`C^2`$ functions, precisely there exists a constant $`\kappa `$ such that
(5.3)
$$\underset{uK}{\mathrm{min}}\{|f^{\prime \prime }(u)|,|g^{\prime \prime }(u)|\}\kappa ,$$
where $`K`$ is a compact interval of $``$ such that $`u_0(x)K`$ for all $`x`$. We recall that by maximum principle also the entropic solutions of (5.1), (5.2) will satisfies the same bound.
We recall that, by , given a point $`(t,x)`$ we can consider the set of characteristics $`\xi (t,x)`$ passing through $`(t,x)`$. If we denote with $`\xi ^{}(t,x)`$ and $`\xi ^+(t,x)`$ the minimal and maximal backward characteristics, then either $`\xi ^{}(t,x)=\xi ^+(t,x)`$ and the solution $`u`$ is continuous in $`(t,x)`$, or we have an admissible shock and the jump is exactly given by the (constant) values of $`u`$ on the characteristics $`\xi ^{}`$, $`\xi ^+`$. By condition (5.3) we have that if at time $`t`$ two characteristics $`\xi ^+(t)`$ and $`\xi ^{}(t)`$ meet, then we have
(5.4)
$$\frac{d}{dt}(\xi ^+\xi ^{})\kappa (u^+u^{}).$$
Suppose that $`\xi ^{}(0)<\xi ^+(0)`$, and consider now an initial datum $`\stackrel{~}{u}_0`$ defined as
(5.5)
$$\stackrel{~}{u}_0(x)=\{\begin{array}{cc}u_0(x)\hfill & x\xi ^{}(0)\hfill \\ u_0^{}\hfill & \xi ^{}(0)<x\mathrm{\Xi }\hfill \\ u_0^+\hfill & \mathrm{\Xi }<x\xi ^+(0)\hfill \\ u_0(x)\hfill & x>\xi ^+(0)\hfill \end{array}$$
where $`\mathrm{\Xi }`$ is chosen such that
(5.6)
$$\mathrm{\Xi }\frac{1}{u_0^{}u_0^+}_\xi ^{}^{\xi ^+}u_0(x)𝑑x+\frac{u_0^{}\xi ^{}u_0^+\xi ^+}{u_0^{}u_0^+}.$$
Using the conservation of mass it is easy to conclude that the solution in unchanged at time $`t`$. In fact consider the triangle $`𝒯`$ whose vertices are $`(0,\xi ^{}(0))`$, $`(0,\xi ^+(0))`$ and $`(t,x)`$. Since the equation (5.1) can be written as
$$\text{div}\left(\begin{array}{c}u\\ f(u)\end{array}\right)=0,$$
and $`u`$ is constant along the lines $`\xi (t)`$, we have
$`{\displaystyle _𝒯(u_t+f(u)_x)𝑑t𝑑x}=`$ $`{\displaystyle _\xi ^{}^{\xi ^+}}u_0(x)𝑑x+{\displaystyle _0^t}(f(u^+)f^{}(u^+)u^+)𝑑t+{\displaystyle _0^t}(f^{}(u^{})u^{}f(u^{}))𝑑t`$
$`=`$ $`{\displaystyle _\xi ^{}^{\xi ^+}}u_0(x)𝑑x+(f^{}(u^{})u^{}f^{}(u^+)u^++f(u^+)f(u^{}))t=0.`$
Using the relation
$$x=\xi ^{}+f^{}(u^{})t=\mathrm{\Xi }+\frac{f(u^+)f(u^{})}{u^+u^{}}t=\xi ^++f^{}(u^+)t,$$
we obtain (5.6). Using (5.6) we can change the initial of (5.1) or (5.2) so that $`u(t)`$ or $`v(t)`$ are unchanged. Since with this procedure we collect all the interactions at time $`0`$, the total variation of $`\stackrel{~}{u}_0`$ has the same value of $`\text{Tot.Var.}(u(t))`$.
To change the initial datum in such a way that both $`u(t)`$ and $`v(t)`$ are the same, consider now the test system
(5.7)
$$w_t+\left(\frac{\kappa }{2}w^2\right)_x=0.$$
By (5.3) and (5.4) if two characteristics meet is $`u(t)`$ and $`v(t)`$, then they also meet in (5.7). Let us denote with $`\stackrel{~}{u}_0`$ the new initial condition, obtained by the above procedure using equation (5.7).
Consider now an interval $`[a,b]`$. By the definition of $`\widehat{\lambda }`$ we have that the values of $`u(t)`$ and $`v(t)`$ in $`[a,b]`$ depends only on $`\stackrel{~}{u}_0`$ in $`[a\widehat{\lambda }t,b+\widehat{\lambda }t]`$. Using the standard estimates we have
(5.8)
$$_a^b|u(t,x)v(t,x)|𝑑xt\underset{uK}{\mathrm{max}}|f^{}(u)g^{}(u)|\text{Tot.Var.}(\stackrel{~}{u}_0;[a\widehat{\lambda }t,b+\widehat{\lambda }t]).$$
To estimate the total variation, an easy computation gives
(5.9)
$$\text{Tot.Var.}(\stackrel{~}{u}_0;[a\widehat{\lambda }t,b+\widehat{\lambda }t])\text{Tot.Var.}(w(t);[a2\widehat{\lambda }t,b+2\widehat{\lambda }t])2\text{diam}(K)\frac{ba+4\widehat{\lambda }t}{\kappa t}.$$
Combining (5.8) and (5.9) we get
(5.10)
$$_a^b|u(t,x)v(t,x)|𝑑x\underset{uK}{\mathrm{max}}\left|f^{}(u)g^{}(u)\right|2\text{diam}(K)\frac{ba+4\widehat{\lambda }t}{\kappa }.$$
###### Remark 5.1.
When $`t0`$ the integral does not converge to $`0`$. This is clear since the initial datum is in $`L^{\mathrm{}}`$, and then the semigroup is continuous but not Lipschitz continuous in time, since the amount of interaction at $`t=0`$ is infinite. Consider for example the following two equations
$$u_t+\left(\frac{u^2}{2}\right)_x=0,v_t+\left(v+\frac{v^2}{2}\right)_x=0,$$
with the periodic initial condition
$$u_{0,n}(x)=\{\begin{array}{cc}1\hfill & \text{if }x[k2^{n+1},k2^{n+1}+2^n],k\hfill \\ 1\hfill & \text{otherwise}\hfill \end{array}$$
At time $`t_n=2^n`$ the two solution are
$$\{\begin{array}{cc}u_n(t_n,x)=2^n(xk2^{n+1})\hfill & \text{if }x(k2^{n+1}2^n,k2^{n+1}+2^n]\hfill \\ v_n(t_n,x)=2^n(xk2^{n+1}2^n)\hfill & \text{if }x(k2^{n+1},(k+1)2^{n+1}]\hfill \end{array}$$
so that
$$_0^1|v_n(t_n,x)u_n(t_n,x)|dx=1.$$
This depends on the fact that the modulus of continuity of the semigroup can be arbitrarily large.
Note moreover that since the solutions $`u`$, $`v`$ are limits of wave front tracking approximations, the continuous dependence of the solution on the flux function $`f`$ can be stated also if $`f`$ non convex. However in general one cannot prove any uniform continuous dependence.
Acknowledgments. The present work was partly accomplished while the authors visited the Max Planck Institute in Leipzig supported by the European TMR Network *“Hyperbolic Conservation Laws”* ERBMRXCT960033.
|
warning/0007/hep-ph0007312.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Reggeization of elementary particles in the non-Abelian gauge theories , particularly in QCD, seems to be one of the most intriguing properties of these theories. Origin and consequences of this property are not completely understood, but its role in the high energy behaviour of scattering amplitudes hardly can be overestimated. The gluon Reggeization determines the high energy behaviour of non-decreasing with energy cross sections in perturbative QCD. In particular, it appears to be the basis of the BFKL approach .
Originally the BFKL equation was derived in the leading logarithm approximation (LLA), that means summation of all terms of perturbation theory, where smallness of the coupling constant $`\alpha _s`$ is compensated by the large logarithm $`\mathrm{ln}s`$ of the squared c.m.s. energy. Two years ago the first radiative correction to the kernel of the BFKL equation for the forward scattering, i.e. $`t=0`$ and colour singlet state in the $`t`$-channel, was found . It gives the possibility to sum also the next-to-leading order (NLO) terms. The correction was derived assuming the validity of the gluon Reggeization in the NLO, which is not proved, and therefore must be carefully tested. That can be done by checking the bootstrap equations appearing as the conditions of the compatibility of the gluon Reggeization with the $`s`$-channel unitarity. The second, not less important reason for this check is that it gives the over-crossed test of the calculations of almost all ingredients used in the BFKL approach. In this approach the high energy scattering amplitudes are given by the convolution of the impact factors of scattered particles and the Green function for the Reggeon-Reggeon scattering. The impact factors and the kernel of the BFKL equation for the Green function, in turn, are expressed through the effective vertices for the Reggeon-particle interactions and the gluon Regge trajectory. The first bootstrap equation ties the kernel of the BFKL equation for the antisymmetric colour octet state of the two Reggeized gluons in the $`t`$-channel $`𝒦^{(8)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})`$ with the gluon trajectory $`j=1+\omega (t)`$. In the NLO the equation takes the form
$$\frac{g^2Nt}{2\left(2\pi \right)^{D1}}\frac{d^{D2}q_1}{\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\stackrel{}{q}_2^2}𝒦^{(8)(1)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})=\omega ^{\left(1\right)}\left(t\right)\omega ^{\left(2\right)}\left(t\right)$$
(1)
where $`g`$ is the coupling constant, $`N`$ is the number of colours ($`N=3`$ for QCD), $`D=4+2ϵ`$ is the space-time dimension taken different from 4 for regularization of infrared divergences. $`𝒦^{(8)(1)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})`$ is the first radiative correction to the octet kernel, the vector sign is used for denoting component momenta transverse to the plane of momenta of initial particles, $`\stackrel{}{q}^{}`$is the momentum transfer in the scattering process. For the sake of brevity we put $`\stackrel{}{q}_i^{}=\stackrel{}{q}_i^{}\stackrel{}{q}`$, $`i=1,2`$; $`\stackrel{}{q}_1^{}`$ and $`\stackrel{}{q}_1^{}`$ ( $`\stackrel{}{q}_2^{}`$ and $`\stackrel{}{q}_2^{}`$) being the transverse momenta of the initial (final) Reggeized gluons in the $`t`$-channel.
The second bootstrap equation connects the colour octet impact factor of any particle and its effective interaction vertex with the Reggeized gluon. This equation was checked for the gluon and quark impact factors. It was proved that the equation is fulfilled both for helicity conserving and non-conserving parts of the impact factors at arbitrary space-time dimension $`D`$. As for the first bootstrap equation, Eq. (1), it was checked (and was found to be satisfied also at arbitrary $`D`$) only in the quark part of this equation . The most important (and much more complicated) gluon part of this equation remained unchecked till now. Recently, the gluon part of the radiative correction to $`𝒦^{\left(8\right)}`$ was obtained , that gives a possibility to analyze Eq. (1) in this part.
In this paper we check the first bootstrap condition (1). Since the fulfillment of this equation for the quark part is established , we consider pure gluodynamics. In the next section we present the explicit expressions for the octet BFKL kernel and the gluon trajectory, and analyse Eq. (1). The proof of its fulfillment is given in Section 3. In the last section we discuss possible generalizations.
## 2 Explicit Form of the Bootstrap Equation
Let us start with the gluon trajectory. Its one-loop expression is well known:
$$\omega ^{(1)}(t)=\frac{g^2Nt}{2\left(2\pi \right)^{D1}}\frac{d^{D2}k}{\stackrel{}{k}^2(\stackrel{}{q}\stackrel{}{k})^2}.$$
(2)
The integral can be easily calculated at arbitrary $`D`$ and one has
$$\omega ^{(1)}(t)=\overline{g}^2(\stackrel{}{q}^2)^ϵ\frac{\mathrm{\Gamma }^2(ϵ)}{\mathrm{\Gamma }(2ϵ)},$$
(3)
where $`\mathrm{\Gamma }`$ is the Euler gamma-function and
$$\overline{g}^2=\frac{g^2N\mathrm{\Gamma }(1ϵ)}{(4\pi )^{D/2}}.$$
(4)
We stress that everywhere in this paper we use the unrenormalized coupling constant $`g`$.
In the two-loop approximation the integral representation for the trajectory is
$$\omega ^{(2)}(t)=\frac{g^2t}{(2\pi )^{D1}}\frac{d^{D2}q_1}{\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\stackrel{}{q}_1^{}_{}{}^{}2}\left[F_G(\stackrel{}{q}_1^{},\stackrel{}{q})2F_G(\stackrel{}{q}_1,\stackrel{}{q}_1)\right],$$
(5)
where
$$F_G(\stackrel{}{q}_1^{},\stackrel{}{q})=\frac{g^2N^2\stackrel{}{q}^{\mathrm{\hspace{0.17em}2}}}{4\left(2\pi \right)^{D1}}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\left(\stackrel{}{q}_2^{}\stackrel{}{q}\right)^2}[\mathrm{ln}\left(\frac{\stackrel{}{q}^{\mathrm{\hspace{0.17em}2}}}{\left(\stackrel{}{q}_1^{}\stackrel{}{q}_2^{}\right)^2}\right)2\psi (1+2ϵ)$$
$$\psi (1ϵ)+2\psi \left(ϵ\right)+\psi \left(1\right)+\frac{1}{1+2ϵ}(\frac{1}{ϵ}+\frac{1+ϵ}{2\left(3+2ϵ\right)})],$$
(6)
and $`\psi (x)=\mathrm{\Gamma }^{}(x)/\mathrm{\Gamma }(x)`$. The integral (5) can be expressed in terms of elementary functions only for $`ϵ0`$. The answer is
$$\omega ^{(2)}(t)\left(\frac{\overline{g}^2\left(\stackrel{}{q}^{\mathrm{\hspace{0.17em}2}}\right)^ϵ}{ϵ}\right)^2\left[\frac{11}{3}+\left(2\psi ^{}(1)\frac{67}{9}\right)ϵ+\left(\frac{404}{27}+\psi ^{\prime \prime }(1)\frac{22}{3}\psi ^{}(1)\right)ϵ^2\right].$$
(7)
The kernel of the BFKL equation for any of the colour states $``$ of two Reggeized gluons in the $`t`$-channel is given by the sum of the “virtual” part, defined by the gluon trajectory, and the “real” part $`𝒦_r^{\left(\right)}`$, related to the real particle production:
$$𝒦^{\left(\right)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})=\left(\omega \left(\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\right)+\omega \left(\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\right)\right)\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\delta ^{(D2)}\left(\stackrel{}{q}_1^{}\stackrel{}{q}_2^{}\right)+𝒦_r^{\left(\right)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q}).$$
(8)
The octet kernel in the Born approximation differs only by the coefficient 1/2 from the singlet (Pomeron) kernel:
$$𝒦_r^{\left(8\right)\left(B\right)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})=\frac{g^2N}{2\left(2\pi \right)^{D1}}f_B,f_B=\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2+\stackrel{}{q}_2^2\stackrel{}{q}_1^2}{\stackrel{}{k}^2}\stackrel{}{q}^2,$$
(9)
where $`\stackrel{}{k}=\stackrel{}{q}_1^{}`$ $`\stackrel{}{q}_2^{}`$. The radiative correction to this kernel is known only in the limit $`ϵ0`$ . It can be presented as the sum of three contributions:
$$𝒦_r^{\left(8\right)\left(1\right)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})=\frac{\overline{g}^4}{\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ)}\left(𝒦_1+𝒦_2+𝒦_3\right).$$
(10)
The first contribution is proportional to the Born kernel (9):
$$𝒦_1=f_B\frac{(\stackrel{}{k}^{\mathrm{\hspace{0.17em}2}})^ϵ}{ϵ}\left[\frac{11}{3}+\left(2\psi ^{}(1)\frac{67}{9}\right)ϵ+\left(\frac{404}{27}+7\psi ^{\prime \prime }(1)\frac{11}{3}\psi ^{}(1)\right)ϵ^2\right].$$
(11)
Note that this part of the kernel contains all explicit singularities in $`ϵ`$, as it should be, because of factorization of singularities in QCD amplitudes. Note also that in Eq. (11) the term $`(\stackrel{}{k}^2)^ϵ`$ is not expanded in $`ϵ`$ and the terms of order $`ϵ`$ are kept in the coefficient. It is done because the kernel is singular at $`\stackrel{}{k}=0`$, so that, at subsequent integrations of the kernel in the BFKL equation, the region of arbitrary small, for $`ϵ0`$, values of $`\stackrel{}{k}^{\mathrm{\hspace{0.17em}2}}`$, where $`ϵ|\mathrm{ln}\stackrel{}{k}^{\mathrm{\hspace{0.17em}2}}|1`$, does contribute and, moreover, leads to the appearance of an extra 1/$`ϵ`$ factor. Consequently, the terms $`ϵ`$ are kept in the coefficient to save all non-vanishing for $`ϵ0`$ terms after the integration. The last two terms in Eq. (10) are known for $`ϵ0`$ only. To stress this circumstance we shall indicate them with the superscript <sup>(0)</sup>. The second contribution can be expressed in terms of logarithms:
$$𝒦_2^{(0)}=\{\stackrel{}{q}^2[\frac{11}{6}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2}{\stackrel{}{q}^2\stackrel{}{k}^2}\right)+\frac{1}{4}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}^2}\right)\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}^2}\right)+\frac{1}{4}\mathrm{ln}\left(\frac{\stackrel{}{q}_2^2}{\stackrel{}{q}^2}\right)\mathrm{ln}\left(\frac{\stackrel{}{q}_2^2}{\stackrel{}{q}^2}\right)+\frac{1}{4}\mathrm{ln}^2\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}_2^2}\right)]$$
$$\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2+\stackrel{}{q}_2^2\stackrel{}{q}_1^2}{2\stackrel{}{k}^2}\mathrm{ln}^2\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}_2^2}\right)+\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2\stackrel{}{q}_2^2\stackrel{}{q}_1^2}{\stackrel{}{k}^2}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}_2^2}\right)(\frac{11}{6}\frac{1}{4}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2}{\stackrel{}{k}^4}\right))\}+\{\stackrel{}{q}_i\stackrel{}{q}_i^{}\}$$
(12)
while the last cannot be written through elementary functions. It has the following integral representation:
$$𝒦_3^{(0)}=\{\frac{1}{2}[\stackrel{}{q}^2(\stackrel{}{k}^2\stackrel{}{q}_1^2\stackrel{}{q}_2^2)+2\stackrel{}{q}_1^2\stackrel{}{q}_2^2\stackrel{}{q}_1^2\stackrel{}{q}_2^2\stackrel{}{q}_2^2\stackrel{}{q}_1^2+\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2\stackrel{}{q}_2^2\stackrel{}{q}_1^2}{\stackrel{}{k}^2}(\stackrel{}{q}_1^2\stackrel{}{q}_2^2)]$$
$$\times _0^1\frac{dx}{(\stackrel{}{q}_1(1x)+\stackrel{}{q}_2x)^2}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2(1x)+\stackrel{}{q}_2^2x}{\stackrel{}{k}^2x(1x)}\right)\}+\{\stackrel{}{q}_i\stackrel{}{q}_i^{}\}.$$
(13)
Knowing the “real” kernel only in the limit $`ϵ0`$, we can think about the check by the bootstrap equation of only the terms in the gluon trajectory non-vanishing for $`ϵ0`$. But we shall see that even for this purpose the knowledge of the kernel (10) is not sufficient. The reason comes from the singular behaviour of the integration measure in the bootstrap condition (1). Let us note that terms with the trajectory contribution in the formula (8) for the kernel can be integrated in Eq. (1) in a general form, even without the knowledge of an explicit expression for the trajectory. The matter is that the dependence of $`\omega ^{(2)}(t)`$ on $`t`$ is dictated by the dimension of the bare coupling constant, so that
$$\omega ^{(2)}(t)=\overline{g}^4(\stackrel{}{q}^2)^{2ϵ}N_ϵ,$$
(14)
where $`N_ϵ`$ is some coefficient depending on $`ϵ`$ (whose expression in the limit $`ϵ0`$ is given by Eq. (7)). Using the generic integral
$$\frac{1}{(2\pi )^{D1}}\frac{d^{D2}k}{(\stackrel{}{k}^2)^{1\delta _1}((\stackrel{}{k}\stackrel{}{q})^2)^{1\delta _2}}=$$
$$\frac{2(\stackrel{}{q}^2)^{ϵ+\delta _1+\delta _21}}{(4\pi )^{2+ϵ}}\frac{B(ϵ+\delta _1,ϵ+\delta _2)\mathrm{\Gamma }(1\delta _1\delta _2ϵ)}{\mathrm{\Gamma }(1\delta _1)\mathrm{\Gamma }(1\delta _2)},$$
(15)
where $`B(x,y)`$ is the Euler beta-function, putting $`\delta _1=2ϵ`$, $`\delta _2=0`$, from Eqs. (1), (3) and (8) we obtain
$$(\stackrel{}{q}^2)^{13ϵ}\frac{d^{D2}q_1}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\stackrel{}{q}_2^2}\frac{𝒦_r^{\left(8\right)\left(1\right)}(\stackrel{}{q}_1,\stackrel{}{q}_2;\stackrel{}{q})}{\overline{g}^4\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ)}=$$
$$N_ϵB(ϵ,ϵ)\left[1+\frac{B(ϵ,3ϵ)}{ϵB(1ϵ,2ϵ)B(ϵ,ϵ)}\right].$$
(16)
The integration measure in Eq. (16) is singular at zero momenta of scattered Reggeons. Fortunately, the kernel (10) turns into zero at this points , as it can be checked by direct inspection of Eqs. (11) - (13), so that at first sight these points could not bring additional singularities in $`ϵ`$. It would be so if integration over only one momentum was performed. But in the region where the momenta of two Reggeons (let us say, $`\stackrel{}{q}_1^{}`$ and $`\stackrel{}{q}_2^{}`$) turn into zero simultaneously, being of the same order, the kernel does not vanish (as it can be easily observed in the example of the Born kernel (9)). Therefore, these regions can give additional singularities in $`ϵ`$; moreover, since integration over these regions leads to singularities, an expansion of, let us say, $`(\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}})^ϵ`$ and $`(\stackrel{}{q}_2^{\mathrm{\hspace{0.17em}2}})^ϵ`$ is not possible anymore, so that we need to know about the kernel more than that is given by Eqs. (10) - (13).
## 3 Proof of the Bootstrap
In the limit $`ϵ0`$ the coefficient $`N_ϵ`$ Eq. (14) is defined from Eq. (7). Our aim is to check the non-vanishing, for $`ϵ0`$, term in $`N_ϵ`$. Using the explicit expression for $`N_ϵ`$ we can rewrite Eq. (16) as
$$(\stackrel{}{q}^2)^{13ϵ}\frac{d^{D2}q_1}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\stackrel{}{q}_2^2}\frac{𝒦_1+𝒦_2+𝒦_3}{(\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ))^2}$$
$$\frac{2}{3ϵ^3}\left[\frac{11}{3}+\left(2\psi ^{}(1)\frac{67}{9}\right)ϵ+\left(\frac{404}{27}11\psi ^{}(1)+\psi ^{\prime \prime }(1)\right)ϵ^2\right],$$
(17)
where in the R.H.S. we have kept only terms singular in $`ϵ`$ (that corresponds to non-vanishing terms in $`\omega ^{(2)}`$). We need to calculate the L.H.S. of Eq. (17) with the same accuracy. Let us consider where the terms singular in $`ϵ`$ can come from. In general, there are two kinds of singularities: the singularity already existing in the kernel (namely, that in the expression (11) for $`𝒦_1`$), and the singularities appearing as the results of integration in Eq. (17). The last singularities, in turn, can come from two kinds of integration regions: one is the region $`|\stackrel{}{k}|0`$, where the kernel (again $`𝒦_1`$, and only this part of the kernel) is singular, and the other consists in the regions where the transverse momenta of the two Reggeons simultaneously tend to zero, as it was explained at the end of the preceding section. These last regions do not overlap each others, but overlap with the region $`|\stackrel{}{k}|0`$. It is clear that the most dangerous is the term $`𝒦_1`$ of the kernel, proportional to the Born kernel, which gives the singularities of all kinds. Nevertheless, it is not very difficult to calculate the contribution of this part in Eq. (17). The main point here is that we know this part of the kernel sufficiently well in the singular regions.
First of all, let us understand, in which of the regions, where the momenta of the two Reggeons are simultaneously tending to zero, the part $`𝒦_1`$ can contribute. Such regions are four, in general:
$`1)`$ $`|\stackrel{}{q}_1||\stackrel{}{q}_2|0;\stackrel{}{q}_1^{}\stackrel{}{q}_2^{}\stackrel{}{q},|\stackrel{}{k}|0,`$ (18)
$`2)`$ $`|\stackrel{}{q}_1^{}||\stackrel{}{q}_2^{}|0;\stackrel{}{q}_1\stackrel{}{q}_2\stackrel{}{q},|\stackrel{}{k}|0,`$
$`3)`$ $`|\stackrel{}{q}_1||\stackrel{}{q}_2^{}|0;\stackrel{}{q}_1^{}\stackrel{}{q}_2\stackrel{}{k}\stackrel{}{q},`$
$`4)`$ $`|\stackrel{}{q}_1^{}||\stackrel{}{q}_2|0;\stackrel{}{q}_1\stackrel{}{q}_2^{}\stackrel{}{k}\stackrel{}{q}.`$
Because of the symmetry of $`𝒦_1`$ with respect to $`\stackrel{}{q}_i\stackrel{}{q}_i^{}`$ it is sufficient to consider only the first and third regions. One can easily see that in the third region $`𝒦_10`$, so that only the first region remains. However, this is the region of small $`|\stackrel{}{k}|`$, where we know the kernel sufficiently well, so that for the calculation we do not need more information than that shown in Eq. (11).
The calculation could be done (and it was done) straightforwardly, but a more sophisticated way of calculation leads to our aim through a much more easy way. Let us first of all put
$$d\rho =\frac{1}{\pi ^{2ϵ+2}\mathrm{\Gamma }^2\left(1ϵ\right)}\frac{d^{D2}q_1}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\stackrel{}{q}_2^2}.$$
(19)
We need to calculate the terms non-vanishing with $`ϵ`$ in the integral
$$J=𝑑\rho f_B\left(\frac{\stackrel{}{k}^2}{\stackrel{}{q}^2}\right)^ϵ𝑑\rho f_B+𝑑\rho f_B\left(\left(\frac{\stackrel{}{k}^2}{\stackrel{}{q}^2}\right)^ϵ1\right)\theta \left(\mu ^2\stackrel{}{k}^2\right)$$
$$+𝑑\rho f_B\left(\left(\frac{\stackrel{}{k}^2}{\stackrel{}{q}^2}\right)^ϵ1\right)\theta \left(\stackrel{}{k}^2\mu ^2\right),$$
(20)
where $`\theta \left(x\right)`$ is the usual step function and $`\mu `$ is chosen in such a way that the ratio $`\mu ^2/\stackrel{}{q}^2ϵ^n1`$, $`n`$ being arbitrary fixed integer number. Then last integral is of order $`ϵ`$ (since we can expand here $`\left(\stackrel{}{k}^2/\stackrel{}{q}^2\right)^ϵ`$ and the regions of singularities are excluded by the $`\theta `$-function) and can be neglected, whereas the first one can be easily calculated with the help of Eq. (15) and gives
$$𝑑\rho f_B\left(\stackrel{}{q}^2\right)^{2ϵ1}\left(\frac{2}{ϵ^2}4\psi ^{}\left(1\right)\right).$$
(21)
In the second integral we can perform the first integration over $`\stackrel{}{q}_1`$ keeping $`\stackrel{}{k}`$ fixed. In terms of these variables one can write
$$\frac{f_B}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2\stackrel{}{q}_2^2\stackrel{}{q}_2^2}=\frac{1}{\stackrel{}{k}^2}\left(\frac{1}{\left(\stackrel{}{q}_1^{}\stackrel{}{q}\right)^2\left(\stackrel{}{q}_1\stackrel{}{k}\right)^2}+\frac{1}{\stackrel{}{q}_1^2\left(\stackrel{}{q}_1\stackrel{}{k}\stackrel{}{q}\right)^2}\right)$$
$$\frac{\stackrel{}{q}^2}{\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\left(\stackrel{}{q}_1^{}\stackrel{}{q}\right)^2\left(\stackrel{}{q}_1^{}\stackrel{}{k}\right)^2\left(\stackrel{}{q}_1^{}\stackrel{}{k}\stackrel{}{q}\right)^2}.$$
(22)
The first two terms can be easily integrated over $`\stackrel{}{q}__1`$; as for the last term, two non-overlapping regions do contribute to the integral at small $`|\stackrel{}{k}|`$: $`|\stackrel{}{q}_1||\stackrel{}{k}|`$ and $`|\stackrel{}{q}_1\stackrel{}{q}||\stackrel{}{k}|`$. They give equal contributions, so it is sufficient considering the first of them, where we can put
$$\frac{\stackrel{}{q}^2}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2\stackrel{}{q}_2^2\stackrel{}{q}_2^2}\frac{1}{\stackrel{}{q}^{\mathrm{\hspace{0.17em}2}}\stackrel{}{q}_1^{\mathrm{\hspace{0.17em}2}}\left(\stackrel{}{q}_1^{}\stackrel{}{k}\right)^2},$$
(23)
and then expand the integration region at all $`\stackrel{}{q}_1`$, due to the convergence of the integral. In such a way we obtain
$$𝑑\rho f_B\left(\left(\frac{\stackrel{}{k}^2}{\stackrel{}{q}^2}\right)^ϵ1\right)\theta \left(\mu ^2\stackrel{}{k}^2\right)$$
$$\frac{2}{\stackrel{}{q}^2}\frac{d^{D2}k}{\pi ^{ϵ+1}\mathrm{\Gamma }\left(1ϵ\right)}\frac{1}{\stackrel{}{k}^2}\left(\left(\frac{\stackrel{}{k}^2}{\stackrel{}{q}^2}\right)^ϵ1\right)\theta \left(\mu ^2\stackrel{}{k}^2\right)B(ϵ,ϵ)\left(\left(\stackrel{}{q}^2\right)^ϵ\left(\stackrel{}{k}^2\right)^ϵ\right).$$
(24)
Using the relation
$$\frac{d^{D2}k}{\pi ^{ϵ+1}\mathrm{\Gamma }\left(1ϵ\right)}=\frac{\left(\stackrel{}{k}^2\right)^ϵd\stackrel{}{k}^2}{\mathrm{\Gamma }(1ϵ)\mathrm{\Gamma }(1+ϵ)}$$
(25)
we obtain
$$𝑑\rho f_B\left(\left(\frac{\stackrel{}{k}^2}{\stackrel{}{q}^2}\right)^ϵ1\right)\theta \left(\mu ^2\stackrel{}{k}^2\right)\left(\stackrel{}{q}^2\right)^{2ϵ1}\left[\frac{4}{3ϵ^2}+\frac{8}{3}\psi ^{}\left(1\right)\right].$$
(26)
This result, together with Eq. (21) gives
$$J\left(\stackrel{}{q}^2\right)^{2ϵ1}\left[\frac{2}{3ϵ^2}\frac{4}{3}\psi ^{}\left(1\right)\right],$$
(27)
and, consequently,
$$\left(\stackrel{}{q}^2\right)^{13ϵ}\frac{d^{D2}q_1}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\stackrel{}{q}_2^2}\frac{𝒦_1}{(\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ))^2}$$
$$\frac{2}{3ϵ^3}\left[\frac{11}{3}+\left(2\psi ^{}(1)\frac{67}{9}\right)ϵ+\left(\frac{404}{27}11\psi ^{}(1)+7\psi ^{\prime \prime }(1)\right)ϵ^2\right].$$
(28)
This result coincides with that of the straightforward calculation which was also performed. Comparing with Eq. (17) we see that the contribution of the term proportional to the Born kernel almost “saturates” the bootstrap condition; the only difference is that, instead of the term $`7\psi ^{\prime \prime }(1)`$ of Eq. (28), in the R.H.S. of Eq. (17) there is $`\psi ^{\prime \prime }(1)`$.
We now turn to the remaining contributions to the kernel, $`𝒦_2`$ and $`𝒦_3`$. They have neither explicit singularities in $`ϵ`$, nor singular behaviour for $`|\stackrel{}{k}|0`$. Therefore, terms of order 1/$`ϵ`$ can be obtained only from the regions 1) - 4) (see Eq. (18)), where the momenta of the two Reggeons tend to zero being of the same order. It is easy to see from Eqs. (12) and (13) that in the regions 3) and 4) both $`𝒦_2^{(0)}`$ and $`𝒦_3^{(0)}`$ turn into zero. Using the symmetry of the kernel with respect to $`\stackrel{}{q}_i\stackrel{}{q}_i^{}`$, we can consider only the first region. Here we have
$$𝒦_2^{(0)}\stackrel{}{q}^2[\frac{11}{6}(\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2}{\stackrel{}{k}^4}\right)+\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2}{\stackrel{}{k}^2}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}_2^2}\right))+(\frac{1}{4}\frac{\stackrel{}{q}_1^2+\stackrel{}{q}_2^2}{2\stackrel{}{k}^2})\mathrm{ln}^2\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}_2^2}\right)$$
$$\frac{1}{4}\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2}{\stackrel{}{k}^2}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2}{\stackrel{}{q}_2^2}\right)\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2\stackrel{}{q}_2^2}{\stackrel{}{k}^4}\right)],$$
(29)
$$𝒦_3^{(0)}\frac{\stackrel{}{q}^2}{2}\left[\stackrel{}{k}^2\stackrel{}{q}_1^2\stackrel{}{q}_2^2+\frac{2\left(\stackrel{}{q}_1^2(\stackrel{}{q}_2\stackrel{}{k})\stackrel{}{q}_2^2(\stackrel{}{q}_1\stackrel{}{k})\right)}{\stackrel{}{k}^2}\right]I,$$
(30)
where
$$I=_0^1\frac{dx}{(\stackrel{}{q}_1(1x)+\stackrel{}{q}_2x)^2}\mathrm{ln}\left(\frac{\stackrel{}{q}_1^2(1x)+\stackrel{}{q}_2^2x}{\stackrel{}{k}^2x(1x)}\right).$$
(31)
If we fix $`\stackrel{}{k}`$ and perform the integration over $`\stackrel{}{q}_1`$, then it is convergent at $`|\stackrel{}{q}_1||\stackrel{}{k}|`$ since
$$d\rho \frac{1}{\pi ^{2ϵ+2}\mathrm{\Gamma }^2\left(1ϵ\right)}\frac{d^{D2}k}{\stackrel{}{q}^4}\frac{d^{D2}q_1}{\stackrel{}{q}_1^2(\stackrel{}{q}_1\stackrel{}{k})^2}.$$
(32)
Therefore with such order of integration at fixed small $`\stackrel{}{k}`$ we can integrate over all $`\stackrel{}{q}_1`$ and nevertheless use the expressions (29) - (31).
All integrals of the terms entering into Eq. (29) for $`𝒦_2^{(0)}`$ can be taken using Eq. (15) and its derivatives with respect to $`\delta _1`$ and $`\delta _2`$. All the integrals of this part can be calculated for arbitrary $`ϵ`$. We are interested in the limit $`ϵ0`$, but in the region of arbitrary small $`\stackrel{}{k}`$, so that we cannot expand $`(\stackrel{}{k}^2)^ϵ`$ in powers of $`ϵ`$. In this limit we obtain
$$\frac{d^{D2}q_1}{\stackrel{}{q}_1^2(\stackrel{}{q}_1\stackrel{}{k})^2}\frac{𝒦_2^{(0)}}{\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ)}\frac{\stackrel{}{q}^2}{\stackrel{}{k}^2}\left(\stackrel{}{k}^2\right)^ϵ3\psi ^{\prime \prime }(1).$$
(33)
The integration of Eq. (30) for $`𝒦_3^{(0)}`$ is not straightforward. We use the representation
$$I=_0^1𝑑x_1^{\mathrm{}}\frac{dz}{z}\frac{1}{\left(\stackrel{}{q}_1x\stackrel{}{k}\right)^2+zx(1x)\stackrel{}{k}^2}=$$
$$_0^1𝑑x_1^{\mathrm{}}\frac{dz}{z}\frac{1}{x(\stackrel{}{q}_1\stackrel{}{k})^2+(1x)\stackrel{}{q}_1^2+(z1)x(1x)\stackrel{}{k}^2}.$$
(34)
After this the integration over $`\stackrel{}{q}_1`$ is reduced to integrations of terms with two denominators, which are performed using the standard Feynman parametrization. Remaining integrations over $`x`$, $`z`$ and the Feynman parameter are rather tedious, although not very complicated. The result for $`ϵ0`$ is the same as for $`𝒦_2^{(0)}`$:
$$\frac{d^{D2}q_1}{\stackrel{}{q}_1^2(\stackrel{}{q}_1\stackrel{}{k})^2}\frac{𝒦_3^{(0)}}{\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ)}\frac{\stackrel{}{q}^2}{\stackrel{}{k}^2}\left(\stackrel{}{k}^2\right)^ϵ3\psi ^{\prime \prime }(1).$$
(35)
Again we do not expand $`\left(\stackrel{}{k}^2\right)^ϵ`$, because at the subsequent integration over $`\stackrel{}{k}`$ the region where $`ϵ\left|\mathrm{ln}\stackrel{}{k}^2\right|1`$ does contribute. Note that appearance of the factors $`\left(\stackrel{}{k}^2\right)^ϵ`$ in Eqs. (33) and (35) is evident without calculations: they are dimensional factors and in the integrals (33) and (35) we have only one dimensional parameter.
We see that scale factors can be important even in calculation of terms singular in $`ϵ`$ which come from the nonsingular contributions $`𝒦_2`$ and $`𝒦_3`$ in Eq. (17). But we know $`𝒦_2`$ and $`𝒦_3`$ only in the limit $`ϵ0`$ at fixed Reggeon momenta, where scale factors are put equal to 1 (see Eqs. (12) and (13)). The problem therefore is to restore the scale factors. Fortunately, we need to know the scale factors only in the first of regions (18), where we have only two essentially different scales since we have $`|\stackrel{}{q}_1||\stackrel{}{q}_2||\stackrel{}{k}|`$ from one side and $`|\stackrel{}{q}_1^{}||\stackrel{}{q}_2^{}||\stackrel{}{q}|`$ from the other. Even without calculations it is clear that the relevant scale should be $`|\stackrel{}{k}|`$, since the scale appears as a result of integration over transverse momenta (transverse momenta of gluons produced in the Reggeon-Reggeon collision for the two-gluon contribution to the kernel, transverse momenta of virtual gluons in the radiative correction to the Reggeon-Reggeon-gluon (RRG) vertex for the one-gluon contribution), and for $`|\stackrel{}{q}_1||\stackrel{}{q}_2||\stackrel{}{k}|`$ in the essential region of the integration these momenta should be of the same order. This conclusion is confirmed by direct inspection of the kernel. Fortunately, for the two-gluon contribution this inspection can be done straightforwardly, since there is the explicit expression for this contribution for arbitrary $`D`$ . The one-gluon contribution consists of two pieces: in one of them radiative corrections are contained in the RRG vertex with momenta $`q_1`$ and $`q_2`$, in the other with momenta $`q_1^{}`$ and $`q_2^{}`$. It follows from the kinematical structure of the RRG vertex that in the region 1) only the first piece does contribute to $`𝒦_2`$ ($`𝒦_3`$ comes totally from the two-gluon production). The evident scale for this piece is $`|\stackrel{}{k}|`$. Therefore, in the first of regions (18) we have
$$𝒦_2=\left(\stackrel{}{k}^2\right)^ϵ𝒦_2^{(0)},𝒦_3=\left(\stackrel{}{k}^2\right)^ϵ𝒦_3^{(0)}.$$
(36)
Using Eqs. (33) and (35), and taking into account the region 2) by doubling the results we obtain
$$\stackrel{}{q}^2\frac{d^{D2}q_1}{\stackrel{}{q}_1^2\stackrel{}{q}_1^2}\frac{d^{D2}q_2}{\stackrel{}{q}_2^2\stackrel{}{q}_2^2}\frac{𝒦_2+𝒦_3}{(\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ))^2}$$
$$12\psi ^{\prime \prime }(1)\frac{d^{D2}k}{\pi ^{1+ϵ}\mathrm{\Gamma }(1ϵ)}\left(\stackrel{}{k}^2\right)^{2ϵ1}\theta \left(\stackrel{}{q}^2\stackrel{}{k}^2\right)\frac{4\psi ^{\prime \prime }(1)}{ϵ}.$$
(37)
Putting Eqs. (28) and (37) into Eq. (17) we see that the bootstrap condition is satisfied.
## 4 Conclusions
We have proved the fulfillment of the first bootstrap condition in the form of Eq. (1) in the limit of the space-time dimension $`D`$ tending to its physical value $`D=4`$. All terms in the two-loop contribution to the gluon trajectory and in the NLO correction to the octet BFKL kernel, non-vanishing in this limit, are involved in this bootstrap. Therefore, the performed verification of the bootstrap gives us not only a powerful confirmation of the gluon Reggeization, but at the same time a stringent test of the calculations of the trajectory and the kernel.
Now we have practically no doubts on the gluon Reggeization in the NLO, as well as on the calculations of gluon trajectory and kernel. Nevertheless it would be interesting to verify if the first bootstrap condition is satisfied for arbitrary space-time dimension $`D`$. That is known to be true for the quark part of the kernel and for the second bootstrap condition in the cases of quark and gluon impact factors. It will become possible to do this verification after the calculation of the one-loop correction to the Reggeon-Reggeon-gluon vertex for arbitrary $`D`$, which is in progress now . Another interesting possibility is to check the so-called “first strong bootstrap condition” for the kernel suggested by Braun and Vacca . This condition is derived from the requirement that the particle-Reggeon scattering amplitudes have a Reggeized form and it is satisfied for the quark part of the kernel, as well as the analogous condition for impact factors in the quark and gluon cases, although the role of the strong bootstrap conditions in the BFKL approach is not completely understood.
Acknowledgment: Two of us (V.S.F. and M.I.K.) thank the Dipartimento di Fisica della Università della Calabria for the warm hospitality while a part of this work was done. V.S.F. acknowledges also the financial support of the Istituto Nazionale di Fisica Nucleare.
|
warning/0007/gr-qc0007016.html
|
ar5iv
|
text
|
# Solution of Dirac equation around a spinning Black Hole
## I. INTRODUCTION
One of the most important solutions of Einstein’s equation is that of the spacetime around and inside an isolated black hole. The spacetime at a large distance is flat and Minkowskian where usual quantum mechanics is applicable, while the spacetime closer to the singularity is so curved that no satisfactory quantum field theory could be developed as yet. An intermediate situation arises when a weak perturbation (due to, say, gravitational, electromagnetic or Dirac waves) originating from infinity impinges on a black hole, interacting with it. The resulting wave is partially transmitted into the black hole through the horizon and partially scatters off to infinity. In the linearized (‘test field’) approximation this problem has been attacked in the past by several authors \[1-4\]. The master equations of Teukolsky which govern these linear perturbations for integral spin (e.g., gravitational and electromagnetic) fields were solved numerically by Press & Teukolsky and Teukolsky & Press . While the equations governing the massive Dirac particles were separated by Chandrasekhar . So far, only the angular eigenfunction and eigenvalue (which happens to be the separation constant) have been obtained . Particularly interesting is the fact that whereas gravitational and electromagnetic radiations were found to be amplified in some range of incoming frequencies, Chandrasekhar predicted that no such amplifications should take place for Dirac waves because of the very nature of the potential experienced by the incoming fields. However, these later conclusions were drawn using asymptotic solutions and no attempt has so far been made to determine the nature of the radial wave functions, both incoming and outgoing, for the Dirac wave perturbations. He also speculated that one needs to look into the problem for negative eigenvalues ($`\lambda `$) where one might come across super-radiance for Dirac waves.
In the present paper, we revisit this important problem to study the nature of the radial wave functions as a function of the Kerr parameter, rest mass and frequency of incoming particle. We also verify that super-radiance is indeed absent for the Dirac field. Unlike the works of Press & Teukolsky and Teukolsky & Press where numerical (shooting) methods were used to solve the master equations governing gravitational and electromagnetic waves, we use an approximate analytical method for the massive Dirac wave. The details of the method would be presented below.
The plan of the paper is as follows: in the next Section, we present the equation governing the Dirac waves (waves for half-integral massive spin particles) as they were separated into radial and angular co-ordinates. We then briefly present the nature of the angular eigenvalues and eigenfunctions. In §3, we present our method of solution and present the spatially complete radial wave functions. Finally, in §4, we draw our conclusions.
## II. THE DIRAC EQUATION IN KERR GEOMETRY
Chandrasekhar separated the Dirac equation in Kerr geometry into radial ($`R`$) and angular ($`S`$) wave functions. Below, we present these equations from Chandrasekhar using the same choice of units: we choose $`\text{}h=1=G=c`$.
The equations governing the radial wave-functions $`R_{\pm \frac{1}{2}}`$ corresponding to spin $`\pm \frac{1}{2}`$ respectively are given by:
$$\mathrm{\Delta }^{\frac{1}{2}}𝒟_0R_{\frac{1}{2}}=(\text{}\lambda +im_pr)\mathrm{\Delta }^{\frac{1}{2}}R_{+\frac{1}{2}},$$
$`(1a)`$
$$\mathrm{\Delta }^{\frac{1}{2}}𝒟_0^{}\mathrm{\Delta }^{\frac{1}{2}}R_{+\frac{1}{2}}=(\text{}\lambda im_pr)R_{\frac{1}{2}},$$
$`(1b)`$
where, the operators $`𝒟_n`$ and $`𝒟_n^{}`$ are given by,
$$𝒟_n=_r+\frac{iK}{\mathrm{\Delta }}+2n\frac{(rM)}{\mathrm{\Delta }},$$
$`(2a)`$
$$𝒟_n^{}=_r\frac{iK}{\mathrm{\Delta }}+2n\frac{(rM)}{\mathrm{\Delta }},$$
$`(2b)`$
and
$$\mathrm{\Delta }=r^2+a^22Mr,$$
$`(3a)`$
$$K=(r^2+a^2)\sigma +am.$$
$`(3b)`$
Here, $`a`$ is the Kerr parameter, $`n`$ is an integer or half integer, $`\sigma `$ is the frequency of incident wave, $`M`$ is the mass of the black hole, $`m_p`$ is the rest mass of the Dirac particle, $`\lambda `$ is the eigenvalue of the Dirac equation and $`m`$ is the azimuthal quantum number.
The equations governing the angular wave-functions $`S_{\pm \frac{1}{2}}`$ corresponding to spin $`\pm \frac{1}{2}`$ respectively are given by:
$$_{\frac{1}{2}}S_{+\frac{1}{2}}=(\text{}\lambda am_p\mathrm{cos}\theta )S_{\frac{1}{2}}$$
$`(4a)`$
$$_{\frac{1}{2}}^{}S_{\frac{1}{2}}=+(\text{}\lambda +am_p\mathrm{cos}\theta )S_{+\frac{1}{2}}$$
$`(4b)`$
where, the operators $`_n`$ and $`_n^{}`$ are given by,
$$_n=_\theta +Q+n\mathrm{cot}\theta ,$$
$`(5a)`$
$$_n^{}=_\theta Q+n\mathrm{cot}\theta $$
$`(5b)`$
and
$$Q=a\sigma \mathrm{sin}\theta +m\mathrm{cosec}\theta .$$
$`(6)`$
Note that both the radial and the angular sets of equations i.e., eqs. 1(a-b) and eqs. 4(a-b) are coupled equations. Combining eqs. 4(a-b), one obtains the angular eigenvalue equations for the spin-$`\frac{1}{2}`$ particles as
$$\left[_{\frac{1}{2}}_{\frac{1}{2}}^{}+\frac{am_p\mathrm{sin}\theta }{\text{}\lambda +am_p\mathrm{cos}\theta }_{\frac{1}{2}}^{}+(\text{}\lambda ^2a^2m_p^2\mathrm{cos}^2\theta )\right]S_{\frac{1}{2}}=0.$$
$`(7)`$
There are exact solutions of this equation for the eigenvalues $`\lambda `$ and the eigenfunctions $`S_{\frac{1}{2}}`$ when $`\rho =\frac{m_p}{\sigma }=1`$ in terms of the orbital quantum number $`l`$ and azimuthal quantum number $`m`$. These solutions are :
$$\text{}\lambda ^2=(l+\frac{1}{2})^2+a\sigma (p+2m)+a^2\sigma ^2\left[1\frac{y^2}{2(l+1)+a\sigma x}\right],$$
$`(8)`$
and
$${}_{\frac{1}{2}}{}^{}S_{lm}^{}={}_{\frac{1}{2}}{}^{}Y_{lm}^{}\frac{a\sigma y}{2(l+1)+a\sigma x}{}_{\frac{1}{2}}{}^{}Y_{l+1m}^{}$$
$`(9)`$
where,
$$p=F(l,l);x=F(l+1,l+1);y=F(l,l+1)$$
and
$$F(l_1,l_2)=[(2l_2+1)(2l_1+1)]^{\frac{1}{2}}<l_21m0|l_1m>$$
$$[<l_21\frac{1}{2}0|l_1\frac{1}{2}>+(1)^{l_2l}<l_21m0|l_1m>[<l_21\frac{1}{2}0|l_1\frac{1}{2}>+(1)^{l_2l}\rho \sqrt{2}<l_21\frac{1}{2}1|l_1\frac{1}{2}>]].$$
$`(10)`$
with $`<\mathrm{}.|..>`$ are the usual Clebsh-Gordon coefficients. For other values of $`\rho `$ one has to use perturbation theories. Solutions upto sixth order using perturbation parameter $`a\sigma `$ is given in Chakrabarti . The eigenfunctions $`\lambda `$ are required to solve the radial equations which we do now.
This radial equations 1(a-b) are in coupled form. One can decouple them and express the equation either in terms of spin up or spin down wave functions $`R_{\pm \frac{1}{2}}`$ but the expression loses its transparency. It is thus advisable to use the approach of Chandrasekhar by changing the basis and independent variable $`r`$ to
$$r_{}=r+\frac{2Mr_++am/\sigma }{r_+r_{}}\mathrm{log}\left(\frac{r}{r_+}1\right)\frac{2Mr_{}+am/\sigma }{r_+r_{}}\mathrm{log}\left(\frac{r}{r_{}}1\right),$$
$`(11)`$
(for $`r>r_+`$),
$$\frac{d}{dr_{}}=\frac{\mathrm{\Delta }}{\omega ^2}\frac{d}{dr},$$
$`(12)`$
$$\omega ^2=r^2+\alpha ^2;\alpha ^2=a^2+am/\sigma ,$$
$`(13)`$
to transform the set of coupled equations 1(a-b) into two independent one dimensional wave equations given by:
$$\left(\frac{d}{dr_{}}i\sigma \right)P_{+\frac{1}{2}}=\frac{\mathrm{\Delta }^{\frac{1}{2}}}{\omega ^2}(\text{}\lambda im_pr)P_{\frac{1}{2}}$$
$`(14)`$
$$\left(\frac{d}{dr_{}}+i\sigma \right)P_{\frac{1}{2}}=\frac{\mathrm{\Delta }^{\frac{1}{2}}}{\omega ^2}(\text{}\lambda +im_pr)P_{+\frac{1}{2}}.$$
$`(15)`$
Here, $`𝒟_0=\frac{\omega ^2}{\mathrm{\Delta }}(\frac{d}{dr_{}}+i\sigma )`$ and $`𝒟_0^{}=\frac{\omega ^2}{\mathrm{\Delta }}(\frac{d}{dr_{}}i\sigma )`$ were used and wave functions were redefined as $`R_{\frac{1}{2}}=P_{\frac{1}{2}}`$ and $`\mathrm{\Delta }^{\frac{1}{2}}R_{+\frac{1}{2}}=P_{+\frac{1}{2}}`$.
We now define a new variable,
$$\theta =tan^1(m_pr/\text{}\lambda )$$
$`(16)`$
which yields,
$$\mathrm{cos}\theta =\frac{\text{}\lambda }{\sqrt{(\text{}\lambda ^2+m_p^2r^2)}},\mathrm{and}\mathrm{sin}\theta =\frac{m_pr}{\sqrt{(\text{}\lambda ^2+m_p^2r^2)}}$$
and
$$(\text{}\lambda \pm im_pr)=exp(\pm i\theta )\sqrt{(\text{}\lambda ^2+m_p^2r^2)},$$
$`(17)`$
so the coupled equations take the form,
$$\left(\frac{d}{dr_{}}i\sigma \right)P_{+\frac{1}{2}}=\frac{\mathrm{\Delta }^{\frac{1}{2}}}{\omega ^2}(\text{}\lambda ^2+m_p^2r^2)^{1/2}P_{\frac{1}{2}}exp\left[i\mathrm{tan}^1\left(\frac{m_pr}{\text{}\lambda }\right)\right],$$
$`(18a)`$
and
$$\left(\frac{d}{dr_{}}+i\sigma \right)P_{\frac{1}{2}}=\frac{\mathrm{\Delta }^{\frac{1}{2}}}{\omega ^2}(\text{}\lambda ^2+m_p^2r^2)^{1/2}P_{+\frac{1}{2}}exp\left[i\mathrm{tan}^1\left(\frac{m_pr}{\text{}\lambda }\right)\right].$$
$`(18b)`$
Then defining,
$$P_{+\frac{1}{2}}=\psi _{+\frac{1}{2}}exp\left[\frac{1}{2}itan^1\left(\frac{m_pr}{\text{}\lambda }\right)\right]$$
$`(19a)`$
and
$$P_{\frac{1}{2}}=\psi _{\frac{1}{2}}exp\left[+\frac{1}{2}itan^1\left(\frac{m_pr}{\text{}\lambda }\right)\right],$$
$`(19b)`$
we obtain,
$$\frac{d\psi _{+\frac{1}{2}}}{dr_{}}i\sigma \left(1+\frac{\mathrm{\Delta }}{\omega ^2}\frac{\text{}\lambda m_p}{2\sigma }\frac{1}{\text{}\lambda ^2+m_p^2r^2}\right)\psi _{+\frac{1}{2}}=\frac{\mathrm{\Delta }^{\frac{1}{2}}}{\omega ^2}(\text{}\lambda ^2+m_p^2r^2)^{1/2}\psi _{\frac{1}{2}}$$
$`(20a)`$
and
$$\frac{d\psi _{\frac{1}{2}}}{dr_{}}+i\sigma \left(1+\frac{\mathrm{\Delta }}{\omega ^2}\frac{\text{}\lambda m_p}{2\sigma }\frac{1}{\text{}\lambda ^2+m_p^2r^2}\right)\psi _{\frac{1}{2}}=\frac{\mathrm{\Delta }^{\frac{1}{2}}}{\omega ^2}(\text{}\lambda ^2+m_p^2r^2)^{1/2}\psi _{+\frac{1}{2}}.$$
$`(20b)`$
Further choosing $`\widehat{r}_{}=r_{}+\frac{1}{2\sigma }\mathrm{tan}^1(\frac{m_pr}{\text{}\lambda })`$ so that $`d\widehat{r}_{}=(1+\frac{\mathrm{\Delta }}{\omega ^2}\frac{\text{}\lambda m_p}{2\sigma }\frac{1}{\text{}\lambda ^2+m_p^2r^2})dr_{}`$, the above equations become,
$$\left(\frac{d}{d\widehat{r}_{}}i\sigma \right)\psi _{+\frac{1}{2}}=W\psi _{\frac{1}{2}},$$
$`(21a)`$
and
$$\left(\frac{d}{d\widehat{r}_{}}+i\sigma \right)\psi _{\frac{1}{2}}=W\psi _{+\frac{1}{2}}.$$
$`(21b)`$
where,
$$W=\frac{\mathrm{\Delta }^{\frac{1}{2}}(\text{}\lambda ^2+m_p^2r^2)^{3/2}}{\omega ^2(\text{}\lambda ^2+m_p^2r^2)+\text{}\lambda m_p\mathrm{\Delta }/2\sigma }.$$
$`(22)`$
Now letting $`Z_\pm =\psi _{+\frac{1}{2}}\pm \psi _{\frac{1}{2}}`$ we can combine the differential equations to give,
$$\left(\frac{d}{d\widehat{r}_{}}W\right)Z_+=i\sigma Z_{},$$
$`(23a)`$
and
$$\left(\frac{d}{d\widehat{r}_{}}+W\right)Z_{}=i\sigma Z_+.$$
$`(23b)`$
¿From these equations, we readily obtain a pair of independent one-dimensional wave equations,
$$\left(\frac{d^2}{d\widehat{r}_{}^{}{}_{}{}^{2}}+\sigma ^2\right)Z_\pm =V_\pm Z_\pm .$$
$`(24)`$
where, $`V_\pm =W^2\pm \frac{dW}{d\widehat{r}_{}}`$
$$=\frac{\mathrm{\Delta }^{\frac{1}{2}}(\text{}\lambda ^2+m_p^2r^2)^{3/2}}{[\omega ^2(\text{}\lambda ^2+m_p^2r^2)+\text{}\lambda m_p\mathrm{\Delta }/2\sigma ]^2}[\mathrm{\Delta }^{\frac{1}{2}}(\text{}\lambda ^2+m_p^2r^2)^{3/2}\pm ((rM)(\text{}\lambda ^2+m_p^2r^2)+3m_p^2r\mathrm{\Delta })]$$
$$\frac{\mathrm{\Delta }^{\frac{3}{2}}(\text{}\lambda ^2+m_p^2r^2)^{5/2}}{[\omega ^2(\text{}\lambda ^2+m_p^2r^2)+\text{}\lambda m_p\mathrm{\Delta }/2\sigma ]^3}[2r(\text{}\lambda ^2+m_p^2r^2)+2m_p^2\omega ^2r+\text{}\lambda m_p(rM)/\sigma ].$$
$`(25)`$
One important point to note: the transformation of spatial co-ordinate $`r`$ to $`r_{}`$ (and $`\widehat{r}_{}`$) is taken not only for mathematical simplicity but also for a physical significance. When $`r`$ is chosen as the radial co-ordinate, the decoupled equations for independent waves show diverging behaviour. However, by transforming those in terms of $`r_{}`$ (and $`\widehat{r}_{}`$) we obtain well behaved functions. The horizon is shifted from $`r=r_+`$ to $`\widehat{r}_{}=\mathrm{}`$ unless $`\sigma \sigma _s=am/2Mr_+`$ (eq. 11). In this connection, it is customary to define $`\sigma _c`$ where $`\alpha ^2=0`$ (eq. 13). Thus, $`\sigma _c=m/a`$. If $`\sigma \sigma _s`$, the region is expected to be super-radiant because for integral spin particles for $`\sigma sigma_s`$ there exhibit super-radiation.
## III. SOLUTION OF THE RADIAL EQUATION
Out of the total physical parameter space, in one region (region I) the total energy of the particle is always greater than the height of potential barrier and in the other region (region II) the energy is less than of the maximum height of the potential barrier. In region II, the wave hits the wall of barrier and tunnels through it. One has to treat these two cases a little differently.
The usual WKB approximation is used to obtain the zeroth order solution. We improve the solution by properly incorporating the inner and outer boundary conditions. After establishing the general solution, we present here the solution of eq. (4) for three sets of parameters as illustrative examples. For those examples the choice of parameters is made in such a way that there is a significant interaction between the particle and the black hole, i.e., when the Compton wavelength of the incoming wave is of the same order as the radius of the outer horizon of the Kerr black hole. So,
$$\frac{G[M+\sqrt{(}M^2a^2)]}{c^2}\frac{\text{}h}{m_pc}.$$
$`(26)`$
We choose as before $`G=\text{}h=c=1`$, so
$$m_p\frac{1}{[M+\sqrt{(}M^2a^2)]}.$$
$`(27)`$
Similarly, the frequency of the incoming particle (or wave) should be of the same order as the inverse of the light crossing time of the radius of the black hole, i.e.,
$$\frac{c^3}{G[M+\sqrt{(}M^2a^2)]}\sigma .$$
$`(28)`$
Using the same units as before, we can write,
$$m_p\sigma [M+\sqrt{(}M^2a^2)]^1.$$
$`(29)`$
In principle, however, one can choose any values of $`\sigma `$ and $`m_p`$ for a particular black hole and the corresponding solution is possible.
One can easily check from equation (25) that for $`r\mathrm{}`$ (i.e., $`\widehat{r}_{}\mathrm{}`$) $`V_\pm m_p^2`$. So we expand the total parameter space in terms of the frequency of the particle (or wave), $`\sigma `$ and the rest mass of the particle, $`m_p`$. It is clear that in half of the parameter space spanned by $`\sigma m_p`$ where, $`\sigma <m_p`$, particles are released at finite distance with so little energy that they cannot escape to infinity. In this case, the total energy $`\sigma ^2`$ of the incoming particle at a large distance is less than the potential energy of the system. We will not discuss solutions in this region. The rest of the parameter space ($`\sigma m_p`$) is divided into two regions – I: $`E>V_m`$ and II: $`E<V_m`$, where $`E`$ is the total energy of the incoming particle and $`V_m`$ is the maximum of the potential. In Region I, the wave is locally sinusoidal because the wave number $`k`$ is real for the entire range of $`\widehat{r}_{}`$. In Region II, on the other hand, the wave is decaying in some region when $`E<V`$, i.e., where the wave ‘hits’ the potential barrier and in the rest of the region, the wave is propagating. We shall show solutions in these two regions separately. In Region-I whatever be the physical parameters, the energy of the particle is always greater than the potential energy and the WKB approximation is generally valid in the whole range (i.e. $`\frac{1}{k^2}\frac{dk}{d\widehat{r}_{}}<<1`$). In cases of Region-II, the energy of the particle is always less than the maximum height of potential barrier. Thus, at two points (where, $`k=0`$) the total energy matches the potential energy and in the neighbourhood of those two points the WKB approximate method is not valid. They have to be dealt with separately. In Fig. 1, we show contours of constant $`w_{max}=`$max($`\frac{1}{k^2}\frac{dk}{d\widehat{r}_{}}`$) for a given set ($`\sigma ,m_p`$) of parameters. The labels show the actual values of $`w_{max}`$. Clearly, in most of the parameter regions the WKB approximation is safely valid for any value of $`\widehat{r}_{}`$. One has to employ a different method (such as using Airy Functions, see below) to find solutions in those regions where $`w_{max}`$ attains a large value which indicates the non-validity of WKB method.
### Solutions of Region I
We re-write equation (24) as,
$$\frac{d^2Z_+}{d\widehat{r}_{}^2}+(\sigma ^2V_+)Z_+=0.$$
$`(30)`$
This is nothing but the Schrödinger equation with total energy of the wave $`\sigma ^2`$. This can be solved by regular WKB method.
Let $`k(\widehat{r}_{})=\sqrt{(\sigma ^2V_+)}`$, $`u(\widehat{r}_{})=k(\widehat{r}_{})𝑑\widehat{r}_{}+\mathrm{constant}`$. $`k`$ is the wavenumber of the incoming wave and $`u`$ as the Eikonal. The solution of the equation (30) is,
$$Z_+=\frac{A_+}{\sqrt{k}}exp(iu)+\frac{A_{}}{\sqrt{k}}exp(iu).$$
$`(31)`$
with
$$A_+^2+A_{}^2=k.$$
$`(32)`$
The motivation of equation (32) is to impose the WKB method at the each space point so that sum of the transmission and reflection coefficients are same at each location. In this case $`\sigma ^2>V_+`$ all along and also $`\frac{1}{k}\frac{dk}{d\widehat{r}_{}}<<k`$, so the WKB approximation is valid in the whole region.
It is clear that a standard WKB solution where $`A_+`$ and $`A_{}`$ are kept constant throughout, can not be accurate in whole range of $`\widehat{r}_{}`$, since the physical inner boundary condition on the horizon must be that the reflected component is negligible there (since there the potential barrier height goes down to zero). Thus the WKB approximation requires a slight modification in which a spatial dependence of $`A_\pm `$ is allowed. On the other hand, at a large distance, where the WKB is strictly valid, $`A_+`$ and $`A_{}`$ should tend to be constants, and hence their difference is also a constant:
$$A_+A_{}=c.$$
$`(33)`$
Here, one can choose also the sum of $`A_+`$ and $`A_{}`$ are constant instead of difference as equation (33), but the final result will not be affected. Here, $`c`$ is determined from the WKB solution at a large distance. For simplicity we choose $`A_\pm `$s are real. This along with (32) gives,
$$A_\pm (r)=\pm \frac{c}{2}+\frac{\sqrt{[2k(r)c^2]}}{2}.$$
$`(34)`$
This spatial variation, strictly valid at large distances only, should not be extendible to the horizon without correcting for the inner boundary condition. These values are to be shifted by, say, $`A_{\pm h}`$ respectively, so that on the horizon one obtains the physical $`R`$ and $`T`$. We first correct the reflection coefficient on the horizon as follows: Let $`A_h`$ be the value of $`A_{}`$ on the horizon (see, eq. 34),
$$A_h=\frac{c}{2}+\frac{\sqrt{[2k(r_+)c^2]}}{2}.$$
It is appropriate to use $`𝒜_{}=A_{}A_h`$ rather than $`A_{}`$ since $`𝒜_{}`$ vanishes at $`r=r_+`$.
Incorporating these conditions, the solution (31) becomes,
$$Z_+=\frac{𝒜_+}{\sqrt{q}}exp(iu)+\frac{𝒜_{}}{\sqrt{q}}exp(iu).$$
$`(35)`$
with the usual normalization condition
$$𝒜_+^2+𝒜_{}^2=q.$$
$`(36)`$
where, $`𝒜_+=A_+A_{+h}`$.
Determination of $`A_{+h}`$ is done by enforcing $`R`$ obtained from eq. (37a), which is shown below, is the same as that obtained by the actual WKB method. The $`q`$ is used to compute the transmission coefficient $`T`$ from eq. (36). In this way, normalization of $`R+T=1`$ is assured.
The normalization factor $`qk`$ as $`\widehat{r}_{}\mathrm{}`$ and the condition $`\frac{1}{q}\frac{dq}{d\widehat{r}_{}}<<q`$ is found to be satisfied whenever $`\frac{1}{k}\frac{dk}{d\widehat{r}_{}}<<k`$ is satisfied. This is the essence of our modification of the WKB. In a true WKB, $`A_\pm `$ are constants and the normalization is with respect to a (almost) constant $`k`$. However, we are using it as if the WKB is instantaneously valid everywhere. Our method may therefore be called ‘Instantaneous’ WKB approximation or IWKB for short. Using the new notations, the instantaneous values (i.e., local values) of the reflection and transmission coefficients are given by (see, eq. 35),
$$R=\frac{𝒜_{}^2}{q}$$
$`(37a)`$
$$T=\frac{𝒜_+^2}{q}.$$
$`(37b)`$
Whatever may be the value of the physical parameters, $`\frac{1}{k}\frac{dk}{d\widehat{r}_{}}<<k`$ is satisfied in whole range of $`\widehat{r}_{}`$ for region I.
The variation of reflection and transmission coefficients would be well understood if we imagine the potential barrier consists of a large number of steps. From simple quantum mechanics, in between each two steps, we can calculate the reflection and transmission coefficients . Clearly these reflection and transmission coefficients at different junctions will be different. This is discussed in detail below. To be concrete, we choose one set of parameters from Region I. Here, the total energy of the incoming particle is greater than the potential barrier height for all values of $`\widehat{r}_{}`$. We use, Kerr parameter, $`a=0.5`$; mass of the black hole, $`M=1`$; Mass of the particle, $`m_p=0.8`$; orbital angular momentum quantum number, $`l=\frac{1}{2}`$; azimuthal quantum number, $`m=\frac{1}{2}`$; frequency of the incoming wave, $`\sigma =0.8`$. The derived parameters are, $`r_+=M+\sqrt{(M^2a^2)}1.86603`$; $`\sigma _c=1`$; $`\sigma _s=0.066987`$; $`\alpha ^2=0.0625`$. For these parameters, the eigenvalue is $`\text{}\lambda =0.92`$ .
Here it is clear that $`\sigma `$ is in between $`\sigma _c`$ and $`\sigma _s`$ and $`\alpha ^2<0`$, $`r_+>|\alpha |`$. So we are in a strictly non super-radiant regime since here, $`\sigma >\sigma _s`$ .
¿From eq. (24) we observe that there are two wave equations for two potentials $`V_+`$ and $`V_{}`$. The nature of the potentials is shown in Fig. 2a. It is clear from the Fig. 2a that the potentials $`V_\pm `$ are well behaved. They are monotonically decreasing as the particle approaches the black hole, and the total energy chosen in this case ($`\sigma ^2`$) is always higher than $`V_\pm `$. For concreteness, we study the equation with potential $`V_+`$. A similar procedure (IWKB method) as explained above can be adopted using the potential $`V_{}`$ to compute $`Z_{}`$ and its form would be
$$Z_{}=\frac{A_+^{}A_{+h}^{}}{\sqrt{q^{}}}exp(iu^{})\frac{A_{}^{}A_h^{}}{\sqrt{q^{}}}exp(iu^{}).$$
$`(35^{})`$
Note the occurance of the negative sign in front of the reflected wave. This is to satisfy the asymptotic property of the wave functions.
In Fig. 2b, we show the nature of $`V_+`$ (solid curve), $`k`$ (dashed curve) and $`E(=\sigma ^2)`$ (short-dashed curve). In the solutions (eq. $`35`$ and $`35^{}`$) the first term corresponds to the incident wave and the second term corresponds to the reflected wave.
In Fig. 2c, the variation of reflection and transmission coefficients are shown. It is seen that as matter comes close to the black hole, the barrier height goes down. As a result, the penetration probability increases, causing the rise of the transmission coefficient.
Local values of the reflection and transmission co-efficients could also be calculated using the well known quantum mechanical approach. First one has to replace the potentials (as shown in Fig. 2a) by a collection of step functions as shown in Fig. 3a. The standard junction conditions of the type,
$$Z_{+,n}=Z_{+,n+1}$$
$`(38a)`$
where,
$$Z_{+,n}=A_nexp[ik_n\widehat{r}_{,n}]+B_nexp[ik_n\widehat{r}_{,n}]$$
and
$$\frac{dZ_+}{d\widehat{r}_{}}|_n=\frac{dZ_+}{d\widehat{r}_{}}|_{n+1}$$
$`(38b)`$
where,
$$\frac{dZ_+}{d\widehat{r}_{}}|_n=ik_nA_nexp(ik_n\widehat{r}_{,n})ik_nB_nexp(ik_n\widehat{r}_{,n})$$
at each of the $`n`$ steps were used to connect solutions at successive steps. ¿From the simple quantum mechanical calculation we obtain the reflection and transmission coefficients at the each junctions. Clearly at different junctions i.e., at different radii this reflection and transmission coefficients will be different. As before, we use the inner boundary condition, to be $`R0`$ at $`\widehat{r}_{}\mathrm{}`$. In reality we use as many steps as possible to follow accurately the shape of the potential. Smaller step sizes were used whenever $`k`$ varies faster. Fig. 3b shows the comparison of the instantaneous reflection coefficients in both the methods. The agreement shows that the WKB can be used at each point quite successfully.
It is to be noted, that, strictly speaking, the terms ‘reflection’ and ‘transmission’ coefficients are traditionally defined with respect to the asymptotic values. The spatial dependence that we show are just the dependence of the instantaneous values. This is consistent with the spirit of IWKB approximation that we are using.
The radial wave functions $`R_{+\frac{1}{2}}`$ and $`R_{\frac{1}{2}}`$ which are of spin up and spin down particles respectively of the original Dirac equation are given below,
$$Re(R_{\frac{1}{2}}\mathrm{\Delta }^{\frac{1}{2}})=\frac{a_+\mathrm{cos}(u\theta )+a_{}\mathrm{cos}(u+\theta )}{2\sqrt{k}}+\frac{a_+^{}\mathrm{cos}(u^{}\theta )a_{}^{}\mathrm{cos}(u^{}+\theta )}{2\sqrt{k^{}}}$$
$`(39a)`$
$$Im(R_{\frac{1}{2}}\mathrm{\Delta }^{\frac{1}{2}})=\frac{a_+\mathrm{sin}(u\theta )a_{}\mathrm{sin}(u+\theta )}{2\sqrt{k}}+\frac{a_+^{}\mathrm{sin}(u^{}\theta )+a_{}^{}\mathrm{sin}(u^{}+\theta )}{2\sqrt{k^{}}}$$
$`(39b)`$
$$Re(R_{\frac{1}{2}})=\frac{a_+\mathrm{cos}(u+\theta )+a_{}\mathrm{cos}(u\theta )}{2\sqrt{k}}\frac{a_+^{}\mathrm{cos}(u^{}+\theta )a_{}^{}\mathrm{cos}(u^{}\theta )}{2\sqrt{k^{}}}$$
$`(39c)`$
$$Im(R_{\frac{1}{2}})=\frac{a_+\mathrm{sin}(u+\theta )a_{}\mathrm{sin}(u\theta )}{2\sqrt{k}}\frac{a_+^{}\mathrm{sin}(u^{}+\theta )+a_{}^{}\mathrm{sin}(u^{}\theta )}{2\sqrt{k^{}}}$$
$`(39d)`$
Here, $`a_+=(A_+A_{+h})/\sqrt{(}q/k)`$ and $`a_{}=(A_{}A_h)/\sqrt{(}q/k)`$. $`\frac{a_+^{}}{\sqrt{k^{}}}`$ and $`\frac{a_{}^{}}{\sqrt{k^{}}}`$ are the transmitted and reflected amplitudes respectively for the wave of corresponding potential $`V_{}`$.
In Fig. 4(a-d) we show the nature of these wavefunctions. The eikonals used in plotting these functions (see, eq. 39\[a-d\]) have been calculated by approximating $`V_\pm `$ in terms of a polynomial and using the definition $`u(\widehat{r}_{})=\sqrt{(}\sigma ^2V_\pm )d\widehat{r}_{}`$. This was done since $`V_\pm `$ is not directly integrable. Note that the amplitude as well as wavelength remain constants in regions where $`k`$ is also constant.
### Solutions of Region II
Here we study the solution of a region where for any set of physical parameters, the total energy of the incoming particle is less than the maximum height of the potential barrier. So the WKB approximation (more precisely, our IWKB approximation) is not valid in the whole range of $`\widehat{r}_{}`$. In the regions where the WKB is not valid, the solutions will be the linear combination of Airy functions because the potential is a linear function of $`\widehat{r}_{}`$ in those intervals. At the junctions one has to match the solutions including Airy functions with the solution obtained by WKB method.
In the region where the WKB approximation is valid, local values of reflection and transmission coefficients and the wave function can be calculated easily by following the same method described in previous sub-section (solutions of region I) and the solution will be same as equation ($`35,35^{}`$). In other regions, the equation reduces to
$$\frac{d^2Z_+}{d\widehat{r}_{}^2}xZ_+=0$$
$`(40)`$
where $`x=\beta ^{\frac{1}{3}}(\widehat{r}_{}p)`$, $`\beta `$ is chosen to be positive and $`p`$ is the critical point where the total energy and potential energy are matching.
Let $`Z_+(x)=x^{\frac{1}{2}}Y(x)`$ and considering region $`x>0`$ the equation (40) reduces to
$$x^2\frac{d^2Y}{dx^2}+x\frac{dY}{dx}\left(x^3+\frac{1}{4}\right)Y(x)=0.$$
$`(41)`$
By making another transformation
$$\xi =\frac{2}{3}x^{\frac{3}{2}}$$
$`(42)`$
we obtain
$$\xi ^2\frac{d^2Y}{d\xi ^2}+\xi \frac{dY}{d\xi }\left(\xi ^2+\frac{1}{9}\right)Y(\xi )=0,$$
$`(43)`$
this is the modified Bessel equation. The solution of this equation is $`I_{+\frac{1}{3}}(\xi )`$ and $`I_{\frac{1}{3}}(\xi )`$. So the solution of eq. (40) will be
$$Z_+(x)=x^{\frac{1}{2}}[C_1I_{+\frac{1}{3}}(\xi )+C_2I_{\frac{1}{3}}(\xi )].$$
$`(44)`$
When $`x<0`$ the corresponding equation is,
$$\xi ^2\frac{d^2Y}{d\xi ^2}+\xi \frac{dY}{d\xi }+\left(\xi ^2\frac{1}{9}\right)Y(\xi )=0,$$
$`(45)`$
which is the Bessel equation. The corresponding solution is
$$Z_+(x)=|x|^{\frac{1}{2}}[D_1J_{+\frac{1}{3}}(\xi )+D_2J_{\frac{1}{3}}(\xi )],$$
$`(46)`$
where $`J_{\pm \frac{1}{3}}`$ and $`I_\pm \frac{1}{3}`$ are the Bessel functions and the modified Bessel functions of order $`\frac{1}{3}`$ respectively.
The Airy functions are defined as
$$Ai(x)=\frac{1}{3}x^{\frac{1}{2}}[I_{\frac{1}{3}}(\xi )I_{+\frac{1}{3}}(\xi )],x>0,$$
$`(47)`$
$$Ai(x)=\frac{1}{3}|x|^{\frac{1}{2}}[J_{\frac{1}{3}}(\xi )+J_{+\frac{1}{3}}(\xi )],x<0,$$
$`(48)`$
$$Bi(x)=\frac{1}{\sqrt{3}}x^{\frac{1}{2}}[I_{\frac{1}{3}}(\xi )+I_{+\frac{1}{3}}(\xi )],x>0,$$
$`(49)`$
$$Bi(x)=\frac{1}{\sqrt{3}}|x|^{\frac{1}{2}}[J_{\frac{1}{3}}(\xi )J_{+\frac{1}{3}}(\xi )],x<0.$$
$`(50)`$
In terms of Airy functions, the solutions (44) and (46) can be written as
$$Z_+=\frac{3}{2}(C_2C_1)Ai(x)+\frac{\sqrt{3}}{2}(C_2+C_1)Bi(x)\mathrm{for}x>0,$$
$`(51)`$
$$Z_+=\frac{3}{2}(D_2+D_1)Ai(x)+\frac{\sqrt{3}}{2}(D_2D_1)Bi(x)\mathrm{for}x<0.$$
$`(52)`$
By matching boundary conditions it is easy to show that the solution corresponding $`x>0`$ and that corresponding $`x<0`$ are continuous when $`C_1=D_1`$ and $`C_2=D_2`$.
As an example of solutions from this region, we choose: $`a=0.95`$, $`M=1`$, $`m_p=0.17`$, $`l=\frac{1}{2}`$, $`m=\frac{1}{2}`$, and $`\sigma =0.21`$. The black hole horizon is at $`r_+=M+\sqrt{(M^2a^2)}1.31225`$, $`\sigma _c=0.526316`$, $`\sigma _s=0.180987`$, $`\alpha ^2=1.356`$ and $`\text{}\lambda =0.93`$ . It is clear that the values of $`\sigma _c`$, $`\sigma _s`$ and $`\alpha ^2`$ indicate the region is non super-radiant. In Fig. 5a, we show the nature of $`V_+`$ and $`V_{}`$, however, while solving, we use the equation containing $`V_+`$ (eq. 24). Unlike the case in the previous sub-section, here $`\sigma ^2`$ is no longer greater than $`V_\pm `$ at all radii. As a result, $`k^2`$ may attain negative values in some region. In Fig. 5b, nature of $`V_+`$ (solid curve), parameter $`k`$ (dashed curve) and energy $`E`$ (short-dashed curve) are shown. Here, WKB approximation can be applied in regions other than $`\widehat{r}_{}4`$ to $`1`$ and $`1`$ to $`7`$ where $`k`$ is close to zero and the condition $`\frac{1}{k}\frac{dk}{d\widehat{r}_{}}<<k`$ is not satisfied.
In the region $`\widehat{r}_{}=7`$ to $`1`$ around the turning point $`\widehat{r}_{}=4.45475`$ the solutions turns out to be ,
$$Z_+=1.087526Ai(x)+0.788968Bi(x).$$
$`(53)`$
Similarly, the solution from $`\widehat{r}_{}1`$ to $`4`$ i.e. around the turning point $`\widehat{r}_{}=2.8053`$ turns out to be ,
$$Z_+=[1.328096Ai(x)+0.774426Bi(x)].$$
$`(54)`$
It is to be noted that in the region $`\widehat{r}_{}1`$ to $`1`$, even though the potential energy dominates over the total energy, WKB approximation method is still valid. Here the solution will take the form $`\frac{exp(u)}{\sqrt{k}}`$ and $`\frac{exp(+u)}{\sqrt{k}}`$. Asymptotic values of the instantaneous reflection and the transmission coefficients (which are traditionally known as the ‘reflection’ and ‘transmission’ coefficients) are obtained from the WKB approximation. This yields the integral constant $`c`$ as in previous case. ¿From eq. 37(a-b) local reflection and transmission coefficients are calculated, behaviours of which are shown in Fig. 5c. The constants $`A_h`$ and $`A_{+h}`$ are calculated as before. Note the decaying nature of the reflection coefficient inside the potential barrier.
### Solutions in the Super-Radiant Region
In this region, the potential diverges at $`r=|\alpha |`$. Here, the barrier height goes up to infinity and then the potential changes sign so that its nature changes from repulsive to attractive and vice versa. This is because $`\sigma <\sigma _s`$ (which is the condition for super-radiance) and $`r_+<|\alpha |`$ \[see equations (11) and (13)\]. Unlike the previous two cases, the relation between $`r`$ and $`\widehat{r}_{}`$ is not single valued. Here, at both $`r=r_+`$ and $`r=\mathrm{}`$, the value of $`\widehat{r}_{}=\mathrm{}`$. With the decrement of $`r`$, $`\widehat{r}_{}`$ is decreased initially up to $`r=|\alpha |`$. Subsequently, $`\widehat{r}_{}`$ starts to rise and at the black hole horizon it diverges. Obviously, in this case particles hit the barrier and we can solve the equation following the same methods as explained in the previous cases, i.e., eqs. (35) and (35) for the region where the WKB method is valid and eqs. (51) and (52) where the WKB method is not valid.
For illustrative example, here, we choose: $`a=0.95`$, $`M=1`$, $`m_p=0.105`$, $`l=\frac{1}{2}`$, $`m=\frac{1}{2}`$, and $`\sigma =0.105.`$ The black hole horizon is located at $`r_+=M+\sqrt{(M^2a^2)}1.31225`$, and $`\sigma _c=0.526316`$, $`\sigma _s=0.180987`$, $`\alpha ^2=3.62`$ and $`\text{}\lambda =0.97`$ . Chandrasekhar showed that for integral spin particles this region exhibits super-radiance and conjectured that for half-integral spins the super-radiance may be absent. We investigate here if this conjecture is valid.
The behaviour of potentials $`V_+`$ and $`V_{}`$ are shown in Fig. 6a. It is clear that at $`r=|\alpha |`$ the potential diverges and the nature of the potential is changed from repulsive to attractive (for $`V_{}`$) and vice versa (for $`V_+`$). Here, we will treat the equation with $`V_{}`$ as the potential (it is equally easy to do the problem with $`V_+`$). We first divide our computations into two parts. In the repulsive part of the potential (i.e., when $`V_{}>0`$), particles come from infinity and most of them reflect back from the infinitely high barrier. In the attractive part of the potential (i.e., when $`V_+<0`$), particle radiates outwards in the $`\widehat{r}_{}`$ co-ordinate (actually, particle goes towards the horizon but due to multivalueness of the radial co-ordinate $`\widehat{r}_{}`$ (with respect to $`r`$) the horizon is mapped to infinity).
In Fig. 6b, nature of $`V_{}`$, $`k`$ and $`E`$ are shown. The WKB approximation (more precisely IWKB approximation) method is valid from infinity to $`\widehat{r}_{}40`$ since, otherwise, $`\frac{1}{k}\frac{dk}{d\widehat{r}_{}}<<k`$ is not satisfied. In those other regions one has to apply a different method (which was also explained in last sub-section) to find solutions. The local values of the reflection and transmission coefficients and the wave function of the particle are calculated as in the previous cases. Since the matter which tunnels through the infinitely high barrier face infinitely strong attractive field, the possibility of extraction of energy would be zero. In Fig. 6c, the variations of local transmission and reflection coefficients are shown. The net transmission of the wave through the horizon is non-negative all along and therefore super-radiation is absent, although $`\sigma `$ is less than $`\sigma _s`$. We believe that the non-existence of super-radiation is due to ($`r|\alpha |)^3`$ variation of the potential near the singular point. Because of the existence of attractive field, the extraction of energy is difficult, so the net transmission of the wave through the horizon from $`\mathrm{}`$ is always positive. This argument is valid for any set of parameters where $`\sigma \sigma _s`$.
## IV. CONCLUSION
In this paper, we studied scattering of massive, spin-half particles from a Kerr black hole, particularly the nature of the radial wave functions and the reflection and transmission coefficients. Our main motivation was to give a general analytical expression of the solution which can be useful for further study. We showed few illustrative cases as examples. We verified that these analytical solutions were indeed correct by explicitly solving the same set of equations using quantum mechanical step-potential approach as described in Section III. We classified the entire parameter space in terms of the physical and unphysical regions and the physical region was further classified into two regions, depending on whether the particle ‘hits’ the potential barrier or not. Again, the region where particle hits on the barrier, is divided into two parts, one is super-radiant region and other is non-super-radiant region. We chose one illustrative example in each of the regions. We emphasize that the most ‘interesting’ region to study would be close to $`m_p\sigma `$. However we pointed out (Fig. 1) that for $`m_p0.35`$, the WKB solutions cannot be trusted, and other methods (such as those using Airy functions) must be employed.
We used the well known WKB approximation method as well as the step-potential method of quantum mechanics to obtain the spatial dependence of the coefficients of the wave function. This in turn, allowed us to determine the reflection and transmission coefficients and the nature of wave functions. The usual WKB method with constant coefficients and (almost) constant wave number $`k`$ is successfully applied even when the coefficients and wave number are not constant everywhere. Solution from this ‘instantaneous’ WKB (or IWKB) method agrees fully with that obtained from a purely quantum mechanical method where the potential is replaced by a collection of steps. Our method of obtaining solutions should be valid for any black hole geometry which are asymptotically flat so that radial waves could be used at a large distance. This way we ensure that the analytical solution is close to the exact solution. In Region II, in some regions, the WKB method cannot be applied and hence Airy function approach or our step-potential approach could be used.
In the literature, reflection and transmission coefficients are defined at a single point. These definitions are meaningful only if the potential varies in a small region while studies are made from a large distance from it. In the present case, the potential changes over a large distance and we are studying in these regions as well. Although we used the words ‘reflection’ and ‘transmission’ coefficients, in this paper very loosely, our definitions are very rigorous and well defined. These quantities are simply the instantaneous values and in our belief more physical. The problem at hand is very similar to the problem of reflection and transmission of acoustic waves from a strucked string of non-constant density where reflection and transmission occurs at each point.
Among other things, we verify Chandrasekhar’s conjecture based on the asymptotic solution, that for spin-$`\frac{1}{2}`$ particle the phenomenon of super-radiance is absent. We believe that this is due to the very way the potential develops the singularity at $`r=|\alpha |`$. Here $`V_{}(\widehat{r}_{})(r|\alpha |)^3`$, which results in an attractive potential in some region very close to the black hole. In contrast, $`V_{}(\widehat{r}_{})(r|\alpha |)^4`$ when electromagnetic and gravitational waves are scattered off the black hole does not create an attractive part in the potential and possibly exhibit the phenomenon of super-radiance. It is noted that all the cases where potential diverge at $`r=\alpha `$ (i.e., so called super-radiation cases) arise for $`\sigma \sigma _s`$ with the negative values of azimuthal quantum number (here, $`m=1/2`$) and the positive Kerr parameter, $`a`$. For positive values of $`m`$ and positive values of $`a`$, potential does not diverge at any point for all values of $`\sigma `$. If we change the spin orientation of the black hole (negative values of $`a`$) and take positive $`m`$ again divergence of the potential will arise. Thus, it seems that the cases with opposite sign of $`a`$ and $`m`$ are physically more interesting.
It is seen that for different physical parameters the solutions are different. The waves scattered off are distinctly different in different parameter regions. In a way, therefore, black holes can act as a mass spectrograph! Another interesting application of our method would be to study interactions of Hawking radiations in regions just outside the horizon.
## FIGURE CAPTIONS
|
warning/0007/hep-ph0007253.html
|
ar5iv
|
text
|
# R-mediation of Dynamical Supersymmetry Breaking
## I Introduction
Low energy supersymmetry may play an important role in solving many problems of the particle physics. If it is the case, supersymmetry must be spontaneously broken, and all superpartners must have appropriate masses, since their effect is not observed yet. Therefore, finding a simple mechanism of the supersymmetry breaking and its mediation without any phenomenological problems is an important task. If we believe low energy supersymmetry, it is natural to consider the supergravity framework.
The simplest scenario of the supersymmetry breaking and its mediation in supergravity theories is the gravity mediation with Polonyi potential in the hidden sector, but supersymmetry is not dynamically broken in this scenario. Moreover, it is well known that the gravity mediation has a phenomenological problem: the degeneracy of squark masses at Planck scale is distorted by the quantum effect at low energies, which causes supersymmetric flavor problem. There is another conceptual problem about the gravity mediation as pointed out in Ref.: it is not the mediation by the gravity, but the mediation by the higher dimensional contact interactions introduced by hand. Although it is possible that the superspace density, which defines the supergravity Lagrangian, contains infinite number of higher dimensional contact interactions so that the Kähler potential has a simple canonical form, the origin of these interactions is mysterious.
There is another possibility that the visible sector and hidden sector are completely separated, namely, no contact interaction among them in the superspace density. This situation would be naturally realized, if two sectors are confined in the different branes separated in the direction of extra dimensions (now the hidden sector should be called as the sequestered sector). In this case the supersymmetry breaking at the sequestered sector is transmitted to the visible sector only through the superconformal anomaly . In this anomaly mediation the masses of squarks highly degenerate at low energies and there is no supersymmetric flavor problem, but sleptons have negative masses ($`m_{\text{slepton}}^2<0`$). There are many attempts to solve this problem , and we usually need some additional fields which communicate two sectors. In this paper we introduce this additional communication by gauging U(1)<sub>R</sub> symmetry in four dimensional supergravity theories . Since the charge of U(1)<sub>R</sub> symmetry does not commute with supercharges, it is natural to consider that the U(1)<sub>R</sub> gauge boson propagates in whole space-time including extra dimensions, and communicates two sectors.
It is also interesting to note that Fayet-Iliopoulos term for U(1)<sub>R</sub> must exist due to the symmetry of supergravity, and this term can act an important role in the supersymmetry breaking. In fact it has been shown that the supersymmetry can be dynamically broken by the interplay between this Fayet-Iliopoulos term and the non-perturbative effect of a gauge interaction . Since the auxiliary field of U(1)<sub>R</sub> gauge multiplet has vacuum expectation value, both squarks and sleptons can have positive masses of the order of the gravitino mass in an appropriate R-charge assignment, and the problem of the anomaly mediation can be avoided.
This paper is organized as follows. In the next section we give a general argument on the supergravity Lagrangian with U(1)<sub>R</sub> gauge symmetry. We give a general formula for the chirality-conserving scalar mass in the presence of U(1)<sub>R</sub> gauge symmetry, which is an extension of the formula given in Ref.. An explicit model is constructed in section 3, and the analysis of the dynamics and mass spectrum is given in section 4. Section 5 contains our conclusions.
## II Supergravity with U(1)<sub>R</sub> gauge symmetry
In the superconformal framework the general supergravity Lagrangian with U(1)<sub>R</sub> gauge symmetry is given by
$`=`$ $``$ $`{\displaystyle \frac{1}{2}}\left[\overline{\mathrm{\Sigma }}_ce^{2g_RV_R}\mathrm{\Sigma }_c\mathrm{\Phi }(S_I,\overline{S}^Ie^{2Q_Ig_RV_R}e^{2g_GV_G})\right]_D`$ (1)
$`+`$ $`\left[W(S_I)\mathrm{\Sigma }_c^3\right]_F{\displaystyle \frac{1}{4}}\left[f_R(S_I)W_RW_R\right]_F{\displaystyle \frac{1}{4}}\left[f_{ab}(S_I)W_G^aW_G^b\right]_F,`$ (2)
where we use the notation in Ref.. Here, $`S_I`$ are matter chiral multiplets with flavor index $`I`$ and U(1)<sub>R</sub> charge $`Q_I`$, $`V_R`$ and $`V_G`$ ($`W_R`$ and $`W_G`$) are vector (chiral) multiplets corresponding to the gauge group of U(1)<sub>R</sub> and $`G`$, respectively. The multiplet $`\mathrm{\Sigma }_c`$ is the compensating multiplet, whose component should be appropriately fixed to obtain Poincaré supergravity. The functions $`\mathrm{\Phi }`$ and $`W`$ are superspace densities in which interactions are described by the products of multiplets. Following the arguments in the previous section, we assume that there is no interaction between the visible sector fields $`S_i`$ and $`V_{Gv}`$ and the hidden (sequestered) sector fields $`S_\alpha `$ and $`V_{Gh}`$ in these superspace densities, namely,
$$\mathrm{\Phi }(S_I,\overline{S}_Ie^{2Q_Ig_RV_R}e^{2g_GV_G})=\mathrm{\Phi }_v(S_i,\overline{S}^ie^{2Q_ig_RV_R}e^{2g_{Gv}V_{Gv}})+\mathrm{\Phi }_h(S_\alpha ,\overline{S}^\alpha e^{2Q_\alpha g_RV_R}e^{2g_{Gh}V_{Gh}}),$$
(3)
$$W(S_I)=W_v(S_i)+W_h(S_\alpha ),$$
(4)
where indices $`i`$ and $`\alpha `$ denote the flavors in the visible and hidden sectors, respectively, and $`Gv`$ and $`Gh`$ are gauge groups in each sector. The gauge kinetic function $`f_{ab}(S_I)`$ should also be restricted as follows.
$$\left[f_{ab}(S_I)W_G^aW_G^b\right]_F\left[f_{ab}^{Gv}(S_i)W_{Gv}^aW_{Gv}^b\right]_F+\left[f_{ab}^{Gh}(S_\alpha )W_{Gh}^aW_{Gh}^b\right]_F.$$
(5)
In the following we assume $`f_R(S_I)=1`$ and $`f_{ab}^{Gv}=f_{ab}^{Gh}=\delta _{ab}`$, for simplicity.
Note that the compensating multiplet $`\mathrm{\Sigma }_c`$ must have R-charge, since the superpotential $`W`$ has R-charge. Therefore, the usual gauge choice to give Poincaré supergravity:
$`z_c=\sqrt{3},\chi _{Rc}=0,b_\mu =0`$ (6)
does not preserve U(1)<sub>R</sub> symmetry, where $`z_c`$ and $`\chi _{Rc}`$ are scalar and spinor components of the compensating multiplet $`\mathrm{\Sigma }_c`$ and $`b_\mu `$ is one of the gauge fields of the superconformal gauge group. We have to rescale the compensating multiplet to obtain the R-symmetric Poincaré supergravity:
$$S_0\mathrm{\Sigma }_c\left(W(S_I)\right)^{1/3}.$$
(7)
The Lagrangian becomes
$`=`$ $``$ $`{\displaystyle \frac{1}{2}}\left[\overline{S}_0S_0\stackrel{~}{\mathrm{\Phi }}(S_I,\overline{S}^Ie^{2Q_Ig_RV_R}e^{2g_GV_G})\right]_D+\left[S_0^3\right]_F`$ (8)
$``$ $`{\displaystyle \frac{1}{4}}\left[W_RW_R\right]_F{\displaystyle \frac{1}{4}}\left[W_{Gv}W_{Gv}\right]_F{\displaystyle \frac{1}{4}}\left[W_{Gh}W_{Gh}\right]_F,`$ (9)
where
$$\stackrel{~}{\mathrm{\Phi }}(S_I,\overline{S}^Ie^{2Q_Ig_RV_R}e^{2g_GV_G})\frac{\mathrm{\Phi }(S_I,\overline{S}^Ie^{2Q_Ig_RV_R}e^{2g_GV_G})}{\left(\overline{W}(\overline{S}^Ie^{2Q_Ig_RV_R}e^{2g_GV_G})W(S_I)\right)^{1/3}}.$$
(10)
The compensating multiplet $`S_0`$ is U(1)<sub>R</sub> singlet now. It was shown in Ref. that the gauge fixing conditions of
$$z_0=\sqrt{3}\stackrel{~}{\mathrm{\Phi }}^{1/2}(z_I,z_I^{}),\chi _{R0}=z_0\stackrel{~}{\mathrm{\Phi }}^1\stackrel{~}{\mathrm{\Phi }}^J\chi _{RJ},b_\mu =0$$
(11)
directly give the standard form of the supergravity Lagrangian given in Ref., where $`z_0`$ and $`\chi _{R0}`$ are scalar and spinor components of the compensating multiplet $`S_0`$, $`\stackrel{~}{\mathrm{\Phi }}^J\stackrel{~}{\mathrm{\Phi }}(z_I,z^I)/z_J`$ and $`z_J`$ is the scalar components of $`S_J`$. After all, the resultant Lagrangian in component fields has the standard form of Ref. including covariant derivatives for U(1)<sub>R</sub> gauge symmetry. The Lagrangian is determined by a function
$$G(z_I,z^I)3\mathrm{ln}\stackrel{~}{\mathrm{\Phi }}(z_I,z^I)=3\mathrm{ln}\mathrm{\Phi }(z_I,z^I)+\mathrm{ln}|W(z_I)|^2,$$
(12)
where $`\mathrm{\Phi }`$ and $`W`$ satisfy the conditions of Eqs.(3) and (4). The difference of R-charges in covariant derivatives for each component field in a multiplet automatically appears due to the fact that $`W`$ has non-trivial R-charge (see Ref.).
The potential for scalar fields is given as follows.
$$V=V_F+V_D,$$
(13)
where F-term contribution is
$$V_F=e^G\left(G_I(G^1)_J^IG^J3\right)$$
(14)
and U(1)<sub>R</sub> D-term contribution is
$$V_D=\frac{g_R^2}{2}\left(G^IQ_Iz_I\right)^2.$$
(15)
We take the reduced Planck scale as a unit of the mass scale. The chirality-conserving scalar mass can be obtained by differentiating this potential by $`z_i`$ and $`z^j`$ and taking its vacuum expectation value. In addition to the conditions of Eqs.(3) and (4), we introduce the conditions of
$$\mathrm{\Phi }_{vi},W_{vi},z_i=0\text{or}1.$$
(16)
These conditions mean the assumption that the breaking scales of gauge symmetries in the visible sector should be much smaller than the reduced Planck scale. We obtain,
$$V_F{}_{j}{}^{i}=m^{ik}\left(G^1\right)_k^lm_{lj}+\frac{2}{3}V_FG_j^i,$$
(17)
$$V_D{}_{j}{}^{i}=\frac{2}{3}V_DG_j^i+g_R^2(Q_i\frac{2}{3})DG_j^i,$$
(18)
where $`m^{ik}`$ is the supersymmetric mass and $`DG^IQ_Iz_I`$. The superpotential $`W`$ has R-charge 2 in our convention. Therefore, the chirality-conserving supersymmetry-breaking scalar mass is obtained as
$$\stackrel{~}{m}_R^2{}_{j}{}^{i}=(\frac{2}{3}V+g_R^2(Q_i\frac{2}{3})D)G_j^i.$$
(19)
The vacuum expectation value of the potential itself corresponds to the cosmological constant which should vanish in realistic models. We see that there is no gravity mediation, but there is “R-mediation” which is the tree-level contribution due to $`D0`$.
## III Constructing a model
We construct an explicit model to show that the scenario which is described in the first section is possible. The particle contents of the model are summarized in Table I. In the following we simply introduce the role of each field without mentioning the dynamics in detail. The dynamics will be discussed in the next section.
As for the hidden sector, we take the same system which was introduced in Ref.. It consists of two fields, $`Q_1`$ and $`Q_2`$, in the fundamental representation of SU(2)<sub>H</sub> gauge group and a Yukawa interaction with a SU(2)<sub>H</sub> singlet field $`S`$:
$$W_h=\lambda S[Q_1Q_2],$$
(20)
where square brackets denote the contraction of SU(2)<sub>H</sub> indices. Supersymmetry is dynamically broken by the interplay between the non-perturbative effect of SU(2)<sub>H</sub> interaction and U(1)<sub>R</sub> Fayet-Iliopoulos term, if there is no other field with R-charge less than $`2/3`$.
The visible sector is based on the system of the minimal supersymmetric standard model. At least the R-charges of leptons must be larger than $`2/3`$ so that sleptons obtain positive masses through R-mediation of Eq.(19) (assuming $`D>0`$ in this section). We simply assume that all the quarks and leptons have unit R-charge. This means that the R-charges of Higgs fields must be zero (note that this is less than $`2/3`$), since we need the Yukawa couplings of
$$W_v^Y=g_uHQ\overline{U}+g_d\overline{H}Q\overline{D}+g_\nu HL\overline{N}+g_e\overline{H}L\overline{E},$$
(21)
where we suppress the generation indices, for simplicity.
To ensure the dynamical supersymmetry breaking we introduce other two Higgs fields, $`H^{}`$ and $`\overline{H}^{}`$, with the mass terms of
$$W_v^{}=\mu _uH\overline{H}^{}+\mu _d\overline{H}H^{}.$$
(22)
Although there are negative contributions to the masses of $`H`$ and $`\overline{H}`$ from Eq.(19), these mass terms can make all masses of Higgs fields positive at the tree level. Therefore, the electroweak symmetry must be broken radiatively.
At this stage, all the gauge anomalies are cancelled out, except for $`\left(\text{SU(3)}_c\right)^2\text{U(1)}_R`$, $`\left(\text{SU(2)}_L\right)^2\text{U(1)}_R`$, $`(\text{U(1)}_R)^3`$ and $`\text{U(1)}_R(\text{gravity})^2`$ anomalies. To cancel $`\left(\text{SU(3)}_c\right)^2\text{U(1)}_R`$ and $`\left(\text{SU(2)}_L\right)^2\text{U(1)}_R`$ anomalies, we further introduce additional fields, $`\mathrm{\Omega }_i`$ and $`\mathrm{\Sigma }_i`$, and Yukawa interactions with $`X`$.
$$W_v^X=gX\left(\mathrm{\Omega }_i\mathrm{\Omega }_i+\mathrm{\Sigma }_i\mathrm{\Sigma }_i\right)+mXX^{}.$$
(23)
The field $`X^{}`$ and the mass term with $`X`$ are required to have positive mass for $`X`$ and to ensure the dynamical supersymmetry breaking, since $`X`$ has R-charge less than $`2/3`$. The fields $`\mathrm{\Omega }_i`$ and $`\mathrm{\Sigma }_i`$ become heavy by the vacuum expectation value of $`X`$ which is generated by the 1-loop effect of the Yukawa coupling in Eq.(23). The remaining anomalies, $`(\text{U(1)}_R)^3`$ and $`\text{U(1)}_R(\text{gravity})^2`$, can be cancelled out by introducing, for example, many fields of R-charge two with appropriate values of $`q_1`$ and $`q_2`$. There may be much more sophisticated and convincing way to cancel these anomalies, but we leave it for further studies.
## IV Dynamics of the model
Before discussing the dynamics of the model in detail, we have to make an assumption about the superspace density, $`\mathrm{\Phi }`$. We simply assume as
$$\mathrm{\Phi }(z_I,z^I)=1\frac{1}{3}\underset{I}{}z^Iz_I$$
(24)
respecting the condition of Eq.(3), where $`z_I`$ are the scalar components of all the chiral multiplets in the model. This gives canonical kinetic terms in the first order of the $`1/M_P`$ expansion, where $`M_P=M_{\mathrm{planck}}/\sqrt{8\pi }`$ is the reduced Planck scale. In this case the scalar potential can be written as
$$V=V_F+V_D$$
(25)
with
$$V_F=\frac{1}{\mathrm{\Phi }^2}\left\{W_I^{}W^I\frac{1}{3}|z_IW^I|^2+\left(W^{}W^Iz_I+WW_I^{}z^I\right)3|W|^2\right\},$$
(26)
$$V_D=\frac{1}{\mathrm{\Phi }^2}\frac{g_R^2}{2}\left\{\left(Q_I\frac{2}{3}\right)z^Iz_I+2\right\}^2,$$
(27)
where we neglect the D-term contributions form other gauge interactions.
First, we discuss the dynamics of the supersymmetry breaking. The instanton effect of SU(2)<sub>H</sub> gauge interaction can be described as a dynamically generated superpotential . The effective superpotential for the hidden sector is
$$W_h^{\mathrm{eff}}=\lambda S[Q_1Q_2]+\frac{\mathrm{\Lambda }^5}{[Q_1Q_2]},$$
(28)
where $`\mathrm{\Lambda }`$ is the scale of the dynamics of SU(2)<sub>H</sub>. If we assume that the vacuum expectation values of $`Q_1`$ and $`Q_2`$ lie on the flat direction of SU(2)<sub>H</sub>, we have
$`V_F`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Phi }^2}}\left\{\left(\lambda v^2\right)^2+2v^2\left(\lambda s{\displaystyle \frac{\mathrm{\Lambda }^5}{v^4}}\right)^2{\displaystyle \frac{25}{3}}\left({\displaystyle \frac{\mathrm{\Lambda }^5}{v^2}}\right)^2+\text{(visible sector)}\right\},`$ (29)
$`V_D`$ $`=`$ $`{\displaystyle \frac{g_R^2}{2}}D^2,`$ (30)
where
$$\mathrm{\Phi }=1\frac{1}{3}s^2\frac{2}{3}v^2\frac{1}{3}\left[z^Iz_I\right]_{\mathrm{visible}},$$
(31)
$$D=\frac{1}{\mathrm{\Phi }}\left\{\left(q_S\frac{2}{3}\right)s^2+\left(q_1+q_2\frac{4}{3}\right)v^2+2+\left[\left(Q_I\frac{2}{3}\right)z^Iz_I\right]_{\mathrm{visible}}\right\},$$
(32)
$`v`$ describes the flat direction of SU(2)<sub>H</sub>, $`s`$ and $`q_S`$ are the vacuum expectation value and R-charge of $`S`$ ($`q_S=4`$ and $`q_1+q_2=2`$). It can be shown that all the visible sector fields do not have vacuum expectation values at the tree level. It is rather trivial for the fields with R-charge larger than $`2/3`$, but it is non-trivial for the fields $`H`$, $`\overline{H}`$ and $`X`$, since the vacuum expectation values of these fields negatively contribute to the vacuum energy in $`V_D`$. These fields do not have vacuum expectation values at the tree level, if the mass parameters $`\mu _u`$, $`\mu _d`$ and $`m`$ are appropriately large, as it will be explained at the end of this section. Therefore, in the following we consider the stationary conditions for $`v`$ and $`s`$ neglecting all the contributions from the visible sector.
The analysis is almost the same as in Ref.. In case of $`g_R^2\lambda \mathrm{\Lambda }^5`$ there should be a solution of stationary conditions so that $`v\sqrt{3/5}`$ and $`s1`$, which results almost vanishing $`D`$. In this case the scalar potential approximately becomes
$$V\frac{1}{\mathrm{\Phi }^2}\left(\frac{3}{5}\right)^2\left\{\lambda ^23\left(\frac{5}{3}\right)^5\mathrm{\Lambda }^{10}\right\}.$$
(33)
Therefore, we can expect that there is a solution of vanishing (or vary small) cosmological constant with $`\lambda \sqrt{5}(5/3)^2\mathrm{\Lambda }^56.2\mathrm{\Lambda }^5`$. Indeed, we can approximately obtain such a solution as
$`v`$ $``$ $`\sqrt{{\displaystyle \frac{3}{5}}}+{\displaystyle \frac{\sqrt{15}}{6}}s^2{\displaystyle \frac{1}{g_R^2}}{\displaystyle \frac{243\lambda ^2+6250\mathrm{\Lambda }^{10}}{900\sqrt{15}}},`$ (34)
$`s`$ $``$ $`{\displaystyle \frac{675\lambda \mathrm{\Lambda }^5}{486\lambda ^2+6250\mathrm{\Lambda }^{10}}}`$ (35)
with vanishing cosmological constant by tuning $`\lambda 6.9\mathrm{\Lambda }^5`$. A complete numerical analysis gives a solution
$$v\sqrt{3/5}+0.012,s0.14$$
(36)
with vanishing cosmological constant, where $`g_R=10^{12}`$, $`\mathrm{\Lambda }=10^3`$ and $`\lambda 6.9\mathrm{\Lambda }^5`$. At this vacuum the gravitino mass $`m_{3/2}`$ becomes
$$m_{3/2}e^{G/2}5.0\times \frac{\mathrm{\Lambda }^5}{M_P^4}.$$
(37)
The contribution to the mass of the scalar field due to $`D0`$ can also be obtained from Eq.(19) as
$$\stackrel{~}{m}_R^2(Q)=g_R^2D\left(Q\frac{2}{3}\right)\left(7.2\times \frac{\mathrm{\Lambda }^5}{M_P^4}\right)^2\left(Q\frac{2}{3}\right),$$
(38)
where $`Q`$ is the R-charge of the scalar field. We see that these supersymmetry-breaking masses are the same order of magnitude. The phenomenologically acceptable values of these masses can be obtained by changing the value of $`\mathrm{\Lambda }`$ within the same order of magnitude.
We summarize the spectrum of the supersymmetry-breaking masses and other supersymmetry breaking terms in the visible sector.
Gauginos in the visible sector can have masses only through the anomaly mediation, since there should be no hidden (sequestered) sector field in the gauge kinetic function. Therefore,
$$m_{\lambda _i}\frac{\beta (g_i^2)}{2g_i^2}m_{3/2},$$
(39)
where $`g_i`$ and $`\beta (g_i^2)`$ are the gauge coupling and its beta function of the gauge group $`i`$ in the visible sector, respectively (see Ref. for more precise formula). There are two contributions to the scalar mass:
$$\stackrel{~}{m}^2\frac{1}{4}\frac{d\gamma }{d\mathrm{ln}\mu }m_{3/2}^2+\stackrel{~}{m}_R^2(Q),$$
(40)
where $`\mu `$ is the renormalization scale, and $`\gamma `$ and $`Q`$ are the anomalous dimension and R-charge of the scalar field, respectively. The first term is the contribution by the anomaly mediation and the second term is the contribution by R-mediation given by Eq.(38). The second contribution always dominates the first contribution, since the second contribution is the tree-level one. Therefore, the scalar filed with $`Q>2/3`$ naturally has positive mass. If we take $`m_{3/2}10`$TeV to have gaugino masses heavier than about $`100`$GeV, the scalar mass becomes of the order of $`(10\text{TeV})^2`$.
Other supersymmetry breaking terms also emerge through the anomaly mediation. The A-term emerges corresponding to each Yukawa coupling through the anomaly mediation:
$$A_{\mathrm{\Phi }_1\mathrm{\Phi }_2\mathrm{\Phi }_3}=\frac{1}{2}\left(\gamma _{\mathrm{\Phi }_1}+\gamma _{\mathrm{\Phi }_2}+\gamma _{\mathrm{\Phi }_3}\right)m_{3/2},$$
(41)
where the Yukawa coupling of $`W_{\mathrm{Yukawa}}=\lambda \mathrm{\Phi }_1\mathrm{\Phi }_2\mathrm{\Phi }_3`$ is considered, and $`\gamma `$ denotes the anomalous dimension of each field. The B-term emerges corresponding to each mass term at the tree level, since the mass term explicitly breaks the superconformal symmetry:
$$Bm_{3/2}.$$
(42)
Note that the order of the magnitude of $`A`$ is always smaller than that of $`B`$, since the B-term emerges at the tree level.
Next, we discuss the radiative electroweak symmetry breaking in our model. When we neglect the hidden sector, Higgs fields do not have vacuum expectation values at the tree level, if the following conditions are satisfied.
$`\left(\mu _u^2+\stackrel{~}{m}_R^2(q_H)\right)\left(\mu _u^2+\stackrel{~}{m}_R^2(q_H^{})\right)\mu _u^2B^2>0,`$ (43)
$`\left(\mu _d^2+\stackrel{~}{m}_R^2(q_H)\right)\left(\mu _d^2+\stackrel{~}{m}_R^2(q_H^{})\right)\mu _d^2B^2>0,`$ (44)
where $`q_H`$ is the R-charge of $`H`$ and $`\overline{H}`$ and $`q_H^{}`$ is the R-charge of $`H^{}`$ and $`\overline{H}^{}`$. This condition is satisfied, if $`\mu _u`$ and $`\mu _d`$ are slightly larger than $`m_{3/2}`$. On the other hand, $`\mu _u^2`$ and $`\mu _d^2`$ must be larger than $`|\stackrel{~}{m}_R^2(q_H)|\frac{2}{3}m_{3/2}^2`$ so that our mechanism of the dynamical supersymmetry breaking works. Therefore, it is natural to consider that the masses $`\mu _u`$ and $`\mu _d`$ are slightly larger than $`m_{3/2}`$ and the electroweak symmetry breaking occurs radiatively at the 1-loop level through the large Yukawa coupling of the top quark and the supersymmetry-breaking mass of the scalar top. We can analytically show that the radiative electroweak symmetry breaking at the weak scale is possible in this model.
Finally, we discuss the radiative mass generation for $`\mathrm{\Omega }_i`$ and $`\mathrm{\Sigma }_i`$ which are introduced for the gauge anomaly cancellation. These fields should become heavy so that running gauge couplings do not blow up before Planck scale. The point is whether $`X`$ has vacuum expectation value or not, since $`X0`$ make these fields massive through the first term of Eq.(23). When we neglect the hidden sector, the vacuum expectation value of $`X`$ vanishes at the tree level, if the following condition is satisfied.
$$\left(m^2+\stackrel{~}{m}_R^2(q_X)\right)\left(m^2+\stackrel{~}{m}_R^2(q_X^{})\right)m^2B^2>0,$$
(45)
where $`q_X`$ and $`q_X^{}`$ are the R-charges of $`X`$ and $`X^{}`$, respectively. This condition is satisfied, if $`m`$ is larger than $`m_{3/2}`$. On the other hand, $`m^2`$ must be larger than $`|\stackrel{~}{m}_R^2(q_X)|\frac{1}{6}m_{3/2}^2`$ so that our mechanism of the dynamical supersymmetry breaking works. Therefore, if we naturally take $`m^2>m_{3/2}^2`$, $`X`$ can not have vacuum expectation value at the tree level. But, it can have vacuum expectation value radiatively at the 1-loop level by the large coupling $`g`$ and the supersymmetry-breaking masses of the scalar components of $`\mathrm{\Omega }_i`$ and $`\mathrm{\Sigma }_i`$. We can analytically show that the value of $`gX`$ can be large enough ($`gX10^2M_P`$, for example) with $`mm_{3/2}`$, $`g1`$ and very small $`g_R10^{12}`$.
## V Conclusion
We have proposed a sequestered sector scenario in which the supersymmetry breaking are mediated by the superconformal anomaly and U(1)<sub>R</sub> gauge interaction without gravity mediation at the tree level. We constructed an explicit model in which supersymmetry is dynamically broken by the interplay between the non-perturbative effect of the gauge interaction and Fayet-Iliopoulos term of U(1)<sub>R</sub>. It was found that the problem of the tachyonic slepton in the anomaly mediation can be avoided in this scenario. The spectrum of the supersymmetry breaking masses is very simple and there is no supersymmetric flavor problem. We have also proposed a mechanism to radiatively generate the mass of the field which should not appear at low energies.
We mention a remarkable fact in this model: R-parity is not necessary. In the minimal supersymmetric standard model R-parity is usually assumed to forbid interactions which violate baryon number symmetry. In our model U(1)<sub>R</sub> gauge symmetry naturally act the same or rather stronger role. It forbids in the superpotential not only renormalizable terms but also all the higher dimensional terms which violate baryon number symmetry. This is a simple realization of the idea proposed in Ref..
We briefly summarize phenomenological consequences of this model. All the gauginos have the masses of the order of $`100`$GeV, and the lightest superparticle would be a neutralino (bino or zino). Further understanding of the gaugino spectrum and the nature of the lightest superparticle requires more detailed analysis on the radiative correction to the spectrum as described in Ref.. All the scalar fermions have the masses of the order of $`10`$TeV. Therefore, in near future collider experiments we could not discover scalar fermions, but gauginos. The Higgs sector in this model is very different from the one in the minimal supersymmetric standard model, since it includes four Higgs doublets. There would be three charged Higgs and seven neutral Higgs, and all of them would have the masses of the order of $`10`$TeV, except for one CP-even neutral Higgs which would have the mass of the order of the weak scale. Therefore, we could see one Higgs boson in near future collider experiments, but it would be impossible to see other Higgs particles.
There is an important point which have to be investigated in future. That is to derive the four dimensional effective theory from the fundamental theory in higher dimension. For example, if we consider a five dimensional theory as a fundamental theory, we have to integrate out the degrees of freedom which can propagate in fifth dimension. In our scenario such degrees of freedom are gravity and R gauge interaction. Especially, it have to be investigated how an U(1) component of the larger R gauge symmetry in five dimensions is projected out to U(1)<sub>R</sub> gauge symmetry in four dimensional effective theories. It is also to be investigated how completely two sectors are separated in the superspace densities in four dimensional effective theories.
Although we still do not have the comprehensive analysis about deriving four dimensional effective theories from higher dimensional gauged supergravity theories, it is possible to expect that the very small value of the U(1)<sub>R</sub> gauge coupling in this model could be naturally obtained in the large extra dimension scenario. The result of Ref. suggests that the natural (dimensionful) value of the gauge coupling in the higher dimensional theory could naturally result very small (dimensionless) value of the gauge coupling in the four dimensional effective theory, if the extra dimensions have relatively large volume. Other gauge and Yukawa couplings in this model are not suppressed by this mechanism, since the visible and sequestered sector are assumed to be confined in each 3-brane and only the supergravity multiplet and U(1)<sub>R</sub> gauge boson can propagate the bulk. If the volume of the extra dimensions is very large, the U(1)<sub>R</sub> gauge boson and graviton would be observed in near future collider experiments.
Finally, we want to emphasize that the proposed scenario is very simple, and we can rather easily construct calculable models which have concrete predictions. We believe that this direction is worth investigating further.
###### Acknowledgements.
This work was supported in part by the Grant-in-aid for Science and Culture Research from the Ministry of Education, Science and Culture of Japan (#11740156, #08557, #2997). N.M. and N.O. are supported by the Japan Society for the Promotion of Science for Young Scientists.
|
warning/0007/cond-mat0007020.html
|
ar5iv
|
text
|
# Exact Ground State of the 2𝐷 Hubbard Model at Half Filling for 𝑈=0⁺
## Abstract
We solve analytically the $`N\times N`$ square lattice Hubbard model for even $`N`$ at half filling and weak coupling by a new approach. The exact ground state wave function provides an intriguing and appealing picture of the antiferromagnetic order. Like at strong coupling, the ground state has total momentum $`K_{tot}=(0,0)`$ and transforms as an $`s`$ wave for even $`N/2`$ and as a $`d`$ wave otherwise.
The $`2D`$ Hubbard model is one of the simplest descriptions of itinerant interacting electrons on a lattice. The Lieb theorem states that at half filling the ground state for a bipartite lattice is unique and has spin $`\frac{||𝒮_1||𝒮_2||}{2}`$ where $`|𝒮_1|`$ ($`|𝒮_2|`$) is the number of sites in the $`𝒮_1`$ ($`𝒮_2`$) sublattice; here and in the following, $`|𝒮|`$ will be the number of elements in the set $`𝒮`$. For a $`N\times N`$ (with even $`N`$) square lattice, the ground state is a singlet. In Ref. it was shown that in the strong coupling limit the ground state has total momentum $`K_{tot}=(0,0)`$ and has $`s`$-wave or $`d`$-wave symmetry for even or odd $`N/2`$, respectively.
In this paper we build the exact ground state of the Hubbard model at half filling in the weak coupling limit. We find the same quantum numbers as predicted by Refs., and this result strongly supports the conjecture that no phase transition takes place for finite values of the Coulomb interaction parameter $`U`$.
Let us consider the Hubbard model with hamiltonian
$$H=H_0+W=t\underset{\sigma }{}\underset{r,r^{}}{}c_{r\sigma }^{}c_{r^{}\sigma }+\underset{r}{}U\widehat{n}_r\widehat{n}_r,U>0$$
(1)
on a square lattice of $`N\times N`$ sites with periodic boundary conditions and even $`N`$. Here $`\sigma =,`$ is the spin and $`r,r^{}`$ the spatial degrees of freedom of the creation and annihilation operators $`c^{}`$ and $`c`$ respectively. We represent sites by $`r=(i_x,i_y)`$ with $`i_x,i_y=0,\mathrm{},N1`$. The sum on $`r,r^{}`$ is over the pairs of nearest neighbor sites and $`\widehat{n}_{r\sigma }`$ is the number operator on the site $`r`$ of spin $`\sigma `$. The point symmetry is $`C_{4v}`$, the Group of a square. We Fourier expand the fermion operators: $`c_{k\sigma }=\frac{1}{N}_re^{ikr}c_{r\sigma }`$, where $`k=(k_x,k_y)=\frac{2\pi }{N}(i_x,i_y)`$; then, $`H_0=_kϵ(k)c_{k\sigma }^{}c_{k\sigma }`$ where $`ϵ(k)=2t[\mathrm{cos}k_x+\mathrm{cos}k_y]`$. The starting point is the following property of the number operator $`\widehat{n}_r=c_r^{}c_r`$ (for the moment we omit the spin index).
Theorem: Let $`𝒮`$ be an arbitrary set of plane-wave eigenstates $`\{|k_i\}`$ of $`H_0`$ and $`(n_r)_{ij}=k_i|\widehat{n}_r|k_j=\frac{1}{N^2}e^{i(k_ik_j)r}`$ the matrix of $`\widehat{n}_r`$ in $`𝒮`$. This matrix has eigenvalues $`\lambda _1=\frac{|𝒮|}{N^2}`$ and $`\lambda _2=\mathrm{}=\lambda _{|𝒮|}=0`$ .
Note that $`|𝒮|N^2`$; if $`|𝒮|=N^2`$ the set is complete, like the set of all orbitals, and the theorem is trivial (a particle on site $`r`$ is the $`\widehat{n}_r`$ eigenstate with eigenvalue 1). Otherwise, if $`|𝒮|<N^2`$, the theorem is an immediate consequence of the fact that
$$det|(n_r)_{ij}\lambda \delta _{ij}|=(\lambda )^{|𝒮|1}(\frac{|𝒮|}{N^2}\lambda ),r.$$
(2)
Let $`𝒮_{hf}`$ denote the set (or shell) of the $`k`$ wave vectors such that $`ϵ(k)=0`$. These $`k`$ vectors lie on the square having vertices $`(\pm \pi ,0)`$ and $`(0,\pm \pi )`$; one can show that the number of solutions, that is, the dimension of the set $`𝒮_{hf}`$, is $`|𝒮_{hf}|=2N2`$. At half filling ($`N^2`$ holes) for $`U=0`$, $`𝒮_{hf}`$ is half occupied. In the $`S_z=0`$ sector there are $`\left(\begin{array}{c}2N2\\ N1\end{array}\right)^2`$ degenerate unperturbed ground state configurations for $`N^2`$ holes, where all $`|k`$ orbitals such that $`ϵ(k)<0`$ are occupied. The first-order splitting of the degeneracy is obtained by diagonalizing the $`W`$ matrix over the unperturbed basis; like in elementary atomic physics, the filled shells just produce a constant shift of all the eigenvalues and for the rest may be ignored in first-order calculations. In other terms, we consider the truncated Hilbert space $``$ spanned by the states of $`N1`$ holes of each spin in $`𝒮_{hf}`$, and we want the exact ground state(s) of $`W`$ in $``$; by construction $``$ is in the kernel of $`H_0`$, so the ground state of $`W`$ is the ground state of $`H`$ as well. We call $`W=0`$ state any vector in $``$ which also belongs to the kernel of $`W`$. Since the lowest eigenvalue of $`W`$ is zero, it is evident that any $`W=0`$ state is a ground state of $`H`$. We want to calculate the unique ground state of the Hubbard Hamiltonian for $`U=0^+`$ at half filling which is a $`W=0`$ singlet with $`2N2`$ holes in $`𝒮_{hf}`$ (filled shells are understood).
To diagonalize the local operator $`W`$ in closed form we need to set up a local basis set of one-body states. If $`𝒮_{hf}`$ were the complete set of plane-wave states $`|k`$, the new basis would be trivially obtained by a Fourier transformation, but this is not the case. The question is: how can we define for each site $`r`$ the local counterparts of $`k`$ states using only those that belong to a degenerate level? The answer is: build a set $`\{|\phi _\alpha ^{(r)}\}`$ of orbitals such that the number operator $`\widehat{n}_r`$ and the Dirac characters of the point symmetry Group $`C_{4v}`$ are diagonal. Using such a basis set for the half-filled shell the unique properties of the antiferromagnetic ground state become simple and transparent. The eigenvectors $`|\phi _\alpha ^{(0)}`$ of $`n_{r=0}`$ and those $`|\phi _\alpha ^{(r)}`$ of other sites $`r`$ are connected by translation and also by a unitary transformation, or change of basis set. Picking $`r=\widehat{e}_l`$, $`l=x`$ means $`\widehat{e}_l=(1,0)`$ or transfer by one step towards the right and $`l=y`$ means $`\widehat{e}_l=(0,1)`$ or transfer by one step upwards. The unitary transformation reads:
$$|\phi _\alpha ^{(\widehat{e}_l)}=\underset{\beta =1}{\overset{2N2}{}}|\phi _\beta ^{(0)}\phi _\beta ^{(0)}|\phi _\alpha ^{(\widehat{e}_l)}\underset{\beta =1}{\overset{2N2}{}}|\phi _\beta ^{(0)}T_{l_{\beta \alpha }}.$$
(3)
The transfer matrix $`T_l`$ knows all the translational and point symmetry of the system, and will turn out to be very special.
For large $`N`$, to find $`\{|\phi _\alpha ^{(r)}\}`$ it is convenient to separate the $`k`$’s of $`𝒮_{hf}`$ in irreducible representations (irreps) of the space Group $`𝐆`$$`=C_{4v}T`$; here $`T`$ is the Abelian Group of the translations and $``$ means the semidirect product. Choosing an arbitrary $`k𝒮_{hf}`$ with $`k_xk_y0`$, the set of vectors $`R_ik`$, where $`R_iC_{4v}`$, is a (translationally invariant) basis for an irrep of $`𝐆`$; the accidental degeneracy of several irreps is due to the presence of extra symmetry, i.e. $`𝐆`$ is a subgroup of the Optimal Group defined in ,. The high symmetry vectors $`(0,\pi )`$ and $`(\pi ,0)`$ always trasform among themselves and are the basis of the only two-dimensional irrep of $`𝐆`$, which exists for any $`N`$. If $`N/2`$ is even, one also finds the high symmetry wavevectors $`k=(\pm \pi /2,\pm \pi /2)`$ which mix among themselves and yield a four-dimensional irrep. In general, the vectors $`R_ik`$ are all different, so all the other irreps of $`𝐆`$ have dimension 8, the number of operations of the point Group $`C_{4v}`$.
Next, we show how to build our local basis set and derive $`W=0`$ states for each kind of irreps of $`𝐆`$. For illustration, we will consider the case $`N=4`$; then $`𝒮_{hf}`$ contains the bases of two irreps of $`𝐆`$, of dimensions 2 and 4. The one with basis $`k_A=(\pi ,0),k_B=(0,\pi )`$ breaks into $`A_1B_1`$ in $`C_{4v}`$ .
The eigenstates of $`(n_{r=0})_{ij}=k_i|\widehat{n}_{r=0}|k_j`$, with $`i,j=A,B`$ , are $`|\psi _{A_1}^{\prime \prime }=\frac{1}{\sqrt{2}}(|k_A+|k_B)`$ with $`\lambda _1=1/8`$ and $`|\psi _{B_1}^{\prime \prime }=\frac{1}{\sqrt{2}}(|k_A|k_B)`$ with $`\lambda _2=0`$. Since under translation by a lattice step $`T_l`$ along the $`l=x,y`$ direction $`|ke^{ik_l}|k`$, using Equation (3) one finds that $`|\psi _{A_1}^{\prime \prime }(1)^{\theta _l^{\prime \prime }}|\psi _{B_1}^{\prime \prime }`$, with $`\theta _x^{\prime \prime }=1,\theta _y^{\prime \prime }=0`$; so $`|\psi _{A_1}^{\prime \prime }`$ has vanishing amplitude on a sublattice and $`|\psi _{B_1}^{\prime \prime }`$ on the other. The two-body state $`|\psi _{A_1}^{\prime \prime }_\sigma |\psi _{B_1}^{\prime \prime }_\sigma `$ has occupation for spin $`\sigma `$ but not for spin $`\sigma `$ on the site $`r=0`$; under a lattice step translation it flips the spin and picks up a (-1) phase factor: $`|\psi _{A_1}^{\prime \prime }_\sigma |\psi _{B_1}^{\prime \prime }_\sigma |\psi _{B_1}^{\prime \prime }_\sigma |\psi _{A_1}^{\prime \prime }_\sigma `$; therefore it has double occupation nowhere and is a $`W=0`$ state ($`W=0`$ pair ).
The 4-dimensional irrep with basis $`k_1=(\pi /2,\pi /2),k_2=(\pi /2,\pi /2),k_3=(\pi /2,\pi /2)k_4=(\pi /2,\pi /2)`$ breaks into $`A_1B_2E`$ in $`C_{4v}`$; letting $`I=1,2,3,4`$ for the irreps $`A_1,B_2,E_x,E_y`$ respectively, we can write down all the eigenvectors of $`(n_{r=0})_{ij}=k_i|\widehat{n}_{r=0}|k_j`$, with $`i,j=1,\mathrm{},4`$, as $`|\psi _I^{}=_{i=1}^4O_{Ii}^{}|k_i`$, where $`O^{}`$ is the following 4$`\times `$4 orthogonal matrix
$$O^{}=\frac{1}{2}\left[\begin{array}{cccc}\hfill 1& \hfill 1& \hfill 1& \hfill 1\\ \hfill 1& \hfill 1& \hfill 1& \hfill 1\\ \hfill 1& \hfill 1& \hfill 1& \hfill 1\\ \hfill 1& \hfill 1& \hfill 1& \hfill 1\end{array}\right].$$
(4)
The state with non-vanishing eigenvalue is again the $`A_1`$ eigenstate. After a little bit of algebra we have shown that under $`T_l`$ the subspace of $`A_1`$ and $`B_2`$ symmetry is exchanged with the one of $`E_x`$ and $`E_y`$ symmetry. Thus we can build a 4-body eigenstate of $`W`$ with vanishing eigenvalue: $`|\psi _{A_1}^{}\psi _{B_2}^{}_\sigma |\psi _{E_x}^{}\psi _{E_y}^{}_\sigma `$. As before under a lattice step translation this state does not change its spatial distribution but $`\sigma \sigma `$ without any phase factor: $`|\psi _{A_1}^{}\psi _{B_2}^{}_\sigma |\psi _{E_x}^{}\psi _{E_y}^{}_\sigma |\psi _{E_x}^{}\psi _{E_y}^{}_\sigma |\psi _{A_1}^{}\psi _{B_2}^{}_\sigma `$.
Now we use these results to diagonalize $`n_{r=0}`$ on the whole set $`𝒮_{hf}`$ (we could have done that directly by diagonalizing $`6\times 6`$ matrices but we wanted to show the general method). The eigenstate of $`n_{r=0}`$ with nonvanishing eigenvalue always belongs to $`A_1`$. The matrix $`n_r`$ has eigenvalues $`3/8`$ and (5 times) $`0`$, as predicted by eq.(2). For $`r=0`$ the eigenvector of occupation $`3/8`$ is $`|\varphi _1^{(0)}|\varphi _{1,A_1}^{(0)}=\frac{1}{\sqrt{3}}|\psi _{A_1}^{\prime \prime }+\sqrt{\frac{2}{3}}|\psi _{A_1}^{}`$. The other $`A_1`$ eigenstate of $`n_{r=0}`$ has 0 eigenvalue and reads: $`|\varphi _2^{(0)}|\varphi _{2,A_1}^{(0)}=\sqrt{\frac{2}{3}}|\psi _{A_1}^{\prime \prime }\frac{1}{\sqrt{3}}|\psi _{A_1}^{}`$.
The other eigenvectors , whose symmetry differs from $`A_1`$, are $`|\varphi _3^{(0)}|\varphi _{B_2}^{(0)}=|\psi _{B_2}^{}`$, $`|\varphi _4^{(0)}|\varphi _{B_1}^{(0)}=|\psi _{B_1}^{\prime \prime }`$, $`|\varphi _5^{(0)}|\varphi _{E_x}^{(0)}=|\psi _{E_x}^{}`$ and $`|\varphi _6^{(0)}|\varphi _{E_y}^{(0)}=|\psi _{E_y}^{}`$. One finds that the transfer matrices $`T_l`$ of Equation (3) such that $`|\varphi _I^{(\widehat{e}_l)}_J|\varphi _J^{(0)}T_{l_{JI}}`$, are:
$$T_x=\left[\begin{array}{cccccc}\hfill 0& \hfill 0& \hfill 0& \hfill \frac{1}{\sqrt{3}}& \hfill i\sqrt{\frac{2}{3}}& \hfill 0\\ \hfill 0& \hfill 0& \hfill 0& \hfill \sqrt{\frac{2}{3}}& \hfill \frac{i}{\sqrt{3}}& \hfill 0\\ \hfill 0& \hfill 0& \hfill 0& \hfill 0& \hfill 0& \hfill i\\ \hfill \frac{1}{\sqrt{3}}& \hfill \sqrt{\frac{2}{3}}& \hfill 0& \hfill 0& \hfill 0& \hfill 0\\ \hfill i\sqrt{\frac{2}{3}}& \hfill \frac{i}{\sqrt{3}}& \hfill 0& \hfill 0& \hfill 0& \hfill 0\\ \hfill 0& \hfill 0& \hfill i& \hfill 0& \hfill 0& \hfill 0\end{array}\right],T_y=\left[\begin{array}{cccccc}\hfill 0& \hfill 0& \hfill 0& \hfill \frac{1}{\sqrt{3}}& \hfill 0& \hfill i\sqrt{\frac{2}{3}}\\ \hfill 0& \hfill 0& \hfill 0& \hfill \sqrt{\frac{2}{3}}& \hfill 0& \hfill \frac{i}{\sqrt{3}}\\ \hfill 0& \hfill 0& \hfill 0& \hfill 0& \hfill i& \hfill 0\\ \hfill \frac{1}{\sqrt{3}}& \hfill \sqrt{\frac{2}{3}}& \hfill 0& \hfill 0& \hfill 0& \hfill 0\\ \hfill 0& \hfill 0& \hfill i& \hfill 0& \hfill 0& \hfill 0\\ \hfill i\sqrt{\frac{2}{3}}& \hfill \frac{i}{\sqrt{3}}& \hfill 0& \hfill 0& \hfill 0& \hfill 0\end{array}\right].$$
(5)
The reason why this choice of the basis set is clever is now apparent. The local basis at any site $`r`$ splits into the subsets $`𝒮_a=\{|\varphi _{1,A_1}^{(r)},|\varphi _{2,A_1}^{(r)},|\varphi _{B_2}^{(r)}\}`$, and $`𝒮_b=\{|\varphi _{B_1}^{(r)},|\varphi _{E_x}^{(r)},|\varphi _{E_y}^{(r)}\}`$; a shift by a lattice step sends members of $`𝒮_a`$ into linear combinations of the members of $`𝒮_b`$, and conversely.
Consider the 6-body eigenstate of $`H_0`$
$$|\mathrm{\Phi }_{AF}_\sigma =|\varphi _{1,A_1}^{(0)}\varphi _{2,A_1}^{(0)}\varphi _{B_2}^{(0)}_\sigma |\varphi _{B_1}^{(0)}\varphi _{E_x}^{(0)}\varphi _{E_y}^{(0)}_\sigma .$$
(6)
In this state there is partial occupation of site $`r=0`$ with spin $`\sigma `$, but no double occupation. It turns out that a shift by a lattice step produces the transformation
$$|\mathrm{\Phi }_{AF}_\sigma |\mathrm{\Phi }_{AF}_\sigma $$
(7)
that is, a lattice step is equivalent to a spin flip, a feature that we call antiferromagnetic property. Since the spin-flipped state is also free of double occupation, $`|\mathrm{\Phi }_{AF}_\sigma `$ is a $`W=0`$ eigenstate. A ground state which is a single determinant is a quite unusual property for an interacting model like this.
Note that $`|\varphi _{1,A_1}^{(0)}\varphi _{2,A_1}^{(0)}`$ is equivalent to $`|\psi _{A_1}^{\prime \prime }\psi _{A_1}^{}`$, because this is just a unitary transformation of the $`A_1`$ wave functions; so $`|\mathrm{\Phi }_{AF}_\sigma `$ can also be written in terms of the old local orbitals (without any mix of the local states of different irreps of $`𝐆`$):
$$|\mathrm{\Phi }_{AF}_\sigma =|\psi _{A_1}^{\prime \prime }\psi _{A_1}^{}\psi _{B_2}^{}_\sigma |\psi _{B_1}^{\prime \prime }\psi _{E_x}^{}\psi _{E_y}^{}_\sigma .$$
(8)
This form of the ground state lends itself to be generalised (see below). For $`N>4`$, $`k`$ vectors arise that do not possess any special symmetry, the vectors $`R_ik`$ are all different for all $`R_iC_{4v}`$, and we get eight-dimensional irreps of $`𝐆`$. Recalling that $`|𝒮_{hf}|=2N2`$, one finds that $`𝒮_{hf}`$ contains $`N_e=\frac{1}{2}(\frac{N}{2}2)`$ irreps of dimension 8, one of dimension 4 and one of dimension 2 if $`N/2`$ is even and $`N_o=\frac{1}{2}(\frac{N}{2}1)`$ irreps of dimension 8 and one of dimension 2 if $`N/2`$ is odd.
To extend the theory to general $`N`$, we note that these $`k`$ vectors, since $`R_ik`$ are all different, are a basis of the regular representation of $`C_{4v}`$. Thus, by the Burnside theorem, each of them breaks into $`A_1A_2B_1B_2EE`$; diagonalizing $`n_{r=0}`$ and the point Group characters on the basis of the $`m`$-th eight-dimensional irrep of $`𝐆`$ one gets one-body states $`|\psi _I^{[m]}`$, where $`I`$ stands for the $`C_{4v}`$ irrep label, $`I`$=$`A_1`$, $`A_2`$, $`B_1`$, $`B_2`$, $`E_x`$, $`E_y`$, $`E_x^{}`$, $`E_y^{}`$; here we denote by $`E^{}`$ the second occourrence of the irrep $`E`$. The ground state wave function in $``$ for the half filled case is a generalized version of Equation (8). For even $`N/2`$, we have
$$|\mathrm{\Phi }_{AF}_\sigma |(\underset{m=1}{\overset{N_e}{}}\psi _{A_1}^{[m]}\psi _{B_2}^{[m]}\psi _{E_x}^{[m]}\psi _{E_y}^{[m]})\psi _{A_1}^{}\psi _{B_2}^{}\psi _{A_1}^{\prime \prime }_\sigma |(\underset{m=1}{\overset{N_e}{}}\psi _{A_2}^{[m]}\psi _{B_1}^{[m]}\psi _{E_x^{}}^{[m]}\psi _{E_y^{}}^{[m]})\psi _{E_x}^{}\psi _{E_y}^{}\psi _{B_1}^{\prime \prime }_\sigma ,$$
(9)
with $`\sigma =,`$. For odd $`N/2`$, $`|\mathrm{\Phi }_{AF}_\sigma `$ is similar but the maximum $`m`$ is $`N_o`$ and the $`|\psi ^{}`$ states do not occour. $`|\mathrm{\Phi }_{AF}_\sigma `$ is a $`W=0`$ state, transforms into $`|\mathrm{\Phi }_{AF}_\sigma `$ for each lattice step translation and manifestly shows an antiferromagnetic order (antiferromagnetic property). Since $`W`$ is a positive semidefinite operator $`|\mathrm{\Phi }_{AF}_\sigma `$ is actually a ground state. In the basis of local states these are the only two determinantal states ($`\sigma =,`$) with the above properties.
A few further remarks about $`|\mathrm{\Phi }_{AF}_\sigma `$ are in order. 1) Introducing the projection operator $`P_S`$ on the spin $`S`$ subspace, one finds that $`P_S|\mathrm{\Phi }_{AF}_\sigma |\mathrm{\Phi }_{AF}^S_\sigma 0,S=0,\mathrm{},N1`$. Then, $`{}_{\sigma }{}^{}\mathrm{\Phi }_{AF}|W|\mathrm{\Phi }_{AF}_{\sigma }^{}=_{S=1}^{N1}{}_{\sigma }{}^{}\mathrm{\Phi }_{AF}^S|W|\mathrm{\Phi }_{AF}^S_{\sigma }^{}=0`$, and this implies that there is at least one ground state of $`W`$ in $``$ for each $`S`$. The actual ground state of $`H`$ at weak coupling is the singlet $`|\mathrm{\Phi }_{AF}^0_\sigma `$. 2) The existence of this singlet $`W=0`$ ground state is a direct consequence of the Lieb theorem. Indeed the maximum spin state $`|\mathrm{\Phi }_{AF}^{N1}_\sigma `$ is trivially in the kernel of $`W`$; since the ground state must be a singlet it should be an eigenvector of $`W`$ with vanishing eigenvalue. 3) The above results and Lieb’s theorem imply that higher order effects split the ground state multiplet of $`H`$ and the singlet is lowest. 4) The Lieb theorem makes no assumptions concerning the lattice structure; adding the ingredient of the $`𝐆`$ symmetry we are able to explicitly display the wave function at weak coupling.
Using the explicit form of $`P_{S=0}`$ one finds that $`P_{S=0}|\mathrm{\Phi }_{AF}_\sigma =P_{S=0}|\mathrm{\Phi }_{AF}_\sigma `$. This identity allows us to study how the singlet component transforms under translations, reflections and rotations. We have found that the total momentum is $`K_{tot}=(0,0)`$. To make contact with Ref. we have also determined how it transforms under the $`C_{4v}`$ operations with respect to the center of an arbitrary placquette. It turns out that it is even under reflections and transforms as an $`s`$ wave if $`N/2`$ is even and as a $`d`$ wave if $`N/2`$ is odd. The present approach lends itself to obtain exact results for other fillings as well, but this will be shown elsewhere.
|
warning/0007/cond-mat0007153.html
|
ar5iv
|
text
|
# Electronic Stopping and Momentum Density of Diamond Obtained from First-Principles Calculations
## I Formulation and Method
The objective of this research is a quantitative, first-principles description of the energy deposition by a bare ion in diamond. The energy loss per unit path length of a massive, punctiform, charged particle with charge number $`Z_1`$ to a target isfulltext
$$\frac{dE}{dx}(𝐯)=\frac{(Z_1e)^2}{4\pi ^3ϵ_0v}d^3q\frac{𝐯𝐪}{q^2}_0^{\mathrm{}}𝑑\omega \delta (\omega 𝐪𝐯)\mathrm{Im}K_{\mathrm{𝟎},\mathrm{𝟎}}(𝐪,\omega ),$$
(1)
with $`𝐯`$ the projectile velocity in the target rest frame, $`e`$ the elementary charge unit, and $`ϵ_0`$ the vacuum permittivity. For reciprocal lattice vectors $`𝐆`$, $`K_{\mathrm{𝟎},\mathrm{𝟎}}`$ is the $`𝐆=𝐆^{}=\mathrm{𝟎}`$ (“head”) component of the inverse dielectric matrix defined by
$$\underset{𝐆^{}}{}ϵ_{𝐆,𝐆^{}}(𝐪,\omega )K_{𝐆^{},𝐆^{\prime \prime }}(𝐪,\omega )=\delta _{\mathrm{𝐆𝐆}^{\prime \prime }}$$
with respect to $`ϵ_{𝐆,𝐆^{}}(𝐤,\omega )`$, the wave-vector and frequency dependent microscopic dielectric matrix. In terms of the irreducible polarizability $`\mathrm{\Pi }`$, the dielectric matrix is
$$ϵ_{𝐆,𝐆^{}}(𝐪,\omega )=\delta _{\mathrm{𝐆𝐆}^{}}\frac{e^2}{ϵ_0|𝐪+𝐆|^2}\mathrm{\Pi }(𝐪+𝐆,𝐪+𝐆^{},\omega ),$$
given in the Random Phase ApproximationAdler (RPA) as a sum over all band pairs $`(\nu ,\nu ^{})`$ and an integral over the first Brillouin zone (BZ)
$`\mathrm{\Pi }(𝐪+𝐆,𝐪+𝐆^{},\omega )`$
$`={\displaystyle \underset{\nu \nu ^{}}{}}{\displaystyle _{\text{BZ}}}{\displaystyle \frac{d^3k}{(2\pi )^3}}{\displaystyle \frac{m_𝐆^{}m_𝐆^{}(f_{\nu 𝐤}f_{\nu ^{}𝐤+𝐪})}{\mathrm{}\omega +i\eta +E_{\nu 𝐤}E_{\nu ^{}𝐤+𝐪}}}.`$ (2)
$`E_{\nu ,𝐤}`$ are the band energies, and $`f_{\nu ,𝐤}`$ are Fermi occupation numbers ($`f_{\nu ,𝐤}=0`$ or 2 for $`E_{\nu ,𝐤}`$ above or below the Fermi energy $`E_F`$). The matrix elements in Eq. (2) are
$$m_𝐆\nu ^{}𝐤+𝐪|e^{i(𝐆+𝐪)𝐫}|\nu 𝐤.$$
We determine the ground-state electronic structure of solids within Density Functional Theory (DFT) as established in the Kohn-Sham (KS) variational procedure and implemented in the computational package gtoff.GTOFF The results of the all-electron, full-potential calculations are eigenfunctions $`\phi _{\nu ,𝐤}(𝐫)`$, expressed as linear combination of Gaussian Type Orbitals (GTO’s), and eigenvalues $`E_{\nu ,𝐤}`$.
The Bloch functions are finally expanded in a (truncated) plane wave (PW) series,
$$\phi _{\nu ,𝐤}=𝐫|\nu 𝐤=e^{i𝐤𝐫}u_{\nu 𝐤}(𝐫)=\frac{1}{V_{\text{UC}}}\underset{𝐆}{}e^{i(𝐤+𝐆)𝐫}u_{\nu 𝐤,𝐆},$$
to represent $`m_𝐆`$ as a simple sum over products of expansion coefficients $`u_{\nu 𝐤,𝐆}`$,fulltext where $`V_{\text{UC}}`$ is the volume of a unit cell (UC). The equivalence of the GTO and PW representations is maintained by monitoring the accumulated norm for each $`|\nu ,𝐤`$ relative to the exact values,
$$\underset{𝐆}{}|u_{\nu ,𝐤,𝐆}|^2=V_{\text{UC}}.$$
(3)
We subdivide the integration region of the integral (2) into $`𝐤`$-space tetrahedra. Recursive further subdivision of a given tetrahedron into smaller tetrahedra is done if $`f_{\nu 𝐤}f_{\nu ^{}𝐤+𝐪}`$ is not constant over all four vertices. Next comes linearization of the product of the matrix elements in the numerator and of the energy denominator inside each tetrahedron for each $`\omega `$. The resulting approximated integral is evaluated analytically.fulltext
The product $`|𝐪+𝐆|^2ϵ_{𝐆,𝐆^{}}(𝐪,\omega )`$ is calculated and, in compensation, the term $`q^2`$ in the denominator of Eq. (1) is dropped. The dielectric function is tabulated for $`𝐪`$ commensurate with the uniform mesh of wave vectors used in the underlying gtoff calculation but covering higher BZ’s as well as the first. \[Values outside the $`(𝐪,\omega )`$-meshes are replaced by the vacuum response, $`\mathrm{Im}K_{\mathrm{𝟎},\mathrm{𝟎}}(𝐪,\omega )=0`$.\] $`\mathrm{Im}K_{𝐆,𝐆^{}}`$ is linearized inside each $`𝐪`$-space tetrahedron. Multiplied by the linear factor $`𝐪𝐯`$, the integrals (1) over tetrahedra are done analytically, then summed.
## II Diamond
### II.1 Basis Sets, Electron Momentum Density
Diamond is studied at the experimental lattice parameter, $`a=6.74071a_0`$ for the cubic unit cell, where $`a_0`$ denotes one bohr. The Moruzzi-Janak-Williams parameterizationMoruzzi of the Hedin-Lundqvist local-density approximation (LDA) to the exchange-correlation potential is used. Small, highly contracted basis sets as used in Ref. Ayma, are generally insufficient to calculate the real parts of dielectric functions and subsequently the energy loss functions. We started from Partridge’s $`16s11p`$ set,PartrC contracted the seven tightest $`s`$-functions and the four tightest $`p`$-functions, removed the most diffuse $`s`$ and the two most diffuse $`p`$-functions to avoid approximate linear dependencies, and added a full set of three $`d`$-functions with exponents equal to the remaining most diffuse $`p`$-functions. (Without $`d`$-orbitals the total energy would rise by 0.24 eV/atom.) Site centered $`s`$\- and $`f`$-type fitting functions were used with exponents for the $`s`$-types as in Ref. TrickeyC, , and for the $`f`$-types $`0.5/a_0^2`$ and $`0.2/a_0^2`$ as in Ref. Lu, . Space group and lattice type are those of silicon; hence we may refer to a prior gtoff studyBoeSi for the symmetry properties of the fitting functions.
The density of states (DOS) computed with this $`9s6p3d`$ basis (84 basis functions in the primitive UC) as shown in Fig. 1 is stable up to $`70`$ eV above the Fermi energy with respect to further de-contraction. Band gaps are too small compared with experiments, as usual for the LDA and known from other DFT calculations on diamond.diabands
The lowest 24 bands were expanded into 531 plane waves, a cut-off at $`|𝐆|=7.3/a_0`$, with a norm in Eq. (3) above $`0.85V_{\text{UC}}`$ for the $`K`$-shell electrons (excitations from which were excluded in the subsequent calculations of $`ϵ_{𝐆,𝐆^{}}`$ and $`dE/dx`$ anyway), and a norm above $`0.98V_{\text{UC}}`$ for the remaining 22 bands. A first result is the all-electron momentum density (EMD)
$$\rho (𝐤+𝐆)\frac{1}{V_{\text{UC}}}\underset{\nu }{}f_{\nu ,𝐤}|u_{\nu ,𝐤,𝐆}|^2$$
in Fig. 2. The values near zero momentum $`𝐤+𝐆=0`$ are the states with long wavelengths and reveal the macroscopic symmetry of the crystal system, the 4-fold axis of the cubic system here. If $`|𝐤+𝐆|`$ is of the order of half a reciprocal lattice vector, the interference with the next nearest neighbors in the lattice becomes visible; the eight foothills in the figure may be interpreted as a projection of the four corners of a carbon tetrahedron onto the $`(001)`$ plane complemented by the inversionFalk
$$u_{\nu 𝐤,𝐆}=u_{\nu 𝐤,𝐆}^{}.$$
If $`|𝐤+𝐆|`$ is large, the spherical symmetry of the core states prevails.
### II.2 Dielectric Function
The element $`ϵ_{\mathrm{𝟎},\mathrm{𝟎}}(𝐪,\omega )`$ of the dielectric matrix is shown in Fig. 3. The main absorption peak at 11 eV for $`|𝐪|0`$ corresponds to direct transitions from the top of the valence to the bottom of the conduction band at X and L.Painter
Local Field Effects are illustrated in Fig. 4. $`\mathrm{Im}ϵ_{\mathrm{𝟎},\mathrm{𝟎}}`$ repeats some values of Fig. 3; $`\mathrm{Im}[1/K_{\mathrm{𝟎},\mathrm{𝟎}}]`$ includes an estimate of the local field by calculating $`K_{𝐆,𝐆^{}}`$ as the inverse of a $`9\times 9`$ dielectric matrix which contains $`𝐆=0`$ and the eight vectors of the closest shell in the bcc reciprocal lattice. The reduction of the values without LFE ($`|\mathrm{Im}ϵ_{\mathrm{𝟎},\mathrm{𝟎}}|`$, open symbols) compared to those with LFE ($`|\mathrm{Im}(1/K_{\mathrm{𝟎},\mathrm{𝟎}})|`$, filled symbols) is of the order reported by Van Vechten and MartinVechten (without their “dynamical correlations”). The different sign of the effect for frequencies above and below the peak has been noticed before.GavriDia The differences are even smaller for the energy loss function. Hence the energy loss in the next paragraph was calculated from $`ϵ_{\mathrm{𝟎},\mathrm{𝟎}}(𝐪,\omega )`$ alone.
### II.3 Computed Electronic Energy Loss
To integrate the energy loss function as described by Eq. (1), the BZ mesh was reduced to $`8\times 8\times 8`$ points, i.e., 35 points in the irreducible Brillouin zone (IBZ). The result is shown in Fig. 5. The dielectric matrix was tabulated on a $`30\times 30\times 30`$ mesh in $`𝐪`$-space (parallelepiped with three edges of length $`6.1/a_0`$), and the stopping power was integrated over the superset of all $`𝐪`$-values obtained from these via point group operations with Seitz symbol $`\{O|𝐰\}`$,
$$K_{𝐆,𝐆^{}}(𝐪,\omega )=e^{iO(𝐆𝐆^{})𝐰}K_{O𝐆,O𝐆^{}}(O𝐪,\omega ).$$
It is about 0–15% lower than experimental resultsKaferb for $`v1\mathrm{}1.3v_0`$ which are shown in Fig. 6. ($`v_02.1910^6`$ m/s is one atomic unit of velocity. Values within the LPDA represent intgrals of the all-electron density of gtoff Fitting Functions weighted with the stopping number of the FEG in the RPA.) Reduction of the integrated $`𝐪`$-region to the parallelepiped decreases the stopping cross section by approximately 10% at $`vv_0`$.wbw18 Within the framework of linear dielectric response an underestimation is appropriate though, because terms of $`O(Z_1^3)`$ will add about 15% to $`S(v)`$ at $`v1.5v_0`$.Jackson
A new result is the “ionic” band gap, the non-linear suppression of $`S(v)`$ for $`v0.2v_0`$, which is approximately the value extracted fromwbw18
$$v<\frac{v_0}{2}\sqrt{\mathrm{}\omega _g/E_0}$$
(4)
for a band gap of $`\mathrm{}\omega _g=4`$ eV. ($`E_0`$ is one rydberg.) By way of contrast, calculations within the local-plasma-density approximation usually integrate over volume elements in real space that are parameterized by the homogeneous electron gas, and inevitably yield $`S(v)v`$ at low velocities. However, an attempt at experimental verification of this reduction of the energy loss in a wide-gap material is handicapped by the additional nuclear energy loss, which has an estimated maximum of $`S0.810^{15}`$ eVcm$`{}_{}{}^{2}/`$atoms at $`v0.08v_0`$.Littmark
Note that Eq. (4) predicts contributions from $`K`$-shell excitations to start at $`v2.2v_0`$, which are not included here. They have been estimated to shift the maximum to higher velocities by about $`0.2v_0`$,SabinDia and to grow until $`v5.3v_0`$.Paul
###### Acknowledgements.
Helpful conversations with S.B. Trickey and J.R. Sabin are gratefully acknowledged. This work was supported by grant DAA-H04-95-1-0326 from the U.S. Army Research Office.
|
warning/0007/cond-mat0007309.html
|
ar5iv
|
text
|
# Lectures on Non Perturbative Field Theory and Quantum Impurity Problems: Part II.
## 1 Some generalities on form-factors
I follow here the notations and conventions of , in particular section 5 of the latter notes.
The space of states is generated by the vectors
$$|\alpha _1,\mathrm{},\alpha _n_{a_1,\mathrm{},a_n}=Z_{a_1}^{}(\alpha _1)\mathrm{}Z_{a_n}^{}(\alpha _n)|0>$$
(1)
and the dual space by
$${}_{}{}^{a_n,\mathrm{},a_1}\alpha _n,\mathrm{},\alpha _1|=0|Z^{a_n}(\alpha _n)\mathrm{}Z^{a_1}(\alpha _1)$$
(2)
Here, the labels $`a`$ stand for instance for solitons/antisolitons ($`\pm `$) and breathers in the sine-Gordon model, $`\alpha `$ denotes the rapidity, and the $`Z,Z^{}`$ are the annihilation creation operators of the Faddeev Zamolodchikov algebra . For a given set of rapidities and quantum numbers, the states with different orderings are related through S matrix elements. One usually call “in” states the states for which $`\alpha _1>\alpha _2\mathrm{}>\alpha _n`$, and “out” states those for which $`\alpha _1<\alpha _2\mathrm{}<\alpha _n`$. In or out states form a complete basis of the set of states, but it is convenient to keep some redundancy and consider all possible rapidity orderings, after division by the appropriate degeneracy factors.
Form-factors (more correctly, “generalized form-factors” since no order of the rapidities is prescribed) of an operator $`𝒪`$ in a bulk theory are defined as :
$$f(\alpha _1,\mathrm{},\alpha _n)_{a_1,\mathrm{},a_n}=<0|𝒪(0,0)|Z_{a_1}^{}(\alpha _1)\mathrm{}Z_{a_n}^{}(\alpha _n)|0>$$
(3)
where $`|0>`$ is the ground state. We chose to take the operator at the origin, since relativistic invariance determines trivially the dependence of the matrix elements on the coordinates of insertion (we set $`\mathrm{}=1`$)
$$<0|𝒪(x,t)|Z_{a_1}^{}(\alpha _1)\mathrm{}Z_{a_n}^{}(\alpha _n)|0>=e^{i(PxEt)}f(\alpha _1,\mathrm{},\alpha _n)_{a_1,\mathrm{},a_n}$$
(4)
$`S`$ matrices have in many cases been obtained by explicit solution of properly regularized quantum field theories. The strategy, which was implemented quite succesfully in the massive Thirring model for instance , is to find out eigenstates, fill the ground state, determine the possible excitations over it, and finally compute their scattering, which is the result of a combination of the bare scattering and “dressing” effects coming from interaction with the ground state. Needless to say, the procedure is laborious, and it has proven much faster to obtain the $`S`$ matrices in a more abstract way, using general axioms of $`S`$ matrix theory, factorizability, and assumptions of maximal analyticity .
It is even harder to obtain the matrix elements from the solution of bare theories. A program to do this is under way , but very few results of interest for our problem have been obtained so far. Form-factors can however be determined (though not quite as easily as $`S`$ matrices) using an axiomatic approach similar in spirit to what was done for the $`S`$ matrices, and this is what we would like to discuss briefly here.
One of the consequences of integrability is that multiple particle processes can be reexpressed in terms of two-particle ones. The basic objects in this approach are therefore those involving only a pair of particles. To start, consider the two particle $`S`$ matrix: by relativistic invariance, it is expected to depend only on the Mandelstam variable
$$s=(p_a(\alpha _1)+p_b(\alpha _2))^2=M_1^2+M_2^2+2M_1M_2\mathrm{cosh}(\alpha _1\alpha _2)$$
(5)
The function $`S(s)`$ is expected to be an analytic function of $`s`$, with a cut along the real axis running from $`s=(M_1+M_2)^2`$ to $`s=\mathrm{}`$, and another running from $`s=\mathrm{}`$ to $`s=(M_1M_2)^2`$. The exchange of in and out states corresponds to exchanging the upper lip of the right cut $`\text{Im }s=0+`$ with the lower lip of the right cut $`\text{Im }s=0`$. Crossing instead exchanges the upper lip of the right cut with the lower lip of the left cut.
This plane with the cuts transforms into the “physical” strip $`0\text{Im }\alpha \pi `$ in terms of the variable $`\alpha =\alpha _1\alpha _2`$. The line $`\text{Im }\alpha =0`$ corresponds to the right cut, while $`\text{Im }\alpha =\pi `$ corresponds to the left cut. Upper and lower lips correspond to $`\text{Re }\alpha >0`$, resp. $`\text{Re }\alpha <0`$.
In terms of the $`\alpha `$ variable, the standard properties of the $`S`$ matrix are:
$``$ $`S_{a_1a_2}^{a_1^{}a_2^{}}(\alpha )=S_{a_2a_1}^{a_2^{}a_1^{}}(\alpha )\text{C}`$ (6)
$``$ $`S_{a_1a_2}^{a_1^{}a_2^{}}(\alpha )=S_{\overline{a}_1\overline{a}_2}^{\overline{a}_1^{}\overline{a}_2^{}}(\alpha )\text{P}`$
$``$ $`S_{a_1a_2}^{a_1^{}a_2^{}}(\alpha )=S_{a_2^{}a_1^{}}^{a_2a_1}(\alpha )\text{T}`$
and
$``$ $`S(\alpha )`$ is real for $`\alpha `$ purely imaginary (7)
$``$ $`\text{Unitarity}S_{a_1a_2}^{b_1b_2}(\alpha )S_{b_1b_2}^{c_1c_2}(\alpha )=\delta _{a_1}^{c_1}\delta _{a_2}^{c_2}`$
$``$ $`\text{Crossing}S_{a_1a_2}^{b_1b_2}(\alpha )=S_{a_1\overline{b}_2}^{b_1\overline{a}_2}(i\pi \alpha )`$
In addition of course, the $`S`$ matrix is a solution of the Yang Baxter equation, which reads in components
$$S_{a_1a_2}^{b_1b_2}(\alpha _1\alpha _2)S_{b_1a_3}^{c_1b_3}(\alpha _1\alpha _3)S_{b_2b_3}^{c_2c_3}(\alpha _2\alpha _3)=S_{a_2a_3}^{b_2b_3}(\alpha _2\alpha _3)S_{a_1b_3}^{b_1c_3}(\alpha _1\alpha _3)S_{b_1b_2}^{c_1c_2}(\alpha _1\alpha _2)$$
(8)
The space of states is conveniently built using the Faddeev Zamolodchikov algebra
$`Z^{a_1}(\alpha _1)Z^{a_2}(\alpha _2)`$ $`=`$ $`S_{a_1^{}a_2^{}}^{a_1a_2}(\alpha _1\alpha _2)Z^{a_2^{}}(\alpha _2)Z^{a_1^{}}(\alpha _1)`$
$`Z_{a_1}^{}(\alpha _1)Z_{a_2}^{}(\alpha _2)`$ $`=`$ $`S_{a_1a_2}^{a_1^{}a_2^{}}(\alpha _1\alpha _2)Z_{a_2^{}}^{}(\alpha _2)Z_{a_1^{}}^{}(\alpha _1)`$
$`Z^{a_1}(\alpha _1)Z_{a_2}^{}(\alpha _2)`$ $`=`$ $`S_{a_2a_1^{}}^{a_2^{}a_1}(\alpha _1\alpha _2)Z_{a_2^{}}^{}(\alpha _2)Z^{a_1^{}}(\alpha _1)+2\pi \delta _{a_2}^{a_1}\delta (\alpha _1\alpha _2)`$ (9)
for which several of the $`S`$ matrix properties (but not those involving crossing) are just consistency relations.
In terms of $`\alpha `$, $`S`$ is a meromorphic function, whose only singularities in the physical strip are poles on the imaginary axis, corresponding in general to bound states.
Consider now the two particle form-factor $`f(\alpha _1,\alpha _2)_{a_1,a_2}`$. Up to some simple dimensionful terms (which are absent if the operator is scalar), this is also a function of the Mandelstam variable $`s`$. Its analytical properties are similar although a bit different than for $`S`$; in particular, the left cut does not exist for form-factors.
From the second relation of (9), it follows immediately that
$$f(\alpha _1,\alpha _2)_{a_1,a_2}=S_{a_1^{}a_2^{}}^{a_1a_2}(\alpha _1\alpha _2)f(\alpha _2,\alpha _1)_{a_2^{},a_1^{}}$$
(10)
If $`\alpha _1>\alpha _2`$, the state $`|\alpha _1,\alpha _2>_{a_1,a_2}`$ is an in state, with the corresponding out state $`|\alpha _2,\alpha _1>_{a_2,a_1}`$. The exchange of the two on the other hand corresponds to going from the upper to the lower lip of the cut in the s-plane, ie from $`\text{Im }\alpha =0`$ to $`\text{Im }\alpha =2\pi `$ in the $`\alpha `$ plane. It follows that
$$f(\alpha _1+2i\pi ,\alpha _2)_{a_1,a_2}=f(\alpha _2,\alpha _1)_{a_2,a_1}$$
(11)
Equations (10,11) are the basic tool to determine form-factors. To see this, suppose for simplicity that the two particle S matrix is diagonal (actually, the following would also hold if it could be diagonalized by a rapidity independent change of basis: such is the case for the sine-Gordon model in the $`+`$ sector). This would correspond, for instance, to the sinh-Gordon case, which we discuss in details below. Assuming the operator of interest is scalar, $`f(\alpha )`$ is thus found to satisfy the constraints
$`f(\alpha )=S(\alpha )f(\alpha )`$
$`f(2i\pi +\alpha )=f(\alpha )`$ (12)
The minimal solution of these equations is obtained by assuming that $`f`$ does not have poles nor zeroes (except the threshold pole at $`\alpha =0`$) in the strip $`\text{Im }\alpha [0,2\pi ]`$. Consider now a closed contour $`𝒞`$ enclosing this strip. By Cauchy’s theorem one has
$`{\displaystyle \frac{d}{d\alpha }}\mathrm{ln}f^{min}(\alpha )`$ $`=`$ $`{\displaystyle \frac{1}{8i\pi }}{\displaystyle _𝒞}{\displaystyle \frac{dz}{\mathrm{sinh}^2(z\alpha )/2}}\mathrm{ln}f(z)`$ (13)
$`=`$ $`{\displaystyle \frac{1}{8i\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dz}{\mathrm{sinh}^2(z\alpha )/2}}\mathrm{ln}{\displaystyle \frac{f(z)}{f(z+2i\pi )}}`$
$`=`$ $`{\displaystyle \frac{1}{8i\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dz}{\mathrm{sinh}^2(z\alpha )/2}}\mathrm{ln}S(z)`$
Very often, the S matrix can be recast in the form
$$S(\alpha )=\mathrm{exp}\left[_0^{\mathrm{}}𝑑xg(x)\mathrm{sinh}(x\alpha /i\pi )\right]$$
(14)
from which one finds
$$f^{min}=\text{cst }\mathrm{sinh}\frac{\alpha }{2}\mathrm{exp}\left\{_0^{\mathrm{}}𝑑x\frac{g(x)}{\mathrm{sinh}x}\mathrm{sin}^2[x(i\pi \alpha )/2\pi ]\right\}$$
(15)
In general, form-factors are expressed in terms of $`f^{min}`$ and simple functions which encode the required pole structure, see below.
Generalizing the two equations (12) we have, for $`n`$ particles,
$`f(\alpha _1,\mathrm{},\alpha _i,\alpha _{i+1},\mathrm{},\alpha _n)_{a_1,\mathrm{},a_i,a_{i+1},\mathrm{},a_n}`$ $`=`$ $`f(\alpha _1,\mathrm{},\alpha _{i+1},\alpha _i,\mathrm{},\alpha _n)_{a_1,\mathrm{},b_{i+1},b_i,\mathrm{},a_n}`$
$`\times `$ $`S_{a_ia_{i+1}}^{b_ib_{i+1}}(\alpha _i\alpha _{i+1})`$
$`f_{a_1\mathrm{}a_n}(\alpha _1+2i\pi ,\mathrm{},\alpha _n)`$ $`=`$ $`f_{a_2\mathrm{}a_n,a_1}(\alpha _2,\mathrm{},\alpha _n,\alpha _1)`$ (16)
The form-factors are meromorphic functions of the rapidity differences $`\alpha _{ij}`$ in the strip $`0\text{Im }\alpha _{ij}2\pi `$, with two kinds of simple poles.
One type of poles is called annihilation, or kinematical, pole. Such poles are always expected, even if the theory has no bound states, when some of the rapidities in the form-factor differ by $`i\pi `$, corresponding physically to the presence of a pair particle-antiparticle in the process. The residue is a form-factor with two fewer particles
$$i\text{ Res}_{\alpha ^{}=\alpha +i\pi }f(\alpha ^{},\alpha ,\alpha _1,\alpha _2,\mathrm{},\alpha _n)_{\overline{a}aa_1\mathrm{}a_n}=\left[\delta _{a_1}^{b_1}\mathrm{}\delta _{a_n}^{b_n}S_{a_1\mathrm{}a_n}^{b_1\mathrm{}b_n}(\alpha _1\mathrm{}\alpha _n|\alpha )\right]f(\alpha _1,\mathrm{},\alpha _n)_{b_1,\mathrm{},b_n},$$
(17)
where, we have defined the $`S`$ matrix element
$$S_{a_1\mathrm{}a_n}^{b_1\mathrm{}b_n}(\alpha _1\mathrm{}\alpha _n|\alpha )=S_{a_1c_n}^{b_1c_1}(\alpha _1\alpha )S_{a_2c_1}^{b_2c_2}(\alpha _2\alpha )\mathrm{}S_{a_nc_{n1}}^{b_nc_n}(\alpha _n\alpha )$$
(18)
This provides a resursive relation between form-factors with $`n+2`$ and $`n`$ particles, which proves crucial in determining the multiple particle form-factors.
The other type of poles has a more physical origin, and corresponds to the appearance of bound states. In words, form-factors obey relations that mimic the bootstrap structure of the theory: if a particle appears as a bound state of two others, its form-factors are residues of form-factors involving these particles. In formulas, things are a bit more complicated. If $`c`$ is a bound state of particles $`a`$ and $`b`$, such that the $`S`$ matrix has a simple pole
$$S_{ab}^{de}(\alpha )\frac{ig_{ab}^cg_c^{de}}{\alpha iu_{ab}^c}$$
(19)
then the form factor $`f(\alpha _1,\alpha _2,\mathrm{},\alpha _{n+2})_{a_1\mathrm{}a_nab}`$ has a pole when $`\alpha _{n+1}\alpha _{n+2}=iu_{ab}^c`$, with residue
$$i\text{Res }_{\alpha _{n+1}\alpha _{n+2}=iu_{ab}^c}f(\alpha _1,\alpha _2,\mathrm{},\alpha _{n+2})_{a_1\mathrm{}a_nab}=g_{ab}^cf(\alpha _1,\alpha _2,\mathrm{},\alpha _n,\alpha _{n+1}iu_{ab}^c/2)_{a_1\mathrm{}a_nc}$$
(20)
where the value of the $`n+1^{th}`$ rapidity follows from conservation of energy and momentum at the three particle vertex $`abc`$. Formulas ( 17,20) are usually called LSZ reduction formulas, from the pioneering paper of Lehmann, Symanzik and Zimmermann .
In the simple case where there is only one type of particle (like in the sinh-Gordon model, see below), the general solution of these equations can always be written in the form
$$f(\alpha _1,\mathrm{},\alpha _n)=K(\alpha _1,\mathrm{},\alpha _n)\underset{i<j}{}f^{min}(\alpha _i\alpha _j)$$
(21)
where $`K`$ is a completely symmetric, doubly periodic function, which contains all the physical poles.
Finally, we point out that other form-factors can be obtained by crossing
$${}_{a_1^{},\mathrm{},a_m^{}}{}^{}\alpha _1^{},\mathrm{},\alpha _m^{}|𝒪|\alpha _1,\mathrm{},\alpha _n_{a_1,\mathrm{},a_n}^{}=f(\alpha _1^{}+i\pi ,\mathrm{},a_m^{}+i\pi ,\alpha _1,\mathrm{},\alpha _n)_{\overline{a}_1^{},\mathrm{},bara_m^{},a_1,\mathrm{},a_n}$$
(22)
Let us now discuss indeed the sinh-Gordon model in more details - here, I follow closely the paper . The action is
$$S=\frac{1}{16\pi g}_{\mathrm{}}^{\mathrm{}}𝑑x𝑑y\left[\left(_x\mathrm{\Phi }\right)^2+\left(_y\mathrm{\Phi }\right)^2+\mathrm{\Lambda }\mathrm{cosh}\mathrm{\Phi }\right].$$
(23)
Note that here the free boson is normalized differently than in : this will prove more convenient below. Eq. (23) defines an integrable massive theory; the conformal weights of the perturbing operator are $`h=\overline{h}=g`$. The spectrum is very simple and consists of a single particle of mass $`M`$, and S matrix :
$$S(\alpha )=\frac{\mathrm{tanh}\frac{1}{2}\left(\alpha i\frac{\pi B}{2}\right)}{\mathrm{tanh}\frac{1}{2}\left(\alpha +i\frac{\pi B}{2}\right)},$$
(24)
where :
$$B=\frac{2g}{1+g}$$
Observe the remarkable duality of the S matrix in $`B2B`$, i.e. in $`g1/g`$. This duality is certainly not obvious at the level of the action, and is deeply non perturbative in nature.
To determine the minimal form-factor, it is convenient to replace trigonometric functions with $`\mathrm{\Gamma }`$ functions using the basic identity
$$\mathrm{\Gamma }(z)\mathrm{\Gamma }(1z)=\frac{\pi }{\mathrm{sin}\pi z}$$
(25)
together with the integral representation
$$\mathrm{ln}\mathrm{\Gamma }(z)=_0^{\mathrm{}}\left[(z1)e^t+\frac{e^{tz}e^t}{1e^t}\right]\frac{dt}{t}$$
(26)
One finds then
$$S(\alpha )=\mathrm{exp}\left[2_0^{\mathrm{}}\frac{dx}{x}\frac{\mathrm{cosh}x(1B)/2}{\mathrm{cosh}x/2}\mathrm{sinh}(\alpha x/i\pi )\right]$$
(27)
Using a similar representation for $`\mathrm{ln}\mathrm{sinh}\alpha /2`$, we find finally
$$f^{min}(\alpha )=𝒩\mathrm{exp}\left\{8_0^{\mathrm{}}\frac{dx}{x}\frac{\mathrm{sinh}(\frac{xB}{4})\mathrm{sinh}[\frac{x(2B)}{4}]\mathrm{sinh}(\frac{x}{2})}{\mathrm{sinh}^2(x)}\mathrm{sin}^2[x(i\pi \alpha )/2\pi ]\right\},$$
(28)
Here, we have put the normalization which is standard in the literature :
$$𝒩=\mathrm{exp}\left[4_0^{\mathrm{}}\frac{dx}{x}\frac{\mathrm{sinh}(\frac{xB}{4})\mathrm{sinh}[\frac{x(2B)}{4}]\mathrm{sinh}(\frac{x}{2})}{\mathrm{sinh}^2(x)}\right]$$
Consider now the form-factors of the field $`\mathrm{\Phi }`$ itself. Since $`\mathrm{\Phi }`$, as well as the creation operators of the sinh-Gordon particle, are odd under the $`Z_2`$ symmetry $`\mathrm{\Phi }\mathrm{\Phi }`$, only form-factors witn an even number of particles are non vanishing. One finds the general formula
$$f(\alpha _1,\mathrm{},\alpha _{2n+1})=\mu \left(\frac{4\mathrm{sin}\frac{\pi B}{2}}{F^{min}(i\pi ,B)}\right)^n\sigma _{2n+1}^{(2n+1)}P_{2n+1}(x_1,\mathrm{},x_{2n+1})\underset{i<j}{}\frac{f^{min}(\alpha _i\alpha _j)}{x_i+x_j},$$
(29)
where we introduced $`xe^\alpha `$ and the $`\sigma `$’s are the basic symmetric polynomials :
$$\sigma _p=\underset{i_1<i_2<\mathrm{}<i_p}{}x_{i_1}x_{i_2}\mathrm{}x_{i_p},$$
with the convention $`\sigma _0=1`$ and $`\sigma _p=0`$ if $`p`$ is greater than the number of variables. The $`P_{2n+1}`$’s are symmetric polynomials, which can be obtained by solving LSZ recursion relations. The first ones read :
$`P_3(x_1,\mathrm{},x_3)=`$ $`1`$
$`P_5(x_1,\mathrm{},x_5)=`$ $`\sigma _2\sigma _3c_1^2\sigma _5`$
$`P_7(x_1,\mathrm{},x_7)=`$ $`\sigma _2\sigma _3\sigma _4\sigma _5c_1^2(\sigma _4\sigma _5^2+\sigma _1\sigma _2\sigma _5\sigma _6+\sigma _2^2\sigma _3c_1^2\sigma _2\sigma _5)`$ (30)
$`c_2(\sigma _1\sigma _6\sigma _7+\sigma _1\sigma _2\sigma _4\sigma _7)+\sigma _3\sigma _5\sigma _6)+c_1c_2^2\sigma _7^2.`$
with $`c_1=2\mathrm{cos}\pi B/2`$, $`c_2=1c_1^2`$. Observe that except for the overall normalization $`\mu (g)`$, these expressions are invariant in the duality transformation $`g\frac{1}{g}`$. This is expected from the duality of the S matrix itself - as for the role of the overall normalization, it will be discussed later.
The conventional normalization is $`\mu =1`$, which corresponds to chosing $`<0|\mathrm{\Phi }(0)|\alpha >=\frac{1}{\sqrt{2}}`$. We shall make a different choice later on, when we consider the massless limit.
The method can be generalized to models with several types of particles, like the sine-Gordon model: this will be discussed in the following sections. Of course, all the foregoing equations for form-factors have been established within the context of ordinary massive integrable field theories. The case of massless theories (first considered in a slightly different context in ) will be handled simply by taking the appropriate massless (ultra violet) limit in all the equations. This is somewhat safer than trying directly to formulate axioms for a massless theory per se, massless scattering presenting some physical ambiguities (see for a detailed discussion of this point).
We shall mostly consider physical properties related with the $`U(1)`$ currents $`\mathrm{\Phi }`$; hence, our discussion will be centered on the form-factors of this operator. In the last section however, we give the example of a correlator involving vertex operators.
We now discuss in more details the questions at stake in the case of the sinh-Gordon model. It has little physical interest in the present context, but is pedagogically quite useful
## 2 Example: The sinh-Gordon model.
### 2.1 Massless form-factors and the bulk current-current correlators.
In most of the following calculations, we shall work in Euclidian space with $`x,y`$ coordinates. Imaginary time is at first considered as running along $`x`$. The action and $`S`$ matrix for the massive sinh-Gordon model were given in the introduction (23).
Let us now try to describe the free boson theory as a massless limit of this model. First, recall the current correlators (we could use here the notation $`\varphi ,\overline{\varphi }`$ for the chiral components, but since these get mixed in the presence of the boundary, we will not do so)
$`<_z\mathrm{\Phi }(z,\overline{z})_z^{}\mathrm{\Phi }(z^{},\overline{z}^{})>={\displaystyle \frac{2g}{(zz^{})^2}}`$
$`<_{\overline{z}}\mathrm{\Phi }(z,\overline{z})_{\overline{z}^{}}\mathrm{\Phi }(z^{},\overline{z}^{})>={\displaystyle \frac{2g}{(\overline{z}\overline{z}^{})^2}}.`$ (31)
To start, we wish to recover these correlators using form-factors. We thus take the massless limit $`\mathrm{\Lambda }0`$ of the factorized scattering description of (23). As discussed in , we start by writing $`\alpha =\pm (A+\theta )`$ and take simultaneously $`A\mathrm{}`$ and $`M0`$ with $`Me^A/2m`$, $`m`$ finite. In that case, the spectrum separates into Right and Left movers with respectively $`E=P=me^\theta `$ and $`E=P=me^\theta `$. The scattering of R and L movers is still given by (24) where $`\alpha \pm \theta `$. The RL and LR scattering becomes a simple phase, $`e^{i\pi B/2}`$. This phase will turn out to cancel out at the end of all computations, but is confusing to keep along. We just set it equal to unity in the following, that is we consider all L and R quantities as commuting.
In this new description of the massless theory, we will need form factors in order to compute (31). By taking the massless limit of the formulas given in the introduction , it is easy to check that $`\mathrm{\Phi }`$ can alter only the right or left content of states; in other words, matrix elements of $`\mathrm{\Phi }`$ between states which have different content both in the left and right sectors vanish (that is, $`\mathrm{\Phi }=\varphi +\overline{\varphi }`$!)
Our conventions are conveniently summarized by giving the one particle form factor of the sinh-Gordon field :
$`<0|\mathrm{\Phi }(x,y)|\theta >_R`$ $`=\mu \mathrm{exp}\left[me^\theta (x+iy)\right]`$
$`<0|\mathrm{\Phi }(x,y)|\theta >_L`$ $`=\mu \mathrm{exp}\left[me^\theta (xiy)\right],`$ (32)
and we will use the obvious notation :
$$f(\theta _1,\mathrm{},\theta _{2n+1})=<0|\mathrm{\Phi }|\theta _1,\mathrm{},\theta _{2n+1}>_{R,\mathrm{},R},$$
(33)
with the normalization of asymptotic states $`{}_{R}{}^{}<\theta |\theta ^{}>_R=2\pi \delta (\theta \theta ^{})`$. Here, $`f`$ depends on the interaction strength $`g`$, but we do not indicate it explicitely for simplicity. We have the following properties :
$`<0|\mathrm{\Phi }|\theta _1,\mathrm{},\theta _{2n+1}>_{R,\mathrm{},R}`$ $`=`$ $`\left(<0|\mathrm{\Phi }|\theta _1,\mathrm{},\theta _{2n+1}>_{L,\mathrm{},L}\right)^{}`$
$`<0|\mathrm{\Phi }|\theta _1,\mathrm{},\theta _{2n+1}>_{R,\mathrm{},R}`$ $`=`$ $`<0|\mathrm{\Phi }|\theta _{2n+1},\mathrm{},\theta _1>_{L,\mathrm{},L}.`$ (34)
These form factors are expressed just as in the massive case
$$f(\theta _1,\mathrm{},\theta _{2n+1})=\mu \left(\frac{4\mathrm{sin}\frac{\pi B}{2}}{F_{min}(i\pi ,B)}\right)^n\sigma _{2n+1}^{(2n+1)}P_{2n+1}(x_1,\mathrm{},x_{2n+1})\underset{i<j}{}\frac{f_{min}(\theta _i\theta _j)}{x_i+x_j},$$
(35)
where we introduced $`xe^\theta `$ and the $`\sigma `$’s are the same symmetric polynomials as before.
We will now chose the normalization $`\mu `$ by demanding that the result (31) be recovered. Using form factors, this two point function expands, assuming $`Rez<Rez^{}`$, as :
$`<0|_z\mathrm{\Phi }(z,\overline{z})_z^{}\mathrm{\Phi }(z^{},\overline{z}^{})|0>=`$ $``$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n+1}}{(2\pi )^{2n+1}(2n+1)!}m^2\left(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}}\right)^2}`$ (36)
$`\times `$ $`\mathrm{exp}\left[m(zz^{})(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}})\right)]|f(\theta _1,\mathrm{},\theta _{2n+1})|^2.`$
Now, by relativistic invariance, all the form factors depend only on differences of rapidities. Setting $`m(zz^{})e^{\theta _0}`$, (where $`\theta _0`$ will in general be complex), one can shift all the $`\beta `$’s by $`\theta _0`$ to factor out, for any $`2n+1`$ particle contributions, a factor $`\frac{1}{(zz^{})^2}`$. Hence, the form factor expansion gives the result (31) provided $`\mu `$ is chosen such that
$$\underset{n=0}{\overset{\mathrm{}}{}}I_{2n+1}=2g,$$
(37)
where
$$I_{2n+1}=\frac{d\theta _1\mathrm{}d\theta _{2n+1}}{(2\pi )^{2n+1}(2n+1)!}\left(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}}\right)^2e^{(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}})}|f(\theta _1,\mathrm{},\theta _{2n+1})|^2.$$
(38)
In practice, this sum cannot be computed analytically, but it can be easily evaluated numerically. The convergence is extremely fast with $`n`$, and for most practical purposes, the consideration of up to five particles is enough to get correct results up to $`10^4`$. Similar convergence properties were observed in in the massive case; note however that in this massless case, contributions with a higher number of particles are not damped-off by exponential mass terms.
It must be emphasized that this result is very peculiar to the current operator. For most other chiral operators, the correct $`(zz^{})`$ dependence involves a non trivial anomalous dimension, instead of the naive engineering dimension. Hence, this dependence is not obtained term by term, as observed here, but rather once the whole series is summed up. Truncating the series to any finite $`n`$ does not, in such cases, give reliable results all the way from short to large distances, unless some additional tricks are performed (see below).
### 2.2 Current current correlators with a boundary
Having fixed the form-factors normalization, let us now consider the theory with a boundary. The geometry of the problem is such that the boundary stands at $`x=0`$, and runs parallel to the $`y=it`$ axis. The action is now :
$$S=\frac{1}{16\pi g}_{\mathrm{}}^0𝑑x_{\mathrm{}}^{\mathrm{}}𝑑y\left[\left(_x\mathrm{\Phi }\right)^2+\left(_y\mathrm{\Phi }\right)^2+\mathrm{\Lambda }\mathrm{cosh}\mathrm{\Phi }\right]+\lambda _{\mathrm{}}^{\mathrm{}}𝑑y\mathrm{cosh}\frac{1}{2}\mathrm{\Phi }(x=0,y).$$
(39)
This model is also integrable for any choice of $`\mathrm{\Lambda },\lambda `$. The boundary dimension of the perturbing operator is $`x=g`$. We can in particular take the limit where $`\mathrm{\Lambda }0`$ while $`\lambda `$ remains finite. It then describes a theory which is conformal invariant in the bulk but has a boundary interaction that breaks this invariance and induces a flow from Neumann boundary conditions at small $`\lambda `$ to Dirichlet boundary conditions at large $`\lambda `$. As discussed in , the boundary interaction is characterized by an energy scale, which one can represent as $`T_B=me^{\theta _B}`$. $`T_B`$ is related with the bare coupling in the action (39) by $`\lambda T_B^{1+g}`$. In the following, since obviously changes of $`m`$ (which is not a physical scale) can be absorbed in rapidity shifts, we set $`m=1`$. The effect of the boundary is then expressed by the reflection matrix :
$$R(\theta )=\mathrm{tanh}\left[\frac{\theta }{2}i\frac{\pi }{4}\right].$$
(40)
In the picture where imaginary time is along $`x`$, the effect of the boundary is represented by a boundary state. Following we can represent it in terms of the boundary scattering matrix :
$$|B>=\mathrm{exp}\left[_{\mathrm{}}^{\mathrm{}}\frac{d\theta }{2\pi }K(\theta _B\theta )Z_L^{}(\theta )Z_R^{}(\theta )\right]|0>.$$
(41)
In this formula, $`Z^{}`$ are again the Zamolodchikov Fateev creation operators, $`K`$ is related with the reflection matrix by :
$$K(\theta )=R\left(\frac{i\pi }{2}\theta \right)=\mathrm{tanh}\frac{\theta }{2}.$$
(42)
We do not prove the expression (41) directly here. It follows however from the compatibility between the present computation, and the one we will do next, in a modular transformed point of view where the direction of imaginary time will have been switched.
One can expand the boundary state into the convenient form :
$`|B>`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle _{0<\theta _1<\mathrm{}<\theta _n}}{\displaystyle \frac{d\theta _1}{2\pi }}\mathrm{}{\displaystyle \frac{d\theta _n}{2\pi }}K(\theta _B\theta _1)\mathrm{}K(\theta _B\theta _n)`$ (43)
$`\times `$ $`Z_L^{}(\theta _1)\mathrm{}Z_L^{}(\theta _n)Z_R^{}(\theta _1)\mathrm{}Z_R^{}(\theta _n)|0>.`$
Here, we have ordered the rapidities, but the result is the same as the unordered integral with the additional symmetry factor $`1/n!`$, the contributions from L and R scattering cancelling out after reordering. Observe now that by analyticity, the matrix elements of $`_z\mathrm{\Phi }`$ between the ground state and any state with at least one L moving particle are identically zero. More generally, the only non vanishing matrix elements of $`_z\mathrm{\Phi }`$ are those where bra and ket have the same L moving part. The same results apply by exchanging $`_z`$ with $`_{\overline{z}}`$ and L with R moving particles. As a result one gets immediately two of the four current correlators :
$`<_z\mathrm{\Phi }(z,\overline{z})_z^{}\mathrm{\Phi }(z^{},\overline{z}^{})>`$ $`={\displaystyle \frac{2g}{(zz^{})^2}}`$
$`<_{\overline{z}}\mathrm{\Phi }(z,\overline{z})_{\overline{z}^{}}\varphi (z^{},\overline{z}^{})>`$ $`={\displaystyle \frac{2g}{(\overline{z}\overline{z}^{})^2}},`$ (44)
which are identical with the ones without a boundary.
The two other correlators are more difficult to get. Let us consider for instance :
$$<0|_{\overline{z}}\mathrm{\Phi }(z,\overline{z})_z^{}\mathrm{\Phi }(z^{},\overline{z}^{})|B>.$$
(45)
The first non trivial contribution comes from the two particle term in the expansion of the boundary state :
$`{\displaystyle _{\mathrm{}}^{\mathrm{}}}`$ $`{\displaystyle \frac{d\theta }{2\pi }}K(\theta _B\theta )<0|_{\overline{z}}\mathrm{\Phi }(z,\overline{z})_z^{}\mathrm{\Phi }(z^{},\overline{z}^{})Z_L^{}(\theta )Z_R^{}(\theta )|0>=`$ (46)
$`\times `$ $`\mu ^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\theta }{2\pi }}K(\theta _B\theta )e^{2\theta }\mathrm{exp}\left[e^\theta (\overline{z}+z^{})\right].`$
More generally, because $`|B>`$ is a superposition of states with equal numbers of left and right moving particles, and $`_z\varphi `$, respectively $`_{\overline{z}}\varphi `$ act only on R, respectively L, particles, the expansion of (45) takes a very simple form :
$`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle }`$ $`{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n+1}}{(2\pi )^{2n+1}(2n+1)!}}K(\theta _B\theta _1)\mathrm{}K(\theta _B\theta _{2n+1})\left(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}}\right)^2`$ (47)
$`\mathrm{exp}\left[(\overline{z}+z^{})\left(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}}\right)\right]|f(\theta _1,\mathrm{},\theta _{2n+1})|^2.`$
This correlation function depends on the product $`e^{\theta _B}(\overline{z}+z^{})`$. It is scale invariant at the UV and IR fixed point. These correspond respectively to sending $`\theta _B`$ to $`\mathrm{}`$, that is the coupling $`\lambda `$ in the action to $`0`$ or $`\mathrm{}`$, in other words Neumann or Dirichlet boundary conditions. In the first case, $`K=1`$, in the second, $`K=1`$. Comparing with (36) and (37) we find, as expected, that :
$$<0|_{\overline{z}}\mathrm{\Phi }(z,\overline{z})_z^{}\mathrm{\Phi }(z^{},\overline{z}^{})|B>=\pm \frac{2g}{(\overline{z}+z^{})^2},$$
(48)
for Neumann, respectively Dirichlet boundary conditions. Although trivial, this result shows that the form factor expansion is well behaved, and allows us to study the correlator all the way from the UV to the IR fixed point when there is a boundary perturbation. In figures 1 and 2 we show the one particle (which is independent of $`B`$) and three particles contributions. We observe that indeed the convergence, by looking at the respective contributions, is very rapid.
The only drawback of this expansion is that it is not suited for studying the correlation of two operators right at the boundary. Indeed in that case, $`Re(\overline{z}+z^{})=0`$, and the integrals in (47) do not converge. To solve this problem, we can introduce a modular transformed picture. We now consider the imaginary time as running along the $`y`$ axis. Now the boundary is not represented as a state; rather, the whole space of states is different, since now we have only a half space to deal with. The asymptotic states are not pure L or R moving, but are mixtures. For instance, one particle states are :
$$||\theta >=|\theta >_R+R(\theta )|\theta >_L.$$
(49)
More generally, asymptotic states are obtained by adding to $`|\theta _1,\mathrm{},\theta _n>_{R,\mathrm{},R}`$ all combinations with different choices of $`R`$ particles transformed into $`L`$ particles, via action of the boundary. Only the following two terms contribute :
$$||\theta _1,\mathrm{},\theta _n>=|\theta _1,\mathrm{},\theta _n>_{R\mathrm{},R}+\mathrm{}+R(\theta _1)\mathrm{}R(\theta _n)|\theta _n,\mathrm{},\theta _1>_{L,\mathrm{},L}+\mathrm{}.$$
(50)
Although we used the same notation as previously, different things are meant by L,R. To make it clear, we now use the conventions :
$`<0|\mathrm{\Phi }(x,y)|\theta >_R=\mu \mathrm{exp}[me^\theta (y+ix)]`$
$`<0|\mathrm{\Phi }(x,y)|\theta >_L=\mu \mathrm{exp}[me^\theta (yix)].`$ (51)
To keep the notations as uniform as possible, we introduce the new coordinates :
$$w(z)iz=y+ix,$$
(52)
so here R movers depend on $`w`$, L movers on $`\overline{w}`$. The normalization $`\mu `$ is of course the same as before, and as before the LL and RR correlators do not depend on the boundary interaction. One finds :
$`<0|_w\mathrm{\Phi }(w,\overline{w})_w^{}\mathrm{\Phi }(w^{},\overline{w}^{})|0>`$ $`={\displaystyle \frac{2g}{(ww^{})^2}}`$
$`<0|_{\overline{w}}\mathrm{\Phi }(w,\overline{w})_{\overline{w}^{}}\mathrm{\Phi }(w^{},\overline{w}^{})|0>`$ $`={\displaystyle \frac{2g}{(\overline{w}\overline{w}^{})^2}},`$ (53)
where we used the fact that $`|R(\theta )|^2=1`$. When compared with (44), these correlators have an overall minus sign due to the dimension $`h=1,\overline{h}=0`$ (resp. $`h=0,\overline{h}=1`$) of the operators.
Let us now consider :
$$<0|_{\overline{w}}\mathrm{\Phi }(w,\overline{w})_w^{}\mathrm{\Phi }(w^{},\overline{w}^{})|0>.$$
(54)
To compute it, we insert a complete set of states which are of the form (49). In the massless case however, since $`_w\mathrm{\Phi }`$ is a R operator, $`_{\overline{w}^{}}\mathrm{\Phi }`$ a L operator, the only terms that contribute are in fact the ones with either all L or all R moving particles, as written in (50). Thus, (54) expands simply as :
$`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle }`$ $`{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n+1}}{(2\pi )^{2n+1}(2n+1)!}}R(\theta _1\theta _B)\mathrm{}R(\theta _{2n+1}\theta _B)\left(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}}\right)^2`$ (55)
$`\mathrm{exp}\left[(\overline{w}w^{})(e^{\theta _1}+\mathrm{}+e^{\theta _{2n+1}})\right]|f(\theta _1,\mathrm{},\theta _{2n+1})|^2.`$
Observe the crucial minus sign when compared to (47). It occurs because in one geometry the correlator depends on $`\overline{z}+z^{}`$, while in the other on $`\overline{w}w^{}`$. This now converges provided $`y>y^{}`$, even if $`x=x^{}=0`$ ie the operators are sitting right on the boundary. Now, using the fact that from factors depend only on differences of rapidities, this expression can be mapped with (47) if we formally set $`\theta =\theta ^{}+i\frac{\pi }{2}`$, provided one has :
$$K(\theta )=R\left(i\frac{\pi }{2}\theta \right),$$
(56)
as claimed above.
To summarize, we can write the left right current current correlator in two possible ways. By using the boundary state one finds :
$$<_{\overline{z}}\mathrm{\Phi }(x,y)_z^{}\mathrm{\Phi }(x^{},y^{})>=_0^{\mathrm{}}𝑑E𝒢(E)\mathrm{exp}\left[E(x+x^{})iE(yy^{})\right],$$
(57)
(recall that $`x,x^{}<0`$). One obtains $`𝒢(E)`$ simply by fixing the energy to a particular value in (47). When this is done, the remaining integrations occur on a finite domain for each of the individual particle energies since $`_{i=1}^{2n+1}e^{\theta _i}=E`$, and there is no problem of convergence anymore. One then gets :
$`𝒢(E)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^{\mathrm{ln}E}}{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n}}{(2\pi )^{2n+1}(2n+1)!}}{\displaystyle \frac{E^2}{Ee^{\theta _1}\mathrm{}e^{\theta _{2n}}}}`$ (58)
$`\times `$ $`K(\theta _B\theta _1)\mathrm{}K(\theta _B\theta _{2n})K\left[\theta _B\mathrm{ln}\left(Ee^{\theta _1}\mathrm{}e^{\theta _{2n}}\right)\right]`$
$`\times `$ $`\left|f[\theta _1\mathrm{}\theta _{2n},\mathrm{ln}\left(Ee^{\theta _1}\mathrm{}e^{\theta _{2n}}\right)]\right|^2,`$
with the constraint $`_{i=1}^{2n}e^{\theta _i}E`$. The denominator might suggest some possible divergences; it is important however to realize that it vanishes if and only if the particle with rapidty $`\theta _{2n+1}`$ has vanishing energy, in which case the form factor vanishes too. We can now shift the integrands to write equivalently :
$`𝒢(E)`$ $`=`$ $`E{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^0}{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n}}{(2\pi )^{2n+1}(2n+1)!}}{\displaystyle \frac{1}{1e^{\theta _1}\mathrm{}e^{\theta _{2n}}}}`$ (59)
$`\times `$ $`K(\mathrm{ln}(T_B/E)\theta _1)\mathrm{}K(\mathrm{ln}(T_B/E)\theta _{2n})K\left[\mathrm{ln}(T_B/E)\mathrm{ln}\left(1e^{\theta _1}\mathrm{}e^{\theta _{2n}}\right)\right]`$
$`\times `$ $`\left|f[\theta _1\mathrm{}\theta _{2n},\mathrm{ln}\left(1e^{\theta _1}\mathrm{}e^{\theta _{2n}}\right)]\right|^2,`$
where the constraint $`_{i=1}^{2n}e^{\theta _i}1`$ is implied, we used the fact that form-factors depend only on rapidity differences, and $`T_Be^{\theta _B}`$.
By using the dual picture, one finds, instead of (57) :
$$<_{\overline{z}}\mathrm{\Phi }(x,y)_z^{}\mathrm{\Phi }(x^{},y^{})>=_0^{\mathrm{}}𝑑E(E)\mathrm{exp}\left[iE(x+x^{})E(yy^{})\right],$$
(60)
where :
$`(E)`$ $`=`$ $`E{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^0}{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n}}{(2\pi )^{2n+1}(2n+1)!}}{\displaystyle \frac{1}{1e^{\theta _1}\mathrm{}e^{\theta _{2n}}}}`$ (61)
$`\times `$ $`R(\theta _1\mathrm{ln}(T_B/E))\mathrm{}R(\theta _{2n}\mathrm{ln}(T_B/E))R\left[\mathrm{ln}\left(1e^{\theta _1}\mathrm{}e^{\theta _{2n}}\right)\mathrm{ln}(T_B/E)\right]`$
$`\times `$ $`\left|f[\theta _1\mathrm{}\theta _{2n},\mathrm{ln}\left(1e^{\theta _1}\mathrm{}e^{\theta _{2n}}\right)]\right|^2,`$
where in (59) and (61) the constraint $`_{i=1}^{2n}e^{\theta _i}1`$ is implied. The two expressions are in correspondence by the simple analytic continuation :
$$𝒢(E)=i(iE).$$
(62)
## 3 The sine-Gordon Model.
In this section we follow the same line of thought for the sine-Gordon model. This is the massive deformation of the free boson which preserves integrability with either boundary interactions used in the fractional quantum Hall problem and the anisotropic Kondo model . Thus the form factors of the sine-Gordon model in the massless limit will be the quantities we need. The solitons/anti-solitons and breathers quasi-excitations make the problem more complicated but the results presented before hold with the addition of a few indices (and rather more complicated form factors).
We start as before with the massive sine-Gordon model, whose action reads :
$$S=\frac{1}{16\pi g}_{\mathrm{}}^{\mathrm{}}𝑑x𝑑y\left[\left(_x\mathrm{\Phi }\right)^2+\left(_y\mathrm{\Phi }\right)^2+\mathrm{\Lambda }\mathrm{cos}\mathrm{\Phi }\right].$$
(63)
Notice again that we have used a different normalization than in : this will avoid carrying factors of $`2\pi `$ all along the following paragraphs. Recall that $`g=\frac{\beta ^2}{8\pi }`$, $`\beta `$ the usual sine-Gordon parameter.
The form factors approach is formally the same, albeit more complicated because the particle content is much richer, and depends on $`g`$. For $`1/2<g<1`$, only solitons/anti-solitons appear in the spectrum of the theory. This is the so called repulsive case, with $`g=1/2`$ the free fermion point (giving rise, in quantum impurity problems, to so called “Toulouse limits”. When $`0<g<1/2`$, the particle content is enriched by $`[1/g2]`$ bound states, called breathers. In the following we will denote by the indices $`a=\pm `$ the solitons and anti-solitons, and $`a=1,2,\mathrm{},[1/g2]`$ the breathers. The solitons form factors in the massive case were written by Smirnov and we obtain the massless form factors by taking the appropriate limit of the massive ones. Only right and left moving form factors survive in this limit, as in the sinh-Gordon case. Moreover, the symmetry of the action dictates that only form factors with total topological charge zero are non-zero for the current operator. As an example, the soliton/anti-soliton form factor is given by :
$`<0|{\displaystyle \frac{1}{2\pi }}_z\mathrm{\Phi }(z,\overline{z})|\theta _1,\theta _2>_{aa^{}}^{RR}`$ $`=`$ $`a^{}\mu mde^{(\theta _1+\theta _2)/2}`$ (64)
$`\times `$ $`{\displaystyle \frac{\zeta (\theta _1\theta _2)}{\mathrm{cosh}\frac{(1g)}{2g}(\theta _1\theta _2+i\pi )}}\mathrm{exp}\left[m(e^{\theta _1}+e^{\theta _2})z\right],`$
with $`a+a^{}=0`$ and $`a=\pm `$ stands for soliton (resp. antisoliton). From one has :
$$\zeta (\theta )=c\mathrm{sinh}\frac{\theta }{2}\mathrm{exp}\left(_0^{\mathrm{}}\frac{\mathrm{sin}^2\frac{x(i\pi \theta )}{2\pi }\mathrm{sinh}\frac{(12g)x}{2(1g)}}{x\mathrm{sinh}\frac{gx}{2(1g)}\mathrm{sinh}x\mathrm{cosh}\frac{x}{2}}𝑑x\right),$$
(65)
(this is essentially the minimum form-factor discussed in the introduction) with the constant $`c`$ given by :
$$c=\left(\frac{4(1g)}{g}\right)^{1/4}\mathrm{exp}\left(\frac{1}{4}_0^{\mathrm{}}\frac{\mathrm{sinh}\frac{x}{2}\mathrm{sinh}\frac{(12g)x}{2(1g)}}{x\mathrm{sinh}\frac{gx}{2(1g)}\mathrm{cosh}^2\frac{x}{2}}𝑑x\right),$$
(66)
and $`d`$ by :
$$d=\frac{1}{2\pi c}\frac{(1g)}{g},$$
(67)
The normalization constant $`\mu `$ can be determined from first principles. Indeed, the operator $`_x\mathrm{\Phi }`$ being related with the $`U(1)`$ charge, we need that
$${}_{R}{}^{+}<\theta _1|_{\mathrm{}}^{\mathrm{}}\frac{1}{2\pi }_x\mathrm{\Phi }|\theta _2>_+^R=2\pi \delta (\theta _1\theta _2),$$
(68)
using the fact that a soliton for the bulk theory (63) obeys $`\mathrm{\Phi }(\mathrm{})\mathrm{\Phi }(\mathrm{})=2\pi `$. On the other hand, using the dependence of the form-factor on spatial coordinates, the left hand side is
$${}_{R}{}^{+}<\theta _1|\frac{1}{2\pi }_x\mathrm{\Phi }|\theta _2>_+^Rdxe^{im(e^{\theta _1}e^{\theta _2})x}=\frac{2\pi }{me^\theta }_R^+<\theta _1|\frac{1}{2\pi }_x\mathrm{\Phi }|\theta _1>_+^R\delta (\theta _1\theta _2),$$
(69)
Comparing the two and using crossing leads to the identity
$$i\mu d\zeta (i\pi )=1$$
(70)
or
$$\mu =2\pi \frac{g}{1g}=\frac{1}{cd}.$$
(71)
The other soliton-antisoliton form factors follow from the sophisticated analysis of . Their expression simplifies in the case $`g=\frac{1}{t}`$, $`t`$ an integer. This is the physically relevant case for the $`\nu =\frac{1}{t}`$ fractional quantum Hall effect. One then finds :
$`<0|{\displaystyle \frac{1}{2\pi }}_z\mathrm{\Phi }(z,\overline{z})|\theta _1,\mathrm{},\theta _{2n}>_{R\mathrm{}R}^{\mathrm{},+\mathrm{}+}=\mu m(2d)^ne^{(\theta _1+\mathrm{}+\theta _{2n})/2}{\displaystyle \underset{i<j}{}}\zeta (\theta _i\theta _j)`$
$`\mathrm{sinh}\left[{\displaystyle \frac{t1}{2}}{\displaystyle \underset{p=1}{\overset{n}{}}}(\theta _{p+n}\theta _pi\pi )\right]{\displaystyle \underset{p=1}{\overset{n}{}}}{\displaystyle \underset{q=n+1}{\overset{2n}{}}}\mathrm{sinh}^1(t1)(\theta _q\theta _p)detH.`$ (72)
The matrix $`H`$ is obtained as follows. First introduce the function :
$$\psi (\alpha )=2^{t2}\underset{j=1}{\overset{t2}{}}\mathrm{sinh}\frac{1}{2}\left(\alpha i\frac{\pi j}{t1}+i\frac{\pi }{4}\right).$$
(73)
One then defines the matrix elements as :
$$H_{ij}=\frac{1}{2i\pi }_{2i\pi }^0𝑑\alpha \underset{k=1}{\overset{k=2n}{}}\psi (\alpha \beta _k)\mathrm{exp}\left[(n2j1)\alpha +(n2i)(t1)\alpha \right],$$
(74)
where $`i,j`$ run over $`1,\mathrm{},n1`$. It is not difficult to convince oneself that this produce a symmetric polynomial of the right degree. Although cumbersome, it is an easy task to extract these determinants, as examples we find for $`g=1/3`$ :
$`detH=\mathrm{exp}\left({\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{2n}{}}}\theta _i\right)\sigma _1(e^{\theta _p}),n=2,`$
$`detH=\mathrm{exp}({\displaystyle \underset{i=1}{\overset{2n}{}}}\theta _i)\sigma _1(e^{\theta _p})\sigma _3(e^{\theta _p}),n=3,`$ (75)
up to irrelevant phases and with the $`\sigma _q`$’s defined previously. Having these expression we can get all form factors using the axiomatics sketched in the introduction. For example, the solitons form factors with different positions of the indices $`a_i`$ we use the symmetry property (16):
$`f(\theta _1,\mathrm{},\theta _i,\theta _{i+1},\mathrm{},\theta _n)_{a_1,\mathrm{},a_i,a_{i+1},\mathrm{},a_n}=`$
$`f(\theta _1,\mathrm{},\theta _{i+1},\theta _i,\mathrm{},\theta _n)_{a_1,\mathrm{},b_{i+1},b_i,\mathrm{},a_n}S_{a_i,a_{i+1}}^{b_i,b_{i+1}}(\theta _i\theta _{i+1}).`$ (76)
Here again, we omit the distinction between left and right moving form factor, they are simply related by complex conjugation. At the points $`g=1/t`$ the soliton $`S`$ matrix used in the last expression is reflectionless and basically just permutes the rapidities up to a phase. When there are breathers, the soliton $`S`$ matrix has poles corresponding to the bound states at the points $`\theta =iu_{ab}^c=i\pi \frac{i\pi g}{(1g)}m`$ for the $`m`$’th breather. In view of the last relation, this induces poles in the form factors. We obtain the breather form factors from these poles :
$`i\mathrm{res}_{\theta _{n1}\theta _n=iu_{ab}^c}f(\theta _1,\mathrm{},\theta _{n1},\theta _n)_{a_1,\mathrm{},a_{n1},a_n}`$ $`=`$ $`r_m(1)^{\frac{a_n+1}{2}m}\delta _{a_{n1}+a_n}`$ (77)
$`\times `$ $`f(\theta _1,\mathrm{},\theta _{n1}+{\displaystyle \frac{i\pi }{2}}i{\displaystyle \frac{i\pi gm}{2(1g)}})_{a_1,\mathrm{},a_{n2},m},`$
and $`r_m`$ is given by the residue at $`iu_{ab}^c=i\pi \frac{i\pi g}{(1g)}m`$ :
$$r_m=\left[S_{++}^{++}\left(i\pi \frac{i\pi mg}{1g}\right)\frac{g}{(1g)}\mathrm{sin}\pi \frac{1g}{g}\right]^{1/2}.$$
(78)
Having these relations, we posess all ingredients to compute all form factors for $`g=1/t`$. Using them for the computation of the current correlations is then merely an extension of the previous results for sinh-Gordon with indices. The normalisation of the form factors, $`\mu `$, should also ensure that (44) is reproduced. This is fixed by introducing a complete basis of states :
$$1=\underset{n=0}{\overset{\mathrm{}}{}}\underset{a_i}{}\frac{d\theta _1\mathrm{}d\theta _n}{(2\pi )^nn!}|\theta _1,\mathrm{},\theta _n>_{a_1,\mathrm{},a_n}{}_{}{}^{a_n,\mathrm{},a_1}<\theta _n,\mathrm{},\theta _1|$$
(79)
and computing the correlations exactly like in the sinh-Gordon case.
Keeping a finite number of form-factors and demanding that the two point function is properly normalized gives rise to values of $`\mu `$ which are slightly different from (71). How different is a good measure of the convergence of the expansion, and the validity of the truncation. For $`g=1/3`$, the one breather and 2 solitons form factors normalise to $`\mu =3.14`$ which is very close to the exact $`\pi `$. Similarly for $`g=1/4`$ we found from the contributions up to two solitons that $`\mu =2.05`$ to compare with $`2.094=2\pi /3`$.
Calculations in the presence of an integrable boundary interaction are also done like in the sinh-Gordon case. The boundary state is now given by :
$`|B>={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}`$ $`{\displaystyle _{0<\theta 1<\mathrm{}<\theta _n}}{\displaystyle \frac{d\theta _1}{2\pi }}\mathrm{}{\displaystyle \frac{d\theta _n}{2\pi }}K^{a_1b_1}(\theta _B\theta _1)\mathrm{}K^{a_nb_n}(\theta _B\theta _n)`$ (80)
$`\times `$ $`Z_L^{a_1}(\theta _1)\mathrm{}Z_L^{a_n}(\theta _n)Z_R^{b_1}(\theta _1)\mathrm{}Z_R^{b_n}(\theta _n),`$
with an implicit sum on the indices. The matrix $`K^{ab}`$ is related to the boundary $`R`$ matrix in the following way :
$$K^{ab}(\theta )=R_{\overline{b}}^a\left(i\frac{\pi }{2}\theta \right).$$
(81)
The $`\overline{b}`$ means that we take the conjugate of the indices ie. $`\pm `$ and $`mm`$.
From the previous expressions, we can compute de current-current correlation function in the presence of a boundary for $`g=1/t`$. The results we will get depends on the boundary interaction, in the next subsection we present some results for the boundary sine-Gordon model, which is of relevance to tunneling experiments in fractional quantum Hall devices .
## 4 Conductance in the fractional quantum Hall effect
### 4.1 General remarks.
The boundary sine-Gordon action is
$$S=\frac{1}{16\pi g}_{\mathrm{}}^0𝑑x_{\mathrm{}}^{\mathrm{}}𝑑y\left[\left(_x\mathrm{\Phi }\right)^2+\left(_y\mathrm{\Phi }\right)^2\right]+\lambda _{\mathrm{}}^{\mathrm{}}𝑑y\mathrm{cos}\frac{1}{2}\mathrm{\Phi }(x=0,y).$$
(82)
The reflection matrices have been worked out in . For generic values of the coupling $`g`$, the amplitude for the processes $`++`$ and $``$ is $`R_\pm ^\pm (\theta \theta _B)`$, and for the processes $`+`$ and $`+`$ it is $`R_{}^\pm (\theta \theta _B)`$ with (the notation is slightly changed with respect to , where outgoing indices were not raised):
$`R_{}^\pm (\theta )`$ $`=e^{\frac{(1g)\theta }{2g}}R(\theta )`$
$`R_\pm ^\pm (\theta )`$ $`=ie^{\frac{(g1)\theta }{2g}}R(\theta )`$ (83)
where the function $`R`$ reads :
$`R(\theta )`$ $`={\displaystyle \frac{e^{i\gamma }}{2\mathrm{cosh}\left[\frac{(1g)\theta }{2g}i\frac{\pi }{4}\right]}}{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{Y_l(\theta )}{Y_l(\theta )}}`$
$`Y_l(\theta )`$ $`={\displaystyle \frac{\mathrm{\Gamma }\left(\frac{3}{4}+l\frac{(1g)}{g}\frac{i(1g)\theta }{2\pi g}\right)\mathrm{\Gamma }\left(\frac{1}{4}+(l+1)\frac{(1g)}{g}\frac{i(1g)\theta }{2\pi g}\right)}{\mathrm{\Gamma }\left(\frac{1}{4}+(l+1/2)\frac{(1g)}{g}\frac{(1g)i\theta }{2\pi g}\right)\mathrm{\Gamma }\left(\frac{3}{4}+(l+1/2)\frac{(1g)}{g}\frac{i(1g)\theta }{2\pi g}\right)}}.`$ (84)
In (83), our conventions are such that in the UV limit ($`\theta _B\mathrm{}`$) the scattering is totally off-diagonal so a soliton bounces back as an anti-soliton, in agreement with classical limit results for Neumann boundary conditions. A useful integral representation of $`R`$ is given by :
$$R(\theta )=\frac{e^{i\gamma }}{2\mathrm{cosh}\left[\frac{(1g)\theta }{2g}i\frac{\pi }{4}\right]}\mathrm{exp}\left(i_{\mathrm{}}^{\mathrm{}}\frac{dy}{2y}\mathrm{sin}\frac{2(1g)\theta y}{g\pi }\frac{\mathrm{sinh}(\frac{12g}{g})y}{\mathrm{sinh}2y\mathrm{cosh}\frac{(1g)y}{g}}\right).$$
(85)
Recall that the spectrum is made of one breather and the pair soliton antisoliton in the whole domain $`1/3g<1/2`$. More breathers appear for $`g<1/3`$ (and the reflection matrix of the 1- breather is always the same as in the sinh-Gordon case.) There are no breathers for $`g>1/2`$.
The physical quantity of interest in this case corresponds to the AC conductance at vanishing temperature in the edge states tunneling problem: this problem has been described in details in , to which we refer the reader.
A standard way of representing the conductance is through the Kubo formula :
$$G(\omega _M)=\frac{1}{8\pi \omega _ML^2}_L^L𝑑x𝑑x^{}_{\mathrm{}}^{\mathrm{}}𝑑ye^{i\omega _My}<j(x,y)j(x^{},0)>,$$
(86)
where $`\omega _M`$ is a Matsubara frequency, $`y`$ is imaginary time, $`y=it`$. One gets back to real physical frequencies by letting $`\omega _M=i\omega `$. In (86) , $`j`$ is the physical current in the unfolded system . Without impurity, the AC conductance of the Luttinger liquid is frequency independent, $`G=g`$. When adding the impurity, it becomes $`G=\frac{g}{2}+\mathrm{\Delta }G`$. After some simple manipulations using the folding , one finds :
$`\mathrm{\Delta }G(\omega _M)`$ $`=`$ $`{\displaystyle \frac{1}{8\pi \omega _ML^2}}{\displaystyle _L^0}𝑑x𝑑x^{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑ye^{i\omega _My}`$ (87)
$`\times `$ $`\left[<_z\mathrm{\Phi }(x,y)_{\overline{z}^{}}\varphi (x^{},0)>+<_{\overline{z}}\mathrm{\Phi }(x,y)_z^{}\mathrm{\Phi }(x^{},0)>\right],`$
where $`z=x+iy`$. The strategy is simply to evaluate the current-current correlator using form-factors, and extract the conductance from (87).
We will do so for special values of $`g`$, but first, we can extract some general features of the UV and IR expansions easily. To do so, consider the soliton antisolitons reflection matrix. Evaluating the integral in 85 by the residues method leads to a double expansion of the elements $`R_\pm ^\pm `$ in powers of $`\mathrm{exp}(\theta )`$ and $`\mathrm{exp}(\frac{1}{g}1)\theta `$ . This leads for the conductance to a double power series in $`(\omega /T_B)^{2+2/g}`$ and $`(\omega /T_B)^2`$ in the IR, $`(T_B/\omega )^{22g}`$ and $`(T_B/\omega )^2`$ in the UV. Breathers do not change this result. For instance for the 1-breather, since the reflection matrix is the same as in the sinh-Gordon case, and therefore $`g`$ independent, the contributions expand as a series in $`(\omega /T_B)^2`$ in the IR, $`(T_B/\omega )^2`$ in the UV. This holds for any coupling $`g`$. Therefore, as first argued by Guinea et al. , at low frequency, the conductance goes as $`\omega ^2`$ for $`g<1/2`$, $`\omega ^{2+2/g}`$ for $`g>1/2`$. The $`\omega ^2`$ power would seem to indicate that there should be a $`T^2`$ term in the DC conductance, but this is not correct because only the modulus square of $`R_\pm ^\pm `$ contribute to the DC conductance, and this expands only as powers of $`\mathrm{exp}(\frac{1}{g}1)\beta `$ .
The presence of analytic terms in the IR is a straightforward consequence of the fact that IR perturbation theory involves an infinity of counter-terms, in particular polynomials in derivatives of $`\mathrm{\Phi }`$ . More surprising maybe is the fact that we find analytical terms in the UV. This requires some discussion. The UV terms follow from the short distance behaviour of the correlation function of the current. For any operator $`O`$ we could write formally,
$$<O(x^{},y^{})O(x,y)>_\lambda =\underset{n=0}{\overset{\mathrm{}}{}}(\lambda )^{2n}d1\mathrm{}𝑑n<O(x^{},y^{})O(x,y)\mathrm{cos}\frac{1}{2}\mathrm{\Phi }(1)\mathrm{}\mathrm{cos}\frac{1}{2}\mathrm{\Phi }(n)>_{\lambda =0}.$$
(88)
From (88), one would naively expect that the two point function of the current expands as a power series in $`\lambda ^2`$, which would lead to a power series in $`(\omega /T_B)^{2g2}`$. This is incorrect however because, even if integrals are convergent at short distance for $`g<1/2`$, they are always divergent at large distances. It is known that these IR divergences give precisely rise to non analyticity in the coupling constant $`\lambda `$. One usually writes :
$$<O(x^{},y^{})O(x,y)>=\underset{i}{}C_{OO}^i(x^{}x,y^{}y)O_i(x,y),$$
(89)
where $`O_i`$ are a complete set of local operators in the theory and the $`C`$’s are structure functions. These, being local quantities, have analytic behaviour in $`\lambda `$. However, $`<O_i(x,y)>`$ being non local is in general non analytic - actually, on dimensional grounds,
$$<O_i(x,y)>\lambda ^{\mathrm{\Delta }/(1g)}T_B^\mathrm{\Delta },$$
(90)
where $`\mathrm{\Delta }=h+\overline{h}`$ is the (bulk) dimension of the field $`O_i`$. If we computed the conductance perturbatively using Matsubara formula, we would use (88) with $`O`$ the electrical current operator. The case $`O_i`$ the identity operator gives rise to an analytical expression in $`\lambda `$, but eg the case $`O_i=_z_{\overline{z}}\mathrm{\Phi }`$ gives $`\lambda ^{2/(1g)}`$ times an analytical expression in $`\lambda `$ (its mean value can be non zero because there is a boundary). More generally, since the only operators $`O_i`$ appearing in the case of the current are polynomials in derivatives of $`\mathrm{\Phi }`$, all with integer dimensions, we expect that the two point function of the current will expand as a double series of the form $`\lambda ^{2n}\lambda ^{2m/(1g)}`$, ie going back to $`T_B`$ variable , that the conductance will expand as a double series of the form $`(T_B/\omega )^{2n(1g)}(T_B/\omega )^{2m}`$, in agreement with the form factors result.
### 4.2 The free case.
In the case $`g=1/2`$ one has simply :
$`R_{}^\pm (\theta )=P(\theta )`$ $`={\displaystyle \frac{e^\theta }{e^\theta +i}}`$
$`R_\pm ^\pm (\theta )=Q(\theta )`$ $`={\displaystyle \frac{i}{e^\theta +i}}.`$ (91)
In that case, only the soliton-antisoliton form factor is non zero, $`f(\theta _1,\theta _2)=i\mu e^{\theta _1/2}e^{\theta _2/2}`$, with the normalization $`\mu =2\pi `$ and we have set $`m=1`$. Hence, $`(\omega )`$ is readily evaluated
$$(\omega )=_{\mathrm{}}^{\mathrm{}}_{\mathrm{}}^{\mathrm{}}𝑑\theta _1𝑑\theta _2\delta (e^{\theta _1}+e^{\theta _2}\omega )[Q(\theta _1)Q(\theta _2)P(\theta _1)P(\theta _2)]e^{\theta _1}e^{\theta _2},$$
(92)
so
$$(\omega )=\omega _0^1𝑑x\frac{x(1x)+1}{\left(x+i\frac{T_B}{\omega }\right)\left(\omega x+i\frac{T_B}{\omega }\right)},$$
(93)
from which it follows that
$$\mathrm{\Delta }G(\omega )=\frac{1}{4}\frac{T_B}{2\omega }\mathrm{tan}^1(\omega /T_B).$$
(94)
Thus we find
$$G(\omega )=\frac{1}{2}\left(1\frac{T_B}{\omega }\mathrm{tan}^1(\omega /T_B)\right).$$
(95)
This is in agreement with the solution of .
### 4.3 $`G(\omega )`$ at $`g=1/3`$.
The conductance for $`g=1/3`$ has a direct application to the quantum Hall effect. Comparing with the free case, previously treated, we now have a breather in the spectrum and non zero form factors for all number of rapidities. Still the convergence is such that evaluating the first few form factors give results to a very good accuracy, independently of the regime, UV or IR, in which we make the computation.
In this case, the first few non zero form factors are $`f_1`$, $`f_{\pm ,}`$, $`f_{\pm ,,1}`$, $`f_{1,1,1}`$, etc… Here the subscript “1” denotes the breather. The first step is to compute the normalisation in order to satisfy (31). When computing this normalisation, we find that the first two form factors account for the whole result to more than one percent accuracy. Then including the 1 breather-2 solitons form factor is sufficient to get the result to a very good accuracy ($`f_{1,1,1}`$ is negligeable). Actually one observes that the speed of convergence of the form factor expansion varies geometrically with the number of solitons (counting the breathers as 2 solitons).
In order to get the conductance we need the reflection matrices, they were given in previous expressions and reduce to a simpler form for this value of $`g`$ :
$$R(\theta )=\frac{1}{2\mathrm{cosh}(\theta \frac{i\pi }{4})}\frac{\mathrm{\Gamma }(3/8\frac{i\theta }{2\pi })\mathrm{\Gamma }(5/8+\frac{i\theta }{2\pi })}{\mathrm{\Gamma }(5/8\frac{i\theta }{2\pi })\mathrm{\Gamma }(3/8+\frac{i\theta }{2\pi })}$$
(96)
and the breather reflection matrix is :
$$R_1^1(\theta )=\mathrm{tanh}\left(\frac{\theta }{2}\frac{i\pi }{4}\right).$$
(97)
From the pole of the 2 solitons form factor, the one breather form factor is found using (77) and its contribution to the conductance is :
$$\mathrm{\Delta }G(\omega )^{(1)}=\mu ^2\frac{\pi d^2}{8}e\mathrm{tanh}\left[\frac{\mathrm{log}(\frac{\omega }{\sqrt{2}T_B})}{2}i\pi /4\right],$$
(98)
here $`\mu =\pi `$ is fixed by (71) and $`d=0.1414\mathrm{}`$. The contribution from the two solitons form factors is computed similarly, we find :
$`\mathrm{\Delta }G(\omega )^{(2)}=`$ $`{\displaystyle \frac{\mu ^2d^2}{2}}e{\displaystyle _{\mathrm{}}^0}𝑑\theta {\displaystyle \frac{R(\theta +\mathrm{log}(\frac{\omega }{T_B}))R(\mathrm{log}[(1e^\theta )\frac{\omega }{T_B}]}{\mathrm{cosh}^2(\theta \mathrm{log}(1e^\theta ))}}|\zeta [\theta \mathrm{log}(1e^\theta )]|^2`$ (99)
$`e^\theta \left[e^\theta (1e^\theta )({\displaystyle \frac{\omega }{T_B}})^2+{\displaystyle \frac{1}{e^\theta (1e^\theta )(\frac{\omega }{T_B})^2}}\right],`$
where $`\zeta (\theta )`$ is the function defined in (64). We can similarly write the following contribution, and we find that these last two expressions are sufficient for any reasonable purpose, they give the frequency dependent conductance to more than one percent accuracy. We give the full function $`G(\omega )`$ in figure 3.
Observe that in the UV and in the IR we obtain the $`\omega `$ dependance discussed previously, even with the truncation to a few form-factors. The form-factors expansion is indeed very different from the perturbative expansion in powers of the coupling constant in the UV, or in powers of the inverse coupling constant in the IR. Each form-factor contribution has by itself the same analytical structure as the whole sum; contributions with higher number of particles simply determine coefficients to a greater accuracy.
## 5 Anisotropic Kondo model and dissipative quantum mechanics.
It turns out that the anisotropic Kondo model is related with the famous two state problem of dissipative quantum mechanics . The bosonised form of the hamiltonian is :
$$H=\frac{1}{2}_{\mathrm{}}^0𝑑x\left[8\pi g\mathrm{\Pi }^2+\frac{1}{8\pi g}(_x\mathrm{\Phi })^2\right]+\frac{\lambda }{2}\left(S_+e^{i\mathrm{\Phi }(0)/2}+S_{}e^{i\mathrm{\Phi }(0)/2}\right),$$
(100)
where $`S`$ are Pauli matrices. As for the quantum Hall problem, we keep using as a basis the massless excitations of the sine-Gordon model; however the boundary interation is different: this will result in different reflection matrices.
### 5.1 A computation of $`C(t)`$.
We first work in imaginary time. We consider therefore the anisotropic Kondo problem at temperature $`T`$. Let us consider the quantity $`X(y)<[S^z(y)S^z(0)]^2>`$. On the one hand, using that $`S^z=\pm 1`$, it reads $`2[1C(y)]`$, where $`C(y)`$ is the usual spin correlation
$$C(y)=\frac{1}{2}\left[<S^z(y)S^z(0)>+<S^z(0)S^z(y)>\right].$$
(101)
On the other hand, we can write a perturbative expansion for $`X(y)`$ by expanding evolution operators in powers of the coupling constant $`\lambda `$. At every order, we get ordered monomials which are are a product of a monomial in $`S^\pm `$ and vertex operators of charge $`\pm 1/2`$. We must then evaluate $`S^z(y)S^z(0)`$ for each such term, trace over the two possible spin states, and average over the quantum field. Since we deal with spin $`1/2`$, terms $`S^+`$ and $`S^{}`$ must alternate, and there must be an overall equal number of $`S^+`$ and $`S^{}`$, and an equal number of $`1/2`$ and $`1/2`$ electric charges.
Now, since each $`S^+(y)`$ comes with a $`e^{i\mathrm{\Phi }(y)/2}`$ and each $`S^{}(y)`$ comes with a $`e^{i\mathrm{\Phi }(y)/2}`$, $`S^z(y)=S^z(0)`$ if there is a vanishing electric charge inserted between $`0`$ and $`y`$, and $`S^z(y)=S^z(0)`$ if the charge inserted between $`0`$ and $`y`$ is non zero (and then it has to be $`\pm 1/2`$). Therefore, we can write the perturbation expansion of $`X(y)`$ in such a way that the spin contributions all disappear:
$`X(y)`$ $`=`$ $`{\displaystyle \frac{1}{Z}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\lambda ^{2n}{\displaystyle \underset{alternatingϵ_i=\pm }{}}{\displaystyle \underset{p=0}{\overset{2n}{}}}{\displaystyle _0^y}𝑑y_1{\displaystyle _{y_1}^y}𝑑y_2\mathrm{}{\displaystyle _{y_{p1}}^y}𝑑y_p{\displaystyle _y^{1/T}}𝑑y_{p+1}\mathrm{}{\displaystyle _{y_{2n1}}^{1/T}}𝑑y_{2n}`$ (102)
$`\times `$ $`4(ϵ_1+\mathrm{}+ϵ_p)^2e^{iϵ_1\varphi (y_1)/2}\mathrm{}e^{iϵ_{2n}\varphi (y_{2n})/2}_N,`$
where $`Z`$ is the partition function, the factor 4 occurs because of the normalization $`S^z=\pm 1`$, for every configuration of $`ϵ`$’s, only one value of $`S^z(0)`$ gives a non vanishing contribution. Here, the label $`N`$ indicates correlation functions for the free boson evaluated with Neumann boundary conditions (the conditions as $`\lambda 0`$).
On the other hand, let us consider the correlator
$$<_x\mathrm{\Phi }(x,y)\mathrm{\Phi }(0,y^{})>_N=8g\frac{x}{x^2+(yy^{})^2},$$
(103)
which goes to $`8g\pi \delta (yy^{})`$ as $`x0`$. We have then, by Wick’s theorem,
$`<e^{iϵ_1\mathrm{\Phi }(y_1)/2}\mathrm{}e^{iϵ_{2n}\mathrm{\Phi }(y_{2n})/2}_x\mathrm{\Phi }(x,y)>_N=`$
$`8ig\left({\displaystyle \underset{i=1}{\overset{2n}{}}}ϵ_i{\displaystyle \frac{x}{x^2+(yy_i)^2}}\right)e^{iϵ_1\mathrm{\Phi }(y_1)/2}\mathrm{}e^{iϵ_{2n}\mathrm{\Phi }(y_{2n}/2)}_N,`$ (104)
and therefore
$`e^{iϵ_1\mathrm{\Phi }(y_1)/2}\mathrm{}e^{iϵ_{2n}\mathrm{\Phi }(y_{2n})/2}:_x\mathrm{\Phi }(x,y)_x\mathrm{\Phi }(x,y^{}):_N=`$
$`(8g)^2\left({\displaystyle \underset{i=1}{\overset{2n}{}}}ϵ_i{\displaystyle \frac{x}{x^2+(yy_i)^2}}\right)\left({\displaystyle \underset{i=1}{\overset{2n}{}}}ϵ_i{\displaystyle \frac{x}{x^2+(y^{}y_i)^2}}\right)e^{iϵ_1\mathrm{\Phi }(y_1)/2}\mathrm{}e^{iϵ_{2n}\mathrm{\Phi }(y_{2n})/2}_N,`$ (105)
where contractions between the dots $`:`$ are discarded. In (105), contractions between the dots would lead to a term factored out as the product of the two point function of $`_x\mathrm{\Phi }`$ and the $`2n`$ point function of vertex operators, both evaluated with N boundary conditions. Now, we are going to be interested in the $`x0`$ limit where, with N boundary conditions, $`_x\mathrm{\Phi }`$ vanishes. As a result we can actually forget the subtraction in 105, and write simply obtain
$$X(y,\lambda )=\frac{1}{(4g\pi )^2}\underset{x0}{lim}_0^y_0^y𝑑y^{}𝑑y^{\prime \prime }<_x\mathrm{\Phi }(x,y^{})_x\mathrm{\Phi }(x,y^{\prime \prime })>_\lambda ,$$
(106)
where the label $`\lambda `$ designates the correlator evaluated at coupling $`\lambda `$, N corresponding to $`\lambda =0`$ . Hence, we can get $`C(y)`$ from the current current correlator. The latter can then be obtained using form factors along the above lines. The only difference is the boundary matrix. If we restrict to the repulsive regime where the bulk spectrum contains only a soliton and an antisoliton, one has :
$$R_\pm ^{}=\mathrm{tanh}\left(\frac{\theta }{2}\frac{i\pi }{4}\right),R_\pm ^\pm =0.$$
(107)
Here again our conventions are such that a soliton bounces back as an antisoliton, in agreement with the UV and the IR limit that have Neumann boundary conditions. In the attractive regime we need to add the breathers with :
$$R_m^m=\frac{\mathrm{tanh}(\frac{\theta }{2}\frac{i\pi m}{4(1/g1)})}{\mathrm{tanh}(\frac{\theta }{2}+\frac{i\pi m}{4(1/g1)})}.$$
(108)
Writing, as in (57) :
$$<_{\overline{z}}\mathrm{\Phi }(x,y^{})_z\mathrm{\Phi }(x,y^{\prime \prime })>_\lambda =_0^{\mathrm{}}𝒢(E,\beta _B)\mathrm{exp}\left[2ExiE(y^{}y^{\prime \prime })\right],$$
(109)
we have that :
$`\underset{x0}{lim}<_x\mathrm{\Phi }(x,y^{})_x\mathrm{\Phi }(x,y^{\prime \prime })>_\lambda =`$
$`{\displaystyle _0^{\mathrm{}}}𝑑E\left[𝒢(E,\beta _B)𝒢(E,\mathrm{})\right]\mathrm{exp}[iE(y^{}y^{\prime \prime })]+c.c.,`$ (110)
where the $`<_z\mathrm{\Phi }_z\mathrm{\Phi }>`$ part and its complex conjugate (which are $`\lambda `$ independent) have been evaluated by requiring that the correlator vanishes as $`\lambda 0`$ due to $`N`$ boundary conditions. Hence, using the fact that $`𝒢`$ is real,
$$X(y)=\frac{1}{2(g\pi )^2}_0^{\mathrm{}}\frac{dE}{E^2}\left[𝒢(E,\theta _B)𝒢(E,\mathrm{})\right]\mathrm{sin}^2(Ey/2).$$
(111)
Therefore, if we write :
$$C(y)1=_0^{\mathrm{}}A(\omega _M)\mathrm{cos}(\omega _My)𝑑\omega _M,$$
(112)
where $`\omega _M`$ is a Matsubara frequency, we have :
$$A(\omega _M)=\frac{1}{(2g\pi )^2}\frac{1}{\omega _M^2}\left[𝒢(\omega _M,\theta _B)𝒢(\omega _M,\mathrm{})\right].$$
(113)
An observation is now in order. From the foregoing results we see that
$$<S^z(0)S^z(y)>1=\frac{1}{(2g\pi )^2}_0^{\mathrm{}}\frac{dE}{E^2}\left[𝒢(E,\theta _B)𝒢(E,\mathrm{})\right]\mathrm{cos}Ey.$$
(114)
On the other hand, consider the expression
$$<_{\mathrm{}}^0𝑑x^{}_{\mathrm{}}^0𝑑x^{\prime \prime }\left[<_x\mathrm{\Phi }(x^{},y)_x\mathrm{\Phi }(x^{\prime \prime },0)>_\lambda <_x\mathrm{\Phi }(x^{},y)_x\mathrm{\Phi }(x^{\prime \prime },0)>_N\right].$$
(115)
By using the same representation (109), this is
$$_{\mathrm{}}^0𝑑x^{}_{\mathrm{}}^0𝑑x^{\prime \prime }_0^{\mathrm{}}𝑑E\left[𝒢(E,\theta _B)𝒢(E,\mathrm{})\right]\mathrm{exp}\left[E(x^{}+x^{\prime \prime })iEy\right]+cc,$$
which coincides with (114) after performing the integrations. We conclude that
$$<S^z(0)S^z(y)>1=<𝒥_x(0)𝒥_x(y)><𝒥_x(0)𝒥_x(y)>_N,$$
(116)
where we defined
$$𝒥_x=\frac{1}{2g\pi }_{\mathrm{}}^0_x\mathrm{\Phi }(x,y)dx.$$
(117)
We find also by the same manipulations that
$$<S^z(0)S^z(y)>1=<𝒥_y(0)𝒥_y(y)><𝒥_y(0)𝒥_y(y)>_N,$$
(118)
where
$$𝒥_y=\frac{1}{2g\pi }_{\mathrm{}}^0_y\varphi (x,y)dx.$$
(119)
We now continue to real frequencies to find the response function :
$$\chi ^{\prime \prime }(\omega )\frac{1}{2}𝑑te^{i\omega t}[S^z(t),S^z(0)],$$
(120)
to find :
$$\chi ^{\prime \prime }(\omega )=\frac{1}{(2g\pi )^2}\frac{1}{\omega ^2}\text{Im}\left[𝒢(i\omega ,\theta _B)𝒢(i\omega ,\mathrm{})\right].$$
(121)
As a first example, let us consider the so called Toulouse limit or free fermion case. Then the only contribution comes from the soliton antisoliton form factors, which as discussed above is $`f(\theta _1,\theta _2)=i\mu e^{\theta _1/2}e^{\theta _2/2}`$. Hence,
$$\chi ^{\prime \prime }(\omega )=\frac{1}{\pi ^2}\text{Re}_{\mathrm{}}^{\mathrm{}}_{\mathrm{}}^{\mathrm{}}𝑑\theta _1𝑑\theta _2\frac{e^{\theta _1}e^{\theta _2}}{(e^{\theta _1}+ie^{\theta _B})(e^{\theta _2}+ie^{\theta _B})}\frac{1}{e^{\theta _1}+e^{\theta _2}}\delta (e^{\theta _1}+e^{\theta _2}\omega ),$$
(122)
that is
$$\chi ^{\prime \prime }(\omega )=\frac{2}{\pi ^2}\frac{T_B}{\omega }\text{Im}\left(_0^\omega 𝑑x\frac{1}{(x+iT_B)(\omega x+iT_B)}\right),$$
(123)
and
$`\chi ^{\prime \prime }(\omega )`$ $`=`$ $`{\displaystyle \frac{4}{\pi ^2}}{\displaystyle \frac{T_B}{\omega }}\text{Im}{\displaystyle \frac{1}{\omega +2iT_B}}\mathrm{ln}\left({\displaystyle \frac{\omega +iT_B}{iT_B}}\right)`$ (124)
$`=`$ $`{\displaystyle \frac{1}{\pi ^2}}{\displaystyle \frac{4T_B^2}{\omega ^2+4T_B^2}}\left[{\displaystyle \frac{1}{\omega }}\mathrm{ln}\left({\displaystyle \frac{T_B^2+\omega ^2}{T_B^2}}\right)+{\displaystyle \frac{1}{T_B}}\mathrm{tan}^1{\displaystyle \frac{\omega }{T_B}}\right].`$
In general, observe that, since the reflection matrix for solitons and antisolitons expands as a series in $`e^\beta `$, $`\frac{\chi ^{\prime \prime }(\omega )}{\omega }`$, will, for any coupling, expand as a series of the form $`(\omega /T_B)^{2n}`$ in the IR. In particular, this leads to a behaviour $`C(t)\frac{1}{t^2},t>>1`$ for any $`g`$. In the UV, one has to split integrals in two pieces. Since the soliton-soliton form factors expansion involves powers of $`\mathrm{exp}(\frac{1}{g}1)\theta `$, $`\frac{\chi ^{\prime \prime }(\omega )}{\omega }`$ expands as a double series in $`(T_B/\omega )^{22g}`$ and $`(T_B/\omega )^2`$ in the UV. Hence at short times, $`C(t)1t^{22g}`$. This is in agreement with the qualitative analysis of .
Results for $`g1/2`$ are more involved because there are non zero form factors at all levels. Still, when working out the first few form factors we observe a very rapid convergence with the number of rapidities and again we can give precise results for different values of $`g`$. As an example, let us show the results for $`g=1/3`$.
The computation for $`g=1/3`$ is very similar to the previous conductance computations. The boundary matrices are much more simpler though. In this case we have :
$$R_\pm ^{}=\mathrm{tanh}\left(\frac{\theta }{2}\frac{i\pi }{4}\right),R_\pm ^\pm =0,$$
(125)
and :
$$R_1^1=\frac{\mathrm{tanh}(\frac{\theta }{2}\frac{i\pi }{8})}{\mathrm{tanh}(\frac{\theta }{2}+\frac{i\pi }{8})}.$$
(126)
Then, as was found for the conductance, we find that the first two contributions are sufficient for most purposes, they are given by :
$$\mathrm{\Delta }\chi ^{\prime \prime }(\omega )^{(1)}=\frac{9\mu ^2d^2}{8\pi \omega }\text{Re }\left[\frac{\mathrm{tanh}\left(\frac{\mathrm{log}(\frac{\omega }{\sqrt{2}T_B})}{2}\frac{i\pi }{8}\right)}{\mathrm{tanh}\left(\frac{\mathrm{log}(\frac{\omega }{\sqrt{2}T_B})}{2}+\frac{i\pi }{8}\right)}1\right]$$
(127)
and :
$`\mathrm{\Delta }\chi ^{\prime \prime }(\omega )^{(2)}=`$ $``$ $`\left({\displaystyle \frac{3\mu d}{2\pi }}\right)^2{\displaystyle \frac{1}{\omega }}\text{Re }{\displaystyle _{\mathrm{}}^0}𝑑\theta {\displaystyle \frac{|\zeta (\theta \mathrm{log}(1e^\theta ))|^2}{\mathrm{cosh}^2(\theta \mathrm{log}(1e^\theta ))}}e^\theta `$ (128)
$`\times `$ $`\left[R_{}^+(\theta +\mathrm{log}(\omega /T_B))R_{}^+(\mathrm{log}[(1e^\theta )\omega /T_B])1\right].`$
Again these two expressions are sufficient to get a very precise result. Similar computations give rise to the results in figure 4 where we plotted $`S(\omega )\chi ^{\prime \prime }(\omega )/\omega `$ for the values, $`g=3/5,1/2,1/3,1/4`$.
When making these calculations we have to be careful about which terms are needed for a good convergence. Our observation is that keeping the form factors up to two rapidities give very good results precise to $`1\%`$. It is possible to go further and get a better precision if needed. The last statements are true for $`g[0.6,0.2]`$ and we believe even further (we can get rough bounds on the higher contributions and have an idea of the precision). Still, for the moment the isotropic Kondo point is difficult to treat.
To a good approximation, $`S(\omega )`$ can be represented for $`g<\frac{1}{2}`$ by a Lorentzian shape built around the four poles of the one-breather term:
$$S(\omega )\frac{cst}{\left[\left(\frac{\omega }{\mathrm{\Omega }}\right)^21+\left(\frac{\mathrm{\Gamma }}{\mathrm{\Omega }}\right)^2\right]^2+4\left(\frac{\mathrm{\Gamma }}{\mathrm{\Omega }}\right)^2}$$
(129)
where the poles are located at $`\omega =\pm \mathrm{\Omega }+\pm i\mathrm{\Gamma }`$. Here, it follows from the one breather reflection matrix that
$$\frac{\mathrm{\Omega }}{\mathrm{\Gamma }}=\mathrm{cot}\frac{\pi g}{2(1g)},$$
(130)
and $`\mathrm{\Omega }=T_B\mathrm{sin}\frac{\pi g}{(1g)}`$. The approximation improves as $`g`$ gets smaller.
According to (129), the pic in $`S(\omega )`$ disappears at $`g=\frac{1}{3}`$, where $`\mathrm{\Omega }=\mathrm{\Gamma }`$. The poles meanwhile reach the imaginary axis at $`g=\frac{1}{2}`$ where $`\frac{\mathrm{\Omega }}{\mathrm{\Gamma }}=0`$. These features are shared by the exact solution, as far as can be established from numerical study of the form-factors series: the pic in $`S(\omega )`$ disappears at $`g=\frac{1}{3}`$, while the tail oscillations of $`C(t)`$ disappear at $`g=\frac{1}{2}`$. Depending on one’s definition of the “transition” from coherent to incoherent regime, the latter therefore occurs at $`g=\frac{1}{3}`$ or $`g=\frac{1}{2}`$. This feature was somewhat unexpected from previous approximate studies, but is fully confirmed by recent numerical work . In any case, there is no real transtion per se, and it is possible that the study of different quantities (for instance multi point correlations) would exhibit a qualitative change of behaviour at still another value of $`g`$.
### 5.2 Shiba’s Relation.
Up untill now, we showed results for certain values of $`g`$ more or less limited by our ability (or tenacity) to write the form factors corresponding to that value of the anisotropy, and make them converge. It is not impossible to find general relations though; for example the behaviours in the UV and the IR in different models were infered in all generality.
Here we present a generalisation of Shiba’s relation which was proven for the Anderson model and generalised to Luttinger liquids by Sassetti and Weiss . The relation states that :
$$\underset{\omega 0}{lim}\frac{\chi ^{\prime \prime }(\omega )}{\omega }=2\pi g\chi _0^2$$
(131)
with $`\chi _0`$ the static succeptibility. If we look at the quantity :
$`𝒢(E)`$ $`=`$ $`E{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^0}{\displaystyle \frac{d\theta _1\mathrm{}d\theta _{2n1}}{(2\pi )^{2n}(2n)!}}{\displaystyle \frac{1}{1e^{\theta _1}\mathrm{}e^{\theta _{2n1}}}}`$ (132)
$`\times `$ $`K^{a_1b_1}(\mathrm{ln}(T_B/E)\theta _1)\mathrm{}K^{a_{n1}b_{2n1}}(\mathrm{ln}(T_B/E)\theta _{n1})`$
$`\times `$ $`K^{a_{2n}b_{2n}}\left[\mathrm{ln}(T_B/E)\mathrm{ln}\left(1e^{\theta _1}\mathrm{}e^{\theta _{2n1}}\right)\right]`$
$`\times `$ $`f_{a_1\mathrm{}a_{2n}}f_{b_1\mathrm{}b_{2n}}^{}[\theta _1\mathrm{}\theta _{2n1},\mathrm{ln}\left(1e^{\theta _1}\mathrm{}e^{\theta _{2n1}}\right)],`$
insert it in the expression for $`\chi ^{\prime \prime }(\omega )`$ and expand it around $`E0`$ we find that the contributions from the $`K`$ matrices all cancel and only a constant is left (we have to take into account the fact that the soliton/anti-solitons $`K`$ matrices always appear in pair). Then comparing this with the UV normalisation we find that :
$$\underset{\omega 0}{lim}\frac{\chi ^{\prime \prime }(\omega )}{\omega }=\frac{1}{\pi ^2gT_B^2}.$$
(133)
The total succeptibility is $`\chi =\chi ^{}+i\chi ^{\prime \prime }`$ and the static succeptibility $`\chi _0`$ which is the zero frequency limit of $`\chi ^{}`$ can also be infered from the previous expressions for the spin-spin correlation. We just need to take the real part when continuing (113) to real frequencies, which leads to :
$$\chi _0=\frac{1}{\pi ^2}gT_B.$$
(134)
Finally, in order to make contact with the usual form of Shiba’s relation, we need to renormalise the spins to $`1/2`$ and use the usual normalization for Fourier transforms, which leads indeed to (131).
## 6 Friedel oscillations: correlations involving Vertex operators
As exemplified in the previous sections, the method works naively indeed for currents, ie for operators with no anomalous dimension (calculations for the stress energy tensor for instance, would be very similar). Many physical properties are however described by more complicated operators, that is operators which have a non trivial anomalous dimension. As an example, I would like to discuss here the equivalent of Friedel oscillations in Luttinger liquids: more precisely, the $`2k_F`$ part of the charge density profile in a one dimensional Luttinger liquid away from an impurity. This is a problem which has attracted a fair amount of interest recently , .
We start with the bosonised form of the model. The hamiltonian is:
$$H=\frac{1}{2}_{\mathrm{}}^{\mathrm{}}𝑑x[8\pi g\mathrm{\Pi }^2+\frac{1}{8\pi g}(_x\varphi )^2]+\lambda \mathrm{cos}\varphi (0),$$
(135)
where we have set $`v_F=g`$. Then for the Friedel oscillations, the charge density operator is just :
$$\rho (x)=\rho _0+2_x\varphi +\frac{k_F}{\pi }\mathrm{cos}[2k_Fx+\varphi (x)].$$
(136)
with $`\rho _0=\frac{k_F}{\pi }`$ the background charge.
We decompose this system into even and odd basis (this is was explained already in ) by writing $`\varphi =\varphi _L+\varphi _R`$ and setting :
$`\phi ^e(x+t)={\displaystyle \frac{1}{\sqrt{2}}}\left[\varphi _L(x,t)+\varphi _R(x,t)\right]`$
$`\phi ^o(x+t)={\displaystyle \frac{1}{\sqrt{2}}}\left[\varphi _L(x,t)\varphi _R(x,t)\right]`$ (137)
Observe that these two field are left movers. We now fold the system by setting (prefactors here are slightly different from () because of the normalizations in the hamiltonian) :
$`\varphi _L^e=\sqrt{2}\phi ^e(x+t),x<0\varphi _R^e=\sqrt{2}\varphi ^e(x+t),x<0`$
$`\varphi _L^o=\sqrt{2}\phi ^e(x+t),x<0\varphi _R^o=\sqrt{2}\varphi ^e(x+t),x<0`$ (138)
and introduce new fields $`\varphi ^{e,o}=\varphi _L^{e,o}+\varphi _R^{e,o}`$. The density oscillations now read :
$$\frac{\rho (x)\rho _0}{\rho _0}=\mathrm{cos}(2k_Fx+\eta _F)\mathrm{cos}\frac{\varphi ^o(x)}{2}\mathrm{cos}\frac{\varphi ^e(x)}{2},$$
(139)
with $`\eta _F`$ the additional phase shift coming from the unitary transformation to eliminate the forward scattering term . $`\varphi ^o`$ is the odd field with Dirichlet boundary conditions, at the origin $`\varphi ^o(0)=0`$ leading to :
$$\mathrm{cos}\frac{\varphi ^o}{2}\left(\frac{1}{x}\right)^{g/2},$$
(140)
and the $`\varphi ^e`$ part is computed with the hamiltonian :
$$H^e=\frac{1}{2}_{\mathrm{}}^0𝑑x[8\pi g\mathrm{\Pi }^{e2}+\frac{1}{8\pi g}(_x\varphi ^e)^2]+\lambda \mathrm{cos}\frac{\varphi ^e(0)}{2}.$$
(141)
On general grounds, we expect the scaling form :
$$\mathrm{cos}\frac{\varphi ^e}{2}\left(\frac{1}{x}\right)^{g/2}F(\lambda x^{1g}),$$
(142)
where $`F`$ is a scaling function to be determined. Note that even the small $`x`$ behaviour of this function was not known in general.
To compute the correlation functions , we use the formalism introduced in the previous section (with $`\varphi ^e\mathrm{\Phi }`$). The massless scattering description and the boundary state are the same; only the operator which is studied changes.
The one point function of interest is $`0|\mathrm{cos}\frac{\mathrm{\Phi }}{2}|B`$. We thus need the matrix elements of the operator $`\mathrm{cos}\frac{\mathrm{\Phi }}{2}`$ in the quasiparticle basis: these still follow easily from the massive sine-Gordon form-factors , like the form-factors of $`\mathrm{\Phi }`$ in the previous sections. Difficulties however arise when one considers the contribution of each form-factor to the one-point function: the rapidity integrals turn out to be all IR divergent! This was not the case for the current operator, whose form factor has the naive engineering dimension of an energy, leading to convergent integrals: some sort of additional regularization is thus needed here.
To explain the strategy, consider first the case $`g=1/2`$. Here, the friedel oscillations are simply related to the spin one point function in an Ising model with boundary magnetic field. By using the same approach as the one outlined before, one finds the following form-factors expansion :
$`\sigma (x)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{n}{}}}\left\{{\displaystyle \frac{d\theta _i}{2\pi }}\mathrm{tanh}{\displaystyle \frac{\theta _B\theta _i}{2}}e^{2mxe^{\theta _i}}\right\}`$ (143)
$`\times `$ $`{\displaystyle \underset{i<j}{}}\left(\mathrm{tanh}{\displaystyle \frac{\theta _i\theta _j}{2}}\right)^2.`$
The integrals are all divergent at low energies, when $`\theta _i\mathrm{}`$ and the integrand goes to a constant. Let us then introduce an IR cut-off ( we chose $`\theta \theta _{min}`$ and set $`\mathrm{\Lambda }e^{\theta _{min}}`$) and take the log of the previous expressions (a similar method has been used in to study the UV limit of massive correlators. See also ,). Ordering this log by increasing number of integrations, one can show that each term diverges as $`\mathrm{ln}\mathrm{\Lambda }`$. Moreover, since the divergence occurs at very low energy, where the $`\mathrm{tanh}`$ goes to unity, the amplitudes of these $`\mathrm{ln}\mathrm{\Lambda }`$ do not depend on $`\theta _B`$ (for $`\theta _B\mathrm{}`$), ie on the boundary coupling. It is then easy to get rid of the cut-off: we simply substract the log of the IR spin function, ie we substract the same formal expression with $`\theta _B=\mathrm{}`$. The first two terms of the resulting expression read :
$`\mathrm{ln}{\displaystyle \frac{\sigma (x)_{T_B}}{\sigma (x)_{IR}}}`$ $`=`$ $`{\displaystyle _\mathrm{\Lambda }^{\mathrm{}}}{\displaystyle \frac{du}{2\pi u}}e^{2ux}\left({\displaystyle \frac{T_Bu}{T_B+u}}1\right)`$ (144)
$`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _\mathrm{\Lambda }^{\mathrm{}}}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle \frac{du_i}{2\pi u_i}}e^{2\mu u_ix}\left({\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle \frac{T_Bu_i}{T_B+u_i}}1\right)`$
$`\times `$ $`\left[\left({\displaystyle \frac{u_1u_2}{u_1+u_2}}\right)^21\right]+\mathrm{}`$
where we have set $`\mu =1`$, $`u_i=e^{\theta _i}`$, $`T_B=e^{\theta _B}\lambda ^{1/(1g)}`$.
Clearly, the integrals are now convergent at low energies, and we can send $`\mathrm{\Lambda }`$ to zero. Since the IR value of the one point function is easily determined by other means, $`<\sigma (x)>_{IR}x^{1/8}`$ , we can now obtain $`<\sigma (x)>_{T_B}`$ from (144). Hence the procedure involves a double regularization. Of course, there remains an infinity of terms to sum over. However, as in the case of current operators, the convergence of the form-factors expansion is very quick, and the first few terms are sufficient to get excellent accuracy all the way from UV to IR. To illustrate this more precisely, we recall that for $`g=1/2`$ (144) can be resummed in closed form , giving rise to :
$$R_{exact}=\frac{\sigma (x)_{T_B}}{\sigma (x)_{IR}}=\frac{1}{\sqrt{\pi }}\sqrt{2xT_B}e^{xT_B}K_0(xT_B).$$
(145)
By reexponentianing the two first terms in (144), one gets a ratio differing from (145) by at most $`1/100`$ for $`xT_B[0,\mathrm{})`$ (see figure 1).
By reexponentiating the first three terms, acuracy is improved to more than $`1/1000`$. Clearly, the form-factors approach thus provides analytical expressions that can be considered as exact for most reasonable purposes.
It is fair to mention however that, at any given order in (144), the exponent controlling the $`x0`$ behaviour is not exactly reproduced, as could be seen on a log-log plot. For instance, the first term is immediately found to produce a behaviour $`R(x)x^{1/\pi }`$, to be compared with the result $`R_{exact}(x)x^{1/2}\mathrm{ln}x`$. The comparison of the exact result (145) and of (144) show that the form factors expression has, term by term, the correct asymptotic expansion ie the IR expansion in powers of $`\frac{1}{xT_B}`$. Adding terms with more form-factors simply gives a more accurate determination of the coefficients. This is to be compared with the results of for eg the frequency dependent conductance, where the form factors expression had the correct functional dependence both in the UV and in the IR. This is not to say that the method is inefficient in the UV, because we know, at least formally, all the terms. In fact, we will show in what follows how the expansion (144) can always be resummed in the UV, and that the exponent can be exactly obtained from the form-factors approach too.
The regularization is the same for other values of $`g`$. Let us discuss here the cases $`g=1/t`$ with $`t`$ integer again. For these values, the scattering is diagonal and the form factors are rather simple. To obtain them, we again take the massless limit of the results in (but this time for vertex operator $`\mathrm{cos}\mathrm{\Phi }/2`$ (in normalizations where the bulk sine-Gordon perturbation is $`\mathrm{cos}\mathrm{\Phi }`$) instead of the current) and impose that half of the quasiparticles become right movers and half become left movers, since the boundary state always involve pairs of right and left moving particles. It is in fact easier to take that limit if we change basis from the solitons and anti-solitons to $`\frac{1}{\sqrt{2}}(|S>\pm |A>)`$. In that case, the boundary scattering matrix becomes diagonal and the isotopic indices always come in pairs. The reflection matrices in this new basis are given by :
$`R_{}(\theta )=e^{i\frac{\pi }{4}(2t)}\mathrm{tanh}\left[{\displaystyle \frac{(t1)\theta }{2}}+i{\displaystyle \frac{\pi (t2)}{4}}\right]R(i{\displaystyle \frac{\pi }{2}}\theta )`$
$`R_+(\theta )=e^{i\frac{\pi }{4}(2t)}R(i{\displaystyle \frac{\pi }{2}}\theta )`$ (146)
with (note the slight change of notation compared with (85)) :
$$R(\theta )=\mathrm{exp}\left(i_{\mathrm{}}^{\mathrm{}}\frac{dy}{2y}\mathrm{sin}\frac{2(t1)y\theta }{\pi }\frac{\mathrm{sinh}(t2)y}{\mathrm{sinh}(2y)\mathrm{cosh}(t1)y}\right).$$
(147)
The breathers reflection matrices follow from (see also (97).
The case $`g=1/2`$ having already been worked out, let us concentrate on $`g=1/3`$. There, in addition to the soliton and anti-soliton, there is also one breather. The first contribution to the one point function comes from the two breathers form-factor, with one right moving and one left moving breather. It is given by a constant :
$$f(\theta ,\theta )_{11}^{LR}=c_1,$$
(148)
and this obviously leads to IR divergences. Other contributions come from $`2n`$ breathers form-factors, and $`4n`$ solitons form-factors The whole expression can be controlled as for $`g=1/2`$, by taking the log, and factoring out the IR part. Setting $`c(x)=\mathrm{cos}\frac{\varphi (x)}{2}`$, we organize the sum as follows :
$$\mathrm{ln}\frac{c(x)_{T_B}}{c(x)_{IR}}=\mathrm{ln}R^{(2)}+\mathrm{ln}R^{(4)}+\mathrm{}$$
(149)
with the subscript denoting the number of intermediate excitations.
Then, using the explicit expressions for $`g=1/3`$ we find :
$$\mathrm{ln}R^{(2)}=2c_1e^{2\sqrt{2}T_Bx}Ei(2\sqrt{2}T_Bx),$$
(150)
where $`E_i`$ is the standard exponential integral. The next term $`\mathrm{ln}R^{(4)}`$ is a bit bulky to be written here, but it is very easy to obtain - similar expressions have been explicitely given in the previous sections.This is all what is needed for an accuracy better than 1 percent. In figure 6 we present the results of the ratio at $`g=1/2,1/3,1/4`$ for the Friedel oscillations. It should be noted that this ratio is just the pinning function of reference and our results agree well qualitatively with the results found there.
As mentioned before, the deep UV behaviour is a little more difficult to obtain: the accuracy is good because the ratio goes to zero anyway, but the numerical evaluation of the power law would not be too accurate with the number of terms we consider. Fortunately, the full form-factors expansion allows the analytic determination of this exponent. First, observe for instance that in (144) the integrals converge for all $`T_B0`$, but strictly at $`T_B=0`$, they do not. To get the dependence of $`c(x)`$ as $`T_B0`$, we will consider, the logarithm of another ratio, $`\mathrm{ln}\frac{c(x)_{T_B}}{c(x^{})_{T_B}}`$, where $`x`$ and $`x^{}`$ are two arbitrary coordinates. For this ratio, even at $`T_B=0`$, the integrals are convergent. But $`T_B=0`$ is the UV fixed point, with Neumann boundary conditions. While the one point function $`c(x)_{UV}`$ vanishes, the ratio of two such one point functions is well defined, and can be computed by putting an IR cut-off (a finite system). One finds that it goes as $`(x/x^{})^{g/2}`$. By regularity as $`T_B0`$, the same is true for the ratio close to $`T_B=0`$, and thus one has
$$c(x)(xT_B)^{g/2},x(T_B)0$$
(151)
This shows that the universal scaling function in (142) behaves as $`F(y)y^{\frac{g}{1g}}`$ for $`g<\frac{1}{2}`$. This exponent can actually be obtained by perturbation theory. Indeed, the first term in the perturbative expansion of $`c(x)`$ is
$$\lambda x^{g/2}_{\mathrm{}}^{\mathrm{}}\frac{dy}{(x^2+y^2)^g}$$
(152)
For $`g<1/2`$, this integral diverges in the IR. To regulate it, we need to put a new cut-off: since there is no other length scale in the problem, this can be nothing but $`1/T_B`$. Changing variables, the leading behaviour is $`x^{g/2}T_B^gx^{g/2}\lambda ^{g/1g}`$, in agreement with the previous discussion. The exponent coincides with the result of the self-consistent harmonic approximation ; but it is important to stress that the latter is valid only for $`g<<1`$.
The function $`F(y)`$ behaves a $`y\mathrm{ln}y`$ for $`g=1/2`$. For $`g>1/2`$, its behaviour is simply $`F(y)y`$, as can be easily shown since the perturbative approach is now convergent. As we approach $`g=1/2`$ this exponent seems to become asymptotic and is more difficult to get numerically .
## 7 Conclusion
The description of the Hilbert space in terms of the integrable quasiparticles basis allows, in a wide variety of cases, extremely accurate computations of time and space dependent correlators. The related form-factors method is not especially elegant, but gives rise to susprisingly good results: in fact, it is not clear whether exact analytical expressions of the correlation functions - if ever obtained - will be more useful. Although we limited ourselves to $`g=1/t`$ with $`t`$ integer in these notes, all values of $`g<1/2`$ are accessible, but general computations are more complicated since the bulk scattering is non diagonal. The methods explained here should be generalizable to other problems, in particular the determination of the screening cloud in the anisotropic Kondo model .
The region $`g>1/2`$ \- in particular the $`SU(2)`$ symmetric point $`g=1`$ <sup>1</sup><sup>1</sup>1Of course the problem of Friedel oscillations for $`g=1`$ can be solved by fermionization , since it corresponds to non interacting fermions. In our approach, this point is non trivial because of the folding. This folding however is necessary for any value $`g1`$: except at $`g=1`$, the problem on the whole line would not be integrable. presents additional difficulties, which are not yet solved. The situation does not look desperate however.
As a last but important comment, I would like to stress that the form-factors technique has also been applied succesfully to the computation of quantities of experimental interest in the case of (bulk) massive theories. See the lectures by F. Essler in this school, or the papers . One point functions in massive theories and in the presence of a boundary have also been recently studied in .
Acknowledgments: These notes result from close collaborations with F. Lesage and S. Skorik, to whom I am very thankful. Support from the DOE, the NSF and the Packard Foundation is also gratefully acknowledged.
|
warning/0007/hep-lat0007038.html
|
ar5iv
|
text
|
# Quenched Lattice QCD with Domain Wall Fermions and the Chiral Limit
## I Introduction
Since spontaneous chiral symmetry breaking is a dominant property of the QCD vacuum and is responsible for much of the low energy physics seen in Nature, having a first principles formulation of lattice QCD which does not explicitly break chiral symmetry has been an important goal. Both Wilson and staggered fermions recover chiral symmetry in the continuum limit but with these techniques the chiral and continuum limits cannot be decoupled. For the QCD phase transition, which is dominantly a chiral symmetry restoring transition, a formulation that is free of violations of chiral symmetry due to lattice artifacts, should give a phase transition more closely approximating that of the continuum limit. For the measurement of matrix elements of operators in hadronic states, a formulation that respects chiral symmetry on the lattice substantially reduces operator mixing through renormalization. Lastly, since much of our analytic understanding of low-energy QCD is formulated in terms of low-energy effective field theories based on chiral symmetry, a lattice formulation preserving chiral symmetry allows controlled comparison with analytic expectations.
Building on the work of Kaplan , who showed how to produce light chiral modes in a $`d`$ dimensional theory as surface states in a $`d+1`$ dimensional theory, a number of attractive lattice formulations have been developed which achieve a decoupling of the continuum and chiral limits. Here we will use Kaplan’s approach as was further developed by Narayanan and Neuberger and by Shamir . It is Shamir’s approach, commonly known as the domain wall fermion formulation, which we adopt. (For reviews of this topic see Refs. and for more extensive recent references see Ref. .) For a physical four-dimensional problem, the domain wall fermion Dirac operator, $`D`$, is a five-dimensional operator with free boundary conditions for the fermions in the new fifth dimension. The desired light, chiral fermions appear as states exponentially bound to the four-dimensional surfaces at the ends of the fifth dimension. The remaining modes for $`D`$ are heavy and delocalized in the fifth dimension.
An additional important feature of the domain wall fermion Dirac operator in the limit $`L_s\mathrm{}`$ is the existence of an “index”, an integer that is invariant under small changes in the background gauge field. Here $`L_s`$ is the extent of the lattice in the fifth dimension. This property, true for all but a set of gauge fields of measure zero, can be readily seen using the overlap formalism . In the smooth background field limit, this index is the normal topological charge but, even for rough fields, it signals the presence of massless fermion mode(s) when non-zero. These zero modes can easily be recognized in numerical studies with semiclassical gauge field backgrounds .
These powerful theoretical developments in fermion formulations require additional study to demonstrate their merit for numerical work. For the case of domain wall fermions, a growing body of numerical results are available. Both quenched and dynamical domain wall fermion simulations have been conducted and the domain wall approach is readily adapted to current algorithms for lattice QCD. (Much work is also being done on the numerical implementation of the overlap formulation and its variations .) A fundamental question, which is a major part of this paper, involves quantifying the residual chiral symmetry breaking effects of finite extent in the fifth dimension.
Due to current limits on computer speed, some lattice QCD studies are only practical when the fermionic determinant is left out of the measure of the path integral. The resulting quenched theory does not suppress gauge field configurations with light fermionic modes, in contrast with the original theory where, for small quark mass, the determinant strongly damps such configurations. The measurement of observables involving fermion propagation through configurations with unsuppressed light fermionic modes can in principle lead to markedly different infrared behavior than that found in full QCD, in the limit of small quark masses. Domain wall fermions, which produce light chiral modes at finite lattice spacing and preserve the global symmetries of continuum QCD, should produce a well-defined chiral limit for full QCD. The central question addressed in this paper is whether a well-controlled chiral limit also exists within the quenched approximation. A thorough theoretical and numerical understanding of the quenched chiral limit is essential if the good chiral properties of domain wall fermions are to be exploited in quenched lattice simulations.
Here we present results from extensive simulations of quenched QCD with domain wall fermions, primarily at two lattice spacings, $`a^11`$ and 2 GeV. Many different values for the fifth dimensional extent, $`L_s`$, and the bare quark mass, $`m_f`$ have been used. Hadron masses, $`f_\pi `$ and the chiral condensate, $`\overline{q}q`$, are the primary hadronic observables we have studied. In calculating physical observables using domain wall fermions, four-dimensional quark fields $`q(x)`$ are defined from the five-dimensional fields $`\mathrm{\Psi }(x,s)`$ by taking the left-handed fields from the four-dimensional hypersurface with smallest coordinate in the fifth dimension and the right-handed fields from the hypersurface with the largest value of this coordinate. We also present results from measuring the lowest eigenvectors and eigenvalues of the hermitian domain wall fermion operator.
Here we list the major topics in each section of this paper. Section II defines our conventions and gives details of the hermitian domain wall fermion operator. Section III discusses our simulation parameters and fitting procedures and includes tables of run parameters and hadron masses for $`m_f0.01`$. In Section IV a precise understanding of how finite $`L_s`$ effects enter $`\overline{q}q`$ is developed and measurements of $`\overline{q}q`$ which show the role of fermionic zero modes are reported. We study the pion mass in the chiral limit in Section V, which requires understanding zero mode effects. Section VI contains two determinations of the residual chiral symmetry breaking for finite $`L_s`$; one from measuring appropriate pion correlators and the other from the explicitly measured small eigenvalues and eigenvectors of the hermitian domain wall fermion operator. Our determination of $`f_\pi `$, an important check of the chiral properties of domain wall fermions, is discussed in Section VII, along with the scaling of hadron masses.
Because of the length of this paper and the number of topics covered, we now give a brief summary of our major results, organized to correspond to the expanded discussion in Sections IV, V VI and VII.
Zero mode effects in $`\overline{q}q`$: As already mentioned, the domain wall fermion operator $`D`$ has an Atiyah-Singer index $`\nu `$ for $`L_s\mathrm{}`$. However, in quenched QCD, $`\nu `$ plays no role in the generation of gauge field configurations. For $`\nu 0`$, both $`D`$ and the hermitian domain wall fermion operator $`D_H`$ have zero modes. Since $`\overline{q}q`$ is an appropriately restricted trace of $`D_H^1`$ it should diverge as $`|\nu |/m_fV`$ for small $`m_f`$ if the ensemble average of $`|\nu |`$ is non-zero. Here $`V`$ is the four-dimensional, space-time volume of the lattice being studied. For large but finite $`L_s`$, the residual chiral symmetry breaking should cut off this divergence.
Figure 1 shows $`\overline{q}q`$ versus the quark mass $`m_f`$ for $`a^11`$ GeV on two different volumes of linear dimensions of about 1.6 and 3.2 Fermi. A divergence for $`m_f0`$ is clearly visible on the smaller volume, but not on the larger. This is expected since $`|\nu |/V`$ should go as $`1/\sqrt{V}`$ and is clear evidence for unsuppressed zero modes in quenched QCD, first reported in Ref. . Notice that there may be other problems with the chiral limit of $`\overline{q}q`$ that are masked by this $`1/m_f`$ divergence.
The chiral limit of $`m_\pi `$: With this clear evidence for zero mode effects in $`\overline{q}q`$, one might expect to see zero mode contributions in any quark propagator $`D^1(x,y)`$ if at both $`x`$ and $`y`$ a single zero eigenvector has reasonable magnitude. For sufficiently large volume, needed to see asymptotic behavior in the limit of large $`|xy|`$, there should be no zero mode effects. Our results for the zero mode effects on the pion mass are presented in Figure 2 which shows $`m_\pi ^2`$ versus $`m_f`$ for $`8^3\times 32`$ lattices with $`L_s=48`$ and Figure 3, where all the parameters are the same except that the volume was increased to $`16^3\times 32`$. The pion mass is determined from three different correlators which are each affected differently by zero modes. For the smaller volume, the pion masses measured disagree for small $`m_f`$, while they agree for the larger volume.
Notice that on the larger volume shown in Figure 3, where zero-mode effects are not apparent, $`m_\pi ^2`$ shows signs of curvature in $`m_f`$ with the three $`m_f=0`$ values lying below the extrapolation from larger masses. In addition, this simple large-mass linear extrapolation vanishes at a value of $`m_f`$ that is more negative than the point $`m_f+m_{\mathrm{res}}=0`$ (shown in the graph by the star) also suggesting downward concavity. While the discrepancy between this $`x`$-intercept and the point $`m_f+m_{\mathrm{res}}=0`$, may be caused by $`O(a^2)`$ effects, we find a considerably larger discrepancy when making a similar comparison at $`\beta =6.0`$. Thus, we have evidence that $`m_\pi ^2`$ does not depend linearly on $`m_f`$ in the chiral limit.
Determining the residual mass: In the limit of small lattice spacing, the dominant chiral symmetry breaking effect, due to the mixing between the domain walls, is the appearance of a residual mass, $`m_{\mathrm{res}}`$ in the low energy effective Lagrangian. The Ward-Takahashi identity for domain wall fermions has an additional contribution representing this explicit chiral symmetry breaking due to finite $`L_s`$. Matrix elements of this additional term between low energy states determine the residual quark mass. Figure 4 shows our results for $`m_{\mathrm{res}}`$ for $`16^3\times 32`$ lattices at $`\beta =6.0`$ as a function of $`L_s`$. $`m_{\mathrm{res}}`$ is clearly falling with $`L_s`$ and reaches a value of $`2`$ MeV for $`L_s24`$. Our data does not resolve the precise behavior of $`m_{\mathrm{res}}`$ for large $`L_s`$, but the very small value makes this less important for current simulations. A similar study on lattices with $`a^11`$ GeV but with larger $`L_s=48`$ finds a value of $`m_{\mathrm{res}}/m_s=0.074(5)`$ or $`m_{\mathrm{res}}8`$ MeV.
We have also used the Rayleigh-Ritz method, implemented using the technique of Kalk-Reuter and Simma , to determine the low-lying eigenvalues and eigenvectors for the hermitian domain wall fermion operator. The results exhibit the approximate behavior expected from low-energy excitations in the domain wall formulation. We use the resulting eigenvalues to provide an independent estimate of the residual mass which is nicely consistent with the more precise value determined from pseudoscalar correlators.
Results for $`f_\pi `$, hadron masses and scaling: With this detailed understanding of the chiral limit of quenched lattice QCD with domain wall fermions, we have calculated $`f_\pi `$ using both pseudoscalar and axial-vector correlators. The results for lattices with $`a^12`$ GeV are shown in Figure 5, where good agreement between the two methods is seen. To do this comparison, the appropriate Z-factor for the local axial current must be determined and a consistent value for $`m_{\mathrm{res}}`$ must be known. The good agreement in the figure is a significant test of these measurements as well as the chiral properties of domain wall fermions. We find very good scaling in the ratio $`f_\pi /m_\rho `$ for $`a^11`$ to 2 GeV. For $`m_N/m_\rho `$ scaling is within 6%. We also find that $`\overline{q}q=(256(8)\mathrm{MeV})^3`$ from our $`\beta =6.0`$ simulations.
## II Domain wall fermions
In this section we first define our notation, including the domain wall fermion Dirac operator, and then derive the precise form of the Banks-Casher relation for domain wall fermions, to second order in the quark mass. In this paper, the variable $`x`$ specifies the coordinates in the four-dimensional space-time volume, with extent $`L`$ along each of the spatial directions and extent $`N_t`$ along the time direction, while $`s=0,1,\mathrm{},L_s1`$ is the coordinate of the fifth direction, with $`L_s`$ assumed to be even. The space-time volume $`V`$ is given by $`V=L^3N_t`$. The domain wall fermion operator acts on a five-dimensional fermion field, $`\mathrm{\Psi }(x,s)`$, which has four spinor components. A generic four-dimensional fermion field, with four spin components, will be denoted by $`\psi (x)`$, while the specific four-dimensional fermion field defined from $`\mathrm{\Psi }(x,s)`$ will be denoted by $`q(x)`$. The space-time indices for vectors will be enclosed in parenthesis while for matrices they will be given as subscripts. Our general formalism follows that developed by Furman and Shamir .
### A Conventions
The domain wall fermion operator is given by
$$D_{x,s;x^{},s^{}}=\delta _{s,s^{}}D_{x,x^{}}^{}+\delta _{x,x^{}}D_{s,s^{}}^{}$$
(1)
$`D_{x,x^{}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{\mu =1}{\overset{4}{}}}\left[(1\gamma _\mu )U_{x,\mu }\delta _{x+\widehat{\mu },x^{}}+(1+\gamma _\mu )U_{x^{},\mu }^{}\delta _{x\widehat{\mu },x^{}}\right]`$ (2)
$`+`$ $`(M_54)\delta _{x,x^{}}`$ (3)
$`D_{s,s^{}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[(1\gamma _5)\delta _{s+1,s^{}}+(1+\gamma _5)\delta _{s1,s^{}}2\delta _{s,s^{}}\right]`$ (4)
$``$ $`{\displaystyle \frac{m_f}{2}}\left[(1\gamma _5)\delta _{s,L_s1}\delta _{0,s^{}}+(1+\gamma _5)\delta _{s,0}\delta _{L_s1,s^{}}\right].`$ (5)
Here, $`U_{x,\mu }`$ is the gauge field at site $`x`$ in direction $`\mu `$, and $`s`$ and $`s^{}`$ lie in the range $`0s,s^{}L_s1`$. The five-dimensional mass, representing the height of the domain wall in Kaplan’s original language, is given by $`M_5`$, while $`m_f`$ directly couples the two domain walls at $`s=0`$ and $`s=L_s1`$. Since the light chiral modes should be exponentially bound to the domain walls, $`m_f`$ mixes the two chiralities and is therefore the input bare quark mass. The value of $`M_5`$ must be chosen to produce these light surface states and, in the free field case, $`0<M_5<2`$ produces a single fermion flavor with the left-hand chirality bound to $`s=0`$ and the right to $`s=L_s1`$. In order to use our pre-existing, high-performance Wilson fermion operator computer program as part of our domain wall fermion operator, we have used the operator $`D`$ above, which is the same as $`D_F^{}`$ of Ref. .
Following Ref. , we define the four-dimensional quark fields $`q(x)`$ by
$`q(x)`$ $`=`$ $`P_\mathrm{L}\mathrm{\Psi }(x,0)+P_\mathrm{R}\mathrm{\Psi }(x,L_s1)`$ (6)
$`\overline{q}(x)`$ $`=`$ $`\overline{\mathrm{\Psi }}(x,L_s1)P_\mathrm{L}+\overline{\mathrm{\Psi }}(x,0)P_\mathrm{R}`$ (7)
where we have used the projection operators $`P_{R,L}=(1\pm \gamma _5)/2`$. Symmetry transformations of the five-dimensional fields yield a four-dimensional axial current
$$𝒜_\mu ^a(x)=\underset{s=0}{\overset{L_s1}{}}\mathrm{sign}\left(s\frac{L_s1}{2}\right)j_\mu ^a(x,s).$$
(8)
Here
$$j_\mu ^a(x,s)=\frac{1}{2}\left[\overline{\mathrm{\Psi }}(x+\widehat{\mu },s)(1+\gamma _\mu )U_{x+\widehat{\mu },\mu }^{}t^a\mathrm{\Psi }(x,s)\overline{\mathrm{\Psi }}(x,s)(1\gamma _\mu )U_{x,\mu }t^a\mathrm{\Psi }(x+\widehat{\mu },s)\right]$$
(9)
while the flavor matrices are normalized to obey $`\mathrm{Tr}(t^at^b)=\delta ^{ab}`$. The divergence of this current satisfies
$$\mathrm{\Delta }_\mu 𝒜_\mu ^a(x)=2m_fJ_5^a(x)+2J_{5q}^a(x)$$
(10)
where $`\mathrm{\Delta }_\mu f(x)=f(x)f(x\widehat{\mu })`$ is a simple finite difference operator and the pseudoscalar density $`J_5^a(x)`$ is
$`J_5^a(x)`$ $`=`$ $`\overline{\mathrm{\Psi }}(x,L_s1)P_\mathrm{L}t^a\mathrm{\Psi }(x,0)+\overline{\mathrm{\Psi }}(x,0)P_\mathrm{R}t^a\mathrm{\Psi }(x,L_s1)`$ (11)
$`=`$ $`\overline{q}(x)t^a\gamma _5q(x).`$ (12)
This equation differs from the corresponding continuum expression by the presence of the $`J_{5q}^a(x)`$ term, which is built from point-split operators at $`L_s/2`$ and $`L_s/21`$ and is given by
$$J_{5q}^a(x)=\overline{\mathrm{\Psi }}(x,L_s/21)P_\mathrm{L}t^a\mathrm{\Psi }(x,L_s/2)+\overline{\mathrm{\Psi }}(x,L_s/2)P_\mathrm{R}t^a\mathrm{\Psi }(x,L_s/21).$$
(13)
We will refer to this term as the “mid-point” contribution to the divergence of the axial current.
This mid-point term adds an additional term to the axial Ward-Takahashi identities and modifies observables, like the pion mass, which are controlled by these identities. The Ward-Takahashi identity is
$$\mathrm{\Delta }_\mu 𝒜_\mu ^a(x)O(y)=2m_fJ_5^a(x)O(y)+2J_{5q}^a(x)O(y)+i\delta ^aO(y).$$
(14)
For operators, $`O`$ made from the fields $`q(y)`$ and $`\overline{q}(y)`$, it has been shown that the $`J_{5q}^a`$ term in Eq. 14 vanishes for flavor non-singlet currents when $`L_s\mathrm{}`$. For the singlet current, this extra term generates the axial anomaly. The mid-point term represents the contribution of finite $`L_s`$ effects on the low-energy physics of domain wall fermions.
### B Definition of the residual mass and the chiral limit
For domain wall fermions, the axial transformation which leads to the Ward-Takahashi identity of Eq. 14 rotates the fermions in the two half-spaces along the fifth direction with opposite charges. For $`m_f=0`$, the action is not invariant under this transformation due to the coupling of the left- and right-handed light surface states at the midpoint of the fifth dimension. This results in the additional term in the divergence of the axial current, as given in Eq. 13. In the $`L_s\mathrm{}`$ limit where the explicit mixing between the $`s0`$ and $`sL_s1`$ states vanishes, this extra “mid-point” contribution will be zero and a continuum-like Ward-Takahashi identity will be realized.
Since we must work at finite $`L_s`$ it is useful to characterize the chiral symmetry breaking effects of mixing between the domain walls as precisely as possible. We do this by adopting the language of the Symanzik improvement program . Here we use an effective continuum Lagrangian $`_n`$ to reproduce to $`O(a^n)`$ the amplitudes predicted by our lattice theory when evaluated at low momenta and finite lattice spacing. Clearly $`_0`$ is simply the continuum QCD Lagrangian, while $`_1`$ will include the dimension-five, clover term: $`\overline{\psi }\sigma ^{\mu \nu }\psi F_{\mu \nu }`$ . The chiral symmetry breaking effects of mixing between the domain walls will appear to lowest order in $`a`$ as an additional, dimension three operator $`a^1e^{\alpha L_s}\overline{\psi }\psi `$. This term represents the residual mass term that remains even after the explicit input chiral symmetry breaking parameter $`m_f`$ has been set to zero. The next chiral symmetry breaking contribution from domain wall mixing will be $`O(a^2)`$ smaller, appearing as a coefficient of order $`a^1`$ for the clover term.
We define the chiral symmetry breaking parameter $`m_{\mathrm{res}}`$ so the complete coefficient of the mass term in $`_0`$ is proportional to the simple sum $`m_f+m_{\mathrm{res}}`$. While this is a precise definition of $`m_{\mathrm{res}}`$, valid for finite lattice spacing, a precise determination of $`m_{\mathrm{res}}`$ in a lattice calculation will be impeded by the need to quantitatively account for the additional chiral symmetry breaking effects of terms of higher order in $`a`$.
Close to the continuum limit, for long distance amplitudes, the Ward-Takahashi identity given in Eq. 14 must agree with the corresponding identity in the effective continuum theory. Thus, for the non-singlet case, the sum of the first two terms on the right-hand side of Eq. 14 must be equivalent to an effective quark mass, $`m_{\mathrm{eff}}=m_f+m_{\mathrm{res}}`$, times the pseudo-scalar density $`J_5^a`$. Thus, the residual mass, $`m_{\mathrm{res}}`$ appears in the low energy identity:
$$J_{5q}^am_{\mathrm{res}}J_5^a$$
(15)
where this equality will hold up to $`O(a^2)`$ in low-momentum amplitudes.
Thus, close to the continuum limit, $`m_{\mathrm{res}}`$ in Eq. 15 is a universal measure of the chiral symmetry breaking effects of domain wall fermions for all low energy matrix elements, with corrections coming from terms of higher order in the lattice spacing. However, away from the continuum limit the $`O(a^2)`$ terms may be appreciable. In addition, if there are high energy scales entering an observable, such a low energy description is not valid and the explicit chiral symmetry breaking effects of finite $`L_s`$ can be more complicated than a simple additive shift of the input quark mass by $`m_{\mathrm{res}}`$.
Many aspects of the chiral behavior of the domain wall theory can be easily understood by reference to the more familiar Wilson fermion formulation. For finite $`L_s`$ the domain wall formulation can be viewed as an “on- and off-shell improved” version of Wilson fermions. The low energy effective Lagrangian for domain wall fermions is the same as that for the Wilson case except the coefficients of the chiral symmetry breaking terms are expected to decrease exponentially with $`L_s`$. Viewed in this way, one might expect to achieve a vanishing pion mass by fine-tuning $`m_f`$ to a critical value, $`m_{fc}`$ in very much the same way as one fine-tunes $`\kappa `$ to $`\kappa _c`$ for Wilson fermions. As the above discussions demonstrates, $`m_{fc}=m_{\mathrm{res}}+O(a^2)`$. Just as in the Wilson case, this limit can be interpreted as the approach to the critical surface of the Aoki phase .
### C The hermitian domain wall fermion operator
A hermitian operator $`D_H`$ can be constructed from $`D`$ through
$$D_H\gamma _5R_5D$$
(16)
where $`(R_5)_{ss^{}}\delta _{s,L_s1s^{}}`$ is the reflection in the fifth dimension around the five-dimensional midpoint, $`s=(L_s1)/2`$. Writing out $`D_H`$ gives
$`D_H`$ $`=`$ $`\gamma _5D_{x,x^{}}^{}\delta _{s+s^{},L_s1}`$ (17)
$`+`$ $`\gamma _5[P_L\delta _{s+s^{},L_s}+P_R\delta _{s+s^{},L_s2}\delta _{s+s^{},L_s1}`$ (18)
$``$ $`m_f(P_L\delta _{s,0}\delta _{s^{},0}+P_R\delta _{s,L_s1}\delta _{s^{},L_s1})]\delta _{x,x^{}}`$ (19)
while as an explicit matrix in the $`s,s^{}`$ indices:
$$D_H=\left(\begin{array}{cccccc}m_f\gamma _5P_L& & & & \gamma _5P_R& \gamma _5(D^{}1)\\ & & & \gamma _5P_R& \gamma _5(D^{}1)& \gamma _5P_L\\ & & \mathrm{}& \mathrm{}& \mathrm{}& \\ & \mathrm{}& \mathrm{}& \mathrm{}& & \\ \gamma _5P_R& \gamma _5(D^{}1)& \gamma _5P_L& & & \\ \gamma _5(D^{}1)& \gamma _5P_L& & & & m_f\gamma _5P_R\end{array}\right).$$
(20)
The eigenfunctions and eigenvectors of $`D_H`$ will be denoted by
$$D_H\mathrm{\Psi }_{\mathrm{\Lambda }_H}=\mathrm{\Lambda }_H\mathrm{\Psi }_{\mathrm{\Lambda }_H}$$
(21)
with the five-dimensional propagator given by
$$S_{x,s;x^{},s^{}}^{(5)}=\underset{\mathrm{\Lambda }_H}{}\frac{\mathrm{\Psi }_{\mathrm{\Lambda }_H}(x,s)\mathrm{\Psi }_{\mathrm{\Lambda }_H}^{}(x^{},\stackrel{~}{s})\gamma _5(R_5)_{\stackrel{~}{s},s^{}}}{\mathrm{\Lambda }_H}.$$
(22)
(Grassmann variables in the Euclidean path integral will be denoted by $`\overline{\mathrm{\Psi }}`$ and $`\mathrm{\Psi }`$, while the eigenfunctions of $`D_H`$ will be denoted $`\mathrm{\Psi }_{\mathrm{\Lambda }_H}^{}`$ and $`\mathrm{\Psi }_{\mathrm{\Lambda }_H}`$.)
We will find it convenient to define three additional matrices
$$(\mathrm{\Gamma }_5)_{s,s^{}}=\delta _{s,s^{}}\mathrm{sign}\left(\frac{L_s1}{2}s\right)$$
(23)
$$Q_{s,s^{}}^{(\mathrm{w})}=P_L\delta _{s,0}\delta _{s^{},0}+P_R\delta _{s,L_s1}\delta _{s^{},L_s1}$$
(24)
and
$$Q_{s,s^{}}^{(\mathrm{mp})}=P_L\delta _{s,L_s/2}\delta _{s^{},L_s/2}+P_R\delta _{s,L_s/21}\delta _{s^{},L_s/21}.$$
(25)
The transformation which generates the current in Eq. 8 is
$`\mathrm{\Psi }`$ $``$ $`\mathrm{exp}(i\alpha ^at^a\mathrm{\Gamma }_5)\mathrm{\Psi }`$ (26)
$`\overline{\mathrm{\Psi }}`$ $``$ $`\overline{\mathrm{\Psi }}\mathrm{exp}(i\alpha ^at^a\mathrm{\Gamma }_5).`$ (27)
The matrices $`Q^{(\mathrm{w})}`$ and $`Q^{(\mathrm{mp})}`$ are the two parts of $`D_H`$ which correspond to terms in $`D=\gamma _5R_5D_H`$ which are not invariant under the transformation in Eq. 27. The matrix $`Q^{(\mathrm{w})}`$ underlies the explicit mass term and, in the original operator $`D`$, explicitly mixes the $`s=0`$ and $`s=L_s1`$ walls. Likewise, the matrix $`Q^{(\mathrm{mp})}`$ is a “mid-point” matrix with non-zero elements only in the center of the fifth dimension. It represents the component of the operator $`D`$ which connects the left and right half regions. These two contributions provide the terms on the right hand side of Eq. 10 and one easily finds
$$\{\mathrm{\Gamma }_5,D_H\}=2m_fQ^{(\mathrm{w})}+2Q^{(\mathrm{mp})}.$$
(28)
Since it is expected that there are eigenvectors of $`D_H`$ which are exponentially localized on the domain walls, we see that with $`m_f=0`$ and the limit $`L_s\mathrm{}`$ taken, $`D_H`$ anticommutes with $`\mathrm{\Gamma }_5`$ in the subspace of these eigenvectors. This is the property expected for massless, four-dimensional fermions in the continuum in Euclidean space.
Using the matrix, $`Q^{(\mathrm{w})}`$, we can write a simple form for the four-dimensional chiral condensate, $`\overline{q}q`$
$`\overline{q}q`$ $`=`$ $`{\displaystyle \frac{1}{12V}}{\displaystyle \underset{x}{}}\overline{q}(x)q(x)`$ (29)
$`=`$ $`{\displaystyle \frac{1}{12V}}{\displaystyle \underset{x}{}}\overline{\mathrm{\Psi }}(x,s)(R_5Q^{(\mathrm{w})})_{s,s^{}}\mathrm{\Psi }(x,s^{})`$ (30)
$`=`$ $`{\displaystyle \frac{1}{12V}}{\displaystyle \underset{x,\mathrm{\Lambda }_H}{}}{\displaystyle \frac{\mathrm{\Psi }_{\mathrm{\Lambda }_H}^{}(x,s)\gamma _5Q_{s,s^{}}^{(\mathrm{w})}\mathrm{\Psi }_{\mathrm{\Lambda }_H}(x,s^{})}{\mathrm{\Lambda }_H}}`$ (31)
$`=`$ $`{\displaystyle \frac{1}{12V}}{\displaystyle \underset{\mathrm{\Lambda }_H}{}}{\displaystyle \frac{\mathrm{\Lambda }_H|\gamma _5Q^{(\mathrm{w})}|\mathrm{\Lambda }_H}{\mathrm{\Lambda }_H}}`$ (32)
where in the last line a bra/ket notation has been used. The large angle brackets indicate the average over an appropriate ensemble of gauge fields.
We define the pion interpolating field as $`\pi ^a(x)i\overline{q}(x)t^a\gamma _5q(x)`$ and then find that the pion two-point function is given by (no sum on $`a`$)
$$\pi ^a(x)\pi ^a(0)=\underset{\mathrm{\Lambda }_H,\mathrm{\Lambda }_H^{}}{}\frac{\mathrm{\Psi }_{\mathrm{\Lambda }_H^{}}^{}(x,s)Q_{s,s^{}}^{(\mathrm{w})}\mathrm{\Psi }_{\mathrm{\Lambda }_H}(x,s^{})\mathrm{\Psi }_{\mathrm{\Lambda }_H}^{}(0,\stackrel{~}{s})Q_{\stackrel{~}{s},\stackrel{~}{s}^{}}^{(\mathrm{w})}\mathrm{\Psi }_{\mathrm{\Lambda }_H^{}}(0,\stackrel{~}{s}^{})}{\mathrm{\Lambda }_H\mathrm{\Lambda }_H^{}}.$$
(33)
Note that the generators, $`t^a`$, do not appear in the spectral sum, since they merely serve to specify the contractions of the quark propagators and that $`\pi ^a(x)=iJ_5^a(x)`$. To investigate the extra term in the axial Ward-Takahashi identity, Eq. 14, we will also have need to measure the correlation function between interpolating pion fields defined on the domain walls and the mid-point contribution to the divergence of the axial current, $`J_{5q}^a`$. We define a mid-point pion interpolating field by $`\pi _{(\mathrm{mp})}^a(x)=iJ_{5q}^a(x)`$ and the spectral decomposition for the correlator between interpolating pion operators on the wall and the midpoint is
$$\pi _{(\mathrm{mp})}^a(x)\pi ^a(0)=\underset{\mathrm{\Lambda }_H,\mathrm{\Lambda }_H^{}}{}\frac{\mathrm{\Psi }_{\mathrm{\Lambda }_H^{}}^{}(x,s)Q_{s,s^{}}^{(\mathrm{mp})}\mathrm{\Psi }_{\mathrm{\Lambda }_H}(x,s^{})\mathrm{\Psi }_{\mathrm{\Lambda }_H}^{}(0,\stackrel{~}{s})Q_{\stackrel{~}{s},\stackrel{~}{s}^{}}^{(\mathrm{w})}\mathrm{\Psi }_{\mathrm{\Lambda }_H^{}}(0,\stackrel{~}{s}^{})}{\mathrm{\Lambda }_H\mathrm{\Lambda }_H^{}}.$$
(34)
We define a local axial current as $`A_\mu ^a(x)\overline{q}(x)t^a\gamma _\mu \gamma _5q(x)`$ and note that it is different from $`𝒜_\mu ^a`$ defined in Eq. 8. The two-point function of the zeroth component of this current, $`A_0^a(x)A_0^a(0)`$, has a form similar to Eq. 33 with a factor of $`\gamma _0`$ multiplying each $`Q^{(\mathrm{w})}`$ and an overall minus sign. Finally, our scalar density is $`\sigma (x)\overline{q}(x)q(x)`$ and the connected correlator $`\sigma (x)\sigma (0)_\mathrm{c}`$ also has the form of Eq. 33 with a factor of $`\gamma _5`$ multiplying each $`Q^{(\mathrm{w})}`$ and an overall minus sign.
## III Hadron Masses for $`m_f0.01`$
In this section we present the results for $`m_\pi `$, $`m_\rho `$ and $`m_N`$ obtained for reasonably heavy input quark mass, $`m_f0.01`$ where the lower limit corresponds to $`m_{\mathrm{quark}}m_{\mathrm{strange}}/4`$. The more challenging study of $`m_\pi `$ for $`m_f0`$ is described later, in Section V. This section is organized as follows. We begin by describing the Monte Carlo runs on which the results in this paper are based. Next the methods used to determine the hadron masses are discussed, both the propagator determinations and our fitting procedures. Finally, we present the results of those calculations for the easier, large mass case, $`m_f0.01`$.
### A Simulation summary
The results reported in this paper were obtained from ensembles of gauge field configuration generated from pure gauge simulations using the standard Wilson action at three values of the coupling parameter, $`\beta =6/g^2`$: 5.7, 5.85 and 6.00. Thus, these ensembles follow the distribution, $`\mathrm{exp}\{6/g^2_𝒫\mathrm{tr}U_𝒫\}`$ where the sum ranges over all elementary plaquettes $`𝒫`$ in the lattice and $`U_𝒫`$ is the ordered product of the four link matrices associated with the edges of the plaquette $`𝒫`$. Some of the $`\beta =5.7`$ simulations and a portion of those at $`\beta =5.85`$ were performed using the hybrid Monte Carlo ‘$`\mathrm{\Phi }`$’ algorithm . These runs were performed on an $`8^3\times 32`$ space-time volume with a domain wall height $`M_5=1.65`$. Each hybrid Monte Carlo trajectory consisted of 50 steps with a step size $`\mathrm{\Delta }t=0.02`$. These runs are summarized in Table I. In each case the first 2,000 hybrid Monte Carlo trajectories were discarded for thermalization before any measurements were made. After these thermalization trajectories, successive measurements of hadron masses and the chiral condensate, $`\overline{q}q`$ were made after each group of 200 trajectories.
A second set of simulations were performed using the heatbath method of Creutz , adapted for $`SU(3)`$ using the two-subgroup technique of Cabibbo and Marinari and improved for a multi-processor machine by the algorithm of Kennedy and Pendleton . The first 5,000 sweeps were discarded for thermalization. These runs are described in Table II where the values of $`M_5`$ used are also given. Finally, the single $`\beta =5.85`$ run with $`M_5=1.9`$ was performed using the MILC code . Here four over-relaxed heatbath sweeps with $`\omega =2`$ were followed by one Kennedy-Pendleton sweep, with 50,000 initial sweeps discarded for thermalization.
A portion of the $`\beta =5.7`$ masses described here appeared earlier in Ref. while the first of the $`\beta =6.0`$ results appear in Refs. and .
### B Mass measurement techniques
We follow the standard procedures for determining the hadron masses from a lattice calculation, extracting these masses from the exponential time decay of Euclidean-space, two-point correlation functions. In our calculation the source may take two forms. The first is a point source
$$O_\mathrm{\Gamma }^a(x)=\overline{q}(x)t^a\mathrm{\Gamma }q(x)$$
(35)
which is usually introduced at the origin. The flavor index $`a`$ is introduced to make clear that we do not study the masses of flavor singlet states. For the nucleon state we use a combination of three quark fields:
$$O_\mathrm{P}(x)=ϵ_{abc}u_a(x)[u_b(x)C\gamma ^5d_c(x)]$$
(36)
where for simplicity we have written the source for a proton in terms of up and down quark fields, $`q=u`$ and $`d`$. Here $`C`$ is the $`4\times 4`$ Dirac charge-conjugation matrix, $`ϵ`$ the anti-symmetric tensor in three dimensions and the color sum over the indices $`a`$, $`b`$ and $`c`$ is shown explicitly. Only these point sources are used in the $`\beta =5.7`$ running.
The second variety of source used in this work is a wall source. Such a source is obtained by a simple generalization of Eqs. 35 and 36 in which we replace the quark fields evaluated at the same space-time point $`x=(\stackrel{}{r},t)`$ with distributed fields, each of which is summed over the entire spatial volume at a fixed time $`t`$. Gauge covariance is maintained by introducing a gauge field dependent color matrix $`V[U](\stackrel{}{r},t_i)`$ which transforms the spatial links in the time slice $`t=t_i`$ into Coulomb gauge. Thus, to construct our wall sources we simply replace the quark field $`q(\stackrel{}{r},t_i)_c`$ by the non-local field
$$q^w(t_i)_c=\underset{\stackrel{}{r}}{}V(\stackrel{}{r},t_i)_{c,c^{}}q(\stackrel{}{r},t_i)_c^{}$$
(37)
where $`c`$ and $`c^{}`$ are color indices. We use these wall sources for the $`\beta =5.85`$ calculations and a combination of both wall and point sources in the $`\beta =6.0`$ studies. The use of wall sources for these weaker coupling runs is appropriate since the physical hadron states are larger in lattice units and better overlap is achieved with the states of interest by using these extended sources.
In all cases we use a zero-momentum-projected point sink for the second operator in the correlation function. This is obtained by simply summing the operators in Eqs. 35 and 36 over all spatial positions $`\stackrel{}{r}`$ in a fixed time plane $`t=t_f`$. Thus, for example, we will extract the mass $`m_\mathrm{\Gamma }`$ of the lightest meson with quantum numbers of the Dirac matrix $`\mathrm{\Gamma }`$ from the large $`t_ft_i`$ expression:
$$\underset{\stackrel{}{r}}{}O_\mathrm{\Gamma }^a(\stackrel{}{r},t_f)\overline{q}^w(t_i)\mathrm{\Gamma }t^aq^w(t_i)A(e^{m_\mathrm{\Gamma }(t_ft_i)}+e^{m_\mathrm{\Gamma }(N_tt_f+t_i)}).$$
(38)
A similar equation is used for the nucleon correlation function except that the second exponent representing the state propagating through the antiperiodic boundary condition connecting $`t=0`$ and $`t=N_t1`$ is reversed in sign and has exchanged upper and lower components for a spinor basis in which $`\gamma ^0`$ is diagonal.
For both the calculation of the quark propagators from which these hadron correlators are constructed and the evaluation of the chiral condensate, $`\overline{q}q`$, we invert the five-dimensional domain wall fermion Dirac operator of Eq. 1, using the conjugate gradient method to solve an equation of the form $`Dy=h`$. This iterative method is run until a stopping condition is satisfied, which requires that the norm squared of the residual be a fixed, small fraction $`ϵ`$ of the norm squared of the source vector $`h`$. At the $`n^{th}`$ iteration, we determine the residual $`r_n`$ as a cumulative approximation to the difference vector obtained by applying the Dirac operator to the present approximate solution $`y_n`$ and $`h`$: $`r_n=Dy_nh`$. We stop the process when $`|r_n|^2/|h|^2<ϵ`$.
For the calculation of $`\overline{q}q`$ we use $`ϵ=10^6`$ for the runs of Table I and $`ϵ=10^8`$ for those in Table II. For the computation of hadron masses in the runs of Table I we use $`ϵ=10^8`$ when $`L_s`$ has the values 10, 16, 24 and 48, the condition $`ϵ=10^{10}`$ for the case $`L_s=32`$. For the hadron masses computed in the runs in Table II we used $`ϵ=10^8`$ for $`\beta =5.7`$ and 6.0, and $`ϵ=10^7`$ for $`\beta =5.85`$. Tests showed that zero-momentum projected hadronic propagators eight time slices from the source, calculated with a stopping condition of $`10^6`$, differed by less than 1% from the same propagators calculated with a stopping condition of $`10^{12}`$ for $`m_f0.01`$ . For a quark mass $`m_f=0.01`$ and a $`16^3\times 32`$ volume with $`L_s=16`$, typically $`1,500`$ conjugate gradient iterations were required to meet the stopping condition. For our very light quark masses ($`m_f0.001`$) up to 10,000 iterations were required for convergence.
The final step in extracting the masses of the lowest-lying hadron states from the exponential behavior of the correlation functions given in Eq. 38 is to perform a fit to this exponential form over a time range chosen so that this single-state description is accurate. Choosing $`t_i=0`$, we use the appropriate $`t_fN_tt_f`$ symmetry of Eq. 38 to fold the correlator data into one-half of the original time range $`0t_f<N_t`$. We then perform a single-state fit of the form in Eq. 38 for the time range $`t_{\mathrm{min}}t_ft_{\mathrm{max}}N_t/2`$. Typically $`t_{\mathrm{max}}`$ is simply set to the largest possible value, $`t_{\mathrm{max}}=N_t/2`$.<sup>*</sup><sup>*</sup>*For the $`\beta =5.7`$ runs we used smaller values of $`t_{\mathrm{max}}`$ for the $`\pi `$ and $`\rho `$ fitting, typically 12 or 14, in order to avoid the effects of rounding errors. These finite-precision errors, caused by a poor choice of initial solution vector, were seen at the largest time separations for the very rapidly falling propagators found at this strong coupling.
The lower limit, $`t_{\mathrm{min}}`$, is decreased to include as large a time range as possible so as to extract the most accurate results. However, $`t_{\mathrm{min}}`$ must be sufficiently large that the asymptotic, single-state formula in Eq. 38 is a good description of the data in the time range studied. These issues are nicely represented by the effective mass, $`m_{\mathrm{eff}}(t)`$, with the parameters $`A`$ and $`mm_{\mathrm{eff}}(t)`$ in Eq. 38 determined to exactly describe the hadron correlator at the times $`t`$ and $`t+1`$. To the extent that $`m_{\mathrm{eff}}(t)`$ is independent of $`t`$, the data are in a time range which is consistent with the desired single state signal. As an illustration, this effective mass is plotted in Figure 6 for the $`\pi `$, $`\rho `$ and nucleon states in the $`16^3\times 32`$, $`\beta =6.0`$, $`L_s=16`$, $`m_f=0.01`$ calculation. Good single-state fits are easy to identify from the plateau regions for the case of $`m_\pi `$ and $`m_\rho `$. For the nucleon the rapidly increasing errors at larger time separations for this relatively light quark mass make it more difficult to determine a plateau. Better nucleon plateaus are seen for larger values of $`m_f`$.
The actual fits are carried out by minimizing the correlated $`\chi ^2`$ to determine the particle mass and propagation amplitude. We then choose $`t_{\mathrm{min}}`$ as small as possible consistent with two criteria. First, the fit must remain sufficiently good that the $`\chi ^2`$ per degree of freedom does not grow above 1-2. Second, we require that the mass values obtained agree with those determined from a larger value of $`t_{\mathrm{min}}`$ within their errors.
In order to keep the fitting procedure as simple and straight forward as possible, we choose values for $`t_{\mathrm{min}}`$ which can be used for as large a range of quark masses, domain wall separations and particle types as possible. Given the large number of Monte Carlo runs and variety of masses and $`L_s`$ values it is possible to employ an essentially statistical technique to determine $`t_{\mathrm{min}}`$. In choosing the appropriate $`t_{\mathrm{min}}`$ we examine two distributions. The first distribution is a simple histogram of values of $`\chi ^2/\mathrm{dof}`$ obtained for all quark masses and a particular physical quantum number. We require that for our choice of $`t_{\mathrm{min}}`$, this distribution is sensibly peaked around the value 1 or lower. An example is shown in Figure 7 for the $`\beta =6.00`$ $`\rho `$ mass determined from a wall source for three values of $`t_{\mathrm{min}}`$: 5, 7 and 9.
In the second distribution we first determine a fitted mass $`m_i(t)`$ and the corresponding error $`\sigma _i(t)`$ for the state $`i`$, where the lower bound on the fitting range is given by $`t`$. We then choose a $`t^{}>t`$ and examine a measure of the degree to which $`m_i(t)`$ and $`m_i(t^{})`$ agree. The measure we choose is
$$\delta _i(t^{},t)=\frac{m_i(t^{})m_i(t)}{\sigma _i(t^{})}.$$
(39)
In Figure 7 we show the distribution of values of $`\delta _i(t^{},t_{\mathrm{min}})`$ for the $`\rho `$ meson for all $`t^{}>t_{\mathrm{min}}`$ and three choices for $`t_{\mathrm{min}}`$: 5, 7 and 9. The distributions include $`\rho `$ mesons with all values of $`m_f0.01`$ and all values for $`L_s`$ used in the calculations.
In our sample Figure 7, we have a reasonable distribution of $`\chi ^2/\mathrm{dof}`$ values for all three choices of $`t_{\mathrm{min}}`$ with only a slight improvement visible as $`t_{\mathrm{min}}`$ increases from 5 to 9. Likewise the distribution of mass values found at $`t^{}>t_{\mathrm{min}}`$ is in reasonable agreement for each value of $`t_{\mathrm{min}}`$ with a slight bias toward larger values being visible at the lowest value $`t_{\mathrm{min}}=5`$. Examining this figure and corresponding figures for the $`\pi `$, for our quoted masses, we chose $`t_{\mathrm{min}}=7`$ for these states. The fact that Figure 7 does not sharply discriminate between these three possible choices of $`t_{\mathrm{min}}`$ implies that we will get essentially equivalent results from each of these three values.
Our choices of $`t_{\mathrm{min}}`$ are as follows. For $`\beta =5.7`$, where only point sources are used, $`t_{\mathrm{min}}`$ was chosen to be 7 for the $`\pi `$, $`\rho `$ and nucleon. For $`\beta =5.85`$, hadron masses were determined only from the doubled $`12^3\times 64`$ configurations using wall sources and the value $`t_{\mathrm{min}}=6`$ for the $`\rho `$ and 7 for the $`\pi `$ and nucleon. Finally for $`\beta =6.0`$ the most accurate mass values were determined using wall sources and it is these mass results which we quote below. Here $`t_{\mathrm{min}}`$ was chosen to be 7 for the $`\pi `$ and $`\rho `$ and 8 for the nucleon. We were able to extract quite consistent results with larger errors using point sources. Here the needed value of $`t_{\mathrm{min}}`$ was 10 for the $`\pi `$ and $`\rho `$ and $`t_{\mathrm{min}}=8`$ for the nucleon. Finally, the errors are determined for each mass by a jackknife analysis performed on the resulting fitted mass.
### C Hadron mass results
The hadron masses that result from the fitting procedures described above are given in Tables IV-XV. Omitted from this tabulation are the masses for the more difficult cases $`m_f=0.0`$ and 0.001 which are discussed later in Section V. In each case the pion mass was determined from the $`A_0^a(x)A_0^a(0)`$ correlator. While the results presented in these tables will be used in later sections of this paper, there are some important aspects of these results which will be discussed in this section. In particular, the dependence on volume and the $`m_f`$ dependence of the $`\rho `$ and nucleon masses will be examined.
We begin by examining the dependence of the $`\rho `$ and nucleon on the input quark mass, $`m_f`$. In Figures 8, 9 and 10 we plot the $`\rho `$ and nucleon masses as a function of $`m_f`$, As the figures show, each case is well described by a simple linear dependence on $`m_f`$. The data plotted in these figures appear in Tables VII XII and XIII, respectively. Also plotted in Figure 10 are our results for $`m_\rho `$ with non-degenerate quarks. The coincidence of these two results implies the familiar conclusion that to a good approximation the meson mass depends on the simple average of the quark masses of which it is composed. For simplicity in obtaining jackknife errors, we have included in these linear fits only that data associated with ensembles of configurations on which all relevant quark mass values were studied. Added configurations where only particular quark masses had been evaluated were not included.
A simple linear fit provides a good approximation to all the masses considered in this section, in particular for $`m_f0.01`$. In Table XVI we assemble the fit parameters for the $`\beta =5.7`$, $`8^3\times 32`$ masses, while Tables XVII, XVIII and XIX contain the fit parameters for the $`\beta =5.7`$, $`16^3\times 32`$, $`\beta =5.85`$, $`12^3\times 32`$ and $`\beta =6.0`$, $`16^3\times 32`$ calculations, respectively. The parameters presented in these three tables were obtained by minimizing a correlated $`\chi ^2`$ which incorporated the effects of the correlation between hadron masses obtained with different valence quark masses, $`m_f`$, but determined on the same ensemble of quenched gauge configurations. The errors quoted follow from the jackknife method and the small values of $`\chi ^2/\mathrm{dof}`$ shown demonstrate how well these linear fits work. Because of the visible curvature in the pion mass for our $`\beta =5.7`$ and 6.0 results, the linear fits for $`m_\pi ^2`$ were made to the lowest three mass values. For the $`\rho `$ and nucleon and all three masses at $`\beta =5.85`$ we fit to the masses obtained for the full range of $`m_f`$ values.
Next we consider the effects of finite volume by comparing the $`8^3\times 32`$ and $`16^3\times 32`$, volumes used in the $`\beta =5.7`$, $`L_s=48`$ calculation. The value of $`m_\pi =0.383(4)`$ found at the lightest $`m_f=0.02`$ mass value for the $`16^3\times 32`$ implies a Compton wavelength of 2.6 in lattice units. This lies between 1/4 and 1/3 of the linear dimension of the smaller lattice, suggesting that we should not expect large finite volume effects. This is borne out by comparing the data in Tables VIII and XI where the two sets of masses agree within errors.
This apparent volume independence within our errors can be nicely summarized by comparing the coefficients of the linear fits of the $`\rho `$ and nucleon. Writing the two $`a`$ and $`b`$ coefficients from the tables as a pair \[a,b\], we can compare the $`16^3\times 32`$ values from Table XVII $`[0.775(18),2.20(7)]`$ and $`[1.03(4),4.13(17)]`$ for the $`\rho `$ and nucleon with the corresponding numbers for the $`8^3\times 32`$ numbers from Table XVI: $`[0.790(13),2.18(5)]`$ and $`[1.13(4),3.79(15)]`$. For $`m_\pi `$ the results on the two volumes agree to within the typical 1% statistical errors. However, for the case of the $`\rho `$ and nucleon masses, finite volume effects may be visible on the two standard deviation or 1-2% level for the more accurate masses obtained for $`m_f0.06`$
Since in lattice units the $`\rho `$ mass decreases by about a factor of two as we change $`\beta `$ from 5.7 to 6.0, the $`16^3`$ spatial volume used at $`\beta =6.0`$ should be equivalent to the $`8^3`$ volume just discussed at $`\beta =5.7`$. Thus, we expect that the $`\rho `$ and nucleon masses that we have found on this $`16^3`$ volume will differ from their large volume limits by an amount on the order of a few percent while the finite-volume pion masses may be accurate on the 0.5% level.
## IV Zero modes and the chiral condensate
### A Banks-Casher formula for domain wall fermions
In the previous section, our results for quark masses $`m_f0.01`$ were given, where the smallest values of $`m_f`$ gave $`m_\pi /m_\rho 0.4`$. Since the domain wall fermion operator with $`m_f=0`$ should give exact fermionic zero modes as $`L_s\mathrm{}`$, observables determined from quark propagators at finite $`L_s`$, when small quark masses are used, should show the effects of topological near-zero modes. For quenched simulations, where zero modes are not suppressed by the fermion determinant, these modes can be expected to produce pronounced effects. One important practical question is the size of the quark mass where the effects are measurable. To begin to investigate this we now turn to the simplest observable where they can occur, $`\overline{q}q`$.
Before considering the domain wall fermion operator, we review the spectral decomposition of the continuum four-dimensional, anti-hermitian Euclidean Dirac operator $`\text{/}D^{(4)}`$.The naive lattice fermion operator $`\text{/}D=\gamma _u(U_{x,\mu }\delta _{x+\widehat{\mu },x^{}}U_{x^{},\mu }^{}\delta _{x\widehat{\mu },x^{}})`$ and the lattice staggered fermion operator have eigenvalues and eigenvectors which also obey Eq. 41. The eigenfunctions and corresponding eigenvalues of such an anti-hermitian operator satisfy
$$(\text{/}D^{(4)}+m)\psi _\lambda =(i\lambda +m)\psi _\lambda .$$
(40)
with $`\lambda `$ real and
$$\gamma _5\psi _\lambda =\{\begin{array}{cc}\pm \psi _\lambda & \lambda =0\\ \text{ }\psi _\lambda & \lambda 0.\end{array}$$
(41)
(We use $`\lambda `$ to label eigenfunctions and eigenvalues of the anti-hermitian operator, saving $`\lambda _H`$ for the “hermitian” case defined below.) In the continuum, the presence of zero modes is guaranteed by the Atiyah-Singer index theorem for a gluonic field background with non-zero winding number .
The four-dimensional quark propagator, $`S_{x,y}^{(4)}`$, can be written as
$$S_{x,y}^{(4)}=\underset{\lambda }{}\frac{\psi _\lambda (x)\psi _\lambda ^{}(y)}{i\lambda +m}$$
(42)
leading directly to the Banks-Casher relation (with our normalization for the chiral condensate)
$`\overline{\psi }\psi `$ $`=`$ $`{\displaystyle \frac{1}{12V}}{\displaystyle \frac{|\nu |}{m}}+{\displaystyle \frac{m}{12V}}{\displaystyle \underset{\lambda 0}{}}{\displaystyle \frac{1}{\lambda ^2+m^2}}`$ (43)
$`=`$ $`{\displaystyle \frac{2m}{12}}{\displaystyle _0^{\mathrm{}}}𝑑\lambda {\displaystyle \frac{\rho (\lambda )}{\lambda ^2+m^2}}`$ (44)
where $`\nu `$ is the winding number and $`\rho (\lambda )`$ is the average density of eigenvalues. For quenched QCD, $`\rho (\lambda )`$ has no dependence on the quark mass. For both quenched and full QCD, one expects that $`|\nu |\sqrt{V}`$, as is the case for a dilute instanton gas model. Thus, zero modes lead to a divergent $`1/m`$ term in $`\overline{q}q`$ whose coefficient decreases as $`1/\sqrt{V}`$. (This contrasts with the behavior seen above the deconfinement transition where it can be shown that the $`1/m`$ term remains non-zero for quenched QCD in the infinite volume limit .) Before discussing the results of our simulations, we first address how this simple expectation of a $`1/m`$ term in $`\overline{q}q`$ due to zero modes should appear for the domain wall fermion operator.
We will find it useful to compare the spectrum and properties of the hermitian domain wall fermion operator $`D_H`$ with the hermitian four-dimensional operator, $`D_H^{(4)}`$, defined by
$$D_H^{(4)}=\gamma _5(\text{/}D^{(4)}+m).$$
(45)
The eigenvalues, $`\lambda _H`$, and eigenvectors, $`\psi _{\lambda _H}`$, for this operator can be given in terms of $`\lambda `$ and $`\psi _\lambda `$ given above. If $`\lambda =0`$ we immediately get an eigenvalue $`\lambda _H=\pm m`$ for the hermitian operator, and an eigenvector with the definite chirality $`+1`$ or $`1`$. For $`\lambda 0`$, the eigenvectors of $`D_H^{(4)}`$ are linear combinations of $`\psi _\lambda `$ and $`\psi _\lambda `$ and the corresponding eigenvalues are $`\lambda _H=\pm \sqrt{\lambda ^2+m^2}`$. Since $`(\text{/}D^{(4)}+m)^1=(D_H^{(4)})^1\gamma _5`$, we have
$`\overline{\psi }\psi `$ $`=`$ $`{\displaystyle \frac{1}{12V}}\mathrm{Tr}(\text{/}D^4+m)^1`$ (46)
$`=`$ $`{\displaystyle \frac{1}{12V}}\mathrm{Tr}(\gamma _5(D_H^{(4)})^1)`$ (47)
$`=`$ $`{\displaystyle \frac{1}{12V}}{\displaystyle \underset{x,\lambda _H}{}}{\displaystyle \frac{\psi _{\lambda _H}(x)^{}\gamma _5\psi _{\lambda _H}(x)}{\lambda _H}}.`$ (48)
Since for $`D_H^{(4)}`$
$$\underset{x}{}\psi _{\lambda _H}^{}(x)\gamma _5\psi _{\lambda _H}(x)=\frac{m}{\lambda _H}$$
(49)
Eq. 48 also reduces to the Banks-Casher relation, Eq. 44. For finite mass, the zero-mode hermitian eigenfunctions are chiral, while other eigenfunctions have a chirality proportional to the mass. This will be important in our comparisons with domain wall fermions.
For large $`L_s`$, it is expected that the spectrum of light eigenvalues of the hermitian domain wall fermion operator, $`D_H`$, should reproduce the features of the operator $`D_H^{(4)}`$. Since $`D_H`$ depends continuously on $`m_f`$, for small $`m_f`$ its $`i`$th eigenvalue must have the form
$$\mathrm{\Lambda }_{H,i}^2=a_i^{}+b_i^{}m_f+c_i^{}m_f^2+\mathrm{}.$$
(50)
To make a connection with the normal continuum form for the eigenvalues we reparameterize $`\mathrm{\Lambda }_{H,i}^2`$ as
$$\mathrm{\Lambda }_{H,i}^2=n_{5,i}^2(\lambda _i^2+(m_f+\delta m_i)^2)+\mathrm{}.$$
(51)
Here $`n_{5,i}`$ is an overall normalization factor and we have defined $`\delta m_i`$, which enters as a contribution to the total quark mass for the $`i`$th eigenvalue. For $`\delta m_i=m_f`$, $`\mathrm{\Lambda }_{H,i}^2`$ is at its minimum. Modes which become precise zero modes when $`L_s\mathrm{}`$ will have non-zero values for $`\lambda _i`$ and $`\delta m_i`$ for finite $`L_s`$. We will refer to such modes as topological near-zero modes.
From perturbation theory in $`m_f`$, one can easily see that
$$\frac{d\mathrm{\Lambda }_{H,i}}{dm_f}=\mathrm{\Lambda }_{H,i}|\gamma _5Q^{(\mathrm{w})}|\mathrm{\Lambda }_{H,i}$$
(52)
while the chain rule applied to Eq. 51 gives
$$\frac{d\mathrm{\Lambda }_{H,i}(m_f)}{dm_f}=\frac{n_{5,i}^2(m_f+\delta m_i)}{\mathrm{\Lambda }_{H,i}}.$$
(53)
Combining this with Eq. 32 gives
$$\overline{q}q=\frac{1}{12V}\underset{i}{}\frac{m_f+\delta m_i}{\lambda _i^2+(m_f+\delta m_i)^2}$$
(54)
which agrees with the Banks-Casher form, Eq. 44 with the addition of the $`i`$ dependent mass contribution $`\delta m_i`$. Thus, the parameter $`\lambda _i`$ in Eq. 51 should be identified with the eigenvalues of the continuum anti-hermitian operator $`\text{/}D^{(4)}`$. As indicated by Eqs. 52 and 53, $`\delta m_i`$ should represent a contribution to the eigenvalue from the chiral symmetry breaking effects of coupling of the domain walls, present for finite $`L_s`$.
These arguments show that the domain wall fermion chiral condensate will grow as $`1/m_f`$ for gauge field configurations with topology, provided $`L_s`$ is large enough to make $`\delta m_i`$ and $`\lambda _i`$ small. The continuum expectation of a $`1/m_f`$ divergence is modified at small $`m_f`$ by the non-zero values of $`\delta m_i`$ and $`\lambda _i`$ for topological near-zero modes. For a single configuration, the precise departure from a $`1/m_f`$ divergence is dominated by the eigenvalues with the smallest values for $`\delta m_i`$ and $`\lambda _i`$; for an ensemble average, the departure from $`1/m_f`$ behavior depends on the distribution of values of $`\delta m_i`$. With this understanding of $`\overline{q}q`$ for domain wall fermions, we turn to our simulation results.
### B Quenched measurements of $`\overline{q}q`$
In this section we discuss our results for $`\overline{q}q`$ for quenched QCD simulations with domain wall fermions. Tables I, II and III give details about the runs where $`\overline{q}q`$ was measured. The most important aspect of the run parameters is the small values for $`m_f`$ used, including $`m_f=0.0`$ where finite $`L_s`$ keeps $`\mathrm{\Lambda }_H`$ non-zero, allowing the conjugate gradient inverter to be used. Of course the number of conjugate gradient iterations becomes quite large.
Equation 54 shows that we should expect large values for $`\overline{q}q`$ for small $`m_f`$ for configurations with topological near-zero modes. Figure 11 shows $`\overline{q}q`$ for $`8^3\times 32`$ lattices at $`\beta =5.7`$ with both $`L_s=32`$ and 48. The quark masses used cover the ranges $`00.04`$ and $`0.000250.008`$, defined in Table III. Both values for $`L_s`$ show an increase in $`\overline{q}q`$ for very small quark mass, an effect expected from the presence of a non-zero value for $`|\nu |`$. (This effect was first reported for domain wall fermions based on quenched simulations done on $`8^3\times 32`$ lattices with $`\beta =5.85`$, $`M_f=1.65`$ and $`L_s=32`$ and listed in Table I .)
Motivated by the form of Eq. 54 we have fit $`\overline{q}q`$ to the following phenomenological form
$$\overline{q}q=\frac{a_1}{m_f+\delta m_{\overline{q}q}}+a_0+a_1m_f$$
(55)
where $`a_1`$, $`a_0`$, $`a_1`$ and $`\delta m_{\overline{q}q}`$ are parameters to be determined. $`\delta m_{\overline{q}q}`$ represents a weighted average of $`\delta m_i`$ over the eigenvalues which dominate $`\overline{q}q`$ for small $`m_f`$. The measurements of $`\overline{q}q`$ for different values of $`m_f`$ are strongly correlated, being done on the same gauge field configurations with, generally, the same random noise estimator used to determine $`\overline{q}q`$ for all the masses. The common noise source makes the signal for the $`1/m_f`$ divergence particularly clean, since the overlap of the topological near-zero mode eigenvectors with the random source does not fluctuate on a single configuration. This strong correlation precludes doing a correlated fit of $`\overline{q}q`$ to $`m_f`$, since the correlation matrix is too singular. Thus, the fits in this section are uncorrelated fits of $`\overline{q}q`$ to $`m_f`$.
Table XX gives the results for fits to the form of Eq. 55 for our $`\beta =5.7`$, 5.85 and 6.0 simulations. All the fits have a value $`\chi ^2/\mathrm{dof}`$ less than 0.1, a consequence of doing uncorrelated fits to such correlated data. In Figure 11, one sees that the fit represents the data quite well. Continuing with $`8^3\times 32`$ lattices at $`\beta =5.7`$, Table XX shows the fit parameters are very similar for $`L_s=32`$ and 48, except for $`\delta m_{\overline{q}q}`$, which drops from 0.0040(4) to 0.0017(2). This indicates a decrease in $`\delta m_i`$ as $`L_s`$ increases.
Figure 12 is a similar plot of $`\overline{q}q`$ for $`16^3\times 32`$ lattices with $`\beta =6.0`$ for $`L_s=16`$ and 24. The rise in $`\overline{q}q`$ for small $`m_f`$ exhibits the same general structure as for the $`\beta =5.7`$ data in Figure 11, but the effect is larger. Here $`\delta m_{\overline{q}q}`$ falls from 0.00056(3) for $`L_s=16`$ to 0.00011(1) for $`L_s=24`$.
To further demonstrate that the divergence for small $`m_f`$ is due to eigenfunctions of $`D_H`$ that represent zero modes of a definite chirality, Figure 13 shows the evolution of both $`\overline{q}q`$ (solid lines) and $`\overline{q}\gamma _5q`$ (dotted lines). These evolutions are for $`16^3\times 32`$ lattices at $`\beta =6.0`$ with $`L_s=16`$. Eigenfunctions with a positive chirality contribute equally to $`\overline{q}q`$ and $`\overline{q}\gamma _5q`$, while negative chirality eigenfunctions contribute with an opposite sign to $`\overline{q}\gamma _5q`$. The topological near-zero modes should be approximately chiral and, for smaller values of $`m_f`$, one see large fluctuations in $`\overline{q}q`$ and $`\overline{q}\gamma _5q`$. Some of the fluctuations have the same sign and some are of opposite sign. Thus, we have configurations with eigenfunctions which are very good approximations to the exact zero modes expected as $`L_s\mathrm{}`$.
As mentioned earlier, $`|\nu |/V`$ should decrease with volume, with the asymptotic dependence given by $`1/\sqrt{V}`$. To investigate this numerically, we have measured $`\overline{q}q`$ on both $`8^3\times 32`$ and $`16^3\times 32`$ lattices at $`\beta =5.7`$ and 5.85 with $`L_s=32`$ and show the $`\beta =5.7`$ results in Figure 1. The graph clearly shows that the $`1/m_f`$ divergence is drastically suppressed by the larger volume. The coefficient of the $`1/m_f`$ term falls from $`6.0(6)\times 10^6`$ to $`2.5(4)\times 10^6`$ as the volume is changed by a factor of 8. This may be somewhat misleading, since $`\delta m_{\overline{q}q}`$ also changes by a factor of about 2, likely due to the phenomenological nature of the fit and the small effects of the $`1/m_f`$ pole for the larger volume. Putting aside this systematic difficulty, the $`1/m_f`$ coefficient decreases by a factor of $`1/\sqrt{5.8}`$, showing the general behavior expected but not in precise agreement with the expected asymptotic form. For $`\beta =5.85`$, where the physical size of the lattices is smaller, the $`1/m_f`$ coefficient falls from $`3.8(3)\times 10^6`$ to $`0.60(9)\times 10^6`$, a factor of 6.3. We have not seen the expected $`1/\sqrt{V}`$ dependence for the $`1/m_f`$ coefficient, but it does decrease with volume in accordance with general ideas. It is possible that on the larger $`16^3\times 32`$ volume, the $`1/m_f`$ rise is not large enough to allow its coefficient to be determined without systematic errors.
Thus, we have clear evidence for topological near-zero modes in our quenched simulations using domain wall fermions. They are revealed through a large $`1/m_f`$ rise in our values for $`\overline{q}q`$, the presence of configurations where $`\overline{q}q`$ and $`\overline{q}\gamma _5q`$ are large and of opposite sign and the volume dependence of the coefficient of the $`1/m_f`$ term. We have extracted a quantity, $`\delta m_{\overline{q}q}`$, from a phenomenological fit to $`\overline{q}q`$, which represents the effects of finite $`L_s`$ on the eigenmodes with small eigenvalues which dominate $`\overline{q}q`$ for $`m_f0`$. Physical values for $`\overline{q}q`$ in the chiral limit, without the contribution of the topological near-zero modes, will be presented in Section VII. We now turn to a discussion of how these zero modes, and the expected light modes responsible for chiral symmetry breaking, are evident in measurements of the pion mass.
## V The pion mass in the chiral limit
For domain wall fermions with $`L_s=\mathrm{}`$, the chiral limit is achieved by taking $`m_f=0`$. For our quenched simulations at finite $`L_s`$, we must investigate the chiral limit in detail to demonstrate that the changes from the $`L_s\mathrm{}`$ limit are under control and of a known size. As is discussed in Sec. II B, for low energy QCD physics the dominant effect of finite $`L_s`$ should be the appearance of an additional chiral symmetry breaking term in the effective Lagrangian describing QCD. This term has the form $`m_{\mathrm{res}}\overline{q}(x)q(x)`$ and in the continuum limit its presence will make $`m_\pi `$ vanish at $`m_f=m_{\mathrm{res}}`$ up to terms of order $`a^2`$. Our investigation of the chiral limit is made more difficult since there are other issues affecting this limit, beyond having $`L_s`$ finite. For domain wall fermion quenched simulations, the chiral limit may be distorted by:
1. Order $`a^2`$ effects. Since we are working at finite lattice spacing chiral symmetry will not be precisely restored even for $`m_f+m_{\mathrm{res}}=0`$. In particular, additional chiral symmetry breaking will come from the effects of higher dimension operators suppressed by factors of $`a^n`$ for $`n2`$. Thus, we cannot not expect $`m_\pi `$ to vanish precisely at the point $`m_f=m_{\mathrm{res}}`$, but perhaps at a nearby point, removed from $`m_{\mathrm{res}}`$ by a terms of $`O(a^2)`$.
2. Finite $`L_s`$. The residual mass, $`m_{\mathrm{res}}`$, should represent the finite $`L_s`$ effects for physics describable by a low-energy effective Lagrangian. However, there will be additional effects of finite $`L_s`$ for observables sensitive to ultraviolet phenomena. Further, a quantity with sufficiently severe infrared singularity may show unphysical sensitivity to those $`L_s`$-dependent eigenfunctions $`|\mathrm{\Lambda }_{H,i}`$ (and the parameters $`n_{5,i}`$, $`\lambda _i`$ and $`\delta m_i`$ of the previous section) with small eigenvalues $`\mathrm{\Lambda }_{H,i}`$.
3. Topological near-zero modes. The previous section has shown these dominate $`\overline{q}q`$ for small quark masses ( $`m_f0.01`$ ) for the volumes we are using. From the Ward-Takahashi identity, these effects must also by present in the pion correlator $`\pi ^a(x)\pi ^a(0)`$.
4. Finite volume. For staggered fermions, where the remnant chiral symmetry at finite lattice spacing requires $`m_\pi ^2=0`$ when the input quark mass is zero, the finite volumes used in simulations have been seen to make $`m_\pi ^2`$ non-zero when extrapolated to the chiral limit from above . Such an effect may also be expected to occur for domain wall fermions.
5. Analytic results argue for the presence of “quenched chiral logs” with the dependence of $`m_\pi ^2`$ on the quark mass in quenched QCD different from that of full QCD .
In this section we study the pion mass in the limit of small quark mass. Demonstrating consistent chiral behavior for the pion mass in the limit $`m_f+m_{\mathrm{res}}0`$ is a critical component in establishing the ability of the domain wall fermion formalism to adequately describe chiral physics. If we discover that the limit $`m_f+m_{\mathrm{res}}0`$ is obscured by large $`O(a^2)`$ effects or large violations of chiral symmetry caused by unanticipated propagation between the domain walls, little may be gained from this new formalism. For the $`\rho `$ and nucleon masses reported in Section III, the masses were shown to be well fit by a linear dependence on the input quark mass, $`m_f`$. Any possible non-linearities are not resolvable within our statistics. For the pion, the statistical errors for these values of $`m_f`$ are smaller and we have also run simulations at smaller values for $`m_f`$ so we might hope to learn more about this important quantity. We begin by investigating the effects of topological near-zero modes on the pion.
### A Topological near-zero mode effects on the pion: analytic considerations
We have seen that topological near-zero modes dominate $`\overline{q}q`$ for small $`m_f`$ and, by continuity, they will also alter the value for $`\overline{q}q`$ determined with larger quark masses. Through the Ward-Takahashi identity, these modes also appear in the pion correlator, $`\pi ^a(x)\pi ^a(0)`$ and therefore can enter in the determination of the pion mass in a lattice simulation. Alternatively, the axial-vector correlator can be used to measure the pion mass and the zero modes may affect this correlator differently. It is vital to understand the role of these topological near-zero modes, since a study of the chiral limit of $`m_\pi ^2`$ depends on an accurate measurement of the mass of the pion state. In this section we will study the way in which topological zero-modes might be expected to effect pion correlation functions for the continuum theory using our results as a guide to the study of the domain wall amplitudes.
Before proceeding, we first establish our notation for susceptibilities and the integrated Ward-Takahashi identity. In general we define
$$\chi _{CD}\frac{1}{12}\underset{x}{}C(x)D(0)$$
(56)
where $`C`$ and $`D`$ are any two hadronic interpolating fields. In particular
$`\chi _{\pi \pi }`$ $``$ $`{\displaystyle \frac{1}{12}}{\displaystyle \underset{x}{}}\pi ^a(x)\pi ^a(0)`$ (57)
$`\chi _{AA}`$ $``$ $`{\displaystyle \frac{1}{12}}{\displaystyle \underset{x}{}}A_0^a(x)A_0^a(0)`$ (58)
$`\chi _{\pi _{(\mathrm{mp})}\pi }`$ $``$ $`{\displaystyle \frac{1}{12}}{\displaystyle \underset{x}{}}\pi _{(\mathrm{mp})}^a(x)\pi ^a(0)`$ (59)
where no sum over $`a`$ is intended and the factor of $`1/12`$ has been introduced to maintain consistency with our somewhat unconventional normalization for the chiral condensate given in Eq. 29. Then the Ward-Takahashi identity, Eq. 14, with $`O(0)=\pi ^a(0)`$ and summed over $`x`$ becomes
$$m_f\chi _{\pi \pi }+\chi _{\pi _{(\mathrm{mp})}\pi }=\overline{q}q$$
(60)
which we will refer to as the integrated Ward-Takahashi identity.
We first consider Eq. 60 for large $`L_s`$, where we should recover the continuum version of the identity. To simplify the presentation, we start with the notation of Section IV A for the continuum four-dimensional anti-hermitian Dirac operator. We immediately deduce from Eq. 60 that a $`1/m`$ divergence in $`\overline{q}q`$ from topological zero modes dictates a $`1/m^2`$ divergence in $`\chi _{\pi \pi }`$. In addition, $`\chi _{\pi \pi }`$ should have a $`1/m`$ divergence for large volumes from the pion pole and, as we will see below, there can also be a $`1/m`$ pole from topological zero modes. However, the volume dependence of these various pole terms should be different. Pole terms from topological near-zero modes should have a coefficient which is $`O(V^{1/2})`$ for large volumes, while the $`1/m`$ term from the pion pole should be volume independent,
Thus, we expect
$$\chi _{\pi \pi }=\frac{1}{V^{1/2}}\frac{a_2}{m^2}+\frac{a_1}{m}+O(m^0).$$
(61)
The coefficients $`a_2`$ and $`a_1`$ should become volume independent in the infinite volume limit. However, the “pion pole” piece, $`a_1`$, may contain an additional $`1/V^{1/2}`$ term arising from zero modes. Note, a particular order of limits must be understood when interpreting Eq. 61. One expects that the usual relation $`m_\pi ^2m`$ will hold only when $`m1/V|\overline{q}q|`$ . Although this prevents our taking the $`m0`$ limit of Eq. 61, it is fully consistent with the domain $`m1/V^{1/2}`$ where the $`1/m^2`$ term in Eq. 61 may be as large as or much larger than the conventional $`1/m`$ term coming from the pion. For domain wall fermions at finite $`L_s`$ these pole terms will be rendered less singular by the presence of the $`\delta m_i`$ terms in the eigenvalues for $`D_H`$.
Lattice measurements of the pion mass come from the exponential decay of a correlator like $`\pi ^a(x)\pi ^a(0)`$ in the limit of large $`|x|`$. Having examined the zero mode effect in the somewhat simpler susceptibilities, we will now investigate the topological zero mode contributions to two-point functions from their spectral decomposition to understand how zero modes can distort measurements of the pion mass. We have
$`\pi ^a(x)\pi ^a(0)`$ $`=`$ $`\mathrm{Tr}[S_{x,0}^{(4)}\gamma _5S_{0,x}^{(4)}\gamma _5]`$ (62)
$`=`$ $`{\displaystyle \underset{\lambda ,\lambda ^{}}{}}{\displaystyle \frac{\psi _\lambda ^{}(x)\psi _\lambda ^{}(x)\psi _\lambda ^{}^{}(0)\psi _\lambda (0)}{(i\lambda +m)(i\lambda ^{}+m)}}.`$ (63)
First we consider the terms in the sum where both $`\lambda `$ and $`\lambda ^{}`$ are zero. This gives a $`1/m^2`$ pole in $`\pi ^a(x)\pi ^a(0)`$, provided the eigenfunctions in the numerator are non-zero at $`x`$ and 0. (Integrating $`\pi ^a(x)\pi ^a(0)`$ over $`x`$ shows that these topological near-zero modes give the $`1/m^2`$ contribution to $`\chi _{\pi \pi }`$.) The terms in the sum where neither $`\lambda `$ or $`\lambda ^{}`$ are zero should include the small eigenvalues which are responsible for the Goldstone nature of the pion. For large $`|x|`$, the total contribution to $`\pi ^a(x)\pi ^a(0)`$ from these modes should be proportional to
$$|0|\pi (0)|\pi |^2\frac{e^{m_\pi |x|}}{m_\pi }\frac{1}{\sqrt{m}}e^{m_\pi |x|}.$$
(64)
(Integrating over $`x`$ gives another factor of $`m_\pi `$ in the denominator, which produces the $`1/m`$ pion pole in $`\chi _{\pi \pi }`$.) Lastly, the terms with either $`\lambda `$ or $`\lambda ^{}`$ zero, but not both, can be written as
$$\underset{\lambda >0,\lambda ^{}=0}{}\frac{\psi _\lambda ^{}(x)\psi _\lambda ^{}(x)\psi _\lambda ^{}^{}(0)\psi _\lambda (0)}{\lambda ^2+m^2}<\underset{\lambda 0}{}\frac{1}{\lambda ^2+m^2}=\frac{\overline{q}q_{\mathrm{nz}}}{m}$$
(65)
where $`\overline{q}q_{\mathrm{nz}}`$ is the chiral condensate measured without zero mode contributions. Here we have used the symmetries in Eq. 41 to combine the $`\pm \lambda `$ terms in the sum over $`\lambda `$ and remove the term odd in $`i\lambda `$. Since $`\overline{q}q_{\mathrm{nz}}`$ should be non-zero as $`m0`$, we see that the contribution to the correlator from terms with zero modes in one of the propagators can produce at most a $`1/m`$ pole term in $`\pi ^a(x)\pi ^a(0)`$.
Thus, we expect
$$\pi ^a(x)\pi ^a(0)=\frac{1}{V^{1/2}}\left(\frac{c_2(x,0)}{m^2}+\frac{c_1(x,0)}{m}\right)+c_{1/2}\frac{e^{m_\pi |x|}}{\sqrt{m}}+\mathrm{}$$
(66)
for small $`m`$. The first two terms represent the possible zero mode contributions. It is important to note that $`c_2(x,0)`$ gets contributions from the modulus squared of the zero-mode eigenfunctions at the points 0 and $`x`$, while $`c_1(x,0)`$ does not. In particular, for a configuration with a single zero mode, $`c_2(x,0)`$ is positive definite, being given by
$$c_2(x,0)=V^{1/2}|\psi _0(x)|^2|\psi _0(0)|^2.$$
(67)
Thus, one could expect $`c_2(x,0)`$ to be a number of order the inverse of the mean zero-mode size squared, while $`c_1(x,0)`$ could be much smaller due to the terms of differing sign appearing in the sum over eigenmodes.
For large enough $`|x|`$, only the true pion state should contribute. Such large $`|x|`$ requires a correspondingly large $`V`$ with the necessary suppression of zero modes. However, at a fixed separation $`|x|`$ in a finite volume and for simulations with small enough $`m`$, the physical pion contribution to $`\pi ^a(x)\pi ^a(0)`$ can be completely negligible. For finite $`L_s`$, the domain wall fermion spectral form, Eq. 33, gives the precise role of the topological near-zero modes. The double sum over $`\mathrm{\Lambda }_H`$ and $`\mathrm{\Lambda }_H^{}`$ decomposes as we have done above for $`D^{(4)}`$ and the dominant contribution of the topological near-zero modes enters as $`1/(\lambda _i^2+(m_f+\delta m_i)^2)`$ provided $`\mathrm{\Lambda }_H|Q^{(\mathrm{w})}|\mathrm{\Lambda }_H^{}`$ is well approximated by $`O(1)\delta _{\mathrm{\Lambda }_H,\mathrm{\Lambda }_H^{}}`$. Thus, for $`\lambda _i`$ and $`\delta m_i`$ small, there should be a region in $`m_f`$ where $`\pi ^a(x)\pi ^a(0)`$ displays a $`1/m_f^2`$ character.
The pion mass can also be measured from the axial current correlator, $`A_0^a(x)A_0^a(0)`$. The susceptibility for this correlator, $`\chi _{AA}`$, is not constrained by the integrated Ward-Takahashi identity as is $`\chi _{\pi \pi }`$. However, there must be a pion pole contribution in addition to any zero mode terms. Therefore
$`\chi _{AA}`$ $``$ $`{\displaystyle \frac{0|A_0(0)|\pi |^2}{m_\pi ^2}}+O(m^n)\mathrm{zero}\mathrm{mode}\mathrm{poles}`$ (68)
$`=`$ $`O(m^0)+O(m^n)\mathrm{zero}\mathrm{mode}\mathrm{poles}`$ (69)
where we have used $`0|A_0(0)|\pi m_\pi `$. The physical pion contribution is independent of $`m`$ for small $`m`$, which is to be compared with the $`1/m`$ contribution of the physical pion to $`\chi _{\pi \pi }`$. We now turn to the question of the zero mode contribution.
Once again we consider the $`L_s=\mathrm{}`$ case and use the notation for $`D^{(4)}`$. For the axial vector correlation function, the spectral decomposition is
$`A_0^a(x)A_0^a(0)`$ $`=`$ $`\mathrm{Tr}[S_{x,0}^{(4)}\gamma _0\gamma _5S_{0,x}^{(4)}\gamma _0\gamma _5]`$ (70)
$`=`$ $`{\displaystyle \underset{\lambda ,\lambda ^{}}{}}{\displaystyle \frac{\psi _\lambda ^{}(x)\gamma _0\psi _\lambda ^{}(x)\psi _\lambda ^{}^{}(0)\gamma _0\psi _\lambda (0)}{(i\lambda +m)(i\lambda ^{}+m)}}.`$ (71)
The terms in the sum where both $`\lambda `$ and $`\lambda ^{}`$ are zero modes vanish here, since the zero modes have a definite chirality and $`\gamma _0`$ couples different chiral components. (On a given configuration, all the exact zero modes must have the same chirality since exact zero modes can only occur through the index theorem.) Thus, there are no $`1/m^2`$ terms in $`A_0^a(x)A_0^a(0)`$. Note that a $`1/m`$ contribution to $`A_0^a(x)A_0^a(0)`$ can appear from terms in the sum with either $`\lambda `$ or $`\lambda ^{}`$ a zero mode (as we saw for $`\pi ^a(x)\pi ^a(0)`$). The size of such a contribution depends on the matrix element of $`\gamma _0`$ between eigenfunctions.
The terms with neither $`\lambda `$ nor $`\lambda ^{}`$ a zero mode give the physical pion contribution, which should have the form
$$m_\pi ^2\frac{e^{m_\pi |x|}}{m_\pi }\sqrt{m}e^{m_\pi |x|}.$$
(72)
Thus, we expect that
$$A_0^a(x)A_0^a(0)=\frac{d_1(x,0)}{mV^{1/2}}+d_{1/2}m^{1/2}e^{m_\pi |x|}+\mathrm{}$$
(73)
with other possible subleading terms from topological zero modes. As for the coefficient $`c_1(x,0)`$ in $`\pi ^a(x)\pi ^a(0)`$, the coefficient $`d_1(x,0)`$ above involves matrix elements between different eigenfunctions and could be quite small from cancellations. Thus, even though in $`A_0^a(x)A_0^a(0)`$, the physical pion contribution can still be $`O(m^{3/2})`$ smaller than the zero mode contribution, the effects of zero modes in this correlator are likely suppressed by the smaller coefficient $`d_1(x,0)`$.
To finish our discussion of the topological zero modes in correlators, we now examine the spectral form of $`\sigma (x)\sigma (0)_\mathrm{c}`$, where the $`c`$ subscript means that we only consider the connected part of the correlator. We find
$$\sigma (x)\sigma (0)_\mathrm{c}=\underset{\lambda ,\lambda ^{}}{}\frac{\psi _\lambda ^{}(x)\gamma _5\psi _\lambda ^{}(x)\psi _\lambda ^{}^{}(0)\gamma _5\psi _\lambda (0)}{(i\lambda +m)(i\lambda ^{}+m)}.$$
(74)
Since zero modes are eigenfunctions of $`\gamma _5`$, their contribution to the $`1/m^2`$ and $`1/m`$ terms in $`\pi ^a(x)\pi ^a(0)`$ and $`\sigma (x)\sigma (0)_\mathrm{c}`$ are equal. Thus, we have
$$\sigma (x)\sigma (0)_\mathrm{c}=\frac{1}{V^{1/2}}\left(\frac{c_2(x,0)}{m^2}+\frac{c_1(x,0)}{m}\right)+c_\sigma e^{m_{\sigma _c}|x|}+\mathrm{}$$
(75)
for small $`m`$. By considering $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$, we obtain a two point function with no zero mode effects, but which contains both the physical pion and a heavier state from $`\sigma (x)\sigma (0)_\mathrm{c}`$. Thus, to reduce the effects of topological near-zero modes in this way requires that one works with correlators where the heavy mass $`\sigma _c`$ state is present.
To summarize this section, we have seen how the topological near-zero modes for domain wall fermions should enter the correlators which are used to determine the pion mass. For $`\pi ^a(x)\pi ^a(0)`$, there must be a $`1/m_f^2`$ contribution from near-zero modes, compared with the $`1/\sqrt{m_f}`$ contribution expected from the physical pion. For $`A_0^a(x)A_0^a(0)`$, the topological near-zero modes can contribute a term of order $`1/m_f`$, while the physical pion should produce a $`\sqrt{m_f}`$ contribution. However, the coefficient of the $`1/m_f`$ term can be small. We also have pointed out that the volume dependence of the contribution of the topological near-zero modes to the correlator is different from the contribution due to the modes responsible for chiral symmetry breaking in QCD so that the zero-mode effects should vanish as the space-time volume increases.
The above discussion explicitly addresses the behavior to be found in a chiral theory. Thus, it will apply to the domain wall theory in the limit $`L_s\mathrm{}`$. We might expect two sorts of modified behavior for a theory with finite $`L_s`$. First, the chiral properties of the exact zero modes which eliminate the most singular terms from the $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ will no longer be exact for finite $`L_s`$ allowing more singular terms suppressed exponentially in $`L_s`$ to appear. Second the zero-mode singularities themselves may be softened by additional mass contributions to the denominators. We now turn to the results of our simulations.
### B Topological near-zero mode effects on the pion: numerical results
The first detailed studies of $`m_\pi ^2`$ as $`m_f0`$, done on $`8^3\times 32`$ lattices with $`\beta =5.7`$ and a variety of values of $`L_s`$, showed that $`m_\pi ^2(m_f=0)`$ was not decreasing exponentially to zero as $`L_s\mathrm{}`$, but rather seemed to be approaching a constant value of $`200`$ MeV . The pion mass was extracted from $`\pi ^a(x)\pi ^a(0)`$ and the resulting $`m_\pi ^2`$ versus $`m_f`$ showed noticeable curvature for the quark masses used, which were in the range $`0.020.22`$. Therefore, the extrapolation to $`m_f=0`$ was done using only the three lightest quark masses: 0.02, 0.06 and 0.10. Figure 14 updates the earlier graph in with more data at $`L_s=48`$ and a new point at $`L_s=64`$. The additional data does show a behavior that is more consistent with a monotonic decrease of $`m_\pi ^2(m_f=0)`$ with increasing $`L_s`$ that than seen in our earlier study . However, the dependence on $`L_s`$ shown in Fig. 14 still cannot be described by a single falling exponential and, for large $`L_s`$ is falling quite slowly. In this section we will probe this issue and others related to the chiral limit, using information from our simulations at both $`\beta =5.7`$ and 6.0 for many values of $`m_f`$ and $`L_s`$.
Figure 2 shows results for $`m_\pi ^2`$ versus $`m_f`$ for $`8^3\times 32`$ lattices at $`\beta =5.7`$ with $`L_s=48`$, including results for $`m_f=0.0`$. The pion mass is extracted from three different correlators: $`\pi ^a(x)\pi ^a(0)`$, $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$. For $`0.02m_f0.15`$, the pion masses extracted from the different correlators are in good agreement. As $`m_f0`$ the masses begin to disagree, presumably due to the differing contributions of the topological near-zero modes to each correlator. Table XXI gives our fitted pion masses for $`m_f<0.01`$. While the different correlators generally have reasonable values for $`\chi ^2/\mathrm{dof}`$ the fitted masses disagree substantially. For large enough separation of the interpolating operators, the three correlators should give the same mass. However, we cannot take this large separation limit in our finite volume. The results in Table XXI are the apparent masses as determined from fitting to the correlators for finite separation of the interpolating operators.
The lines drawn in Figure 2 are from correlated linear fits to $`m_\pi ^2`$ using $`m_f=0.02`$ to 0.1. The dotted line is for $`m_\pi ^2`$ from $`\pi ^a(x)\pi ^a(0)`$, the solid line for $`A_0^a(x)A_0^a(0)`$ and the dashed line for $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$. The fit results are
$`m_\pi ^2=0.053(8)+4.76(12)m_f`$ $`\chi ^2/\mathrm{dof}=0.7\pm 2.5`$ (76)
$`m_\pi ^2=0.042(8)+4.90(9)m_f`$ $`\chi ^2/\mathrm{dof}=0.01\pm 0.25`$ (77)
$`m_\pi ^2=0.037(8)+5.04(6)m_f`$ $`\chi ^2/\mathrm{dof}=1.7\pm 2.8`$ (78)
for $`\pi ^a(x)\pi ^a(0)`$, $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ respectively. Note for large mass, $`m_f>0.1`$, $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ gives a mass that is systematically higher than that implied by the other two correlators, likely due to contamination from heavy states present in $`\sigma (x)\sigma (0)_\mathrm{c}`$.
Figure 15 shows the pion effective mass from the three different correlators for $`8^3\times 32`$ lattices at $`\beta =5.7`$ with $`L_s=48`$ and $`m_f=0.0`$. Reasonable plateaus are present in $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)`$, although the value for the effective mass is markedly different. The $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ effective mass becomes very small for intermediate values of $`t`$. Figure 16 is a similar plot, except for $`m_f=0.04`$. Here the effective mass plots show nice plateaus and consistent results. This supports the presence of topological near-zero modes affecting the various correlators in different ways and provides an example where nice plateaus do not assure a correct asymptotic result.
As a final step in demonstrating the zero mode effects in the various correlators, in Figure 17 the evolution of $`\pi ^a(x)\pi ^a(0)`$, $`\sigma (x)\sigma (0)_\mathrm{c}`$ and $`A_0^a(x)A_0^a(0)`$ is shown for $`8^3\times 32`$ lattices at $`\beta =5.7`$ with $`L_s=48`$ and $`m_f=0.0`$. These correlators are from a point source to a point sink and the zero spatial momentum component is taken for the sink. The sink is at a separation $`t=8`$ from the source. The correlators $`\pi ^a(x)\pi ^a(0)`$ and $`\sigma (x)\sigma (0)_\mathrm{c}`$ show very large fluctuations, which are common to both correlators, showing the presence of topological near-zero modes. These large fluctuations are clearly dominating the ensemble average for the correlators at this separation, $`t=8`$. The $`A_0^a(x)A_0^a(0)`$ correlator does not show large fluctuations where $`\pi ^a(x)\pi ^a(0)`$ and $`\sigma (x)\sigma (0)_\mathrm{c}`$ do, making the topological near zero mode effects smaller for this correlator, as expected from the theoretical discussion of the previous subsection. Figure 18 is a similar plot, for the same configurations, except with $`m_f=0.04`$. There is no evidence for a large role being played by the topological near-zero modes.
Similar results have been obtained for simulations on $`16^3\times 32`$ lattices at $`\beta =6.0`$ with $`L_s=16`$. These lattices have essentially the same spatial volume, in physical units, as the previous $`8^3\times 32`$, $`\beta =5.7`$ lattices since the lattice spacing is half that for $`\beta =5.7`$. Figure 19 shows $`m_\pi ^2`$ for $`A_0^a(x)A_0^a(0)`$, $`\pi ^a(x)\pi ^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$. For the smallest $`m_f`$ points, 0.0 and 0.001, all three correlators give different results. For larger values of $`m_f`$, the pion mass from $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)`$ agree, while the $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ pion mass is systematically high, likely due to the contribution of the heavy states in $`\sigma (x)\sigma (0)_\mathrm{c}`$. The lines drawn in Figure 19 are from correlated linear fits to $`m_\pi ^2`$ using $`m_f=0.01`$ to 0.04. The dotted line is for $`m_\pi ^2`$ from $`\pi ^a(x)\pi ^a(0)`$, the solid line for $`A_0^a(x)A_0^a(0)`$ and the dashed line for $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$. The fit results are
$`m_\pi ^2=0.0132(20)+3.07(6)m_f`$ $`\chi ^2/\mathrm{dof}=5.0\pm 5.0`$ (79)
$`m_\pi ^2=0.0098(20)+3.14(9)m_f`$ $`\chi ^2/\mathrm{dof}=0.03\pm 0.30`$ (80)
$`m_\pi ^2=0.0020(26)+3.56(8)m_f`$ $`\chi ^2/\mathrm{dof}=0.06\pm 0.51`$ (81)
for $`\pi ^a(x)\pi ^a(0)`$, $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ respectively.
Figure 20 shows effective mass plots for the pion from the three correlators for $`m_f=0.001`$ and Figure 21 is for $`m_f=0.01`$. Both figures show reasonable plateaus, even though there are differences in the final fitted masses. We have also studied the evolution of the correlators at a fixed $`t`$ for these $`\beta =6.0`$ lattices and see clear topological near-zero mode effects as were seen at $`\beta =5.7`$.
Thus, investigating the chiral limit of domain wall fermions in quenched QCD by measuring the pion mass is made difficult by the presence of topological near-zero modes. One component in the somewhat large values of $`m_\pi ^2`$ plotted in Fig. 14 is the effect of topological near zero modes. As can be seen by a comparison with Table XVI the results we find from the correlator $`A_0^a(x)A_0^a(0)`$ for $`m_\pi ^2(m_f=0)`$ are about 1 1/2 standard deviations lower for $`L_s=32`$ and $`48`$. This is likely a systematic bias caused by the greater influence of the topological near-zero modes on the $`\pi ^a(x)\pi ^a(0)`$ correlator. Unfortunately, these effects may also enter in the other correlators that can give $`m_\pi `$, at least for the source-sink time separations currently accessible. With this large distortion due to the topological near-zero modes, there we cannot determine the chiral limit by extrapolating to the point where $`m_\pi ^2`$ vanishes. Subtler finite volume effects and possible quenched chiral logarithms are completely overshadowed by the singular nature of the basic quark propagators for small $`m_f`$.
In many ways, the presence of these topological near-zero modes is a welcome change from other lattice fermion formulations because they are a vital part of the spectrum of any continuum Dirac operator. However, in order to further investigate the chiral limit, they must be removed, or at least suppressed. Without adding the fermionic determinant to the path integral, we can suppress the effect of topological near-zero modes by going to large volumes.
### C The pion mass for larger volume
Having seen clear evidence for topological near-zero modes in the measurements of the pion mass for lattices with a physical size of $`2`$ Fermi, we have worked on a larger physical volume, $`4`$ Fermi, to suppress the effects of these modes. As we saw in Section IV from studying $`\overline{q}q`$, the effects of the topological near-zero modes were dramatically reduced for larger volumes. Here we present results for the pion mass from simulating with $`16^3\times 32`$ lattices at $`\beta =5.7`$ and $`L_s=48`$.
Figure 3 shows $`m_\pi ^2`$ plotted against $`m_f`$ for these runs. In contrast to the smaller volume $`8^3\times 32`$ result shown in Figure 2, all three correlators now give the same results for the pion mass, within statistics (Table XXI). The larger volume has clearly reduced the effects of the zero modes. Further evidence of the consistency of the mass from the three correlators is shown in Figure 22. Here, for each $`m_f`$, the average value of $`m_\pi `$ is calculated and then the deviation from that average, for each correlator, is plotted. For each $`m_f`$, $`A_0^a(x)A_0^a(0)`$ is offset to the left and the $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ to the right for clarity.
Figure 23 shows the effective mass from each of the three correlators for $`m_f=0`$. In contrast to the smaller volume case, the effective masses have quite similar values and lead to the same fitted mass, within errors. As a last comparison with the small volume, Figure 24 shows the three correlators at a time separation of $`t=8`$ as a function of configuration number. Little if any effect of topological near-zero modes is seen. Thus, we conclude that this larger volume has suppressed these effects as expected.
Having established that a consistent pion mass can be determined from our fitting range, we discuss the result of linear fits of $`m_\pi ^2`$ as a function of $`m_f`$. We have done correlated linear fits of $`m_\pi ^2`$ to $`m_f`$ for each of the correlators, using a variety of different ranges for $`m_f`$ in the fit. The resulting $`\chi ^2`$ per degree of freedom is shown in Figure 25, including the jackknife error on the $`\chi ^2`$. (The plotted error bars are the $`\pm 1\sigma `$ errors from the jackknife procedure and do not mean that $`\chi ^2`$ can become negative.) The pion propagator for $`m_f=0.0`$ and 0.04 was measured on the same set of configurations, with some of the $`m_f=0.08`$ propagators also measured on those configurations. The $`m_f=0.02`$, 0.06 and 0.10 points were all measured on the same configurations, along with the remaining 0.08 propagators. Thus, these points are less correlated in $`m_f`$ than the corresponding measurements on the smaller volumes.
Now let us discuss the quality of these fits. Given the significant upward curvature of $`m_\pi ^2`$ for $`m_f0.1`$, seen for example in Figure 2, we limit the mass range to $`m_f0.08`$. If we do not include the lightest masses and fit the points with $`0.02m_f0.08`$, as shown in Figure 25 we obtain acceptable values for $`\chi ^2`$ per degree of freedom for all three correlators. Specifically using the mass range $`m_f=0.02`$ to 0.08, the fits to $`m_\pi ^2`$ from the correlators $`\pi ^a(x)\pi ^a(0)`$, $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ are
$$\begin{array}{cc}m_\pi ^2=0.044(5)+4.75(5)m_f& \chi ^2/\mathrm{dof}=1.4\pm 3.6\\ m_\pi ^2=0.051(3)+4.68(4)m_f& \chi ^2/\mathrm{dof}=1.4\pm 1.4\\ m_\pi ^2=0.049(3)+4.70(5)m_f& \chi ^2/\mathrm{dof}=2.5\pm 2.4.\end{array}$$
(82)
However, given our confidence that this larger $`16^3`$ volume permits the reliable calculation of the pion mass for smaller values of $`m_\pi `$ we can also attempt a linear fit in the entire range $`0.0m_f0.08`$. For this mass range, we find
$$\begin{array}{ccc}m_\pi ^2=0.042(3)+4.77(3)m_f& \chi ^2/\mathrm{dof}=1.3\pm 3.5& \\ m_\pi ^2=0.044(3)+4.75(5)m_f& \chi ^2/\mathrm{dof}=4.3\pm 2.6& \\ m_\pi ^2=0.042(4)+4.82(6)m_f& \chi ^2/\mathrm{dof}=4.4\pm 3.0& \end{array}$$
(83)
for the correlators $`\pi ^a(x)\pi ^a(0)`$, $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ respectively. The $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ fits suggest that $`m_\pi ^2`$ is not linear in this mass range. While the $`\pi ^a(x)\pi ^a(0)`$ fit is acceptable, as can be seen from a careful examination of Figure 3, this acceptable fit comes because the $`m_f=0.02`$ point lies somewhat below while the $`m_f=0.0`$ lies somewhat above the masses obtained from the other two correlators. Since the smaller volume studies suggest that the $`\pi ^a(x)\pi ^a(0)`$ correlator is most sensitive to zero modes and such an upturn for small mass is the effect of zero modes seen at smaller volume, this could easily be a remaining zero mode distortion.
It is difficult to draw a firm conclusion from the relatively large correlated $`\chi ^2`$/dof presented in Eq. 83. As is indicated by the errors shown, these $`\chi ^2`$/dof are not reliably known. However, the comparison of the $`\chi ^2`$/dof between Eqs. 82 and 83 may be more meaningful. We attribute significant weight to the fact that the lightest $`m_f=0`$ point lies below the value predicted by a linear extrapolation from larger masses as can be easily seen in Figure 3.
We conclude that a linear fit does not well represent our data over the full mass range $`m_f=0.0`$ to 0.1. Of course, non-linearities for larger masses can come from a variety of sources including terms from the naive analytic expansion in powers of $`m_f`$. However, for small $`m_f`$, linearity is expected for large volumes in full QCD. In contrast, in the quenched approximation the absence of the fermion determinant may result in complex and more singular infrared behavior. For example, it has been argued that a quenched chiral logarithm can appear in $`m_\pi ^2`$ versus $`m_f`$ for quenched QCD . The results just presented may be evidence for some non-linear behavior of this sort.
Because of the poor linear fits found for small $`m_f`$, our data does not allow a determination of the location of the chiral limit for quenched domain wall fermions by a simple extrapolation of $`m_\pi ^2`$. Even with the suppression of topological near-zero mode effects that has been achieved by going to larger volume, further theoretical input may needed if we are to deduce $`m_{\mathrm{res}}`$ from these measurements of $`m_\pi ^2`$. In the next section we will discuss our determination of the location of the chiral limit using other techniques and then return to the question of the behavior of $`m_\pi ^2`$ with $`m_f`$.
## VI The residual mass
### A Determining the residual mass
In this section, we discuss our determination of $`m_{\mathrm{res}}`$ using the low-momentum identity in Eq. 15. This can done by calculating the ratio
$$R(t)=\frac{_\stackrel{}{x}J_{5q}^a(\stackrel{}{x},t)\pi ^a(0)}{_\stackrel{}{x}J_5^a(\stackrel{}{x},t)\pi ^a(0)}$$
(84)
as a function of $`t`$ (no sum on $`a`$), where $`\pi ^a(0)`$ is a source evaluated at $`t=0`$ but possibly extended in spatial position. This ratio was first used to determine $`m_{\mathrm{res}}`$ in Ref. and later in Refs. . Our results are consistent with this earlier work, but a much more detailed study is undertaken here. For $`t`$ outside some short-distance region, $`tt_{\mathrm{min}}`$, $`R(t)`$ should be simply equal to $`m_{\mathrm{res}}`$. Using $`R(t)`$ for very large $`t`$ gives $`m_{\mathrm{res}}`$ as the coupling of the pion to the mid-point pseudoscalar density divided by its coupling to the wall pseudoscalar density. Of course, $`m_{\mathrm{res}}`$ is an additive contribution to the effective quark mass at low energies which effects all low-energy physics, not just the pion. To understand how large $`t`$ must be, Figure 26 shows a typical good plateau and a poor one. Results are shown for $`8^3\times 32`$ lattices with $`m_f=0.04`$ and $`\beta =5.7`$ for $`L_s=32`$ and 48. The good plateau is obtained from 335 configurations for $`L_s=48`$, while the poor plateau is obtained from 184 configurations for $`L_s=32`$. The fewer measurements for $`L_s=32`$ likely is the cause for the upturn in the data at large $`t`$ and adding more configurations at this $`L_s`$ should improve the signal.
From observing the onset of the plateaus in our data, we calculate $`m_{\mathrm{res}}`$ from the ratio in Eq. 84 using the range $`4t16`$ for $`\beta =5.7`$, $`6t26`$ for $`\beta =5.85`$, and $`2t16`$ for $`\beta =6.0`$. The jackknife method is used to measure the statistical uncertainty and our $`m_{\mathrm{res}}`$ results at $`\beta =5.7,5.85`$ and $`6.0`$ are listed in Tables XXII and XXIII. For most data sets, nice plateaus can be seen over the selected range, while for the few others with the poor plateaus, using a different range could change the results by $`<5\%`$. We have also measured $`m_{\mathrm{res}}`$ for different values of $`m_f`$ for $`\beta =5.7`$ on $`16^3\times 32`$ lattices with $`L_s=48`$. Table XXIII gives the results and shows that the residual mass has little dependence on the input quark mass, reflecting the expected universal character of $`m_{\mathrm{res}}`$. Our $`\beta =6.0`$ results for $`m_{\mathrm{res}}`$ appear to be a consistent extension of the values plotted in Figure 5 of Ref. for $`L_s=4`$, 6 and 10.
The $`L_s`$ dependence of $`m_{\mathrm{res}}`$ is of vital importance to numerical simulations with domain wall fermions. Without the effects of topological near-zero modes, quenched chiral logs and finite volume, $`m_\pi ^2(m_f=0)`$ should be proportional to $`m_{\mathrm{res}}`$ and should vanish with $`m_{\mathrm{res}}`$ as $`L_s\mathrm{}`$. However, in Section V we discussed how topological near-zero mode effects alter $`\pi ^a(x)\pi ^a(0)`$ and can distort the value of $`m_\pi ^2(m_f=0)`$ for large $`L_s`$ shown in Figure 14. By measuring the ratio in Eq. 84, we can determine $`m_{\mathrm{res}}`$ for non-zero $`m_f`$ and suppress all these effects which make the $`m_f0`$ limit problematic. This allows us to study the $`L_s`$ dependence of $`m_{\mathrm{res}}`$, to which we now turn.
From the two values of $`L_s`$ shown in Figure 26, we see that the residual mass for $`8^3\times 32`$ lattices at $`\beta =5.7`$ falls from $`0.0105(2)`$ to $`0.00688(13)`$ as $`L_s`$ is increased from $`32`$ to $`48`$. This is in sharp contrast to the almost identical results for $`m_\pi ^2(m_f=0)`$ at these two values for $`L_s`$ (Figure 14). The overlap of the surface states is significantly suppressed, as expected, even at this relatively strong coupling. We have not pursued the asymptotic behavior for large $`L_s`$ at $`\beta =5.7`$, due to the large values for $`L_s`$ required, but instead have studied this question for $`\beta =6.0`$.
Figure 4 shows a similar study of the $`L_s`$ dependence of the residual mass for $`16^3\times 32`$ lattices with $`m_f=0.02`$ and $`\beta =6.0`$. The number of configurations used is modest for the larger values of $`L_s`$. We have used the factor $`Z_S(\overline{\mathrm{MS}},2\mathrm{G}\mathrm{e}\mathrm{V})=0.619(25)`$ obtained by a combination of non-perturbative renormalization and standard perturbation theory to convert the plotted values of $`m_{\mathrm{res}}`$ into MeV. The value of $`m_{\mathrm{res}}`$ is decreasing with $`L_s`$ for all values of $`L_s`$, but is poorly fit by a simple exponential. In particular, an exponential fit using all values of $`L_s`$ gives
$$m_{\mathrm{res}}=0.0068(4)\mathrm{exp}(0.094(4)L_s),\chi ^2/\mathrm{dof}(3)=32$$
(85)
which clearly does not match the measured values. Adding a constant to the fit gives
$$m_{\mathrm{res}}=0.00032(3)+0.018(3)\mathrm{exp}(0.181(13)L_s),\chi ^2/\mathrm{dof}(2)=4.1$$
(86)
where again all values for $`L_s`$ were used. Even if this is the correct asymptotic form, the value of $`m_{\mathrm{res}}`$ for $`L_s\mathrm{}`$ is very small, 1 MeV.
We have also tried fitting the largest three $`L_s`$ points to a simple exponential and find
$$m_{\mathrm{res}}=0.0012(2)\mathrm{exp}(0.032(6)L_s),\chi ^2/\mathrm{dof}(1)=0.074.$$
(87)
Our data is consistent with the residual mixing vanishing exponentially as $`L_s\mathrm{}`$, but the 0.032 coefficient in the exponent of Eq. 87 is quite small. Of course, we can easily obtain an excellent fit to our five points if we include a second exponential. For example, as shown in the figure, the five points fit well to two-exponential function
$$m_{\mathrm{res}}=0.038(16)\mathrm{exp}(0.26(4)L_s)+0.0010(3)\mathrm{exp}(0.027(7)L_s),\chi ^2/\mathrm{dof}(1)=0.1.$$
(88)
Our measurements do not demonstrate a precise asymptotic form for $`m_{\mathrm{res}}`$ as a function of $`L_s`$. However, we do see $`m_{\mathrm{res}}`$ decreasing for large $`L_s`$ until, for $`L_s24`$, it has become so small as to be essentially negligible for current numerical work. For $`\beta =5.7`$ at $`L_s`$ = 48, $`m_{\mathrm{res}}`$ is $`0.074(5)`$ in units of the strange quark mass, while for $`\beta =6.0`$ at $`L_s=16`$ it is $`0.033(3)m_s`$. In the latter case, where we know the renormalization factors, $`m_{\mathrm{res}}`$ in the $`\overline{\mathrm{MS}}`$ scheme at 2 GeV is 3.87(16) MeV. Thus, even though more simulations will be needed to get the precise asymptotic form, we find domain wall fermions having the expected chiral properties for large $`L_s`$, even for lattice spacings of around 1 Fermi.
In the next subsection, we will use the values of $`m_{\mathrm{res}}`$ that have just been determined to investigate further the $`m_f`$ dependence of $`m_\pi ^2`$, looking in particular at possible non-linear behavior as $`m_f+m_{\mathrm{res}}0`$. Here we would like to discuss a simpler consistency check on the values of $`m_{\mathrm{res}}`$ just obtained. For the $`\pi `$, $`\rho `$ and nucleon we have established good linear $`m_f`$ behavior for larger values of $`m_f`$ with slopes and intercepts given in Tables XVI and XVII. If the only effect on these masses of changing $`L_s`$ is to change the effective quark mass through the corresponding change in $`m_{\mathrm{res}}`$, then we should be able to relate the the differences in the intercepts given in these tables to the product of the corresponding slope times the change in $`m_{\mathrm{res}}`$ given in Tables XXII and XXIII.
While this comparison shows no inconsistencies, the errors in the intercepts are typically too large to permit a detailed confirmation. For example, the difference in intercepts for $`m_\pi ^2`$ at $`\beta =6.0`$ between $`L_s=16`$ and 24 is 0.0004(30) while the difference predicted from the slope and the measured change in $`m_{\mathrm{res}}`$ is 0.0020(2). The best test of this sort can be made using the actual value for $`m_\pi `$ determined at $`\beta =5.7`$ and $`m_f=0.04`$ for $`L_s=32`$ and 48. Here the difference of the masses squared is 0.012(6) while the prediction from the slope and change in $`m_{\mathrm{res}}`$ is 0.0176(12). Thus, we can demonstrate consistency with the expected behavior but cannot make a definitive test.
### B The residual mass and $`m_\pi ^2`$ versus $`m_f`$
The definition of $`m_{\mathrm{res}}`$ and its measurement mean that we have determined the value of $`m_f`$ for which the pion should become massless if the domain wall method is successfully representing the chiral limit of the underlying theory. We can now return to the question of the dependence of $`m_\pi ^2`$ on $`m_f`$, starting with the $`16^3\times 32`$ simulations at $`\beta =5.7`$ and $`L_s=48`$. Recalling Figure 3, we found that the larger volume gave consistent pion mass measurements from the three correlators, but $`m_\pi ^2`$ was not well fit as a linear function of $`m_f`$ for two of the correlators if the $`m_f=0.0`$ point was included. In Figure 3, we have included the value of $`m_{\mathrm{res}}`$ (the starred point) as measured from Eq. 84. (Its error bar on the horizontal axis is a vertical line on this scale.) The solid line is the fit to the $`A_0^a(x)A_0^a(0)`$ correlator for $`m_f=0.02`$ to 0.08 given in Eq. 82 while the dotted line is for the $`\pi ^a(x)\pi ^a(0)`$ correlator for $`m_f=0.0`$ to 0.08 as given in Eq. 83. Thus, we see that linear fits poorly represent the data when the $`m_f=0.0`$ point is included for the $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ case and fail for all three correlators when the pion mass is required to vanish at $`m_f=m_{\mathrm{res}}`$.
We can make this conclusion more quantitative by comparing our accurate value for $`m_{\mathrm{res}}=0.00688(13)`$ at $`L_s=48`$ determined on an $`8^3\times 32`$ lattice with the naive linear extrapolation of $`m_\pi ^2(m_f)`$ to the point $`m_\pi ^2=0`$. Using the most reliable linear fits obtained by excluding the $`m_f=0`$ point in Eq. 82 we obtain the $`x`$-intercept values shown in Table XXIV: -0.0092(12), -0.0108(7) and -0.0104(7) for the $`\pi ^a(x)\pi ^a(0)`$, $`A_0^a(x)A_0^a(0)`$ and $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ correlators respectively. These differ from this value of $`m_{\mathrm{res}}`$ by $`50\%`$ and 2, 5 and 6 standard deviations respectively. We conclude that there is a significant discrepancy between the $`m_f`$-dependence of these $`m_\pi `$ results and the hypothesis that $`m_\pi ^2(m_{\mathrm{res}})=0`$. However, notice that if the $`m_f=0`$ points are included in the linear fits, and the less accurate $`m_{\mathrm{res}}`$ from the same volume is used, this discrepancy can be reduced. For example, a linear fit to the data from the $`\pi ^a(x)\pi ^a(0)`$ correlator in Eq. 83 has an intercept at -0.0088(5) while $`m_{\mathrm{res}}=0.0072(9)`$ on the same volume. We believe that such an interpretation should be discounted as failing to exploit all the available information.
Is this significant discrepancy caused by essential non-linearities in the quenched approximation or by a breakdown of the domain wall method, for example, large $`O(a^2)`$ effects? We can address this question by making a similar comparison for $`\beta =6.0`$ where $`O(a^2)`$ effects should be significantly reduced. Since we have not investigated a large volume at this weaker coupling, we propose to examine the $`A_0^a(x)A_0^a(0)`$ correlator because reduced zero-mode effects were seen for this correlator in our $`\beta =5.7`$ studies. Using the three lightest masses we find $`x`$-intercepts of -0.0031(7) and -0.0030(9) for the $`L_s=16`$ and 24 cases respectively. Again, these are dramatically farther from the origin than the corresponding values of $`m_{\mathrm{res}}=0.00124(5)`$ and 0.00059(4). These are each three standard deviation effects. However, they are obtained on independent configurations and together can be viewed as a 6 standard deviation discrepancy. Thus, if possible finite-volume difficulties are ignored, we have again strong evidence for a discrepancy. Rather than decreasing by a factor of four as would be expected from an $`O(a^2)`$ error, this fractional discrepancy is substantially larger in this $`\beta =6.0`$ comparison. Thus, it is natural to conclude that domain wall fermions are accurately representing the chiral behavior of quenched QCD.
At the beginning of Section V we listed possible systematic effects influencing the chiral limit for $`m_\pi `$. With a measurement of $`m_{\mathrm{res}}`$ we have quantified the role of finite $`L_s`$ and with the larger volume used for $`\beta =5.7`$ we have reduced, if not eliminated the topological near-zero modes. We should also have minimized other finite volume distortions of the density of eigenvalues, which also influence the pion mass. Finally with the comparison above, we have examined the possibility of $`O(a^2)`$ errors. Thus, we now address the question of quenched chiral logarithms. Predictions of this particular pathology of quenched simulations were made some time ago. There is certainly much data indicating possible support for the predictions, but there is disagreement about its conclusiveness, see for example refs. . Since many other effects must be removed before these subtle logarithms are convincingly seen, it is a challenging numerical issue.
The natural first place to look for quenched chiral logarithm effects is in $`m_\pi `$, but this is difficult for Wilson fermions, where the chiral point is not crisply defined for finite lattice spacing. For staggered fermions, where the chiral limit occurs when the input quark mass is zero, the issue is complicated by the presence of only a single Goldstone pion. In some respects, domain wall fermions are an ideal place to look for these effects, except that the statistical resolution needed is difficult to achieve with the additional computational load of the fifth dimension. In addition, the topological near-zero modes are a much larger quenched pathology at moderate volumes.
As one way of probing the non-linearity in $`m_\pi ^2`$ versus $`m_f`$, we have fitted our data for $`m_\pi ^2`$ for $`16^3\times 32`$ lattices at $`\beta =5.7`$ and $`L_s=48`$ to the form
$$m_\pi ^2=a_0(m_f+a_1)(1+a_2\mathrm{ln}(m_f+a_1))$$
(89)
and the results are given in Table XXIV. The fit yields a value for the residual mass (the parameter $`a_1`$ above) and the results are quite close to those measured from the ratio of Eq. 84. Figure 27 shows the result from fitting $`A_0^a(x)A_0^a(0)`$ for $`m_f=0.0`$ to 0.08 to the quenched chiral logarithm form given in Eq. 89. We have excluded the larger values of $`m_f`$ from our fits, since higher order terms are needed in Eq. 89 to accommodate the upward curvature of our $`m_\pi ^2`$ data. While the $`\chi ^2`$/dof for the logarithmic fit is only marginally better than those obtained for the simple linear fits described earlier in Eq. 83 for this same mass range, the ability of the logarithmic fit to predict the appropriate $`m_{\mathrm{res}}`$ value is significant.
For the simulations at smaller physical volumes, $`8^3\times 32`$ at $`\beta =5.7`$ and $`16^3\times 32`$ at $`\beta =6.0`$, the values for $`m_{\mathrm{res}}`$ measured from Eq. 84 are generally smaller than the $`x`$ intercepts for the linear fits shown in Figures 2 and 19. This indicates curvature in the direction given by a chiral logarithm, but the other phenomena that may be affecting these chiral limits make quantitative analysis ambiguous. We note that $`m_\pi ^2`$ from $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ seems to smoothly curve towards the value of $`m_{\mathrm{res}}`$ from the previous subsection. However, we are not sufficiently certain of the absence of zero mode effects in the $`\pi ^a(x)\pi ^a(0)+\sigma (x)\sigma (0)_\mathrm{c}`$ correlator to describe a logarithmic fit to these cases.
This nice agreement between the values of $`m_{\mathrm{res}}`$ determined from the location of the $`m_\pi ^2=0`$ point in these fits and that computed by other means earlier in the paper implies consistency between our results and the logarithmic form of Eq. 89. Of course, other non-linear terms could be used to explain this curvature and, given our statistics, would provide an equally consistent description of our data.
However, our most important conclusion is not related to quenched chiral logarithms, but rather to having seen all the expected properties for the chiral limit with domain wall fermions. Once the topological near-zero mode effects are reduced or eliminated, consistent pion masses can be measured. A precise measurement of $`m_{\mathrm{res}}`$ is consistent with our $`m_\pi ^2`$ versus $`m_f`$ dependence if, for example, a chiral logarithm term is included. In short, domain wall fermions are showing sensible chiral properties, even on lattices with a lattice spacing of $`1`$ Fermi.
We have chosen not to pursue an additional method of determining $`m_{\mathrm{res}}`$ that has been proposed in two of our previous publications . In that method, one examines the integrated Ward-Takahashi identity in Eq. 60 and uses the location of the pion pole in $`\chi _{\pi \pi }`$ to determine $`m_{\mathrm{res}}`$. While this technique should be reliable for dynamical fermion calculations, e.g. as used in Ref , it does not explicitly allow for the effects of topological near-zero modes or possible non-linear behavior of $`m_\pi ^2(m_f)`$ that we have found to be important in the quenched approximation. Thus, even though this method gave a result for $`\beta =5.7`$ quite close to the $`L_s=48`$ value $`m_{\mathrm{res}}=0.00688(13)`$ presented in this paper, more analysis is needed to adequately justify its use in this quenched case.
### C Eigenvalue properties and $`m_{\mathrm{res}}`$
A comparison of the approximate form of the Banks-Casher relation for domain wall fermions given in Eq. 54 with the usual 4-dimensional expression in Eq. 44 suggests a close relationship between the parameter $`\delta m_i`$ deduced from the $`i^{th}`$ eigenvalue $`\mathrm{\Lambda }_{H,i}`$ of $`D_H`$ and the residual mass $`m_{\mathrm{res}}`$. In this section we will explore this relation further making use of an exploratory study of the low-lying spectrum of $`D_H`$ .
These eigenvalues were calculated for 32 configurations obtained at $`\beta =6.0`$ on a $`16^4`$ lattice with $`L_s=16`$ listed in Table II and beginning with an equilibrated configuration from an earlier run. We used the Kalkreuter-Simma method to find the 19 lowest eigenvalues on each configurationWe thank Robert Edwards whose program formed the basis of the code used in this part of the calculation.. We apply this method to the positive matrix $`D_H^2`$, and then determine the eigenvectors and eigenvalues of $`D_H`$ by a final explicit diagonalization of $`D_H`$ in the subspace of the eigenvectors of $`D_H^2`$ just determined. The details of our application of this method and a more complete description of these results will be presented in a later publication .
While this method determines both the eigenvalues and eigenvectors, we have chosen to examine only the $`s`$-dependent, four-dimensional inner products:
$$\mathrm{\Gamma }_{R/L}(s)_{i,j}=\underset{x}{}\mathrm{\Psi }^{}(x,s)_{\mathrm{\Lambda }_{H,i}}P_{R/L}\mathrm{\Psi }(x,s)_{\mathrm{\Lambda }_{H,j}},$$
(90)
where the indices $`i,j`$ run over all of the 19 eigenvalues while $`P_R`$ and $`P_L`$ are the left and right spin projection operators defined above Eq. 8. In order to be able to make use of the mass dependence of the eigenvalues, we have repeated the calculation of $`\mathrm{\Lambda }_i(m_f)`$ and $`\mathrm{\Gamma }_{R/L}(s,m_f)_{i,j}`$ five times on each configuration for the five different mass values $`m_f=0.0`$, 0.0025, 0.005, 0.0075 and 0.001.
Here we will describe some of the overall features of this calculation and then examine more closely the relation between the parameters $`\delta m_i`$ and the value of $`m_{\mathrm{res}}`$ determined earlier in this paper. First we examine the diagonal elements of the matrix $`\mathrm{\Gamma }(s)`$
$$𝒩(s)_i=\mathrm{\Gamma }_R(s)_{i,i}+\mathrm{\Gamma }_L(s)_{i,i}.$$
(91)
This is the contribution to the norm of the 5-dimensional wave function from the 4-dimensional hyperplane with a specific value of $`s`$. For these low lying eigenvalues, we expect that this norm should be concentrated on the $`s=0`$ and $`s=L_s1`$ walls, which we find to be true to good accuracy. For the entire group of $`32\times 19\times 5=3,040`$ eigenvectors computed, the ratio of the sum of the norm on the two walls to the minimum value of this norm between the walls was always greater than 34, $`𝒩(0)+𝒩(L_s1)>34𝒩(s_{\mathrm{min}})`$. The median value for this ratio was 744. Thus, the general framework upon which the domain wall formalism rests appears approximately valid.
As a test of our method for determining the eigenvectors, we evaluate the left- and right-hand side of the symmetry relation, Eq. 28, between pairs of eigenvectors on a given configuration. The resulting equality:
$$(\mathrm{\Lambda }_{H,i}+\mathrm{\Lambda }_{H,j})\mathrm{\Lambda }_{H,i}|\mathrm{\Gamma }_5|\mathrm{\Lambda }_{H,j}=\mathrm{\Lambda }_{H,i}|\left(2m_fQ^{(\mathrm{w})}+2Q^{(\mathrm{mp})}\right)|\mathrm{\Lambda }_{H,j}$$
(92)
provides a good test of our diagonalization procedure. The vectors $`|\mathrm{\Lambda }_{H,j}`$ needed to evaluate this expression are eigenvectors of the Dirac operator $`D_H`$, not $`D_H^2`$. We determine the eigenvectors of $`D_H`$ by diagonalization within the 19-dimensional subspace found by applying the Kalkreuter-Simma method to $`D_H^2`$. In the event that the $`19^{th}`$ and $`20^{th}`$ eigenvalues of $`D_H^2`$ are nearly degenerate (not entirely unlikely given the expectation that the eigenvalues of $`D_H`$ occur in $`\pm \mathrm{\Lambda }_H`$ pairs), this truncated, 19-dimensional subspace will not be spanned by eigenvectors of $`D_H`$. It will contain 18 valid eigenvectors and a $`19^{th}`$ vector, orthogonal to the rest but not an eigenvector of $`D_H`$. This “spurious” eigenvector can be reliably removed since it will give an “eigenvalue” whose square does not agree with any found for $`D_H^2`$. We remove such eigenvectors from our test of Eq. 92 and, for uniformity, the $`19^{th}`$ eigenvector in the case that no spurious eigenvector occurs. There are then $`32\times 5=160`$ instances where we can check $`18^2`$ independent elements of Eq. 92. We find that 95% of these 51,840 comparisons have a fractional error below 5%. The few cases with significantly worse agreement, result from infrequent near degeneracies which challenge the Rayleigh-Ritz method on which the Kalkreuter-Simma algorithm is based.
For most configurations there are easily identified zero modes. Typically the few lowest eigenvalues have eigenvectors all of which are bound to the same wall, either $`s=0`$ or $`s=L_s1`$. The corresponding matrix elements $`\mathrm{\Lambda }_{H,i}|\mathrm{\Gamma }_5|\mathrm{\Lambda }_{H,i}`$ all have the same sign and are within a few percent of 1, showing precisely the structure expected in a four-dimensional theory as summarized in Eq. 41.
The potential of the domain wall method is nicely displayed by examining the properties of one of our better configurations. In Fig 28, we show the magnitude of the elements of the matrix $`\mathrm{\Lambda }_{H,i}|\mathrm{\Gamma }_5|\mathrm{\Lambda }_{H,j}`$ in a three-dimensional plot. Note the five zero-modes in this configuration are easily recognized. Each has a diagonal matrix element of $`\mathrm{\Gamma }^5`$ within 1.5% of 1 and matrix elements with other vectors all of magnitude below 0.06. The values of $`\lambda _i`$ for these five eigenvalues all lie in magnitude below $`3.6\times 10^4`$ while the remaining paired eigenvalues lie between 0.028 and 0.093. In Figures 29 and 30 we show the $`s`$-dependence of the first two zero-modes and the first pair of non-zero eigenvectors, numbers 5 and 6. One sees precisely the expected behavior. Both zero-modes are bound to the same wall (as are the other three zero modes) while the two paired non-zero modes are nearly symmetrical between right and left. This is clearly identified as a configuration with topological charge $`\nu =+5`$.
Of direct interest in this section is the mass dependence of $`\mathrm{\Lambda }(m_f)_{H,i}`$ and a quadratic fit of the sort proposed in Eq. 51. For the small masses we have used, this quadratic form provides an excellent fit, after some re-sorting of eigenvalues is performed to account for infrequent level crossings as $`m_f`$ is varied. In order to avoid the possibility that these level crossings may have pushed a needed eigenvalue up to beyond number 19, we have excluded those quadratic fits which contain the largest eigenvalue at $`m_f=0`$ for each of the 32 configurations. The resulting root-mean-square of the fractional differences between the left- and right-hand sides of Eq. 51 is very small. The average root-mean-square of the fractional difference is $`1.3\times 10^4`$ while the largest value is $`4.3\times 10^3`$.
In Figure 31 we present a histogram of the distribution of fit parameters for the $`18\times 32=576`$, $`\delta m_i`$ values that we obtain. The majority of values are quite small, very much on the order of $`m_{\mathrm{res}}`$. While a few larger values of $`\delta m_i`$ are seen (the largest is 0.0660), the median of the distribution is $`\overline{\delta m}=0.00147`$ which is remarkably close to the value of $`m_{\mathrm{res}}=0.00124`$ found earlier for this value of $`\beta `$ and $`L_s`$.
The 4-dimensional expression for $`\overline{\psi }\psi `$ in Eq. 44 and the 5-dimensional result in Eq. 54 as a function of $`m=Zm_f`$ must agree in the continuum limit after a rescaling and overall subtraction. This must be true even if $`m_{\mathrm{res}}`$ is held fixed in physical units as $`a0`$. Therefore, in the limit of zero lattice spacing, the histogram shown in Figure 31 must approach a delta function so that $`\delta m_i`$ has the unique value $`m_{\mathrm{res}}`$. Thus, we might interpret the width of the distribution in Figure 31 as a result of $`O(a^2)`$ effects. The large size of the fluctuations relative to the central value is presumably a result of the small central value produced by our quite large separation of 16 between the walls.
## VII Hadronic Observables
We can now use the results of the previous sections to compute a variety of hadronic properties. In this section we will discuss two topics: the evaluation of the pion decay constant $`f_\pi `$ and the scaling properties of the nucleon to $`\rho `$ mass ratio. The first topic is of greatest interest since we can compute the pion decay constant using two independent methods, one of which depends directly on the residual mass determined in Section VI. The close agreement between these two approaches provides a very important consistency check of the analysis and results presented in this paper.
### A Calculation of $`f_\pi `$
In the conventional continuum formulation, the pion decay constant $`f_\pi `$ is defined through the equation
$$0\left|i\overline{\psi }\gamma _M^\mu \gamma _5t^a\psi \right|\pi ^b(\stackrel{}{p})f_\pi \frac{p^\mu \delta ^{a,b}}{\sqrt{2E_\pi (\stackrel{}{p})}}$$
(93)
where the fields $`\psi `$ and $`\overline{\psi }`$ are interpreted as conventional, Hilbert space quark operators and the pion state obeys the non-covariant normalization $`\pi (\stackrel{}{p}^{})|\pi (\stackrel{}{p})=\delta ^3(\stackrel{}{p}\stackrel{}{p}^{})`$. To be concrete we adopt the Minkowski metric $`g^{\mu \nu }`$ with signature $`(1,+1,+1,+1)`$ and a Minkowski gamma matrix convention in which $`\gamma _M^0`$ is anti-hermitian and $`\{\gamma _M^\mu ,\gamma _M^\nu \}=2g^{\mu \nu }`$. With this normalization, $`f_\pi 130`$ MeV.
Following the usual methods of lattice gauge theory, we evaluate matrix elements of the two-quark operator appearing in Eq. 93 with the Euclidean time dependence resulting from use of the evolution operator $`e^{Ht}`$ where $`H`$ is the QCD Hamiltonian. Thus, we choose to evaluate
$$f_\pi ^2\frac{m_\pi }{2}e^{m_\pi t}=\underset{T\mathrm{}}{lim}\frac{Tr\left\{e^{H(Tt)}d^3xi\overline{\psi }\gamma _M^0\gamma _5t^a\psi (\stackrel{}{x},t)e^{Ht}i\overline{\psi }\gamma _M^0\gamma _5t^a\psi (\stackrel{}{0},0)\right\}}{Tr\left\{e^{HT}\right\}}$$
(94)
where no sum over the flavor index $`a`$ is intended and the time $`t`$ is assumed sufficiently large that only the pion intermediate state contributes.
The continuum operators in Eq. 94 are easily represented as lattice, Euclidean-space expressions once the usual transition to a Euclidean-space path integral has been performed. In particular, the operators $`\psi (\stackrel{}{x})`$ and $`\overline{\psi }(\stackrel{}{x})=\psi ^{}(\stackrel{}{x})\gamma ^0`$ are replaced by the Grassmann variables $`q(\stackrel{}{x},t)`$ and $`\overline{q}(\stackrel{}{x},t)`$ respectively. Thus, we extract $`f_\pi ^2`$ from the usual Euclidean correlation function:
$$\frac{f_\pi ^2}{Z_A^2}\frac{m_\pi }{2}e^{m_\pi t}=d^3x\overline{q}\gamma ^0\gamma ^5t^aq(\stackrel{}{x},t)\overline{q}\gamma ^0\gamma ^5t^aq(\stackrel{}{0},0)$$
(95)
where now Euclidean gamma matrices appear, obeying $`\{\gamma ^\mu ,\gamma ^\nu \}=2\delta ^{\mu \nu }`$. Here we have introduced the Grassmann variables $`q`$ and $`\overline{q}`$ defined earlier in this paper so the axial current appearing in Eq. 95 is explicitly constructed from the five-dimensional quark fields $`\mathrm{\Psi }`$ and $`\overline{\mathrm{\Psi }}`$ restricted to the $`s=0`$ and $`s=L_s1`$ walls. This “local” current, $`A_\mu ^a`$ is not conserved in the full five-dimensional theory so the factor $`Z_A`$ appearing on the left hand side of Eq. 95 is needed to make a connection to the continuum axial current.
The conserved current $`𝒜_\mu ^a`$ defined in Eq. 8 must approach the corresponding, partially conserved continuum current with unit normalization, when the continuum limit is taken. Thus, to order $`a^2`$, the low energy matrix elements of $`𝒜_\mu ^a`$ and $`A_\mu ^a`$ must be proportional: $`𝒜_\mu ^a=Z_AA_\mu ^a`$. While we have computed $`f_\pi `$ using the local current $`A_\mu ^a`$ we have also compared that current to the partially conserved domain wall axial current $`𝒜_\mu ^a`$, allowing an accurate determination $`Z_A`$.
In addition to the procedure just described, there is a second, independent method that we have used to compute $`f_\pi `$. Here we use the Ward-Takahashi identity to relate $`𝒜_\mu ^a`$ and the pseudo-scalar density $`J_5^a`$:
$$\mathrm{\Delta }_\mu 𝒜_\mu ^a(x)2(m_f+m_{\mathrm{res}})J_5^a(x)$$
(96)
an expression valid for low energy matrix amplitudes . In particular, we have replaced the usual midpoint term in the exact identity of Eq. 10 by its low energy limit: $`2m_{\mathrm{res}}J_5^a(x)`$. Thus, we can also obtain $`f_\pi `$ from the correlation function:
$$\frac{f_\pi ^2}{(m_f+m_{\mathrm{res}})^2}\frac{m_\pi ^3}{8}e^{m_\pi t}=d^3x\overline{q}\gamma ^5t^aq(\stackrel{}{x},t)\overline{q}\gamma ^5t^aq(\stackrel{}{0},0)$$
(97)
where again no sum over the flavor index $`a`$ is intended. This formula involves no renormalization factors but requires knowledge of the residual mass $`m_{\mathrm{res}}`$ induced by mixing between the walls. Thus, a comparison of the values for $`f_\pi `$ obtained from Eqs. 95 and 97 provides a critical test of the analysis presented in this paper.
We will now discuss these two calculations of $`f_\pi `$ in detail. To measure the value for the renormalization factor $`Z_A`$, we compare the amplitudes of two-point functions $`C(t)`$ and $`L(t)`$ defined as
$`C(t+1/2)`$ $`=`$ $`{\displaystyle \underset{\stackrel{}{x}}{}}𝒜_0^a(\stackrel{}{x},t)\pi ^a(\stackrel{}{0},0)`$ (98)
$`L(t)`$ $`=`$ $`{\displaystyle \underset{\stackrel{}{x}}{}}A_0^a(\stackrel{}{x},t)\pi ^a(\stackrel{}{0},0).`$ (99)
The $`1/2`$ in the argument of $`C(t+1/2)`$ in Eq. 99 comes from the fact the conserved axial current $`𝒜_\mu ^a(x)`$ is not the current at lattice site $`x`$ but instead the current carried by the link between $`x`$ and $`x+\widehat{\mu }`$. We take appropriate arithmetic averages to solve the problem that $`C(t+1/2)`$ and $`L(t)`$ are not at the same location. To avoid as much systematic error as possible, we define $`Z_A(t)`$ as
$$Z_A(t)=\frac{1}{2}\left\{\frac{C(t+1/2)+C(t1/2)}{2L(t)}+\frac{2C(t+1/2)}{L(t)+L(t+1)}\right\}.$$
(100)
For $`ta^1`$, $`C(t)/L(t)`$ behaves like a constant which can be identified with $`Z_A`$. Both terms in Eq. 100 estimate this value without $`O(a)`$ error. The average of these two, incorporated in Eq. 100, further eliminates a portion of the $`O(a^2)`$ error.
Figure 32 shows the ratio $`Z_A(t)`$ defined in Eq. 100 for both a $`16^3\times 32`$ lattice with $`L_s=16`$, and $`\beta =6.0`$ as well as the same quantity for a $`8^3\times 32`$ lattice with $`L_s=48`$ and $`\beta =5.7`$. We determine the value for the renormalization factor $`Z_A`$ by calculating the average over two ranges of $`t`$: $`4t14`$ and $`18t28`$, chosen to avoid the largest time separation $`t16`$ where the errors are quite large. A jackknife error is determined, to compensate for possible correlation between the numerator and denominator in Eq. 100.
The results for $`Z_A`$ at $`\beta =6.0`$, $`16^3\times 32`$, $`M_5=1.8`$ and with different values of $`L_s`$ are listed in Table XXV. The data shows little $`L_s`$ dependence, as should be expected. Figure 32, also shows our result of $`Z_A=0.7732(14)`$ found for the $`8^3\times 32`$ lattice with $`\beta =5.7`$, $`L_s=48`$, $`M_5=1.65`$, $`m_f=0.02`$.
The results for the amplitudes for the axial vector current correlator and the pseudoscalar density correlator at $`\beta =5.7`$ and $`6.0`$ are given in Tables XXVI-XXX. They are obtained from the point-source correlators using a conventional 2-parameter fit with the pion masses extracted concurrently. We also list in the same tables the results for $`f_\pi `$ as a function of $`m_f`$ determined from the corresponding correlators with the help of $`Z_A`$ and $`m_{\mathrm{res}}`$ (Tables XXII-XXIII). These values of $`f_\pi `$ have been converted to physical units using the measured $`\rho `$ mass discussed in Section III, extrapolated to the chiral limit $`m_f+m_{\mathrm{res}}=0`$.
Next, we use a linear fit in $`m_f`$ to evaluate $`f_\pi `$ for two values of $`m_f`$. To obtain a value of $`f_\pi `$ close to that for the physical pion, we go to the chiral limit $`m_f+m_{\mathrm{res}}=0`$. For $`f_K`$ we choose for $`m_f`$ that value which gives $`m_\pi /m_\rho =0.645`$. In determining $`f_\pi `$ for the physical pion state, we did not attempt to use a value of $`m_f`$ giving the physical value for the ratio $`m_\pi /m_\rho =0.18`$ since we do not adequately know the $`m_f`$ dependence of this ratio in the relevant region. These linear fit parameters as well as the resulting values for $`f_\pi `$ and $`f_K`$ are summarized in Table XXXI. The errors given in the tables are obtained from the jackknife method.
Figure 33 shows the values for $`f_\pi `$ at $`\beta =5.7`$, $`8^3\times 32`$, $`L_s=48`$ as a function of $`m_f`$ and the linear fits through all the $`m_f`$ points. The results obtained from the pseudoscalar-pseudoscalar correlator are higher than those from the axial-axial correlator. The two linear fits give $`f_\pi =127(4)\mathrm{MeV}`$, $`f_K=145(4)\mathrm{MeV}`$ and $`f_\pi =132(4)\mathrm{MeV}`$, $`f_K=154(4)\mathrm{MeV}`$ respectively. When the lattice volume is increased to $`16^3\times 32`$ (Figure 34), the difference between the linear fits from the two methods becomes smaller. We obtain $`f_\pi =133(4)\mathrm{MeV}`$, $`f_K=149(2)\mathrm{MeV}`$ and $`f_\pi =125(4)\mathrm{MeV}`$, $`f_K=149(2)\mathrm{MeV}`$ from the two correlators. The values for $`f_\pi (m_f)`$ obtained from the two methods should agree for all values of $`m_f`$ since they are related by a Ward-Takahashi identity that should become exact in the continuum limit. Presumably the visibly different slopes seen in Figures 33 and 34 are the result of order $`a^2`$ errors.
We also calculate $`f_\pi `$ at a weaker coupling. Figure 5 shows our results for $`\beta =6.0`$, $`16^3\times 32`$, $`L_s=16`$ on 85 configurations. The two independent calculations give very consistent results. We have $`f_\pi =137(11)\mathrm{MeV}`$, $`f_K=156(8)\mathrm{MeV}`$ from the axial vector current correlator and almost the same values from the pseudoscalar correlator. Our results for $`f_\pi `$ at both $`\beta =5.7`$ and $`6.0`$ agree well with the experimental value of $`130\mathrm{MeV}`$, while the values for $`f_K`$ may be somewhat smaller than the experimental value of $`160\mathrm{MeV}`$ as is expected from quenched chiral perturbation theory arguments and naive scaling considerations . Note, in Table XXXI we also list $`f_K/f_\pi `$ with jackknifed errors for the ratio. Here the statistical errors are now well below the systematic errors that might be expected in the $`m_f+m_{\mathrm{res}}0`$ extrapolation. The values shown for $`f_K/f_\pi `$ agree on the 5% level between methods of determination and different lattice spacings but are systematically below the experimental value of 1.21.
This same analysis was done using the amplitudes calculated from the point-source correlators but making a 1-parameter fit using the pre-determined pion masses computed from the more mass accurate measurements based on the wall-source correlators. This method gives consistent results with slightly smaller errors. The results are not listed here.
The reasonable agreement of our domain wall results with the experimental values and their relative insensitivity to $`a`$ is encouraging. Similar results were obtained at $`\beta =6.0`$ for smaller values of $`L_s`$ with somewhat larger errors in Ref. . Of special interest here is the comparison that we make between the two methods of determining $`f_\pi `$, which is done here for the first time. As can be seen from Eq. 97, the determination of $`f_\pi `$ from $`\pi ^a(x)\pi ^a(0)`$ depends directly on $`m_{\mathrm{res}}`$. Thus, the comparison of these two methods is an important check of our understanding of the chiral properties of the domain wall formulation. The ratio of these two quantities extrapolated to the point $`m_f+m_{\mathrm{res}}=0`$ provides an interesting figure of merit for the present calculation. We find $`(f_\pi )_{PP}/(f_\pi )_{AA}=1.00(10)`$ and 0.96(10) for $`L_s=16`$ and 24 respectively. However, if instead of the values of $`m_{\mathrm{res}}`$ given in Table XXII, we use the $`x`$-intercepts -0.0031(7) and -0.0030(9) quoted earlier and obtained from the $`A_0^a(x)A_0^a(0)`$ values of $`m_\pi ^2`$, we find $`(f_\pi )_{PP}/(f_\pi )_{AA}=1.20(12)`$ for both the $`L_s=16`$ and 24 cases. While these ratios each differ from 1 by two standard deviations, they are independent calculations and demonstrate the good chiral properties of domain wall fermions.
### B Continuum limit of $`m_N/m_\rho `$
Here we combine the hadron mass results tabulated in Section III to examine the behavior of the nucleon to $`\rho `$ mass ratio as $`\beta `$ varies between 5.7 to 6.0. First we evaluate $`m_N`$ and $`m_\rho `$ in the limit $`m_f+m_{\mathrm{res}}=0`$. We did not use the value of $`m_f`$ which gives the physical ratio, $`m_\pi /m_\rho =0.18`$ for the reasons outlined in the previous section. In Table XXXII we give the resulting mass ratios as well as the lattice spacings in physical units as determined from $`m_\rho `$ evaluated at $`m_f+m_{\mathrm{res}}=0`$. Note, no contribution to the quoted error for these mass ratios arising from the uncertainty in this choice of $`m_f`$ has been included.
The relatively large variation of $`m_N/m_\rho `$ with $`\beta `$ suggests that the errors shown in Table XXXII may be underestimated and makes a simple $`a^2`$ extrapolation to the continuum limit somewhat uncertain. Nevertheless the result of such an extrapolation to $`a0`$ is $`m_N/m_\rho =1.37(5)`$. Perhaps more interesting is a comparison with similar quantities computed at comparable lattice spacings and volumes using Wilson and staggered fermions. For staggered fermions at $`\beta =6.0`$ on comparable volumes, one finds $`m_N/m_\rho =1.47(3)`$, a somewhat larger and less physical value than the 1.42(4) and 1.38(4) results obtained here for $`L_s=16`$ and 24. However, this comparison is made somewhat ambiguous by the significant finite size effects seen in staggered calculations when going from our $`16^3\times 32`$ to larger volumes . For Wilson fermions, as reported in Ref. , one deduces $`m_N/m_\rho =1.37(2)`$ by linear interpolation between the $`\beta =5.93`$ and 6.17 values presented, a number remarkably close to our domain wall value. When comparing these values, it is important to recall that our $`8^3`$ and $`16^3`$ spatial volumes are not yet infinite and, as discussed in Section III, corrections on the order of a few percent are expected.
### C Determining the chiral condensate $`\overline{q}q`$
Finally we use the results presented earlier to estimate the size of the chiral condensate $`\overline{q}q`$. Naively, one might expect that a physical value for $`\overline{q}q`$ could be easily identified in Table XX as the $`m_f`$-independent term $`a_0`$, defined in Eq. 55. This quantity represents a simple extrapolation of $`\overline{q}q(m_f)`$ from large mass down to the point $`m_f=0`$. Given the volume independence seen for the parameter $`a_0`$ when comparing the $`\beta =5.7`$, $`8^3`$ and $`16^3`$ volumes in Table XX, it is natural to expect that such a choice minimizes the sensitivity to the finite-volume zero mode effects that give rise to the more singular $`a_1`$ term.
However, there are other issues that must be addressed. Perhaps most obvious is the fact that the point $`m_f=0`$ is not the physical chiral limit because the effects of $`m_{\mathrm{res}}`$ have been ignored. This is easily remedied by using the slope $`a_1`$, to extrapolate to the physical point $`m_f+m_{\mathrm{res}}=0`$. The resulting estimate of $`\overline{q}q`$, in lattice units, is given as the fourth column in Table XXXIII. However, because $`\overline{q}q`$ is a quadratically divergent quantity, we cannot expect that all the chiral symmetry breaking effects of domain wall mixing are removed by this choice of $`m_f`$. In contrast to many physical quantities, $`\overline{q}q`$ receives contributions from energy scales much larger than those for which $`m_{\mathrm{res}}`$ represents the complete effect of chiral symmetry breaking. Thus, we should expect additional contributions to $`\overline{q}q`$ of order $`e^{\alpha L_s}/a^3m_{\mathrm{res}}/a^2`$. This is born out in Table XXXIII where we see that the differences between $`\overline{q}q`$ for the two different values of $`L_s`$ at a given $`\beta `$ are of the same order as the difference between the values with and without the extrapolation to $`m_f=m_{\mathrm{res}}`$.
This unwanted $`m_{\mathrm{res}}/a^2`$ contribution to $`\overline{q}q`$ can only be controlled by explicitly taking the limit $`L_s\mathrm{}`$. We do not at present have the numerical results to permit such an extrapolation. Therefore, we will use the $`\beta =6.0`$, $`L_s=24`$ result as our best approximation to such a limit and interpret the difference between the $`L_s=16`$ and 24 values as an estimate of the systematic error, $`10\%`$. Given the value of $`Z_S(\overline{\mathrm{MS}},2\mathrm{G}\mathrm{e}\mathrm{V})=0.619(25)`$ for $`\beta =6.0`$ quoted earlier and the results for the lattice spacing in physical units in Table XXXII, we can determine $`\overline{q}q`$ in physical units. The results for $`L_s=16`$ and 24, $`(245(7)\mathrm{MeV})^3`$ and $`(256(8)\mathrm{MeV})^3`$, are included in Table XXXIII, where only the statistical error is displayed. The agreement between these numbers and phenomenological estimates of the chiral condensate is satisfactory, for example the value of $`\frac{1}{2}(\overline{u}u+\overline{d}d)_{\overline{\mathrm{MS}},1\mathrm{G}\mathrm{e}\mathrm{V}}=(229\pm 9\mathrm{M}\mathrm{e}\mathrm{V})^3`$ obtained in Ref. . Note the $`e^{\alpha L_s}/a^3`$ uncertainty present in our calculation does not have an analogue in the properly regulated continuum theory. While $`\overline{\psi }\psi `$ does contain a quadratically divergent piece in the continuum theory, this is eliminated for the chirally symmetric choice $`m_{\mathrm{quark}}=0`$. This choice is not available in a domain wall fermion calculation without taking the $`L_s\mathrm{}`$ limit. Of course, the other lattice methods for directly computing $`\overline{\psi }\psi `$ have equal or more severe difficulties.
Finally it is interesting to compare the $`\beta =5.7`$ and $`\beta =6.0`$ results for $`\overline{q}q`$. Since we do not at present have a reliable determination of the needed renormalization constant, $`Z_S`$, for the stronger $`\beta =5.7`$ coupling, we do not attempt to quote a physical value. However, the ratio of the unrenormalized lattice numbers given in Table XXXIII for $`\overline{q}q_{(L_s=32,\beta =5.7)}/\overline{q}q_{(L_s=24,\beta =6.0)}=4.8(2)`$ is reasonably consistent with the ratio expected from naive scaling $`a_{(L_s=32,\beta =5.7)}^3/a_{(L_s=24,\beta =6.0)}^3=7.4(4)`$.
Given the values now determined for $`\overline{q}q`$, $`f_\pi `$ and quark mass, it is natural to test the degree to which the Gell-Mann-Oakes-Renner relation
$$f_\pi ^2\frac{m_\pi ^2}{48(m_f+m_{\mathrm{res}})}=\overline{q}q$$
(101)
is obeyed. However, the form of this equation reveals an important difficulty. At what value of $`m_f`$ should the ratio $`m_\pi ^2/(m_f+m_{\mathrm{res}})`$ be computed? In full QCD, this ratio becomes a constant for small quark mass. As we have seen earlier, this is not the case in the quenched approximation where one expects non-linearities.
We might try to determine the proper treatment of these non-linearities by returning to the underlying equation, Eq. 60, from which the Gell-Mann-Oakes-Renner relation is derived. However, this is somewhat complex. Both sides of this original equation have a mass dependence which comes from the contribution of the pion pole term and other physical states, all influenced by the quenched approximation, as well as the quadratically divergent terms in $`\overline{q}q`$ and the contact term in $`\chi _{\pi \pi }`$. Thus, while the underlying Eq. 60 will be obeyed exactly in our calculation, there is considerable ambiguity in deciding how to extract a quenched generalization of the Gell-Mann-Oakes-Renner relation, Eq. 101.
Here we will simply compare the right- and left-hand-sides of Eq. 101 by replacing the ratio $`m_\pi ^2/(m_f+m_{\mathrm{res}})`$ by the slope $`b`$ obtained at larger masses, $`m_f0.01`$ and given in Tables XVI, XVII and XIX. The results from the left hand side of Eq. 101 are given in Table XXXIII. Given our uncertainty in determining $`\overline{q}q`$ and the significant non-linearities we see in $`m_\pi ^2`$, the agreement seen between the fourth and fifth columns in Table XXXIII is within our errors.
## VIII Conclusions
We have presented the results of detailed studies of quenched lattice QCD using domain wall fermions, with particular attention paid to the lowest order chiral symmetry breaking effects of finite $`L_s`$ and the behavior of the theory for small values of $`m_f`$. A major difficulty in studying the small $`m_f`$ behavior of the theory is the presence of topological near-zero modes which are unsuppressed in the quenched theory. These are a result of the improved character of the domain wall fermion operator, which has an Atiyah-Singer index at finite lattice spacing and $`L_s\mathrm{}`$. However, these zero-modes complicate the quenched theory and demonstrate that the quenched approximation is considerably more treacherous than might have been originally expected. We have seen how these modes produce the expected $`1/m_f`$ divergence in $`\overline{q}q`$ for small $`m_f`$ and distort correlation functions used to measure the properties of the pion. By working on larger volumes, we found that the effects of these modes were dramatically reduced, as expected. We were then able to see a common pion mass determined from different correlators.
We have determined or constrained the value for the residual mass, $`m_{\mathrm{res}}`$, which enters the effective quark mass for low-energy physics as $`m_{\mathrm{eff}}=m_f+m_{\mathrm{res}}`$, a number of ways and found good agreement. The residual mass was measured from the extra, finite $`L_s`$ term in the divergence of the conserved axial current and from the explicitly determined lowest eigenvalues of the hermitian domain wall fermion operator. These two determinations agree within errors. We have also determined the difference in $`m_{\mathrm{res}}`$ for two values of $`L_s`$ from the pion mass and find this agrees with the results from our explicitly calculated $`m_{\mathrm{res}}`$. Lastly, agreement for $`f_\pi `$ as calculated from axial vector and pseudoscalar correlators requires knowledge of $`m_{\mathrm{res}}`$ and the agreement serves as a further check.
While our data for weaker couplings does not clearly demonstrate that $`m_{\mathrm{res}}0`$, we have seen it fall to 1 MeV for $`L_s=48`$ at $`\beta =6.0`$. For $`L_s=16`$, a practical value for studies of low energy hadronic physics and matrix elements, $`m_{\mathrm{res}}`$ has a value of 3.87(16) MeV, roughly 1/30 of the strange quark mass. Even at stronger couplings, where the lattice spacing is $`a^11`$ GeV, we have measured $`m_{\mathrm{res}}`$ to also be about 1/14 of the strange quark mass, although here $`L_s=48`$ was required. Thus, we see domain wall fermions producing the desired light surface states with small mixing, even for relatively strong couplings.
We have measured hadron masses and $`f_\pi `$ for lattice scales 1 GeV $`<a^1<`$ 2 GeV and have studied scaling in this region. Our determinations of $`f_\pi `$ involve not only $`m_{\mathrm{res}}`$ as mentioned above but also the measurement of the Z-factor for the local axial current. We find $`f_\pi /m_\rho `$ evaluated at the $`m_f+m_{\mathrm{res}}=0`$ point to be scaling very well, while for $`m_N/m_\rho `$ the scaling violations may be at the 6% level. However, scaling seems at least as good as that seen for staggered fermions at similar lattice spacings and similar to that found for Wilson fermions with a clover term . This is in accord with general expectations that finite lattice spacing errors will enter domain wall fermion amplitudes at $`O(a^2)`$ .
Our results demonstrate that quenched domain wall fermions do exhibit the desired good chiral properties, even at relatively strong couplings. The residual quark mass effects, which break the full global symmetries to leading order in $`a`$, can be eliminated by an appropriate choice of $`m_f`$, so that low energy physics should be well described by an effective theory with the continuum global symmetries. Quenched chiral logarithm effects may appear for quenched domain wall fermion simulations, as they do for other fermion formulations, but present no new difficulties. For large enough volumes, the effects of topological near-zero modes are suppressed and the small $`m_f`$ region can be investigated. For larger values of $`m_f`$, where these zero mode effects are suppressed by the quark mass, one has a formulation of lattice QCD with the full global symmetries realized to order $`a^2`$ and an effective quark mass of $`m_f+m_{\mathrm{res}}`$. Thus, the domain wall formulation provides a powerful new tool which can be used, even within the quenched approximation, to study many of the outstanding problems in particle and nuclear physics for which chiral symmetry plays an important role.
Note added: After this paper was essentially complete, the recent work of the CP-PACS collaboration became available . The reader is referred to this paper for another discussion of some of the topics presented here.
## Acknowledgments
The authors would like to acknowledge useful discussion with Shoichi Sasaki, Thomas Manke, T. D. Lee, Robert Edwards and Mike Creutz. We thank RIKEN, Brookhaven National Laboratory and the U.S. Department of Energy for providing the facilities essential for the completion of this work. Finally, we acknowledge with gratitude the Ritz diagonalization program provided by Robert Edwards.
The numerical calculations at $`\beta =5.7`$ and 6.0 were done on the 400 Gflops QCDSP computer at Columbia University and the 600 Gflops QCDSP computer at the RIKEN-BNL Research Center . The $`\beta =5.85`$ results were calculated at NERSC. We acknowledge the MILC collaboration whose software provided the basis for this $`\beta =5.85`$ calculation. This research was supported in part by the DOE under grant # DE-FG02-92ER40699 (Columbia), in part by the NSF under grant # NSF-PHY96-05199 (Vranas), in part by the DOE under grant DE-AC02-98CH10886 (Soni-Dawson), in part by the RIKEN-BNL Research Center (Blum-Wingate-Ohta) and in part by the Max-Kade Foundation (Siegert).
|
warning/0007/hep-th0007244.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Ever since ’t Hooft and Polyakov found a monopole solution in the SU(2) Yang-Mills-Higgs theory, solitons in field theories have been studied extensively. Our understanding of monopole solutions has been greatly enhanced by an existence proof for static solutions by Taubes and the construction of monopole solutions started by Ward . This process was not matched by quite the same progress in our understanding of the Abrikosov solutions of the Ginzburg-Landau theory, although one might have expected that the Abelian Higgs theory in 2+1 dimensions is actually simpler than the SU(2) Yang-Mills-Higgs theory in 3+1 dimensions. Again an existence proof was given by Taubes . However, only superimposed vortices can be described explicitly and no explicit construction of separated vortices is known. In this paper, we want to give the solution for two vortices close together in terms of an expansion in the parameters which describe the relative location.
In Sections 2 and 3, we study a model for one complex field. Here the calculations are simpler than in the Ginzburg-Landau theory which is our second model. The first model has, however, some peculiar (unphysical) features. Assuming the most symmetric form in terms of angular dependence, only two smooth vortices can be superimposed, and when ’pulled apart’, they develop a singularity at third order. In the Ginzburg-Landau model this does not happen. In fact, delicate cancellations take place to make the expansion smooth, at least up to third order. In this model the radial functions are given as solutions of certain linear ordinary differential equations. This is discussed in Section 4.
## 2 Vortex solutions and zero modes in a simple model
Our first model is a model for a pair of real fields $`\varphi ^a(\stackrel{}{x})`$, a,b = 1,2, or equivalently, for a complex field $`\varphi =\varphi _1+i\varphi _2`$. The Lagrangian density of the model reads
$$=_{[i}\varphi ^a_{j]}\varphi ^b^{[i}\varphi _a^{j]}\varphi _b+(1|\varphi |^2)^2|\varphi |^2,$$
(2.1)
where $`a,b=1,2`$ labels the components of the Higgs field and $`i,j=1,2`$ are the space indices. The square brackets mean antisymmetrization,
$$_{[i}\varphi ^a_{j]}\varphi ^b=(_i\varphi ^a)(_j\varphi ^b)(_j\varphi ^a)(_i\varphi ^b).$$
(2.2)
We are working in 2-dimensional Euclidean space, i.e., the space indices can be raised and lowered without any change in the formulas. The indices which label the components of the Higgs field can also be raised and lowered without any change. In terms of the complex field $`\varphi `$ the Euler-Lagrange equation reads
$$_i\varphi ^{}_j(^{[i}\varphi ^{j]}\varphi ^{})=(1|\varphi |^2)|\varphi |\frac{}{\varphi }(1|\varphi |^2)|\varphi |.$$
(2.3)
Any solution of the equation
$$2det(\frac{\varphi ^a}{x^i})=\pm (1|\varphi |^2)|\varphi |$$
(2.4)
solves the equation of motion (2.3). Note that Eq. (2.4) is a first order equation whereas Eq. (2.3) is of second order. So we would expect that (2.4) is somewhat easier to solve than (2.3). For different types of models, this reduction of order was first introduced by Bogomolnyi . That is why we call Eq. (2.4) the Bogomolnyi equation here. Any solution of (2.4) also attains the lower bound in the following inequality,
$$A=_{𝐑^2}d^2x\frac{16\pi }{15}|Q|,$$
(2.5)
where
$$Q=\frac{15}{8\pi }_{𝐑^2}iϵ_{ij}(1|\varphi |^2)|\varphi |(^i\varphi )(^j\varphi ^{})d^2x$$
(2.6)
is the winding number. Finally, all finite-action solutions actually solve the Bogomolnyi equation, so we do not miss out on any by concentrating on the first order equation.
We now seek to attain a smooth finite-action solution of Eq. (2.4). For
$$\varphi =f(r)e^{in\theta }$$
(2.7)
Eq. (2.4) reduces to
$$\frac{nf(r)f^{^{}}(r)}{r}=\frac{1}{2}(1f^2)f.$$
(2.8)
Since $`f0`$ as $`r0`$ (otherwise $`\varphi `$ in (2.7) is not defined at the origin), we have
$$f=\mathrm{tanh}\frac{r^2}{4n}.$$
(2.9)
The solution $`\varphi `$ in (2.7) with f(r) given by (2.9) is defined in the whole of $`𝐑^2`$ and is clearly a $`C^{\mathrm{}}`$ function in $`𝐑^2\{0\}`$. Since
$`f12\mathrm{exp}{\displaystyle \frac{r^2}{2n}}\mathrm{as}r\mathrm{},`$ (2.10)
$`\varphi `$ has the right asymptotic behaviour for a solution with winding number n. We still have to ensure that $`\varphi `$ is $`C^{\mathrm{}}`$ at the origin. There we use the Taylor expansion of $`f`$,
$`f={\displaystyle \underset{K=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2^{2k}(2^{2k}1)B_{2k}}{(2k)!}}({\displaystyle \frac{r^2}{4n}})^{2k1}={\displaystyle \frac{r^2}{4n}}{\displaystyle \frac{1}{3}}({\displaystyle \frac{r^2}{4n}})^3+\mathrm{}`$ (2.11)
where $`B_k`$ is the $`k^{th}`$ Bernoulli number. We see that for $`n=2`$ and only for $`n=2`$, $`\varphi `$ is a polynomial in $`x^i`$. In this model, we have the (somewhat peculiar) situation that within the most natural ansatz (2.7) smooth finite action solutions exist only for $`n=2`$, i.e., we only have a solution of the form (2.7) for 2 vortices.
We have found the solution for two vortices sitting on top of each other, which we now denote by $`\widehat{\varphi }`$. To extend our study to two vortices slightly apart we consider $`\varphi =\widehat{\varphi }+\gamma `$, where $`\gamma `$ is very small, and we solve the Bogomolyni equation, linearized in $`\gamma `$. Equation (2.4) becomes
$$(f^{^{}}\mathrm{cos}\theta \mathrm{cos}2\theta +\frac{2}{r}f\mathrm{sin}\theta \mathrm{sin}2\theta )\frac{\gamma ^2}{x^2}+(f^{^{}}\mathrm{sin}\theta \mathrm{sin}2\theta +\frac{2}{r}f\mathrm{cos}\theta \mathrm{cos}2\theta )\frac{\gamma ^1}{x^1}$$
$$(f^{^{}}\mathrm{sin}\theta \mathrm{cos}2\theta \frac{2}{r}f\mathrm{cos}\theta \mathrm{sin}2\theta )\frac{\gamma ^2}{x^1}(f^{^{}}\mathrm{cos}\theta \mathrm{sin}2\theta \frac{2}{r}f\mathrm{sin}\theta \mathrm{cos}2\theta )\frac{\gamma ^1}{x^2}$$
$$=\frac{1}{2}(13f^2)(\gamma ^1\mathrm{cos}2\theta +\gamma ^2\mathrm{sin}2\theta )$$
(2.12)
We find a 2-parameter family of zero modes,
$`\gamma (r)=[\alpha +\beta +ı(\alpha \beta )]h(r)\mathrm{with}h(r)={\displaystyle \frac{\mathrm{sinh}\frac{r^2}{8}}{\mathrm{cosh}^3\frac{r^2}{8}}}.`$ (2.13)
These zero modes are $`C^{\mathrm{}}`$ functions which vanish exponentially at infinity. By a rotation, one of the parameters could be removed and the vortices could be positioned on the $`x`$-axis, say. Since this does not simplify the calculations significantly, we will retain both parameters. Retaining the two parameters would also be necessary for a study of vortex scattering in the slow-motion approximation. This study is not done in this paper.
## 3 The quadratic and cubic terms
We now consider $`\varphi =\widehat{\varphi }+\gamma +\delta `$, and equate the second order terms in the Bogomolyni equation (2.4). This leads to the equation
$`{\displaystyle \frac{2}{r}}(f^{^{}}\mathrm{cos}2\theta {\displaystyle \frac{\delta ^2}{\theta }}+2f\mathrm{sin}2\theta {\displaystyle \frac{\delta ^2}{r}}f^{^{}}\mathrm{sin}2\theta {\displaystyle \frac{\delta ^1}{\theta }}+2f\mathrm{cos}2\theta {\displaystyle \frac{\delta ^1}{r}})`$
$`=(\alpha ^2+\beta ^2)fh^2({\displaystyle \frac{1}{f^2}}3){\displaystyle \frac{1}{2}}fh^2(3+{\displaystyle \frac{1}{f^2}})[\alpha ^2(\mathrm{cos}2\theta +\mathrm{sin}2\theta )^2`$
$`+2\alpha \beta (\mathrm{cos}^22\theta \mathrm{sin}^22\theta )+\beta ^2(\mathrm{cos}2\theta \mathrm{sin}2\theta )^2]`$
$`+(13f^2)(\delta ^1\mathrm{cos}2\theta +\delta ^2\mathrm{sin}2\theta )`$ (3.1)
with $`f(r)`$ given in (2.9) and $`h(r)`$ given in (2.13).
With $`\delta `$ of the form
$$\delta =\alpha ^2F(r,\theta )+2\alpha \beta G(r,\theta )+\beta ^2H(r,\theta ),$$
(3.2)
we obtain the following equation for $`F(r,\theta )`$,
$`{\displaystyle \frac{2}{r}}(f^{^{}}\mathrm{cos}2\theta {\displaystyle \frac{F^2}{\theta }}+2f\mathrm{sin}2\theta {\displaystyle \frac{F^2}{r}}f^{^{}}\mathrm{sin}2\theta {\displaystyle \frac{F^1}{\theta }}+2f\mathrm{cos}2\theta {\displaystyle \frac{F^1}{r}})`$
$`=h^2({\displaystyle \frac{1}{f}}3f){\displaystyle \frac{h^2}{2}}(3f+{\displaystyle \frac{1}{f}})(\mathrm{cos}2\theta +\mathrm{sin}2\theta )^2`$
$`+(13f^2)(F^1\mathrm{cos}2\theta +F^2\mathrm{sin}2\theta ).`$ (3.3)
To solve this equation we seek a solution of the form
$$F=f_1(r)\mathrm{exp}^{ı2\theta }ıf_2(r)\mathrm{exp}^{ı2\theta }.$$
(3.4)
The ansatz (3.4) leads to two decoupled equations for $`f_1`$ and $`f_2`$. In terms of the variable $`\xi =r^2/8`$, they read
$$\frac{df_1}{d\xi }+\frac{1}{f}(3f^21\frac{df}{d\xi })f_1=\frac{h^2}{2f^2}(19f^2),$$
(3.5)
$$\frac{df_2}{d\xi }+\frac{1}{f}(3f^21+\frac{df}{d\xi })f_2=\frac{h^2}{2f}(1+3f^2).$$
(3.6)
The general solutions to equation (3.5) is
$$f_1=\frac{1}{\mathrm{cosh}^2\xi }(\frac{3\mathrm{sinh}\xi }{2\mathrm{cosh}^3\xi }\frac{\mathrm{sinh}\xi }{\mathrm{cosh}\xi }+C_1)$$
(3.7)
The function $`f_1`$ is a $`C^{\mathrm{}}`$ function for $`0<\xi <\mathrm{}`$. For $`\xi 0`$, $`f_1C_1`$ holds. This implies that $`C_1=0`$; otherwise $`F`$ in (3.4) is not defined at the origin. Therefore, $`f_1`$ reads
$$f_1=\frac{3\mathrm{sinh}\xi }{2\mathrm{cosh}^5\xi }\frac{\mathrm{sinh}\xi }{\mathrm{cosh}^3\xi }.$$
(3.8)
The expansion of $`f_1`$ near the origin is of the form
$$f_1=\underset{k=1}{\overset{\mathrm{}}{}}a_k\xi ^k=\underset{k=1}{\overset{\mathrm{}}{}}a_k(\frac{r^2}{8})^k.$$
(3.9)
Hence, the first term in (3.4) is a $`C^{\mathrm{}}`$ function of $`x^1`$ and $`x^2`$ at the origin. We also see that $`f_1`$ vanishes exponentially at infinity. So its contribution to $`\varphi `$ does not change the winding number (2.6) which is a multiple of the action.
A similar calculation yields a one parameter family of solutions to Eq. (3.6), namely
$$f_2=\frac{\mathrm{sinh}\xi }{2\mathrm{cosh}^3\xi }\frac{3\mathrm{sinh}^3\xi }{2\mathrm{cosh}^5\xi }+C_2\frac{\mathrm{sinh}^2\xi }{\mathrm{cosh}^4\xi }.$$
(3.10)
In contrast to $`f_1`$, all the solutions $`f_2`$ are acceptable. In fact, for all $`C_2`$, $`f_2`$ is of the form
$$f_2=\underset{k=1}{\overset{\mathrm{}}{}}b_k\xi ^k=\underset{k=1}{\overset{\mathrm{}}{}}b_k(\frac{r^2}{8})^k$$
(3.11)
near the origin, and therefore the second term in (3.4) is in $`C^{\mathrm{}}(𝐑^2)`$. The winding number and the action are also not altered because $`f_2`$ decays exponentially at infinity.
The functions $`G`$ and $`H`$ in (3.2) can be found in the same way. If we put all results together, we obtain the second order terms,
$$\delta =(\alpha ^2+\beta ^2)f_1(r)\mathrm{exp}^{ı2\theta }+ı(\alpha i\beta )^2f_2(r)\mathrm{exp}^{ı2\theta },$$
(3.12)
where $`f_1`$ and $`f_2`$ are given by (3.8) and (3.10), respectively.
To find the cubic terms, we consider $`\varphi =\widehat{\varphi }+\gamma +\delta +ϵ`$, with $`\gamma `$ given in (2.13) and $`\delta `$ given by (3.12). We set $`\beta =0`$ and concentrate on
$$ϵ=\alpha ^3I(r,\theta ).$$
(3.13)
For the Bogomolnyi equation to hold, $`I`$ must satisfy
$$\frac{2}{r}(f^{^{}}\mathrm{cos}2\theta \frac{I^2}{\theta }+2f\mathrm{sin}2\theta \frac{I^2}{r}f^{^{}}\mathrm{sin}2\theta \frac{I^1}{\theta }+2f\mathrm{cos}2\theta \frac{I^1}{r})$$
$$+h^{^{}}(2f_1\mathrm{cos}2\theta +2f_2\mathrm{sin}2\theta )+h^{^{}}(2f_1\mathrm{sin}2\theta +2f_2\mathrm{cos}2\theta )$$
$$=3f^2(I^1\mathrm{cos}2\theta +I^2\mathrm{sin}2\theta )f_2(\mathrm{cos}2\theta +\mathrm{sin}2\theta )$$
$$3fh(\mathrm{cos}2\theta +\mathrm{sin}2\theta )(f_12f_2\mathrm{cos}2\theta \mathrm{sin}2\theta )3(\mathrm{cos}2\theta +\mathrm{sin}2\theta )h^2$$
$$+(I^1\mathrm{cos}2\theta +I^2\mathrm{sin}2\theta )+\frac{h}{2}[f_1(\mathrm{cos}2\theta +\mathrm{sin}2\theta )f_2(\mathrm{cos}2\theta +\mathrm{sin}2\theta )]$$
$$+\frac{h^3}{2}(\mathrm{cos}2\theta +\mathrm{sin}2\theta )^3\frac{h^3}{f^2}(\mathrm{cos}2\theta +\mathrm{sin}2\theta )$$
$$\frac{1}{f}(\mathrm{cos}2\theta +\mathrm{sin}2\theta )(f_12f_2\mathrm{cos}2\theta \mathrm{sin}2\theta )+\frac{1}{2f^2}(I^1\mathrm{cos}2\theta +I^2\mathrm{sin}2\theta )^3$$
(3.14)
To solve equation (3.14) we seek a solution of the form
$$I^1=g_1(\xi )+g_2(\xi )(\mathrm{cos}4\theta \mathrm{sin}4\theta ),$$
$$I^2=g_1(\xi )g_2(\xi )(\mathrm{cos}4\theta +\mathrm{sin}4\theta ).$$
(3.15)
This implies that $`g_1`$ and $`g_2`$ must satisfy the equations
$$\frac{dg_1}{d\xi }+(3f\frac{1}{f})g_1=\frac{f_1+f_2}{f}\frac{dh}{d\xi }6hf_1+\frac{9}{2}hf_2\frac{hf_2}{2f^2}\frac{h^3}{4f^3}\frac{9h^3}{4f},$$
(3.16)
$$\frac{dg_2}{d\xi }(\frac{1}{f}3f+\frac{2}{f}\frac{df}{d\xi })g_2=\frac{hf_2}{2f^2}\frac{3hf_2}{2}\frac{h^3}{4f^3}\frac{h^3}{4f}.$$
(3.17)
The general solution to Eq. (3.17) is
$$g_2=\frac{\mathrm{sinh}\xi }{4\mathrm{cosh}^5\xi }\frac{5\mathrm{sinh}^3\xi }{4\mathrm{cosh}^7\xi }+C_2\left(\frac{\mathrm{sinh}^2\xi }{2\mathrm{cosh}^4\xi }\frac{3\mathrm{sinh}^4\xi }{2\mathrm{cosh}^6\xi }\right)+C_3\frac{\mathrm{sinh}^3\xi }{\mathrm{cosh}^5\xi }.$$
(3.18)
All solutions (3.18) decay exponentially at infinity. For $`r0`$, however,
$$g_2(r)=\frac{1}{24}r^2+\mathrm{}$$
(3.19)
Hence, $`I`$ in (3.15) is not a $`C^{\mathrm{}}`$ function on $`𝐑^2`$. Our expansion gets singular at third order for the ansatz (3.15). In the next section we will discuss a realistic model in which a similar pattern emerges but no singularities occur.
## 4 Abrikosov vortices
The Ginzburg-Landau theory of a superconductor in a magnetic field in direction $`z`$ is given by the Lagrangian density
$$=\frac{1}{4}F_{ij}F^{ij}+\frac{1}{2}(D_i\varphi )(D^i\varphi )^{}+\frac{\lambda }{8}(\varphi ^21)^2,$$
(4.1)
where $`\varphi `$ is the complex Higgs field, and $`D_i\varphi =_i\varphi iA_i\varphi `$ and $`F_{ij}=_iA_j_jA_i`$ in terms of the gauge potentials $`A_i,i=1,2`$. The Euler-Lagrange equations are
$$D_iD^i\varphi =\frac{\lambda }{2}\varphi (1\varphi ^2),_iF^{ij}=\frac{ı}{2}[\varphi (D^j\varphi )^{}\varphi ^{}D^j\varphi ]$$
(4.2)
In the special case $`\lambda =1`$ it can be shown that all finite action solutions of Eq. (4.2) satisfy the first-order Bogmolnyi equations ,
$$F_{12}=\frac{1}{2}(1\varphi ^2),D_1\varphi =iD_2\varphi .$$
(4.3)
It has also been shown that a $`2n`$-parameter family of solution of (4.3) exists with winding number
$$n=\frac{1}{2\pi }_{𝐑^2}F_{12}d^2x.$$
(4.4)
This family describes $`n`$ vortices sitting at $`n`$ position in space.
Even for $`n`$ vortices sitting on top of each other, the solution is not known explicitly in terms of elementary functions. It is known , however, that this solution is of the form
$$\varphi =f(r)e^{ın\theta },A_i=\frac{na(r)}{r^2}\epsilon _{ij}x^j,$$
(4.5)
where $`f`$ and $`a`$ satisfy
$$rf^{}n(1a)f=(2n/r)a^{}+f^21=0$$
(4.6)
and
$$f(0)=a(0)=0,\underset{r\mathrm{}}{lim}f(r)=\underset{r\mathrm{}}{lim}a(r)=1.$$
(4.7)
In the following, we restrict our attention to $`n=2`$ and use the solution (4.5) as the zero order term in an expansion in the separation parameters. The first order terms are given by the two zero modes describing the separation of the votices. These were found by Weinberg . Using his results we can write, up to quadratic terms,
$$\varphi =fe^{2ı\theta }+2(\alpha +ı\beta )kf+\alpha ^2\psi +\alpha \beta \varphi +\beta ^2\chi +\mathrm{},$$
(4.8)
$$A_1+ıA_2=ı\frac{2a}{r}e^{ı\theta }2ı(\alpha +ı\beta )(k^{}+\frac{2k}{r})e^{ı\theta }$$
$$+\alpha ^2(B_1+ıB_2)+\alpha \beta (C_1+ıC_2)+\beta ^2(D_1+ıD_2)+\mathrm{}$$
(4.9)
Here the radial function $`k(r)`$ satisfies
$$k^{\prime \prime }+\frac{1}{r}k^{}(f^2+\frac{4}{r^2})k=0,$$
(4.10)
with
$$\underset{r0}{lim}r^2k=1,\underset{r\mathrm{}}{lim}k(r)=0.$$
(4.11)
Our task is to determine $`\psi ,\varphi ,\chi ,B_i,C_i,D_i`$, which are functions of $`r`$ and $`\theta `$.
Equating the $`\alpha ^2`$-terms in the Bogomolnyi equations (4.3), we obtain
$$(_1+ı_2)\psi +\frac{2a}{r}\psi e^{ı\theta }ıf(B_1+ıB_2)e^{2ı\theta }=4kf(k^{}+\frac{2k}{r})e^{ı\theta },$$
(4.12)
$$_1B_2_2B_1+\frac{1}{2}(f\psi e^{2ı\theta }+f\psi e^{2ı\theta })=2k^2f^2.$$
(4.13)
A Fourier expansion with the minimal number of nonzero terms leads to the ansatz
$$\psi =g(r)f(r)e^{2ı\theta }+\stackrel{~}{g}(r)e^{2ı\theta },$$
$$B_1+ıB_2=\stackrel{~}{b}(r)e^{ı\theta }+ıb(r)f(r)e^{3ı\theta },$$
(4.14)
and to equations for $`g(r),\stackrel{~}{g}(r),b(r)`$ and $`\stackrel{~}{b}(r).`$ The equations for $`\stackrel{~}{g}(r)`$ and $`\stackrel{~}{b}(r)`$ read
$$\stackrel{~}{g}=\frac{1+2a}{r}bb^{},\stackrel{~}{b}=ıh^{}.$$
(4.15)
The functions $`g(r)`$ and $`b(r)`$ must satisfy the equations
$$g^{\prime \prime }+\frac{1}{r}g^{}f^2g=2k^2f^2,$$
(4.16)
$$b^{\prime \prime }+\frac{1}{r}b^{}(\frac{1+f^2}{2}+\frac{1+4a+4a^2}{r^2})b=4kf(k^{}+\frac{2k}{r}).$$
(4.17)
Equating the $`\alpha \beta `$-terms and the $`\beta ^2`$-terms in the Bogonolnyi equation (4.3), we obtain equations for $`\varphi `$ and $`C_i`$, and for $`\chi `$ and $`D_i`$ respectively. These equations, which are very similar to equations (4.12) and (4.13), can again be solved by functions with the same $`\theta `$-dependence as in (4.14) but with slightly different radial functions. Collecting all results, we can write the solution, up to quadratic terms, in the form
$$\varphi =fe^{2ı\theta }+2(\alpha +ı\beta )kf$$
$$+(\alpha ^2+\beta ^2)gfe^{2ı\theta }+(\alpha +ı\beta )^2(\frac{1+2a}{r}bb^{})e^{2ı\theta }+\mathrm{}$$
$$A_1+ıA_2=ı\frac{2a}{r}e^{ı\theta }2ı(\alpha +ı\beta )(k^{}+\frac{2k}{r})e^{ı\theta }$$
$$ı(\alpha ^2+\beta ^2)g^{}e^{ı\theta }+ı(\alpha +ı\beta )^2bfe^{3ı\theta }+\mathrm{}$$
(4.18)
It remains to be shown that the quadratic terms in (4.18) are $`C^{\mathrm{}}`$ functions on $`𝐑^2`$ which do not change the action (and the winding number). To this end we use the asymptotic expansions of $`f,a`$ and $`k`$ at zero ,
$$f(r)=f_1r^2+\frac{1}{8}f_1r^4+\mathrm{},a(r)=\frac{1}{8}r^2\frac{1}{24}f_1^2r^6+\mathrm{},k(r)=r^2+k_1r^2+\mathrm{},$$
(4.19)
where $`f_1=.236`$ and $`k_1=.025`$ from the numerical analysis. We find that the solutions of (4.16) and (4.17) have the following expansions at the origin,
$$g(r)=g_1\mathrm{log}r+g_1+\frac{1}{2}f_1^2r^2+\mathrm{}$$
$$b(r)=b_1r^1+b_1r+(\frac{1}{8}b_12f_1k_1)r^3+\mathrm{}$$
(4.20)
The higher order terms in $`g(r)`$ are even powers of $`r`$, whereas the higher order term in $`b(r)`$ are odd powers of $`r`$. Hence, the quadratic terms in (4.18) are $`C^{\mathrm{}}`$ near the origin if and only if $`h_1=b_1=0.`$ So far the constants $`g_1`$ and $`b_1`$ are arbitrary.
For large $`r`$ the functions $`f,a,`$ and $`k`$ have the following asymptotic behavior :
$$f(r)=1+\stackrel{~}{f}_1(r)e^r+\mathrm{},$$
$$a(r)=1+\stackrel{~}{a}_1(r)e^r+\mathrm{},$$
(4.21)
$$k(r)=\stackrel{~}{k}_1(r)e^r+\mathrm{},$$
with coefficient functions which are polynomially bounded. This leads to the existence of exponentially decaying solutions which asymptotically are of the form
$$g(r)=\stackrel{~}{g}_1(r)e^r+\mathrm{},b(r)=\stackrel{~}{b}_1(r)e^r+\mathrm{}$$
(4.22)
Here $`\stackrel{~}{g}_1`$ and $`\stackrel{~}{b}_1`$ are polynomially bounded.
By numerical integration, the coefficients $`g_1`$ and $`b_1`$ which lead to an exponential fall-off at infinity, are found to be $`g_1=.144`$ and $`b_1=.026`$. The existence of such functions can be explained analytically as follows: Equation (4.16) shows that for positive $`g_1`$, g cannot have a maximum for any $`r`$. So the function diverges exponentially. For very small $`g_1`$, the term on the right-hand side of (4.16) will force the function to cross the $`r`$-axis, and then, as before, diverge exponentially. For very large negative $`g_1`$, the third term in (4.16) will force $`g`$ to go through a maximum for large $`r`$. After that, the function cannot have a minimum and must go to minus infinity. Because of the continuous dependence on the initial data, we have an open set of data for which $`g`$ crosses the $`r`$-axis, and an open set of data for which $`g`$ goes through a maximum below the $`r`$-axis. Therefore, we have at least one value of $`g_1`$ for which the function does neither. This function must converge and does so to zero, exponentially.
A similar argument explains the existence of an acceptable solution $`b(r)`$ to Eq. (4.17). The right-hand side of that equation is positive. So again $`b`$ cannot have a maximum above the $`r`$-axis. Also, for very small negative $`b_1`$, the right-hand side will force $`b`$ to go through a minimum and then cross the $`r`$-axis. For very large negative $`b_1`$, the third term in (4.17) prevents $`b`$ from going through a minimum. In between these two possibilities we find the desired solution which goes through a minimum but does not cross the $`r`$-axis. Such a solution must decay exponentially.
The cubic terms can be calculated in the same manner. We find, at third order,
$$\varphi =\mathrm{}+(\alpha +ı\beta )(\alpha ^2+\beta ^2)fh+(\alpha +ı\beta )^3(c^{}+\frac{3+2a}{r}c)e^{4ı\theta }+\mathrm{},$$
$$A_1+ıA_2=\mathrm{}$$
$$+ı(\alpha +ı\beta )(\alpha ^2+\beta ^2)[h^{}\frac{2}{r}h+2g(k^{}+\frac{2k}{r})+2kg^{}]e^{ı\theta }+ı(\alpha +ı\beta )^3fce^{5ı\theta }+\mathrm{}$$
(4.23)
The new radial functions, $`h(r)`$ and $`c(r)`$, satisfy the equations,
$$h^{\prime \prime }+\frac{1}{r}h^{}(f^2+\frac{4}{r^2})h=4k^{}g^{}+2fk(2fk^2+3fg+\frac{1+2a}{r}bb^{}),$$
(4.24)
$$c^{\prime \prime }+\frac{1}{r}c^{}(\frac{1+f^2}{2}+\frac{9+12a+4a^2}{r^2})c=2kf^2b2(k^{}+\frac{2k}{r})(\frac{1+2a}{r}bb^{}).$$
(4.25)
Near the origin, Eq. (4.25) has a series solution in powers of $`r^2`$ of the form
$$h(r)=f_1^2+h_1r^2+h_2r^4+\mathrm{}$$
(4.26)
The constant term is given in terms of the coefficient $`f_1`$ of the leading term in the expansion (4.19) of $`f(r)`$. The form of this term leads to the cancellation of the $`r^1`$-terms in the radial function multiplying $`e^{ı\theta }`$ in (4.24), and thus ensures that this term in (4.23) is $`C^{\mathrm{}}`$ on $`𝐑^2`$. The series in odd powers of $`r`$ for $`c(r)`$ which solves Eq. (4.25) near the origin, is
$$c(r)=c_1r^3+c_2r^5+\mathrm{}$$
(4.27)
The form of the series solutions at the origin guarantees that the cubic terms in (4.23) are $`C^{\mathrm{}}`$ functions on $`𝐑^2`$. For large $`r`$, Eqs (4.24) and (4.25) have exponentially decaying solutions.
## 5 Conclusions
Our expansions show a simple $`\theta `$-dependence in terms of trigonometric functions. In both models, the expansion of $`\varphi `$ exhibits the following pattern:
$$\begin{array}{ccccccc}& & & & & & e^{2ı\theta }\\ & & & & & e^{0ı\theta }& \\ & & & & e^{2ı\theta }& & e^{2ı\theta }\\ & & & e^{4ı\theta }& & e^{0ı\theta }& \\ & & e^{6ı\theta }& & e^{2ı\theta }& & e^{2ı\theta }\\ & e^{8ı\theta }& & e^{4ı\theta }& & e^{0ı\theta }\\ \mathrm{}& & \mathrm{}& & \mathrm{}& & \mathrm{}\end{array}$$
Here the first line gives the $`\theta `$ dependence of the zero order term; the second line gives the first order term, and so on. We get a similar triangular pattern for the $`\theta `$ dependence of $`A_1+ıA_2`$ at any order. For the radial functions we find differences between the two models. In the model for one complex field, the radial functions can be given explicitly in terms of exponential functions. However, for the angular dependence (3.15), a singularity occurs at the origin. (We have found no solution to (3.14) which is not of the form (3.15); we have no proof that there is none.)
For the Ginzburg-Landau theory on the other hand, the expansion is smooth, at least up to the order to which we carried out our calculations. In this model the radial functions are not given in terms of well-known functions. Having used the technique to calculate the terms up to third order, it is quite clear how to proceed to any order, and also how to proceed in the case of more than two vortices. We expect these expansions to converge for small separation parameters in the physical Ginzburg-Landau model. However, we do not have an estimate of the radius of convergence.
|
warning/0007/hep-th0007135.html
|
ar5iv
|
text
|
# Magnetic monopoles, alive. ITEP-TH-36/00
## 1 Generalities
### 1.1 Introduction
Magnetic monopoles is undoubtedly a fascinating subject. Not a new one, though. The Dirac magnetic monopole is 70 years old soon . And the first 50 years of development of the theory of the magnetic monopoles were summarized in an illuminating review by Coleman . Thus, the question may arise why it is instructive to come back to the monopoles now.
The main development since the Coleman’s review is that monopoles were copiously observed (for review and further references see, e.g., ) and the theory can be now confronted with the data. True, the monopoles observed are not exactly those introduced by Dirac, but rather their close akins, that is monopoles of non-Abelian gauge theories (moreover, for the sake of definiteness we concentrate on the simplest gauge group, that is $`SU(2)`$). Also true, the data are numerical mostly and obtained on the lattice so that their interpretation in terms of the continuum theory may not be so straightforward. Nevertheless, it is a direct challenge to theory to explain the ample data on the magnetic monopoles which have already been accumulated in the lattice simulations.
Moreover, the issue of the so to say lattice monopoles is very much rich and varied by itself. Let us mention here three topics:
(a) The numerical data refer mostly to the monopoles with a double magnetic charge, $`|Q_m|=2`$ where the units are fixed by the Dirac quantization condition for the gluons. Classically, there are no stable solutions with $`|Q_m|=2`$ and, therefore, quantum effects seem to be absolutely crucial even to introduce such monopoles. As a result, the theory of these monopoles is in its infancy.
(b) There are recent measurements of the interaction potential between the fundamental monopoles with $`|Q_m|=1`$ on the lattice , which are introduced through the so called ’t Hooft loop . Unlike the case of the $`|Q_m|=2`$ monopoles the interaction of the fundamental monopoles is in fact quite well understood. The fact, which might be not well appreciated by the community.
(c) There exists surprisingly simple phenomenological description of the properties of the $`|Q_m|=2`$ monopoles which are so poorly understood on the purely theoretical side. We mean here models like the Abelian Higgs model which provide quantitative support to the old idea of the dual-superconductor mechanism and work surprisingly well at least in some cases, for review and further references see .
In this mini-review we will emphasize some new points related to each of the items (a)-(c) listed above and which are based mostly on the original papers . The new points, although they refer to various topics, are unified by a common approach. The starting point is that we consider monopoles within the fundamental gluodynamics while the more traditional approach is to introduce monopoles within an effective theory intended to mimic QCD in the infrared region . Also, we understand the continuum gluodynamics rather as the limiting case of the lattice formulation. As a result, one allows for certain singular gauge transformations which are not included in more traditional frameworks.
### 1.2 Dirac monopole and Dirac string.
The Dirac monopole, by definition, is associated with a radial magnetic field similar to the electric field of a point-like charge, $`𝐇=(4\pi )^1Q_m(𝐫/r^3)`$. One can easily construct a corresponding vector potential:
$$A_r=A_\theta =0,A_\varphi =\frac{Q_m}{4\pi }\frac{(1+\mathrm{cos}\theta )}{r\mathrm{sin}\theta }.$$
(1)
The analogy between the electric and magnetic charges is somewhat formal, however. Namely, because of the conservation of the magnetic flux, the radial magnetic field of the monopole should be supplemented by the magnetic field of a string which brings in the flux spread out uniformly by the radial component of the field. Thus, we actually have
$$𝐇=𝐇_{rad}+𝐇_{string}.$$
(2)
The presence of the string is exhibited, in particular, by the explicit expression for the potential $`𝐀`$ above.
The Dirac string is unphysical and there is a number of constraints imposed on the theory to ensure that the string does not produce any physical effect. First, there is the Dirac veto which forbids any direct interaction with the string. The best known constraint is the Dirac quantization condition which ensures the absence of the Aharonov-Bohm effect for the electrons scattered on the string:
$$Q_e_{string}𝐀d𝐱=Q_eQ_m=2\pi k,$$
(3)
where $`Q_e`$ is the electric charge of the electron and $`k`$ is an integer number. Let us also emphasize that naively the energy of the string is infinite in the ultraviolet:
$$ϵ_{string}(𝐇_{string})^2\mathrm{d}^3r\frac{(Length)}{(Area)}\mathrm{\Lambda }_{UV}^2(Length),$$
(4)
where we used the fact that the magnetic flux is quantized (see above) and that the cross section of the string denoted by $`(Area)`$ should tend to zero at the end of the calculation. Thus, we substituted $`(Area)^1`$ by $`\mathrm{\Lambda }_{UV}^2`$.
The radial part of the magnetic field is also associated with an infinite energy:
$$ϵ_{rad}(𝐇_{rad})^2\mathrm{d}^3r\frac{1}{r_0}\mathrm{\Lambda }_{UV}.$$
(5)
Note that this ultraviolet divergence is linear, i.e. somewhat weaker than the divergence due to the string, see Eq. (4).
The infinite magnetic field of the string may have more subtle manifestations as well. Consider interaction of two magnetic monopoles with magnetic charge $`\pm Q_m`$ placed at distance $`R`$ from each other. Then, by the analogy with the the case of two electric charges, we would like to have the following expression for the interaction energy:
$$ϵ_{int}=𝐇_{1,rad}𝐇_{2,rad}\mathrm{d}^3r=\frac{Q_m^2}{4\pi }\frac{1}{R}.$$
(6)
Note, however, that if we substitute the sum of the radial and string fields for $`𝐇_{1,2}`$, then we would have an extra term in the interaction energy:
$$\stackrel{~}{ϵ}_{int}=\left(𝐇_{1,rad}𝐇_{2,string}+𝐇_{1,string}𝐇_{2,rad}\right)\mathrm{d}^3r=+2\frac{Q_m^2}{4\pi }\frac{1}{R}.$$
(7)
In other words, the account of the string field would flip the sign of the interaction energy! This contribution, although looks absolutely finite, is of course a manifestation of the singular nature of the string magnetic field, $`|𝐇_{string}|(Flux)/(Area)`$. Note that the integral in (7) does not depend on the shape of the string.
To maintain the unphysical nature of the Dirac string we should use a regularization scheme which would allow to get rid of these singularities.
### 1.3 Lattice regularization.
Since the monopoles are naively having divergent energy (or action) in the ultraviolet, the regularization is a crucial issue. Moreover, we would like to follow the lattice formulation since the monopoles are observed on the lattice.
Consider first the $`U(1)`$ case. As is emphasized in Ref. , the lattice formulation implies that Dirac string which produces no Aharonov-Bohm scattering costs no action as well. The reason is very simple. The lattice action is written originally in terms of the contour integrals like (3) rather than field strength $`F_{\mu \nu }`$:
$$S=\underset{p}{}\mathrm{Re}\mathrm{exp}\{iQ_e_pA_\mu 𝑑x^\mu \},$$
(8)
where the sum is taken over all the plaquettes $`p`$. Thus, the condition (3) means absence of both the Aharonov-Bohm effect and the quadratic divergence (4) in the lattice regularization. Moreover, it is straightforward to see that the interference term (7) also vanishes. Later, we will also discuss the case of the Dirac string which in the limit $`g0`$ correspond to negative plaquettes in the lattice formulation. Its energy is infinite in the continuum limit, in agreement with the naive estimate (4). The interference term (7), however, disappears in the lattice formulation in this case as well.
Moreover, the lattice formulation naturally leads to the monopole–antimonopole potential (6) without the unphysical string contribution (7).
The radial field, $`𝐇_{rad}`$ may also cause problems with infinite energy, see (5). The lattice regularization is not much specific in that case, however. The role of $`r_0`$ is simply played by the lattice spacing $`a`$. Thus, the probability to find a monopole on the lattice is suppressed by the action as:
$$e^S\mathrm{exp}(constQ_e^2L/a),$$
(9)
where $`L`$ is the length of the monopole trajectory, and the $`Q_e^2`$ factor appears because of the Dirac quantization condition (3) which relates the magnetic charge $`Q_m`$ to the inverse electric charge.
Although the Eq. (9), at first sight, rules out monopoles as physically significant excitations, the fate of the monopoles in the $`U(1)`$ case depends in fact on the value of the charge $`Q_e`$. The point is that the entropy factor grows also exponentially with the length of the monopole trajectory:
$$(Entropy)\mathrm{exp}(+const^{}L/a),$$
(10)
where the $`const^{}`$ is a pure geometric factor, not related to any coupling constant like $`Q_e`$. As a result for $`Q_e1`$ there is a phase transition corresponding to the condensation of the monopoles. This phase transition, which is well studied on the lattice, is the first and striking example of importance of the UV regularization in the non-perturbative sector. Indeed, once the UV divergence (4) is removed by the lattice regularization the monopoles can modify the physics completely (for further comments see ).
### 1.4 Classification of monopoles in non-Abelian theories.
From now on, we will discuss monopoles in unbroken non-Abelian gauge theories, having in mind primarily gluodynamics, i.e. quantum chromodynamics without dynamical quarks. Moreover, for the sake of simplicity we will consider only the $`SU(2)`$ gauge group.
A natural starting point to consider monopoles in non-Abelian theories is their classification. There are actually a few approaches to the monopole classification and it is important to realize both similarities and differences between them.
The dynamical, or $`U(1)`$ classification. Within this approach , one looks for monopole-like solutions of the classical Yang-Mills equations. Where by the “monopole-like” solutions one understands potentials which fall off as $`1/r`$ at large $`r`$, see Eq. (1). The basic finding is that there are no specific non-Abelian solutions and all the monopoles can be viewed as Abelian-like embedded into the $`SU(2)`$ group. Moreover, using the gauge invariance one can always choose the corresponding $`U(1)`$ group as, say, the rotation group around the third direction in the color space. According to this classification, the monopoles are characterized by their charge with respect to a $`U(1)`$ group and may have, therefore, charges,
$$|Q_m|=0,1,2,\mathrm{}.$$
(11)
The topological, or $`Z_2`$ classification. The $`Z_2`$ classification is based entirely on topological arguments. Namely, independent types of monopoles can be enumerated by considering the first homotopy group of the gauge group. The SU(2) gauge group is trivial since $`\pi _1(SU(2))=0`$, while in the case of the $`SO(3)`$, however,
$$\pi _1(SU(2)/Z_2)=Z_2$$
(12)
and there exists a single non-trivial topological monopole. We will denote the magnetic charge of such monopoles as $`|Q_m|=1`$. Note, however, that the charges $`Q_m=\pm 1`$ are indistinguishable in fact. As for the charges $`Q_m=2`$ they are equivalent, from this point of view, to no magnetic charge at all.
The topological classification (12) is readily understood if one tries to enumerate various types of the Dirac strings whose end points represent monopoles under consideration. Then there is only one non-trivial string, that is the one for which Eq. (3) is satisfied for gluons but not for quarks. Namely, because the $`U(1)`$ charge associated with gluons is twice as big as that of the particles in the fundamental representation (quarks) we may have
$$\mathrm{exp}\{igA_\mu 𝑑x^\mu \}=1$$
(13)
and such a string is not visible for the isospin one particles. On the other hand, the standard plaquette action is based on the phase factor evaluated for particles in the fundamental representation. Which means, in turn, that the Dirac string is piercing the negative plaquettes. This observation is the basis for introducing the $`|Q_m|=1`$ monopoles via the ’t Hooft loop: one changes the sign of $`\beta `$ ($`\beta 4/g^2`$) on a world sheet. The boundary of this sheet corresponds to the end points of the Dirac string, or the monopole trajectory.
### 1.5 $`Z_2`$ monopoles.
In principle, the $`U(1)`$ and $`Z_2`$ classifications are different. Indeed, while the $`U(1)`$ classification allows for any integer charge, the $`Z_2`$ classification leaves space only for a single non-trivial charge:
$$Q_m=0,1.$$
(14)
The reconciliation of the two classifications is that the $`U(1)`$ solutions with $`|Q_m|2`$ are in fact unstable because of the presence massless charged vector particles (gluons) . The instability of the solutions implies that even if the external sources with $`|Q_m|2`$ were introduced into the vacuum state of the gluodynamics, charged gluons would fall onto the center because of the strong magnetic interactions. Moreover, one can imagine that as result of this instability the charges fields $`A^\pm `$ are build up as well.
In a somewhat related way, one can demonstrate the apparent irrelevance of the $`|Q_m|=2`$ monopoles by producing an explicit non-Abelian field configuration which looks as a $`|Q_m|=2`$ monopole in its Abelian part but has no $`SU(2)`$ action at all . This field configuration is a Dirac string with open ends, which correspond to the monopole-anti-monopole pair separated by the distance $`R`$. In more detail, such a configuration is generated from the vacuum by the following gauge rotation matrix:
$$\mathrm{\Omega }=\left(\begin{array}{cc}e^{i\phi }\sqrt{A_D}& \sqrt{1A_D}\\ \sqrt{1A_D}& e^{i\phi }\sqrt{A_D}\end{array}\right),$$
(15)
where $`\phi `$ is the angle of rotation around the axis connecting the monopoles and $`A_D`$ is the $`U(1)`$ potential representing pure Abelian monopole pair:
$$A_\mu dx_\mu =\frac{1}{2}\left(\frac{z_+}{r_+}\frac{z_{}}{r_{}}\right)d\phi A_D(z,\rho )d\phi ,$$
(16)
where $`z_\pm =z\pm R/2`$, $`\rho ^2=x^2+y^2`$, $`r_\pm ^2=z_\pm ^2+\rho ^2`$. Note that the action associated with the Dirac string is considered in this case zero, in accordance with the lattice version of the theory (for details see ).
In this example, the monopoles with $`|Q_m|=2`$ are a kind of a pure gauge field configurations carrying no action. Note that the Abelian flux is still transported along the Dirac string and is still conserved for the radial field. What is lost, however, is the relation between the Abelian flux and action. In the Abelian case non-vanishing flux means non-vanishing magnetic field and non-vanishing action since the action density is simply $`𝐇^2`$. Now the action is $`(F_{\mu \nu }^a)^2`$ and the Abelian part of the $`F_{\mu \nu }^3`$ can be canceled by the commutator term. This is exactly what happens in the example (15) above.
It is somewhat more difficult to visualize dynamically the equivalence of the $`Q_m=\pm 1`$ monopoles, also implied by the $`Z_2`$ classification. The mechanism mixing the $`Q_m=\pm 1`$ solutions seems to be the following. Imagine that we start with, say, $`Q_m=+1`$ solution. Then a Dirac string carrying the flux corresponding to the $`Q_m=2`$ can be superimposed on this solution. It is important at this point that such a Dirac string costs no action (or energy). Then the radial magnetic field can also change its direction since it does not contradict the flux conservation any longer. In a related language, one could say that the $`|Q_m|=2`$ monopoles are condensed in the vacuum and that is why the magnetic charge can be changed freely by two units.
As far as interaction of two $`|Q_m|=1`$ monopoles is concerned, one might expect that they would behave themselves as an monopole-antimonopole pair. Indeed monopole and antimonopole would attract each other and thus represent the lowest energy state of the system.
### 1.6 Conclusions # 1
Thus, the physics of the monopoles in the first approximation turns very simple.
Namely, there exist only monopoles with $`|Q_m|=12\pi /g`$ where $`g`$ is the coupling constant of the non-Abelian $`SU(2)`$ theory. The monopoles are infinitely heavy and can be introduced only as external object through the ’t Hooft loop. Their interaction is Abelian like:
$$V_{m\overline{m}}=\frac{Q_m^2}{4\pi R}=\frac{\pi }{g^2R},$$
(17)
where $`R`$ is the separation between the monopoles.
Clearly enough, this first, or classical approximation falls far beyond an adequate description of the empirical data on the monopoles, see the Introduction. Thus, we are invited to go into more advanced approaches which we would try to introduce step by step.
## 2 Lagrangian approach.
### 2.1 The Zwanziger Lagrangian.
There is a long standing interest in constructing the dual gluodynamics, for review and further references see . The dual gluon, by definition, interacts with monopoles. The motivation is to realize in the field theoretical language the dual superconductor model of the quark confinement according to which the quarks are connected at large distances by an Abrikosov-type vortex . The key element is the construction of the non-Abelian monopoles, which are usually modeled after the ’t Hooft–Polyakov solution. Namely, one introduces first non-Abelian dual gluons interacting with Higgs fields and then assumes condensation of the Higgs fields which mimics the condensation of the monopoles. In the realistic case of the $`SU(3)`$ gauge group one needs an octet of dual gluons and three octets of the Higgs fields, all of them understood in terms of effective field theory valid in the infrared region.
While such a construction might be viable as an effective theory, we need in fact tools to describe interaction of non-Abelian monopoles at arbitrary short distances as well . Indeed, in the lattice version of the theory external monopoles can be introduced via the ’t Hooft loop operator and in the continuum limit these monopoles are point like. Thus, we are encouraged to consider the dual gluodynamics at short distances, or at the fundamental level.
It is natural to try a Lagrangian approach to the dual gluodynamics. Indeed, in case of the same ’t Hooft loop operator it is known that its expectation value depends only on the boundary and not on the shape of the Dirac string. Thus, it seems natural to introduce a dual gluon which would interact directly with point-like monopoles. In the context of electrodynamics, the idea is of course very old and goes back to papers in Ref. . There are successes and problems inherent to this approach, for a review see .
A well known example of Lagrangian which describes interaction of a U(1) gauge fields with Abelian point-like monopoles is due to Zwanziger :
$$L_{Zw}(A,B)=\frac{1}{2}(m[A])^2+\frac{1}{2}(m[B])^2+$$
(18)
$$+\frac{i}{2}(m[A])(m{}_{}{}^{}[B])\frac{i}{2}(m[B])(m{}_{}{}^{}[A])+ij_eA+ij_mB,$$
where $`j_e,j_m`$ are electric and magnetic currents, respectively, $`m_\mu `$ is a constant vector, $`m^2=1`$ and
$$[AB]_{\mu \nu }=A_\mu B_\nu A_\nu B_\mu ,(m[AB])_\mu =m_\nu [AB]_{\mu \nu },$$
$${}_{}{}^{}[AB]_{\mu \nu }^{}=\frac{1}{2}\epsilon _{\mu \nu \lambda \rho }[AB]_{\lambda \rho }.$$
At first sight, we have introduced two different vector fields, $`A,B`$ to describe interaction with electric and magnetic charges, respectively. If it were so, however, we would have solved a wrong problem because we need to have a single photon interacting both with electric and magnetic charges. And this is what is achieved by the construct (18). Indeed, the action (18) is not diagonal in the $`A`$, $`B`$ fields and one can convince oneself that the form of the bilinear in $`A,B`$ interference terms in (18) is such that the field strength tensors constructed on the potentials $`A`$ and $`B`$ are in fact related to each other:
$$F_{\mu \nu }(A)={}_{}{}^{}F_{\mu \nu }^{}(B).$$
(19)
Which means in turn that there are only two physical degrees of freedom corresponding to the transverse photons which can be described either in terms of the potential $`A`$ or $`B`$. Topological excitations, however, can be different in terms of $`A`$ and $`B`$.
The physical content of (18) is revealed by the propagators for the fields $`A,B`$. In the $`\alpha `$-gauge one can derive:
$$\begin{array}{c}A_\mu A_\nu =B_\mu B_\nu =\frac{1}{k^2}(\delta _{\mu \nu }(1\alpha )\frac{k_\mu k_\nu }{k^2}),\\ \\ A_\mu B_\nu =B_\mu A_\nu =\frac{i}{k^2(km)}{}_{}{}^{}[mk]_{\mu \nu }^{}.\end{array}$$
(20)
The propagators should reproduce, as usual the classical solutions. And indeed, the $`AA`$, $`BB`$ propagators describe the Coulomb-like interaction of two charges and magnetic monopoles, respectively. While the $`AB`$ propagator reproduces interaction of the magnetic field of a monopole with a moving electric charge.The appearance of the poles in $`(km)`$ is a manifestation of the Dirac strings.
To summarize, the Zwanziger Lagrangian in electrodynamics reproduces the classical interaction of monopoles and charges. Upon the quantization, it describes the correct number of the degrees of freedom associated with the photon.
### 2.2 Dual gluon as an Abelian vector field.
Now, if we would approach the problem of constructing a Zwanziger-type Lagrangian for the dual gluodynamics, we immediately come to a paradoxical conclusion that the dual field, if any, is Abelian. Indeed, monopoles associated with, say, $`SU(N)`$ gauge group are classified according to $`U(1)^{N1}`$ subgroups and might be realized as a pure Abelian objects. Thus, there is no place for a non-Abelian dual gluon because the monopoles do not constitute representations of the non-Abelian group.
The function of the classical Lagrangian is, first of all, to reproduce the classical interactions of the monopoles and charges. It is rather obvious that the potential (17) can be derived in the classical approximation from the Lagrangian:
$$L_{dual}(A^a,B)=\frac{1}{4}(F_{\mu \nu }^a)^2+\frac{1}{2}(m[Bi{}_{}{}^{}G])^2+ij_mB+ij_e^aA^a,$$
(21)
where $`a=1,2,3`$ is the color index, $`j_m`$ is the magnetic current and $`F_{\mu \nu }^a`$ is the non-Abelian field strength tensor. The Lagrangian (21) also contains vector field $`n^a`$, $`n^2=1`$ in the adjoint representation and antisymmetric tensor $`G_{\mu \nu }`$ is the ’t Hooft tensor :
$$G_{\mu \nu }=n^aF_{\mu \nu }^a\epsilon ^{abc}n^a(D_\mu n)^b(D_\nu n)^c.$$
(22)
Let us add a few comments on the meaning and rules of using the Lagrangian (21).
(a) First, if the magnetic current is vanishing, $`j_m=0`$ then the integration over the field $`B`$ reproduces the standard Lagrangian of the gluodynamics.
(b) As far as the quantization is concerned, the Lagrangian (21) reproduces the correct degrees of freedom of the free gluons. Indeed, in the limit $`g0`$ and for $`n^a=\delta ^{a,3}`$ the Lagrangian (21) becomes:
$$L_{dual}(A^a,B)=\frac{1}{4}(A^1)^2+\frac{1}{4}(A^2)^2+ij_mB+ij_e^aA^a+$$
$$+\frac{1}{2}[m(A^3)]^2+\frac{1}{2}[m(B)]^2+$$
(23)
$$+\frac{i}{2}[m(A^3)][m{}_{}{}^{}(B)]\frac{i}{2}[m{}_{}{}^{}(A^3)][m(B)],$$
which is essentially the Zwanziger Lagrangian (18). Quantization at this point is the same as in the case of a single photon.
(c) Already in the Zwanziger example (18) we have seen that the fields that are mixed up in the Lagrangian have a common source. Namely, in case of the electrodynamics $`{}_{}{}^{}F(A)=F(B)=j_m`$. Since it is known that the monopoles in non-Abelian theories serve as source for the ’t Hooft tensor (22) one expect from the very beginning that in case of the gluodynamics the (dual) field strength tensor build up on the dual gluon field $`B`$ is mixed up with the ’t Hooft tensor constructed in terms of the gluon filed $`A`$. And, indeed, this is true for (21).
(d) The emergence of the vector $`n^a`$ is of crucial importance in the Lagrangian (21). The point is that the origin of the vector $`n^a`$ goes back to choosing the color orientation of the monopoles. As is emphasized above the monopole solutions are Abelian in nature which means, in particular, that they can be rotated to any direction in the color space by gauge transformations. Thus, picking up a particular $`n^a`$ is nothing else but using the gauge fixing freedom. Therefore, we can either average over the directions of $`n^a`$ or fix $`n^a`$ but evaluate only gauge invariant quantities, like the Wilson loop (note somewhat similar remarks in Ref. ).
(e) The $`Z_2`$ nature of the monopoles is manifested in the freedom of changing $`n^an^a`$, $`B_\mu B_\mu `$. Indeed, under such transformation the monopole with the charge $`Q_m=+1`$ is transformed into a monopole with $`Q_m=1`$ and vice versa. In the language we used above such a transformation corresponds to adding a Dirac string with a double magnetic flux. We see that the averaging over $`\pm n^a`$ is a part of the overall averaging over all possible embedding of the $`U(1)`$ into the $`SU(2)`$ gauge group.
An apparent application of (21) would be evaluating the running of the coupling $`g`$ in the expression (17). And, indeed, exploiting the Lagrangian (21) one can approach the problem of the running of the coupling in a way similar to the case of pure electrodynamics, for a review and further references see . We comment on this approach below.
### 2.3 Radiative corrections.
We will consider now the radiative corrections to the Coulomb-like interaction (17) at short distances. Obviously enough, one would expect that the radiative corrections result in the standard, non-Abelian running of the coupling $`g^2`$. Which is indeed our main conclusion. Moreover, since for a constant vector $`n^a`$ the non-Abelian monopole essentially coincides with the Dirac monopole, there is no much specific about the derivation of the running of the coupling. And, indeed, our considerations overlap to a great extent with those given in the original papers and in the reviews . Still,we feel that it is useful to present the arguments, may be in a new sequence, to emphasize the points crucial for our purposes.
Let us emphasize from the very beginning that the evaluation of the radiative corrections addresses in fact two different, although closely related problems. That is, running of the coupling and stability of the classical solutions. Both aspects are unified, of course, into evaluation of a single loop in the classical background. However, the running of the coupling can be clarified by keeping track of the ultraviolet logs, $`\mathrm{ln}\mathrm{\Lambda }_{UV}`$ alone and are universal since in the ultraviolet all the external fields can be neglected. Therefore, the coefficient in front of $`\mathrm{ln}\mathrm{\Lambda }_{UV}`$ can be found by evaluating the loop graph with two external legs, i.e. the graph corresponding to the standard polarization operator in perturbation theory. This is true despite of the fact that the monopole field is strong (i.e. the product of the magnetic and electric coupling is of order unity). On the other hand the stability of the classical solution is decided by the physics in the infrared. Here one needs to consider the particular dynamical system, monopoles in our case, and the fact that the magnetic charge is of order $`1/g`$ can be crucial.
Consider first the running of the coupling. Moreover, for the sake of definiteness we concentrate on the Dirac monopole with the minimal magnetic charge interacting with electrons and in one-loop approximation . The crucial point here is that only loops with insertion of two external (i.e., monopole) fields can be considered despite of the fact that there is no perturbative expansion at all. Indeed, considering more insertions make the graphs infrared sensitive, with no possibility for $`\mathrm{ln}\mathrm{\Lambda }_{UV}`$ to emerge.
Then, the evaluation of, say, first radiative correction to the propagator $`B_\mu B_\nu `$ in the Zwanziger formalism (20) seems very straightforward and reduces to taking a product of two $`AB`$ propagators and inserting in between the standard polarization operator of two electromagnetic currents. The result is :
$$B_\mu B_\nu (k)=\frac{\delta _{\mu \nu }}{k^2}(1L)+\frac{1}{(km)^2}(\delta _{\mu \nu }m_\mu m_\nu )L,$$
(24)
$$L=\frac{\alpha _{el}}{6}\mathrm{ln}\mathrm{\Lambda }_{UV}^2/k^2$$
and we neglect the electron masses so that the infrared cut-off is provided, in the logarithmic approximation, by the momentum $`k`$.
At first sight, there is nothing disturbing about the result (24). Indeed, we have a renormalization of the original propagator which is to be absorbed into the running coupling, and a new structure with the factor $`(km)^2`$ which is non-vanishing, however, only on the Dirac string. The latter term would correspond renormalization of the Dirac-string self-energy which we do not follow in any case since it is included into self-energy of the external monopoles. What is, actually, disturbing is that according (24) the magnetic coupling would run exactly the same as the electric charge,
$$A_\mu A_\nu (k)=(1L)\frac{\delta _{\mu \nu }}{k^2},$$
violating the Dirac quantization condition.
The origin of the trouble is not difficult to figure out. Indeed, using the propagator $`AB`$ while evaluating the radiative corrections is equivalent, of course, to using the full potential corresponding to the Dirac monopole $`A_D^{cl}`$. Then, switching on the interaction with electrons would bring terms like $`A_D^{cl}\overline{\psi }\gamma \psi `$. Since $`A_D`$ includes the potential of the string electrons do interact with the Dirac string and we are violating the Dirac “veto” which forbids any direct interaction with the string.
Let us demonstrate that, indeed, the incorrect treatment of the Dirac string changes the sign of the radiative correction. This can be done in fact in an amusingly simple way. First, let us note that it is much simpler to remove the string if one works in terms of the field strength tensor, not the potential. Indeed, we have $`𝐇=𝐇_{rad}+𝐇_{string}`$ while in terms of the potential $`𝐀`$ any separation of the string would be ambiguous (see Eq. (1)).
Thus, we start with relating the potential, or energy to the interference term in the $`𝐇^2`$ field:
$$V_{m\overline{m}}=𝐇_1𝐇_2\mathrm{d}^3r.$$
(25)
Now, it is not absolutely trivial, how we should understand the product $`𝐇_1𝐇_2`$. Indeed, we emphasized in section 1.2 that the string field is to be removed from this interference term, see Eq (7). Thus, in the zero, or classical approximation we have:
$$𝐇_1𝐇_2𝐇_{1,rad}𝐇_{2,rad}.$$
(26)
However, if we use the standard technique of an external field:
$$A_\mu =A_\mu ^{class}+a_\mu $$
(27)
and substitute (1) as the classical background then the first radiative correction would bring the product of the total $`𝐇_1𝐇_2`$ which includes also the string contribution<sup>1</sup><sup>1</sup>1At this point we assume in fact that $`\mathrm{\Lambda }_{UV}`$ is smaller than the inverse size of the string, which is convenient for our purposes here. Other limiting procedures could be considered as well, however.. Indeed, the result in the log approximation would be as follows:
$$\delta (𝐇_1𝐇_2)=L(𝐇_{1,string}+𝐇_{1,rad})(𝐇_{2,string}+𝐇_{2,rad})=L𝐇_{1,rad}𝐇_{2,rad},$$
(28)
where at the last step we have used the observation (7).
Now, it is clear how we could ameliorate the situation. Namely, to keep the Dirac string unphysical we should remove the string field from the expression (28) which arises automatically if we use the propagators (20) following from the Zwanziger Lagrangian. Thus, we introduce:
$$(𝐇_1𝐇_2)^{^{}}𝐇_{1,rad}𝐇_{2,rad}$$
(29)
and change $`𝐇_1𝐇_2`$ in the expression (28) into (29) so to say by hand. The justification is that we should remove the effect of the string field from any observable.
Then we reverse the sign of the radiative correction and the final result is
$$(V_{m\overline{m}})_{class}\frac{\pi }{g_0^2}\frac{1}{R}\frac{\pi }{g^2(R)}\frac{1}{R}.$$
(30)
One might wonder, how it happens that the couplings in the electric and magnetic potential run in opposite ways. Indeed, now we reduced the product $`𝐇_1𝐇_2`$ to exactly the same form as the product $`𝐄_1𝐄_2`$ in case of two electric charges (since the radial magnetic and electric fields are the same, up to a change of the overall constants). The resolution of the paradox is that the renormalization of the electric and magnetic fields are indeed similar in the language of the Lagrangian. However, the small corrections to the Lagrangian and Hamiltonian are related as:
$$\delta L=\delta H.$$
(31)
Since $`𝐄^2`$ and $`𝐇^2`$ enter with the same sign into the expression for the Hamiltonian and with the opposite signs into the Lagrangian, Eq (31) implies that the running of the couplings in the electric $`V_e`$ and magnetic $`V_m`$ potentials are opposite in sign. Which is, of course, in full agreement with expectations since $`V_eg^2`$ and $`V_mg^2`$.
Thus, it is not difficult to derive the running of the magnetic coupling following only the ultraviolet log, $`\mathrm{ln}\mathrm{\Lambda }_{UV}`$. Note, however, that the same arguments would go through without change if we started with, say, monopoles with $`Q_m=2`$. But such monopoles are unstable and this is a much more drastic effect than the would-be running of the coupling. There are also more subtle mechanisms which can be brought in by radiative corrections. In case of the same Dirac monopole interacting with electrons consideration of the modes reveals that the Hamiltonian is in fact non Hermitian. As a result the classical field approximation is not adequate and one should consider the corresponding field theory, or the monopole catalysis .
Thus, to investigate the stability of the classical soluition one has, generally speaking, to consider all orders in $`g_eg_m1`$. It is known that single monopoles with $`Q_m=1`$ are stable. The stability of the monopole-antimonopole system, which we are interested in, has never been investigated analytically in detail because of the complexity of the problem. However, there is no known mechanism which could cause instability of the classical monopole-antimonopole solution. Moreover, we checked numerically that the classical solution is indeed stable .
### 2.4 Why the “right way” is correct?
Thus, our exercise with evaluating the running of the magnetic coupling has brought mixed results. On one hand, we were able to derive that the product of electric and magnetic coupling constants is not renormalized, as one would expect. On the other hand, to derive this we had to go actually beyond the Lagrangian approach and remove the effect of the magnetic field of the string. Now we ask the next question, why this removing was the correct procedure.
Let us reexamine the grounds for the Lagrangian approach, in their generality. Any monopole involves also a Dirac string and, as a result, a world sheet, not just particle trajectories. If we stop here, then the conclusion would be that there is no Lagrangian approach to the problem. However, we are aware that the ’t Hooft loop operator depends only on its boundary, which is the monopoles trajectory, $`j_m`$. And this is the real basis for the hopes for the Lagrangian formulation. Now, we see that the Dirac veto is not respected by the Lagrangian formulation and, therefore, the possibility arise that the world sheet swept by the Dirac string is still somehow important. Thus, we will outline in this subsection an approach which is based on derivation of a continuum analog of the lattice ’t Hooft loop operator and avoids any direct use of Lagrangians.
The general one-plaquette action of $`SU(2)`$ lattice gauge theory (LGT) can be represented as:
$$S_{lat}(U)=\frac{4}{g^2}\underset{p}{}S_p\left(1\frac{1}{2}TrU[p]\right),$$
(32)
where $`g`$ is the bare coupling, $`p`$ is the boundary of an elementary plaquette $`p`$, the sum is taken over all $`p`$, $`U[p]`$ is the ordered product of link variables $`U_l`$ along $`p`$. In particular, if $`S_P(x)=x`$ then (32) is the standard Wilson action. The exponent of the lattice field strength tensor $`F_p`$ is defined in terms of $`U[p]`$:
$$U[p]=e^{i\widehat{F}_p}=\mathrm{cos}\left[\frac{1}{2}|F_p|\right]+i\tau ^an^a\mathrm{sin}\left[\frac{1}{2}|F_p|\right],$$
(33)
where $`\widehat{F}=F^a\tau ^a/2,|F|=\sqrt{F^aF^a}`$ and we define $`n_p^a=F_p^a/|F_p|`$ for $`|F_p|0`$, while $`n_p^a`$ is an arbitrary unit vector for $`|F_p|=0`$.
The lattice action (32) depends only on $`\mathrm{cos}\left[\frac{1}{2}|F_p|\right]`$. Therefore the action of the $`SU(2)`$ LGT possesses not only the usual gauge symmetry, but allows also for the gauge transformations which shift the field strength tensor by $`4\pi k`$, $`|F_p||F_p|+4\pi k`$, $`kZ`$:
$$e^{i\widehat{F}_p}=\mathrm{exp}\{i|F_p|\widehat{n}_p\}=\mathrm{exp}\{i(|F_p|+4\pi )\widehat{n}_p\}=\mathrm{exp}\{i(F_p^a+4\pi n_p^a)\tau ^a/2\}.$$
(34)
Thus, the symmetry inherent to the lattice formulation can be represented as:
$$F_p^aF_p^a+4\pi n_p^a,\stackrel{}{F}_p\times \stackrel{}{n}_p=0,n_p^2=1.$$
(35)
The symmetry (35) is absent in the conventional continuum limit, $`(F_{\mu \nu }^a)^2d^4x`$. Note that in the continuum limit $`n_p^a`$ becomes a singular two-dimensional structure $`{}_{}{}^{}\mathrm{\Sigma }_{\mu \nu }^{a}`$ which is representing the Dirac string world sheet.
So far we discussed an invisible Dirac string, which is nothing else but a generalized (or singular) gauge transformation. The Dirac string corresponding to the fundamental monopole corresponds to the phase factor $`1`$ and we can obtain, therefore, an expression for a continuum analog of the ’t Hooft loop by substituting:
$$F_{\mu \nu }^aF_{\mu \nu }^a+2\pi {}_{}{}^{}\mathrm{\Sigma }_{\mu \nu }^{a}.$$
(36)
In this way we come to the following definition of the ’t Hooft loop operator in the continuum:
$$H(\mathrm{\Sigma }_𝒞)=\mathrm{exp}\left\{\frac{1}{4g^2}d^4x\left[\left(F_{\mu \nu }^a\right)^2\left(F_{\mu \nu }^a+2\pi {}_{}{}^{}\mathrm{\Sigma }_{𝒞\mu \nu }^{a}\right)^2\right]\right\},$$
(37)
$$\mathrm{\Sigma }_{𝒞\mu \nu }^a=d^2\sigma _{\mu \nu }n^a(\sigma )\delta ^{(4)}(x\stackrel{~}{x}(\sigma )),$$
(38)
where the surface $`\mathrm{\Sigma }_𝒞^a`$ spanned on the contour $`𝒞`$ is assumed to be non-intersecting. The unit three-dimensional vector field $`n^a(\sigma )`$, $`\stackrel{}{n}^2=1`$ is defined on the world-sheet:
$$n^a(\sigma )=(t{}_{}{}^{}F_{}^{a})\left[(t{}_{}{}^{}F_{}^{b})^2\right]^{1/2},(tF^a)=t_{\mu \nu }(\sigma )F_{\mu \nu }^a(\stackrel{~}{x}),$$
(39)
$$t_{\mu \nu }(\sigma )=\frac{1}{\sqrt{g}}\epsilon ^{\alpha \beta }_\alpha \stackrel{~}{x}_\mu _\beta \stackrel{~}{x}_\nu ,t_{\mu \nu }^2=2,g(\sigma )=\mathrm{Det}[_\alpha \stackrel{~}{x}_\mu _\beta \stackrel{~}{x}_\mu ].$$
(40)
Therefore $`n^a(\sigma )`$ is not an independent variable, it is completely determined by the components of the field strength tensor $`F_{\mu \nu }^a`$. On the set of points where $`(t{}_{}{}^{}F_{}^{a})=0`$ the direction of $`n^a(\sigma )`$ is arbitrary. It can be shown that the Eq. (37)-(40) define the correct ’t Hooft loop operator the expectation value of which depends only on the contour $`𝒞`$, not on the particular position of the surface $`\mathrm{\Sigma }_𝒞`$.
Consider now the equations of motion in presence of the ’t Hooft loop operator:
$$D_\nu \left(F_{\mu \nu }(A)+2\pi {}_{}{}^{}\mathrm{\Sigma }_{\mu \nu }^{}\right)=0,$$
(41)
which should be supplemented by the Bianchi identities:
$$D_\nu {}_{}{}^{}F_{\mu \nu }^{}=0.$$
(42)
To appreciate the meaning of the equation of motion (41) let us choose the gauge such that $`\mathrm{\Sigma }_{\mu \nu }^a`$ has a constant color orientation characterized by the vector $`n_0^a`$. A particular solution of (41) may be found within the anzatz $`A_\mu ^a=n_0^aA_\mu `$, for which Eq. (41) reduces to:
$$_\nu \left(_{[\mu }A_{\nu ]}\right)=2\pi _\nu {}_{}{}^{}\mathrm{\Sigma }_{\mu \nu }^{}.$$
(43)
The solution of this equation in the Landau gauge,
$$A_\mu ^a=n_0^a2\pi \frac{1}{\mathrm{\Delta }}_\nu {}_{}{}^{}\mathrm{\Sigma }_{\mu \nu }^{},$$
(44)
corresponds to the gauge potential of an Abelian monopoles current $`\mathrm{\Sigma }`$ embedded into the $`SU(2)`$ group. Thus, $`\mathrm{\Sigma }_{\mu \nu }`$ is the Dirac world sheet.
Derivation of the classical equations of motion (43) is the first step in deriving the interaction of the fundamental monopoles outside any Lagrangian framework.
One could consider along these lines also the radiative corrections . We will not go into details here but let us mention, how it comes about that the “Dirac veto” is observed and virtual particles do not interact with the Dirac string. We will substantiate this point here on the example of the spin interaction. Since the Yang-Mills quanta possess spin, there exists interaction which is generalization of the non-relativistic expression $`\sigma 𝐇`$. In particular, if there exists classical field directed in third direction in the color space, $`(F_{\mu \nu }^3)_{cl}`$ then its interaction with the quantum charged fields $`a_\mu ^\pm `$ contains the term
$$g(F_{\mu \nu }^3)_{cl}a_\mu ^+a_\nu ^{}.$$
(45)
In the Zwanziger formalism, the $`(F_{\mu \nu }^3)_{cl}`$ means the whole magnetic field, the field of the string including. Then the interaction (45) brings the term $`𝐇_{1,string}𝐇_{2,radial}`$ on the level of the quantum corrections. As we emphasized in sect. 1.2, this term actually proportional to $`𝐇_{1,radial}𝐇_{2,radial}`$ which is responsible for the coupling running. In this way the term $`𝐇_{1,string}𝐇_{2,radial}`$, if it arises, brings in a “wrong” sign of the radiative correction.
On the other hand, in our formulation of the continuum analog of the ’t Hooft loop operator, see Eq. (37), there is no spin interaction of the virtual particles with the string magnetic field. This is crucial to prove that the coupling governing the monopole-antimonopole interaction indeed runs as $`g^2`$.
### 2.5 Conclusions # 2
We considered in fact two different points. First we argued that the dual gluon is a $`U(1)`$ gauge boson. The $`SU(2)`$ invariance is to be maintained either by integrating over all the possible embedding of the (dual) $`U(1)`$ into $`SU(2)`$ or by constraining calculations to gauge invariant quantities, like the Wilson loops.
Second, we discussed how far one can go with a Lagrangian formulation of the dual gluodynamics a la Zwanziger. To test the Lagrangian approach we evaluated the running of the coupling in the monopole-antimonopole potential. The conclusion is that one can get the correct running of the coupling by imposing the Dirac veto which forbids the interaction of virtual particles with the Dirac string. This requirement is not inherent to the Lagrangian approach (the same is true for the Zwanziger Lagrangian in the $`U(1)`$ case), however. It can be derived by studying the continuum analog of the ’t Hooft loop operator.
## 3 Monopoles with $`Q_m`$=2.
### 3.1 $`Q_m=2`$ monopoles as quantum objects.
So far we discussed the fundamental monopoles $`|Q_m|=1`$ which can be visualized as classical infinitely heavy objects. Because of infinite mass, they can be used only as probes of the QCD vacuum but play no dynamical role by themselves. The monopoles with the double charge $`|Q_m|=2`$ are very different. As discussed above, they do not exist on the classical level. On the other hand, there exist very simple arguments that they can play dynamical role on the quantum level. As far as the fundamental Lagrangian is concerned, the only role of the quantum corrections is the running of the non-Abelian coupling. In particular, if we consider a lattice coarse enough then $`g_{SU(2)}`$ becomes of order unity. Obviously, the same coupling governs the physics associated with any $`U(1)`$ subgroup of the $`SU(2)`$. However, if the coupling $`g_{U(1)}`$ becomes of order unity, then there is a phase transition associated with the monopole condensation . Thus, one can argue that the running will be stopped by the monopole condensation, if not by something else already at smaller values of $`g_{SU(2)}`$.
Thus, it is very natural to assume that the monopole condensation occurs also in QCD since the running of the coupling allows to scan the physics at all the values of $`g_{SU(2)}`$ until one runs into a phase transition.
However, even if one accepts such speculations, there remains a very important unresolved question. Namely, it is not clear which $`U(1)`$ subgroup of the full non-Abelian group is to be selected as the classification group for the monopoles. The most common approach here is to rely on the empirical data. In a way, it is forced on us since the phase transition is expected to happen at $`g_{SU(2)}^21`$ where analytical approaches are hardly possible. From the lattice simulations it is knwon that the monopoles in the Maximal Abelian projection appear to be most relevant, see for review and further references.
Instead of reviewing this material once more – which would take us far beyond the scope of the present article – we will highlight some features of new kind of monopoles introduced in Ref. . The basic idea behind this construction is to make monopoles as much geometrical objects as possible.
### 3.2 “Geometrical” monopoles.
The construction of the new kind monopoles is in few steps which we will briefly outline now.
(i) The usual starting point to introduce monopoles is to fix some $`U(1)`$ for the whole lattice and then look for the Dirac strings and monopoles with respect to this $`U(1)`$. The starting point of is somewhat different. Namely, it is the observation that each Wilson loop defines in a natural way its own $`U(1)`$. Indeed, turn back to the expression (33) for the plaquette action which is actually true for any Wilson loop. Then, it is clear that each Wilson loop defines the vector $`\widehat{F}_p`$ and the “natural” $`U(1)`$ is the group of rotation around this vector (in the color space). In this way, one can define a $`U(1)`$ group for each plaquette. The definition of the $`U(1)`$ subgroup varies from one plaquette to another, emphasizing the non-Abelian nature of the underlying theory.
(ii) The plaquette action is $`1/2\mathrm{cos}\varphi `$ and is invariant under $`\varphi _p\varphi _p+2\pi k`$, as is emphasized in sect. 1.3. Now to detect the Dirac strings we should be able to somehow define the integer $`k`$. For a plaquette, the natural decomposition is
$$\varphi _p=\varphi _1+\varphi _2+\varphi _3+\varphi _4,$$
(46)
where the phases $`\varphi _i(i=1,\mathrm{},4)`$ are associated with the corresponding links. The decomposition (46) comes about naturally in the basis of the coherent states. Indeed, for a particular coherent state the whole evolution may be reduced to a phase factor:
$$|\psi (t)=e^{i\varphi (t)}|\psi (0).$$
(47)
Moreover, the coherent states can be explicitly constructed in terms of the link matrices, for details and further references see . As a result, for any given lattice fields configuration, one can determine $`k`$ and detect the Dirac strings in this way. The monopoles are defined then as the end points of the strings.
(iii) The phase $`\varphi (t)`$ can in fact be decomposed into the dynamical and Berry phase. It is useful for this purpose to introduce a single-valued state vector $`|\stackrel{~}{\psi }`$ defined as
$$|\stackrel{~}{\psi }(T)=|\stackrel{~}{\psi }(0),$$
(48)
where $`T`$ is the period of the motion so that at $`t=T`$ the system comes back to the same point in the parameter space as at the moment $`t=0`$. Then
$$\varphi (T)=\delta +\gamma =_0^T\stackrel{~}{\psi }|H|\stackrel{~}{\psi }+i_C\stackrel{~}{\psi }|\frac{}{\lambda _i}|\stackrel{~}{\psi }𝑑\lambda ^i,$$
(49)
where $`\lambda _i`$ are parameters, $`\lambda _i(T)=\lambda _i(0)`$ and $`C`$ is a closed contour in the parameter space.
### 3.3 Choice of the gauge and the numerical results.
The steps (i) – (iii) described above fully determine monopoles as geometrical objects. As a mathematical construct, it certainly appears very appealing. However, from the physical point of view the crucial observation is the gauge dependence of the monopoles constructed in this way. As a result, the monopoles are devoid, generally speaking of any physical meaning. It is amusing that one can actually specify the conditions for the monopoles to be physical objects. In particular, the monopole density $`\rho `$ should satisfy the renormgroup equation:
$$\rho =\mathrm{const}\beta ^{153/121}\mathrm{exp}\left(\frac{9\pi ^2}{11}\beta \right),$$
(50)
where $`\beta 4/g^2`$. The condition (50) is a very strong constraint and there is no much surprise that the monopoles defined according to the procedure outlined in the preceding subsection, generally speaking, do not satisfy (50).
To continue with the physics, we need a physically motivated choice of the gauge. At first sight, such a choice is impossible. However, one can argue that the Lorenz gauge is a proper gauge. The Lorenz gauge on the lattice is defined by the requirement that the functional
$$R=\underset{l}{}\left(1\frac{1}{2}\mathrm{Tr}U_l\right)$$
(51)
is minimal on the gauge orbit ($`U_l`$ denote the link matrices). In the naive continuum limit (51) reduces to $`R=1/4(A_\mu ^a)^2`$.
The logic behind the choice (51) is as follows. In the continuum limit both the Dirac strings and monopoles correspond to singular gauge potentials $`A_\mu ^a`$. It is easy to imagine, therefore, that one can generate an arbitrary number of spurious strings and monopoles by going to arbitrary large potentials $`A`$, so to say inflated by the gauge transformations. On the other hand, by minimizing potentials one may hope to squeeze the number of the topological defects to its minimum and these topological defects may be physically significant.
And, indeed, the numerical simulations indicate that the geometrical monopoles defined in the Lorenz gauge are physical objects, i.e. their density satisfies the condition (50). There are also other indications that the geometrical monopoles are physical . For example, there is an excess of the non-Abelian action associated with them.
### 3.4 Conclusions # 3.
Monopoles with $`Q_m=2`$ unify properties of field-theoretical and statistical objects. Namely, on one hand the monopoles are defined locally in terms of the link matrices. However, the link matrices are gauge dependent and existence or non-existence of a monopole at a particular point is devoid of physical meaning for this reason. On the other hand, if one introduces gauge fixing in a physically reasonable way, the statistical properties of the monopoles, such as their density, satisfy very non-trivial renormgroup constraints and demonstrate their physical significance.
## 4 Phenomenological applications.
### 4.1 The effective Lagrangian and the Casimir scaling.
The standard way to develop a phenomenology is to assume that the monopoles condense. We will follow the suit and modify the Zwanziger Lagrangian (21) by adding the effective Higgs interaction where the role of the Higgs field is played by the monopole field $`\varphi _m`$:
$$S_{eff}=S_{dual}(A^a,B)+S_{Higgs}(B,\varphi _m),$$
(52)
where $`S_{Higgs}`$ is the standard action of the Abelian Higgs model. The vacuum expectation value of the Higgs, or monopole field is, of course, of order $`\mathrm{\Lambda }_{QCD}`$.
Despite its apparent simplicity, Eq. (52) is highly speculative. Namely, it unifies so to say fundamental gluons, $`A^a`$, their dual counterpart $`B`$ which is an Abelian gauge boson, and $`\varphi _m`$ which is presumably an effective scalar field. One may justify the use of (52) by assuming that the effective size of the monopoles with $`Q_m=2`$ is in fact numerically small, although generically it is of order $`\mathrm{\Lambda }_{QCD}`$. While in our presentation here we follow mostly the lines of Refs. , let us note that similar consequences arise within the models also introducing a new mass scale.
What is also specific about the Lagrangian (52) is that the dual gluon is a $`U(1)`$ gauge boson. The color symmetry is maintained by averaging over all possible embeddings of the (dual) $`U(1)`$ into $`SU(2)`$, see the discussion in section 2.2. The confinement mechanism inherent to (52) is the formation of the Abrikosov-Nielsen-Olesen string which can be considered alredy on the classical level. More generally, the Lagrangian (52) exibits the Abelian dominance in the confining region which is the dominance of Abelian-like field configurations in the full non-Abelian theory. This dominance is common to all the realizations of the dual-superconductor model of confinement and is strongly supported by the lattice data . What we avoid, however, is the breaking of $`SU(2)`$ to $`U(1)`$ which is inherent to the models which start with the dual gluons in the adjoint representation and then add effective isospin one Higgs fields. Such models have well-known principal difficulties with, say, describing interaction of the adjoint sources, see, e.g. .
To the contrary, the model (52) can be applied to consider the static interaction of the sources belonging to various representations of $`SU(2)`$. One of the basic facts here, established through the numerical simulations on the lattice , is the so called Casimir scaling. The phenomenon of the Casimir scaling is that the static potential is described by a sum of the Coulomb-like and linear terms:
$$V_j(r)j(j+1)\frac{\alpha _s}{\pi r}+j(j+1)\sigma r,$$
(53)
where $`j`$ labels the representation (we consider the $`SU(2)`$ case) and $`\sigma `$ is independent of $`j`$. Note that at large distances one expects qualitatively different behavior of the potential for integer and half-integer spins $`j`$ because of the string breaking in case of the integer representations. However, at presently measured distances Eq. (53) turns to be a very good approximation to the potential.
The potential of the type (53) does arise in the classical approximation in the model (52) because there are classical string solutions. However, the tension of the string is now a dynamical quantity which can be found as a function of the parameters of the model, that is vector and Higgs masses:
$$\sigma =\sigma _j(m_H/m_V).$$
(54)
In particular, the Casimir scaling holds in the London limit,
$$\frac{\sigma _{j_1}}{\sigma _{j_2}}=\frac{j_1(j_1+1)}{j_2(j_2+1)},\mathrm{if}m_Hm_V.$$
(55)
Thus, the model (52) can incorporate the Casimir scaling.
However, the description of the profile of the confining string is the best if $`m_Hm_V`$ . Thus, there is mismatch with (55). Since the functions of $`m_H/m_V`$ involved in the fits are rather smooth, it is possible to get a compromise description which is valid in both cases, say, with 20% accuracy. Which is not bad at all keeping in mind that we are using a classical approximation. Nevertheless, the fact that the Casimir scaling works at a per cent level remains a kind of an unexplained mystery in the classical approximation. Further analysis of this point might be needed.
### 4.2 Unconventional power corrections.
We introduced (52) as an effective Lagrangian. Now we will describe a rather paradoxical situation that this Lagrangian seem to provide with better phenomenology of the power corrections to the parton approximation than the conventional QCD approach. Namely, there are novel $`1/Q^2`$ corrections inherent to Higgs models which are absent in the standard considerations<sup>2</sup><sup>2</sup>2 Literally, the model (52) which introduces averaging over all the embedding of the Abelian dual gluon into $`SU(2)`$ has not been ever discussed so far. However, as far as the power corrections are concerned there is no difference from the cases considered in . . Moreover, these corrections seem to fit the data at all the distances measured so far, that is $`r0.1\mathrm{fm}`$. Example of this type was found in Refs. . Further support came from the instanton physics . We have already reviewed the unconventional power corrections in and will be brief here.
In the standard approach, the power corrections are given by matrix elements of various operators constructed on the quark and gluon fields . For our set up, the central point is that for the vacuum state the simplest matrix element had dimension $`d=4`$,
$$0|\alpha _s(G_{\mu \nu }^a)^2|0\mathrm{\Lambda }_{QCD}^4,$$
(56)
and, as a result, there are no $`\mathrm{\Lambda }_{QCD}^2/Q^2`$ corrections . On the other hand, the Higgs model has a mass parameter built it. This mass parameter can thought of as the mass of the vector particle, $`m_V^2`$. Let us list a few examples where the two approaches lead to different predictions:
(1) First, within the Higgs model the dual gluon gets mass and the monopole potential becomes Yukawa type:
$$V_{m\overline{m}}\frac{\pi }{g^2r}e^{m_Vr}.$$
(57)
The prediction has already been confirmed by the data . In the conventional approach, one should have to remove at least the term $`m^2r`$ at short distances. The quality of the data might be not so good as to rule this out, but the possibility looks quite bizarre.
(2) Since the dual and “ordinary” gluon are in fact the same particles (see discussion in section 2) one would assume that the massiveness of the dual gluon implies the massiveness of the gluon interacting with the color. But this is not true ! There is no analyticity in this sense. And the reason is again problems with the Dirac veto which we had already chance to discuss in connection with the radiative corrections (see section 2.3). Namely, as far as we discuss only the “dual world”, one can forget about the Dirac strings. However, if we introduce color sources $`Q\overline{Q}`$ into the vacuum with $`\varphi _M0`$ we should respect the Dirac veto. The ordinary operator product expansion, or perturbative expansion in $`m_V^2/Q^2`$ do not respect this veto – like ordinary perturbation theory does not do this either (see section 2.1). The correct treatment demonstrates that their is a linear correction to the quark potential at short distances:
$$\delta V_{Q\overline{Q}}=\sigma _0r,$$
(58)
where $`\sigma _0`$ is calculable function of $`m_H,m_V`$. Within the standard approach there is no such term, for explanations and further references see . The data do support the presence of the linear term. Amusingly enough, the data refer exclusively to the the nonperturbative potential, and there is no need for painful separation of (small) power corrections against the perturbative “background”.
(3) The linear term (58) can be rephrased as the statement that the gluon has a tachyonic mass. Indeed, the ordinary mass would give a negative $`\sigma _0`$, as seen from the expansion of the Yukawa potential at short distances. The introduction of a tachyonic gluon mass in the framework of the QCD sum rules allows to resolve in an absolutely natural way long standing problems with the phenomenology based on the QCD sum rules .
(4) The last but not the least point in our discussion concerns the instanton density . The conventional approach predicts that the deviations from the ’t Hooft instanton density due to non-trivial background vacuum fields is of the fourth order in the instanton size $`\rho `$:
$$dn(\rho )=dn_{pert.}(\rho )(1+\frac{\pi ^4\rho ^4}{2g^4}0|g^2(G_{\mu \nu }^a)^2|0+\mathrm{})$$
(59)
The data, on the other hand are beautifully fitted by a quadratic correction, inherent to (52). Note however, that the coefficient in front of the quadratic term has been fitted rather than calculated from (52) so far.
### 4.3 Conclusions # 4
We have proposed in this section a phenomenological Lagrangian (52) which unifies the Higgs mechanism for the Abelian dual gluon with full $`SU(2)`$ symmetry of the ordinary gluodynamics. The full study of the consequences from this formulation is still awaiting its time to come.
However, it seems promising that the color SU(2) is not broken at any step despite the Higgs mechanism. This allows to broaden applications of the effective Lagrangian and incorporate, to certain accuracy, the Casimir scaling.
Also, emergence of the mass of the dual gluon in the effective Lagrangian approach provides with a natural framework to introduce the novel $`1/Q^2`$ corrections. Phenomenologically, these corrections bring crucial improvements to the existing phenomenology. Moreover, generically the corrections are of the same type as those associated with ultravioloet renormalons (see, e.g., ). However, within the effective Lagrangian approach these corrections should disappear in the limit of infinite $`Q^2`$ which is not true for the ultraviolet renormalons and has not been supported by any data so far.
## 5 Conclusions.
In this review we considered various effects related to the monopoles in unbroken non-Abelian gauge theories. In conclusion, let us reiterate the main poins (see also conclusions to the Sections 1–4):
(i) Fundamental monopoles with the magnetic charge $`|Q_m|=1`$ are introduced as external objects via the ’t Hooft loop. The corresponding intermonopole potential $`V_{m\overline{m}}(r)`$ can be evaluated at short distances from first principles. In the Lagrangian approach similar to that of Zwanziger, the dual gluon interacting with point-like external monopoles appears as an Abelian gauge field (see Section 2 and for details).
(ii) Monopoles with the magnetic charge $`|Q_m|=2`$ are pure quantum objects which can be studied so far only numerically. We discussed briefly the newly intoduced geometrical monopoles which appear to be physical objects.
(iii) Effective Lagrangian which assumes condensation of the monopoles incorporates the Abelian dominance at distances, where the effects of confinement are crucial, without breaking $`SU(2)`$ to $`U(1)`$. In the London limit, it reproduces the the Casimir scaling phenomenon. There are further phenomenological consequences, in particular, the evaluation of the potential $`V_{m\overline{m}}(r)`$ at larger distances (see Section 4 and for details).
## Acknowledgments
The review is based to a large extent on talks presented by the authors on various conferences this year. The authors are gratefull to the organizers for the invitations. We are thankful to V.A. Shevchenko, Yu.A. Simonov, L. Stodolsky, T. Suzuki, V.A. Rubakov for valuable discussions. M.N.Ch. and M.I.P. acknowledge the kind hospitality of the staff of the Max-Planck Institut für Physik (München), where the work was initiated. Work of M.N.C., F.V.G. and M.I.P. was partially supported by grants RFBR 99-01230a and INTAS 96-370. M.N.Ch. and M.I.P. are supported by Monbushu grant and CRDF award RP1-2103.
|
warning/0007/math0007189.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
We recall that a pseudo-Riemannian manifold $`(M,g)`$ is called a symmetric space if any point $`xM`$ is an isolated fixed point of an involutive isometry $`s_x`$ (called central symmetry with centre $`x`$). Since the product of two central symmetries $`s_x`$ and $`s_y`$ with sufficiently close centres is a shift along the geodesic $`(xy)`$, the group generated by central symmetries acts transitively on $`M`$ and one can identify $`M`$ with the quotient $`M=G/K`$ where $`G`$ is the connected component of the isometry group $`\mathrm{Isom}(M,g)`$ and $`K`$ is the stabilizer of a point $`oM`$.
A symmetric space $`(M=G/K,g)`$ is called Kähler (respectively, hyper-Kähler) if its holonomy group Hol$`(M,g)`$ is a subgroup of the pseudo-unitary group $`\mathrm{U}(p,q)`$ (respectively, of the pseudo-symplectic group $`\mathrm{Sp}(p,q)\mathrm{SU}(2p,2q)`$). Any hyper-Kähler symmetric space is in particular a homogeneous hypercomplex manifold. Homogeneous hypercomplex manifolds of compact Lie groups were constructed by Ph. Spindel, A. Sevrin, W. Troost, A. Van Proeyen \[SSTVP\] and by D. Joyce \[J\] and homogeneous hypercomplex structures on solvable Lie groups by M.L. Barberis and I. Dotti-Miatello \[BD\].
The classification of simply connected symmetric spaces reduces to the classification of involutive automorphisms $`\sigma `$ of a Lie algebra g, such that the adjoint representation $`\mathrm{ad}_\text{k}|\text{m}`$ preserves a pseudo-Euclidean scalar product $`g`$, where
$$\text{g}=\text{k}+\text{m},\sigma |\text{k}=1,\sigma |\text{m}=1,$$
is the eigenspace decomposition of the involution $`\sigma `$. Note that the eigenspace decomposition of an involutive automorphism is characterized by the conditions
$$[\text{k},\text{k}]\text{k},[\text{k},\text{m}]\text{m},[\text{m},\text{m}]\text{k}.$$
Such a decomposition is called a symmetric decomposition.
In fact, for any pseudo-Riemannian symmetric space $`M=G/K`$ the conjugation with respect to the central symmetry $`s_o`$ with centre $`o=eK`$ is an involutive automorphism of the Lie group $`G`$, which induces an involutive automorphism $`\sigma `$ of its Lie algebra g. The pseudo-Riemannian metric of $`M`$ induces a k-invariant scalar product on $`\text{m}T_oM`$, where $`\text{g}=\text{k}+\text{m}`$ is the symmetric decomposition defined by $`\sigma `$. Conversely, a symmetric decomposition $`\text{g}=\text{k}+\text{m}`$ together with a k-invariant scalar product on m determines a pseudo-Riemannian symmetric space $`M=G/K`$ where $`G`$ is the simply connected Lie group with the Lie algebra g, $`K`$ is the connected (and closed) subgroup of $`G`$ generated by k, the pseudo-Riemannian metric on $`M`$ is defined by $`g`$ and the central symmetry is defined by the involutive automorphism $`\sigma `$ associated to the symmetric decomposition.
Naturally identifying the space m with the tangent space $`T_oM`$, the isotropy group is identified with $`\mathrm{Ad}_K|\text{m}`$ and the holonomy algebra is identified with $`\mathrm{ad}_\text{h}`$ where $`\text{h}=[\text{m},\text{m}]`$. If one assumes that the holonomy algebra is irreducible then one can prove that the Lie algebra g is semisimple. Hence the classification of pseudo-Riemannian symmetric spaces with irreducible holonomy reduces to the classification of involutive authomorphisms of semisimple Lie algebras. Such a classification was obtained by M. Berger \[B1, B2\] and A. Fedenko \[F\]. It includes the classification of Riemannian symmetric spaces (obtained earlier by E. Cartan), since according to de Rham’s theorem any simply connected complete Riemannian manifold is a direct product of Riemannian manifolds with irreducible holonomy algebra and a Euclidean space.
A classification of pseudo-Riemannian symmetric spaces with non completely reducible holonomy is known only for signature $`(1,n)`$ (Cahen-Wallach \[C-W\]) and for signature $`(2,n)`$ under the assumption that the holonomy group is solvable (Cahen-Parker \[C-P\]). The classification problem for arbitrary signature looks very complicated and includes, for example, the classification of Lie algebras which admit a nondegenerate ad-invariant symmetric bilinear form. Some inductive construction of solvable Lie algebras with such a form was given by V. Kac.
In this paper we give a classification of pseudo-Riemannian hyper-Kähler symmetric spaces. In particular, we prove that any simply connected hyper-Kähler symmetric space $`M`$ has signature (4m,4m) and its holonomy group is commutative. The main result is the following, see Theorem 6.
Let $`(E,\omega ,j)`$ be a complex symplectic vector space of dimension $`4m`$ with a quaternionic structure $`j`$ such that $`\omega (jx,jy)=\overline{\omega (x,y)}`$ for all $`x,yE`$ and $`E=E_+E_{}`$ a $`j`$-invariant Lagrangian decomposition. Such a decomposition exists if and only if the Hermitian form $`\gamma =\omega (,j)`$ has real signature $`(4m,4m)`$. We denote by $`\tau `$ the real structure in $`S^{2r}E`$ defined by $`\tau (e_1e_2\mathrm{}e_{2r}):=j(e_1)j(e_2)\mathrm{}j(e_{2r})`$, $`e_iE`$. Then any element $`S(S^4E_+)^\tau `$ defines a hyper-Kähler symmetric space $`M_S`$ which is associated with the symmetric decomposition
$$\text{g}=\text{h}+\text{m},$$
where $`\text{m}=(\mathrm{}^2E)^\rho `$ is the fixed point set of the real structure $`\rho `$ on m given by $`\rho (he)=j_Hhje`$, where $`j_H`$ is the standard quaternionic structure on $`\mathrm{}^2=\mathrm{}`$, $`\text{h}=\mathrm{span}\{S_{ee^{}}|e,e^{}E\}^\tau sp(E)^\tau \mathrm{sp}(m,m)`$ with the natural action on $`\text{m}\mathrm{}^2E`$ and the Lie bracket $`\text{m}\text{m}\text{h}`$ is given by
$$[he,h^{}e^{}]=\omega _H(h,h^{})S_{e,e^{}}.$$
Moreover, we establish a natural 1-1 correspondence between simply connected hyper-Kähler symmetric spaces (up to isomorphism) and orbits of the group $`GL(m,\mathrm{})`$ in $`(S^4E_+)^\tau `$.
We define also the notion of complex hyper-Kähler symmetric space as a complex manifold $`(M,g)`$ of complex dimension $`4n`$ with holomorphic metric $`g`$ such that for any point $`xM`$ there is a holomorphic central symmetry $`s_x`$ with centre $`x`$ and which has holonomy group $`\mathrm{Hol}(M,g)\mathrm{Sp}(n,\mathrm{})`$ ($`\mathrm{Sp}(n,\mathrm{})\mathrm{Sp}(n,\mathrm{})\times \mathrm{Sp}(n,\mathrm{})\mathrm{O}(4n,\mathrm{})`$ is diagonally embedded) and give a classification of such spaces. We establish a natural 1-1 correspondence between simply connected complex hyper-Kähler symmetric spaces and homogeneous polynomials of degree 4 in the vector space $`\mathrm{}^{2n}`$ considered up to linear transformations from $`GL(2n,\mathrm{})`$.
## 2 Symmetric spaces
### 2.1 Basic facts about pseudo-Riemannian symmetric spaces
A pseudo-Riemannian symmetric space is a pseudo-Riemannian manifold $`(M,g)`$ such that any point is an isolated fixed point of an isometric involution. Such a pseudo-Riemannian manifold admits a transitive Lie group of isometries $`L`$ and can be identified with $`L/L_o`$, where $`L_o`$ is the stabilizer of a point $`o`$. More precisely, any simply connected pseudo-Riemannian symmetric space $`M=G/K`$ is associated with a symmetric decomposition
$$\text{g}=\text{k}+\text{m},[\text{k},\text{k}]\text{k},[\text{k},\text{m}]\text{m},[\text{m},\text{m}]\text{k}$$
(2.1)
of the Lie algebra $`\text{g}=LieG`$ together with an $`\mathrm{Ad}_K`$-invariant pseudo-Euclidean scalar product on m. We will assume that G acts almost effectively on $`M`$, i.e. k does not contain any nontrivial ideal of g, that $`M`$ and $`G`$ are simply connected and that $`K`$ is connected. Then, under the natural identification of the tangent space $`T_oM`$ at the canonical base point $`o=eK`$ with m, the holonomy group $`\mathrm{Hol}`$ Ad$`{}_{K}{}^{}|\text{m}`$. We will denote by h the holonomy Lie algebra. Since the isotropy representation is faithfull it is identified with the subalgebra $`\text{h}=[\text{m},\text{m}]:=\mathrm{span}\{[x,y]|x,y\text{m}\}\text{k}`$. Recall that the curvature tensor $`R`$ of a symmetric space $`M`$ at $`o`$ is h-invariant and determines the Lie bracket in the ideal $`\text{h}+\text{m}\text{g}`$ as follows:
$$\text{h}=R(\text{m},\text{m}):=\mathrm{span}\{R(x,y)|x,y\text{m}\}\text{and}[x,y]=R(x,y),x,y\text{m}.$$
The following result is well known:
###### Proposition 1
The full Lie algebra of Killing fields of a symmetric space has the form
$$\mathrm{isom}(M)=\stackrel{~}{\text{h}}+\text{m},$$
where the full isotropy subalgebra is given by
$$\stackrel{~}{\text{h}}=\mathrm{aut}(R)=\{A\mathrm{so}(\text{m})|AR=[A,R(,)]R(A,)R(,A)=0\}.$$
(2.2)
### 2.2 Symmetric spaces of semisimple Lie groups
We will prove that in the case when $`(M=G/K,g)`$ is a pseudo-Riemannian symmetric space of a (connected) semisimple Lie group $`G`$ then $`G`$ is the maximal connected Lie group of isometries of $`M`$.
###### Proposition 2
Let $`(M=G/K,g)`$ be a pseudo-Riemannian symmetric space associated with a symmetric decomposition $`\text{g}=\text{k}+\text{m}`$. If $`G`$ is semisimple and almost effective then
1. the restriction of the Cartan-Killing form $`B`$ of g to k is nondegenerate and hence k is a reductive subalgebra of g and $`\text{g}=\text{k}+\text{m}`$ is a $`B`$-orthogonal decomposition,
2. $`\text{k}=[\text{m},\text{m}]`$ and
3. $`\text{g}=\mathrm{isom}(M,g)`$ is the Lie algebra of the full isometry group of $`M`$.
Proof: For (i) see \[O-V\] Ch. 3 Proposition 3.6.
(ii) It is clear that $`\overline{\text{g}}=[\text{m},\text{m}]+\text{m}`$ is an ideal of g. The $`B`$-orthogonal complement $`\text{a}:=\overline{\text{g}}^{}\text{k}`$ is a complementary ideal of g. Since $`[\text{a},\text{m}]=0`$ the Lie algebra a acts trivially on $`M`$. From the effectivity of g we conclude that $`\text{a}=0`$.
(iii) By Proposition 1, $`\stackrel{~}{\text{g}}=\mathrm{isom}(M,g)=\stackrel{~}{\text{h}}+\text{m}`$, where $`\stackrel{~}{\text{h}}=\mathrm{aut}(R)=\{A\mathrm{so}(\text{m})|AR=0\}`$. Now $`\stackrel{~}{\text{h}}`$ preserves m and by the identity $`AR=[A,R(,)]R(A,)R(,A)`$ it also normalizes k. This shows that $`\stackrel{~}{\text{h}}`$ normalizes g and hence $`\text{g}\stackrel{~}{\text{g}}`$ is an ideal. Since g is semisimple there exists a g-invariant complement b in $`\stackrel{~}{\text{g}}`$. Note that $`[\text{g},\text{b}]\text{g}\text{b}=0`$. We can decompose any $`X\text{b}`$ as $`X=Y+Z`$, where $`Y\stackrel{~}{\text{h}}`$ and $`Z\text{m}`$. From $`[\text{g},\text{b}]=0`$ it follows that $`[\text{g},Y]=[\text{g},Z]=0`$ and in particular $`[\text{m},Y]=0`$. This implies that $`Y=0`$ and $`X=Z\text{b}\text{m}=0`$. This shows that $`\text{b}=0`$ proving (iii).
We recall that a pseudo-Riemannian Hermitian symmetric space is pseudo-Riemannian symmetric space $`(M=G/K,g)`$ together with an invariant (and hence parallel) $`g`$-orthogonal complex structure $`J`$.
###### Proposition 3
Let $`(M=G/K,g,J)`$ be a pseudo-Riemannian Hermitian symmetric space of a semisimple and almost effective Lie group $`G`$. Then the Ricci curvature of $`M`$ is not zero.
Proof: From Proposition 2 it follows that $`\text{g}=\mathrm{isom}(M,g)=\stackrel{~}{\text{h}}+m`$, where $`\stackrel{~}{\text{h}}=\text{k}=[\text{m},\text{m}]`$. It is well known that the curvature tensor $`R`$ of any pseudo-Kähler manifold (and in particular of any pseudo-Riemannian Hermitian symmetric space) is invariant under the operator $`J`$. This shows that $`J\stackrel{~}{\text{h}}=\mathrm{aut}(R)=[\text{m},\text{m}]=\text{h}`$ (holonomy Lie algebra), which implies that the holonomy Lie algebra is not a subalgebra of $`\mathrm{su}(\text{m})\mathrm{su}(p,q)`$. Hence $`M`$ is not Ricci-flat. In fact, we can write $`J=\mathrm{ad}[X_i,Y_i]`$, for $`X_i,Y_i\text{m}`$. Then using the formulas $`\mathrm{Ric}(X,JY)=\mathrm{tr}JR(X,Y)`$ for the Ricci curvature of a pseudo-Kähler manifold and $`R(X,Y)=ad_{[X,Y]}|\text{m}`$ for the curvature of a symmetric space we calculate:
$$\mathrm{Ric}(X_i,JY_i)=\mathrm{tr}JR(X_i,Y_i)=\mathrm{tr}J\mathrm{ad}[X_i,Y_i]=\mathrm{tr}J^20.$$
## 3 Structure of hyper-Kähler symmetric spaces
### 3.1 Definitions
A (possibly indefinite) hyper-Kähler manifold is a pseudo-Riemannian manifold $`(M^{4n},g)`$ of signature $`(4k,4l)`$ together with a compatible hypercomplex structure, i.e. three $`g`$-orthogonal parallel complex structures $`(J_1,J_2,J_3=J_1J_2)`$. This means that the holonomy group $`\mathrm{Hol}\mathrm{Sp}(k,l)`$. Two hyper-Kähler manifolds $`(M,g,J_\alpha )`$ ($`\alpha =1,2,3`$) and $`(M^{},g^{},J_\alpha ^{})`$ are called isomorphic if there exists a triholomorphic isometry $`\phi :MM^{}`$, i.e. $`\phi ^{}J_\alpha ^{}=J_\alpha `$ and $`\phi ^{}g^{}=g`$.
A hyper-Kähler symmetric space is a pseudo-Riemannian symmetric space $`(M=G/K,g)`$ together with an invariant compatible hypercomplex structure. Consider now a simply connected hyper-Kähler symmetric space $`(M=G/K,g,J_\alpha )`$. Without restriction of generality we will assume that $`G`$ acts almost effectively. $`M`$ being hyper-Kähler is equivalent to Ad$`{}_{K}{}^{}|\text{m}\mathrm{Sp}(k,l)`$, or, since $`K`$ is connected, to ad$`{}_{\text{k}}{}^{}|\text{m}\mathrm{sp}(k,l)`$. This condition means that k commutes with the Lie algebra $`Q=\mathrm{sp}(1)\mathrm{so}(\text{m})=\mathrm{so}(4k,4l)`$ spanned by three anticommuting complex structures $`J_1,J_2,J_3`$.
### 3.2 Existence of a transitive solvable group of isometries and solvability of the holonomy
In this subsection we prove that any simply connected hyper-Kähler symmetric space $`(M,g,J_\alpha )`$ admits a transitive solvable Lie group $`G\mathrm{Aut}(g,J_\alpha )`$ of automorphisms and has solvable holonomy group.
###### Proposition 4
Let $`(M=G/K,g,J_\alpha )`$ be a simply connected hyper-Kähler symmetric space and $`A=\mathrm{Aut}_0(g,Q)\mathrm{Aut}_0(g,J_\alpha )G`$ the connected group of isometries which preserve the quaternionic structure $`Q=\mathrm{span}\{J_\alpha \}`$. Then
1. the stabilizer $`A_o`$ of a point $`oM`$ contains a maximal semisimple subgroup of $`A`$,
2. the radical $`R`$ of $`A`$ acts transitively and triholomorphically on $`M`$ and
3. the holonomy group of $`M`$ is solvable.
Proof: We consider the quaternionic Kähler symmetric space $`(M=A/A_o,g,Q)`$. The Lie algebra $`\text{a}_o`$ of the stabilizer is given by
$$\text{a}_o=\mathrm{aut}(R,Q)=\{A\mathrm{so}(T_oM)|AR=0,[A,Q]Q\}.$$
Since the curvature tensor of a quaternionic Kähler manifold is invariant under the quaternionic structure $`Q`$ we conclude that $`Q\text{a}_o`$ and $`\text{a}_o=QZ_\text{a}(Q)`$, where $`Z_\text{a}(Q)`$ denotes the centralizer of $`Q`$ in a. Since $`Q\mathrm{sp}(1)`$ is simple, we may choose a Levi-Malcev decomposition $`\text{a}=\text{s}+\text{r}`$ such that the Levi subalgebra $`\text{s}Q`$. We put $`\text{m}_r:=[Q,\text{r}]`$ and denote by $`\text{m}_s`$ a $`QZ_\text{s}(Q)`$-invariant complement of $`Q`$ in $`[Q,\text{s}]`$. The stabilizer has the decomposition $`\text{a}_o=Q(Z_\text{s}(Q)+Z_\text{r}(Q))`$.
###### Lemma 1
The complement $`\text{m}=\text{m}_s+\text{m}_r`$ to $`\text{a}_o`$ in a is $`\text{a}_o`$-invariant and the decomposition
$$\text{a}=\text{a}_o+\text{m}$$
is a symmetric decomposition.
Proof: It is clear that $`\text{m}_r`$ is $`\text{a}_o`$-invariant and $`\text{m}_s`$ is invariant under $`QZ_\text{s}(Q)`$ by construction. It remains to check that $`[Z_\text{r}(Q),\text{m}_s]\text{m}`$. Since $`\text{m}_s=[Q,\text{m}_s]`$, we have
$$[Z_\text{r}(Q),\text{m}_s]=[Z_\text{r}(Q),[Q,\text{m}_s]]=[Q,[Z_\text{r}(Q),\text{m}_s]][Q,\text{r}]=\text{m}_r\text{m}.$$
This shows that $`\text{a}=\text{a}_o+\text{m}`$ is an $`\text{a}_o`$-invariant decomposition. We denote by $`\text{a}=\text{a}_o+\text{p}`$ a symmetric decomposition. Any other $`\text{a}_o`$-invariant decomposition is of the form $`\text{a}=\text{a}_o+\text{p}_\phi `$, where $`\phi :\text{p}\text{a}_o`$ is an $`\text{a}_o`$-equivariant map and $`\text{p}_\phi =\{X+\phi (X)|X\text{p}\}`$. If such non-zero equivariant map $`\phi `$ exists then p and $`\text{a}_o`$ contain non-trivial isomorphic $`Q`$-submodules. Since p is a sum of 4-dimensional irreducible $`Q`$-modules and $`\text{a}_o`$ is the sum of the 3-dimensional irreducible $`Q`$-module $`Q`$ and the trivial complementary $`Q`$-module $`Z_{\text{a}_o}(Q)`$, we infer that there exists a unique $`\text{a}_o`$-invariant decomposition, which coincides with the symmetric decomposition $`\text{a}=\text{a}_o+\text{p}`$.
To prove (i) we have to check that $`\text{m}_s=0`$. We note that by the previous lemma $`\text{s}=(QZ_\text{s}(Q))+\text{m}_s`$ is a symmetric decomposition of the semisimple Lie algebra s. Since $`[\text{m}_s,\text{m}_s]Z_\text{s}(Q)`$ it defines a hyper-Kähler symmetric space $`N=L/L_o`$, where $`L`$ is the simply connected semisimple Lie group with Lie algebra $`\text{l}=Z_\text{s}(Q)+\text{m}_s`$ and $`L_o`$ is the Lie subgroup generated by the subalgebra $`Z_\text{s}(Q)\text{l}`$. Since $`N`$ is in particular a Ricci-flat pseudo-Riemannian Hermitian symmetric space, from Proposition 3 we obtain that $`N`$ is reduced to point. Therefore $`\text{m}_s=0`$. This proves (i) and (ii). Finally, since the holonomy Lie algebra h is identified with $`\text{h}=[\text{m},\text{m}]=[\text{m}_r,\text{m}_r]\text{r}`$ it is solvable as subalgebra of the solvable Lie algebra r.
### 3.3 Hyper-Kähler symmetric spaces and second prolongation of symplectic Lie algebras
Let $`(M=G/K,g,J_\alpha )`$ be a simply connected hyper-Kähler symmetric space associated with a symmetric decomposition (2.1). Without restriction of generality we will assume that $`G`$ acts almost effectively and that $`\text{k}=[\text{m},\text{m}]=\text{h}`$ (holonomy Lie algebra). The complexification $`\text{m}^{\mathrm{}}`$ as $`\text{h}^{\mathrm{}}`$-module can be written as $`\text{m}^{\mathrm{}}=HE`$, such that $`\text{h}^{\mathrm{}}\mathrm{Id}\mathrm{sp}(E)\mathrm{sp}(E)`$, where $`H=\mathrm{}^2`$ and $`E=\mathrm{}^{2n}`$ are complex symplectic vector spaces with symplectic form $`\omega _H`$ and $`\omega _E`$, respectively, such that $`g^{\mathrm{}}=\omega _H\omega _E`$ is the complex bilinear metric on $`\text{m}^{\mathrm{}}`$ induced by $`g`$. Note that the symplectic forms are unique up to the transformation $`\omega _H\lambda \omega _H`$, $`\omega _E\lambda ^1\omega _E`$, $`\lambda \mathrm{}^{}`$. We have also quaternionic structures $`j_H`$ and $`j_E`$ on $`H`$ and $`E`$, such that $`\omega _H(j_Hx,j_Hy)=\overline{\omega _H(x,y)}`$ for all $`x,yH`$ and $`\omega _E(j_Ex,j_Ey)=\overline{\omega _E(x,y)}`$ for all $`x,yE`$, where the bar denotes complex conjugation. This implies that $`\gamma _H:=\omega _H(,j_H)`$ and $`\gamma _E:=\omega _E(,j_E)`$ are Hermitian forms on $`H`$ and $`E`$. For fixed $`\omega _H`$ and $`\omega _E`$ the quaternionic structures $`j_H`$ and $`j_E`$ are uniquely determined if we require that $`\gamma _H`$ is positive definite and that $`\rho =j_Hj_E`$ is the real structure on $`\text{m}^{\mathrm{}}`$, i.e. the complex conjugation with respect to m. The metric $`g^{\mathrm{}}`$ and the Hermitian form $`g^{\mathrm{}}(,\rho )=\gamma _H\gamma _E`$ restrict to a real valued scalar product $`g`$ of some signature $`(4k,4l)`$ on $`\text{m}=(HE)^\rho `$, where $`(2k,2l)`$ is the (real) signature of the Hermitian form $`\gamma _E=\omega _E(,j_E)`$. Note that for the holonomy algebra we have the inclusion
$$\text{h}=\mathrm{Id}(\text{h}^{\mathrm{}})^{j_E}\mathrm{sp}(E)^{j_E}=\{A\mathrm{sp}(E)|[A,j_E]=0\}=\mathrm{aut}(E,\omega _E,j_E)\mathrm{aut}(\text{m},g,J_\alpha )\mathrm{sp}(k,l).$$
Using the symplectic forms we identify $`H=H^{}`$ and $`E=E^{}`$. Then the symplectic Lie algebras are identified with symmetric tensors as follows:
$$\mathrm{sp}(H)=S^2H,\mathrm{sp}(E)=S^2E.$$
Since the curvature tensor $`R`$ of any hyper-Kähler manifold $`M^{4n}`$ can be identified at a point $`pM`$ with an element $`RS^2\mathrm{sp}(k,l)`$ it is invariant under the Lie algebra $`\mathrm{sp}(1)=\mathrm{span}\{J_1,J_2,J_3\}`$. Let $`M=G/K`$ be a hyper-Kähler symmetric space as above. By Proposition 1 we can extend the Lie algebra $`\text{g}=\text{k}+\text{m}=\text{h}+\text{m}`$ to a Lie algebra
$$\stackrel{~}{\text{g}}=\mathrm{sp}(1)+\text{h}+\text{m}$$
of Killing vector fields such that $`[\mathrm{sp}(1),\text{h}]=0`$. In the $`HE`$-formalism the Lie algebra $`\mathrm{sp}(1)`$ is identified with $`\mathrm{sp}(H)^{j_H}\mathrm{Id}\mathrm{so}(\text{m})`$.
###### Lemma 2
Denote by $`\stackrel{~}{\text{g}}^{\mathrm{}}=\mathrm{sp}(1,\mathrm{})+\text{h}^{\mathrm{}}+\text{m}^{\mathrm{}}`$ the complexification of the Lie algebra $`\stackrel{~}{\text{g}}`$. Then the Lie bracket $`[,]:^2\text{m}^{\mathrm{}}\text{h}^{\mathrm{}}`$ can be written as
$$[he,h^{}e^{}]=\omega _H(h,h^{})S_{e,e^{}},$$
(3.1)
where $`S(\text{h}^{\mathrm{}})^{(2)}:=\text{h}^{\mathrm{}}S^2E^{}ES^3E^{}=\text{h}^{\mathrm{}}\text{h}^{\mathrm{}}S^4E`$. Moreover $`S`$ is $`\mathrm{sp}(1,\mathrm{})\text{h}^{\mathrm{}}`$-invariant and satisfies the following reality condition: $`[S_{j_Ee,e^{}}S_{e,j_Ee^{}},j_E]=0`$.
Proof: The Lie bracket $`[,]:^2\text{m}^{\mathrm{}}\text{h}^{\mathrm{}}`$ is an $`\mathrm{sp}(1,\mathrm{})\text{h}^{\mathrm{}}`$-equivariant map, due to the Jacobi identity. We decompose the $`\mathrm{sp}(H)\mathrm{sp}(E)`$-module $`^2\text{m}^{\mathrm{}}`$:
$$^2\text{m}^{\mathrm{}}=^2(HE)=^2HS^2ES^2H^2E=\omega _HS^2ES^2H^2E.$$
Since $`\text{h}^{\mathrm{}}S^2E`$ the Lie bracket defines an $`\mathrm{sp}(1,\mathrm{})\text{h}^{\mathrm{}}`$-invariant element of the space $`\omega _HS^2ES^2ES^2H^2ES^2E`$. The second summand has no nontrivial $`\mathrm{sp}(1,\mathrm{})`$-invariant elements. Hence the bracket is of the form (3.1), where $`SS^2E^{}\text{h}^{\mathrm{}}S^2ES^2E`$. The Jacobi identity reads:
$$0=[he,[h^{}e^{},h^{\prime \prime }e^{\prime \prime }]][[he,h^{}e^{}],h^{\prime \prime }e^{\prime \prime }][h^{}e^{},[he,h^{\prime \prime }e^{\prime \prime }]]$$
$$=\omega _H(h^{},h^{\prime \prime })hS_{e^{},e^{\prime \prime }}e\omega _H(h,h^{})h^{\prime \prime }S_{e,e^{}}e^{\prime \prime }+\omega _H(h,h^{\prime \prime })h^{}S_{e,e^{\prime \prime }}e^{}.$$
Since $`dimH=2`$ we may assume that $`h,h^{}=h^{\prime \prime }`$ is a symplectic basis, i.e. $`\omega _H(h,h^{})=1`$, and the equation implies: $`S_{e,e^{\prime \prime }}e^{}=S_{e,e^{}}e^{\prime \prime }`$, i.e. $`S(\text{h}^{\mathrm{}})^{(2)}`$. The Lie bracket of two real elements $`he+j_Hhj_Ee`$ and $`he^{}+j_Hhj_Ee^{}\text{m}\text{m}^{\mathrm{}}`$ is an element of h. This gives:
$$[he+j_Hhj_Ee,he^{}+j_Hhj_Ee^{}]=\omega _H(h,j_Hh)(S_{e,j_Ee^{}}S_{j_Ee,e^{}})\text{h}.$$
From the fact that the Hermitian form $`\gamma _H=\omega _H(,j_H)`$ is positive definite it follows that $`\omega _H(h,j_Hh)0`$. This establishes the reality condition since $`\text{h}=\{A\text{h}^{\mathrm{}}|[A,j_E]=0\}`$.
In fact any tensor $`SS^4E`$ satisfying the conditions of the above lemma can be used to define a hyper-Kähler symmetric space as the following theorem shows. We can identify $`S^4E`$ with the space $`\mathrm{}[E]^{(4)}`$ of homogeneous quartic polynomials on $`EE^{}`$.
###### Theorem 1
Let $`SS^4E`$, $`E=\mathrm{}^{2n}`$, be a quartic polynomial invariant under all endomorphisms $`S_{e,e^{}}S^2E=\mathrm{sp}(E)`$ and satisfying the reality condition
$$[S_{j_Ee,e^{}}S_{e,j_Ee^{}},j_E]=0.$$
(3.2)
Then it defines a hyper-Kähler symmetric space, which is associated with the following complex symmetric decomposition
$$\text{g}^{\mathrm{}}=\text{h}^{\mathrm{}}+HE,\text{h}^{\mathrm{}}=\mathrm{span}\{S_{e,e^{}}|e,e^{}E\}\mathrm{sp}(E).$$
(3.3)
The bracket $`^2(HE)\text{h}^{\mathrm{}}`$ is given by (3.1). The real symmetric decomposition is defined as $`\rho `$-real form $`\text{g}=\text{h}+\text{m}`$ of (3.3), where
$$\text{k}=\text{h}=\{A\text{h}^{\mathrm{}}|[A,j_E]=0\}=\mathrm{span}\{S_{j_Ee,e^{}}S_{e,j_Ee^{}}|e,e^{}E\},\text{m}=(HE)^\rho .$$
The hyper-Kähler symmetric space $`M`$ associated to this symmetric decomposition is the quotient $`M=M_S=G/K`$, where $`G`$ is the simply connected Lie group with Lie algebra g and $`KG`$ is the connected (and closed) subgroup with Lie algebra $`\text{k}=\text{h}`$.
Moreover any simply connected hyper-Kähler symmetric space can be obtained by this construction. Two hyper-Kähler symmetric spaces $`M_S`$ and $`M_S^{}`$ defined by quartics $`S`$ and $`S^{}`$ are isomorphic if and only if $`S`$ and $`S^{}`$ are in the same orbit of the group $`\mathrm{Aut}(E,\omega _E,j_E)=\{A\mathrm{Sp}(E)|[A,j_E]=0\}\mathrm{Sp}(k,l)`$.
Proof: First of all we note that $`\text{h}^{\mathrm{}}=S_{E,E}:=\mathrm{span}\{S_{e,e^{}}|e,e^{}E\}`$ is a subalgebra of $`\mathrm{sp}(E)`$ because
$$[S_{e,e^{}},S_{f,f^{}}]=(S_{e,e^{}}S)_{f,f^{}}S_{S_{e,e^{}}f,f^{}}S_{f,S_{e,e^{}}f^{}}=S_{S_{e,e^{}}f,f^{}}S_{f,S_{e,e^{}}f^{}}\text{h}^{\mathrm{}}.$$
Since $`S`$ is $`\text{h}^{\mathrm{}}`$-invariant and completely symmetric we can check, as in Lemma 2, that the Jacobi identity is satisfied and that (3.3) defines a complex symmetric decomposition. We prove that $`\text{h}:=\mathrm{span}\{S_{j_Ee,e^{}}S_{e,j_Ee^{}}|e,e^{}E\}\{A\text{h}^{\mathrm{}}|[A,j_E]=0\}`$ defines a real form of $`\text{h}^{\mathrm{}}`$. Indeed for $`e,e^{}E`$ we have
$$S_{e,e^{}}=\frac{1}{2}(S_{e,e^{}}+S_{j_Ee,j_Ee^{}})\frac{\sqrt{1}}{2}(\sqrt{1}S_{e,e^{}}\sqrt{1}S_{j_Ee,j_Ee^{}})$$
$$=\frac{1}{2}(S_{j_Ee^{\prime \prime },e^{}}S_{e^{\prime \prime },j_Ee^{}})\frac{\sqrt{1}}{2}(S_{j_Ee^{\prime \prime },\sqrt{1}e^{}}S_{e^{\prime \prime },j_E\sqrt{1}e^{}}),$$
where $`e^{\prime \prime }=j_Ee`$. Due to the reality condition the restriction of the Lie bracket $`[,]:^2\text{m}^{\mathrm{}}\text{h}^{\mathrm{}}`$ to $`^2\text{m}`$ has values in h and $`\text{g}=\text{h}+\text{m}`$ is a symmetric decomposition with $`[\text{m},\text{m}]=\text{h}`$. The metric $`g^{\mathrm{}}=\omega _H\omega _E`$ defines a real valued scalar product $`g`$ of some signature $`(p,q)`$ on $`\text{m}=(HE)^\rho `$, which is invariant under the Lie algebra h. Since $`[\text{h},j_E]=0`$ the holonomy algebra $`\text{h}\mathrm{sp}(k,l)`$, $`p=4k`$, $`q=4l`$. Hence this symmetric decomposition defines a hyper-Kähler symmetric space.
By Lemma 2 any hyper-Kähler symmetric space can be obtained by this construction. It is well known that a simply connected symmetric space $`M`$ of signature $`(p,q)`$ is determined by its abstract curvature tensor $`RS^2(^2V)`$, $`V=\mathrm{}^{p,q}`$, and two tensors $`R`$ and $`R^{}`$ define isometric symmetric spaces if and only if they belong to the same $`\mathrm{O}(V)`$ orbit. Similarly a simply connected hyper-Kähler symmetric space is determined up to isometry by its abstract curvature tensor $`RS^2(^2V)`$, where $`V=\mathrm{}^{4k,4l}`$ is the pseudo-Euclidean vector space with fixed hypercomplex structure $`J_\alpha O(V)`$. For a hyper-Kähler symmetric space the complexified curvature tensor has the form
$$R(he,h^{}e^{})=\omega _H(h,h^{})S_{e,e^{}}.$$
where $`SS^4E`$ is the quartic of Lemma 2. Two such curvature tensors define isomorphic hyper-Kähler symmetric spaces if and only if they belong to the same orbit of $`\mathrm{Aut}(\mathrm{}^{4k,4l},J_\alpha )=\mathrm{Sp}(k,l)`$. The group $`\mathrm{Sp}(k,l)`$ acts on $`V^{\mathrm{}}=HE`$ as $`\mathrm{Id}\mathrm{Sp}(E)^{j_E}=\mathrm{Id}\mathrm{Aut}(E,\omega _E,j_E)`$. Hence two curvature tensors $`R=\omega _HS`$ and $`R^{}=\omega _HS^{}`$ are in the same $`\mathrm{Sp}(k,l)`$-orbit if and only if $`S`$ and $`S^{}`$ are in the same $`\mathrm{Sp}(k,l)`$-orbit on $`S^4E`$.
## 4 Complex hyper-Kähler symmetric spaces
### 4.1 Complex hyper-Kähler manifolds
A complex Riemannian manifold is a complex manifold $`M`$ equipped with a complex metric $`g`$, i.e. a holomorphic section $`g\mathrm{\Gamma }(S^2T^{}M)`$ which defines a nondegenerate complex quadratic form. As in the real case any such manifold has a unique holomorphic torsionfree and metric connection (Levi-Cività connection). A complex hyper-Kähler manifold is a complex Riemannian manifold $`(M^{4n},g)`$ of complex dimension $`4n`$ together with a compatible hypercomplex structure, i.e. three $`g`$-orthogonal parallel complex linear endomorphisms $`(J_1,J_2,J_3=J_1J_2)`$ with $`J_\alpha ^2=1`$. This means that the holonomy group $`\mathrm{Hol}\mathrm{Sp}(n,\mathrm{})=Z_{O(4n,\mathrm{})}(\mathrm{Sp}(1,\mathrm{}))`$. The linear group $`\mathrm{Sp}(n,\mathrm{})`$ is diagonally embedded into $`\mathrm{Sp}(n,\mathrm{})\times \mathrm{Sp}(n,\mathrm{})\mathrm{GL}(4n,\mathrm{})`$. Two complex hyper-Kähler manifolds $`(M,g,J_\alpha )`$ ($`\alpha =1,2,3`$) and $`(M^{},g^{},J_\alpha ^{})`$ are called isomorphic if there exists a holomorphic isometry $`\phi :MM^{}`$ such that $`\phi ^{}J_\alpha ^{}=J_\alpha `$ and $`\phi ^{}g^{}=g`$.
We will show that the complex hyper-Kähler structure can be described as a half-flat Grassmann structure of a certain type. A Grassmann structure on a complex Riemannian manifold $`(M,g)`$ is a decomposition of the (holomorphic) tangent bundle $`TMHE`$ into the tensor product of two holomorphic vector bundles $`H`$ and $`E`$ of rank $`2m`$ and $`2n`$ with holomorphic nondegenerate 2-forms $`\omega _H`$ and $`\omega _E`$ such that $`g=\omega _H\omega _E`$. The Grassmann structure will be called parallel if the Levi-Cività connection $`=^{TM}`$ can be decomposed as:
$$=^H\mathrm{Id}+\mathrm{Id}^E,$$
where $`^H`$ and $`^E`$ are (uniquely defined) symplectic connections in the bundles $`H`$ and $`E`$. A parallel Grassmann structure will be called half-flat if $`^H`$ is flat. Note that a parallel Grassmann structure on a simply connected manifold is half-flat if and only if the holonomy group of the Levi-Cività connection is contained in $`\mathrm{Id}\mathrm{Sp}(n,\mathrm{})\mathrm{Sp}(m,\mathrm{})\mathrm{Sp}(n,\mathrm{})\mathrm{O}(\mathrm{}^{2m}\mathrm{}^{2n})`$.
###### Proposition 5
A complex hyper-Kähler structure $`(g,J_\alpha )`$ on a simply connected complex manifold $`M`$ is equivalent to the following geometric data:
1. a half-flat Grassmann structure $`(TM,g,)(H,\omega _H,^H)(E,\omega _E,^E)`$ and
2. an isomorphism of flat symplectic vector bundles $`HM\times \mathrm{}^2`$. Under this isomorphism $`\omega _H=h_1^{}h_2^{}`$, where $`(h_1,h_2)`$ is the standard basis of $`\mathrm{}^2`$ considered as parallel frame of the trivial bundle $`H=M\times \mathrm{}^2`$.
More precisely,
$$J_1=R_i\mathrm{Id},J_2=R_j\mathrm{Id},\text{and}J_3=R_k\mathrm{Id},$$
(4.1)
where we have identified $`\mathrm{}^2=\mathrm{}h_1\mathrm{}h_2`$ with $`\mathrm{}=\mathrm{span}_{\mathrm{}}\{1,i,j,k\}=\mathrm{span}_{\mathrm{}}\{1,j\}=\mathrm{}1\mathrm{}j`$ with the complex structure defined by left-multiplication by $`i`$ and $`R_x`$ denotes the right-multiplication by the quaternion $`x\mathrm{}`$.
Proof: It is easy to check that the geometric data (i) and (ii) define a complex hyper-Kähler structure on $`M`$. Conversely let $`(g,J_\alpha )`$ be a complex hyper-Kähler structure on $`M`$. The endomorphism $`J_1`$ has eigenvalues $`\pm i`$ and the tangent space can be decomposed into a sum of eigenspaces
$$TM=E_+E_{}.$$
From the $`J_1`$-invariance of the metric $`g`$ it follows that $`g(E_\pm ,E_\pm )=0`$ and we can identify $`E_{}=E^{}`$ with the dual space of $`E=E_+`$. Since $`J_2`$ anticommutes with $`J_1`$ it interchanges $`E`$ and $`E^{}`$ and hence defines an isomorphism $`E\stackrel{}{}E^{}`$. Now $`g(,J_2)`$ defines a symplectic form $`\omega _E`$ on $`E`$. Let $`H=M\times \mathrm{}^2=M\times (\mathrm{}h_1\mathrm{}h_2)`$ be the trivial bundle with 2-form $`\omega _H=h_1^{}h_2^{}`$. Then we can identify
$$TM=EE^{}=EE=h_1Eh_2E=HE.$$
We check that under this identification we have $`g=\omega _H\omega _E`$. Note that both sides vanish on $`h_1E`$ and $`h_2E`$ and $`\omega _H(h_1,h_2)=1`$. We calculate for $`e,e^{}E=E_+=h_1E`$:
$$g(e,J_2e^{})=\omega _E(e,e^{})=\omega _H(h_1,h_2)\omega _E(e,e^{})=(\omega _H\omega _E)(h_1e,h_2e^{}).$$
Hence we have a Grassmann structure. The eigenspaces $`E_\pm `$ of the parallel endomorphism $`J_1`$ are invariant under parallel transport. Therefore the Levi-Cività connection $``$ induces a connection $`^E`$ in the bundle $`E`$. Since $`g=0`$ and $`J_2=0`$ we have $`^E\omega _E=0`$. We define a flat connection $`^H`$ on the trivial bundle $`H=M\times \mathrm{}^2`$ by the condition $`^Hh_1=^Hh_2=0`$. Then $`=^H\mathrm{Id}+\mathrm{Id}^E`$. So the Grassmann structure is half-flat.
Finally, using the standard identification $`\mathrm{}^2=\mathrm{}`$, one can easily check that the $`J_\alpha `$ are given by (4.1).
### 4.2 Complexification of real hyper-Kähler manifolds
Let $`(M,g,J_\alpha )`$ be a (real) hyper-Kähler manifold. We will assume that it is real analytic. This is automatically true if the metric $`g`$ is positive definite since it is Ricci-flat and a fortiori Einstein. Using analytic continuation we can extend $`(M,g,J_\alpha )`$ to a complex hyper-Kähler manifold $`(M^{\mathrm{}},g^{\mathrm{}},J_\alpha ^{\mathrm{}})`$ equipped with an antiholomorphic involution $`T`$. In complex local coordinates $`z^j=x^j+iy^j`$ which are extension of real analytic coordinates $`x^j,y^j`$ the involution is given by the complex conjugation $`z^j\overline{z}^j=x^jiy^j`$. We can reconstruct the (real) hyper-Kähler manifold as the fixed point set of $`T`$. We will call $`(M,g,J_\alpha )`$ a real form of $`(M^{\mathrm{}},g^{\mathrm{}},J_\alpha ^{\mathrm{}})`$ and $`(M^{\mathrm{}},g^{\mathrm{}},J_\alpha ^{\mathrm{}})`$ the complexification of $`(M,g,J_\alpha )`$.
In general a complex hyper-Kähler manifold has no real form. A necessary condition is that the holonomy group of $`^E`$ is contained in $`\mathrm{Sp}(k,l)`$, $`n=k+l`$, and hence preserves a quaternionic structure. Then we can define a parallel antilinear endomorphism field $`j_E:EE`$ such that $`j_E^2=1`$ and $`\omega _E(j_Ex,j_Ey)=\overline{\omega _E(x,y)}`$ for all $`x,yE`$, where the bar denotes complex conjugation. We define a parallel antilinear endomorphism field $`j_H:HH`$ as the left-multiplication by the quaternion $`j`$ on $`H=M\times \mathrm{}`$. Then $`\rho =j_Hj_E`$ defines a field of real structures in $`TM=HE`$. We denote by $`𝒟TM`$ the real eigenspace distribution of $`\rho `$ with eigenvalue $`1`$. Here $`TM`$ is considered as real tangent bundle of the real manifold $`M`$. If $`M^\rho M`$ is a leaf of $`𝒟`$ of real dimension $`4n`$ then the data $`(g,J_\alpha )`$ induce on $`M^\rho `$ a (real) hyper-Kähler structure.
### 4.3 Complex hyper-Kähler symmetric spaces
A complex Riemannian symmetric space is a complex Riemannian manifold $`(M,g)`$ such that any point is an isolated fixed point of an isometric holomorphic involution. Like in the real case one can prove that it admits a transitive complex Lie group of holomorphic isometries and that any simply connected complex Riemannian symmetric $`M`$ is associated to a complex symmetric decomposition
$$\text{g}=\text{k}+\text{m},[\text{k},\text{k}]\text{k},[\text{k},\text{m}]\text{m},[\text{m},\text{m}]=\text{k}$$
(4.2)
of a complex Lie algebra g together with an $`\mathrm{ad}_\text{k}`$-invariant complex scalar product on m. More precisely $`M=G/K`$, where $`G`$ is the simply connected complex Lie group with the Lie algebra g and $`K`$ is the (closed) connected subgroup associated with k. The holonomy group of such manifold is $`H=\mathrm{Ad}_K|\text{m}`$. Any pseudo-Riemannian symmetric space $`M=G/K`$ associated with a symmetric decomposition $`\text{g}=\text{k}+\text{m}`$ has a canonical complexification $`M^{\mathrm{}}=G^{\mathrm{}}/K^{\mathrm{}}`$ defined by the complexification $`\text{g}^{\mathrm{}}=\text{k}^{\mathrm{}}+\text{m}^{\mathrm{}}`$ of the symmetric decomposition. Proposition 1 remains true for complex Riemannian symmetric spaces. Ignoring the reality condition we obtain the following complex version of Theorem 1.
###### Theorem 2
Let $`SS^4E`$, $`E=\mathrm{}^{2n}`$, be a quartic polynomial invariant under all endomorphisms $`S_{e,e^{}}S^2E=\mathrm{sp}(E)`$. Then it defines a complex hyper-Kähler symmetric space, which is associated with the following complex symmetric decomposition
$$\text{g}=\text{h}+HE,\text{h}=S_{E,E}=\mathrm{span}\{S_{e,e^{}}|e,e^{}E\}\mathrm{sp}(E).$$
(4.3)
The bracket $`^2(HE)\text{h}`$ is given by (3.1). The complex hyper-Kähler symmetric space $`M`$ associated to this symmetric decomposition is the quotient $`M=M_S=G/K`$, where $`G`$ is the (complex) simply connected Lie group with Lie algebra g and $`KG`$ is the connected (and closed) subgroup with Lie algebra $`\text{k}=\text{h}`$.
Moreover any simply connected complex hyper-Kähler symmetric space can be obtained by this construction. Two complex hyper-Kähler symmetric spaces $`M_S`$ and $`M_S^{}`$ defined by quartics $`S`$ and $`S^{}`$ are isomorphic if and only if $`S`$ and $`S^{}`$ are in the same orbit of $`\mathrm{Aut}(E,\omega _E)=\mathrm{Sp}(E)\mathrm{Sp}(n,\mathrm{})`$.
###### Corollary 1
There is a natural bijection between simply connected complex hyper-Kähler symmetric spaces of dimension $`4n`$ up to isomorphism and $`\mathrm{Sp}(n,\mathrm{})`$-orbits on the space of quartic polynomials $`SS^4E`$ in the symplectic vector space $`E=\mathrm{}^{2n}`$ such that
$$S_{e,e^{}}S=0\text{for all}e,e^{}E.$$
(4.4)
### 4.4 Classification of complex hyper-Kähler symmetric spaces
The following complex version of Proposition 4 (with similar proof) will be a crucial step in the classification of complex hyper-Kähler symmetric spaces.
###### Proposition 6
Let $`(M=G/K,g,J_\alpha )`$ be a simply connected complex hyper-Kähler symmetric space. Then the holonomy group of $`M`$ is solvable and $`M`$ admits a transitive solvable Lie group of automorphisms.
Due to Corollary 1 the classification of simply connected complex hyper-Kähler symmetric spaces reduces to the determination of quartic polynomials $`S`$ satisfying (4.4). Below we will determine all such polynomials. We will prove that the following example gives all such polynomials.
Example 1: Let $`E=E_+E_{}`$ be a Lagrangian decomposition, i.e. $`\omega (E_\pm ,E_\pm )=0`$, of the symplectic vector space $`E=\mathrm{}^{2n}`$. Then any polynomial $`SS^4E_+S^4E`$ satisfies the condition (4.4) and defines a simply connected complex hyper-Kähler symmetric space $`M_S`$ with Abelian holonomy algebra $`\text{h}=S_{E_+,E_+}S^2E_+S^2E=\mathrm{sp}(E)`$.
In fact, since $`E_+`$ is Lagrangian the endomorphisms from $`S^2E_+`$ form an Abelian subalgebra of $`\mathrm{sp}(E)`$, which acts trivially on $`E_+`$ and hence on $`S^4E_+`$.
###### Theorem 3
Let $`SS^4E`$ be a quartic polynomial satisfying (4.4). Then there exists a Lagrangian decomposition $`E=E_+E_{}`$ such that $`SS^4E_+`$.
Proof: According to Theorem 2 the quartic $`S`$ defines a hyper-Kähler symmetric space with holonomy Lie algebra $`\text{h}=S_{E,E}`$. Since, by Proposition 6, h is solvable, Lie’s theorem implies the existence of a one-dimensional h-invariant subspace $`P=\mathrm{}pE`$. There exists an $`\omega `$-nondegenerate subspace $`WE`$ such that the $`\omega `$-orthogonal complement of $`P`$ is $`P^{}=PW`$. We choose a vector $`qE`$ such that $`\omega (p,q)=1`$ and $`\omega (W,q)=0`$ and put $`Q:=\mathrm{}q`$. Then we have
$$E=PWQ.$$
Since h preserves $`P`$ we have the following inclusion
$$\text{h}PE+W^2=P^2+PW+PQ+W^2,$$
where we use the notation $`XY=XY`$ for the symmetric product of subspaces $`X,YE`$. Then the second prolongation $`\text{h}^{(2)}=\{TS^4E|T_{e,e^{}}\text{h}`$ for all $`e,e^{}E\}`$ has the following inclusion
$$\text{h}^{(2)}P^3E+P^2W^2+PW^3+W^4=P^4+P^3Q+P^3W+P^2W^2+PW^3+W^4.$$
(4.5)
Indeed $`\text{h}^{(2)}\text{h}^2=P^4+P^3Q+P^3W+P^2Q^2+P^2WQ+P^2W^2+PQW^2+PW^3+W^4`$. The projection $`\text{h}^{(2)}P^2Q^2+P^2WQ+PQW^2`$ is zero because otherwise $`S_{q,q}\text{h}PE+W^2`$ would have a nonzero projection to $`Q^2+WQ`$ or $`S_{w,q}\text{h}`$ would have a nonzero projection to $`QW`$ for appropriate choice of $`wW`$. By (4.5) we can write the quartic $`S`$ as
$$S=p^3(\lambda p+\mu q+w_0)+p^2B+pC+D,$$
where $`\lambda ,\mu \mathrm{}`$, $`w_0W`$, $`BS^2W`$, $`CS^3W`$ and $`DS^4W`$. From now on we will identify $`S^dE`$ with the space $`\mathrm{}[E^{}]^{(d)}`$ of homogeneous polynomials on $`E^{}`$ of degree $`d`$. Then the $`\omega `$-contraction $`T_x=\iota _{\omega x}T=T(\omega x,\mathrm{})`$ of a tensor $`TS^dE`$ with a vector $`xE`$ is identified with the following homogeneous polynomial of degree $`d1`$:
$$T_x=\frac{1}{d}_{\omega x}T,$$
where $`_{\omega x}T`$ is the derivative of the polynomial $`T\mathrm{}[E^{}]^{(d)}`$ in the direction of $`\omega x=\omega (x,)E^{}`$. For example $`p_q=p,\omega q=\omega (q,p)=_{\omega q}p=_p^{}p=1=q_p`$.
From $`S_{p,q}=\frac{1}{4}\mu p^2`$ and the condition $`S_{p,q}S=0`$ we obtain $`\mu =0`$, since $`p^2S=\mu p^4`$. This implies $`S_{p,}=0`$. Next we compute:
$`S_{q,q}`$ $`=`$ $`{\displaystyle \frac{1}{6}}(6\lambda p^2+3pw_0+B)`$
$`S_{q,w}`$ $`=`$ $`{\displaystyle \frac{1}{12}}(3p^2\omega (w_0,w)+2p_{\omega w}B+_{\omega w}C)`$
$`=`$ $`{\displaystyle \frac{1}{12}}(3p^2\omega (w_0,w)+4pB_w+3C_w)`$
$`S_{w,w^{}}`$ $`=`$ $`{\displaystyle \frac{1}{6}}(p^2B_{w,w^{}}+3pC_{w,w^{}}+6D_{w,w^{}})`$
for any $`w,w^{}W`$.
Now the condition (4.4) can be written as follows:
$`0`$ $`=`$ $`6S_{q,q}S=(3pw_0+B)S=({\displaystyle \frac{3}{2}}(pw_0+w_0p)+B)S`$
$`=`$ $`{\displaystyle \frac{3}{2}}(2p^3B_{w_0}+3p^2C_{w_0}+4pD_{w_0})+p^3Bw_0+p^2BB+pBC+BD`$
$`=`$ $`2p^3Bw_0+{\displaystyle \frac{9}{2}}p^2C_{w_0}+p(6D_{w_0}+BC)+BD.`$
Note that $`Bw_0=B_{w_0}`$ and $`BB=[B,B]=0`$.
$`0`$ $`=`$ $`12S_{q,w}S=(4pB_w+3C_w)S`$
$`=`$ $`2(p^4\omega (B_w,w_0)+2p^3B^2w3p^2C_{Bw}4pD_{Bw})`$
$`+3(p^3C_ww_0+p^2C_wB+pC_wC+C_wD)`$
$`=`$ $`2p^4\omega (B_w,w_0)+p^3(4B^2w+3C_ww_0)+p^2(6C_{Bw}+3C_wB)`$
$`+p(8D_{Bw}+3C_wC)+3C_wD`$
$`0`$ $`=`$ $`2S_{w,w^{}}S={\displaystyle \frac{1}{2}}(pC_{w,w^{}}+C_{w,w^{}}p)S+2D_{w,w^{}}S`$
$`=`$ $`{\displaystyle \frac{1}{2}}(p^4\omega (C_{w,w^{}},w_0)2p^3BC_{w,w^{}}+3p^2C_{C_{w,w^{}}}+4pD_{C_{w,w^{}}})+`$
$`2(p^3D_{w,w^{}}w_0+p^2D_{w,w^{}}B+pD_{w,w^{}}C+D_{w,w^{}}D)`$
$`=`$ $`{\displaystyle \frac{1}{2}}p^4\omega (C_{w,w^{}},w_0)+p^3(BC_{w,w^{}}+2D_{w,w^{}}w_0)+`$
$`p^2({\displaystyle \frac{3}{2}}C_{C_{w,w^{}}}+2D_{w,w^{}}B)+p(2D_{C_{w,w^{}}}+2D_{w,w^{}}C)+2D_{w,w^{}}D`$
This gives the following system of equations:
$`(1)`$ $`Bw_0=0`$
$`(2)`$ $`C_{w_0}=0`$
$`(3)`$ $`6D_{w_0}+BC=0`$
$`(4)`$ $`BD=0`$
$`(5)`$ $`\omega (B_w,w_0)=0`$
$`(6)`$ $`4B^2w+3C_ww_0=0`$
$`(7)`$ $`2C_{Bw}+C_wB=0`$
$`(8)`$ $`8D_{Bw}+3C_wC=0`$
$`(9)`$ $`C_wD=0`$
$`(10)`$ $`\omega (C_ww^{},w_0)=0`$
$`(11)`$ $`BC_ww^{}+2D_{w,w^{}}w_0=0`$
$`(12)`$ $`{\displaystyle \frac{3}{2}}C_{C_ww^{}}+2D_{w,w^{}}B=0`$
$`(13)`$ $`D_{C_ww^{}}+D_{w,w^{}}C=0`$
$`(14)`$ $`D_{w,w^{}}D=0`$
Note that (5) and (10) follow from (1) and (2) and that using (2) equation (6) says that the endomorphism $`B`$ has zero square:
$$(6^{})B^2=0.$$
Eliminating $`D_{w_0}`$ in equations (3) and (11) we obtain:
$$(15)0=(BC)_ww^{}+3BC_ww^{}=BC_ww^{}C_{Bw}w^{}C_wBw^{}+3BC_ww^{}=4BC_ww^{}C_{Bw}w^{}C_wBw^{}.$$
We can rewrite (7) as:
$$(7^{})2C_{Bw}w^{}+C_wBw^{}BC_ww^{}=0.$$
Eliminating $`C_{Bw}w^{}`$ in ($`7^{}`$) and (15) we obtain:
$$(16)3BC_ww^{}+C_wBw^{}=0.$$
Since the first summand is symmetric in $`w`$ and $`w^{}`$ we get
$$(17)C_wBw^{}=C_w^{}Bw=C_{Bw}w^{}.$$
Now using (17) we can rewrite (15) as:
$$(15^{})2BC_ww^{}C_wBw^{}=0.$$
The equations ($`15^{}`$) and (16) show that $`BC_ww^{}=C_wBw^{}=C_{Bw}w^{}=0`$ and hence also $`BC=0`$. This implies $`D_{w_0}=0`$, by (3). Now we can rewrite (1-14) as:
$$Bw_0=C_{w_0}=D_{w_0}=0$$
(4.6)
$$BC=BD=0$$
(4.7)
$$B^2=0$$
(4.8)
$$C_{Bw}=C_wB=BC_w=0$$
(4.9)
$$8D_{Bw}+3C_wC=0$$
(4.10)
$$C_wD=0$$
(4.11)
$$\frac{3}{2}C_{C_ww^{}}+2D_{w,w^{}}B=0$$
(4.12)
$$D_{C_ww^{}}+D_{w,w^{}}C=0$$
(4.13)
$$D_{w,w^{}}D=0$$
(4.14)
Now to proceed further we decompose $`K:=\text{ker}`$ $`B=W_0W^{}`$, where $`W_0=\text{ker}`$ $`\omega |K`$ and $`W^{}`$ is a (nondegenerate) complement. Let us denote by $`W_1`$ a complement to $`K`$ in $`W`$ such that $`\omega (W^{},W_1)=0`$. Then $`W_0+W_1`$ is the $`\omega `$-orthogonal complement to the $`B`$-invariant nondegenerate subspace $`W^{}`$. This shows that $`BW_1(W_0+W_1)K=W_0`$. Moreover since $`W_1K=0`$ the map $`B:W_1W_0`$ is injective and hence $`dimW_1dimW_0`$. On the other hand $`dimW_1dimW_0`$, since $`W_0`$ is an isotropic subspace of the symplectic vector space $`W_0+W_1`$. This shows that $`B:W_1W_0`$ is an isomorphism.
###### Lemma 3
$`CS^3K`$ and $`DS^4K`$.
Proof: Since $`W_0=BW`$ the equation (4.9) shows that $`C_{W_0}=0`$, which proves the first statement. From (4.10) and the identity
$$(C_xC)_y=[C_x,C_y]C_{C_xy}$$
(4.15)
we obtain
$$D_{Bx,y}+D_{By,x}=\frac{3}{8}((C_xC)_y+(C_yC)_x)=\frac{3}{4}C_{C_xy}.$$
(4.16)
The equation $`BD=0`$ (4.7) reads:
$$0=(BD)_{x,y}=[B,D_{x,y}]D_{Bx,y}D_{x,By}.$$
Using this (4.12) yields:
$$D_{Bx,y}+D_{By,x}=[B,D_{x,y}]=D_{x,y}B=\frac{3}{4}C_{C_xy}.$$
(4.17)
Now from (4.16) and (4.17) we obtain that
$$0=C_{C_xy}z=C_zC_xy$$
for all $`x,y,zW`$. This implies $`[C_x,C_y]=0`$ for all $`x,yW`$ and hence
$$C_xC=0$$
(4.18)
for all $`xW`$, by (4.15). Finally this shows that $`D_{W_0}=0`$ by (4.10). This proves the second statement.
###### Lemma 4
$`D_{x,y}C_z=C_zD_{x,y}=0`$ for all $`x,y,zW`$.
Proof: Using (4.13) we compute:
$$D_{x,y}C_zw=D_{C_zw,x}y=(D_{z,w}C)_xy=([D_{z,w},C_x]yC_{D_{z,w}x}y).$$
(4.19)
From (4.11) we get:
$$0=(C_xD)_{z,w}y=[C_x,D_{z,w}]yD_{C_xz,w}yD_{z,C_xw}y=C_xD_{z,w}yD_{z,w}C_xyD_{y,w}C_xzD_{z,y}C_xw,$$
and hence:
$$[D_{z,w},C_x]y=D_{y,w}C_xzD_{z,y}C_xw,$$
and
$$C_{D_{z,w}x}y=C_yD_{z,w}x=D_{z,w}C_xy+D_{x,w}C_yz+D_{z,x}C_yw.$$
Now we eliminate the $`CD`$-terms from (4.19) arriving at:
$$D_{x,y}C_zw=(D_{y,w}C_xz+D_{z,y}C_xw+D_{z,w}C_xy+D_{x,w}C_yz+D_{z,x}C_yw).$$
(4.20)
Considering all the permutations of $`(x,y,z,w)`$ we get 6 homogeneous linear equations for the 6 terms of equation (4.20) with the matrix:
$$\left(\begin{array}{cccccc}1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1\end{array}\right)$$
This is the matrix of the endomorphism $`2\text{Id}+ee`$ in the arithmetic space $`\mathrm{}^6`$, where $`e=e_1+\mathrm{}+e_6`$; $`(e_i)`$ the standard basis. It has eigenvalues $`(4,2,2,2,2,2)`$. This shows that the matrix is nondegenerate and proves the lemma.
For a symmetric tensor $`TS^dW`$ we denote by
$$\mathrm{\Sigma }_T:=\mathrm{span}\{T_{x_1,x_2,\mathrm{},x_{d2}}x_{d1}|x_1,x_2,\mathrm{}x_{d1}W\}W\}$$
the support of $`T`$.
###### Lemma 5
The supports of the tensors $`BS^2W`$, $`CS^3W`$ and $`DS^4W`$ admit the following inclusions
$$\mathrm{\Sigma }_B+\mathrm{\Sigma }_C\mathrm{ker}B\mathrm{ker}C\mathrm{ker}D,\mathrm{\Sigma }_D\mathrm{ker}B\mathrm{ker}C.$$
Moreover $`\mathrm{\Sigma }_B+\mathrm{\Sigma }_C`$ is isotropic and $`\omega (\mathrm{\Sigma }_D,\mathrm{\Sigma }_B+\mathrm{\Sigma }_C)=0`$.
Proof: The first statement follows from $`B^2=BC_x=BD_{x,y}=C_xB=C_xC_y=C_xD_{y,z}=D_{x,y}B=D_{x,y}C_z=0`$ for all $`x,y,zW`$. The second statement follows from the first and the definition of support, e.g. if $`z=C_xy\mathrm{\Sigma }_C`$ and $`w\mathrm{\Sigma }_B+\mathrm{\Sigma }_C+\mathrm{\Sigma }_D\mathrm{ker}C`$ we compute:
$$\omega (z,w)=\omega (C_xy,w)=\omega (y,C_xw)=0.$$
###### Lemma 6
The Lie algebra $`D_{W,W}S^2W\mathrm{sp}(W)`$ is solvable.
Proof: This follows from Proposition 6, since $`DS^4W`$ satisfies (4.4) and hence defines a complex hyper-Kähler symmetric space with holonomy Lie algebra $`D_{W,W}`$. It also follows from the solvability of $`S_{E,E}`$ as we show now. In terms of the decomposition $`E=P+W+Q`$ an endomorphism
$$S_{x,y}=(\lambda p^4+p^3w_0+p^2B+pC+D)_{x,y}=(p^2B+pC+D)_{x,y}=B(x,y)p^2pC_xy+D_{x,y},$$
where $`x,yW`$, is represented by
$$\left(\begin{array}{ccc}0& (C_xy)^t& B(x,y)\\ 0& D_{x,y}& C_xy\\ 0& 0& 0\end{array}\right).$$
Since the Lie algebra $`S_{E,E}`$ is solvable this implies that the Lie algebra $`D_{W,W}`$, which corresponds to the induced representation of $`S_{E,E}`$ on $`P^{}/PW`$, is also solvable.
###### Lemma 7
$$\omega (w_0,\mathrm{\Sigma }_B+\mathrm{\Sigma }_C+\mathrm{\Sigma }_D)=0.$$
Proof: Note that $`w_0\mathrm{ker}B\mathrm{ker}C\mathrm{ker}D`$, due to equations (1-3) and (4.7). This implies the lemma. In fact, if e.g. $`y=Bx\mathrm{\Sigma }_B`$ then
$$\omega (w_0,y)=\omega (w_0,Bx)=\omega (Bw_0,x)=0,$$
which shows that $`\omega (w_0,\mathrm{\Sigma }_B)=0`$
Now to finish the proof of Theorem 3 we will use induction on the dimension $`dimE=2n`$. If $`n=1`$ the (solvable) holonomy algebra h is a proper subalgebra of $`S^2E\mathrm{sl}(2,\mathrm{})`$. Without loss of generality we may assume that either
a) $`\text{h}=\mathrm{}p^2`$ or
b) $`\text{h}=\mathrm{}pq`$ or
c) $`\text{h}=\mathrm{}p^2+\mathrm{}pq`$,
where $`(p,q)`$ is a symplectic basis of $`E`$. In the all three cases the Lie algebra $`S_{E,E}=\text{h}\mathrm{}p^2+\mathrm{}pq`$ and hence
$$S=\lambda p^4+\mu p^3q.$$
In the cases b) and c) we have that $`pq\text{h}`$ and since
$$pqS=\frac{1}{2}(pq+qp)S=2\lambda p^4\mu p^3q=0$$
it follows that $`S=0`$. In the case a) from $`S_{E,E}=\text{h}=\mathrm{}p^2`$ we have that $`S=\lambda p^4`$. This tensor is invariant under $`\text{h}=\mathrm{}p^2`$ and belongs to the fourth symmetric power of the Lagrangian subspace $`\mathrm{}pE`$. This establishes the first step of the induction. Now by induction using equation (4.14) and Lemma 6 we may assume that $`\mathrm{\Sigma }_D`$ is isotropic. Now Lemma 5 and Lemma 7 show that $`\mathrm{}w_0+\mathrm{\Sigma }_B+\mathrm{\Sigma }_C+\mathrm{\Sigma }_D`$ is isotropic and hence is contained in some Lagrangian subspace $`E_+E`$. This implies that $`SS^4E_+`$.
Now we give a necessary and sufficient condition for a symmetric manifold $`M=M_S`$, $`SS^4E_+`$, to have no flat de Rham factor.
###### Proposition 7
The complex hyper-Kähler symmetric space $`M_S`$, $`SS^4E_+`$, has no flat de Rham factor if and only if the support $`\mathrm{\Sigma }_S=E_+`$.
Proof: If $`M=M_S=G/K`$ has a flat factor $`M_0`$, such that $`M=M_1\times M_0`$, then this induces a decomposition $`E=E^1E^0`$ and $`SS^4E_1`$; hence $`\mathrm{\Sigma }_SE_1E_+E_+`$. Conversely let $`SS^4E_+`$, assume that $`E_+^1=\mathrm{\Sigma }_SE_+`$ is a proper subspace and choose a complementary subspace $`E_+^0`$. We denote by $`E_{}^1`$ and $`E_{}^0`$ the annihilator of $`\omega E_+^0`$ and $`\omega E_+^1`$ respectively. Let us denote $`E^1=E_+^1E_{}^1`$, $`E^0=E_+^0E_{}^0`$, $`\text{m}^1=HE^1`$ and $`\text{m}^0=HE^0`$. Then $`E^0,E^1E`$ are $`\omega `$-nondegenerate complementary subspaces and $`\text{m}^0,\text{m}^1\text{m}=T_oM`$ are $`g`$-nondegenerate complementary subspaces. Since $`SS^4E_+^1`$ the Lie algebra $`\text{g}=\text{h}+\text{m}=(\text{h}+\text{m}^1)\text{m}^0`$ has the Abelian direct summand $`\text{m}^0`$, see (3.1), which gives rise to a flat factor $`M^0M_S=M^1\times M^0`$.
###### Theorem 4
Any simply connected complex hyper-Kähler symmetric space without flat de Rham factor is isomorphic to a complex hyper-Kähler symmetric space of the form $`M_S`$, where $`SS^4E_+`$ and $`E_+E`$ is a Lagrangian subspace of the complex symplectic vector space $`E=\mathrm{}^{2n}`$. Moreover there is a natural 1-1 correspondence between simply connected complex hyper-Kähler symmetric spaces without flat factor up to isomorphism and orbits $`𝒪`$ of the group $`\mathrm{Aut}(E,\omega ,E_+)=\{ASp(E)|AE_+=E_+\}\mathrm{GL}(E_+)\mathrm{GL}(n,\mathrm{})`$ on the space $`S^4E_+`$ such that $`\mathrm{\Sigma }_S=E_+`$ for all $`S𝒪`$.
Proof: This is a corollary of Theorem 2, Theorem 3 and Proposition 7.
Let $`M=G/K`$ be a simply connected complex hyper-Kähler symmetric space without flat factor. By Theorem 3 and Proposition 7 it is associated to quartic polynomial $`SS^4E_+`$ with support $`\mathrm{\Sigma }_S=E_+`$. Now we describe the Lie algebra $`\mathrm{aut}(M_S)`$ of the full group of automorphisms, i.e. isometries which preserve the hypercomplex structure, of $`M_S`$.
###### Theorem 5
Let $`M_S=G/K`$ be as above. Then the full automorphism algebra is given by
$$\mathrm{aut}(M_S)=\mathrm{aut}(S)+\text{g},$$
where $`A\mathrm{aut}(S)=\{B\mathrm{gl}(E_+)|BS=0\}`$ acts on $`\text{g}=\text{h}+\text{m}`$ as follows. It preserves the decomposition and acts on $`\text{h}=S_{E,E}`$ by
$$[A,S_{x,y}]=S_{Ax,y}+S_{x,Ay}$$
for all $`x,yE`$ and on $`\text{m}=HE`$ by
$$[A,he]=hAe,$$
where $`\mathrm{gl}(E_+)`$ is canonically embedded into $`\mathrm{sp}(E)`$.
Proof: By (the complex version of) Proposition 1 it is sufficient to determine the centralizer c of $`\mathrm{sp}(1,\mathrm{})`$ in the full isotropy algebra $`\stackrel{~}{\text{h}}=\mathrm{aut}(R)\mathrm{sp}(1,\mathrm{})\text{h}`$. Equation (2.2) shows that
$$\text{c}=\{\mathrm{Id}A|A\mathrm{sp}(E),AS=[A,S(,)]S(A,)S(,A)=0\}.$$
From $`AS=0`$ we obtain that the commutator $`[A,S_{x,y}]=S_{Ax,y}+S_{x,Ay}`$ for all $`x,yE`$ and $`A\mathrm{\Sigma }_S=AE_+E_+`$. This implies $`\text{c}=\mathrm{aut}(S)`$.
## 5 Classification of hyper-Kähler symmetric spaces
Using the description of complex hyper-Kähler symmetric spaces given in Theorem 4 we will now classify (real) hyper-Kähler symmetric spaces. Recall that a simply connected pseudo-Riemannian manifold is called indecomposable if it is not a Riemannian product of two pseudo-Riemannian manifolds. Any simply connected pseudo-Riemannian manifold can be decomposed into the Riemannian product of indecomposable pseudo-Riemannian manifolds. By Wu’s theorem \[W\] a simply connected pseudo-Riemannian manifold is indecomposable if and only if its holonomy group is weakly irreducible, i.e. has no invariant proper nondegenerate subspaces. Therefore it is sufficient to classify (real) hyper-Kähler symmetric spaces with indecomposable holonomy.
Let $`(M=G/K,g,J_\alpha )`$ be a hyper-Kähler symmetric space associated to a symmetric decomposition (2.1). The complexified tangent space of $`M`$ is identified with $`\text{m}^{\mathrm{}}=HE`$, the tensor product of to complex symplectic vector spaces with quaternionic structure $`j_H`$ and $`j_E`$ such that $`\rho =j_Hj_E`$ is the complex conjugation of $`\text{m}^{\mathrm{}}`$ with respect to m. By Theorem 1 it is defined by a quartic polynomial $`SS^4E`$ satisfying the conditions of the theorem. Moreover the holonomy algebra h acts trivially on $`H`$ and is identified with the real form of the complex Lie algebra $`S_{E,E}\mathrm{sp}(E)`$ given by $`\text{h}=\mathrm{span}\{S_{je,e^{}}S_{e,je^{}}|e,e^{}E\}=\{AS_{E,E}|[A,j_E]=0\}\mathrm{sp}(E)^{j_E}`$.
The quartic polynomial $`S`$ defines also a complex hyper-Kähler symmetric space $`M^{\mathrm{}}=G^{\mathrm{}}/K^{\mathrm{}}`$, which is the complexification of $`M=G/K`$. By Theorem 3, $`SS^4L`$ for some Lagrangian subspace $`LE`$. Recall that the symplectic form $`\omega =\omega _E`$ together with the quaternionic structure $`j=j_E`$ define a Hermitian metric $`\gamma =\gamma _E=\omega _E(,j_E)`$ of (real) signature $`(4k,4l)`$, $`n=k+l`$, which coincides with the signature of the pseudo-Riemannian metric $`g`$ (we normalize $`\gamma _H=\omega _H(,j_H)`$ to be positive definite). We may decompose $`\gamma `$-orthogonally $`L=L^0L^+L^{}`$, such that $`\gamma `$ vanishes on $`L^0`$ is positive definite on $`L^+`$ and negative definite on $`L^{}`$.
###### Lemma 8
1. $`jL^0=L^0`$ and
2. $`L^++L^{}+jL^++jL^{}E`$ is an $`\omega `$-nondegenerate and $`\gamma `$-nondegenerate h-invariant subspace (with trivial action of h).
Proof: We show first that $`L+jL^0`$ is $`\omega `$-isotropic and hence $`L+jL^0=L`$ since $`L`$ is Lagrangian. Indeed $`LL^0`$ is $`\omega `$-isotropic and also $`jL^0`$ because $`\omega `$ is $`j`$-invariant. So it suffices to remark that $`\omega (L,jL^0)=0`$:
$$\omega (L,jL^0)=\gamma (L,L^0)=0.$$
This implies that $`jL^0L`$. Since $`\gamma (L,jL^0)=\omega (L,L^0)=0`$, we conclude that $`jL^0\mathrm{ker}\gamma |L=L^0`$. This proves (i).
To prove (ii) it is sufficient to check that the subspace $`L^++L^{}+jL^++jL^{}E`$ is nondegenerate with respect to $`\gamma `$, since it is $`j`$-invariant. First we remark that $`\gamma `$ is positive definite on $`L^+`$ and $`jL^+`$ and negative definite on $`L^{}`$ and $`jL^{}`$, due to the $`j`$-invariance of $`\gamma `$: $`\gamma (jx,jx)=\gamma (x,x)`$, $`xE`$. So to prove (ii) it is sufficient to check that $`jL^+jL^{}`$ is $`\gamma `$-orthogonal to the $`\gamma `$-nondegenerate vector space $`L^++L^{}`$:
$$\gamma (L^++L^{},jL^++jL^{})=\omega (L^++L^{},L^++L^{})=0.$$
By Theorem 1 the quartic polynomial $`S`$ must satisfy the reality condition $`[S_{je,e^{}}S_{e,je^{}},j]=0`$. Now we describe all such polynomials.
The quaternionic structure $`j`$ on $`E`$ is compatible with $`\omega `$, i.e. $`\omega (jx,jy)=\overline{\omega (x,y)}`$ for all $`x,yE`$ and it induces a real structure (i.e. an antilinear involution) $`\tau :=jj\mathrm{}j`$ on all even powers $`S^{2r}EEE\mathrm{}E`$. For $`SS^{2r}E`$ and $`x_1,\mathrm{},x_{2r}E`$ we have
$$(\tau S)(x_1,\mathrm{},x_{2r})=\overline{S(jx_1,\mathrm{},jx_{2r})}.$$
Note that the fixed point set $`\mathrm{sp}(E)^\tau =\{A\mathrm{sp}(E)|[A,j]=0\}\mathrm{sp}(k,l)`$.
###### Proposition 8
Let $`(E,\omega ,j)`$ be a complex symplectic vector space with a quaternionic structure $`j`$ such that $`\omega (jx,jy)=\overline{\omega (x,y)}`$ for all $`x,yE`$. Then a quartic polynomial $`SS^4E`$ satisfies the reality condition $`[S_{je,e^{}}S_{e,je^{}},j]=0`$ if and only if $`S(S^4E)^\tau =\mathrm{span}\{T+\tau T|TS^4E\}`$.
Proof: The reality condition for $`SS^4E`$ can be written as
$$[S_{jx,jy}+S_{x,y},j]z=0$$
for all $`x,y,zE`$. Contracting this vector equation with $`jwE`$ by means of $`\omega `$ and using the compatibility between $`j`$ and $`\omega `$ we obtain the equivalent condition
$`0`$ $`=`$ $`\omega (jw,[S_{jx,jy}+S_{x,y},j]z)`$ (5.1)
$`=`$ $`S(jx,jy,jz,jw)\overline{S(jx,jy,z,w)}+S(x,y,jz,jw)\overline{S(x,y,z,w)}.`$
Now putting $`x=y=z=w=u`$ we obtain:
$$0=S(ju,ju,ju,ju)\overline{S(ju,ju,u,u)}+S(u,u,ju,ju)\overline{S(u,u,u,u)}$$
and putting $`x=iu`$ and $`y=z=w=u`$ we obtain:
$$0=iS(ju,ju,ju,ju)i\overline{S(ju,ju,u,u)}+iS(u,u,ju,ju)+i\overline{S(u,u,u,u)}.$$
Comparing these two equations we get $`S(ju,ju,juj,u)=\overline{S(u,u,u,u)}`$, i.e. $`S=\tau S`$. This shows that the reality condition implies that $`S(S^4E)^\tau `$. Conversely the condition $`S=\tau S`$ can be written as
$$S(jx,jy,jz,jw)=\overline{S(x,y,z,w)}\text{for all}x,y,z,wE.$$
Changing $`zjz`$ and $`wjw`$ in this equation we obtain
$$S(jx,jy,z,w)=\overline{S(x,y,jz,jw)}\text{for all}x,y,z,wE.$$
These two equations imply (5.1) and hence the reality condition.
Now we are ready to classify simply connected hyper-Kähler symmetric spaces. We will show that the following construction gives all such symmetric spaces.
Let $`(E,\omega ,j)`$ be a complex symplectic vector space of dimension $`2n`$ with a quaternionic structure $`j`$ such that $`\omega (jx,jy)=\overline{\omega (x,y)}`$ for all $`x,yE`$ and $`E=E_+E_{}`$ a $`j`$-invariant Lagrangian decomposition. Such a decomposition exists if and only if the Hermitian form $`\gamma =\omega (,j)`$ has real signature $`(4m,4m)`$, where $`dim_{\mathrm{}}E=2n=4m`$. Then any polynomial $`S(S^4E_+)^\tau =S^4E_+(S^4E)^\tau `$ satisfies the condition (4.4) and the reality condition, by Proposition 8. Hence by Theorem 1 it defines a (real) simply connected hyper-Kähler symmetric space $`M_S`$ with Abelian holonomy algebra $`\text{h}=(S_{E_+,E_+})^\tau =S_{E_+,E_+}(S^2E)^\tau =\mathrm{span}\{S_{je,e^{}}S_{e,je^{}}|e,e^{}E\}\mathrm{sp}(E)^\tau \mathrm{sp}(m,m)`$.
###### Theorem 6
Any simply connected hyper-Kähler symmetric space without flat de Rham factor is isomorphic to a hyper-Kähler symmetric space of the form $`M_S`$, where $`S=T+\tau T`$, $`TS^4E_+`$ and $`E_+E`$ is a $`j`$-invariant Lagrangian subspace of the complex symplectic vector space $`E`$ with compatible quaternionic structure $`j`$. A hyper-Kähler symmetric space of the form $`M_S`$ has no flat factor if and only if it complexification has no flat factor, which happens if and only if the support $`\mathrm{\Sigma }_S=E_+`$. Moreover there is a natural 1-1 correspondence between simply connected hyper-Kähler symmetric spaces without flat factor up to isomorphism and orbits $`𝒪`$ of the group $`\mathrm{Aut}(E,\omega ,j,E_+)=\{ASp(E)|[A,j]=0,AE_+=E_+\}\mathrm{GL}(m,\mathrm{})`$ on the space $`(S^4E_+)^\tau `$ such that $`\mathrm{\Sigma }_S=E_+`$ for all $`S𝒪`$.
Proof: Let $`M`$ be a simply connected hyper-Kähler symmetric space. We first assume that it is indecomposable. Then the holonomy algebra h is weakly irreducible. By Theorem 1, $`M=M_S`$ for some quartic polynomial $`SS^4E`$ satisfying (4.4) and the reality condition (3.2). By Proposition 8 the reality condition means that $`S(S^4E)^\tau `$. On the other hand, by Theorem 3 $`SS^4L`$ for some Lagrangian subspace $`L`$ of $`E`$. Now the weak irreducibility of h and Lemma 8 imply that $`L=L^0`$ is $`j`$-invariant. This proves that $`S(S^4E_+)^\tau `$, where $`E_+=L=L^0`$ is a $`j`$-invariant Lagrangian subspace of $`E`$. This shows that $`M`$ is obtained from the above construction. Any simply connected hyper-Kähler symmetric space $`M`$ without flat factor is the Riemannian product of indecomposable ones, say $`M=M_1\times M_2\times \mathrm{}\times M_r`$, and we may assume that $`M_i=M_{S_i}`$, $`S_iS^4E_i`$. Therefore $`M`$ is associated to the quartic polynomial $`S=S_1S_2\mathrm{}S_rS^4E`$, $`E=E_1E_2\mathrm{}E_r`$. Moreover $`S`$ satisfies (4.4) and (3.2) if the $`S_i`$ satisfy (4.4) and (3.2). This shows that any simply connected hyper-Kähler symmetric space is obtained from the above construction.
It is clear that the complexification $`M_S^{\mathrm{}}`$ has a flat factor if $`M_S`$ has a flat factor. Conversely let us assume that $`M_S^{\mathrm{}}`$ has a flat factor, hence $`\mathrm{\Sigma }_SE_+`$ is a proper subspace. Since $`j\mathrm{\Sigma }_S=jS_{E,E}E=S_{E,E}jE=S_{E,E}E=\mathrm{\Sigma }_S`$ there exists a $`j`$-invariant complementary subspace $`E_+^{}`$ in $`E_+`$. Denote by $`E_{}^{}`$ the annihilator of $`\mathrm{\Sigma }_S`$ in $`E_{}`$ then $`E^{}=E_+^{}E_{}^{}`$ is an $`\omega `$-nondegenerate and $`j`$-invariant subspace of $`E`$ on which the holonomy $`\text{h}^{\mathrm{}}=S_{E,E}S^2\mathrm{\Sigma }_S`$ acts trivially. Then the corresponding real subspace $`(HE^{})^\rho \text{m}=(HE)^\rho `$ is a $`g`$-nondegenerate subspace on which the holonomy h acts trivially. By Wu’s theorem \[W\] it defines a flat de Rham factor.
Now the last statement follows from the corresponding statement in Theorem 1.
###### Corollary 2
Any hyper-Kähler symmetric space without flat factor has signature $`(4m,4m)`$. In particular its dimension is divisible by $`8`$.
###### Corollary 3
Let $`M=M_S`$ be a complex hyper-Kähler symmetric space without flat factor associated with a quartic $`SS^4E_+`$, where $`E_+E`$ is a Lagrangian subspace. It admits a real form if and only if there exists a quaternionic structure $`j`$ on $`E`$ compatible with $`\omega `$ preserving $`E_+`$ such that $`\tau S=S`$, where $`\tau `$ is the real structure on $`S^4E`$ induced by $`j`$. In particular $`dim_{\mathrm{}}M`$ has to be divisible by $`8`$.
Let $`M=G/K`$ be a simply connected hyper-Kähler symmetric space without flat factor. By Theorem 6 it is associated to a quartic polynomial $`S(S^4E_+)^\tau `$ with support $`\mathrm{\Sigma }_S=E_+`$. Now we describe the Lie algebra $`\mathrm{aut}(M_S)`$ of the full group of automorphisms, i.e. isometries which preserve the hypercomplex structure of $`M_S`$.
###### Theorem 7
Let $`M_S=G/K`$ be as above. Then the full automorphism algebra is given by
$$\mathrm{aut}(M_S)=\mathrm{aut}(S)+\text{g},$$
where $`\mathrm{aut}(S)=\{A\mathrm{gl}(E_+)|[A,j]=0,AS=0\}`$ acts on
$$\text{g}=\text{h}+\text{m},\text{h}=\{AS_{E,E}|[A,j]=0\}=\mathrm{span}\{S_{jx,y}S_{x,jy}|x,yE\},\text{m}=(HE)^\rho $$
as in Theorem 5.
Proof: The proof is similar to that of Theorem 5.
## 6 Low dimensional hyper-Kähler symmetric spaces
### 6.1 Complex hyper-Kähler symmetric spaces of dimension $`8`$
Dimension 4
Assume that $`M`$ is a simply connected complex hyper-Kähler symmetric space of dimension 4. Applying Theorem 4 we conclude that $`M=M_S`$ for some $`SS^4E_+`$, where $`E_+E`$ is a one-dimensional subspace $`E_+=\mathrm{}e`$. This proves:
###### Theorem 8
There exists up to isomorphism only one non-flat simply connected complex hyper-Kähler symmetric space of dimension 4: $`M=M_S`$ associated with the quartic $`S=e^4`$.
Dimension 8
Any eight-dimensional simply connected complex hyper-Kähler symmetric space is associated with a quartic $`SS^4E_+`$, where $`E_+E`$ is a Lagrangian subspace of $`E=\mathrm{}^4`$. We denote by $`(e,e^{})`$ a basis of $`E_+`$.
###### Theorem 9
Eight-dimensional simply connected complex hyper-Kähler symmetric space are in natural 1-1 correspondence with the orbits of the group $`\mathrm{CO}(3,\mathrm{})=\mathrm{}^{}\mathrm{SO}(3,\mathrm{})`$ on the space $`S_0^2\mathrm{}^3`$ of traceless symmetric matrices. The complex hyper-Kähler symmetric space associated with a traceless symmetric matrix $`A`$ is the manifold $`M_{S(A)}`$, where $`S(A)S^4\mathrm{}^2`$ is the quartic polynomial which corresponds to $`A`$ under the $`\mathrm{SO}(3,\mathrm{})`$-equivariant isomorphism $`S_0^2\mathrm{}^3`$ $`S_0^2^2\mathrm{}^3S_0^2S^2\mathrm{}^2=S^4\mathrm{}^2`$.
The classification of $`\mathrm{SO}(3,\mathrm{})`$-orbits on $`S_0^2\mathrm{}^3`$ was given by Petrov \[P\] in his classification of Weyl tensors of Lorentzian 4-manifolds.
Proof: By Theorem 4 the classification of eight-dimensional simply connected complex hyper-Kähler symmetric spaces reduces to the description of orbits of the group $`\mathrm{GL}(E_+)=\mathrm{GL}(2,\mathrm{})`$ on $`S^4\mathrm{}^2S^2S^2\mathrm{}^2`$. Fixing a volume form $`\sigma `$ on $`\mathrm{}^2`$ we can identify $`S^2\mathrm{}^2`$ with $`\mathrm{sp}(1,\mathrm{})\mathrm{so}(3,\mathrm{})`$. Then the Killing form $`B`$ is an $`\mathrm{SL}(2,\mathrm{})`$-invariant and we have the $`\mathrm{GL}(2,\mathrm{})`$-invariant decomposition: $`S^2S^2\mathrm{}^2=S_0^2S^2\mathrm{}^2\mathrm{}B`$. The action of $`\mathrm{SL}(2,\mathrm{})`$ on $`S^2\mathrm{}^2`$ is effectively equivalent to the adjoint action of $`\mathrm{SO}(3,\mathrm{})`$. The problem thus reduces essentially to the determination of the orbits of $`\mathrm{SO}(3,\mathrm{})`$ on $`S_0^2\mathrm{}^3`$.
### 6.2 Hyper-Kähler symmetric spaces of dimension $`8`$
By Corollary 3 the minimal dimension of non-flat hyper-Kähler symmetric spaces is 8.
###### Theorem 10
Eight-dimensional simply connected hyper-Kähler symmetric space are in natural 1-1 correspondence with the orbits of the group $`\mathrm{}^+\mathrm{SO}(3)`$ on the space $`S_0^2\mathrm{}^3`$ of traceless symmetric matrices. The hyper-Kähler symmetric space associated with a traceless symmetric matrix $`A`$ is the manifold $`M_{S(A)}`$, where $`S(A)(S^4\mathrm{}^2)^\tau `$ is the quartic polynomial which corresponds to $`A`$ under the $`\mathrm{SO}(3)`$-equivariant isomorphism $`S_0^2\mathrm{}^3(S^4\mathrm{}^2)^\tau `$.
|
warning/0007/hep-th0007149.html
|
ar5iv
|
text
|
# Finite Temperature Tunneling and Phase Transitions in 𝑆𝑈(2)-Gauge Theory
## 1 Introduction
One of the amazing phenomena of quantum physics is the barrier penetration due to tunneling processes. The occurrence of such processes in different areas of physics (solid state physics, high energy multiparticle scattering with baryon number violation, low-temperature physics, nuclear reactions, the formation of the Universe etc.) does not leave any doubt that they do actually take place. The theory of tunneling has been studied in many ways. It has become evident that such processes are due to classical configurations, which are solutions of the classical equation of motion with Euclidean time, namely stable vacuum configurations, now called instantons, which are responsible for transitions between topologically distinct vacua (relevant in explaining the high energy multiparticle collisions accompanied by baryon number violation) and unstable periodic configurations called periodic instantons and periodic bounces, which determine the decay of metastable physical systems. On the basis of the latter a theory of barrier penetration at finite energies has been developed. In particular it was shown, that the transition from temperature assisted tunneling to thermal activation can be considered as a phase transition which takes place as the temperature (or energy) of the system increases. Dissipative forces do not affect the general features of transitions. At zero temperature the barrier penetration is determined by tunneling with a rate controlled by vacuum instantons and is proportional to $`\mathrm{exp}(S)`$ where $`S`$ is the action. As the temperature increases the tunneling process (temperature assisted tunneling) begins to be suppressed and at sufficiently high energies (comparable with the height of the potential barrier) the system overrides the barrier (by thermal activation) and the penetration is governed by the Boltzman factor $`\mathrm{exp}(E_0/kT)`$ where $`E_0`$ is the energy of the system, corresponding to a particle sitting at the top of the barrier (the sphaleron). The configurations, which interpolate between these two processes are the periodic instantons . It was recently discovered, that depending on the shape of the potential barrier a transition of the first order (i.e. a sharp transition) is also possible. In the context of the Higgs model with some effective potential this type of transition was known before. In refs. criteria for the occurrence of the transition of the first order have been derived and examined for various quantum mechanical models. Furthermore it has been shown, that the periodic instantons may have bifurcations, which qualitatively change the behaviour of the phase transitions at finite temperature.
Recently large-spin systems turned out to be of increased interest, as these exhibit the first order phase transitions. The specific feature of these systems is a nonlinearity of the kinetic term. This could hint at the existence of sharp first order transitions in $`\sigma `$-models. One should keep in mind, that although the general theory of tunneling is well understood , only few examples are known, which make it possible to analyze the problem of phase transitions by explicit analytical calculations. In field theory the problem is even more complicated. In spite of these drawbacks one has succeeded to investigate the problem in some field-theoretical models numerically or by reducing the problem to a quantum mechanical one . Recent investigations in scalar theories in $`D+1`$\- dimensions with $`D=1,2,3`$ have shown the existence of both types of phase transitions depending on the value of the parameter, which violates a certain symmetry of the theory. From the point of view of the electroweak processes the $`\sigma `$-model and Abelian-Higgs model in two dimensions are of more interest, since as toy models they exhibit some features of the electroweak theory. It has been shown that whereas the $`\sigma `$-model with the symmetry-breaking term allows the first-order phase transition (adding the Skyrme-term allows the second-order transition too), the Abelian-Higgs model admits only second-order phase transitions. It is interesting, that in the model the sharp first-order transition takes place, if the spatial coordinate is compactified. Recently complex periodic instantons have been investigated in $`SU(2)`$-Yang-Mills-Higgs theory. New classes of bifurcating periodic instantons of the classical equations of motion were found numerically, depending on the self-coupling constant of the scalar field.
In the present work we investigate a theory with a vector field, namely globally $`SU(2)`$-invariant theory. The study of vector fields is, in our opinion, of interest from the point of view of such a complicated theory as the electroweak theory. The theory we are interested in is a pure Yang-Mills theory extended by addition of a quartic term. This term violates the local gauge invariance of theory, but there is still a global invariance. Although adding this term will destroy renormalizability this is irrelevant to the problem we are going to investigate. Considering the model in the framework of the field theory constructed on the sphere $`S^3`$ one succeeds to reduce the problem to the quantum mechanical one with the effective potential which is an asymmetric double-well potential and depends on the extra parameter, introduced with the quartic term. We show, that for all values of this parameter transitions are of second-order. The model we consider here is one of very few in more than two dimensions which can be treated analytically and therefore deserves particular attention even in spite of its idealization.
## 2 The model
The model we consider is described by the Euclidean action
$$S=d^4x\left\{\frac{1}{4}F_{\mu \nu a}F_{\mu \nu a}\frac{\mathrm{\Lambda }g^2}{12}(A_{\mu a}A_{\mu a})^2\right\}$$
(1)
in which as usual
$$F_{\mu \nu a}=_\mu A_{\nu a}_\nu A_{\mu a}+g\epsilon _{abc}A_{\mu b}A_{\nu c}.$$
(2)
Although the additional quartic term breaks the local gauge invariance the model is still globally $`SU(2)`$-invariant and is of interest from a field-theoretical point of view. The model admits pseudoparticle-antipseudoparticle classical field configurations. We present below periodic field configurations, which are responsible for quantum-classical phase transitions. In what follows we shall work in the framework of a field-theoretical approach , which is constructed on the sphere $`S^3`$ embedded in a 4-dimensional Euclidean space. This approach is especially convenient for conformally invariant theories, as the system can be considered to evolve along the radius of $`S^3,`$ in which case the operator of scale transformations becomes an evolution operator. Thus the radius $`r`$, namely the parameter $`\sigma =\mathrm{ln}r,`$ is a proper time of the physical system and the operator of scale transformations is considered as the “scaled energy” of the system.
We shall illustrate this by considering the simplest example of the scalar field $`\mathrm{\Phi }(x)`$. The general conformal group contains a dilatation operator $`D`$ defined in terms of the field $`\mathrm{\Phi }(x)`$ as
$$D=d^3xx_\nu T_{4\nu }$$
where $`T_{\mu \nu }`$ is the energy-momentum tensor. The scaling transformation for the field is
$$i[D,\mathrm{\Phi }(x)]=x_\mu _\mu \mathrm{\Phi }(x)+\mathrm{\Phi }(x).$$
We introduce spherical coordinates in 4-dimensional Euclidean space by setting
$$x_\mu =rn_\mu $$
(3)
with a unit vector
$$n_\mu =(\mathrm{sin}\psi \mathrm{sin}\theta \mathrm{cos}\varphi ,\mathrm{sin}\psi \mathrm{sin}\theta \mathrm{sin}\varphi ,\mathrm{sin}\psi \mathrm{cos}\theta ,\mathrm{cos}\psi ),$$
(4)
and define the field $`\chi (r,n_\mu )=r\mathrm{\Phi }(x)`$. In terms of the new coordinates and the field $`\chi (r,n_\mu )`$ the transformation law reads:
$$i[D,\chi (r,n_\mu )]=\frac{\chi (r,n_\mu )}{\mathrm{ln}r},$$
which makes the role of $`D`$ as the evolution operator evident. One can also find the following integral representation for D:
$$D=𝑑\mathrm{\Omega }\left\{\frac{}{\chi /\mathrm{ln}r}\frac{\chi }{\mathrm{ln}r}\right\},$$
which is the Legendre-transform of the Lagrangean $``$ integrated over the angles. The action of the system is then
$$S=d\mathrm{ln}rd\mathrm{\Omega }.$$
Thus we conclude, that in field theory constructed on $`S^3`$ the energy is replaced by the eigenvalues of the operator of the scaling transformations. The temperature is defined as the inverse of the period of periodic field configurations expressed in terms of the proper time $`\mathrm{ln}r`$. The period is determined as the derivative of the action with respect to the “scaled energy”.
We now return to our model. We look for periodic field configurations of period $`P(E)`$, where $`E`$ is a “scaled energy” of the system, which interpolate between the vacuum and the unstable field configuration named sphaleron as the proper time $`\tau `$ varies from $`P(E)/2`$ to $`P(E)/2.`$ We make the Ansatz
$$A_{\mu a}=\frac{1}{g}\eta _{a\mu \sigma }n_\sigma \frac{\varphi (r)}{r}$$
(5)
with the spherically symmetric function $`\varphi (r)`$ to be determined. Performing the integration frac12the angle variables we obtain (with $`\tau =\mathrm{ln}r`$):
$$S=\frac{6\pi ^2}{g^2}_{P(E)/2}^{P(E)/2}𝑑\tau \left\{\frac{1}{2}\left(\frac{\varphi (\tau )}{\tau }\right)^2+V(\varphi (\tau ))\right\},$$
(6)
in which $`V(\varphi (\tau ))`$ is an effective potential in terms of the function $`\varphi (\tau )`$:
$$V(\varphi (\tau ))=\frac{(2\mathrm{\Lambda })}{4}\varphi ^4(\tau )2\varphi ^3(\tau )+2\varphi ^2(\tau )$$
(7)
The shape of the potential depends on the parameter $`\mathrm{\Lambda }`$.
Fig. 1. Different shapes of the effective potential
for different values of $`\mathrm{\Lambda }`$.
The potential $`V(\varphi (\tau ))`$ is quartic except for $`\mathrm{\Lambda }=2`$ in which case it becomes cubic. For nonzero values of $`\mathrm{\Lambda }`$ the potential is asymmetric, indeed for $`0<\mathrm{\Lambda }<2`$ there are two unequal minima at
$$\varphi _{1min}=0,\varphi _{2min}=\frac{3+\sqrt{1+4\mathrm{\Lambda }}}{2\mathrm{\Lambda }}>0$$
(8)
and one maximum at
$$\varphi _{max}=\frac{3\sqrt{1+4\mathrm{\Lambda }}}{2\mathrm{\Lambda }}>0.$$
(9)
For $`\mathrm{\Lambda }>2`$ the shape of the potential is changed and there are two maxima at
$$\varphi _{lmax}=\frac{3\sqrt{1+4\mathrm{\Lambda }}}{\mathrm{\Lambda }2}<0,\varphi _{smax}=\frac{3+\sqrt{1+4\mathrm{\Lambda }}}{\mathrm{\Lambda }2}>0,$$
(10)
with the small and large barriers $`V_{smax}`$ and $`V_{lmax}`$ and one minimum at
$$\varphi _{min}=0.$$
(11)
In the case of $`\mathrm{\Lambda }=2`$ the potential is cubic with the minimum at $`\varphi _c=0`$ and maximum at $`\varphi _c=2/3`$ . The case of $`\mathrm{\Lambda }=0`$ is just the gauge theory with local $`SU(2)`$-symmetry and the effective potential is symmetric in this case with two minima at $`\varphi =0,2`$ and one maximum at $`\varphi =1`$. All these cases are shown in Fig. 1.
The points of intersection of the line $`E=const`$ with the potential determine the turning points of the periodic motion.
## 3 Quantum-classical transitions.
### 3.1 The case $`0<\mathrm{\Lambda }<2`$
The equation of motion for nonzero constant of integration $`E`$ reads:
$$\frac{1}{2}\left(\frac{\varphi }{\tau }\right)^2=V(\varphi (\tau ))E.$$
(12)
The solution, which satisfies the periodicity condition $`\varphi (\tau +P(E))=\varphi (\tau )`$ is found to be:
$$\varphi (\tau )=\frac{c(bd)d(bc)\mathrm{sn}^2(B(\mathrm{\Lambda },E)\tau +\mathrm{K}(k),k)}{(bd)(bc)\mathrm{sn}^2(B(\mathrm{\Lambda },E)\tau +\mathrm{K}(k),k)},$$
(13)
in which
$$B(\mathrm{\Lambda },E)=\sqrt{\frac{(2\mathrm{\Lambda })}{8}}\sqrt{(bd)(ac)},$$
(14)
and $`k`$ is the modulus of the Jacobian elliptic function
$$k=\sqrt{\frac{(bc)(ad)}{(ac)(bd)}}.$$
(15)
The function $`\mathrm{K}(k)`$ is the complete elliptic integral of the first kind. The quantities $`a,b,c,d`$ with $`a>b>c>d`$ are turning points of the motion (see Appendix). The solution corresponds to the periodic motion from the point $`c`$ via the maximum of the potential to $`b`$ and back. The period of motion is
$$P(E)=\frac{2\mathrm{K}(k)}{B(\mathrm{\Lambda },E)},$$
(16)
so that
$$\varphi (P(E)/2)=\varphi (P(E)/2)=c\mathrm{and}\varphi (0)=b.$$
The period $`P(E)`$ is a monotonous function of the energy, as shown in Fig. 2. This indicates that the quantum-classical phase transition is of second order.
Fig. 2. Energy dependence of the period
for $`\mathrm{\Lambda }=1/8`$, which is typical for all $`\mathrm{\Lambda }`$.
Substituting the solution into the action and integrating over the period gives the following expression:
$`S(\mathrm{\Lambda },E)={\displaystyle \frac{6\pi ^2}{g^2}}\{EP(E)+D(\mathrm{\Lambda },E)\mathrm{K}(k)+G(\mathrm{\Lambda },E)\mathrm{E}(k)+`$
$`H(\mathrm{\Lambda },E)\mathrm{\Pi }(\alpha ^2,k)\},`$ (17)
with
$$\alpha ^2=\frac{bc}{bd}>0$$
The functions $`D(E),G(E),H(E)`$ are defined in the Appendix. The functions $`\mathrm{K}(k),\mathrm{E}(k)`$ and $`\mathrm{\Pi }(\alpha ^2,k)`$ are complete elliptic integrals of the first, second and third kind respectively. In semiclassical approximation the solution describes the quantum tunneling process. As the solution responsible for the thermal activation we choose the constant solution of the equation of motion, namely $`\varphi (\tau )=\varphi _{max}`$ which corresponds to the “particle” with the maximal energy $`E=V_{max}=V(\varphi _{max})`$ sitting at the top of the potential barrier. The action of this configuration is
$$S_0(T)=\frac{6\pi ^2}{g^2}\frac{V_{max}}{T},$$
(18)
where $`T`$ is the temperature. In Fig. 3 we display the action-versus-temperature plot (with $`P(E)=1/T(E)`$ in (17)). One can see, that the transition is of the second order with the transition temperature $`T_{cr}=0.24`$.
Fig. 3. The action-versus-temperature diagram: the solid line
represents the action of the the sphaleron $`\varphi (\tau )=\varphi _{max}`$
and the dashed line the action of the periodic instanton
for $`\mathrm{\Lambda }=1/8`$.
The case of $`\mathrm{\Lambda }=0`$ (theory with local symmetry) is included in our formulae for $`0<\mathrm{\Lambda }<2`$ by setting $`\mathrm{\Lambda }=0`$. Nevertheless we give the exact expressions for the solution of the problem. The turning points for the periodic motion with finite“energy” as the solutions of the equation
$$V(\varphi )=\frac{1}{2}\varphi ^42\varphi ^3+\varphi ^2=E$$
are:
$$1\sqrt{1+\sqrt{E}},\mathrm{\hspace{0.33em}1}\sqrt{1\sqrt{E}},\mathrm{\hspace{0.33em}1}+\sqrt{1\sqrt{E}},\mathrm{\hspace{0.33em}1}+\sqrt{1+\sqrt{E}}.$$
The periodic field configuration, which describes the motion between the points $`1\sqrt{1\sqrt{E}}`$ and $`1\sqrt{1+\sqrt{E}}`$ is
$$\varphi (\tau )=1+\sqrt{1\sqrt{E}}\mathrm{sn}\sqrt{1+\sqrt{E}}(\tau +P_0(E)/2)$$
(19)
and satisfies the condition $`\varphi (P_0(E)/2)=\varphi (P_0(E)/2)`$, in which
$$P_0(E)=\frac{4}{\sqrt{1+\sqrt{E}}}\mathrm{K}(k_0)$$
is a period of motion with modulus
$$k_0=\sqrt{\frac{1\sqrt{E}}{1+\sqrt{E}}}.$$
The values of the actions corresponding to the periodic motion and the static field configuration $`\varphi =1/2`$ (at the top of the barrier) are then
$$S_0(E)=\frac{6\pi ^2}{g^2}\left\{\frac{E}{T(E)}+\frac{4}{3}(1+\sqrt{E})\sqrt{1\sqrt{E}}\left[(1+k_0^2)\mathrm{E}(k_0)(1k_0^2)\mathrm{K}(k_0)\right]\right\},$$
$$S_{0st}=\frac{3\pi ^2}{g^2T}.$$
with temperature $`T_0(E)=P_0^1(E)`$. The period is a decreasing function of the energy and the phase transition is of second order. Finally we mention, that although we have restricted ourselves to positive values of the parameter $`\mathrm{\Lambda }`$, some negative values may also be allowed, in particular one sees from the expressions of the extrema of the potential $`V(\varphi (\tau ))`$ that the maximum exists for $`1/4<\mathrm{\Lambda }`$. For the values of $`\mathrm{\Lambda }1/4`$ the potential $`V(\varphi (\tau ))`$ does not have a maximum any more.
### 3.2 The case $`\mathrm{\Lambda }>2`$
In the case of $`\mathrm{\Lambda }>2`$ the “particle” sitting in the potential well with minimal energy can move in the direction of either a bigger or smaller barrier. The solutions of the field equations in both cases will be presented in the form, which is different from that given by (13). The turning points $`a_s>b_s>a_l>b_l`$ are given by
$$a_s=m_s+n_s,b_s=m_sn_s,a_l=m_l+n_l,b_l=m_ln_l,$$
where $`=m_s,n_s,m_l,n_l`$ are determined in the Appendix. We consider first the motion to the small barrier. In this case the periodic field configuration is:
$$\varphi _s(\tau )=\frac{q_sb_s+p_sa_s(q_sb_sp_sa_s)\mathrm{cn}(\sqrt{\frac{(\mathrm{\Lambda }2)p_sq_s}{2}}(\tau +P_s(E)/2),k_s)}{q_s+p_s(q_sp_s)\mathrm{cn}(\sqrt{\frac{(\mathrm{\Lambda }2)p_sq_s}{2}}(\tau +P_s(E)/2),k_s)},$$
(20)
with the conditions
$`\varphi _s(P_s(E)/2)`$ $`=`$ $`\varphi _s(P_s(E)/2))=a,`$
$`\varphi _s(P_s(E)/4)`$ $`=`$ $`\varphi _s(P_s(E)/4)=V_{\mathrm{smax}},`$
$`\varphi _s(0)`$ $`=`$ $`b.`$
The real quantities $`q_s,p_s`$ and the modulus $`k_s`$ of the Jacobian elliptic functions are defined as
$`q_s^2=(m_la_s)^2n_l^2,p_s^2=(m_lb_s)^2n_l^2,k_s={\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{4n_s^2(p_sq_s)^2}{p_sq_s}}}.`$
The period of the motion
$$P_s(E)=4\sqrt{\frac{2}{(\mathrm{\Lambda }2)p_sq_s}}\mathrm{K}(k_s)$$
(21)
is again a monotonically decreasing function of the energy $`E`$. The action integrated out over the period is a linear combination of complete elliptic integrals:
$`S_s(\mathrm{\Lambda },E)={\displaystyle \frac{6\pi ^2}{g^2}}\{EP_s(E)+D_s(\mathrm{\Lambda },E)\mathrm{K}(k_s)+G_s(\mathrm{\Lambda },E)\mathrm{E}(k_s)+.`$
$`.H_s(\mathrm{\Lambda },E)\mathrm{\Pi }({\displaystyle \frac{a_s^2}{\alpha _s^21}},k_s)\},`$ (22)
with
$$\alpha _s^2=\frac{p_sq_s}{p_s+q_s}.$$
The coefficients $`D_s(E),G_s(E),H_s(E)`$) are given in the Appendix. Fig. 4 a) shows the action-versus-temperature plot, in which
$$S_{s0}(T)=\frac{6\pi ^2}{g^2}\frac{V_{smax}}{T}$$
is an action corresponding to the top of the small barrier.
a) b)
Fig. 4. The action-versus-temperature diagram: a)For the small barrier the solid line represents the action $`S_{s0}`$ of the the sphaleron $`\varphi (\tau )=\varphi _{smax}`$ and the dashed line the action $`S_s(\mathrm{\Lambda },E)`$ of the periodic instanton at $`\mathrm{\Lambda }=14`$. b)For the large barrier the solid line represents the action $`S_{l0}`$ of the the sphaleron $`\varphi (\tau )=\varphi _{lmax}`$ and the dashed line the action $`S_l(\mathrm{\Lambda },E)`$ of the periodic instanton at $`\mathrm{\Lambda }=14`$.
In the case of motion to the large barrier the solution is defined by formula (20) with appropriately replaced coefficients, in particular
$$\varphi _l(\tau )=\frac{q_lb_l+p_la_l(q_lb_lp_la_l)\mathrm{cn}(\sqrt{\frac{(\mathrm{\Lambda }2)p_lq_l}{2}}(\tau +P_l(E)/2),k_l)}{q_l+p_l(q_lp_l)\mathrm{cn}(\sqrt{\frac{(\mathrm{\Lambda }2)p_lq_l}{2}}(\tau +P_l(E)/2),k_l)},$$
(23)
This solution corresponds to periodic motion starting at the point $`a_l`$, going to $`b_l`$ and back to $`a_l`$. The quantities $`p_l,q_l`$ are given by
$$q_l^2=(m_sa_l)^2n_s^2,p_l^2=(m_sb_l)^2n_s^2.$$
The period $`P_l(E)`$ is a monotonically decreasing function of $`E`$ and is given by (21) with $`k_s`$ replaced as
$$k_sk_l=\frac{1}{2}\sqrt{\frac{4n_l^2(p_lq_l)^2}{p_lq_l}}.$$
The action $`S_l(\mathrm{\Lambda },E)`$ is also defined by (22) with different coefficients $`D_l(E),G_l(E),H_l(E)`$ (see Appendix) and the parameter
$$\alpha _l^2=\frac{p_lq_l}{p_l+q_l}$$
instead of $`\alpha _s^2`$. The action of the sphaleron in this case is expressed through the value $`V_{lmax}`$:
$$S_{l0}=\frac{6\pi ^2}{g^2}\frac{V_{lmax}}{T}.$$
The corresponding diagrams, again showing the existence of the smooth second order phase transition, are shown in Fig. 4 b). The critical temperatures are $`T_{lcr}=0.58`$ and $`T_{scr}=0.38`$. One can see, that tunneling through the large potential barrier dominates that of the small barrier ($`S_l(\mathrm{\Lambda },E)>S_s(\mathrm{\Lambda },E)`$ for a given “energy” $`E`$). With increase of the parameter $`\mathrm{\Lambda }`$ this difference is diminished.
### 3.3 The case $`\mathrm{\Lambda }=2`$.
In the case of $`\mathrm{\Lambda }=2`$ the finite “energy” periodic solution of the equation of motion is
$$\varphi _c(\tau )=\frac{b_c(a_cd_c)d_c(a_cb_c)\mathrm{sn}^2(\sqrt{\frac{a_cd_c}{2}}(\tau +\frac{P(E)}{2}))}{(a_cd_c)(a_cb_c)\mathrm{sn}^2(\sqrt{\frac{a_cd_c}{2}}(\tau +\frac{P(E)}{2}))},$$
(24)
in which
$$k_c=\sqrt{\frac{a_cb_c}{a_cd_c}}$$
is the modulus of the elliptic functions and the quantities $`a_c>b_c>d_c`$ are turning points given by the solutions of the cubic equation
$$2\varphi ^3+2\varphi ^2=E.$$
(25)
The solution describes a motion from the point $`b_c`$ to $`a_c`$ and back. The period of motion
$$P(E)=\frac{2\sqrt{2}\mathrm{K}(k_c)}{\sqrt{a_cd_c}}$$
is a monotonic function of $`E`$. The action integrated out over the period reads:
$`S_c(\mathrm{\Lambda },E)={\displaystyle \frac{6\pi ^2}{g^2}}\{{\displaystyle \frac{E}{T(E)}}+{\displaystyle \frac{8(b_cd_c)^2}{5}}\sqrt{{\displaystyle \frac{a_cd_c}{2}}}[\left({\displaystyle \frac{2(2k_c^2)}{3(1k_c^2)^2}}\right)`$
$`{\displaystyle \frac{2}{1k_c^2}}\mathrm{E}(k_c){\displaystyle \frac{(2k_c^2)}{3(1k_c^2)}}\mathrm{K}(k_c)]\}.`$ (26)
The maximum of the potential is $`8/27`$ and thus the action of the static field $`\varphi _c`$ is $`S_{c0}=16\pi ^2/9T`$. The“energy” dependence of the period and the action-versus-temperature diagrams confirm the existence of the phase transitions of the second order and we do not reproduce the corresponding diagrams.
## 4 Conclusions
Above we demonstrated that in the model considered the smooth phase transitions of the second order take place. This is not surprising, since it is known from general arguments that in theories with cubic and quartic terms the phase transitions of the second order occur. Although our considerations are explicit they are based on the Ansatz (5), which singles out a class of periodic field configurations (which are nonselfdual) reducing the problem to the one-dimensional one. In order to check the criteria for occurrence of a transition of the first order one has to investigate fluctuations around a static (in our model constant) configuration. The nonselfdual character of the solutions complicates the second order fluctuation differential equations for derivation of analytical solutions considerably, so that this is not attempted here.
Acknowledgment. J.-Q.L.and D. K. Park are indebted to The Deutsche Forschungsgemeinschaft (Germany) for financial support of visits to Kaiserslautern.
Appendix
We here give some explicit formulae referred to above for the $`3`$ possible domains of the parameter $`\mathrm{\Lambda }`$.
a) $`0<\mathrm{\Lambda }<2.`$
The turning points of the periodic motion as solutions of the equation
$$\frac{(2\mathrm{\Lambda })}{4}\varphi ^42\varphi ^3+\varphi ^2=E$$
are:
$`a`$ $`=`$ $`{\displaystyle \frac{1}{2}}(r+\sqrt{r^22r+2y})+{\displaystyle \frac{1}{2}}\sqrt{2(r^2ry)+2r{\displaystyle \frac{r^22r}{\sqrt{r^22r+2y}}}},`$
$`b`$ $`=`$ $`{\displaystyle \frac{1}{2}}(r+\sqrt{r^22r+2y}){\displaystyle \frac{1}{2}}\sqrt{2(r^2ry)+2r{\displaystyle \frac{r^22r}{\sqrt{r^22r+2y}}}},`$
$`c`$ $`=`$ $`{\displaystyle \frac{1}{2}}(r\sqrt{r^22r+2y})+{\displaystyle \frac{1}{2}}\sqrt{2(r^2ry)2r{\displaystyle \frac{r^22r}{\sqrt{r^22r+2y}}}},`$
$`d`$ $`=`$ $`{\displaystyle \frac{1}{2}}(r\sqrt{r^22r+2y}){\displaystyle \frac{1}{2}}\sqrt{2(r^2ry)2r{\displaystyle \frac{r^22r}{\sqrt{r^22r+2y}}}},`$
in which
$$r=\frac{4}{2\mathrm{\Lambda }},$$
$$y=\frac{1}{3}r+2\sqrt{\frac{1}{3}\left(\frac{r^2}{3}rE\right)}\mathrm{cos}\left(\frac{1}{3}\mathrm{arctan}\frac{2\sqrt{\frac{r^3}{27}E(EV_{max})(EV_{2min})}}{\frac{2}{27}r^3+\frac{1}{2}r^2(r\frac{4}{3})E}\right).$$
One checks, that the quantities $`a,b,c,d`$ are real for “energies” $`E`$ in $`0<E<V_{max}`$. The coefficients in the expression for the action $`S(\mathrm{\Lambda },E)`$ are defined as
$$D(\mathrm{\Lambda },E)=\frac{4B(\mathrm{\Lambda },E)}{\alpha ^2}(cd)^2(\frac{\alpha ^2k^2+\alpha ^23k^2}{3(1\alpha ^2)}+\frac{(2\alpha ^2k^2+2\alpha ^2\alpha ^43k^2)^2}{8(1\alpha ^2)^2(k^2\alpha ^2)}$$
$$+\frac{k^2(2\alpha ^2k^2+2\alpha ^2\alpha ^43k^2)}{12(1\alpha ^2)(k^2\alpha ^2)}),$$
$$G(\mathrm{\Lambda },E)=\frac{4B(\mathrm{\Lambda },E)(cd)^2}{k^2\alpha ^2}\left(\frac{\alpha ^2k^2+\alpha ^23k^2}{3(1\alpha ^2)}+\frac{(2\alpha ^2k^2+2\alpha ^2\alpha ^43k^2)^2}{8(1\alpha ^2)^2(k^2\alpha ^2)}\right),$$
$$H(\mathrm{\Lambda },E)=\frac{4B(\mathrm{\Lambda },E)(cd)^2}{\alpha ^2}(k^2+$$
$$\frac{(\alpha ^2k^2+\alpha ^23k^2)(2\alpha ^2k^2+2\alpha ^2\alpha ^43k^2)}{2(1\alpha ^2)(k^2\alpha ^2)}+\frac{(2\alpha ^2k^2+2\alpha ^2\alpha ^43k^2)^3}{8(1\alpha ^2)^2(k^2\alpha ^2)^2}).$$
b) $`\mathrm{\Lambda }>2.`$
The quantities $`m_s,n_s,m_l,n_l`$ which determine the turning points in this case, are given by
$`m_s`$ $`=`$ $`{\displaystyle \frac{1}{2}}(r+\sqrt{r^2+2r+2y}),`$
$`n_s`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{4r(r+1)(r\sqrt{r^2+2r+2y})^2+4\sqrt{y^2rE}},`$
$`m_l`$ $`=`$ $`{\displaystyle \frac{1}{2}}(r\sqrt{r^2+2r+2y}),`$
$`n_l`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{4r(r+1)(r+\sqrt{r^2+2r+2y})^24\sqrt{y^2rE}},`$
with
$$r=\frac{4}{\mathrm{\Lambda }2}$$
and
$$y=\frac{1}{3}r+2\sqrt{\frac{1}{3}\left(\frac{r^2}{3}+rE\right)}\mathrm{cos}\left(\frac{1}{3}\mathrm{arctan}\frac{2\sqrt{\frac{r^3}{27}E(EV_{smax})(EV_{\mathrm{lmax}})}}{\frac{2}{27}r^3\frac{1}{2}r^2(r+\frac{4}{3})E}\right).$$
The coefficients which appear in the action $`S_s`$ read:
$$D_s(\mathrm{\Lambda },E)=\frac{8p_s^2q_s^2(a_sb_s)^2}{3\alpha _s^4(p_s+q_s)^4}\sqrt{\frac{(\mathrm{\Lambda }2)p_sq_s}{2}}[4k_s^2\frac{2(6k_s^2+\alpha _s^22\alpha _s^2k_s^2)}{(\alpha _s^21)(\alpha _s^2+k_s^2\alpha _s^2k_s^2)}$$
$$\frac{3(2\alpha _s^2k_s^2\alpha _s^22k_s^2)^2}{(\alpha _s^21)^2(\alpha _s^2+k_s^2\alpha _s^2k_s^2)^2}+\frac{2k_s^2(2\alpha _s^2k_s^2\alpha _s^22k_s^2)}{(\alpha _s^21)(\alpha _s^2+k_s^2\alpha _s^2k_s^2)}],$$
$$G_s(\mathrm{\Lambda },E)=\frac{8p_s^2q_s^2(a_sb_s)^2}{3\alpha _s^4(p_s+q_s)^4}\sqrt{\frac{(\mathrm{\Lambda }2)p_sq_s}{2}}[\frac{2\alpha _s^2(6k_s^2+\alpha _s^22\alpha _s^2k_s^2)}{(\alpha _s^21)(\alpha _s^2+k_s^2\alpha _s^2k_s^2)}+$$
$$\frac{3\alpha _s^2(2\alpha _s^2k_s^2\alpha _s^22k_s^2)}{(\alpha _s^21)^2(\alpha _s^2+k_s^2\alpha _s^2k_s^2)^2}],$$
$$H_s(\mathrm{\Lambda },E)=\frac{8p_s^2q_s^2(a_sb_s)^2}{3\alpha _s^4(p_s+q_s)^4}\sqrt{\frac{(\mathrm{\Lambda }2)p_sq_s}{2}}(\frac{12k_s^2}{1\alpha _s^2}+$$
$$\frac{3(2\alpha _s^2k_s^2\alpha _s^22k_s^2)^3}{(\alpha _s^21)^3(\alpha _s^2+k_s^2\alpha _s^2k_s^2)^2}+\frac{3(2\alpha _s^2k_s^2\alpha _s^22k_s^2)(6k_s^2+\alpha _s^22\alpha _s^2k_s^2)}{(\alpha _s^21)^2(\alpha _s^2+k_s^2\alpha _s^2k_s^2)}).$$
The corresponding quantities in $`S_l(\mathrm{\Lambda },E)`$ are obtained by the following replacements:
$$a_s,b_sa_l,b_l,\alpha _s\alpha _lk_sk_l,p_s,q_sp_l,q_l.$$
c) $`\mathrm{\Lambda }=2.`$
The turning points in this case are the solutions of the cubic equation (25), in particular:
$$a_c=\frac{1}{3}+\mathrm{cos}\frac{\varphi }{3},b_c=\frac{1}{3}+\mathrm{cos}\frac{\varphi +2\pi }{3},d_c=\frac{1}{3}+\mathrm{cos}\frac{\varphi +4\pi }{3},$$
in which the angle $`\varphi `$ is defined by
$$\mathrm{tan}\frac{\varphi }{3}=\frac{2\sqrt{\frac{1}{4}(\frac{2}{27}+\frac{E}{2})^2+\frac{1}{27}}}{\frac{2}{27}+\frac{E}{2}}.$$
|
warning/0007/cond-mat0007054.html
|
ar5iv
|
text
|
# Broken Generalized Kohn Theorem in Harmonic Dot Lattices due to Coulomb Interaction between the Dots: Exact solution of the Schrödinger equation in the dipole approximation
## I Introduction
The Generalized Kohn Theorem (GKTh) plays a crucial role in quantum dot physics with far reaching consequences. It considers interacting electron systems in a harmonic confinement and a constant magnetic field, and it states that excitations by long wavelength radiation are not effected by the electron electron (e e) interaction. This statement applies to arrays of identical harmonic dot confinements (with e e interaction between the dots) as well (see Appendix of ). This does not mean that all excitations are independent of e e interaction, but only the optically active ones (Kohn modes), and it does not mean that e e interaction is not important for the other excitations. However, this fact prevents the e e interaction from beeing seen and investigated e.g. by far infrared (FIR) spectroscopy. The FIR absorption spectrum of the whole system agrees exactly with the spectrum of a single particle. The GKTh does not hold for arrays of different dot confinements, e.g. periodic dot lattices with two different harmonic dot confinements per unit cell . Then, all collective modes are excited by FIR radiation and effected by e e interaction, or in other words, there is no Kohn mode. The calculation and investigation of absorption frequencies and probabilities in the latter case is the subject of this work.
In order to obtain a visual picture, let us first consider a classical model for the Kohn mode for vanishing magnetic field. (This preliminary consideration will be replaced by a rigorous quantum mechanical treatment in the following.) Classically, the charge distributions of all dots oscillate rigidly in– phase with the bare confinement frequency, and the e e interaction contributes only a constant term to the total energy (independent of elongation). If we have more than one identical dots per unit cell, there are additional collective modes, in which the individual dots oscillate out of phase, and which are affected by dot interaction, but which are not optically active. Consequently, the dot interaction is not observable with FIR spectroscopy in arrangements of identical dots. One way to trick Kohn’s theorem is to include different dot species. Then, there is no coherent oscillation mode for all dots, which does not change the e e interaction energy of the system in elongation, because there is no common bare confinement frequency. As a consequence, all collective modes (two modes per dot in the unit cell) are effected by dot interaction and excited by FIR radiation with a finite probability. In other words, the Generalized Kohn Theorem for dot arrays is broken.
Other systems, where Kohn’s Theorem does not hold, comprise: i) anharmonic confinements (circular dots with $`r^4`$ and higher order terms in the radial dependence or cubic dots with terms of type $`x^2y^2`$) , ii) hole dots with different effective masses . One point of this paper is that the GKTh can be broken despite an exactly harmonic Hamiltonian. A further possibility to observe the e e interaction in the excitations is to consider finite wave length .
## II Magnetophonon Hamiltonian
The first part of the calculation of the eigenstates of the Hamiltonian follows closely the procedure described in Ref.. We only have to consider that now the confinement potentials and electron numbers can be different in different dots. After introducing center– of – mass (c.m.) and relative coordinates in each dot and applying the dipole approximation for the Coulomb interaction between the dots, we observe that the Hamiltonian of all c.m. coordinates is decoupled from individual dot Hamiltonians in the relative coordinates. That’s why all excitations can be classified into i) collective (c.m.) excitations, and ii) intra-dot excitations. The latter are not considered here because they are not optically active. The Hamiltonian in the c.m. coordinates $`𝐑_{n,\alpha }`$ reads in atomic units $`\mathrm{}=m=e=1`$ (see also Sect. IV A in Ref.)
$`H_{c.m.}`$ $`=`$ $`{\displaystyle \underset{n,\alpha }{}}{\displaystyle \frac{1}{2m^{}}}\left[{\displaystyle \frac{𝐏_{n,\alpha }}{\sqrt{N_\alpha }}}+{\displaystyle \frac{\sqrt{N_\alpha }}{c}}𝐀\left(𝐔_{n,\alpha }\right)\right]^2`$ (2)
$`+{\displaystyle \frac{1}{2}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{n,\alpha }{n^{},\alpha ^{}}}{}}\sqrt{N_\alpha N_\alpha ^{}}𝐔_{n,\alpha }𝐂_{n,\alpha ;n^{},\alpha ^{}}𝐔_{n^{},\alpha ^{}}`$
where $`𝐔_{n,\alpha }=𝐑_{n,\alpha }𝐑_{n,\alpha }^{(0)}`$ is the elongation of the c.m. at lattice site $`(n,\alpha )`$ and $`𝐏_{n,\alpha }=i_{𝐔_{n,\alpha }}`$ is the corresponding canonical momentum operator. $`n`$ runs over the unit cells and $`\alpha `$ over the dot species within a cell. $`N_\alpha `$ is the number of electrons in dot $`\alpha `$, and $`m^{}`$ the effective mass. It is clear already from inspection of (2) that the eigenvalues of $`H_{c.m.}`$ do not depend on the explicitly shown electron numbers $`N_\alpha `$, because the factors $`\sqrt{N_\alpha }`$ can be considered just as a rescaling factor of the coordinates $`𝐔_{n,\alpha }`$. However, the eigenfunctions (and quantities derived from them) do depend on the explicit $`N_\alpha `$. The force constant tensor reads
$`C_{n,\alpha ;n,\alpha }`$ $`=`$ $`𝛀_\alpha +ϵ^1N_\alpha {\displaystyle \underset{n^{},\alpha ^{}(n,\alpha )}{}}𝐓\left(𝐑_{n,\alpha }^{(0)}𝐑_{n^{},\alpha ^{}}^{(0)}\right)`$ (3)
$`C_{n,\alpha ;n^{},\alpha ^{}}`$ $`=`$ $`ϵ^1\sqrt{N_\alpha N_\alpha ^{}}𝐓\left(𝐑_{n,\alpha }^{(0)}𝐑_{n^{},\alpha ^{}}^{(0)}\right)\text{for}(n,\alpha )(n^{},\alpha ^{})`$ (4)
where $`ϵ^1`$ is the inverse background dielectric constant and $`𝛀_\alpha `$ the bare confinement tensor, which produces a harmonic confinement. The dipole tensor is defined as $`𝐓(𝒂)=\frac{1}{a^5}\left[3𝒂𝒂a^2𝐈\right]`$ where ($``$) denotes the dyad product and $`𝐈`$ the unit tensor. Observe that $`𝐂`$ depends on $`N_\alpha `$ implicitely which effects the energy eigenvalues.
A unitary transformation to collective magnetophonon coordinates
$`𝐔_{n,\alpha }`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{N_c}}}{\displaystyle \underset{𝐪}{\overset{BZ}{}}}e^{i𝐪R_n^{(0)}}𝐔_{𝐪,\alpha }`$ (5)
$`𝐏_{n,\alpha }`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{N_c}}}{\displaystyle \underset{𝐪}{\overset{BZ}{}}}e^{+i𝐪R_n^{(0)}}𝐏_{𝐪,\alpha }`$ (6)
where $`N_c`$ is the number of unit cells, leaves us with a sum on $`N_c`$ decoupled subsystems $`H_{c.m.}=_𝐪H_𝐪`$
$`H_𝐪`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m^{}}}\left[{\displaystyle \frac{𝐏_{𝐪,\alpha }}{\sqrt{N_\alpha }}}+{\displaystyle \frac{\sqrt{N_\alpha }}{c}}𝐀(𝐔_{𝐪,\alpha }^{})\right]^{}\left[{\displaystyle \frac{𝐏_{𝐪,\alpha }}{\sqrt{N_\alpha }}}+{\displaystyle \frac{\sqrt{N_\alpha }}{c}}𝐀(𝐔_{𝐪,\alpha }^{})\right]`$ (8)
$`+{\displaystyle \frac{1}{2}}{\displaystyle \underset{\alpha ,\alpha ^{}}{}}\sqrt{N_\alpha N_\alpha ^{}}𝐔_{𝐪,\alpha }^{}𝐂_{𝐪;\alpha ,\alpha ^{}}𝐔_{𝐪,\alpha ^{}}`$
which includes the dynamical matrix
$$𝐂_{𝐪;\alpha ,\alpha ^{}}=\underset{n}{}e^{i𝐪𝐑_n^{(0)}}𝐂_{\alpha ,\alpha ^{}}\left(𝐑_n^{(0)}\right);𝐂_{\alpha ,\alpha ^{}}\left(𝐑_n^{(0)}\right)=𝐂_{n,\alpha ;\mathrm{\hspace{0.17em}0},\alpha ^{}}$$
(9)
With (3) and (4), we obtain
$`C_{𝐪;\alpha ,\alpha }`$ $`=`$ $`𝛀_\alpha +ϵ^1N_\alpha {\displaystyle \underset{\alpha ^{}(\alpha )}{}}𝐓\left(𝒂_\alpha 𝒂_\alpha ^{}\right)`$ (10)
$`+`$ $`{\displaystyle \underset{n0}{}}\left[{\displaystyle \underset{\alpha ^{}}{}}𝐓\left(𝐑_n^{(0)}+𝒂_\alpha 𝒂_\alpha ^{}\right)e^{i𝐪𝐑_n^{(0)}}𝐓\left(𝐑_n^{(0)}\right)\right]`$ (11)
$`C_{𝐪;\alpha ,\alpha ^{}}`$ $`=`$ $`ϵ^1\sqrt{N_\alpha N_\alpha ^{}}{\displaystyle \underset{n}{}}e^{i𝐪𝐑_n^{(0)}}𝐓\left(𝐑_n^{(0)}+𝒂_\alpha 𝒂_\alpha ^{}\right)\text{for}\alpha \alpha ^{}`$ (12)
where $`𝐑_{n,\alpha }^{(0)}=𝐑_n^{(0)}+𝒂_\alpha `$ and $`n0`$ under the sum means that the term $`𝐑_n^{(0)}=0`$ is excluded.
Now, we focus our attention to long– wavelength modes (the index $`𝐪=0`$ is dropped henceforth) and consider a simple cubic lattice, alternatively occupied by two different dot species. The minimum unit cell is face centered cubic (see Fig.1) with lattice constant $`a`$. After performing the lattice sum involved in (9) numerically, we obtain the four $`2\times 2`$ dynamical matrices
$`𝐂_{11}`$ $`=`$ $`𝛀_1+dp_1𝐈`$ (13)
$`𝐂_{22}`$ $`=`$ $`𝛀_2+dp_2𝐈`$ (14)
$`𝐂_{12}`$ $`=`$ $`𝐂_{21}=dp_{12}𝐈`$ (15)
with the interaction parameters
$`p_i`$ $`=`$ $`2N_iϵ^1/(n.n.distance)^3=4\sqrt{2}N_iϵ^1/a^3,(i=1,2)`$ (16)
$`p_{12}`$ $`=`$ $`2\sqrt{N_1N_2}ϵ^1/(n.n.distance)^3=4\sqrt{2}\sqrt{N_1N_2}ϵ^1/a^3`$ (17)
and $`d=1.460`$. From the preceding definitions it follows that $`p_{12}=\sqrt{p_1p_2}`$.
## III Eigenstates
Now we are going to find eigenvalues and eigenfunction of (8). For avoiding divergences for $`B=0`$, we add an isotropic oscillator potential $`\frac{1}{2}_\alpha \omega _0^2𝐔_\alpha ^2`$ to the kinetic energy in (8) and subtract it from the interaction term. $`\omega _0`$ is in principle arbitrary, but we chose the mean value of the bare confinement frequencies included in $`\mathrm{\Omega }_1`$ and $`\mathrm{\Omega }_2`$. Now we replace the coordinates in (8) (for $`𝐪=0`$) by Boson ladder operators. This is analogous to the usual text book transformation (see e.g. Ref. Sect. 3.3) apart from the factors $`\sqrt{N_\alpha }`$. It is obvious that this modification can be taken into account by introducing scaled coordinates $`𝐔_\alpha \stackrel{~}{𝐔}_\alpha =\sqrt{N_\alpha }𝐔_\alpha `$ (what implies $`𝐏_\alpha \stackrel{~}{𝐏}_\alpha =𝐏_\alpha /\sqrt{N_\alpha }`$).
$`\sqrt{N_\alpha }U_{\alpha x}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{2}{\stackrel{~}{\omega }_c^{}}}}\left(a_{\alpha 1}^++a_{\alpha 2}^++a_{\alpha 1}+a_{\alpha 2}\right)`$ (18)
$`\sqrt{N_\alpha }U_{\alpha y}`$ $`=`$ $`{\displaystyle \frac{i}{2}}\sqrt{{\displaystyle \frac{2}{\stackrel{~}{\omega }_c^{}}}}\left(a_{\alpha 1}^+a_{\alpha 2}^+a_{\alpha 1}+a_{\alpha 2}\right)`$ (19)
where the first subscript $`(\alpha =1,2)`$ indicates the dot number and the second one the component. The transformation of the c.m. momentum operators is analogous.
$`{\displaystyle \frac{P_{\alpha x}}{\sqrt{N_\alpha }}}`$ $`=`$ $`{\displaystyle \frac{i}{2}}\sqrt{{\displaystyle \frac{\stackrel{~}{\omega }_c^{}}{2}}}\left(a_{\alpha 1}^++a_{\alpha 2}^+a_{\alpha 1}a_{\alpha 2}\right)`$ (20)
$`{\displaystyle \frac{P_{\alpha y}}{\sqrt{N_\alpha }}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{\stackrel{~}{\omega }_c^{}}{2}}}\left(a_{\alpha 1}^+a_{\alpha 2}^++a_{\alpha 1}a_{\alpha 2}\right)`$ (21)
The cyclotron frequency is $`\omega _c^{}=B/m^{}c`$ and $`\stackrel{~}{\omega }_c^{}=\sqrt{\omega _c^2+4\omega _0^2}`$. Firstly, it is clear that the Hamiltonian in these ladder operators does not show an explicit $`N_\alpha `$– dependence anymore (apart from that implicit in the dynamical matrix). This implies that the eigenvalues of the Hamiltonian do not depend on those $`N_\alpha `$ explicitly seen in (8). Secondly, the commutators of the ladder operators are not influenced by the $`N_\alpha `$–factors and agree with those of Bosons: $`[a_{\alpha i},a_{\alpha i}^+]=1`$ and all other commutators vanish. (This is because the commutators of the $`\stackrel{~}{𝐔}_\alpha `$ and $`\stackrel{~}{𝐏}_\alpha ^{}`$ agree with the commutators of the untilded quantities.) Now, the total Hamiltonian can be written in matrix notation in the following compact form
$$H=\left[𝒂^+𝒂\right]𝐇\left[\begin{array}{c}𝒂\\ 𝒂^+\end{array}\right]$$
(22)
where
$$\left[𝒂^+𝒂\right]=\left[a_{11}^+a_{12}^+a_{21}^+a_{22}^+|a_{11}a_{12}a_{21}a_{22}\right]$$
(23)
and $`\left[\begin{array}{c}𝒂\\ 𝒂^+\end{array}\right]`$ is the transposed and Hermitian conjugate of (23). The $`8\times 8`$ Hamiltonian matrix is not unique, but can be cast into the following form
$$𝐇=\left[\begin{array}{cc}\alpha \beta & \\ \beta ^{}\alpha ^{}& \end{array}\right]\text{with}\alpha ^+=\alpha ;\beta ^T=\beta $$
(24)
consisting of the $`4\times 4`$ matrices
$$\alpha =\frac{1}{2}\left[\begin{array}{cc}\omega \mathrm{\hspace{0.33em}0}& \\ 0\omega & \end{array}\right]+\frac{1}{4\stackrel{~}{\omega }_c^{}}𝐄^+\left[\begin{array}{cc}\stackrel{~}{𝐂}_{11}𝐂_{12}& \\ 𝐂_{21}\stackrel{~}{𝐂}_{22}& \end{array}\right]𝐄$$
(25)
$$\beta =\frac{1}{4\stackrel{~}{\omega }_c^{}}𝐄^+\left[\begin{array}{cc}\stackrel{~}{𝐂}_{11}𝐂_{12}& \\ 𝐂_{21}\stackrel{~}{𝐂}_{22}& \end{array}\right]𝐄^{}$$
(26)
with $`𝐄=\left[\begin{array}{cc}\epsilon \mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ 0\epsilon & \end{array}\right]`$ and the $`2\times 2`$ matrices
$$\stackrel{~}{𝐂}_{kk}=𝐂_{kk}\frac{1}{2}\omega _0^2𝐈;\omega =\left[\begin{array}{cc}\omega _+\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ 0\omega _{}& \end{array}\right];\epsilon =\left[\begin{array}{cc}\mathrm{1\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}& \\ ii& \end{array}\right]$$
(27)
with
$$\omega _\pm =\sqrt{\omega _o^2+\left(\frac{\omega _c^{}}{2}\right)^2}\pm \left(\frac{\omega _c^{}}{2}\right)$$
(28)
Finding the eigenstates of the Boson Hamiltonian (22) is provided by mathematical physics and described in Ref. in full detail. The goal is to find a linear transformation $`\left[\begin{array}{c}𝒃\\ 𝒃^+\end{array}\right]=𝐀\left[\begin{array}{c}𝒂\\ 𝒂^+\end{array}\right]`$ which preserves Boson commutators and diagonalizes $`H`$. We shall only summarize the recipe here.
The eigenvalues are given by $`E_{n_1,n_2,n_3,n_4}=_k^{(1\mathrm{}4)}\left(n_k+\frac{1}{2}\right)\omega _k`$ with $`n_k`$ being non–negative integers and $`\omega _k=2\gamma _k`$ with $`\gamma _k`$ being the four positive eigenvalues of the matrix $`𝐇𝐉`$. The $`8\times 8`$ matrix $`𝐉=\left[\begin{array}{cc}𝐈\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ \mathrm{\hspace{0.33em}\hspace{0.33em}0}𝐈& \end{array}\right]`$ is made up of $`4\times 4`$ unit matrices. All eigenvalues of $`𝐇𝐉`$ come in pairs $`(\gamma _k,\gamma _k)`$.
The eigenfunctions of $`H`$ are constructed as usual for Bosons
$$|n_1,n_2,n_3,n_4>=\underset{k}{\overset{(1\mathrm{}4)}{}}\frac{\left(b_k^+\right)^{n_k}}{\sqrt{n_k!}}|0>$$
(29)
The four eigenvectors belonging to the positive eigenvalues are written in the form $`𝐱_k=\left[\begin{array}{c}𝒖_k\\ 𝒗_k\end{array}\right]`$. The column vectors of $`𝐀^+`$ are given by the vectors $`𝐱_k`$, and by the vectors $`\widehat{𝐱}_k=\left[\begin{array}{c}𝒗_k^{}\\ 𝒖_k^{}\end{array}\right]`$, which are the eigenvectors belonging to $`\gamma _k`$. The eigenvectors have to be properly orthonormalized $`𝐱_i^+𝐉𝐱_k=\delta _{i,k}`$. Without degeneracy, the orthogonality is guaranteed automatically. The inverse of this particular transformation is obtained from $`𝐀^1=𝐉𝐀^+𝐉`$ which shows that the linear transformation is not unitary (but unitary in a non– Euklidian metric).
## IV Oscillator strength
Optical oscillator strength between the states $`|n>=|n_1,n_2,n_3,n_4>`$ and $`|n^{}>=|n_1^{},n_2^{},n_3^{},n_4^{}>`$ for polarization in $`\eta =(x\text{or}y)`$ direction are defined as
$$f_{n,n^{};\eta }=2m^{}\omega _{n,n^{}}|<n|𝐔_{\eta ;tot}|n^{}>|^2$$
(30)
where $`\omega _{n,n^{}}`$ is the corresponding excitation energy, and $`𝐔_{\eta ;tot}`$ is the $`\eta `$–component of the total c.m. of the electrons in a unit cell (apart from a constant term). In formulae, this means $`𝐔_{tot}=\frac{N_1}{N_{tot}}𝐔_1+\frac{N_2}{N_{tot}}𝐔_2`$, where $`N_{tot}=N_1+N_2`$. After expressing the vectors $`𝐔`$ by ladder operators $`b_k,b_k^+`$ and using (29), we obtain the usual selection rules, i.e., only one quantum with energy $`\omega _k`$ can be absorped or emitted, so that we obtain only four absorption lines. The result for the oscillator strength for the four possible transitions $`(k=1\mathrm{}4)`$ and for $`\eta `$– polarization reads
$$f_{k,\eta }=\frac{m^{}\omega _k}{N_{tot}\stackrel{~}{\omega }_c^{}}|S_{k,\eta }|^2\{\begin{array}{c}(n_k+1)\\ n_k\end{array}\text{for}\{\begin{array}{c}absorption\\ emission\end{array}$$
(31)
where $`n_k`$ denotes the initial state, and
$`S_{k,x}`$ $`=`$ $`{\displaystyle \underset{i}{\overset{(1,2)}{}}}\sqrt{{\displaystyle \frac{N_1}{N_{tot}}}}\left(u_{ki}v_{ki}\right)+{\displaystyle \underset{i}{\overset{(3,4)}{}}}\sqrt{{\displaystyle \frac{N_2}{N_{tot}}}}\left(u_{ki}v_{ki}\right)`$ (32)
$`S_{k,y}`$ $`=`$ $`{\displaystyle \underset{i}{\overset{(1,2)}{}}}\sqrt{{\displaystyle \frac{N_1}{N_{tot}}}}(1)^{(i+1)}\left(u_{ki}+v_{ki}\right)+{\displaystyle \underset{i}{\overset{(3,4)}{}}}\sqrt{{\displaystyle \frac{N_2}{N_{tot}}}}(1)^{(i+1)}\left(u_{ki}+v_{ki}\right)`$ (33)
In the last definition, $`u_{ki}`$ and $`v_{ki}`$ for i=1…4 are the components of the vectors $`𝐮_k`$ and $`𝐯_k`$, respectively. The oscillator strength defined in (30) fulfill the following exact $`f`$–sum rule $`_kf_{k,\eta }=\frac{1}{N_{tot}}`$. It is worth pointing out that for equal electron numbers in either dot ($`N_1=N_2=N`$), the oscillator strength depends explicitly on $`N`$ (contrary to the optical excitation energies). In all figures presented below the oscillator strength are for $`N_1=N_2=N`$.
## V Results
Now the two simplest cases are discussed in more detail: two different circles and two identical, but rotated ellipses. The ratio of the two bare confinement frequencies involved in either case is 1:1.5 which means, that the two confinemnet frequencies in units of the mean frequency $`\omega _0`$ are 1.2 1nd 0.8. In our figures, all frequencies (energies) are given in units of the mean confinement frequency $`\omega _0`$ and the interaction parameters $`p`$ in units of $`\omega _0^2`$. The magnetic field is given in terms of the effective cyclotron frequency $`\omega _c^{}`$ in units of $`\omega _0`$ (upper scale) and in $`Tesla`$ (lower scale). The conversion between both scales is provided by
$$\omega _c^{}[\omega _0]=\frac{0.913410^2}{m^{}\omega _0[a.u.^{}]}B[Tesla]$$
(34)
In our figures we used $`\omega _0=0.2a.u.^{}=2.53meV`$ and $`m^{}`$ of GaAs for this conversion. (We want to stress that this parameter choice effects only the magnetic field scale and not the curves.) The definitions of the interaction parameters (16) for GaAs in more convenient units reads
$$p_i[\omega _0^2]=\frac{2.2610^7N_i}{(n.n.distance[\AA ])^3\left(\omega _0[meV]\right)^2}$$
(35)
(For a more detailed discussion of order– of – magnitude estimates see Ref. .)
For two different circular dots with bare confinement frequencies $`\omega _1`$ and $`\omega _2`$ and $`N_1=N_2`$, the absorption spectrum and the oscillator strength are shown in Fig.2. Although all absorption lines are effected by the dot interaction (represented by the interaction parameter $`p`$), and all modes are optically active, there is no qualitative effect of interaction in the position of the absorption lines. The reason can be understood easily. In this particular case, the four eigenmodes can be calculated analytically providing
$$\omega _{1,2,3,4}=\sqrt{\omega _{eff,i}^2+\left(\frac{\omega _c^{}}{2}\right)^2}\pm \left(\frac{\omega _c^{}}{2}\right),(i=1,2)$$
(36)
where
$$\omega _{eff,1,2}^2=\frac{(\omega _1^2+\omega _2^2)}{2}+\frac{(p_1+p_2)}{2}d\pm \sqrt{\left[\frac{(\omega _1^2+\omega _2^2)}{2}+\frac{(p_1+p_2)}{2}d\right]^2(\omega _1^2p_2d+\omega _2^2p_1d+\omega _1^2\omega _2^2)}$$
(37)
(The upper and lower sign belongs to $`\omega _{eff,1}`$ and $`\omega _{eff,2}`$, respectively). Consequently, if we had to interpret an experimental spectrum, we could do this using the formula (36) for non– interacting dots, but with the effective (i.e. interaction affected) confinement parameters defined in (37). Only if we take the intensities into account, we see some qualitative effect. Whereas for non– interacting dots (with $`p=0`$) and for $`B=0`$ the oscillator strength of all modes agree (for a single oscillator, $`f`$ is independent of the oscillator frequency), there is a large difference for interacting dots at $`p=0.5`$. This large difference can be understood as follows. In the limit $`p\mathrm{}`$, the upper pair of modes develops into the spurious Brillouin zone boundary mode, which has vanishing oscillator strength and the sum rule has to be fulfilled only by the lower pair (see also the discussion below).
In Fig.2 both dot species bare the same number of electrons. Therefore, only one interaction parameter $`p`$ is involved. Calculations with different $`N_i`$ (and $`p_i`$) do not show any qualitative difference. In the limit of large $`p`$ (and equal electron numbers) we obtain from (37)
$$\omega _{eff,1,2}^2=\frac{(\omega _1^2+\omega _2^2)}{2}+\{\begin{array}{c}2pd\\ 0\end{array}\pm \frac{(\omega _1^2\omega _2^2)}{8pd}+O(p^3)$$
(38)
Consequently, the square of the smaller effective confinement frequency (which is the only one giving rise to modes with a finite oscillator strength for large $`p`$) approaches the mean value of both squared bare confinement frequencies, whereas the larger one grows continously for large $`p`$.
In Fig.3 and 4b we show the results for two identical, but mutually rotated, elliptical dots. Without dot interaction ($`p=0`$), we have two doubly degenerate lines. With increasing interaction strength, we observe a splitting of degenerate modes and an anti-crossing behavior for finite $`B`$. As in the case of circular dots, the oscillator strength at $`B=0`$ for non– interacting dots ($`p=0`$) agree for all four modes. The dot interaction lifts this degeneracy. Additionally, we observe at $`p=0.5`$ that the oscillator strength in the limits of small and large magnetic fields is considerable only for two of the modes, except in the gap region, where three modes contribute. By comparison of Fig.s 3a and 4b we see that the magnetic field for minimum gap (between the second and third mode) increases with increasing $`p`$, whereas the gap width decreases. Consequently, the location and width of the gap provides information on the interaction strength.
By comparison of Fig.s 2 and 3 with Fig.4, and more clearly by consideration of formula (38) and Fig.5, it becomes clear that in either case the lower pair of degenerate modes at $`B=0`$ converges to a constant (the mean square bare confinement frequency $`\sqrt{(\omega _1^2+\omega _2^2)/2}`$, which amounts to $`1.02\omega _0`$ in our numerical example). Even for finite $`B`$, there are two branches, which converge to a finite ($`B`$-dependent) value for $`p\mathrm{}`$, or in other words, which become independent of $`p`$ in this limit. At first sight this looks surprising because the e e interaction does not show any saturation, if we increase the interaction parameter, but it continues to compress the dot state. However, there is a simple visual explanation for this feature: Generally, the dot interaction adds an additional second order contribution to the confinement, which has the same symmetry as the lattice, i.e. it is circular for a cubic lattice. For large $`p`$, this additional term outweighs the bare confinement, and the effective confinement in both dots becomes isotropic and equal. Thus, we approach the case of a lattice of identical dots, for which a pair of Kohn modes exists. Because these Kohn modes do not exactly agree with the modes of noninteracting dots, we call them pseudo Kohn modes. In a sense, the Generalized Kohn Theorem reentries for dot lattices with strong interdot interaction. In other lattices with lower symmetry, the effective confinement in the strong interaction limit might be elliptical, leading pseudo Kohn modes with a gap at $`B=0`$. The other pair of modes (which diverge for $`p\mathrm{}`$) turns into the in-folded modes at the Brillouin zone corner (because the units cell halves if all dots become equivalent). These modes become spurious in the long wavelength and the large–$`p`$ limit and the oscillator strength of them converge to zero.
In Fig.s 3a and 4b we observe an additional qualitative effect of dot interaction. For isolated elliptical dots we expect a gap between the two excitation branches at $`B=0`$. However, for larger $`p`$ only the pseudo Kohn mode might be observable, because the oscillator strength of the BZ boundary mode decrease rapidly. On the other hand, the two lower modes for finite $`p`$ develope out of the degenerate lower mode for $`p=0`$, whereby the degeneracy at $`B=0`$ survives. Therefore, at $`B=0`$ it looks as if we had a circular dot. The closing of the gap between the two most intensive branches at $`B=0`$ is not a gradual effect proceeding with increasing $`p`$, but initiated by symmetry. (For a deeper understanding see also the additional figures in Ref. \[Taut-condmat\].)
## VI Summary
We have shown that breaking the GKTh by constructing quantum dot lattices with at least two different dot confinements per unit cell has experimentally observable consequences. Generally speaking, there are no Kohn modes, i.e. interaction independent modes, anymore. In both of the considered cases, the degeneracy in the FIR intensities at $`B=0`$ between the upper and lower absorption lines is lifted due to dot interaction. For two mutually rotated elliptical dots (per cell), we observe also a splitting of formerly degenerate absorption frequencies and the appearance of an anticrossing. For two different circular dots no qualitative effect of e e interaction in the absorption frequencies is observed. Instead, the absorption spectrum can be mimiced by two noninteracting dots with modified (effective) confinements. We also pointed out that an extensively strong interaction destroys the effect of interaction by producing pseudo– Kohn modes. Although this limit cannot be reached experimentally, it might be important to take this tendency into consideration.
Only in the case of two circular dots there is a simple analytical closed form solution. However, with the formulae presented above, the absorption frequencies and oscillator strength for any cubic lattice with two harmonic dot species can be easily calculated. The only numerical task is to find the eigenvalues of an explicitly given non– Hermitian $`8\times 8`$ matrix and to perform a special sum over the eigenvector components.
## VII Acknowledgment
I am indebted to D.Heitmann, J.Kotthaus, and H.Eschrig, and their groups, as well as G.Paasch for helpful discussion.
|
warning/0007/hep-ph0007081.html
|
ar5iv
|
text
|
# Transverse lattice
## 1 Introduction
A light-front hamiltonian has wavefunctions initialised on a null plane $`x^+=(x^0+x^3)/\sqrt{2}=0`$. They are manifestly Lorentz-boost invariant and therefore suitable for highly relativistic problems. Indeed, many high energy QCD scattering or decay processes factorize into a non-perturbative light-front wavefunction convoluted with a perturbative scattering kernel . The calculation of light-front wavefunctions in QCD is not easy, however. One has to address the usual problems of confinement and chiral symmetry breaking in an unusual theoretical context, where renormalisation issues are non-standard and unavoidably non-perturbative.
Path integral lattice gauge theory simulation has been the most comprehensive method to compute hadronic observables from first principles QCD. Some indirect information on light-front wavefunctions has also been obtained from this method, but far less than one would like. It is therefore natural to attempt a light-front hamiltonian quantisation of lattice gauge theory. Treating $`x^+`$ as canonical time and $`\{x^{}=(x^0x^3)/\sqrt{2},x^1,x^2\}`$ as the spatial variables, $`x^+`$ must be continuous for a light-cone hamiltonian formulation. A high-energy lattice cut-off can only reasonably be applied to the transverse spatial directions $`𝐱=\{x^1,x^2\}`$, since large values of the momentum $`p^+`$ conjugate to $`x^{}`$ correspond to small energies $`p^{}`$ conjugate to $`x^+`$. A high energy cut-off may be applied to $`x^{}`$ by making it periodic, $`x^{}x^{}+L`$. For a state of total longitudinal momentum $`P^+`$, the choice $`L=2\pi K/P^+`$, with $`K`$ a fixed integer, is sometimes convenient (discrete light-cone quantisation (DLCQ) ). Naively, one might try to obtain the corresponding transverse lattice gauge theory by taking the small distance continuum limit of Wilson’s lattice gauge theory in the $`x^+`$ and $`x^{}`$ directions. However, there are some difficulties associated with this; in particular the light-cone quantisation of the resulting coupled gauged non-linear sigma models at each transverse lattice site is awkward . Bardeen and Pearson suggested that, for large values of the transverse lattice spacing $`a`$, a linear sigma model approximation could be used. Instead of formulating the lattice part of the theory in terms of link variables $`USU(N)`$, one uses link variables $`MGL(N)`$, where the complex matrices $`M`$ transform in the same way under gauge transformation. Use of such linearized variables is expected to be efficient on coarse lattices, as they represent the important degrees of freedom at such scales. Further discussion of the origin and meaning of these ‘colour-dielectric’ variables can be traced in ref. .
Thus, in transverse lattice gauge theory the Lorentz indices $`\mu ,\nu \{0,1,2,3\}`$ are split into LF indices $`\alpha ,\beta \{+,\}`$ and transverse indices $`r,s\{1,2\}`$. One has link variables $`M_r(𝐱)`$ associated with the link from $`𝐱`$ to $`𝐱+a\widehat{𝐫}`$ on a square transverse lattice, together with continuum $`SU(N)`$ gauge potentials $`A_\alpha (𝐱)`$ and Dirac fermions $`\mathrm{\Psi }(𝐱)`$ associated with a site $`𝐱`$. These variables transform under transverse lattice gauge transformations $`VSU(N)`$ as
$`A_\alpha (𝐱)`$ $``$ $`V(𝐱)A_\alpha (𝐱)V^{}(𝐱)+\mathrm{i}\left(_\alpha V(𝐱)\right)V^{}(𝐱)`$
$`M_r(𝐱)`$ $``$ $`V(𝐱)M_r(𝐱)V^{}(𝐱+a\widehat{𝐫})`$ (1)
$`\mathrm{\Psi }(𝐱)`$ $``$ $`V(𝐱)\mathrm{\Psi }(𝐱).`$ (2)
While a lattice gauge theory at large $`a`$ may have simple solutions when formulated in terms of disordered variables $`M`$, this will cease to be true when $`a`$ is made small. Although one may only be interesting in solving the theory at large $`a`$ because of this, traditionally the only way to derive the appropriate effective theory was to start at small $`a`$ and ‘integrate out’ short distance degrees of freedom. This problem unfortunately remains largely unsolved in the present case. A more radical approach attempts to remove cut-off artifacts already at large $`a`$, by tuning the effective theory to respect the continuum symmetries (gauge, Poincaré, chiral) required of the theory. This is the method that will be followed here.
The hamiltonian will consist of all operators invariant under lattice gauge symmetries (2) and Poincaré symmetries unviolated by the cut-offs. We will always have in mind that the transverse lattice spacing $`a`$ is the only cut-off to which this applies, all others, such as DLCQ, having been extrapolated (in principle). As a result, we may also use dimensional counting in the $`x^\pm `$ co-ordinates to limit the number of allowed operators. We search for a domain of coupling constants<sup>3</sup><sup>3</sup>3We must allow these couplings to be functions of longitudinal momentum $`p^+`$ in general . where Poincaré symmetries violated by the cut-off $`a`$ are restored. In a general lattice gauge theory, gauge and Poincaré symmetry alone is not sufficient to yield a trajectory of couplings flowing to the continuum QCD. However, the transverse lattice gauge theory is actually ‘half-continuum’ ($`A^+`$ and $`A^{}`$ appear explicitly), so it is inconceivable that anything else is obtained provided Lorentz invariance can be convincingly demonstrated. Symmetries that may be ‘spontaneously broken’, such as chiral symmetry, are treated in an unconventional way in light-front quantization. Because cut-offs may, and usually are chosen to, render the vacuum state trivial, symmetries that would conventionally be referred to as spontaneously broken will appear explicitly broken in the light-front hamiltonian. The behaviour of the explicit symmetry-breaking couplings is governed by the underlying symmetry invariance and, in the case of ‘dynamical symmetry breaking’, by stability of the vacuum.
In a real calculation at a given lattice spacing $`a`$ one must make further approximations, both on the Hamiltonian and its Hilbert (Fock) space. The Fock space at fixed lattice spacing may be truncated with the DLCQ cut-off $`K`$ and also a Tamm-Dancoff cut-off on the maximum number of quanta in a state. In principle the latter two cut-offs should be extrapolated, as has been done in studies of glueballs , if we are to follow the line of argument above. The QCD hamiltonian may be expanded in gauge-invariant powers of $`M`$ and the quark fields $`\mathrm{\Psi }`$. The intuitive justification for this is that both should behave like massive degrees of freedom (at large $`a`$); the former because of its association with the colour-dielectric mechanism of confinement and the latter because of spontaneous chiral symmetry breaking. We may also assume some transverse locality by expanding the hamiltonian in $`a𝐏`$, where $`𝐏`$ is transverse momentum. All these approximations have a physical justification and may be systematically relaxed. A final couple of approximations, which may be made for convenience, are to keep as many Poincaré generators as possible in kinematic form (independent of interactions and their renormalisation) and to neglect non-trivial longitudinal momentum dependence of coupling ‘constants’. The consequences of this will show up in the extent to which symmetry can be restored. The above approximations provide a general prescription for reducing the allowed operators in the hamiltonian down to a finite set, which may be systematically enlarged, and reducing the Hilbert space to a finite dimension, which may be extrapolated. The reward for a non-standard quantization with non-standard variables and approximations will be a dramatically simplified hadronic wavefunction.
## 2 Light meson calculations
Let us consider the first non-trivial approximation to the problem of meson boundstates in transverse lattice QCD. Applying the previous considerations, the leading-order lagrangian for a single quark flavour is
$`L`$ $`=`$ $`{\displaystyle \underset{𝐱}{}}{\displaystyle 𝑑x^{}\underset{\alpha ,\beta =+,}{}\underset{r=1,2}{}}{\displaystyle \frac{1}{2G^2}}\mathrm{Tr}\left\{F^{\alpha \beta }F_{\alpha \beta }\right\}`$
$`+\mathrm{Tr}\left\{[(_\alpha +iA_\alpha (𝐱))M_r(𝐱)iM_r(𝐱)A_\alpha (𝐱+a\widehat{𝐫})][\mathrm{h}.\mathrm{c}.]\right\}`$
$`\mu ^2\mathrm{Tr}\left\{M_rM_r^{}\right\}+i\overline{\mathrm{\Psi }}\gamma ^\alpha (_\alpha +iA_\alpha )\mathrm{\Psi }\mu _F\overline{\mathrm{\Psi }}\mathrm{\Psi }`$
$`+i\kappa _A\left(\overline{\mathrm{\Psi }}(𝐱)\gamma ^rM_r(𝐱)\mathrm{\Psi }(𝐱+a\widehat{𝐫})\overline{\mathrm{\Psi }}(𝐱)\gamma ^rM_r^{}(𝐱a\widehat{𝐫})\mathrm{\Psi }(𝐱a\widehat{𝐫})\right)`$
$`+\kappa _S\left(\overline{\mathrm{\Psi }}(𝐱)M_r(𝐱)\mathrm{\Psi }(𝐱+a\widehat{𝐫})+\overline{\mathrm{\Psi }}(𝐱)M_r^{}(𝐱a\widehat{𝐫})\mathrm{\Psi }(𝐱a\widehat{𝐫})\right)`$
The Hilbert space of the corresponding light-front hamiltonian will also be expanded in fields. For the present calculation we keep only the $`|\overline{\mathrm{\Psi }}(𝐱)\mathrm{\Psi }(𝐱)>`$ and $`|\overline{\mathrm{\Psi }}(𝐱)M_r(𝐱)\mathrm{\Psi }(𝐱+a\widehat{𝐫})>`$ states and their translates. Thus, in this ‘one-link’ approximation, a quark and antiquark can be at the same transverse lattice site, or separated by one link.<sup>4</sup><sup>4</sup>4This approximation to the Hamiltonian and Fock space was first considered in ref.. Those authors subsequently treated it as a phenomenological model, fixing coupling constants by hand and comparison to experimental mass ratios. However, see also Seal’s talk in these proceedings. This is the most severe approximation that still allows a meson to propagate on the transverse lattice. The calculations I have done so far have been for fixed $`K`$. This cut-off (at least) must eventually be extrapolated.
Using the chiral representation we decompose $`\mathrm{\Psi }^{}=(u_+^{},v_+^{},v_{}^{},u_{}^{})/2^{1/4}`$ into left (right) movers $`v`$ ($`u`$) with a helicity subscript. In light-cone gauge $`A_{}=0`$, one can eliminate non-dynamical degrees of freedom $`A_+`$ and $`v_\pm `$ at the classical level, to derive a light-front hamiltonian in terms of transverse polarizations only:
$`P^{}`$ $`=`$ $`{\displaystyle 𝑑x^{}\underset{𝐱}{}\frac{G^2}{4}\left(\mathrm{Tr}\left\{J^+\frac{1}{(\mathrm{i}_{})^2}J^+\right\}\frac{1}{N}\mathrm{Tr}\left\{J^+\right\}\frac{1}{(\mathrm{i}_{})^2}\mathrm{Tr}\left\{J^+\right\}\right)}`$ (3)
$`+{\displaystyle \frac{\mu _F^2}{2}}\left(F_+^{}{\displaystyle \frac{1}{\mathrm{i}_{}}}F_++F_{}^{}{\displaystyle \frac{1}{\mathrm{i}_{}}}F_{}\right)+\mu ^2{\displaystyle \underset{r=1}{\overset{2}{}}}\mathrm{Tr}\left\{M_rM_r^{}\right\}`$
$`F_\pm (𝐱)`$ $`=`$ $`u_\pm (𝐱)+{\displaystyle \frac{\kappa _S}{\mu _F}}{\displaystyle \underset{r}{}}\left(M_r(𝐱a\widehat{𝐫})+M_r^{}(𝐱)\right)u_\pm (𝐱)`$ (4)
$`\pm {\displaystyle \frac{\mathrm{i}\kappa _A}{\sqrt{2}\mu _F}}((M_1(𝐱a\widehat{𝐫})iM_2(𝐱a\widehat{𝐫}))`$
$`(M_1^{}(𝐱)\pm iM_2^{}(𝐱)))u_{}(𝐱)`$
$`J^+(𝐱)`$ $`=`$ $`i{\displaystyle \underset{r}{}}\left(M_r(𝐱)\underset{}{\overset{}{}}M_r^{}(𝐱)+M_r^{}(𝐱a\widehat{𝐫})\underset{}{\overset{}{}}M_r(𝐱a\widehat{𝐫})\right)`$ (5)
$`+u_+(𝐱)u_+^{}(𝐱)+u_{}(𝐱)u_{}^{}(𝐱)`$
With the exception of DLCQ self-energies, which are logarithmically divergent as $`K\mathrm{}`$ and needed for finite physical answers, the quartic terms in $`F^{}(\mathrm{i}_{})^1F`$ are dropped for the following calculation since the one-link approximation treats them asymmetrically . They may be included in higher link approximations and are essential for recovering parity invariance . Let us define $`\overline{G}=G\sqrt{(N^21)/N}`$, which has the dimensions of mass, and introduce the dimensionless variables
$$mb=\frac{\mu }{\overline{G}};mfs1=\left(\frac{\mu _F^{(1)}}{\overline{G}}\right)^2;mf2=\frac{\mu _F^{(2)}}{\overline{G}};$$
(6)
$$ka=\frac{\kappa _A}{\overline{G}};ks=\frac{\kappa _S}{\overline{G}}.$$
(7)
A Fock-sector dependent bare fermion mass $`\mu _F^{(i)}`$ has been allowed because the one-link approximation produces sector-dependent self-energies. $`i=1`$ is the $`\overline{\mathrm{\Psi }}\mathrm{\Psi }`$ sector and $`i=2`$ corresponds to the $`\overline{\mathrm{\Psi }}M\mathrm{\Psi }`$ sector. All $`N`$-dependence of the theory has been absorbed into $`\overline{G}`$ now.
Following a successful series of glueball studies with van de Sande , I investigated the coupling constant space (6)(7) for signals of enhanced Lorentz covariance using similar methods. The preliminary results I present here are mainly intended as an illustration of the procedure, and in no way should be considered definitive. Ideally, after DLCQ and Tamm-Dancoff cut-offs are extrapolated, the theory should exhibit enhanced Lorentz covariance on an approximately one-dimensional trajectory of couplings, along which $`a`$ varies. This trajectory represents the best approximation to the (unique) renormalised trajectory in the infinite-dimensional space of couplings. The estimate itself is not unique because it depends on the quantification of violations of Lorentz covariance. Before proceeding one must address chiral symmetry. There are two related issues. Firstly, cut-offs tend to break chiral symmetry in light-front quantisation because it is a dynamical symmetry. Secondly, spontaneous chiral symmetry breaking cannot appear as a non-trivial vacuum. The first issue means we should allow chiral-symmetry breaking terms in the cut-off hamiltonian. In principle they should be tuned so that the appropriate chiral Ward identities are satisfied. Spontaneous breaking would then have to appear as explicit breaking in the hamiltonian (see for an instructive example). Since the boundstate spectrum is at our disposal, a convenient criterion that addresses both issues at once is ’t Hooft’s anomaly matching condition . In a Lorentz-covariant theory, chiral symmetry is realised in the boundstate spectrum either as a goldstone boson (massless $`\pi `$) or massless composite fermions. Therefore, we could try to find the renormalised trajectory by optimising both Lorentz covariance and one or other of these chiral symmetry conditions. An incorrect choice for the manifestation of chiral symmetry will make Lorentz covariance difficult to obtain in a theory with a stable vacuum (no tachyons).
I used a $`\chi ^2`$ test with variables to measure the anisotropy of $`\pi `$ and $`\rho `$ dispersion relations and the deviation from zero of the pion to $`\rho `$ mass ratio.<sup>5</sup><sup>5</sup>5To complete the argument one should also study baryons. This requires more links and fermions for lattice propagation and chiral anomalies. The QCD scale was set from the (experimental) $`\rho `$ mass and the lattice spacing by demanding isotropy of the $`\pi `$’s dispersion in continuum and on-axis lattice directions. Solving the $`P^{}`$ eigenvalue problem for about 1000 combinations of the couplings (6)(7) and at various momenta, figure 1 shows the result for the one-link approximation at $`K=8`$ and a particular link-field mass $`mb=0.2`$. The transverse lattice spacing $`a=0.35`$fm (with a large error). This creates a problem in the one-link approximation, since it is smaller that true $`\pi `$’s radius. The $`\pi `$ is artificially squashed in the transverse direction or, in more technical language, the higher Fock state structure will be prone to cut-off artifacts. The $`|\overline{\mathrm{\Psi }}(𝐱)\mathrm{\Psi }(𝐱)>`$ sector may be less sensitive to cut-off artifacts, however.
Although the results are still quite crude and further work is needed to search couplings more finely and with more realistic criteria, extrapolate $`K`$ and add more links/quarks, some results are given here for experimentally known observables. The spin projections of the $`\rho `$ are not all degenerate because of residual breaking of Lorentz symmetry. Setting the averaged mass to 770 MeV, at the couplings with minimum $`\chi ^2`$ I find $`m_\pi =23`$ MeV, $`m_\rho (J_z=0)=642`$MeV, $`m_\rho (J_z=\pm 1)=898`$MeV. <sup>6</sup><sup>6</sup>6As a consistency check on the scale, the string tension calculated in the same approximation is found to be $`\sqrt{\sigma }=0.643m_\rho =495`$MeV. The $`\pi \rho `$ splitting is generated by the helicity-flip term $`\kappa _A`$ (see Perry’s talk in these proceedings). The $`|\overline{\mathrm{\Psi }}(𝐱)\mathrm{\Psi }(𝐱)>`$ component of the meson wavefunction is related to the leading order perturbative QCD expression for many exclusive processes . Fitting the transverse lattice pion wavefunction to the conformal expansion of this component, one finds
$`\varphi _\pi (x,Q^21GeV^2)`$ $`=`$ $`{\displaystyle \frac{2.653}{a}}x(1x)\{C_0^{3/2}+0.237C_2^{3/2}0.102C_4^{3/2}`$ (8)
$`0.05C_6^{3/2}+\mathrm{}\}.`$
The transverse scale $`1`$GeV is a rough estimate based on $`\pi /a`$, and $`C_n^{3/2}(12x^2)`$ are the appropriate Gegenbauer polynomials. Eq. (8) directly confirms that the conformal expansion is a good one. The overall normalisation yields $`f_\pi =101`$MeV compared with the experimental value $`f_\pi (exp.)=93`$MeV.
Using leading order perturbative QCD evolution to other scales $`Q^2`$, and assuming leading order perturbative QCD factorization, we can compare (8) with direct and indirect experimental measurements. The cleanest extraction from a $`\pi `$ form factor comes from $`\gamma \gamma ^{}\pi ^0`$; $`Q^2F_{\pi ^0}`$ is approximately constant in the range $`1GeV^2<Q^2<10GeV^2`$ and at the higher end $`Q^2F_{\pi ^0}(Q^2=8GeV^2)=0.16\pm 0.03`$GeV has been measured at CLEO-II . This compares with the theoretical result
$$Q^2F_{\pi ^0}(Q^2=8GeV^2)=\frac{4}{\sqrt{3}}_0^1𝑑x\frac{\varphi _\pi (x,Q^28GeV^2)}{x}=0.21\mathrm{GeV}$$
(9)
Direct tests of $`\varphi _\pi `$ have recently become possible from diffractive dissociation on a nucleus $`\pi +AA+\mathrm{jets}`$ . Figure 2 compares the quark amplitude with the one quoted in ref. as that which best fits the jet data after hadronization and experimental acceptance (see Ashery’s talk in these proceedings for an updated account of this experiment.) Although hadronization tends to wash out any fine structure in the quark amplitude, making true comparison between similar curves difficult, the theoretical curve seems a bit too peaked away from $`x=1/2`$. This is consistent with the result (9) being slightly too high. Of course there is no reason why the finite-$`K`$ calculation should give exactly the right result. Use of a nearly massless pion, ambiguous normalisation scale $`Q^2`$, leading order evolution from 1 to $`10GeV^2`$, etc. leads to further errors.
Inclusive processes are typically sensitive to higher Fock state structure. Here, we may not expect to do so well because of the severity of our cut-offs. This suspicion is confirmed by the standard $`x^\alpha (1x)^\beta `$ fits to the non-singlet quark probability distribution in the $`\pi `$, obtained from $`\pi N`$ Drell-Yan ; see fig.3. The deviation of the lattice results at large $`x`$ is due at least partly to cut-off artifacts. Relaxing the DLCQ and one-link approximation will allow some of the quark momentum at large $`x`$ to radiate into small $`x`$ quarks and gluons as more gluonic channels are opened. The gluons carry only about $`10\%`$ of the $`\pi `$ momentum at present — much lower than the accepted experimental value — confirming there is a problem with higher Fock states.
|
warning/0007/math0007028.html
|
ar5iv
|
text
|
# Lagrangian torus fibration and mirror symmetry of Calabi-Yau hypersurface in toric variety
## 1 Introduction
In this paper we give a construction of Lagrangian torus fibration for Calabi-Yau hypersurface in toric variety via the method of gradient flow. Using our construction of Lagrangian torus fibration, we are able to prove the symplectic topological version of SYZ mirror conjecture for generic Calabi-Yau hypersurface in toric variety. We will also be able to give precise formulation of SYZ mirror conjecture in general (including singular locus and duality of singular fibres).
The motivation of our work comes from the study of Mirror Symmetry. Mirror Symmetry conjecture originated from physicists’ work in conformal field theory and string theory. It proposes that for a Calabi-Yau 3-fold $`X`$ there exists a Calabi-Yau 3-fold $`Y`$ as its mirror manifold. The quantum geometry of $`X`$ and $`Y`$ are closely related. In particular one can compute the number of rational curves in $`X`$ by solving the Picard-Fuchs equation coming from variation of Hodge structure of $`Y`$.
Despite the great impact and success mirror symmetry conjecture brings to the understanding of geometry of Calabi-Yau manifolds and their moduli spaces, the fundamental question that in general given a Calabi-Yau manifold how to construct the corresponding mirror Calabi-Yau manifold is not clear at all from the original mirror conjecture. Although mirror manifolds are worked out in certain cases, the construction of mirror for general Calabi-Yau is still very elusive and mysterious. The most general construction so far was given by Batyrev for Calabi-Yau hypersurface in toric variety. From the toric geometry standing point, Batyrev propose that for Calabi-Yau hypersurface $`X`$ in the toric variety $`P_\mathrm{\Delta }`$ corresponding to a reflexive polyhedron $`\mathrm{\Delta }`$, the mirror should be the Calabi-Yau hypersurface $`Y`$ in the toric variety $`P_\mathrm{\Delta }^{}`$ corresponding to the dual reflexive polyhedron $`\mathrm{\Delta }^{}`$. Batyrev computed among other things the Hodge numbers of Calabi-Yau hypersurface $`X`$ and the mirror $`Y`$ and showed that they behave as predicted by mirror symmetry. Batyrev’s mirror construction includes many special cases of mirror construction discussed previously by many physicists and mathematicians. Later, there are lots of important work surrounding Batyrev mirror construction, including computing number of rational curves, etc.
In 1996 Strominger, Yau and Zaslow () proposed a geometric construction of mirror manifold via special Lagrangian torus fibration. According to their program (we will call it SYZ construction), a Calabi-Yau 3-fold should admit a special Lagrangian torus fibration. The mirror manifold can be obtained by dualizing the fibers. Or equivalently, the mirror manifold of $`M`$ is the moduli space of special Lagrangian 3-torus in $`M`$ with a flat $`U(1)`$ connection.
In a sense, SYZ construction appear to be more fundamental to mirror symmetry phenomenon and more classical than quantum mirror symmetry. More importantly, SYZ construction has the potential to explain the mathematical reason behind mirror symmetry. For example, SYZ construction gives us a possible way to construct the mirror manifold if we understand how to construct dual singular fibers.
Remark: The original SYZ mirror conjecture is rather sketchy in nature. Many detail knowledge of the fibraton like singular locus, singular fibres and duality of singular fibres are necessary to be worked out to formulate the precise SYZ mirror conjecture. Without the precise formulation, one would not really be able to construct the mirror manifold completely.
According to the SYZ construction, special Lagrangian submanifold and special Lagrangian fibration for Calabi-Yau manifolds seem to play very important roles in understanding mirror symmetry. However, despite its great potential in solving the mirror symmetry conjecture, our understanding on special Lagrangian submanifolds is very limited. The known examples are mostly explicit local examples or examples coming from $`n=2`$. There are very few example of special Lagrangian submanifold or special Lagrangian fibration for dimension higher than two. M. Gross, P.M.H. Wilson and N. Hitchin () did some important work in this area in recent years. On the other extreme, in , Zharkov constructed some non-Lagrangian fibration of Calabi-Yau hypersurface in toric variety. Despite all these efforts, SYZ construction still remains to be a beautiful dream to us.
For our discussion, we will relax the special Lagrangian condition, yet still keep the Lagrangian condition. We will consider Lagrangian torus fibration of Calabi-Yau manifolds. According to our discussion in , There are major differences between $`C^{\mathrm{}}`$-Lagrangian fibration and general Lagrangian fibration, let alone non-Lagrangian fibration. In a sense, $`C^{\mathrm{}}`$-Lagrangian fibration should already capture symplectic topological structure of the corresponding special Lagrangian fibration. We will construct Lagrangian torus fibration that exhibit the same topological structure as $`C^{\mathrm{}}`$-Lagrangian torus fibration. In particular, singular locus of the fibration will be of codimension 2. In we smooth out our Lagrangian fibration to a great extent, yet we still fall short of making it into a $`C^{\mathrm{}}`$-Lagrangian fibration everywhere.
In this paper, we will construct Lagrangian torus fibration of generic Calabi-Yau hypersurface in toric variety correponding to a reflexive polyhedron in complete generality. With these detailed understanding of Lagrangian torus fibration of generic Calabi-Yau hypersurface in toric variety, we will be able to prove the symplectic topological SYZ mirror conjecture for Calabi-Yau hypersurface in toric variety. More precisely
###### Theorem 1.1
For generic Calabi-Yau hypersurface $`X`$ and its mirror Calabi-Yau hypersurface $`Y`$ near their corresponding large complex limit and large radius limit, there exist corresponding Lagrangian torus fibrations
$$\begin{array}{ccccccc}X_{\varphi (b)}& & X& & Y_b& & Y\\ & & & & & & \\ & & \mathrm{\Delta }_v& & & & \mathrm{\Delta }_w\end{array}$$
with singular locus $`\mathrm{\Gamma }\mathrm{\Delta }_v`$ and $`\mathrm{\Gamma }^{}\mathrm{\Delta }_w`$, where $`\varphi :\mathrm{\Delta }_w\mathrm{\Delta }_v`$ is a natural homeomorphism, $`\varphi (\mathrm{\Gamma }^{})=\mathrm{\Gamma }`$, and $`b\mathrm{\Delta }_w\backslash \mathrm{\Gamma }^{}`$. The corresponding fibres $`X_{\varphi (b)}`$ and $`Y_b`$ are naturally dual to each other.
For detail notations and results, please refer to later sections of this paper.
Our work essentially indicate that Batyrev mirror construction, which was proposed purely from toric geometry stand point, can also be understood and justified by the SYZ mirror construction. This should give us greater confidence on SYZ mirror conjecture for general Calabi-Yau manifolds.
This paper is a sequel of a series of papers on Lagrangian torus fibration of quintic Calabi-Yau hypersurface in $`𝐏^\mathrm{𝟒}`$. In , we described a very simple and natural construction via gradient flow, which will in principle be able to produce Lagrangian torus fibration for general Calabi-Yau hypersurface in toric variety. For simplicity, we described the case of Fermat type quintic Calabi-Yau threefold familly $`\{X_\psi \}`$ in $`\mathrm{𝐂𝐏}^\mathrm{𝟒}`$
$$p_\psi =\underset{1}{\overset{5}{}}z_k^55\psi \underset{k=1}{\overset{5}{}}z_k=0$$
near the large complex limit $`X_{\mathrm{}}`$
$$p_{\mathrm{}}=\underset{k=1}{\overset{5}{}}z_k=0.$$
in great detail. Most of the essential features of the general case already show up there. We also discussed expected special Lagrangian torus fibration structure, especially we computed the monodromy transformations of the expected special Lagrangian fibration and also discussed the expected singular fibre structure implied by monodromy information in this case. Then we compared our Lagrangian fibration constructed via gradient flow with the expected special Lagrangian fibration. Finally, we discussed its relevance to mirror construction for Calabi-Yau hypersurface in toric variety.
In order to apply our gradient flow construction effectively, it is necessary to address several technical problems concerning to the flow. First of all, the gradient flow in our situation is highly non-conventional gradient flow. The critical points of our function are usually highly degenerate and often non-isolated. Worst of all our function is not even smooth (it has infinities along some subvariety). In we discussed the local behavior of our gradient flow near critical points and infinities of our function, also the dependence on the metric to make sure that they behave the way we want.
Secondly, SYZ construction requires special Lagrangian fibration. The corresponding singular locus is expected of codimension 2. The Lagrangian torus fibration naturally produced by the gradient flow and the natural Lagrangian torus fibration of the large complex limit has singular locus of codimension 1. As described in , it is not hard to deform our Lagrangian fibration to a $`C^{\mathrm{}}`$ non-Lagrangian fibration with the same topological structure as the expected $`C^{\mathrm{}}`$ special Lagrangian fibration. To get a Lagrangian fibration of the same topological structure is more delicate. In , we are able to squeeze the codimension 1 singular locus to codimension 2 by symplectic geometry technique to get Lagrangian torus fibration with the same topological structure as the expected special Lagrangian fibration.
Thirdly, our gradient flow approach naturally produces piecewise smooth (Lipschitz) Lagrangian fibration. In our opinion, for symplectic topological aspect of SYZ construction, one should consider $`C^{\mathrm{}}`$ Lagrangian fibration. This is indicated by somewhat surprising fact, discussed in , that a general piecewise smooth (Lipschitz) Lagrangian fibration can not be smoothed to a $`C^{\mathrm{}}`$ Lagrangian fibration by small perturbation. More precisely, the singular locus of a general piecewise smooth (Lipschitz) Lagrangian fibration is usually of codimension 1 while the singular locus of the corresponding $`C^{\mathrm{}}`$ Lagrangian fibration necessarily has codimension 2. In , we constructed many local examples of piecewise smooth (Lipschitz) Lagrangian fibrations that are not $`C^{\mathrm{}}`$. In , we discussed method to deform a piecewise smooth (Lipschitz) Lagrangian fibration to a $`C^{\mathrm{}}`$ Lagrangian fibration. Although we still fall short to make our Lagrangian fibration $`C^{\mathrm{}}`$ everywhere.
With all the techniques developed in , in we are able to construct Lagrangian torus fibration with codimension 2 singular locus for general quintic Calabi-Yau hypersurface in $`\mathrm{𝐂𝐏}^\mathrm{𝟒}`$. Here the main difficulty do not come from the gradient flow. As we mentioned before that the gradient flow will always give us a Lagrangian fibration, even for more general situation. But as we learn from the Fermat type quintic case, Lagrangian fibration so constructed is usually not of the topological type of the expected $`C^{\mathrm{}}`$ Lagrangian (or special Lagrangian) torus fibration. We need to modify the fibration. Unlike the case of Fermat type quintic Calabi-Yau, where we can always modify the Lagrangian fibration to the expected topological type due to its nicely behaved singular locus. It is hopeless to do the same for general quintic Calabi-Yau also due to generally badly behaved singular locus. It turns out that for general quintic near the large complex limit (in certain sense), singular locus will be much better behaved than general quintic Calabi-Yau. And we can do the same thing as in the fermat type Calabi-Yau case to get the desired Lagrangian fibration. In we also constructed Lagrangian torus fibration for the mirror of quintic Calabi-Yau.
With all these construction of Lagrangian torus fibrations of general quintic Calabi-Yau hypersurfaces and their mirror, we are able to prove the symplectic topological SYZ mirror conjecture for generic quintic Calabi-Yau hypersurface. In we also discussed a very simple computation of monodromy of Lagrangian torus fibration and singular fibres that will appear in generic Lagrangian torus fibration. These discussion in principle give us a general construction of mirror Calabi-Yau from a given generic Lagrangian torus fibration of a Calabi-Yau manifold.
In this paper, we generalize our work in to the case of general Calabi-Yau hypersurface in toric variety with respect to a reflexive polyhedron, which is exactly the situation of the Batyrev dual polyhedron mirror construction. Compared to the quintic case, in general toric case, usually both the Kähler moduli and the complex moduli of a Calabi-Yau hypersurface are non-trivial. The construction of the Lagrangian torus fibration has to depend on both the Kähler form and the complex structure of the Calabi-Yau hypersurface. We also need the most general monomial-divisor map to carry out the discussion of symplectic topological SYZ mirror construction for general Calabi-Yau hypersurface in toric variety.
In section 2, we start by reviewing basic facts from toric geometry necessary for our discussion. Kähler moduli of the Calabi-Yau hypersurface is discussed in section 3. Our symplectic topological SYZ construction is based on the monomial-divisor mirror map. The monomial-divisor mirror map of Aspinwall, Greene and Morrison was first constructed in with certain restriction. In sections 4 and 5, using slicing theorem, we generalize their construction of monomial-divisor mirror map to the case of Calabi-Yau hypersurface in toric variety with respect to general reflexive polyhedron. We also pin down the suitable toric compactification of the complex moduli near the large complex limit.
In section 6, we discuss the construction of the Lagrangian torus fibration of Calabi-Yau hypersurfaces in toric variety that depends on both the Kähler form and the complex structure of the Calabi-Yau hypersurface. At first, the gradient flow produce a Lagrangian torus fibration with codimension 1 singular locus for general Calabi-Yau hypersurface. With the miraculous fact that near the large complex and the large radius limit the singular locus behave like a fattening of a suitable graph, using techniques developed in , we are able to modify our construction to get a Lagrangian torus fibration with codimension 2 singular locus.
The symplectic topological SYZ mirror conjecture is proved in section 7 and 8. There are two major ingradients, the identification of bases (topologically $`S^3`$) and the duality relation of the fibres of the Lagrangian torus fibrations of the Calabi-Yau hypersurface of the toric variety and its mirror. The bases of the two fibrations are naturally boundaries of the two convex polyhedrons. Their identification can be understood purely from the combinatoric properties of the two convex polyhedrons. This identification is discussed in section 7. In section 8, we discuss the duality relation of the fibres of the Lagrangian torus fibrations of the Calabi-Yau hypersurface of the toric variety and its mirror.
As mentioned in the previous remark, the original SYZ conjecture is not enough to construct the mirror manifold completely. The main problem is the lack of knoledge of singular locus, singular fibres and duality of singular fibres. With our construction of Lagrangian torus fibration in and this paper, we have much better knowledge of the structure of singular locus, generic singular fibres and how singular fibres are dual to each other. This detailed knowledge enable us to precisely formulate the SYZ construction in section 9, which will enable us to construct mirror manifold topologically if a SYZ Lagrangian torus fibration of a Calabi-Yau manifold is given.
Note: Partly based on our previous monodromy compuation in , at around the same time as our paper , in his intersting work , M. Gross was able to use the classical knowledge of resolution of singularity of $`𝐂^3/(𝐙_5\times 𝐙_5)`$ and monodromy information to figure out the expected fibration structure of certain quintic. This is in the same line with the idea in our previous paper to use monodromy information to figure out the expected fibration. Gross went one significant step further to compute the cubic form etc. of the constructed topological manifold and use theorems of C.T.C. Wall to verify that the manifold consructed is indeed diffeomorphic to quintic, which is very different from our approach, therefore constructing a torus fibration for certain quintic Calabi-Yau (not necessarily Lagrangian fibration). In our work , using gradient flow approach, we were able to naturally construct Lagrangian torus fibration for general quintic Calabi-Yau and its mirror. Then we show that there is a natural duality relation between the fibres, therefore proving the symplectic SYZ completely for quintic Calabi-Yau.
## 2 Background in toric geometry
In this section, we will first review some basic constructions in toric geometry, especially the toric variety associated with a reflexive polyhedron. Then we will discuss possible singularities of simplicial toric varieties. For our gradient flow method to work effectively, this knowledge of singularities is very crucial. Finally, we will discuss some facts necessary for getting toric quotients and performing toric blow up.
### 2.1 Basic
Consider a rank $`r`$ lattice $`N𝐙^r`$ and its dual lattice $`M=\mathrm{hom}_𝐙(N,𝐙)`$ with the natural pairing:
$$,:M\times N𝐙.$$
A fan in $`N`$ is a nonempty collection $`\mathrm{\Sigma }`$ of strongly convex ($`\sigma (\sigma )=\{0\}`$) rational polyhedral cones in $`N`$, such that subcones of a polyhedral cone $`\sigma \mathrm{\Sigma }`$ is also in $`\mathrm{\Sigma }`$.
For a cone $`\sigma \mathrm{\Sigma }`$, the dual cone is defined as
$$\sigma ^{}=\{mM|m,n0\mathrm{for}\mathrm{all}n\sigma \}.$$
Let $`U_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])`$. An inclusion $`\sigma \sigma ^{}`$ induce an inclusion $`U_\sigma U_\sigma ^{}`$. Gluing $`U_\sigma `$ together for $`\sigma \mathrm{\Sigma }`$, we get the toric variety $`P_\mathrm{\Sigma }`$. In particular, $`T=U_{\{0\}}=\mathrm{Spec}(𝐂[M])N_𝐂/N=N_𝐙𝐂^{}`$ is a complex $`r`$-torus. There is a natural action of $`T=U_{\{0\}}`$ on $`P_\mathrm{\Sigma }`$ and $`U_{\{0\}}`$ is the unique Zariski dense orbit in $`P_\mathrm{\Sigma }`$.
Let’s fix $`\sigma \mathrm{\Sigma }`$ and try to understand one affine piece $`U_\sigma `$ better. Now the problem can be rephrased as given an $`r`$ dimensional rational convex polyhedral cone $`\sigma ^{}M`$, discuss the structure of $`U_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])`$. For any cone $`\tau M`$, denote $`X_\tau =\mathrm{Spec}(𝐂[\tau ])`$. (Relate to the old notation $`X_\sigma ^{}=U_\sigma `$.)
For a subcone $`\tau \sigma ^{}`$, There is a seemingly unnatural semi-group map $`r:\sigma ^{}\tau `$ ($`r|_\tau =id`$, $`r|_{\sigma ^{}\backslash \tau }=0`$) that induce a very natural embedding $`X_\tau X_\sigma ^{}=U_\sigma `$. $`r`$ is just the corresponding restriction map. On the other hand, the natural embedding $`\tau \sigma ^{}`$ induces a seemingly rather unnatural projection $`X_\sigma ^{}=U_\sigma X_\tau `$. (This projection can be understood as every torus in $`X_\sigma ^{}=U_\sigma `$ is projected to the largest adjacent torus in $`X_\tau `$. In particular, fibre dimension could jump.) Let $`\overline{\tau }`$ denote the span of $`\tau `$, and $`T_\tau =X_{\overline{\tau }}`$. We have
Let $`\mathrm{\Sigma }(k)`$ denote the collection of $`k`$-dimensional cones in the fan $`\mathrm{\Sigma }`$, then for any $`\sigma \mathrm{\Sigma }(1)`$, $`X_\sigma ^{}U_\sigma P_\mathrm{\Sigma }`$ give us a $`T`$-invariant irreducible Weil divisor $`D_\sigma `$. Let $`e_\sigma `$ denote the unique primitive element in $`\sigma `$. Consider $`M`$ as the set of $`T`$-invariant meromorphic functions on $`P_\mathrm{\Sigma }`$, then for $`mM`$, $`m,e_\sigma `$ is the vanishing order of $`m`$ along $`D_\sigma `$. This important observation motivates the representation of line bundle by piecewise linear function on the fan. The most important line bundle is the anti-canonical bundle $`𝒪(D_\mathrm{\Sigma })`$, where
$$D_\mathrm{\Sigma }=\underset{\sigma \mathrm{\Sigma }(1)}{}D_\sigma $$
$`𝒪(D_\mathrm{\Sigma })`$ can be characterized by the piecewise linear function $`p_\mathrm{\Sigma }`$ that satisfy $`p_\mathrm{\Sigma }(e_\sigma )=1`$ for any $`\sigma \mathrm{\Sigma }(1)`$.
Given a fan $`\mathrm{\Sigma }`$, there are two naturally associated polyhedrons. The Newton polyhedron of $`𝒪(D_\mathrm{\Sigma })`$ defined as
$$\mathrm{\Delta }_\mathrm{\Sigma }=\{mM|m,np_\mathrm{\Sigma }(n),\mathrm{for}\mathrm{any}nN\}$$
The second polyhedron $`\mathrm{\Delta }_\mathrm{\Sigma }^{}`$ is the convex hull of $`e_\sigma `$ for $`\sigma \mathrm{\Sigma }(1)`$. Since $`p_\mathrm{\Sigma }(e_\sigma )=1`$ for $`\sigma \mathrm{\Sigma }(1)`$, if we think of $`M`$, $`N`$ as real vector spaces, the two real polyhedrons are naturally dual to each other. Namely
$$\mathrm{\Delta }_\mathrm{\Sigma }^{}=\{nN|m,n1,\mathrm{for}\mathrm{any}m\mathrm{\Delta }_\mathrm{\Sigma }\}$$
In our situation, we think of $`M`$, $`N`$ as two lattices. $`\mathrm{\Delta }_\mathrm{\Sigma }`$ and $`\mathrm{\Delta }_\mathrm{\Sigma }^{}`$ are two integral polyhedrons. In general, we only have $`\mathrm{\Delta }_\mathrm{\Sigma }`$ is dual of $`\mathrm{\Delta }_\mathrm{\Sigma }^{}`$, not vise versa. When the two integral polyhedrons are dual to each other, we call $`\mathrm{\Delta }_\mathrm{\Sigma }`$ reflexive.
A dimemsion $`r`$ cone $`\sigma `$ is called simplicial cone if it has exactly $`r+1`$ 1-dimensional subcones, if further more the primitive elements in those 1-dimensional subcones generate $`\sigma `$, then $`\sigma `$ is called primitive simplicial cone.
We can also introduce volume $`v_\sigma `$ of a simplicial cone $`\sigma `$ to be the volume of the parallelgram generated by generators of 1-dimensional subcones of $`\sigma `$. Then a simplicial cone $`\sigma `$ is primative if and only if $`v_\sigma =1`$.
$`P_\mathrm{\Sigma }`$ is usually not smooth, since $`\sigma \mathrm{\Sigma }`$ may not be simplicial. To make $`P_\mathrm{\Sigma }`$ smoother, we need to subdivide the fan $`\mathrm{\Sigma }`$. We usually do not want the polyhedron $`\mathrm{\Delta }_\mathrm{\Sigma }`$ to be affected by subdivision. The best way to archieve this is to consider subdivision $`\mathrm{\Sigma }^{}`$ with $`\mathrm{\Sigma }^{}(1)`$ consists of only the boundary integral points of $`\mathrm{\Delta }_\mathrm{\Sigma }^{}`$. This kind of subdivision is called crepant subdivision.
Clearly, through crepant subdivision, we can always make the cones in $`\mathrm{\Sigma }`$ simplicial. It is usually impossible to make them primitive in general only by crepant subdivision. Recall that $`P_\mathrm{\Sigma }`$ is smooth if and only if each $`\sigma \mathrm{\Sigma }`$ is primitive simplicial cone.
### 2.2 Toric varieties associated with a reflexive polyhedron
A good way to avoid many unpleasant features is to start with a reflexive polyhedron $`\mathrm{\Delta }M`$ (when both $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$ are integral). Fan $`\mathrm{\Sigma }`$ should be constructed in two steps. First construct all the normal cones, for a face $`\alpha `$ of the polyhedron $`\mathrm{\Delta }`$, define the normal cone of $`\alpha `$
$$\sigma _\alpha :=\{nN|m^{},nm,n\mathrm{for}\mathrm{all}m^{}\alpha ,m\mathrm{\Delta }\}$$
When taking $`\alpha `$ to be vertices of $`\mathrm{\Delta }`$, we get the top dimension cones. The collection of all the normal cones $`\sigma _\alpha `$ forms the anti-canonical fan $`\mathrm{\Sigma }_\mathrm{\Delta }`$ of $`\mathrm{\Delta }`$. In the second step, through crepant subdivision, we subdivide these cones to get a simplicial fan $`\mathrm{\Sigma }`$ with as many cones as possible. In this way, $`\mathrm{\Sigma }(1)`$ still consists of only the boundary integral points of $`\mathrm{\Delta }^{}`$. We will denote divisors come from the first step principle divisors and denote by $`\mathrm{\Sigma }(1)_p`$, divisors come from the second step exceptional divisors and denote by $`\mathrm{\Sigma }(1)_e`$.
Notice the second step is not unique. In this way, we are getting a collection of toric varieties $`P_\mathrm{\Sigma }`$ with common $`\mathrm{\Sigma }(1)`$ that is associated with reflexive polyhedron $`\mathrm{\Delta }`$. They differ only by codimension two birational modification.
A fan $`\mathrm{\Sigma }`$ so constructed is a maximal simplicial fan that is compatible with $`\mathrm{\Delta }`$, where $`\mathrm{\Sigma }`$ is called compatible with $`\mathrm{\Delta }`$ if $`\mathrm{\Sigma }`$ is a crepant subdivision of the anti-canonical fan $`\mathrm{\Sigma }_\mathrm{\Delta }`$. In this case, $`P_\mathrm{\Sigma }`$ is a crepant resolution of $`P_{\mathrm{\Sigma }_\mathrm{\Delta }}`$.
In our point of view, for the fan $`\mathrm{\Sigma }`$, the set of 1-dimensional cones $`\mathrm{\Sigma }(1)`$ is more fundamental. Let $`e_\sigma `$ be the primitive generator of $`\sigma \mathrm{\Sigma }(1)`$, then $`\sigma e_\sigma `$ defines an identification of $`\mathrm{\Sigma }(1)`$ with integral points in $`\mathrm{\Delta }^{}`$. Namely, $`\mathrm{\Sigma }(1)`$ contains the same information as $`\mathrm{\Delta }`$.
### 2.3 Subtoric varieties
Recall for $`\sigma \mathrm{\Sigma }`$, the restriction map $`r:\sigma ^{}\sigma ^{}`$ induces the embedding $`T_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])U_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])P_\mathrm{\Sigma }`$. $`dim_𝐂T_\sigma +dim\sigma =r`$. (Notice here we changed notation of $`T_\sigma `$ a little bit.) In this way, we have a decomposition of $`P_\mathrm{\Sigma }`$ into disjoint union of complex tori.
$$P_\mathrm{\Sigma }=\underset{\sigma \mathrm{\Sigma }}{}T_\sigma $$
$`T_\tau `$ is adjacent to $`T_\sigma `$ (namely, $`T_\tau \overline{T_\sigma }`$) if and only if $`\sigma `$ is a subcone of $`\tau `$. For $`\sigma \mathrm{\Sigma }`$, consider projection $`\pi _\sigma :NN_\sigma :=N/\overline{\sigma }`$. Then $`N_\sigma `$ is naturally dual to $`M_\sigma :=\sigma ^{}`$. We can define a new fan
$$\mathrm{\Sigma }_\sigma =\{\pi _\sigma (\tau )N_\sigma |\sigma \tau \mathrm{\Sigma }\}$$
The corresponding toric variety $`P_{\mathrm{\Sigma }_\sigma }`$ is the closure of $`T_\sigma `$ in $`P_\mathrm{\Sigma }`$.
On the other hand, the natural injection $`\sigma ^{}\sigma ^{}`$ induces the fibration $`U_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])T_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])`$.
###### Proposition 2.1
The fibres of the fibration $`U_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])T_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])`$ are naturally affine toric variety with fan $`\mathrm{\Sigma }^\sigma `$ consists of subcones of $`\sigma \overline{\sigma }=\mathrm{Span}(\sigma )`$.
It is interesting to understand the restriction of a line bundle on $`P_\mathrm{\Sigma }`$ to a subvariety $`P_{\mathrm{\Sigma }_\sigma }`$. As we know, a piecewise linear function $`p`$ on $`N`$ with respect to the fan $`\mathrm{\Sigma }`$ will give us a meromorphic section $`s_p`$ of the corresponding line bundle $`L_p`$. $`s_p`$ is non-vanishing (vanishing, has a pole) when restrict to the torus $`T_\sigma `$ if and only if $`p`$ restrict to the interior of $`\sigma `$ is zero (negative, positive). In particular, $`s_p`$ is holomorphic if and only if $`p0`$. To make restriction to $`P_{\mathrm{\Sigma }_\sigma }`$ meaningful, we would want section $`s_p`$ of $`L_p`$ to be holomorphic and non-vanishing along $`T_\sigma `$. Namely, we want $`p`$ restrict to zero on $`\sigma `$. Since $`p`$ is convex, it is easy to modify $`p`$ by a linear function $`mM`$ to ensure $`p0`$ and $`p|_\sigma =0`$. Under these conditions, $`p`$ will naturally induce a piecewise linear convex function $`p_\sigma `$ on $`N_\sigma `$ with respect to fan $`\mathrm{\Sigma }_\sigma `$. Now assume cones in $`\mathrm{\Sigma }`$ are all simplicial cones.
### 2.4 Singularities of a simplicial toric variety
As we know, $`P_\mathrm{\Sigma }`$ is an union of complex torus (orbits of $`T`$ action) that is parametrized by cones in the fan $`\mathrm{\Sigma }`$. In particular, $`r`$ dimensional cones $`\sigma \mathrm{\Sigma }(r)`$ give rise to zero dimensional orbits (fixed points of the $`T`$ action) $`q_\sigma `$. On the other hand, an $`r`$ dimensional cone $`\sigma \mathrm{\Sigma }(r)`$ also gives rise to an affine toric subvariety $`U_\sigma P_\mathrm{\Sigma }`$, which is naturally a neighborhood of $`q_\sigma `$. All this neighborhoods $`\{U_\sigma \}_{\sigma \mathrm{\Sigma }(r)}`$ form an open covering of $`P_\mathrm{\Sigma }`$.
###### Proposition 2.2
Assume $`e_1,e_2,e_3`$ are the generators of a 3-dimensional simplicial cone $`\sigma `$ in a lattice $`M`$ and $`\{0,e_1,e_2,e_3\}=\sigma \mathrm{\Delta }`$ for reflexive polyhedron $`\mathrm{\Delta }M`$. Then $`\sigma `$ is primative.
Proof: Assume that $`\sigma `$ is not primative, there exist a prime number $`p`$ and non-negative integers $`a,b,c`$ such that
$$m=\frac{a}{p}e_1+\frac{b}{p}e_2+\frac{c}{p}e_30$$
is in $`\sigma `$. Without loose of generality, we may assume $`c=1`$ and $`0a,bp1`$. Then $`1a+b+c2p1`$.
Since $`\{0,e_1,e_2,e_3\}=\sigma \mathrm{\Delta }`$, by theorem 2.2, $`e_1,e_2,e_3`$ are all in one codimension 1 face of reflexive polyhedron $`\mathrm{\Delta }M`$, there exist an element $`n`$ in the dual lattice of $`M`$ such that $`n|_\mathrm{\Delta }1`$ and $`e_k,n=1`$ for $`k=1,2,3`$. Since $`mM`$,
$$m,n=\frac{a+b+c}{p}$$
should be an integer. Therefore
$$a+b+c=p$$
$`m`$ as a convex combination of $`e_1,e_2,e_3`$ is in $`\sigma \mathrm{\Delta }`$. A contradiction!
$`\mathrm{}`$
Remark: This theorem is equivalent to the fact that in any rank two lattice, a simplex with integral vertex that do not contain integral points in inside and edges is always the standard simplex with area $`\frac{1}{2}`$.
By proposition 2.1 and 2.2, we have
###### Corollary 2.1
If $`\mathrm{\Sigma }`$ is a maximal simplicial fan on a 4-dimensional lattice $`N`$ that is compatible with a reflexive polyhedron $`\mathrm{\Delta }M`$, Then $`P_\mathrm{\Sigma }`$ is smooth away from the zero dimension subtorus (fixed points under the action of $`T`$) of $`P_\mathrm{\Sigma }`$.
$`\mathrm{}`$
###### Corollary 2.2
If $`\mathrm{\Sigma }`$ is a maximal simplicial fan on a 4-dimensional lattice $`N`$ that is compatible with a reflexive polyhedron $`\mathrm{\Delta }M`$, then a generic section of the anti-canonical bundle $`L_\mathrm{\Delta }`$ on $`P_\mathrm{\Sigma }`$ cut out a smooth 3-dimensional Calabi-Yau manifold.
$`\mathrm{}`$
Or more generally
###### Corollary 2.3
If $`\mathrm{\Sigma }`$ is a maximal simplicial fan on a lattice $`N`$ that is compatible with a reflexive polyhedron $`\mathrm{\Delta }M`$, Then $`P_\mathrm{\Sigma }`$ is smooth away from the codimension 4 subtorus of $`P_\mathrm{\Sigma }`$. A generic section of the anti-canonical bundle $`L_\mathrm{\Delta }`$ on $`P_\mathrm{\Sigma }`$ cut out a Calabi-Yau hypersurface with codimension 4 singularities.
$`\mathrm{}`$
### 2.5 Toric action of a torus on a toric variety
As we know, $`M`$ can be viewed as monomial functions on $`T=\mathrm{Spec}(𝐂[M])N_𝐂/N=N_𝐙𝐂^{}`$. Given a subspace $`WN`$, $`W^{}M`$ can be viewed as $`T_W=W_𝐂/W=W_𝐙𝐂^{}`$ invariant functions on $`T`$, or on another word, functions on $`T/T_W`$.
For $`\sigma \mathrm{\Sigma }`$, the action of $`T_W`$ on $`T`$ can be naturally extended to an action on $`T_\sigma `$, that is indicated by $`W_\sigma N_\sigma =N/\overline{\sigma }`$. The stablizer of the action is corresponding to $`W\overline{\sigma }`$. We have the exact sequence
$$0T_{W\overline{\sigma }}T_WT_{W_\sigma }0$$
$`W^{}\sigma ^{}\sigma ^{}`$ can be viewed as $`T_{W_\sigma }=W_\sigma _𝐙𝐂^{}`$ invariant functions on $`T_\sigma `$, or on another word, functions on $`T_\sigma /T_{W_\sigma }`$.
###### Lemma 2.1
$`T_W`$ acts freely on $`T_\sigma `$ if and only if $`\overline{\sigma }W=\{0\}`$.
$`\mathrm{}`$
### 2.6 Blow up
Given two rational strongly convex polyhedron cones $`\tau \sigma `$ in $`N`$ ($`\tau `$ is not necessaryly a subcone of $`\sigma `$), then
###### Proposition 2.3
$`\tau ^{}\sigma ^{}`$ is a subface of $`\sigma ^{}`$, and the following diagram commutes
where $`\sigma _\tau `$ is the smallest subcone of $`\sigma `$ that contains $`\tau `$ and clearly, $`\sigma _\tau ^{}=\mathrm{Span}(\tau ^{}\sigma ^{})`$.
$`\mathrm{}`$
###### Corollary 2.4
If $`dim\tau =dim\sigma `$, then $`\sigma _\tau =\sigma `$, $`T_\tau T_\sigma `$ is an isomorphism.
$`\mathrm{}`$
This proposition implies that $`U_\tau `$ is blow up from $`U_\sigma `$ and also indicates that how subtorus in $`U_\tau `$ is blow down to subtorus of $`U_\sigma `$.
## 3 Kähler moduli of Calabi-Yau hypersurface in toric variety
It is interesting to consider Calabi-Yau hypersurface in a toric variety $`P_\mathrm{\Sigma }`$ that comes from a reflexive polyhedron $`\mathrm{\Delta }M`$. As we know, anti-canonical bundle $`L_\mathrm{\Delta }`$ is semi-ample. A generic section $`s`$ of $`L_\mathrm{\Delta }`$ cut out a Calabi-Yau hypersurface $`X=X_sP_\mathrm{\Sigma }`$. Integral points in $`\mathrm{\Delta }`$ naturally give us $`T`$-invariant sections of $`L_\mathrm{\Delta }`$, $`s`$ can be taken to be generic linear combination of these $`T`$-invariant sections. For the purpose of mirror symmetry, it is important to understand the complex moduli and Kähler cone of these Calabi-Yau hypersurfaces. Before describing the Kähler moduli of Calabi-Yau hypersurface, first we need to discuss the Kähler cone of a toric variety.
### 3.1 Kähler cone of a toric variety
It is interesting to understand the Kähler cone of $`P_\mathrm{\Sigma }`$. Namely the cone of convex piecewise linear functions modulo linear functions. Recall that $`mM`$ (linear function on $`N`$) corresponds to $`T`$ invariant meromorphic functions. An integral piecewise linear function $`p`$ corresponds to a meromorphic section (up to multiplication by a non-zero constant) of the corresponding line bundle, those $`p0`$ correspond to holomorphic sections. In particular, $`m=00`$ corresponds to constant functions that are the only holomorphic functions on $`P_\mathrm{\Sigma }`$. More precisely, an integral piecewise linear function corresponds to a divisor. (A divisor is equivalent to a line bundle together with a non-trivial meromorphic section up to non-zero constant multiple.)
A piecewise linear function $`p`$ is called compatible with a fan $`\mathrm{\Sigma }`$ if the restriction of $`p`$ to each cone $`\sigma \mathrm{\Sigma }`$ is linear. Fix value of $`p`$ on $`\mathrm{\Sigma }(1)`$, there are many ways to extend $`p`$ to piecewise linear function on $`N`$. For any simplicial fan $`\mathrm{\Sigma }`$ that is compatible with the Newton polyhedron $`\mathrm{\Delta }_\mathrm{\Sigma }`$ (determined by $`\mathrm{\Sigma }(1)`$), there is an unique extension of $`p`$ as a piecewise linear function on $`N`$ that is compatible with $`\mathrm{\Sigma }`$.
Conversely, a piecewise linear function $`p`$ will canonically determine a fan $`\mathrm{\Sigma }_p`$ such that $`\mathrm{\Sigma }_p`$ is compatible with $`p`$ and any other $`\mathrm{\Sigma }`$ compatible with $`p`$ is a subdivision of $`\mathrm{\Sigma }_p`$. A (convex) piecewise linear function $`p`$ is called strongly (convex) piecewise linear with respect to a fan $`\mathrm{\Sigma }`$ compatible with $`p`$ if $`\mathrm{\Sigma }=\mathrm{\Sigma }_p`$, we will also say $`p`$ is strongly (convex) piecewise linear without mentioning the fan $`\mathrm{\Sigma }`$ if $`\mathrm{\Sigma }=\mathrm{\Sigma }_p`$ is simplicial.
Let $`\mathrm{\Sigma }`$ be a simplicial cone compatible with the Newton polyhedron $`\mathrm{\Delta }`$, then the classical Kähler cone is just the set of (strongly) convex piecewise liear function $`p`$ (modulo linear functions $`M`$) that is compatible with $`\mathrm{\Sigma }`$. For our purpose, we are more interested in the movable cone, which is solely determined by the polyhedron $`\mathrm{\Delta }`$. For a piecewise linear function $`p`$, let $`p^1=p|_{\mathrm{\Sigma }(1)}`$. We call $`p^1`$ the 1-skeleton of $`p`$. To define the movable cone, in general we also use $`p^1`$ to denote a general piecewise linear finction on $`\mathrm{\Sigma }(1)`$. Then the movable cone
$$K_\mathrm{\Delta }=\{p^1|\mathrm{for}\mathrm{any}\sigma \mathrm{\Sigma }(1),\mathrm{exist}m_\sigma M,\mathrm{s}.\mathrm{t}.m_\sigma p^1,m_\sigma |_\sigma =p^1|_\sigma \}/M$$
It is easy to see that $`K_\mathrm{\Delta }`$ is a convex cone. For any $`p^1K_\mathrm{\Delta }`$, we can define
$$p(n)=\mathrm{min}\{m,n|mp^1,mM\}$$
Clearly, $`p`$ is a convex piecewise linear extension of $`p^1`$, and $`p^1`$ is the 1-skeleton of $`p`$. Namely, if the 1-skeleton $`p^1`$ of $`p`$ is in $`K_\mathrm{\Delta }`$, then $`p`$ is completely determined by $`p^1`$ and $`p`$ is convex. For generic $`p^1K_\mathrm{\Delta }`$, $`p`$ is strongly convex, $`\mathrm{\Sigma }_p`$ is a simplicial fan. According to type of $`\mathrm{\Sigma }_p`$, the movable cone $`K_\mathrm{\Delta }`$ is divided into subcones that are the Kähler cones of different birational equivalent models.
For $`p^10`$, a piecewise linear function on $`\mathrm{\Sigma }(1)`$, for each $`\sigma \mathrm{\Sigma }(1)`$, we can define halfspace
$$H_\sigma =\{mM|m,e_\sigma p^1(e_\sigma )\}$$
$$\mathrm{\Delta }_p=\underset{\sigma \mathrm{\Sigma }(1)}{}H_\sigma $$
is a convex polyhedron in $`M`$ containing origion. $`p^1K_\mathrm{\Delta }`$ if and only if for each $`\sigma \mathrm{\Sigma }(1)`$, $`\mathrm{\Delta }_p`$ touch boundary of $`H_\sigma `$.
In particular, when $`\mathrm{\Delta }`$ is reflexive the anti-canonical divisor corresponds to $`p_\mathrm{\Delta }^1(e_\sigma )=1`$ for $`\sigma \mathrm{\Sigma }(1)`$. The corresponding convex polyhedron $`\mathrm{\Delta }_{p_\mathrm{\Delta }}`$ is exactly $`\mathrm{\Delta }`$. $`H_\sigma `$ will touch $`\mathrm{\Delta }`$ in dimension $`r1`$ when $`\sigma \mathrm{\Sigma }(1)_p`$, and will touch in lower dimension when $`\sigma \mathrm{\Sigma }(1)_e`$. Therefore anti-canonical class is in $`K_\mathrm{\Delta }`$. Further more, $`p_\mathrm{\Delta }`$ is compatible with any simplicial fan $`\mathrm{\Sigma }`$ that is compatible with $`\mathrm{\Delta }`$. Therefore, anti-canonical class is in the Kähler cone of every birational equivalent model that dominate the anti-canonical model. Namely in the intersection of all the Kähler cones of the birational equivalent models that dominate the anti-canonical model.
Remark: In general, for $`p^1K_\mathrm{\Delta }`$, apriorily, the corresponding $`p`$ is not necessary compatible with the Newton polyhedron $`\mathrm{\Delta }`$. This corresponds to the question, weither the corresponding toric variety will dominate anti-canonical model or in another word, weither the anti-canonical bundle is semi-ample. It turns out the anti-canonical bundle is always semi-ample.
###### Theorem 3.1
For strongly convex $`p^1K_\mathrm{\Delta }`$, the corresponding $`p`$ is always compatible with the Newton polyhedron $`\mathrm{\Delta }`$. Equivalently, $`P_{\mathrm{\Sigma }_p}`$ dominate the anti-canonical model.
This result is easyly implied by the following
###### Theorem 3.2
For reflexive polyhedron $`\mathrm{\Delta }M`$, if two integral points $`m_1,m_2\mathrm{\Delta }`$ is not in the same face of $`\mathrm{\Delta }`$, then $`m_1+m_2\mathrm{\Delta }`$.
Proof: Recall that corresponging to the reflexive polyhedron $`\mathrm{\Delta }`$ there is an integer valued non-positive convex function $`p_\mathrm{\Delta }`$, such that
$$\mathrm{\Delta }=\{mM|p_\mathrm{\Delta }(m)1\}.$$
$$\mathrm{\Delta }=\{mM|p_\mathrm{\Delta }(m)=1\}.$$
Hence $`p_\mathrm{\Delta }(m_1)=p_\mathrm{\Delta }(m_2)=1`$. Convexity of $`p_\mathrm{\Delta }`$ implies that
$$p_\mathrm{\Delta }(m_1+m_2)p_\mathrm{\Delta }(m_1)+p_\mathrm{\Delta }(m_2)=2.$$
Equality hold if and only if $`m_1,m_2`$ are in the same face of $`\mathrm{\Delta }`$. Therefore, when $`m_1,m_2`$ are not in the same face of $`\mathrm{\Delta }`$, we have $`p_\mathrm{\Delta }(m_1+m_2)=1`$ and $`m_1+m_2`$ is in $`\mathrm{\Delta }`$.
$`\mathrm{}`$
Remark: The theorem 3.2 proved more than what is needed for the theorem 3.1, and is interesting in its own.
###### Corollary 3.1
The movable cone $`K_\mathrm{\Delta }`$ of $`P_\mathrm{\Sigma }`$ is divided into union of Kähler cones of different birational models. All of these Kähler cones have a common boundary point, the anti-canonical class of $`P_\mathrm{\Sigma }`$.
Given $`\mathrm{\Sigma }(1)`$ consists of integral points on the boundary of $`\mathrm{\Delta }^{}`$, take a convex body $`B`$ containing origion (e.g. standard ball), then for any $`\sigma \mathrm{\Sigma }(1)`$, $`B`$ intersects $`\sigma `$ at a vector $`v_\sigma `$. $`p^1(v_\sigma )=1`$ will give us $`p^1K_\mathrm{\Delta }`$.
### 3.2 Restriction to a Calabi-Yau hypersurface
The toric part of the Kähler cone of $`X`$ is the restriction of the Kähler cone of $`P_\mathrm{\Sigma }`$ to $`XP_\mathrm{\Sigma }`$. Just like in the toric case, it is more natural to consider movable cone of $`X`$. Recall that the Kähler cone of $`P_\mathrm{\Sigma }`$ corresponds to convex piecewise linear function $`p`$ on $`N`$ with respect to $`\mathrm{\Sigma }`$. Function $`p`$ are determined by its restriction $`p^1`$ of $`p`$ to $`\sigma \mathrm{\Sigma }(1)`$, whose primitive elements $`e_\sigma `$ are vertices of $`\mathrm{\Delta }^{}`$. The union of all the Kähler cones of $`P_\mathrm{\Sigma }`$ for $`\mathrm{\Sigma }`$ compatible with $`\mathrm{\Delta }`$ forms the movable cone. Anti-canonical class is determined by $`p_\mathrm{\Delta }`$ that satisfies $`p_\mathrm{\Delta }(e_\sigma )=1`$ for $`\sigma \mathrm{\Sigma }(1)`$. $`p_\mathrm{\Delta }`$ also determine a section of $`L_\mathrm{\Delta }`$ that corresponds to $`0\mathrm{\Delta }`$, which vanishes to order one on each divisor $`D_\sigma =P_{\mathrm{\Sigma }_\sigma }P_\mathrm{\Sigma }`$ for $`\sigma \mathrm{\Sigma }(1)`$. Recall to get the restriction of $`L_\mathrm{\Delta }`$ to $`D_\sigma `$, we need to adjust $`p_\mathrm{\Delta }`$ by linear functions in $`M`$ to get a $`p_\sigma `$ such that $`p_\sigma |_\sigma =0`$ and $`p_\sigma 0`$. $`p_\sigma `$ will naturally give us a piecewise linear function (still denote by $`p_\sigma `$) on $`N_\sigma `$ with respect to the fan $`\mathrm{\Sigma }_\sigma `$. It is easy to see that
###### Lemma 3.1
$`p_\sigma =0`$ on $`N_\sigma `$ if and only if $`e_\sigma `$ is an inner point of a $`r1`$ dimensional face of $`\mathrm{\Delta }^{}`$.
###### Corollary 3.2
$`L_\mathrm{\Delta }`$ restricts to $`𝒪_{D_\sigma }`$ on $`D_\sigma `$ if and only if $`e_\sigma `$ is an inner point of a $`r1`$ dimensional face of $`\mathrm{\Delta }^{}`$.
###### Corollary 3.3
For a generic section $`s`$ of $`L_\mathrm{\Delta }`$, $`X_s`$ does not intersect $`D_\sigma `$ if and only if $`e_\sigma `$ is an inner point of a $`r1`$ dimensional face of $`\mathrm{\Delta }^{}`$.
Let $`\mathrm{\Sigma }(1)_0`$ be the subset of $`\mathrm{\Sigma }(1)`$ consists of $`\sigma \mathrm{\Sigma }(1)`$ such that $`e_\sigma `$ is not an inner point of a $`r1`$ dimensional face of $`\mathrm{\Delta }^{}`$. Let $`p^0=p|_{\mathrm{\Sigma }(1)_0}`$, then the movable cone $`K_\mathrm{\Delta }^0`$ of $`X`$ can be defined as
$$K_\mathrm{\Delta }^0=\{p^0|\mathrm{for}\mathrm{any}\sigma \mathrm{\Sigma }(1)_0,\mathrm{exist}m_\sigma M,\mathrm{s}.\mathrm{t}.m_\sigma p^0,m_\sigma |_\sigma =p^0|_\sigma \}/M$$
It is easy to see that $`K_\mathrm{\Delta }^0`$ is a convex cone. For any $`p^0K_\mathrm{\Delta }^0`$, we can define
$$p(n)=\mathrm{min}\{m,n|mp^0,mM\}$$
Clearly, $`p`$ is a convex piecewise linear extension of $`p^0`$. For generic $`p^0K_\mathrm{\Delta }^0`$, $`p`$ is strongly convex, $`\mathrm{\Sigma }_p`$ is a simplicial fan. According to type of $`\mathrm{\Sigma }_p`$, the movable cone $`K_\mathrm{\Delta }^0`$ is divided into subcones that are the Kähler cones of different birational equivalent models of $`X`$.
The two movable cones $`K_\mathrm{\Delta }`$ and $`K_\mathrm{\Delta }^0`$ are closely related. There are natural projection map $`\pi :K_\mathrm{\Delta }K_\mathrm{\Delta }^0`$ and inclusion map $`i:K_\mathrm{\Delta }^0K_\mathrm{\Delta }`$. $`\pi `$ is defined as $`p^0=p^1|_{\mathrm{\Sigma }(1)_0}`$, and $`i`$ is defined as the restriction of the extension $`p`$ of $`p^0`$ to $`\mathrm{\Sigma }(1)`$. Clearly, $`\pi i=id`$.
## 4 Automorphism group of toric variety and slicing theorem
The toric part of the complex moduli of $`X=X_s`$ is parametrized by section $`s`$ of the anti-canonical bundle $`L_\mathrm{\Delta }`$ modulo the action of the automorphism group of $`P_\mathrm{\Sigma }`$. The space of sections modulo constant multiple is a projective space. To understand the complex moduli, the key point is to understand the automorphism group of $`P_\mathrm{\Sigma }`$. In this section, we will first discuss the automorphism group of $`P_\mathrm{\Sigma }`$, then we will prove a slicing theorem that enable us to reduce the complex moduli near the large complex limit from a quotient by $`\mathrm{Aut}(P_\mathrm{\Sigma })`$ to a toric quotient by the maximal torus of $`\mathrm{Aut}(P_\mathrm{\Sigma })`$.
### 4.1 Automorphism group of a toric variety
The automorphism group $`\mathrm{Aut}(P_\mathrm{\Sigma })`$ of $`P_\mathrm{\Sigma }`$ is a complex Lie group. Infinitesimally, automorphism group can be decomposed into two parts, the part that keep the subtorus of $`P_\mathrm{\Sigma }`$ invariant (the toric part) and the part that move those subtorus (the non-toric part). Elements in the first part are determined by the automorphisms of the big torus $`TU_0`$. Elements in the second part (more precisely, modulo the first part) are determined by their action on each divisor $`D_\sigma `$ for $`\sigma \mathrm{\Sigma }(1)`$, which can be characterized as sections of the normal bundles $`𝒪(D_\sigma )|_{D_\sigma }`$ of $`D_\sigma P_{\mathrm{\Sigma }_\sigma }`$ for $`\sigma \mathrm{\Sigma }(1)`$. On the other hand, the action of non-toric part of $`\mathrm{Aut}(P_\mathrm{\Sigma })`$ on divisors $`D_\sigma `$ for $`\sigma \mathrm{\Sigma }(1)`$ are independent. Namely, any configuration of sections of $`𝒪(D_\sigma )|_{D_\sigma }`$ for $`\sigma \mathrm{\Sigma }(1)`$ can be realized from an infinitesimal automorphism of $`P_\mathrm{\Sigma }`$. More precisely, this discription of the infinitesimal automorphism group of $`P_\mathrm{\Sigma }`$ (the Lie algebra) can be expressed by the following exact sequence:
###### Lemma 4.1
$$0\mathrm{\Theta }_{P_\mathrm{\Sigma }}(\mathrm{log}D)\mathrm{\Theta }_{P_\mathrm{\Sigma }}\underset{\sigma \mathrm{\Sigma }(1)}{}𝒪(D_\sigma )|_{D_\sigma }0$$
is exact, where $`D=_\sigma D_\sigma `$, $`\mathrm{\Theta }_{P_\mathrm{\Sigma }}`$ denote the tangent sheaf of $`P_\mathrm{\Sigma }`$.
###### Corollary 4.1
$$0H^0(\mathrm{\Theta }_{P_\mathrm{\Sigma }}(\mathrm{log}D))H^0(\mathrm{\Theta }_{P_\mathrm{\Sigma }})\underset{\sigma \mathrm{\Sigma }(1)}{}H^0(𝒪(D_\sigma )|_{D_\sigma })0$$
is exact.
###### Proposition 4.1
$$\mathrm{\Theta }_{P_\mathrm{\Sigma }}(\mathrm{log}D)𝒪_{P_\mathrm{\Sigma }}_𝐙N$$
$$H^0(\mathrm{\Theta }_{P_\mathrm{\Sigma }}(\mathrm{log}D))𝐂_𝐙N$$
Let $`H_\sigma =\{mM|m,e_\sigma =1\}`$, $`\mathrm{\Delta }H_\sigma `$ is a subface of $`\mathrm{\Delta }`$ of dimension $`r1`$. When $`\mathrm{\Delta }H_\sigma `$ is of dimension $`r1`$, let $`\mathrm{\Delta }_\sigma `$ denotes the polyhedron consisting of interior points of $`\mathrm{\Delta }H_\sigma `$. When $`\mathrm{\Delta }H_\sigma `$ is of dimension $`<r1`$, let $`\mathrm{\Delta }_\sigma =\mathrm{}`$. $`\mathrm{\Delta }_\sigma `$ can also be interpreted as the set of supporting hyperplanes of $`\mathrm{\Delta }^{}`$ that touch $`\mathrm{\Delta }^{}`$ only at $`e_\sigma `$.
###### Proposition 4.2
For $`\sigma \mathrm{\Sigma }(1)`$, $`\mathrm{\Delta }_\sigma `$ is exactly the Newton polyhedron of the line bundle $`𝒪(D_\sigma )|_{D_\sigma }`$. Namely,
$$H^0(𝒪(D_\sigma )|_{D_\sigma })\mathrm{Span}_𝐂(\mathrm{\Delta }_\sigma ).$$
As we know, vector fields on $`T=U_0=\mathrm{Spec}(𝐂[M])`$ can be understood as differentials from $`𝐂[M]`$ to itself. First, let’s consider differentials that respect the toric structure. Namely, those with $`mM`$ as eigenvectors. This kind of $`𝐂`$-linear maps on $`𝐂[M]`$ is differential if and only if the eigenvalue $`\lambda _m`$ of $`mM`$ as a function on $`M`$ is linear with repect to the linear structure on $`M`$. Hence, for each $`nN`$, we can define a differential (vector field) $`v_n`$ that respect the toric structure by $`v_n(m)=m,nm`$. This naturally gives us the embedding
$$𝐂_𝐙NH^0(\mathrm{\Theta }_{U_0})$$
It is easy to see that these vector fields can naturally be extended to holomorphic vector fields on $`P_\mathrm{\Sigma }`$ and form all the sections of $`\mathrm{\Theta }_{P_\mathrm{\Sigma }}(\mathrm{log}D)`$. These vector fields vanish to order at most 1 at each divisor $`D_\sigma `$ for $`\sigma \mathrm{\Sigma }(1)`$. $`v_n`$ vanishes at $`D_\sigma `$ if and only if $`n`$ is a multiple of $`e_\sigma `$. General holomorphic vector fields on $`U_0`$ can be characterized as
###### Proposition 4.3
$$H^0(\mathrm{\Theta }_{U_0})𝐂[M]_𝐙N$$
In general, holomorphic vector fields on $`P_\mathrm{\Sigma }`$ can be characterized by image of the restriction map $`H^0(\mathrm{\Theta }_{P_\mathrm{\Sigma }})H^0(\mathrm{\Theta }_{U_0})`$.
###### Proposition 4.4
$`mv_n`$ can be extended holomorphically to $`P_\mathrm{\Sigma }`$ if and only if $`n`$ is a multiple of $`e_\sigma `$ for a $`\sigma \mathrm{\Sigma }(1)`$ and $`m`$ defines a supporting function of $`\mathrm{\Delta }^{}`$, whose supporting hyperplane touches $`\mathrm{\Delta }^{}`$ only at $`e_\sigma `$.
Now we have a natural splitting of the exact sequence in corollary 4.1
###### Corollary 4.2
$$H^0(\mathrm{\Theta }_{P_\mathrm{\Sigma }})H^0(\mathrm{\Theta }_{P_\mathrm{\Sigma }}(\mathrm{log}D))\underset{\sigma \mathrm{\Sigma }(1)}{}H^0(𝒪(D_\sigma )|_{D_\sigma })$$
with the first summand generated by $`v_n`$ for $`nN`$ and the second summand having a basis $`\{mv_{e_\sigma }\}_{m\mathrm{\Delta }\backslash \mathrm{\Delta }^0}`$, where $`m,e_\sigma =1`$ for a unique $`\sigma \mathrm{\Sigma }(1)`$ determined by $`m`$.
Remark: The automorphism group of $`P_\mathrm{\Sigma }`$ apriorily has nothing to do with the toric structure on $`P_\mathrm{\Sigma }`$. The effect of the toric structure is to specify a particular maximal torus in the automorphism group.
### 4.2 Slicing theorem
For the sake of the mirror symmetry, we need to understand complex moduli near the large complex limit. Notice that the large complex limit (corresponds to $`0\mathrm{\Delta }M`$) is not invariant under the full automorphism group, and is only invariant under the toric automorphism group (the maximal torus). To explicitly describe the complex moduli near the large complex limit, it is important to find an explicit slice of the space of sections of $`L_\mathrm{\Delta }`$ that contains the large complex limit, is invariant under toric automorphism group and intersects each nearby orbit of the automorphism group at an orbit of the toric automorphism group (the maximal torus). Let $`\mathrm{\Delta }^0`$ denote the set of integral points in the $`(r2)`$-skeleton of $`\mathrm{\Delta }`$.
###### Theorem 4.1
Near the large complex limit, $`\mathrm{Span}_𝐂(\mathrm{\Delta }^0)`$ is a slice of $`\mathrm{Span}_𝐂(\mathrm{\Delta })`$ that contains the large complex limit, is invariant under toric automorphism group and intersects each nearby orbit of the automorphism group at an orbit of the toric automorphism group (the maximal torus).
To understand the construction better, let’s analyze the example of cubics in $`\mathrm{𝐂𝐏}^2`$ in detail.
Example: In the case of cubics in $`\mathrm{𝐂𝐏}^2`$, the large complex limit corresponds to $`z_0z_1z_2`$. We want to reduce
$`F(z)`$ $`=`$ $`a_0z_0^3+a_1z_1^3+a_2z_2^3+\psi z_0z_1z_2`$
$`+a_{01}z_1^2z_2+a_{02}z_1z_2^2`$
$`+`$ $`a_{10}z_0^2z_2+a_{12}z_0z_2^2`$
$`+`$ $`a_{20}z_0^2z_1+a_{21}z_0z_1^2`$
to
$$F_0(z)=a_0z_0^3+a_1z_1^3+a_2z_2^3+\psi z_0z_1z_2$$
Consider the transformation
$$T_1:\left(\begin{array}{c}z_0\hfill \\ z_1\hfill \\ z_2\hfill \end{array}\right)\left(\begin{array}{ccc}1& a_{01}/\psi & a_{02}/\psi \\ a_{10}/\psi & 1& a_{12}/\psi \\ a_{20}/\psi & a_{21}/\psi & 1\end{array}\right)\left(\begin{array}{c}z_0\hfill \\ z_1\hfill \\ z_2\hfill \end{array}\right)$$
Then
$$F_1(z)=F(T_1z)=\psi z_0z_1z_2+a_0z_0^3+a_1z_1^3+a_2z_2^3+O(1/\psi )$$
Or write
$`F_1(z)`$ $`=`$ $`a_0z_0^3+a_1z_1^3+a_2z_2^3+\psi z_0z_1z_2`$
$`+a_{01}z_1^2z_2+a_{02}z_1z_2^2`$
$`+`$ $`a_{10}z_0^2z_2+a_{12}z_0z_2^2`$
$`+`$ $`a_{20}z_0^2z_1+a_{21}z_0z_1^2`$
with $`a_{jk}=O(1/\psi )`$. (Here for simplicity, we are abusing the notation a little bit by using the same notation for the coefficients of $`F_1`$ as for the coefficients of $`F`$.) Repeat this process inductively, we will get $`T_2,T_3,\mathrm{}`$ such that
$`F_l(z)=F_{l1}(T_lz)`$ $`=`$ $`a_0z_0^3+a_1z_1^3+a_2z_2^3+\psi z_0z_1z_2`$
$`+a_{01}z_1^2z_2+a_{02}z_1z_2^2`$
$`+`$ $`a_{10}z_0^2z_2+a_{12}z_0z_2^2`$
$`+`$ $`a_{20}z_0^2z_1+a_{21}z_0z_1^2`$
with $`a_{jk}=O(1/\psi ^l)`$. Let $`T=T_1T_2\mathrm{}`$, we get
$$F_0(z)=F(Tz)=a_0z_0^3+a_1z_1^3+a_2z_2^3+\psi z_0z_1z_2$$
To prove theorem 4.1 in general, we need to understand the action of $`mv_{e_\sigma }`$ on sections of the anti-canonical bundle $`L_\mathrm{\Delta }`$. Section of $`L_\mathrm{\Delta }`$, in general, can be expressed as $`m(v_{n_1}\mathrm{}v_{n_r})`$, for $`m\mathrm{\Delta }`$, where $`n_1,\mathrm{},n_r`$ is a basis of $`N`$. We have
###### Lemma 4.2
$$_{m_1v_{n_1}}m_2(v_{n_1}\mathrm{}v_{n_r})=m_2m_1,n_1(m_1+m_2)(v_{n_1}\mathrm{}v_{n_r})$$
In particular, for $`mv_{e_\sigma }`$, $`m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0`$, $`m,e_\sigma =1`$,
$$_{mv_{e_\sigma }}m_1(v_{n_1}\mathrm{}v_{n_r})=(m_1,e_\sigma +1)(m+m_1)(v_{n_1}\mathrm{}v_{n_r})$$
$$_{mv_{e_\sigma }}(v_{n_1}\mathrm{}v_{n_r})=m(v_{n_1}\mathrm{}v_{n_r})$$
To simplify the notation, let $`s_m=m(v_{n_1}\mathrm{}v_{n_r})`$, for $`mM`$ and $`v_m=mv_{e_\sigma }`$ for $`m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0`$, $`m,e_\sigma =1`$. Then the above formula can be expressed as
###### Corollary 4.3
For $`m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0`$, $`m_1M`$,
$$_{v_m}s_0=s_m,_{v_m}s_{m_1}=(m_1,e_\sigma +1)s_{m+m_1}$$
where $`\sigma \mathrm{\Sigma }(1)`$ is determined by $`m,e_\sigma =1`$.
Proof of theorem 4.1: Let
$$s=\underset{m(0)\mathrm{\Delta }}{}a_ms_m+\psi s_0,$$
Then the theorem asserts that when $`\psi `$ is large, automorphism of $`P_\mathrm{\Sigma }`$ can reduce $`s`$ to the following standard form
$$s^0=\underset{m\mathrm{\Delta }^0}{}a_ms_m+\psi s_0,$$
Consider the automorphism of $`P_\mathrm{\Sigma }`$
$$T_1=\mathrm{exp}(\underset{m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0}{}a_mv_m)$$
Then by using corollary 4.3, we have
$`s^1=T_1(s)`$ $`=`$ $`\psi s_0+{\displaystyle \underset{m(0)\mathrm{\Delta }}{}}a_ms_m{\displaystyle \underset{m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0}{}}a_m_{v_m}s_0+O({\displaystyle \frac{1}{\psi }})`$
$`=`$ $`\psi s_0+{\displaystyle \underset{m\mathrm{\Delta }^0}{}}a_ms_m+{\displaystyle \underset{m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0}{}}a_ms_m{\displaystyle \underset{m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0}{}}a_ms_m+O({\displaystyle \frac{1}{\psi }})`$
$`=`$ $`\psi s_0+{\displaystyle \underset{m\mathrm{\Delta }^0}{}}a_ms_m+O({\displaystyle \frac{1}{\psi }})`$
Adjust $`\psi `$ and $`a_m`$ accordingly, above equation can also be expressed as
$$s^1=\underset{m(0)\mathrm{\Delta }}{}a_ms_m+\psi s_0,$$
where $`a_m=O(1/\psi )`$ for $`m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0`$. Repeat this process inductively, we will get $`T_2,T_3,\mathrm{}`$ such that
$$s^l=T_l(s^{l1})=\underset{m(0)\mathrm{\Delta }}{}a_ms_m+\psi s_0,$$
with $`a_m=O(1/\psi ^l)`$ for $`m(0)\mathrm{\Delta }\backslash \mathrm{\Delta }^0`$. Let $`T=T_1T_2\mathrm{}`$, we get
$$s^0=T(s)=\underset{m\mathrm{\Delta }^0}{}a_ms_m+\psi s_0$$
is in standard form (belong to our slice).
$`\mathrm{}`$
## 5 Complex moduli of Calabi-Yau hypersurface and the monomial-divisor mirror map
According to theorem 4.1, the complex moduli near the large complex limit can be characterized by the space of sections
$$s=s_0+\underset{m\mathrm{\Delta }^0}{}a_ms_m$$
of $`L_\mathrm{\Delta }`$ for $`a_m`$ small modulo the action of $`T=U_0=N_𝐂/N`$. Let
$$\stackrel{~}{M}_0=\left\{a^I=\underset{m\mathrm{\Delta }^0}{}a_m^{i_m}\right|I=(i_m)_{m\mathrm{\Delta }^0}𝐙^{\mathrm{\Delta }^0}\}𝐙^{\mathrm{\Delta }^0},$$
then its dual $`\stackrel{~}{N}_0`$ is naturally the space of all the weights
$$\stackrel{~}{N}_0=\{w=(w_m)_{m\mathrm{\Delta }^0}𝐙^{\mathrm{\Delta }^0}\}𝐙^{\mathrm{\Delta }^0}.$$
By lemma 4.2, we have
$$_{v_n}s_m=m,ns_m$$
$`nm,n`$ define an embedding $`N\stackrel{~}{N}_0`$. Let $`WN_0`$ be the image of this embedding. Then $`\stackrel{~}{M}=W^{}`$ is the space of $`T`$ invariant functions on complex torus $`\stackrel{~}{T}_0=\mathrm{Spec}(𝐂[\stackrel{~}{M}_0])(𝐂^{})^{\mathrm{\Delta }^0}`$. $`\stackrel{~}{N}=\stackrel{~}{N}_0/W`$ is dual to $`\stackrel{~}{M}`$. For a convex cone $`\sigma _0\stackrel{~}{N}_0`$, let $`\sigma ^{}=\sigma _0^{}\stackrel{~}{M}`$, then $`\sigma =(\sigma ^{})^{}`$ is naturally the projection of $`\sigma _0`$ to $`\stackrel{~}{N}`$.
Let
$$\sigma _0=\{w=(w_m)_{m\mathrm{\Delta }^0}𝐙_0^{\mathrm{\Delta }^0}\}\stackrel{~}{N}_0.$$
Then the fan $`\stackrel{~}{\mathrm{\Sigma }}_0`$ consists of subcones of $`\sigma _0`$ corresponds to the affine toric variety $`𝐂^{\mathrm{\Delta }^0}`$. Subcones $`\sigma _{S,0}`$ of $`\sigma _0`$ is indexed by subsets $`S\mathrm{\Delta }^0`$.
$$\sigma _{S,0}=\{w\sigma _0|w_m=0,\mathrm{for}mS\}$$
Let $`\sigma _S`$, $`\sigma `$ be the cones that are the projection to $`N`$ of the cones $`\sigma _{S,0}`$, $`\sigma _0`$ in $`\stackrel{~}{\mathrm{\Sigma }}_0`$. Notice that $`\stackrel{~}{N}_0`$ can be interpreted as the space of restriction to $`\mathrm{\Delta }^0`$ of the piecewise linear functions on $`M`$, which we encountered, when we discussed movable cone of a Calabi-Yau hypersurface. This can be easyly seen by defining $`p_w^0(m)=w_m`$. In this way, $`\sigma _0`$ corresponds to piecewise linear functions that are non-positive, namely, convex at origion. $`\sigma _{S,0}`$ corresponds to piecewise linear functions that are non-positive and vanish at $`S`$, namely, convex at the cone generated by $`S`$ (if there are no more other vertices in this cone). $`W`$ as image of $`N`$ corresponds to the space of linear function on $`M`$. $`\stackrel{~}{N}`$ can be viewed as space of restriction to $`\mathrm{\Delta }^0`$ of the piecewise linear functions modulo linear functions on $`M`$. $`\sigma _S`$, $`\sigma `$ also have corresponding convexity interpretation modulo linear functions.
We are interested in the complex moduli, which is, near the large complex limit, by our slicing theorem, the quotient of $`𝐂^{\mathrm{\Delta }^0}`$ by the action of $`T`$. But the set theoritical quotient is usually not seperable, let along an algebraic variety. To get a seperable algebraic variety as quotient, we have to take the quotient in geomeric invariant theory sense. Namely, we may need to throw away some of the subtorus and collapse some non-generic orbits together when we do the quotient. Then there is always the problem that what to throw away and what to collapse. There are many ways to do it. For our purpose, we want to find a “canonical” quotient that suits our need for mirror symmetry.
There are some general guideline for the choice, we usually want to keep the good generic orbits, in particular, the stablizer is better trivial. A sequence of generic orbits at limit, could converge to a union of several orbits. If some of these limiting orbits are to be keeped, this orbits should be collapsed to one point in the quotient.
Example: $`𝐂^{}`$ acts on $`𝐂^\mathrm{𝟐}`$ by
$$t(z_1,z_2)=(tz_1,t^1z_2).$$
the generic orbits are $`O_c=\{z_1z_2=c\}`$. When $`c`$ approach zero, $`O_c`$ approach the union of three orbits
$$\{z_1=z_2=0\}\{z_1=0,z_20\}\{z_2=0,z_10\}$$
The first orbit has the stablizer $`𝐂^{}`$ need to be throw away, the last two orbits have trivial stablizer and can be keeped, but have to be collapsed to one point in the quotient. In this way we get the quotient to be $`𝐂`$ parametrized by $`c`$.
According to lemma 2.1, the stablizer of the action of $`T`$ on $`T_{\sigma _{S,0}}`$ is corresponding to $`W\mathrm{Span}_𝐂(\sigma _{S,0})`$. The stablizer is trivial if and only if $`W`$ projects injectively into $`\stackrel{~}{N}_{\sigma _{S,0},0}=\stackrel{~}{N}_0/\mathrm{Span}_𝐂(\sigma _{S,0})`$. In another word,
###### Lemma 5.1
The stablizer of the action of $`T`$ on $`T_{\sigma _{S,0}}`$ is trivial if and only if the cone generated by $`S\mathrm{\Delta }^0`$ is of top dimension in $`M`$.
In general, we are intersted in the kind of toric varieties whose fan $`\mathrm{\Sigma }`$ is determined by the top dimensional cones $`\mathrm{\Sigma }(r)`$, namely each cone in $`\mathrm{\Sigma }`$ is a subcone of a top dimensional cone in $`\mathrm{\Sigma }`$. Assume our quotient is a toric variety of this type. Then we only need to determine top dimensional cones in its fan $`\stackrel{~}{\mathrm{\Sigma }}`$.
###### Lemma 5.2
$`\sigma _{S,0}`$ is of top dimensional in $`\stackrel{~}{N}`$ and $`T`$ acts freely on $`T_{\sigma _{S,0}}`$ if and only if
$$\stackrel{~}{N}_0=W\mathrm{Span}_𝐂(\sigma _{S,0})$$
or equivalently, $`S`$ contain exactly $`r`$ linear independent elements in $`M`$.
For our purpose, we want our construction to be compatible with the Newton polyhedron $`\mathrm{\Delta }`$, namely, we require $`S`$ belongs to one of the $`(r1)`$-dimensional face of $`\mathrm{\Delta }`$. This is the way complex moduli of Calabi-Yau hypersurface in $`P_\mathrm{\Sigma }`$ and Kähler moduli of Calabi-Yau hypersurface in $`P_\mathrm{\Sigma }^{}`$ are related. Namely, the complex moduli of Calabi-Yau hypersurface in $`P_\mathrm{\Sigma }`$ near the large complex limit is naturally a toric variety, its fan can be naturally identified with the fan made up with the varieous Kähler cones of different birational models of Calabi-Yau hypersurface in $`P_\mathrm{\Sigma }^{}`$.
The mirror symmetry actually requires something more, a precise identification of complex moduli $`P_{\stackrel{~}{\mathrm{\Sigma }}}`$ near the large complex limit and the complexified Kähler moduli $`(\stackrel{~}{N}_𝐙𝐑+iK_\mathrm{\Delta }^{}^0)/\stackrel{~}{N}`$ near the large radius limit.
In general, recall that for $`nN_𝐂=N_𝐙𝐂`$, $`mM`$, we have the Lie algebra action $`v_n(m)=m,nm`$. By the expoential map, we have the Lie group action
$$n(m)=e^{2\pi im,n}m$$
where we assume that $`nT=(N_𝐙𝐂)/N`$. For any $`\sigma \mathrm{\Sigma }`$, we can define $`\sigma _𝐂=N_𝐑+i\sigma _𝐑N_𝐂`$, where $`N_𝐑=N_𝐙𝐑`$ and $`\sigma _𝐑`$ is the real cone in $`N_𝐑`$ generated by $`\sigma `$. $`\sigma ^{}M`$ are the monomials that can be extended to the affine toric variety $`U_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])`$. $`\sigma ^{}\backslash \sigma ^{}M`$ are the functions on $`U_\sigma `$ that vanish along $`T_\sigma U_\sigma `$. For $`n=n_1+in_2\sigma _𝐂^0`$, $`m\sigma ^{}\backslash \sigma ^{}`$, we have $`n_2\sigma _𝐑^0`$ and $`m,n_2>0`$. For $`t𝐑_+`$, we have
$$(tn)(m)=e^{2\pi im,tn}m=e^{2\pi im,tn_1}e^{2\pi tm,n_2}m$$
clearly
$$\underset{t+\mathrm{}}{lim}(tn)(m)=0$$
This implies that for any $`n=\sigma _𝐂^0`$ and $`xT=U_0`$, the ray $`tn(x)`$ approaches a point in $`T_\sigma `$ when $`t𝐑_+`$ approaches $`+\mathrm{}`$. The limit is actually indipendent of $`n\sigma _𝐂^0`$.
$$\underset{t+\mathrm{}}{lim}(tn)(x)=x_\sigma $$
where $`x_\sigma `$ is the image of $`x`$ under the projection $`\pi _\sigma :T=\mathrm{Spec}(𝐂[M])T_\sigma =\mathrm{Spec}(𝐂[\sigma ^{}])`$ that is determined by the embeding $`\sigma ^{}M`$. Actually $`n(x)\pi _\sigma ^1(x_\sigma )`$ for $`n\sigma _𝐂\overline{\sigma }_𝐂`$.
Apply this construction to our situation, we get the monomial divisor map
$$(\stackrel{~}{N}_𝐙𝐑+iK_\mathrm{\Delta }^{}^0)/\stackrel{~}{N}P_{\stackrel{~}{\mathrm{\Sigma }}}$$
which is an approximation of the actual mirror map.
Question: Fan $`\stackrel{~}{\mathrm{\Sigma }}`$ determine a sense of being near the large complex limit. What is precise characterization of being near the large complex limit determined by the fan $`\stackrel{~}{\mathrm{\Sigma }}`$?
For $`\sigma \mathrm{\Sigma }`$, $`\sigma _𝐂`$ is a subsemigroup of the complex torus $`TU_0`$. For any $`xU_0`$, $`O_x=\sigma _𝐂(x)`$ is an orbit of $`\sigma _𝐂`$. Two orbits $`O_x`$, $`O_y`$ of $`\sigma _𝐂`$ are called equivalent, if $`O_xO_y`$ is non-empty. We denote the equivalent class by $`[O_x]`$. Use these equivalence classes of orbits of semigroups, we have another interpretation of $`P_\mathrm{\Sigma }`$.
$$P_\mathrm{\Sigma }\{[O]|OU_0\mathrm{is}\mathrm{an}\mathrm{orbit}\mathrm{of}\sigma _𝐂\mathrm{for}\mathrm{some}\sigma \mathrm{\Sigma }\}$$
From this point of view, for $`w=(w_m)_{m\mathrm{\Delta }^0}K_\mathrm{\Delta }^{}^0`$, and
$$s=s_0+\underset{m\mathrm{\Delta }^0}{}a_ms_m$$
$$t^ws=s_0+\underset{m\mathrm{\Delta }^0}{}(t^{w_m}a_m)s_m$$
should be near the large complex limit if $`t𝐑_+`$ is sufficiently small. In another word, for $`a=(a_m)_{m\mathrm{\Delta }^0}`$ small,
$$s=s_0+\underset{m\mathrm{\Delta }^0}{}a_ms_m$$
is near the large complex limit $`s_0`$ if $`(\mathrm{log}a_m)_{m\mathrm{\Delta }^0}`$ is in $`(K_\mathrm{\Delta }^{}^0)_𝐂`$, or equivalently, $`(\mathrm{log}|a_m|)_{m\mathrm{\Delta }^0}`$ is in $`K_\mathrm{\Delta }^{}^0`$.
## 6 Lagrangian torus fibration of Calabi-Yau hypersurface
In , , we constructed Lagrangian fiberation for quintic Fermat type Calabi-Yau familly $`\{X_\psi \}`$ in $`𝐏^\mathrm{𝟒}`$ by gradient flow method. In , we deal with the case of generic quintic. In generic quintic case, the difficulty is not gradient flow. In very general situation, based on the technique result in , the gradient flow will deform the canonical Lagrangian torus fibration of the large complex limit to form Lagrangian torus fibration on general quintic Calabi-Yau. But the singular locus of the fibration so constructed is usually of codimension 1, while the SYZ conjecture expects that the singular locus to be of codimension 2. In the case of Fermat type quintic case, due to the explicit nature of the singular locus, it is easy to observe that the singular locus is naturally a fattening of a 1-dimensional graph $`\mathrm{\Gamma }`$. Use some explicit computation and symplectic geometry technique, in we were able to squize the singular locus to construct a Lagrangian torus fibration with codimension 2 singular locus $`\mathrm{\Gamma }`$. In the generic quintic case, the singular locus is usually very complicated and in general does not necessaryly resemble a fattening of a 1-dimensional graph. However, as observed in , when the quintic Calabi-Yau is near the large complex limit in a suitable sense, miraculously, the singular locus will resemble a fattening of a 1-dimensional graph $`\mathrm{\Gamma }`$. In , we discussed this phenomenon in great detail and used similar technique as in to finally achieve codimension 2 singular locus for Lagrangian torus fibration of generic Calabi-Yau quintic near the large complex limit in .
In this section, we will discuss how to generalize the results in , , , to construct Lagrangian torus fibration with codimension 2 singular locus for generic Calabi-Yau hypersurface in toric variety. Although Calabi-Yau hypersurface in general toric variety sounds much more general than quintic in $`𝐏^\mathrm{𝟒}`$. Actually there are no essentially new technical difficulties for the gradient flow. Generalization is a very natural one.
### 6.1 Lagrangian fibration with codimension 1 singular locus
Before getting into the detail, let’s first fix some notation. Assume that $`\mathrm{\Delta }`$ is a reflexive polyhedron, $`p_w`$ ($`p_v`$) is a convex piecewise linear function on $`M`$ ($`N`$) with respect to $`\mathrm{\Delta }^0`$ ($`(\mathrm{\Delta }^{})^0`$). $`p_w(m)=w_m`$ ($`p_v(n)=v_n`$) for $`m\mathrm{\Delta }^0`$ ($`n(\mathrm{\Delta }^{})^0`$). $`p_w`$ ($`p_v`$) naturally determines a fan $`\mathrm{\Sigma }^w`$ ($`\mathrm{\Sigma }^v`$) for $`M`$ ($`N`$) that is compatible with $`\mathrm{\Delta }^{}`$ ($`\mathrm{\Delta }`$). $`p_w`$ ($`p_v`$) also naturally determines a polyhedron $`\mathrm{\Delta }_wN`$ ($`\mathrm{\Delta }_vM`$) consists of $`nN`$ ($`mM`$) that as a linear function on $`M`$ ($`N`$) is greater or equal to $`p_w`$ ($`p_v`$). In particular, the fan $`\mathrm{\Sigma }`$ ($`\mathrm{\Sigma }^{}`$) of the anti-canonical model $`P_\mathrm{\Sigma }`$ ($`P_\mathrm{\Sigma }^{}`$) is determined by $`p_\mathrm{\Delta }`$ ($`p_\mathrm{\Delta }^{}`$), $`p_\mathrm{\Delta }(n)=1`$ ($`p_\mathrm{\Delta }^{}(m)=1`$) for $`n\mathrm{\Delta }^{}\backslash \{0\}`$ ($`m\mathrm{\Delta }\backslash \{0\}`$), for which the corresponding polyhedron is $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$).
The fan $`\mathrm{\Sigma }^v`$ ($`\mathrm{\Sigma }^w`$) is a crepant subdivision of the fan $`\mathrm{\Sigma }`$ ($`\mathrm{\Sigma }^{}`$). The corresponding toric variety $`P_{\mathrm{\Sigma }^v}`$ ($`P_{\mathrm{\Sigma }^w}`$) is a crepant resolution of the anti-canonical model $`P_\mathrm{\Sigma }`$ ($`P_\mathrm{\Sigma }^{}`$), with the natural map
$$\pi _v:P_{\mathrm{\Sigma }^v}P_\mathrm{\Sigma }$$
$$\pi _w:P_{\mathrm{\Sigma }^w}P_\mathrm{\Sigma }^{}$$
Before getting into the gradient flow construction, it is important to construct the suitable $`T_𝐑`$-invariant Kähler metric $`g`$ on $`P_{\mathrm{\Sigma }^v}`$. According to philosophy of SYZ conjecture, our purpose is not merely getting Lagrangian fibration for $`X_s`$. Instead, we need to construct Lagrangian fibration for $`X_s`$ that is depnding suitablly on any specific complex structure and Kähler structure on $`X_s`$. Our construction of $`T_𝐑`$-invariant Kähler metric $`g`$ on $`P_\mathrm{\Sigma }`$ should reflect this philosophy.
In the case of quintic Calabi-Yau that we dealt with in our previous papers, we essentially used the Fubini-Study metric. The reason is that the Kähler moduli for quintic Calabi-Yau is of dimension 1. The restriction of the Fubini-Study metric to quintic Calabi-Yau is enough to capture the Kähler moduli of quintic. Even in that case, to make the singular locus better understood, it is a good idea to modify the the Fubini-Study metric in the same Kähler class according to the complex structure of the particular quintic Calabi-Yau. It is helpful to review this construction for the quintic Calabi-Yau here to help us to understand our construction of $`T_𝐑`$-invariant Kähler metric here in the general toric case.
Consider a generic quintic defined by
$$p(z)=\underset{|m|=5}{}a_mz^m=0$$
in $`𝐏^\mathrm{𝟒}`$. Let $`\mathrm{\Delta }`$ denote the Newton polygon for quintic polynomials. Then we can introduce the Kähler potential
$$h_a=\mathrm{log}(|z|_a^2)=\mathrm{log}\left(\underset{m\mathrm{\Delta }}{}|a_mz^m|^2\right).$$
Let $`z^m=|z^m|e^{i\theta _m}`$, then
$$\omega =\overline{}h_a=i\underset{m\mathrm{\Delta }}{}d\theta _md\left(\frac{|a_mz^m|^2}{|z|_a^2}\right)$$
define a $`T_𝐑`$-invariant Kähler form on $`𝐏^\mathrm{𝟒}`$ representing the Kähler class corresponding to $`𝒪(5)`$. Notice that $`\mathrm{\Delta }`$ is the Newton polygon of $`𝒪(5)`$ on $`𝐏^\mathrm{𝟒}`$.
###### Lemma 6.1
The moment map is
$$F_a(x)=\underset{m\mathrm{\Delta }}{}\frac{|a_mz^m|^2}{|z|_a^2}m,$$
which maps $`𝐏^\mathrm{𝟒}`$ to $`\mathrm{\Delta }`$.
Near the large complex limit, under this moment map, the singular locus behave very nicely. This enable us to construct Lagrangian torus fibration with codimension 2 singular locus for generic quintic Calabi-Yau in .
For the case of general Calabi-Yau hypersurface in toric variety, $`p_w`$ ($`p_v`$) as a convex piecewise linear function on $`M`$ ($`N`$) with respect to $`\mathrm{\Delta }^{}`$ ($`\mathrm{\Delta }_{}^{}{}_{}{}^{}`$) can be naturally understood as a Kähler class for Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }^w}`$ ($`P_{\mathrm{\Sigma }^v}`$). Recall that under the monomial-divisor mirror map, the complex moduli of Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }^v}`$ is naturally identified with complexified Kähler cone of Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }^w}`$. Therefore $`p_w`$ also characterize complex moduli of Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }^v}`$. More precisely, for the Calabi-Yau hypersurface $`X_s`$ defined by
$$s=\underset{m(0)\mathrm{\Delta }}{}a_ms_m+\psi s_0=0,$$
$$p_w(m)=w_m=\mathrm{log}|a_m/\psi |.$$
The pair $`(p_w,p_v)`$ together characterize the whole moduli (including both the complex and the Kähler moduli) of a Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }^v}`$. For the purpose of our gradient flow, we need to construct a $`T_𝐑`$-invariant Kähler metric $`g`$ on $`P_{\mathrm{\Sigma }^v}`$ that depends on both $`p_v`$ and $`p_w`$. More precisely, we want the Kähler form of $`g`$ to correspond to $`p_v`$. Also to make the singular locus well behaved, we want the Kähler form to depend on complex moduli $`p_w`$ in a way similar to the quintic case.
Now we can define a natural $`T`$-invariant Kähler form on $`P_{\mathrm{\Sigma }^v}`$. $`mM`$ naturally define a holomorphic function $`s_m`$ on $`T=N_𝐂/N`$, $`|s_m|`$ is a $`T_𝐑`$-invariant function on $`T`$. For general $`mM_𝐑`$, $`s_m`$ will not make sense as function, but we always can define $`|s_m|`$ as a $`T_𝐑`$-invariant function on $`T`$. Let
$$h_{w,v}=\mathrm{log}(|s|_{w,v}^2)=\mathrm{log}\left(_{m\mathrm{\Delta }_v}|s_m|_w^2\right),$$
where $`|s_m|_w^2=|e^{w_m}s_m|^2`$.
$$\omega =\overline{}h_{w,v}=i_{m\mathrm{\Delta }_v}dm,\theta d\left(\frac{|s_m|_w^2}{|s|_{w,v}^2}\right)$$
define a $`T`$-invariant Kähler form on $`P_{\mathrm{\Sigma }^v}`$ representing the Kähler class corresponding to $`p_v`$.
###### Lemma 6.2
The moment map is
$$F_{w,v}(x)=_{m\mathrm{\Delta }_v}\frac{|s_m|_w^2}{|s|_{w,v}^2}m,$$
which maps $`P_{\mathrm{\Sigma }^v}`$ to $`\mathrm{\Delta }_v`$.
By this map, $`T_𝐑`$-invariant functions $`h_{w,v}`$, $`\frac{|s_m|_w^2}{|s|_{w,v}^2}`$ can all be viewed as function on $`\mathrm{\Delta }_v`$. We have
###### Lemma 6.3
$`\rho _m=\frac{|s_m|_w^2}{|s|_{w,v}^2}`$ as a function on $`\mathrm{\Delta }_v`$ archieves maximum exactly at $`m\mathrm{\Delta }_v`$.
Proof: By
$$x_k\frac{|s_m|_w^2}{x_k}=m,n_k|s_m|_w^2$$
$`\rho _m=\frac{|s_m|_w^2}{|s|_{w,v}^2}`$ archieves maximal implies
$$\underset{k=1}{\overset{r}{}}x_k\frac{\rho _m}{x_k}m_k=\frac{|s_m|_w^2}{|s|_{w,v}^2}_{m^{}\mathrm{\Delta }}(m,n_km^{},n_k)\frac{|s_m^{}|_w^2}{|s|_{w,v}^2}m_k$$
$$=\frac{|s_m|_w^2}{|s|_{w,v}^2}_{m^{}\mathrm{\Delta }}(mm^{})\frac{|s_m^{}|_w^2}{|s|_{w,v}^2}=\frac{|s_m|_w^2}{|s|_{w,v}^2}(mF_{w,v}(x))=0$$
Therefore
$$F_{w,v}(x)=m$$
when $`\rho _m=\frac{|s_m|_w^2}{|s|_{w,v}^2}`$ archieves maximal.
$`\mathrm{}`$
The moment maps naturally induce the following diagram
Recall that $`L_\mathrm{\Delta }`$ is the anti-canonical bundle of $`P_\mathrm{\Delta }`$. $`s_m`$ for $`m\mathrm{\Delta }`$ give us the set of $`T`$-invariant sections of $`L_\mathrm{\Delta }`$. $`\pi _v^{}L_\mathrm{\Delta }`$ is the anti-canonical bundle of $`P_{\mathrm{\Sigma }^v}`$. Consider a generic section of $`L_\mathrm{\Delta }`$
$$s=\underset{m(0)\mathrm{\Delta }}{}a_ms_m+\psi s_0,$$
$`s`$ can also be thought of as a section of $`\pi _v^{}L_\mathrm{\Delta }`$. Zero set $`X_s`$ of $`s`$ as a section of $`\pi _v^{}L_\mathrm{\Delta }`$ is a Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }^v}`$. When we vary $`s`$, we get the family of Calabi-Yau hypersurfaces corresponds to sections of $`L_\mathrm{\Delta }`$. When $`\psi `$ approach $`\mathrm{}`$, the corresponding Calabi-Yau hypersurface $`X_s`$ approach the “large complex limit” $`X_{\mathrm{}}`$ defined by $`s_0=0`$. The moment map $`F_{w,v}`$ naturally give us the Lagrangian fibration
$$F_{w,v}:X_{\mathrm{}}\mathrm{\Delta }_v$$
of the large complex limit $`X_{\mathrm{}}`$.
Consider the meromorphic function
$$q(z)=\frac{s_0}{s\psi s_0}$$
defined on $`P_\mathrm{\Sigma }`$. Let $`\omega `$ denote the Kähler form of a $`T_𝐑`$-invariant Kähler metric $`g`$ on $`P_\mathrm{\Sigma }`$, and $`f`$ denote the gradient vector field of real function $`f=Re(q)`$ with respect to the Kähler metric $`g`$. Clearly, $`X_{\mathrm{}}`$ is invariant under the action of $`T_𝐑`$. The fibres of the natural Lagrangian fibration of $`X_{\mathrm{}}`$ are exactly the orbits of $`T_𝐑`$-action. From our previous discussion, we know that the flow of $`V=\frac{f}{|f|^2}`$ moves $`X_{\mathrm{}}`$ to $`X_s`$ symplectically, therefore gives rise to Lagrangian fiberation $`F_s`$ of $`X_s`$ over $`\mathrm{\Delta }_v`$.
From our previous experience with Fermat type quintic Calabi-Yau and generic quintic Calabi-Yau, we know that the Lagrangian fibration we get by this way is not a $`C^{\mathrm{}}`$ Lagrangian fibration and usually with wrong type of singular fibers.
In the Fermat type quintic case, due to the highly symmetric nature of Fermat type quintic Calabi-Yau, it was relatively easy to figure out what is the expected $`C^{\mathrm{}}`$ Lagrangian (even special Lagrangian) fibration structure and corresponding singular fibres. Also in that case, it is easy to see explicitly that the singular locus of the Lagrangian fibration we constructed by gradient flow is a fattened version of the singular locus of the expected Lagrangian fibration. Therefore, even without perturbing to the $`C^{\mathrm{}}`$ Lagrangian fibration, we can already compute the expected monodromy.
In the case of general Calabi-Yau hypersurfaces in toric variety, just like the case of generic quintic Calabi-Yau discussed in , to carry out similar construction we just described for the Fermat type quintic, there are several obstacles. One of the major difficulty to generalize the discussion to the case of general Calabi-Yau hypersurfaces in toric variety is that for general Calabi-Yau hypersurfaces in toric variety the singular locus of the Lagrangian fibration constructed from deforming the standard Lagrangian fibration of $`X_{\mathrm{}}`$ via the flow of $`V`$ can be fairly arbitrary and does not necessarily resemble the fattening of any “expected singular locus”. Worst of all, in the case of general Calabi-Yau hypersurfaces in toric variety, there are no obvious guess what the “expected” $`C^{\mathrm{}}`$ Lagrangian (special Lagrangian) fiberation should be. One clearly expect the expected singular locus to be some graph in $`\mathrm{\Delta }S^3`$. But it seems to take some miracle (at least to me when I first dream about it) for a general singular set $`C`$ (which is an algebraic curve) to project to the singular locus $`\stackrel{~}{\mathrm{\Gamma }}=F(C)`$ that resemble a fattened graph.
Interestingly, miracle happens here! It largely relies on better understanding of what it means to be near the large complex limit. Philosophically speaking, it is commonly believed that the large complex limit corresponds to classical situation (in comparison to quantum situation). The closer one gets to the large complex limit, the smaller the quantum effects will be. Therefore we may explain the possible chaotic behavior of the singular locus $`\stackrel{~}{\mathrm{\Gamma }}=F(C)`$ by quantum effect. To make the singular locus behave nicely to resemble a one-dimensional graph, we simply need to get really close to the large complex limit $`X_{\mathrm{}}`$. Therefore the qestion remain is: how to get close to the large complex limit?
Before getting into the detail, let’s recall facts of our gradient flow. One way to understand $`V`$ is to realize that $`dq`$ is a meromorphic section of $`N_X^{}`$, $`(dq)^1`$ is a natural meromorphic section of $`N_X`$. Notice the exact sequence
$$0T_XT_{P_\mathrm{\Sigma }}|_XN_X0$$
With respect to the Kähler metric $`g`$ on $`T_{P_\mathrm{\Sigma }}`$, the exact sequence has a natural (non-holomorphic) splitting
$$T_{P_\mathrm{\Sigma }}|_X=T_XN_X$$
$`V`$ is just real part of the natural lift of $`(dq)^1`$ via this splitting. $`V`$ is singular exactly when $`(dq)^1`$ is singular or equivalently, when $`dq=0`$, which corresponds to singular part of $`X_{\mathrm{}}`$. On the other hand, the union of the Lagrangian 3-torus non-singular fibers of $`X_{\mathrm{}}`$ is exactly the smooth part of $`X_{\mathrm{}}`$. Since $`V`$ is non-singular on the smooth part of $`X_{\mathrm{}}`$, all these 3-torus fibers will be carried to $`X_s`$ nicely by the flow of $`V`$.
Now we will try to understand how the gradient flow of $`f`$ behaves at singularities of $`X_{\mathrm{}}`$. As we know, $`X_{\mathrm{}}`$ has only normal crossing singularities. $`\mathrm{Sing}(X_{\mathrm{}})`$ is a union of complex torus (orbits of $`T`$) of dimension less or equal to $`r2`$. According to general results in about the gradient vector field, we have
###### Proposition 6.1
Under the flow of $`V`$, points in $`X_{\mathrm{}}X_s`$ will be fixed. Points in $`\mathrm{Sing}(X_{\mathrm{}})\backslash X_s`$ that is located in a $`l`$-dimensional complex torus in $`\mathrm{Sing}(X_{\mathrm{}})`$ will bubble off to a $`(rl1)`$-dimensional real torus in $`X_s`$.
From this general structure result on gradient flow, it is easy to pin down the topological singular set $`C`$ of the fibration $`F_s`$.
###### Corollary 6.1
The topological singular set of the Lagrangian torus fibration $`F_s`$ is $`C=\mathrm{Sing}(X_{\mathrm{}})X_s`$, and the corresponding singular locus is $`\stackrel{~}{\mathrm{\Gamma }}=F_s(C)`$.
Recall that $`T_𝐑`$-orbits form the fibres of the Lagrangian fibration $`F_{w,v}:X_{\mathrm{}}\mathrm{\Delta }_v`$. By proposition 6.1, a $`T_𝐑`$-orbit in $`\mathrm{Sing}(X_{\mathrm{}})`$ that does not intersect $`C`$ will deform to a Lagrangian 3-torus in $`X_s`$ under the flow of $`V`$. $`T_𝐑`$-orbits intersecting $`C`$ will deform to singular fibres of the Lagrangian fibration $`F_s:X_s\mathrm{\Delta }_v`$. The actural singular fibre type is determined by the way the $`T_𝐑`$-orbit is intersecting $`C`$. To further discuss singular fibre type, it is important to understand better the singular set $`C`$ and the corresponding singular locus $`\stackrel{~}{\mathrm{\Gamma }}`$.
In general, for any $`\sigma \mathrm{\Sigma }`$, we have the quotient fan $`\mathrm{\Sigma }_\sigma `$, $`P_{\mathrm{\Sigma }_\sigma }`$ is the clossure of $`T_\sigma `$ in $`P_\mathrm{\Sigma }`$. $`\mathrm{\Delta }_\sigma ^{}=\sigma ^{}(\mathrm{\Delta }\backslash \{0\})`$ is a subface of $`\mathrm{\Delta }`$ and is the Newton polyhedron of $`T_\sigma `$-invariant sections of $`L_\mathrm{\Delta }|_{P_{\mathrm{\Sigma }_\sigma }}`$. $`s_w`$ is restrict to
$$s_{w,\sigma }=\underset{m\mathrm{\Delta }_\sigma ^{}}{}e^{w_m}s_m$$
on $`P_{\mathrm{\Sigma }_\sigma }`$. Let $`C_\sigma =\{s_{w,\sigma }^1(0)\}`$.
The singular set $`C`$ can be understood as the curve in the 2-skeleton $`X_{\mathrm{}}^{(2)}=\mathrm{Sing}(X_{\mathrm{}})`$ of $`X_{\mathrm{}}`$ that is cut out by $`s=0`$. For 2-skeleton, we have the following diagram
$`P_\mathrm{\Sigma }^{(2)}`$ is a union of complex torus of dimension less or equal to 2. Generic section $`s`$ can avoid verteces (the zero dimensional tori) and will intersect complex torus of dimension 1 and 2. In general
$$dim\mathrm{\Delta }_\sigma ^{}+dim\sigma =r.$$
Therefore, for $`\sigma \mathrm{\Sigma }(3)`$, $`dim_𝐂(C_\sigma )=0`$, for $`\sigma \mathrm{\Sigma }(2)`$, $`dim_𝐂(C_\sigma )=1`$. Notice that $`\mathrm{Sing}(X_{\mathrm{}})=P_{\mathrm{\Sigma }^v}^{(2)}`$. The singular set $`C`$ can be decomposed into two parts $`C=C_1C_2`$, where
$$C_1=\underset{\sigma \mathrm{\Sigma }(3)}{}(\pi _v^{(2)})^1(C_\sigma )$$
$$C_2=\underset{\sigma \mathrm{\Sigma }(2)}{}(\pi _v^{(2)})^1(C_\sigma )$$
The corresponding singular locus $`\stackrel{~}{\mathrm{\Gamma }}=\stackrel{~}{\mathrm{\Gamma }}_1\stackrel{~}{\mathrm{\Gamma }}_2`$. It is easy to see that $`dim\stackrel{~}{\mathrm{\Gamma }}_1=1`$ and $`dim\stackrel{~}{\mathrm{\Gamma }}_2=2`$. Let $`\stackrel{~}{\mathrm{\Gamma }}^0`$ denotes the set of verteces of $`\stackrel{~}{\mathrm{\Gamma }}`$, $`\stackrel{~}{\mathrm{\Gamma }}_1^1=\stackrel{~}{\mathrm{\Gamma }}_1\backslash \stackrel{~}{\mathrm{\Gamma }}^0`$, $`\stackrel{~}{\mathrm{\Gamma }}_2^2`$ denotes the interior of $`\stackrel{~}{\mathrm{\Gamma }}_2`$, and $`\stackrel{~}{\mathrm{\Gamma }}_2^1=\stackrel{~}{\mathrm{\Gamma }}_2\backslash (\stackrel{~}{\mathrm{\Gamma }}_2^0\stackrel{~}{\mathrm{\Gamma }}_2^2)`$. When $`X_s`$ is near the large complex limit (we will descuss in more detail in a moment), under the moment map, $`C_2`$ is 2-1 over $`\stackrel{~}{\mathrm{\Gamma }}_2^2`$ and 1-1 over $`\stackrel{~}{\mathrm{\Gamma }}_2^1`$.
For the fibration $`F_{w,v}:X_{\mathrm{}}\mathrm{\Delta }_v`$, when $`X_s`$ is near the large complex limit, fibre over a point in $`\stackrel{~}{\mathrm{\Gamma }}^0`$ is a circle that intersects $`C`$ at a point, fibre over a point in $`\stackrel{~}{\mathrm{\Gamma }}_1^1`$ is a 2-torus that intersects $`C`$ at a circle, fibre over a point in $`\stackrel{~}{\mathrm{\Gamma }}_2^1`$ is a 2-torus that intersects $`C`$ at a point, fibre over a point in $`\stackrel{~}{\mathrm{\Gamma }}_2^2`$ is a 2-torus that intersects $`C`$ at 2 points. By proposition 6.1, we can pin down all the singular fibres of the Lagrangian fibration $`F_s`$.
###### Theorem 6.1
When $`X_s`$ is near the large complex limit, flow of $`V`$ will produce a Lagrangian fibration $`F_s:X_s\mathrm{\Delta }_v`$. There are 5 types of fibers.
(i). For $`p\mathrm{\Delta }\backslash \stackrel{~}{\mathrm{\Gamma }}`$, $`F^1(p)`$ is a Lagrangian 3-torus.
(ii). For $`p\stackrel{~}{\mathrm{\Gamma }}_2^2`$, $`F_s^1(p)`$ is a Lagrangian 3-torus with $`2`$ circles collapsed to $`2`$ singular points.
(iii). For $`p\stackrel{~}{\mathrm{\Gamma }}_2^1`$, $`F_s^1(p)`$ is a Lagrangian 3-torus with $`1`$ circle collapsed to $`1`$ singular point.
(iv). For $`p\stackrel{~}{\mathrm{\Gamma }}^0`$, $`F_s^1(p)`$ is a Lagrangian 3-torus with $`1`$ 2-torus collapsed to $`1`$ singular point (the type III singular fibre).
(v). For $`p\stackrel{~}{\mathrm{\Gamma }}_1^1`$, $`F_s^1(p)`$ is a type I singular fibre (a circle times a rational curve with one node).
The Lagrangian torus fibration $`F_s:X_s\mathrm{\Delta }_v`$ we constructed by flow of $`V`$ is not really the kind of Lagrangian torus fibration that is expected by the SYZ mirror conjecture. SYZ construction generally requires that singular locus of the special Lagrangian fibration to be some one dimensional graph in $`S^3`$. For our purpose we want to construct $`C^{\mathrm{}}`$ Lagrangian fibrations with dimension one singular locus that we believe will capture symplectic topological structure of special Lagrangian fibrations in general. Or at least, we want to construct a Lagrangian fibration with dimension one singular locus. To achieve these, in the Fermat type quintic case, the singular locus can easily be seen as a fattening of a one dimensional graph $`\mathrm{\Gamma }`$. In , we use symplectic technique to explicitely deform the Lagrangian fibration before we deform acccording the flow to ensure that the singular locus $`F(C)=\mathrm{\Gamma }`$. Then the corresponding Lagrangian fibration of $`X_s`$ we construct by the flow of $`V`$ will have dimension one singular locus. In , we also discussed the question of how to deform this Lagrangian fibration into a $`C^{\mathrm{}}`$ Lagrangian fibration.
In the case of general Calabi-Yau hypersurface in toric variety, we can do exactly the same thing, granted that we can realize the singular locus $`F_s(C)`$ as a fattening of some graph $`\mathrm{\Gamma }`$ and be able to explicitly construct symplectic deformation, which deform $`C`$ to satisfy $`F_s(C)=\mathrm{\Gamma }`$. Recall that $`C=C_1C_2`$, and $`\stackrel{~}{\mathrm{\Gamma }}_1=F_s(C_1)`$ is already of dimension 1. We only need to concentrate on the $`C_2`$ part. Since $`C_2`$ is reducible, and each irreducible component is in a 2-dimensional toric subvariety $`P_{\mathrm{\Sigma }_\tau ^v}`$ of $`P_{\mathrm{\Sigma }^v}`$ that is over a 2-dimensional toric subvariety $`P_{\mathrm{\Sigma }_\sigma }`$ of $`P_\mathrm{\Sigma }`$. according to , the problem can be isolated to each $`P_{\mathrm{\Sigma }_\tau ^v}`$. More precisely, for $`\tau \sigma `$, $`\tau \mathrm{\Sigma }^v(2)`$, $`\sigma \mathrm{\Sigma }(2)`$, $`\pi _v:P_{\mathrm{\Sigma }_\tau ^v}P_{\mathrm{\Sigma }_\sigma }`$ is a resolution. For $`X_s`$ generic, the restriction $`\pi _v:C_2P_{\mathrm{\Sigma }_\tau ^v}C_2P_{\mathrm{\Sigma }_\sigma }`$ is an isomorphism. The problem can further be reduced to discussion over $`P_{\mathrm{\Sigma }_\sigma }P_\mathrm{\Sigma }`$ as the following.
Problem: Given an integral polyhedron $`\mathrm{\Delta }M`$, one can naturally construct a fan $`\mathrm{\Sigma }`$ and an ample line bundle $`L_\mathrm{\Delta }`$ over the toric variety $`P_\mathrm{\Sigma }`$, such that the Newton polygon of sections of $`L_\mathrm{\Delta }`$ is exactly $`\mathrm{\Delta }`$. Let $`F:P_\mathrm{\Sigma }\mathrm{\Delta }`$ be the moment map with respect to the class of $`L_\mathrm{\Delta }`$. We need a curve $`C_s=\{s^1(0)\}`$ in $`P_\mathrm{\Sigma }`$, where $`s`$ is a section of $`L_\mathrm{\Delta }`$, such that $`F(C_s)`$ is a fattening of some graph $`\mathrm{\Gamma }`$ and we want to explicitly construct symplectic deformation of $`P_\mathrm{\Sigma }`$, which deform $`C`$ to satisfy $`F(C_s)=\mathrm{\Gamma }`$.
Clearly, one can not expect that $`C_s`$ for all section $`s`$ of $`L_\mathrm{\Delta }`$ will have such a nice property. As we point out earlier, it turns out, when the Calabi-Yau hypersurface in toric variety is generic and close to the large complex limit in a certain sense, the corresponding singular set curves $`C_\sigma `$ in $`P_{\mathrm{\Sigma }_\sigma }`$ have the properties described in above problem. This kind of curve and more general situation have been discussed intensively in our paper . To describe the result from there, let’s first introduce some notation and definations.
### 6.2 Newton polygon and string diagram
Given an integral polyhedron $`\mathrm{\Delta }M`$, one can naturally construct a fan $`\mathrm{\Sigma }`$ and an ample line bundle $`L_\mathrm{\Delta }`$ over the toric variety $`P_\mathrm{\Sigma }`$, such that the Newton polygon of sections of $`L_\mathrm{\Delta }`$ is exactly $`\mathrm{\Delta }`$. In this context, we assume that $`r=\mathrm{rk}(M)=2`$.
Consider a general section of $`L_\mathrm{\Delta }`$
$$s=\underset{m\mathrm{\Delta }}{}a_ms_m$$
Let $`|a_m|=e^{w_m}`$, then $`w=(w_m)_{m\mathrm{\Delta }}`$ define a function on integral points in $`\mathrm{\Delta }`$ (thinking of $`\mathrm{\Delta }`$ as a real triangle in $`M_𝐑=M𝐑`$).
###### Definition 6.1
$`w=(w_m)_{m\mathrm{\Delta }}`$ is called convex on $`\mathrm{\Delta }`$, if for any $`m^{}\mathrm{\Delta }`$ there exist an affine function $`n`$ such that $`n(m^{})=w_m^{}`$ and $`n(m)w_m`$ for $`m\mathrm{\Delta }\backslash \{m^{}\}`$.
We will always assume that $`w`$ is convex. With $`w`$ we can define the moment map
$$F_w(x)=\underset{m\mathrm{\Delta }}{}\frac{|s_m|_w^2}{|x|_w^2}m,$$
which maps $`P_\mathrm{\Sigma }`$ to $`\mathrm{\Delta }`$, where $`|s_m|_w^2=|e^{w_m}s_m|^2`$, $`|s|_w^2=_{m\mathrm{\Delta }}|s_m|_w^2`$.
$`\mathrm{\Delta }`$ can also be thought of as a real triangle in $`M_𝐑=M𝐑`$. Then $`w=(w_m)_{m\mathrm{\Delta }}`$ define a function on integral points in $`\mathrm{\Delta }`$. If $`w`$ is convex, then $`w`$ can be extended to a piecewise linear convex function on $`\mathrm{\Delta }`$. We will denote the extension also by $`w`$. Generic $`w`$ will determine a simplicial decomposition of $`\mathrm{\Delta }`$, with zero simplices being integral points in $`\mathrm{\Delta }`$. In this case, we say the piecewise linear convex function $`w`$ is compatible with the simplicial decomposition of $`\mathrm{\Delta }`$, and the simplicial decomposition of $`\mathrm{\Delta }`$ is determined by $`w`$. Consider the baricenter subdivision of this simplicial decomposition of $`\mathrm{\Delta }`$, let $`\mathrm{\Gamma }_w`$ denote the union of simplices in the baricenter subdivision that do not intersect integral points in $`\mathrm{\Delta }`$. Then it is not hard to see that $`\mathrm{\Gamma }_w`$ is an one dimensional graph. And $`\mathrm{\Gamma }_w`$ divide $`\mathrm{\Delta }`$ into regions, with a unique integral point of $`\mathrm{\Delta }`$ located at the center of each of this region. In particular we can think of the regions as parametrized by integral points in $`\mathrm{\Delta }`$.
Let $`C`$ denotes the curve defined by $`s=0`$, the results in give the following.
###### Theorem 6.2
For $`w=(w_m)_{m\mathrm{\Delta }}`$ convex and positive, and $`|w|`$ large enough, $`F_w(C)`$ will be a fattening of graph $`\mathrm{\Gamma }_w`$. There exist a symplectic diffeomorphism that map $`C`$ to $`\stackrel{~}{C}`$ such that $`F_w(\stackrel{~}{C})=\mathrm{\Gamma }_w`$.
Remark: The moment map $`F_w`$ is invariant under the real 2-torus action. For any other moment map $`F`$ that is invariant under the real 2-torus action, we also have that $`F(\stackrel{~}{C})`$ is a 1-dimensional graph.
Example: For the standard simplicial decomposition of the Newton polygon $`\mathrm{\Delta }_5`$ of quintic polyonmials,
Figure 1: the standard simplicial decomposition
We have the corresponding $`\widehat{\mathrm{\Gamma }}=F(C)`$
Figure 2: $`\stackrel{~}{\mathrm{\Gamma }}`$ for the standard simplicial decomposition
which is a fattening of the following graph $`\mathrm{\Gamma }`$. By results in , we can simplectically deform $`C`$ to symplectic curve $`\widehat{C}`$ such that $`F(\widehat{C})=\mathrm{\Gamma }`$
Figure 3: $`\mathrm{\Gamma }`$ for the standard simplicial decomposition
If the simplicial decomposition is changed to
Figure 4: alternative simplicial decomposition
We have the corresponding $`\mathrm{\Gamma }`$
Figure 5: $`\mathrm{\Gamma }`$ for alternative simplicial decomposition
Change the simplicial decomposition further, we will get
Figure 6: another alternative simplicial decomposition
We have the corresponding $`\mathrm{\Gamma }`$
Figure 7: the corresponding $`\mathrm{\Gamma }`$
### 6.3 Lagrangian fibration with codimension 2 singular locus
With all this preparation, now we would like to address the meaning of near the large complex limit. Consider the Calabi-Yau hypersurface $`X_s`$ in $`P_\mathrm{\Sigma }`$ defined by
$$s=\underset{m(0)\mathrm{\Delta }}{}a_ms_m+\psi s_0=0$$
Let $`\mathrm{\Delta }^0`$ denote the set of integral points in the 2-skeleton of $`\mathrm{\Delta }`$. Or in another word, the integral points in $`\mathrm{\Delta }`$ that is not in the interior of 3-faces and also not the origion. For $`m\mathrm{\Delta }`$, let $`|a_m|=e^{w_m}`$, $`w_m^{}=w_mw_0`$ (recall $`\psi =a_0`$). We have two functions $`w=(w_m)_{m\mathrm{\Delta }}`$, $`w^{}=(w_m^{})_{m\mathrm{\Delta }}`$ defined on the integral points in $`\mathrm{\Delta }`$.
###### Definition 6.2
$`w^{}=(w_m^{})_{m\mathrm{\Delta }}`$ is called convex with respect to $`\mathrm{\Delta }^0`$, if for any $`m^{}\mathrm{\Delta }^0`$ there exist a linear function $`n`$ such that $`n(m^{})=w_m^{}^{}`$ and $`n(m)w_m`$ for $`m\mathrm{\Delta }^0\backslash \{m^{}\}`$.
The two convexities we defined are closely related. The 2-skeleton
$$\mathrm{\Delta }^0=\underset{\sigma \mathrm{\Sigma }(2)}{}\mathrm{\Delta }_\sigma ^{}.$$
Let
$$w_\sigma =w|_{\mathrm{\Delta }_\sigma ^{}}.$$
Then we have the following lemma
###### Lemma 6.4
$`w^{}=(w_m^{})_{m\mathrm{\Delta }}`$ is convex with respect to $`\mathrm{\Delta }^0`$ for $`w_{m_0}`$ large enough, if and only if $`w_\sigma `$ is convex on $`\mathrm{\Delta }_\sigma ^{}`$ in the sense of definition 6.1 for all $`\sigma \mathrm{\Sigma }(2)`$.
With this lemma in mind, we can make the concept of near the large complex limit more precise as follows
###### Definition 6.3
The quintic Calabi-Yau hypersurface $`X_s`$ is said to be near the large complex limit, if $`w^{}=(w_m^{})_{m\mathrm{\Delta }}`$ is convex with respect to $`\mathrm{\Delta }^0`$ and $`|w|`$ is large.
When $`X_s`$ is near the large complex limit, the corresponding $`w_\sigma `$ is convex on $`\mathrm{\Delta }_\sigma ^{}`$. When $`w`$ is generic, by previous construction, $`w_\sigma `$ determine a 1-dimensional graph $`\mathrm{\Gamma }_\sigma `$ on $`\mathrm{\Delta }_\sigma ^{}`$. Let
$$\mathrm{\Gamma }=\underset{\sigma \mathrm{\Sigma }(2)}{}(\widehat{\pi }_v^{(2)})^1(\mathrm{\Gamma }_\sigma )$$
By theorem 6.2, $`\stackrel{~}{\mathrm{\Gamma }}_\sigma =F(C_\sigma )`$ is a fattening of $`\mathrm{\Gamma }_\sigma `$. Therefore,
$$\stackrel{~}{\mathrm{\Gamma }}=\underset{\sigma \mathrm{\Sigma }(2)}{}(\widehat{\pi }_v^{(2)})^1(\stackrel{~}{\mathrm{\Gamma }}_\sigma )$$
is a fattening of 1-dimensional graph $`\mathrm{\Gamma }`$. Let $`\mathrm{\Gamma }=\mathrm{\Gamma }^1\mathrm{\Gamma }^2\mathrm{\Gamma }^3`$, where $`\mathrm{\Gamma }^1`$ is the smooth part of $`\mathrm{\Gamma }`$, $`\mathrm{\Gamma }^2`$ is the singular part of $`\mathrm{\Gamma }`$ that will map to the interior of the 2-skeleton of $`\mathrm{\Delta }`$ under $`\widehat{\pi }_v`$, $`\mathrm{\Gamma }^3`$ is the singular part of $`\mathrm{\Gamma }`$ that will map to the 1-skeleton of $`\mathrm{\Delta }`$ under $`\widehat{\pi }_v`$. Apply second part of the theorem 6.2, also with help of some other result in , we can produce a Lagrangian fibration $`\widehat{F}:X_{\mathrm{}}\mathrm{\Delta }_v`$ such that $`\widehat{F}(C)=\mathrm{\Gamma }`$. Then we have
###### Theorem 6.3
When $`X_\psi `$ is near the large complex limit, start with Lagrangian fibration $`\widehat{F}`$ the flow of $`V`$ will produce a Lagrangian fibration $`\widehat{F}_s:X_s\mathrm{\Delta }_v`$. There are 4 types of fibres.
(i). For $`p\mathrm{\Delta }_v\backslash \mathrm{\Gamma }`$, $`\widehat{F}_s^1(p)`$ is a smooth Lagrangian 3-torus.
(ii). For $`p\mathrm{\Gamma }^1`$, $`\widehat{F}_s^1(p)`$ is a type I singular fibre.
(iii). For $`p\mathrm{\Gamma }^2`$, $`\widehat{F}_s^1(p)`$ is a type II singular fibre.
(iv). For $`p\mathrm{\Gamma }^3`$, $`\widehat{F}_s^1(p)`$ is a type III singular fibre.
Remark: Although the Lagrangian fibration produced by this theorem have exactly the same topological structure as the expected special Lagrangian fibration, this fibration map is still not $`C^{\mathrm{}}`$ map (merely Lipschitz). But I believe that a small pertubation of this map will give a $`C^{\mathrm{}}`$ Lagrangian fibration with the same topological structure. In , we were able to make the map $`C^{\mathrm{}}`$ away from $`C`$. We also indicated how to modify singular fibres around $`C`$. Yet, we still fall short to completely make the fibration $`C^{\mathrm{}}`$. In any case, we do not really use $`C^{\mathrm{}}`$ property of the fibration map in the following discussion.
## 7 Duality map of convex polyhedrons
As we know, $`(p_v,p_w)`$ parametrizes Kähler moduli and complex moduli of Calabi-Yau hypersurface $`X`$ in $`P_{\mathrm{\Sigma }^v}`$. From symmetric point of view, $`(p_v,p_w)`$ also parametrizes complex moduli and Kähler moduli of the mirror Calabi-Yau hypersurface $`Y`$ in $`P_{\mathrm{\Sigma }^w}`$. According to the previous section, we can construct Lagrangian torus fibrations
$$\widehat{F}_{w,v}:X\mathrm{\Delta }_v,\widehat{F}_{v,w}:Y\mathrm{\Delta }_w.$$
with singular locuses $`\mathrm{\Gamma }_{w,v}`$ and $`\mathrm{\Gamma }_{v,w}`$.
One key ingredient of SYZ mirror conjecture is an identification of the bases of the fibrations $`\varphi :\mathrm{\Delta }_w\mathrm{\Delta }_v`$, such that $`\varphi (\mathrm{\Gamma }_{v,w})=\mathrm{\Gamma }_{w,v}`$. It turns out that this identification can be constructed purely combinatorically depending on the two convex piecewise linear convex functions $`p_v,p_w`$, without refering to symplectic form, Calabi-Yau etc. We will descuss this purely combinatoric construction in this section.
Assume that $`\mathrm{\Delta }`$ is a reflexive polyhedron, $`\mathrm{\Sigma }`$ the corresponding fan. Faces of $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$) is in 1-1 correspondence with cones in $`\mathrm{\Sigma }^{}`$ ($`\mathrm{\Sigma }`$). For a cone $`\sigma `$ in $`\mathrm{\Sigma }^{}`$ ($`\mathrm{\Sigma }`$), let $`\mathrm{\Delta }_\sigma =\sigma (\mathrm{\Delta }\backslash \{0\})`$ ($`(\mathrm{\Delta }^{})_\sigma =\sigma (\mathrm{\Delta }^{}\backslash \{0\})`$) be the corresponding face of $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$). This correspondence preserve the inclusion relation. There is a natural dual correspondence between faces of $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$. For $`\mathrm{\Delta }_\sigma `$ a face of $`\mathrm{\Delta }`$, the dual face
$$(\mathrm{\Delta }_\sigma )^{}=\{n\mathrm{\Delta }^{}|m,n=1,\mathrm{for}m\mathrm{\Delta }_\sigma \}.$$
is a face of $`\mathrm{\Delta }^{}`$. Then there exist cone $`\sigma ^{}\mathrm{\Sigma }`$ such that $`(\mathrm{\Delta }_\sigma )^{}=\mathrm{\Delta }_\sigma ^{}^{}`$ In this way, there is a natural identification of faces of $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$, or equivalently, a natural identification of the sets $`\mathrm{\Sigma }^{}`$ and $`\mathrm{\Sigma }`$. These identifications reverse the inclusion relation, which is a partial relation.
A simplex in the baricenter subdivision of $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$) can be expressed by an inclusion related sequence of cones $`\{\sigma _k\}_{k=0}^l`$ in $`\mathrm{\Sigma }^{}`$ ($`\mathrm{\Sigma }`$) that satisfies $`\sigma _k\sigma _{k+1}`$. Here $`\sigma _k`$ is best understood to represent the baricenter of $`\mathrm{\Delta }_{\sigma _k}`$. $`\widehat{\sigma }=\{\sigma _k\}_{k=0}^l`$ denotes the $`l`$-simplex $`\widehat{\sigma }`$ with vertices being the baricenters of $`\mathrm{\Delta }_{\sigma _k}`$ for $`k=0,\mathrm{},l`$.
###### Proposition 7.1
The dual correspondence induces naturally a piecewise linear homeomorphism $`\mathrm{\Delta }\mathrm{\Delta }^{}`$ with respect to the baricenter subdivision. Where baricenter of $`\mathrm{\Delta }_\sigma `$ is map to baricenter of $`\mathrm{\Delta }_\sigma ^{}^{}`$, and simplex $`\widehat{\sigma }=\{\sigma _k\}_{k=0}^l`$ is mapped to $`\widehat{\sigma }^{}=\{\sigma _k^{}\}_{k=0}^l`$ linearly.
From now on, we will introduce the notation $`\widehat{\mathrm{\Sigma }}^{}`$ ($`\widehat{\mathrm{\Sigma }}`$) (not as a fan) to index simpleces $`\widehat{\sigma }=\{\sigma _k\}_{k=0}^l`$ ($`\widehat{\sigma }^{}=\{\sigma _k^{}\}_{k=0}^l`$) in the baricenter subdivision of $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$).
Assume that $`p_w`$ ($`p_v`$) is a convex piecewise linear function on $`M`$ ($`N`$) with respect to $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$). $`p_w(m)=w_m`$ ($`p_v(n)=v_n`$) for $`m\mathrm{\Delta }\backslash \{0\}`$ ($`n\mathrm{\Delta }^{}\backslash \{0\}`$). $`p_w`$ ($`p_v`$) naturally determines a fan $`\mathrm{\Sigma }^w`$ ($`\mathrm{\Sigma }^v`$) for $`M`$ ($`N`$) that is compatible with $`\mathrm{\Delta }^{}`$ ($`\mathrm{\Delta }`$). $`p_w`$ ($`p_v`$) also naturally determines a polyhedron $`\mathrm{\Delta }^wN`$ ($`\mathrm{\Delta }^vM`$) consists of $`nN`$ ($`mM`$) that as a linear function on $`M`$ ($`N`$) is greater or equal to $`p_w`$ ($`p_v`$). In particular, the fan $`\mathrm{\Sigma }`$ ($`\mathrm{\Sigma }^{}`$) of the anti-canonical model $`P_\mathrm{\Sigma }`$ ($`P_\mathrm{\Sigma }^{}`$) is determined by $`p_\mathrm{\Delta }`$ ($`p_\mathrm{\Delta }^{}`$), $`p_\mathrm{\Delta }(n)=1`$ ($`p_\mathrm{\Delta }^{}(m)=1`$) for $`n\mathrm{\Delta }^{}\backslash \{0\}`$ ($`m\mathrm{\Delta }\backslash \{0\}`$), for which the corresponding polyhedron is $`\mathrm{\Delta }`$ ($`\mathrm{\Delta }^{}`$).
$`p_w`$ ($`p_v`$) determines a polyhedron decomposition $`Z^w`$ ($`Z^v`$) of $`\mathrm{\Delta }`$ ($`(\mathrm{\Delta }^{})`$). Let $`\widehat{Z}^w`$ ($`\widehat{Z}^v`$) be the baricenter subdivision of $`Z^w`$ ($`Z^v`$).
Restrict to each face, for $`\sigma \mathrm{\Sigma }^{}`$, $`p_w`$ ($`p_v`$) determines a polyhedron decomposition $`Z_\sigma ^w`$ ($`Z_\sigma ^{}^v`$) of $`\mathrm{\Delta }_\sigma `$ ($`(\mathrm{\Delta }^{})_\sigma ^{}`$). Here $`Z_\sigma ^w`$ ($`Z_\sigma ^{}^v`$) only consists of those open polyhedrons whose baricenters are in the interior of $`\mathrm{\Delta }_\sigma `$ ($`(\mathrm{\Delta }^{})_\sigma ^{}`$). We have
$$Z^w=\underset{\sigma \mathrm{\Sigma }}{}Z_\sigma ^{}^w$$
$$Z^v=\underset{\sigma \mathrm{\Sigma }}{}Z_\sigma ^v$$
Let $`\widehat{Z}_\sigma ^w`$ ($`\widehat{Z}_\sigma ^{}^v`$) be the baricenter subdivision of $`Z_\sigma ^w`$ ($`Z_\sigma ^{}^v`$), and simplices in $`\widehat{Z}_\sigma ^w`$ ($`\widehat{Z}_\sigma ^{}^v`$) have vertices in $`Z_\sigma ^w`$ ($`Z_\sigma ^{}^v`$).
For $`\widehat{\sigma }=\{\sigma _k\}_{k=0}^l\widehat{\mathrm{\Sigma }}^{}`$ ($`\widehat{\sigma }^{}=\{\sigma _k^{}\}_{k=0}^l\widehat{\mathrm{\Sigma }}`$), define
$$\widehat{Z}_{\widehat{\sigma }}^w=\{\widehat{\alpha }=\{\widehat{\alpha }_0,\mathrm{},\widehat{\alpha }_l\}|\widehat{\alpha }_k\widehat{Z}_{\sigma _k}^w,\widehat{\alpha }_k<\widehat{\alpha }_{k+1},\mathrm{for}\mathrm{all}k.\}$$
$$\widehat{Z}_{\widehat{\sigma }^{}}^v=\{\widehat{\alpha }=\{\widehat{\alpha }_0,\mathrm{},\widehat{\alpha }_l\}|\widehat{\alpha }_k\widehat{Z}_{\sigma _k^{}}^v,\widehat{\alpha }_k<\widehat{\alpha }_{k1},\mathrm{for}\mathrm{all}k.\}$$
where $`\widehat{\alpha }<\widehat{\beta }`$ means that each polyhedron in $`\widehat{\alpha }`$ is a face of every polyhedron in $`\widehat{\beta }`$. Then
$$\widehat{Z}^w=\underset{\widehat{\sigma }\widehat{\mathrm{\Sigma }}}{}\widehat{Z}_{\widehat{\sigma }^{}}^w$$
$$\widehat{Z}^v=\underset{\widehat{\sigma }\widehat{\mathrm{\Sigma }}}{}\widehat{Z}_{\widehat{\sigma }}^v$$
Now we are ready to discuss the polyhedron decomposition on $`\mathrm{\Delta }^v`$. Consider the piecewise linear map $`\pi _v:\mathrm{\Delta }^v\mathrm{\Delta }`$, for $`\sigma \mathrm{\Sigma }^{}`$, $`\pi _v`$ over the interior of $`\mathrm{\Delta }_\sigma `$ is a fibration with compact fibres. A simplex in a fibre is corresponding to a simplex of complimentary dimension in $`Z_\sigma ^{}^v`$. In this way, we may identify the baricenter subdivision of a fibre by $`\widehat{Z}_\sigma ^{}^v`$. (the reason for baricenter subdivision is that after it the correspondence will preserve the dimension of the simplex, instead of go to the complimentary dimension.) On the other hand, $`\widehat{Z}_\sigma ^w`$ form a simplicial decomposition of interior of $`\mathrm{\Delta }_\sigma `$. For $`\widehat{\alpha }\widehat{Z}_\sigma ^w`$, $`\widehat{\beta }\widehat{Z}_\sigma ^{}^v`$, $`\pi _v^1(\widehat{\alpha })\widehat{\beta }\widehat{\alpha }\times \widehat{\beta }`$ form a polyhedron $`\mathrm{\Delta }_{\widehat{\alpha },\widehat{\beta }}^v`$ in $`\mathrm{\Delta }^v`$. Let $`\widehat{Z}_\sigma ^{v,w}\widehat{Z}_\sigma ^w\times \widehat{Z}_\sigma ^{}^v`$ represent the set of this kind of polyhedron in $`\mathrm{\Delta }^v`$. Then for $`\widehat{\sigma }=\{\sigma _k\}_{k=0}^l\widehat{\mathrm{\Sigma }}^{}`$, define
$$\widehat{Z}_{\widehat{\sigma }}^{v,w}=\{\widehat{\gamma }=\{\widehat{\alpha }_0\times \widehat{\beta }_0,\mathrm{},\widehat{\alpha }_l\times \widehat{\beta }_l\}|\widehat{\alpha }_k\times \widehat{\beta }_k\widehat{Z}_{\sigma _k}^{v,w},\widehat{\alpha }_k<\widehat{\alpha }_{k+1},\widehat{\beta }_{k+1}<\widehat{\beta }_k\mathrm{for}\mathrm{all}k.\}$$
$$\widehat{Z}^{v,w}=\underset{\widehat{\sigma }\widehat{\mathrm{\Sigma }}^{}}{}\widehat{Z}_{\widehat{\sigma }}^{v,w}$$
Then $`\widehat{Z}^{v,w}`$ forms a polyhedron decomposition of $`\mathrm{\Delta }^v`$.
$$\widehat{\gamma }^{}=\{\widehat{\alpha }_0^{}\times \widehat{\beta }_0^{},\mathrm{},\widehat{\alpha }_l^{}^{}\times \widehat{\beta }_l^{}^{}\}\widehat{Z}_{\widehat{\sigma }^{}}^{v,w}$$
is a face of
$$\widehat{\gamma }=\{\widehat{\alpha }_0\times \widehat{\beta }_0,\mathrm{},\widehat{\alpha }_l\times \widehat{\beta }_l\}\widehat{Z}_{\widehat{\sigma }}^{v,w}$$
if and only if $`\widehat{\sigma }^{}\widehat{\sigma }`$ (more precisely, $`l^{}l`$, $`\sigma _k^{}=\sigma _k`$ for $`1kl^{}`$) and $`\widehat{\alpha }_k^{}\widehat{\alpha }_k`$, $`\widehat{\beta }_k^{}\widehat{\beta }_k`$ for $`1kl^{}`$.
In completely symmetric way, we can define
$$\widehat{Z}^{w,v}=\underset{\widehat{\sigma }\widehat{\mathrm{\Sigma }}}{}\widehat{Z}_{\widehat{\sigma }}^{w,v}$$
as a polyhedron decomposition of $`\mathrm{\Delta }^w`$. Since
$$\widehat{Z}_\sigma ^{v,w}\widehat{Z}_\sigma ^w\times \widehat{Z}_\sigma ^{}^v\widehat{Z}_\sigma ^{}^{w,v}$$
and $`\sigma \sigma ^{}`$ naturally identify $`\mathrm{\Sigma }`$ with $`\mathrm{\Sigma }^{}`$, we have
###### Proposition 7.2
$$\widehat{Z}^{w,v}\widehat{Z}^{v,w}$$
and subface is identified with subface.
This identification naturally induce
###### Theorem 7.1
There is a natural piecewise linear homeomorphism
$$\mathrm{\Delta }^w\mathrm{\Delta }^v$$
with respect to polyhedron decompositions $`\widehat{Z}^{w,v}\widehat{Z}^{v,w}`$.
There is another way to see the theorem. for $`\sigma \mathrm{\Sigma }^{}`$, $`\pi _v`$ over the interior of $`\mathrm{\Delta }_\sigma `$ is a fibration with compact fibres. A simplex $`\beta ^{}`$ in a fibre is corresponding to a simplex of complimentary dimension $`\beta `$ in $`Z_\sigma ^{}^v`$. On the other hand, $`Z_\sigma ^w`$ form a symplicial decomposition of interior of $`\mathrm{\Delta }_\sigma `$. For $`\alpha Z_\sigma ^w`$, $`\beta Z_\sigma ^{}^v`$, the closure of $`\pi _v^1(\alpha )\beta ^{}\alpha \times \beta ^{}`$ in $`\mathrm{\Delta }^v`$ forms a polyhedron $`\mathrm{\Delta }_{\alpha ,\beta }^v`$ in $`\mathrm{\Delta }^v`$. Let $`Z_\sigma ^{v,w}Z_\sigma ^w\times Z_\sigma ^{}^v`$ represent the set of this kind of polyhedron in $`\mathrm{\Delta }^v`$. Then
$$Z^{v,w}=\underset{\sigma \mathrm{\Sigma }^{}}{}Z_\sigma ^{v,w}$$
forms a polyhedron decomposition of $`\mathrm{\Delta }^v`$.
$$\alpha ^{}\times (\beta ^{})^{}Z_\sigma ^{}^{v,w}Z_\sigma ^{}^w\times Z_{(\sigma ^{})^{}}^v$$
is a subface of
$$\alpha \times \beta Z_\sigma ^{v,w}Z_\sigma ^w\times Z_\sigma ^{}^v$$
if and only if $`\sigma ^{}`$, $`\alpha ^{}`$, $`\beta ^{}`$ is a subface of $`\sigma `$, $`\alpha `$, $`\beta `$.
In completely symmetric way, we can define $`Z_\sigma ^{w,v}Z_\sigma ^v\times Z_\sigma ^{}^w`$ and
$$Z^{w,v}=\underset{\sigma \mathrm{\Sigma }^{}}{}Z_\sigma ^{w,v}$$
forms a polyhedron decomposition of $`\mathrm{\Delta }^w`$.
The natural identification of $`Z^{v,w}`$ and $`Z^{w,v}`$ is a duality relation (inclusion relation is reversed). Then in the spirit of proposition 7.1, this dual relation induce a idnetification $`\mathrm{\Delta }^w\mathrm{\Delta }^v`$ that is piecewise linear with respect to the baricenter subdivisons $`\widehat{Z}^{v,w}`$ and $`\widehat{Z}^{w,v}`$.
## 8 SYZ mirror construction for Calabi-Yau hypersurface
To fully establish SYZ construction, we also need to establish the duality relation of the Lagrangian torus fibres in the Lagrangian fibrations of Calabi-Yau hypersurface and its mirror. There are many way to see this, for example, one may compute monodromy operators of the two fibrations, and show that they are dual to each other. We will use a more direct method, we will establish a canonical characterization of Lagrangian torus fibre of the Lagrangian fibrations of Calabi-Yau hypersurface in $`P_{\mathrm{\Sigma }_v}`$ and its mirror in $`P_{\mathrm{\Sigma }_w}`$, and prove that under this canonical characterization the two fibres related via the base identification $`\varphi `$ constructed in the previous section are canonically dual to each other.
Recall the Kähler moduli (the movable cone) of Calabi-Yau hypersurface $`YP_{\mathrm{\Sigma }^w}`$ can be characterized as
$$\tau =\underset{Z\stackrel{~}{Z}}{}\tau _Z$$
where $`\stackrel{~}{Z}`$ denotes the set of simplicial decomposition of $`\mathrm{\Delta }^0`$. For a particular simplicial decomposition $`Z\stackrel{~}{Z}`$, $`\tau _Z`$ is the set of $`w=(w_m)_{m\mathrm{\Delta }^0}`$ (understood as a piecewise linear function on $`M`$) that is convex with respect to $`\mathrm{\Delta }^0`$ in the sense of definition 6.2, and determine the simplicial decomposition $`Z`$ of $`\mathrm{\Delta }^0`$, modulo linear function on $`M`$.
The monomial divisor map gives us a natural identification of the complex moduli of Calabi-Yau hypersurface $`XP_{\mathrm{\Sigma }^v}`$ near the large complex limit with the complexified Kähler moduli of Calabi-Yau hypersurface $`YP_{\mathrm{\Sigma }^w}`$, $`\tau _𝐂=(\stackrel{~}{N}_𝐙𝐑+i\tau )/\stackrel{~}{N}`$.
For any $`u=(u_m)_{m\mathrm{\Delta }^0}(\stackrel{~}{N}_𝐙𝐑+i\tau )/\stackrel{~}{N}`$, consider the Calabi-Yau hypersurface $`X_uP_{\mathrm{\Sigma }^v}`$ defined by
$$s_u(z)=\underset{m\mathrm{\Delta }^0}{}e^{2\pi iu_m}s_m+s_0=0.$$
Let $`w_m=\mathrm{Im}(u_m)`$, then $`w=(w_m)_{m\mathrm{\Delta }^0}\tau `$.Lagrangian torus fibration $`X_u\mathrm{\Delta }_v`$ is constructed by deforming the natural Lagrangian torus fibration of the large complex limit $`X_{\mathrm{}}`$ via gradient flow. Where $`X_{\mathrm{}}`$ is defined by $`s_0=0`$.
For any integral $`n\mathrm{\Delta }^{}N`$, there is a corresponding dimension 3 face $`\alpha _n`$ of $`\mathrm{\Delta }_v`$ defined as
$$\alpha _n=\{m\mathrm{\Delta }_v|v_nm,n=1.\}$$
Namely $`v_nn`$ is the unique supporting function of $`\alpha _n`$. Clearly, fibres of the fibration $`X_{\mathrm{}}\mathrm{\Delta }_v`$ over $`\alpha _n^0`$ (interior of $`\alpha _n`$) are naturally identified with
$$T_n(N_n_𝐙𝐑)/N_n$$
where $`N_n=N/\{𝐙n\}`$. Since Lagrangian torus fibration $`X_u\mathrm{\Delta }_v`$ is a deformation of fibration $`X_{\mathrm{}}\mathrm{\Delta }_v`$, we have
###### Proposition 8.1
3-torus fibres of the Lagrangian fibration $`X_u\mathrm{\Delta }_v`$ over interior of $`\alpha _n`$ can be naturally identified with $`T_n`$.
The dual torus of $`T_n`$ is naturally
$$T_n^{}(M_n_𝐙𝐑)/M_n$$
where $`M_n=n^{}M`$.
Similarly, for any integral $`m\mathrm{\Delta }`$, we can define
$$T_m^{}(N_m_𝐙𝐑)/N_m$$
where $`N_m=m^{}N`$. We have
###### Proposition 8.2
3-torus fibre $`X_b`$ of the Lagrangian fibration $`X_u\mathrm{\Delta }_v`$ over $`b`$ in a small neighborhood $`U_m`$ of $`\widehat{\pi }_v^1(m)\mathrm{\Delta }_v`$ can be naturally identified with $`T_m^{}`$. In addition, if $`b\alpha _n^0`$, then the following diagram commutes
$$\begin{array}{ccc}X_b& & T_m^{}\\ & & \\ T_n& & T_𝐑\end{array}$$
Proof: $`gT`$ naturally acts on $`P_{\mathrm{\Sigma }_v}`$, and also on the space of Calabi-Yau hypersurfaces in $`P_{\mathrm{\Sigma }_v}`$. The action acturally preserve our slice. Under the action of $`g`$, Calabi-Yau hypersurfaces $`X_u`$ defined by $`s_u(z)=0`$ is mapped to $`X_{g(u)}`$ defined by $`s_{g(u)}(z)=s_u(g^1(z))=0`$. Recall that the moment map
$$F_{w,v}(x)=_{m\mathrm{\Delta }_v}\frac{|s_m|_w^2}{|s|_{w,v}^2}m,$$
maps $`P_{\mathrm{\Sigma }_v}`$ to $`\mathrm{\Delta }_v`$. The action of $`g`$ on $`u`$ will naturally induce an action on $`w`$. Clearly, the image $`F_{w,v}(X_u)\mathrm{\Delta }_v`$ is invariant under the action of $`g`$. More precisely, the following diagram commutes.
Think of $`w`$ as a convex piecewise linear function on $`M`$, the action of $`gT`$ is just modifying $`w`$ by linear function. Since $`w`$ is strongly convex, for integral $`m^{}\mathrm{\Delta }`$, by suitable modification of linear function, we may assume that $`u_m^{}=0`$ and $`w_m>0`$ for $`m\mathrm{\Delta }\backslash \{m^{}\}`$. Then
$$s_u(z)=s_0+s_m^{}+\underset{m\mathrm{\Delta }_0\backslash \{m^{}\}}{}e^{2\pi iu_m}s_m.$$
Let $`s_u=s_0+\stackrel{~}{s}_u`$, then
$$\stackrel{~}{s}_u(z)=s_m^{}+\underset{m\mathrm{\Delta }_0\backslash \{m^{}\}}{}e^{2\pi iu_m}s_m.$$
For the purpose of gradient flow, consider meromorphic function
$$q=\frac{s_0(z)}{\stackrel{~}{s}_u(z)}$$
$`f`$ for $`f=\mathrm{Re}(q)`$ is our gradient vector field. We actually use the flow of $`V=\frac{f}{|f|^2}`$.
When $`s_u(z)`$ is near the large complex limit, for $`z`$ near $`F_{w,v}^1\widehat{\pi }_v^1(m^{})`$, we have $`\stackrel{~}{s}_u(z)=s_m^{}+o(1)`$. Hence, $`\stackrel{~}{s}_u(z)`$ is non-vanishing near $`F_{w,v}^1\widehat{\pi }_v^1(m^{})`$, and the flow of $`V`$ is moving Lagrangian 3-torus fibre near $`F_{w,v}^1\widehat{\pi }_v^1(m^{})`$ entirely away from $`X_{\mathrm{}}`$ into the large torus $`T`$. We need to show that this 3-torus in $`T`$ is of the same homotopy class as $`T_m^{}^{}T`$.
In general, a face $`\alpha _n`$ of $`\mathrm{\Delta }_v`$ intersects $`U_m^{}`$ is and only if $`m^{},n=1`$. Under this condition, the composition of
(8.1)
$$T_m^{}^{}T_𝐑T_n$$
is an isomorphism. This condition also ensure that we can choose $`T`$-invariant coordinate $`x=(x_1,x_2,x_3,x_4)`$ near $`F_{w,v}^1(\alpha _n)`$ such that $`F_{w,v}^1(\alpha _n)`$ is defined by $`x_4=0`$ and
$$q=\frac{s_0(z)}{\stackrel{~}{s}_u(z)}=\frac{s_0(z)}{s_m^{}(z)}(1+o(1))=x_4(1+o(1)).$$
Namely, $`x_4`$ is the $`T`$-invariant function on $`T`$ corresponding to $`m^{}M`$. Let $`x_k=r_ke^{i\theta _k}`$ for $`i=1,\mathrm{},4`$, then under this coordinate (8.1) can be expressed as
$$T_m^{}=\{(\theta _1,\theta _2,\theta _3,0)\}T_𝐑=\{(\theta _1,\theta _2,\theta _3,\theta _4)\}T_n=\{(\theta _1,\theta _2,\theta _3)\}$$
Start with Lagrangian 3-torus
$$L_0=\{x||x_i|=r_i,\mathrm{for}i=1,2,3,x_4=0\}$$
The flow of $`f`$ in $`ϵ`$ time will deform $`L_0`$ to approximately
$$L_ϵ=\{x||x_i|=r_i,\mathrm{for}i=1,2,3,x_4=ϵ\}$$
This 3-torus is clearly naturally identified with $`T_m^{}^{}`$. With this explicit description of $`X_b`$, it is easy to see that the diagram in the proposition commute.
$`\mathrm{}`$
Now, let’s consider the mirror Calabi-Yau $`YP_{\mathrm{\Sigma }^w}`$. Also by gradient flow method, we can construct Lagrangian fibration $`Y\mathrm{\Delta }_w`$. For any $`m\mathrm{\Delta }M`$, there is a corresponding dimension 3 face $`\alpha _m`$ of $`\mathrm{\Delta }_w`$ defined as
$$\alpha _m=\{n\mathrm{\Delta }_w|w_mm,n=1.\}$$
Namely $`w_mm`$ is the unique supporting function of $`\alpha _m`$. Clearly, fibres of the fibration $`Y_{\mathrm{}}\mathrm{\Delta }_w`$ over $`\alpha _m^0`$ (interior of $`\alpha _m`$) are naturally identified with
$$T_m(M_m_𝐙𝐑)/M_m$$
where $`M_m=M/\{𝐙m\}`$. Since Lagrangian torus fibration $`Y\mathrm{\Delta }_w`$ is a deformation of fibration $`Y_{\mathrm{}}\mathrm{\Delta }_w`$, we have
###### Proposition 8.3
3-torus fibres of the Lagrangian fibration $`Y\mathrm{\Delta }_w`$ over interior of $`\alpha _m`$ can be naturally identified with $`T_m`$.
Recall the natural map $`\widehat{\pi }_w:\mathrm{\Delta }_w\mathrm{\Delta }^{}`$. For any integral point $`n`$ in $`\mathrm{\Delta }^{}`$, we have
###### Proposition 8.4
3-torus fibre $`Y_b`$ of the Lagrangian fibration $`Y\mathrm{\Delta }_w`$ over $`b`$ in a small neighborhood $`U_n`$ of $`\widehat{\pi }_w^1(n)\mathrm{\Delta }_w`$ can be naturally identified with $`T_n^{}`$. In addition, if $`b\alpha _m^0`$, then the following diagram commutes
$$\begin{array}{ccc}Y_b& & T_n^{}\\ & & \\ T_m& & T_𝐑\end{array}$$
Proof: This is just the mirror statement of proposition 8.2.
$`\mathrm{}`$
Recall from the theorem 7.1, we have the natural piecewise linear identification of the two base of the Lagrangian fibrations
$$\varphi :\mathrm{\Delta }_w\mathrm{\Delta }_v$$
Let $`\alpha _n^0`$, $`\alpha _m^0`$ denote the interior of $`\alpha _n`$, $`\alpha _m`$. Then we have
###### Proposition 8.5
$`U_m`$, $`U_n`$ can be suitablly chosen such that
$$\varphi (U_n)=\alpha _n^0,\varphi (\alpha _m^0)=U_m.$$
And
$$\mathrm{\Delta }_w\backslash \mathrm{\Gamma }^{}=\left(\underset{m\mathrm{\Delta }^0}{}\alpha _m^0\right)\left(\underset{n\mathrm{\Delta }^{}}{}U_n\right)$$
$$\mathrm{\Delta }_v\backslash \mathrm{\Gamma }=\left(\underset{n\mathrm{\Delta }^{}}{}\alpha _n^0\right)\left(\underset{m\mathrm{\Delta }^0}{}U_m\right)$$
We are now ready to establish the dual relation of the fibres. For any $`b\mathrm{\Delta }_w\backslash \mathrm{\Gamma }^{}`$, let $`Y_b`$ ($`X_{\varphi (b)}`$) denote the fibre of the Lagrangian fibration $`Y\mathrm{\Delta }_w`$ ($`X\mathrm{\Delta }_v`$) over $`b\mathrm{\Delta }_w\backslash \mathrm{\Gamma }^{}`$ ($`\varphi (b)\mathrm{\Delta }_v\backslash \mathrm{\Gamma }`$). Then we have
###### Theorem 8.1
For any $`b\mathrm{\Delta }_w\backslash \mathrm{\Gamma }^{}`$, $`Y_b`$ is naturally dual to $`X_{\varphi (b)}`$.
Proof: Based on above propositions, duality is very easy to establish. Only thing that need to be addressed is that duality defined in two ways according to $`U_m`$ or $`\alpha _n^0`$ for $`bU_m\alpha _n^0`$ coincide. For this purpose, one only need to show that the following diagram commutes.
$$\begin{array}{ccccccc}X_{\varphi (b)}& & T_m^{}& & Y_b& & T_n^{}\\ & & & & & & \\ T_n& & T_𝐑& & T_m& & T_𝐑^{}\end{array}$$
This is proved in proposition 8.2 and 8.4.
$`\mathrm{}`$
With explicit identification of fibres in place, monodromy computation becomes a piece of cake! Consider the path $`\gamma _{nmn^{}m^{}}=\alpha _n^0U_m\alpha _n^{}^0U_m^{}\alpha _n^0`$ on $`\mathrm{\Delta }_v`$, where $`n,n^{}\mathrm{\Delta }^{}`$, $`m,m^{}\mathrm{\Delta }^0`$ satisfying
$$m,n=m,n^{}=m^{},n=m^{},n^{}=1.$$
This condition implies that $`\alpha _n`$ and $`\alpha _n^{}`$ have common face that contains $`m,m^{}`$. Correspondingly we have the diagram
$$\begin{array}{ccc}N_n& & N_m\\ & & \\ N_m^{}& & N_n^{}\end{array}$$
$$xx+m,xnx+m,xn+m^{},x+m,xnn^{}=x+m,xn+m^{}m,xn^{}$$
Compose the four operators and modulo $`n`$, we get
###### Theorem 8.2
The monodromy operator along $`\gamma _{nmn^{}m^{}}`$ is
$$[x][x]+m^{}m,x[n^{}]\mathrm{for}[x]N_n$$
Remark: Now we have find an extremely simple way to compute monodromy. All our monodromy computation in can be much easily performed by this method.
We can similarly compute monodromy along the corresponding path for the mirror Lagrangian torus fibration of the mirror Calabi-Yau hypersurface. Consider the corresponding path $`\varphi (\gamma _{nmn^{}m^{}})=U_n\alpha _m^0U_n^{}\alpha _m^{}^0U_n`$ on $`\mathrm{\Delta }_w`$. We have the corresponding diagram
$$\begin{array}{ccc}M_n& & M_m\\ & & \\ M_m^{}& & M_n^{}\end{array}$$
$$yy+y,n^{}my+y,n^{}m+y+y,n^{}m,nm^{}=y+y,n^{}(mm^{})$$
Compose the four operators and modulo $`n`$, we get
###### Theorem 8.3
The monodromy operator along $`\varphi (\gamma _{nmn^{}m^{}})`$ is
$$yy+y,n^{}(mm^{})\mathrm{for}yM_n.$$
Compare with the monodromy operator along $`\gamma _{nmn^{}m^{}}`$, we get
###### Theorem 8.4
The monodromy operator along $`\varphi (\gamma _{nmn^{}m^{}})`$ is naturally dual to the monodromy operator along $`\gamma _{nmn^{}m^{}}`$.
Remark: Although this duality result can be easily shown by the two monodromy formulas, more fundamentally, it is directly implied by the explicit description of fibres described in the previous propositions.
Summerize our results, we have proved the symplectic topological version of SYZ conjecture for quintic Calabi-Yau.
###### Theorem 8.5
For generic quintic Calabi-Yau $`X`$ near the large complex limit, and its mirror Calabi-Yau $`Y`$ near the large radius limit, there exist corresponding Lagrangian torus fibrations
$$\begin{array}{ccccccc}X_{\varphi (b)}& & X& & Y_b& & Y\\ & & & & & & \\ & & \mathrm{\Delta }_v& & & & \mathrm{\Delta }_w\end{array}$$
with singular locus $`\mathrm{\Gamma }\mathrm{\Delta }_v`$ and $`\mathrm{\Gamma }^{}\mathrm{\Delta }_w`$, where $`\varphi :\mathrm{\Delta }_w\mathrm{\Delta }_v`$ is a natural homeomorphism, $`\varphi (\mathrm{\Gamma }^{})=\mathrm{\Gamma }`$ and $`b\mathrm{\Delta }_w\backslash \mathrm{\Gamma }^{}`$. The corresponding fibres $`X_{\varphi (b)}`$ and $`Y_b`$ are naturally dual to each other.
Our construction also give the dual relation of the singular fibres. Recall the discussion of singular fibres of generic Lagrangian torus fibration in section 7. For generic Lagrangian torus fibration, the singular fibres over smooth part of $`\mathrm{\Gamma }`$ are of type $`I`$, singular fibres over verteces of $`\mathrm{\Gamma }`$ are of type $`II`$ or type $`III`$.
###### Theorem 8.6
For $`b\mathrm{\Gamma }^{}`$, if $`Y_b`$ is a singular fibre of type $`I`$, $`II`$, $`III`$, then $`X_{\varphi (b)}`$ is a singular fibre of type $`I`$, $`III`$, $`II`$.
Remark: The last piece of the SYZ puzzle we have not yet discussed is the construction of a section of the Lagrangian fibration. With the explicit description of Lagrangian fibres in this section, it is not hard to construct the section. Roughly, one can take the identity section on each piece with explicit description. When all the coefficients are real, they almost piece together to form an approximate global section, with error depending on how close the Calabi-Yau is near the Lagrangian complex limit. With a more careful construction, we will get a global section. In the case of general coefficients, similar construction will give us a canonical global section if understood in a suitable sense. We will describe the precise construction of gloal section in , where more generally we will discuss monodromy near the large complex limit of general Calabi-Yau hypersurface in toric variety.
## 9 Construction of mirror manifold via Lagrangian torus fibration
As we mentioned in the introduction, the original SYZ mirror conjecture is rather sketchy in nature, with no mentioning of singular locus, singular fibres and duality of singular fibres, which is essential if one wants, for example, to use SYZ to construct mirror manifold. Our discussion of construction of Lagrangian torus fibration and symplectic topological SYZ of generic Calabi-Yau hypersurface in toric variety explicitly produce for us the 3 types of generic singular fibres (type $`I`$, $`II`$, $`III`$ as described in ) and the way they are dual to each other under the mirror symmetry. This together with the knowledge of singular locus from our construction will enable us to give a more precise formulation of SYZ nirror conjecture. This precise formulation will naturally suggest a way to construct mirror manifold in general starting from a generic Lagrangian torus fibration of a Calabi-Yau manifold.
Precise SYZ mirror conjecture For any Calabi-Yau 3-fold $`X`$, with Calabi-Yau metric $`\omega _g`$ and holomorphic volume form $`\mathrm{\Omega }`$, there exists a special Lagrangian fibration of $`X`$ over $`S^3`$
$$\begin{array}{ccc}T^3& & X\\ & & \\ & & S^3\end{array}$$
with a special Lagrangian section and codimension 2 singular locus $`\mathrm{\Gamma }S^3`$, such that general fibres (over $`S^3\backslash \mathrm{\Gamma }`$) are 3-torus. For generic such fibration, $`\mathrm{\Gamma }`$ is a graph with only 3-valent verteces. Let $`\mathrm{\Gamma }=\mathrm{\Gamma }^1\mathrm{\Gamma }^2\mathrm{\Gamma }^3`$, where $`\mathrm{\Gamma }^1`$ is the smooth part of $`\mathrm{\Gamma }`$, $`\mathrm{\Gamma }^2\mathrm{\Gamma }^3`$ is the set of the verteces of $`\mathrm{\Gamma }`$. For any leg $`\gamma \mathrm{\Gamma }^1`$, the monodromy of $`H_1(X_b)`$ of fibre under suitable basis is
$$T_\gamma =\left(\begin{array}{ccc}1& 1& 0\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)$$
Singular fibre along $`\gamma `$ is of type $`I`$.
Consider a vertex $`P\mathrm{\Gamma }^2\mathrm{\Gamma }^3`$ with legs $`\gamma _1`$, $`\gamma _2`$, $`\gamma _3`$. Correspondingly, we have monodromy operators $`T_1`$, $`T_2`$, $`T_3`$.
For $`P\mathrm{\Gamma }^2`$, under suitable basis we have
$$T_1=\left(\begin{array}{ccc}1& 1& 0\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)T_2=\left(\begin{array}{ccc}1& 0& 1\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)T_3=\left(\begin{array}{ccc}1& 1& 1\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)$$
Singular fibre over $`P`$ is of type $`II`$.
For $`P\mathrm{\Gamma }^3`$, under suitable basis we have
$$T_1=\left(\begin{array}{ccc}1& 0& 0\\ 1& 1& 0\\ 0& 0& 1\end{array}\right)T_2=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 1& 0& 1\end{array}\right)T_3=\left(\begin{array}{ccc}1& 0& 0\\ 1& 1& 0\\ 1& 0& 1\end{array}\right)$$
Singular fibre over $`P`$ is of type $`III`$.
The special Lagrangian fibration for the mirror Calabi-Yau manifold $`Y`$ has the same base $`S^3`$ and singular locus $`\mathrm{\Gamma }S^3`$. For $`bS^3\backslash \mathrm{\Gamma }`$, $`Y_b`$ is the dual torus of $`X_bT^3`$. In another word, the $`T^3`$-fibrations
$$\begin{array}{ccc}T^3& & X\\ & & \\ & & S^3\backslash \mathrm{\Gamma }\end{array}\begin{array}{ccc}T^3& & Y\\ & & \\ & & S^3\backslash \mathrm{\Gamma }\end{array}$$
are dual to each other. In particular the monodromy operator will be dual to each other.
For the fibration of $`Y`$, singular fibres over $`\mathrm{\Gamma }^1`$ should be type $`I`$, singular fibres over $`\mathrm{\Gamma }^2`$ should be type $`III`$, singular fibres over $`\mathrm{\Gamma }^2`$ should be type $`II`$. Namely, dual singular fibre of a type $`I`$ singular fibre is still type $`I`$. Type $`II`$ and $`III`$ singular fibres are dual to each other.
$`\mathrm{}`$
Our work (in and this paper) proved the symplectic topological version of this precise SYZ mirror conjecture for quintic, Calabi-Yau hypersurface and more generally Calabi-Yau complete intersection in toric varieties.
Remark: Although it is widely believed that the special Lagangian fibration is a $`C^{\mathrm{}}`$-map and the singular locus is of codimension 2. It is still possible that the singular locus of the actual special Lagrangian is of codimension 1 (as the Lagrangian fibration we constructed before we modify the singular locus to codimension 2). If that is the case, the special Lagangian fibration will not be a $`C^{\mathrm{}}`$-map. Even if this is the case, we believe the symplectic topological version of this precise SYZ conjecture will still hold true, which is sufficient for us to construct the mirror manifold symplectic topologically.
$`\mathrm{}`$
This precise formulation of SYZ conjecture naturally suggests a way to construct mirror manifold $`Y`$ in general starting from a generic Lagrangian torus fibration of a Calabi-Yau manifold $`X`$. The mirror manifold $`Y`$ can be constructed as the dual Lagrangian torus fibration over $`S^3\backslash \mathrm{\Gamma }`$. Filling singular fibres is a delicate issue. Roughly speaking, dual singular fibre of a type $`I`$ singular fibre is still type $`I`$. Type $`II`$ and $`III`$ singular fibres are dual to each other.
To make the construction of mirror manifold precise, it is necessary to construct explicit local model for type $`II`$ and $`III`$ singular fibres. (Our gradient flow construction naturally provide this kind of local models.) Then one need to fit things together using the horizontal section and horizontal foliation determined by the Lagrangian fibration and the horizontal section. We intend to discuss this construction in greater detail in the future.
Current address:
Wei-Dong Ruan
Department of Mathematics
University of Illinois at Chicago
Chicago, IL 60607-7045
email: ruan@math.uic.edu
|
warning/0007/math0007130.html
|
ar5iv
|
text
|
# Symplectic maps to projective spaces and symplectic invariants
## 1. Introduction
Let $`(X^{2n},\omega )`$ be a compact symplectic manifold. We will throughout this text assume that the cohomology class $`\frac{1}{2\pi }[\omega ]H^2(X,)`$ is integral. This assumption makes it possible to define a complex line bundle $`L`$ over $`X`$ such that $`c_1(L)=\frac{1}{2\pi }[\omega ]`$. We also endow $`X`$ with a compatible almost-complex structure $`J`$, and endow $`L`$ with a Hermitian metric and a Hermitian connection of curvature $`i\omega `$.
The line bundle $`L`$ should be thought of as a symplectic version of an ample line bundle over a complex manifold. Indeed, although the lack of integrability of $`J`$ prevents the existence of holomorphic sections, it was observed by Donaldson in that, for large $`k`$, the line bundles $`L^k`$ admit many approximately holomorphic sections.
Observe that all results actually apply as well to the case where $`\frac{1}{2\pi }[\omega ]`$ is not integral, with the only difference that the choice of the line bundle $`L`$ is less natural : the idea is to perturb $`\omega `$ into a symplectic form $`\omega ^{}`$ whose cohomology class is rational, and then work with a suitable multiple of $`\omega ^{}`$. One chooses an almost-complex structure $`J^{}`$ which simultaneously is compatible with $`\omega ^{}`$ and satisfies the positivity property $`\omega (v,J^{}v)>0`$ for all tangent vectors. All the objects that we construct are then approximately $`J^{}`$-holomorphic, and therefore symplectic with respect to not only $`\omega ^{}`$ but also $`\omega `$.
Donaldson was the first to show in that, among the many approximately holomorphic sections of $`L^k`$ for $`k0`$, there is enough flexibility in order to obtain nice transversality properties ; this makes it possible to imitate various classical topological constructions from complex algebraic geometry in the symplectic category. Let us mention in particular the construction of smooth symplectic submanifolds (, see also and ), symplectic Lefschetz pencils (, see also ), branched covering maps to $`^2`$ (,), Grassmannian embeddings and determinantal submanifolds ().
Intuitively, the main reason why the approximately holomorphic framework is suitable to imitate results from algebraic geometry is that, for large values of $`k`$, the increasing curvature of $`L^k`$ provides access to the geometry of $`X`$ at very small scale ; as one zooms into $`X`$, the geometry becomes closer and closer to a standard complex model, and the lack of integrability of $`J`$ becomes negligible.
In all cases the strategy is more or less the same : the aim is, starting with a sequence of given approximately holomorphic sections of $`L^k`$ for all $`k0`$, to perturb them in order to ensure uniform transversality properties that will guarantee the desired topological properties.
For example, to construct submanifolds, one obtains bounds of the type $`|s_k|_{g_k}>\eta `$ along the zero set of $`s_k`$ for a fixed constant $`\eta >0`$ independent of $`k`$, while approximate holomorphicity implies a bound of the type $`|\overline{}s_k|_{g_k}=O(k^{1/2})`$ everywhere. Here $`g_k=kg`$ is a rescaled metric which dilates everything by a factor of $`k^{1/2}`$ in order to adapt to the decreasing “characteristic scale” imposed by the increasing curvature $`ik\omega `$ of the line bundles $`L^k`$. The desired topological picture, similar to the complex algebraic case, emerges for large $`k`$ as an inequality of the form $`|\overline{}s_k||s_k|`$ becomes satisfied : this can easily be shown to imply that the zero set of $`s_k`$ is smooth, approximately pseudo-holomorphic, and symplectic.
The starting points for the construction, in all cases, are the existence of very localized approximately holomorphic sections of $`L^k`$ concentrated near any given point $`xX`$, and an effective transversality result for approximately holomorphic functions defined over a ball in $`^n`$ with values in $`^r`$ due to Donaldson (see for the case $`r=1`$ and for the general case). These two ingredients imply that a small localized perturbation can be used to ensure uniform transversality over a small ball. Combining this local result with a globalization argument (, see also and ), one obtains transversality everywhere.
The interpretation of the construction of submanifolds as an effective transversality result for sections extends verbatim to the more sophisticated constructions (Lefschetz pencils, branched coverings) : in these cases the transversality properties also concern the covariant derivatives of the sections, and this can be thought of as an effective analogue in the approximately holomorphic category of the standard generalized transversality theorem for jets.
This is especially clear when looking at the arguments in , or : the perturbative argument is now used to obtain uniform transversality of the holomorphic parts of the $`1`$-jets or $`2`$-jets of the sections with respect to certain closed submanifolds in the space of holomorphic jets. Successive perturbations are used to obtain transversality to the various strata describing the possible singular models ; one uses that each stratum is smooth away from lower dimensional strata, and that transversality to these lower dimensional strata is enough to imply transversality to the higher dimensional stratum near its singularities.
An extra step is necessary in the constructions : recall that desired topological properties only hold when the antiholomorphic parts of the derivatives are much smaller than the holomorphic parts. In spite of approximate holomorphicity, this can be a problem when the holomorphic part of the jet becomes singular. Therefore, a small perturbation is needed to kill the antiholomorphic part of the jet near the singularities ; this perturbation is in practice easy to construct. The reader is referred to and for details.
Although no general statement has yet been formulated and proved, it is completely clear that a very general result of uniform transversality for jets holds in the approximately holomorphic category. Therefore, the observed phenomenon for Lefschetz pencils and maps to $`^2`$, namely the fact that near every point $`xX`$ the constructed maps are given in approximately holomorphic coordinates by one of the standard local models for generic holomorphic maps, should hold in all generality, independently of the dimensions of the source and target spaces. This approach will be developed in a forthcoming paper .
In the remainder of this paper we focus on the topological monodromy invariants that can be derived from the various available constructions. In Section 2 we study symplectic Lefschetz pencils and their monodromy, following the results of Donaldson and Seidel . In Section 3 we describe symplectic branched covers of $`^2`$ and their monodromy invariants, following and ; we also discuss the connection with 4-dimensional Lefschetz pencils. In Section 4 we extend this framework to the higher dimensional case, and investigate a new type of monodromy invariants arising from symplectic maps to $`^2`$. We finally show in Section 5 that a dimensional induction process makes it possible to describe a compact symplectic manifold of any dimension by a series of words in braid groups and a word in a symmetric group.
Acknowledgement. The author wishes to thank Ludmil Katzarkov, Paul Seidel and Bob Gompf for stimulating discussions, as well as Simon Donaldson for his interest in this work.
## 2. Symplectic Lefschetz pencils
Let $`(X^{2n},\omega )`$ be a compact symplectic manifold as above, and let $`s_0,s_1`$ be suitably chosen approximately holomorphic sections of $`L^k`$. Then $`X`$ is endowed with a structure of symplectic Lefschetz pencil, which can be described as follows.
For any $`\alpha ^1=\{\mathrm{}\}`$, define $`\mathrm{\Sigma }_\alpha =\{xX,s_0+\alpha s_1=0\}`$. Then the submanifolds $`\mathrm{\Sigma }_\alpha `$ are symplectic hypersurfaces, smooth except for finitely many values of the parameter $`\alpha `$ ; for these parameter values $`\mathrm{\Sigma }_\alpha `$ contains a singular point (a normal crossing when $`dimX=4`$). Moreover, the submanifolds $`\mathrm{\Sigma }_\alpha `$ fill all of $`X`$, and they intersect transversely along a codimension 4 symplectic submanifold $`Z=\{xX,s_0=s_1=0\}`$, called the set of base points of the pencil.
Define the projective map $`f=(s_0:s_1):XZ^1`$, whose level sets are precisely the hypersurfaces $`\mathrm{\Sigma }_\alpha `$. Then $`f`$ is required to be a complex Morse function, i.e. its critical points are isolated and non-degenerate, with local model $`f(z_1,\mathrm{},z_n)=z_1^2+\mathrm{}+z_n^2`$ in approximately holomorphic coordinates.
The following result due to Donaldson holds :
###### Theorem 2.1 (Donaldson ).
For $`k0`$, two suitably chosen approximately holomorphic sections of $`L^k`$ endow $`X`$ with a structure of symplectic Lefschetz pencil, canonical up to isotopy.
This result is proved by obtaining uniform transversality with respect to the strata $`s_0=s_1=0`$ (of complex codimension 2) and $`f=0`$ (of complex codimension $`n`$) in the space of holomorphic 1-jets of sections of $`^2L^k`$, by means of the techniques described in the introduction. A small additional perturbation ensures the compatibility requirement that $`\overline{}f`$ vanishes at the points where $`f=0`$. These properties are sufficient to ensure that the structure is that of a symplectic Lefschetz pencil. For details, the reader is referred to .
The statement that the constructed pencils are canonical up to isotopy for $`k0`$ is to be interpreted as follows. Consider two sequences $`(s_k^0)_{k0}`$ and $`(s_k^1)_{k0}`$ of approximately holomorphic sections of $`^2L^k`$ for increasing values of $`k`$. Assume that they satisfy the three above-described transversality and compatibility properties and hence define symplectic Lefschetz pencils. Then, for large enough $`k`$ (how large exactly depends on the estimates on the given sections), there exists an interpolating family $`(s_k^t)_{t[0,1]}`$ of approximately holomorphic sections, depending continuously on the parameter $`t`$, such that for all values of $`t`$ the sections $`s_k^t`$ satisfy the transversality and compatibility properties. In particular, for large enough $`k`$ the symplectic Lefschetz pencils defined by $`s_k^0`$ and $`s_k^1`$ are isotopic to each other. Moreover, the same result remains true if the almost-complex structures $`J_0`$ and $`J_1`$ with respect to which $`s_k^0`$ and $`s_k^1`$ are approximately holomorphic differ, so the topology of the constructed pencils depends only on the topology of the symplectic manifold $`X`$ (and on $`k`$ of course). However, because isotopy holds only for large values of $`k`$, this is only a weak (asymptotic) uniqueness result.
A convenient way to study the topology of a Lefschetz pencil is to blow up $`X`$ along the submanifold $`Z`$. The resulting symplectic manifold $`\widehat{X}`$ is the total space of a symplectic Lefschetz fibration $`\widehat{f}:\widehat{X}^1`$. Although in the following description we work on the blown up manifold $`\widehat{X}`$, it is actually preferrable to work directly on $`X`$ ; verifying that the discussion applies to $`X`$ itself is a simple task left to the reader.
The fibers of $`\widehat{f}`$ can be identified with the submanifolds $`\mathrm{\Sigma }_\alpha `$, made mutually disjoint by the blow-up process. It is then possible to study the monodromy of the fibration $`\widehat{f}`$ around its singular fibers.
One easily checks that this monodromy consists of symplectic automorphisms of the fiber $`\mathrm{\Sigma }_\alpha `$. Moreover, the exceptional divisor obtained by blowing up the set of base points $`Z`$ is a subfibration of $`\widehat{f}`$, with fiber $`Z`$, which is unaffected by the monodromy. Therefore, it is natural to consider that the monodromy of $`\widehat{f}`$ takes values in the symplectic mapping class group $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)=\pi _0(\{\varphi \mathrm{Symp}(\mathrm{\Sigma },\omega ),\varphi _{|U(Z)}=\mathrm{Id}\})`$, i.e. the set of isotopy classes of symplectomorphisms of the generic fiber $`\mathrm{\Sigma }`$ which coincide with the identity near $`Z`$.
In the four-dimensional case, $`Z`$ consists of a finite number $`n`$ of points, and $`\mathrm{\Sigma }`$ is a compact surface with a certain genus $`g`$ (note that $`\mathrm{\Sigma }`$ is always connected because it satisfies a Lefschetz hyperplane type property) ; $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$ is then the classical mapping class group $`\mathrm{Map}_{g,n}`$ of a genus $`g`$ surface with $`n`$ boundary components.
In fact, the image of the monodromy map is contained in the subgroup of exact symplectomorphisms in $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$ : the connection on $`L^k`$ induces over $`\mathrm{\Sigma }Z`$ a 1-form $`\alpha `$ such that $`d\alpha =\omega `$. This endows $`\mathrm{\Sigma }Z`$ with a structure of exact symplectic manifold. Monodromy transformations are then exact symplectomorphisms in the sense that they preserve not only $`\omega `$ but also the 1-form $`\alpha `$ (see for details).
It is well-known (see e.g. , ) that the singular fibers of a Lefschetz fibration are obtained from the generic fiber by collapsing a vanishing cycle to a point. The vanishing cycle is an embedded closed loop in $`\mathrm{\Sigma }`$ in the four-dimensional case ; more generally, it is an embedded Lagrangian sphere $`S^{n1}\mathrm{\Sigma }`$. Then, the monodromy of $`\widehat{f}`$ around one of its singular fibers consists in a generalized Dehn twist in the positive direction along the vanishing cycle.
The picture is the following :
However, because the normal bundle to the exceptional divisor is not trivial, the monodromy map cannot be defined over all of $`^1`$, and we need to restrict ourselves to the preimage of an affine subset $``$ (the fiber at infinity can be assumed regular). The monodromy around the fiber at infinity of $`\widehat{f}`$ is given by a mapping class group element $`\delta _Z`$ corresponding to a twist around $`Z`$. In the four-dimensional case $`Z`$ consists of $`n`$ points, and $`\delta _Z`$ is the product of positive Dehn twists along $`n`$ loops each encircling one of the base points ; in the higher-dimensional case $`\delta _Z`$ is a positive Dehn twist along the unit sphere bundle in the normal bundle of $`Z`$ in $`\mathrm{\Sigma }`$ (i.e. it restricts to each fiber of the normal bundle as a Dehn twist around the origin).
It follows from the above observations that the monodromy of the Lefschetz fibration $`\widehat{f}`$ with critical levels $`p_1,\mathrm{},p_d`$ is given by a group homomorphism
$$\psi :\pi _1(\{p_1,\mathrm{},p_d\})\mathrm{Map}^\omega (\mathrm{\Sigma }^{2n2},Z)$$
(1)
which maps the geometric generators of $`\pi _1(\{p_1,\mathrm{},p_d\})`$, i.e. loops going around one of the points $`p_i`$, to Dehn twists.
Alternately, choosing a system of generating loops in $`\{p_1,\mathrm{},p_d\}`$, we can express the monodromy by a factorization of $`\delta _Z`$ in the mapping class group :
$$\delta _Z=\underset{i=1}{\overset{d}{}}\tau _{\gamma _i},$$
(2)
where $`\gamma _i`$ is the image in a chosen reference fiber of the vanishing cycle of the singular fiber above $`p_i`$ and $`\tau _{\gamma _i}`$ is the corresponding positive Dehn twist. The identity (2) in $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$ expresses the fact that the monodromy of the fibration around the point at infinity in $`^1`$ decomposes as the product of the elementary monodromies around each of the singular fibers.
The monodromy morphism (1), or equivalently the mapping class group factorization (2), completely characterize the topology of the Lefschetz fibration $`\widehat{X}`$. However, they are not entirely canonical, because two choices have been implicitly made in order to define them.
First, a base point in $`\{p_1,\mathrm{},p_d\}`$ and an identification symplectomorphism between $`\mathrm{\Sigma }`$ and the chosen reference fiber of $`\widehat{f}`$ are needed in order to view the monodromy transformations as elements in the mapping class group of $`\mathrm{\Sigma }`$. The choice of a different identification affects the monodromy morphism $`\psi `$ by conjugation by a certain element $`g\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$. The corresponding operation on the mapping class group factorization (2) is a simultaneous conjugation of all factors : each factor $`\tau _{\gamma _i}`$ is replaced by $`\tau _{g(\gamma _i)}=g^1\tau _{\gamma _i}g`$.
Secondly, a system of generating loops has to be chosen in order to define a factorization of $`\delta _Z`$. Different choices of generatic systems differ by a sequence of Hurwitz operations, i.e. moves in which two consecutive generating loops are exchanged, one of them being conjugated by the other in order to preserve the counterclockwise ordering. On the level of the factorization, this amounts to replacing two consecutive factors $`\tau _1`$ and $`\tau _2`$ by respectively $`\tau _2`$ and $`\tau _2^1\tau _1\tau _2`$ (or, by the reverse operation, $`\tau _1\tau _2\tau _1^1`$ and $`\tau _1`$).
It is quite easy to see that any two factorizations of $`\delta _Z`$ describing the Lefschetz fibration $`\widehat{f}`$ differ by a sequence of these two operations (simultaneous conjugation and Hurwitz moves). Therefore, Donaldson’s uniqueness statement implies that, for large enough values of $`k`$, the mapping class group factorizations associated to the symplectic Lefschetz pencil structures obtained in Theorem 2.1 are, up to simultaneous conjugation and Hurwitz moves, symplectic invariants of the manifold $`(X,\omega )`$.
Conversely, given any factorization of $`\delta _Z`$ in $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$ as a product of positive Dehn twists, it is possible to construct a symplectic Lefschetz fibration with the given monodromy. It follows from a result of Gompf that the total space of such a fibration is always a symplectic manifold. In fact, because the monodromy preserves the symplectic submanifold $`Z\mathrm{\Sigma }`$, it is also possible to reconstruct the blown down manifold $`X`$. More precisely, the following result holds :
###### Theorem 2.2 (Gompf).
Let $`(\mathrm{\Sigma },\omega _\mathrm{\Sigma })`$ be a compact symplectic manifold, and $`Z\mathrm{\Sigma }`$ a codimension $`2`$ symplectic submanifold such that $`[Z]=PD([\omega _\mathrm{\Sigma }])`$. Consider a factorization of $`\delta _Z`$ as a product of positive Dehn twists in $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$. In the case $`dim(\mathrm{\Sigma })=2`$, assume moreover that all the Dehn twists in the factorization are along loops that are not homologically trivial in $`\mathrm{\Sigma }Z`$.
Then the total space $`X`$ of the corresponding Lefschetz pencil carries a symplectic form $`\omega _X`$ such that, given a generic fiber $`\mathrm{\Sigma }_0`$ of the pencil, $`[\omega _X]`$ is Poincaré dual to $`[\mathrm{\Sigma }_0]`$, and $`(\mathrm{\Sigma }_0,\omega _{X|\mathrm{\Sigma }_0})`$ is symplectomorphic to $`(\mathrm{\Sigma },\omega _\mathrm{\Sigma })`$. This symplectic structure on $`X`$ is canonical up to symplectic isotopy.
The strategy of proof is to first construct a symplectic structure in the correct cohomology class on a neighborhood of any fiber of the pencil, which is easily done as $`\mathrm{\Sigma }`$ already carries a symplectic structure and the monodromy lies in the symplectomorphism group. It is then possible to combine these local symplectic structures and obtain a globally defined symplectic form, singular near the base locus $`Z`$. Since the total monodromy is $`\delta _Z`$, the structure of $`X`$ near $`Z`$ is completely standard, and so a non-singular symplectic form on $`X`$ can be recovered (this process can also be viewed as a symplectic blow-down along the exceptional hypersurface $`^1\times Z`$ in the total space of the corresponding Lefschetz fibration). This operation changes the cohomology class of the symplectic form on $`X`$, but one easily checks that the resulting class is a nonzero multiple of the Poincaré dual to a fiber ; scaling the symplectic form by a suitable factor then yields $`\omega _X`$. The proof that this process is canonical up to symplectic isotopy is a direct application of Moser’s stability theorem. The reader is referred to for details.
In conclusion, the study of the monodromy of symplectic Lefschetz pencils makes it possible to define invariants of compact symplectic manifolds, which in principle provide a complete description of the topology. However, the complexity of mapping class groups and the difficulties in computing the invariants in concrete situations greatly decrease their usefulness in practice. This motivates the introduction of other similar topological constructions which may lead to more usable invariants.
## 3. Branched covers of $`^2`$ and invariants of symplectic 4-manifolds
Throughout §3, we assume that $`(X,\omega )`$ is a compact symplectic 4-manifold. In that case, three generic approximately holomorphic sections $`s_0`$, $`s_1`$ and $`s_2`$ of $`L^k`$ never vanish simultaneously, and so they define a projective map $`f=(s_0:s_1:s_2):X^2`$. It was shown in that, if the sections are suitably chosen, this map is a branched covering, whose branch curve $`RX`$ is a smooth connected symplectic submanifold in $`X`$.
There are two possible local models in approximately holomorphic coordinates for the map $`f`$ near the branch curve. The first one, corresponding to a generic point of $`R`$, is the map $`(x,y)(x^2,y)`$ ; locally, both the branch curve $`R`$ and its image by $`f`$ are smooth. The other local model corresponds to the isolated points where $`f`$ does not restrict to $`R`$ as an immersion. The model map is then $`(x,y)(x^3xy,y)`$, and the image of the smooth branch curve $`R:3x^2y=0`$ has equation $`f(R):27z_1^2=4z_2^3`$ and presents a cusp singularity. These two local models are the same as in the complex algebraic setting.
It is easy to see by considering the two model maps that $`R`$ is a smooth approximately holomorphic (and therefore symplectic) curve in $`X`$, and that $`f(R)`$ is an approximately holomorphic symplectic curve in $`^2`$, immersed away from its cusps. After a generic perturbation, we can moreover require that the branch curve $`D=f(R)`$ satisfies a self-transversality property, i.e. that its only singular points besides the cusps are transverse double points (“nodes”). Even though $`D`$ is approximately holomorphic, it is not immediately possible to require that all of its double points correspond to a positive intersection number with respect to the standard orientation of $`^2`$ ; the presence of (necessarily badly transverse) negative double points is a priori possible.
It was also shown in that the branched coverings obtained from sections of $`L^k`$ are, for large values of $`k`$, canonical up to isotopy (this weak uniqueness statement holds in the same sense as that of Theorem 2.1). Therefore, the topology of the branch curve $`D=f(R)`$ can be used to define symplectic invariants, provided that one takes into account the possibility of cancellations or creations of pairs of nodes with opposite orientations in isotopies of branched coverings.
Most of the results cited below were obtained in a joint work with L. Katzarkov .
### 3.1. Quasiholomorphic maps to $`^2`$
In order to study the topology of the singular plane curve $`D`$, it is natural to try to adapt the braid group techniques previously used by Moishezon and Teicher in the algebraic case (see e.g. , , ). However, in order to apply this method it is necessary to ensure that the branch curve satisfies suitable transversality properties with respect to a generic projection map from $`^2`$ to $`^1`$. This leads naturally to the notion of quasiholomorphic covering introduced in , which we now describe carefully.
We slightly rephrase the conditions listed in in such a way that they extend naturally to the higher dimensional case ; the same definitions will be used again in §4. It is important to be aware that these concepts only apply to sequences of objects obtained for increasing values of the degree $`k`$ ; the general strategy is always to work simultaneously with a whole family of sections indexed by the parameter $`k`$, in order to ultimately ensure the desired properties for large values of $`k`$. We start with the following terminology :
###### Definition 3.1.
A sequence of sections $`s_k`$ of complex vector bundles $`E_k`$ over $`X`$ (endowed with Hermitian metrics and connections) is asymptotically holomorphic if there exist constants $`C_j`$ independent of $`k`$ such that $`|^js_k|_{g_k}C_j`$ and $`|^{j1}\overline{}s_k|_{g_k}C_jk^{1/2}`$ for all $`j`$, all norms being evaluated with respect to the rescaled metric $`g_k=kg`$ on $`X`$.
The sections $`s_k`$ are uniformly transverse to $`0`$ if there exists a constant $`\gamma >0`$ such that, at every point $`xX`$ where $`|s_k(x)|\gamma `$, the covariant derivative $`s_k(x)`$ is surjective and has a right inverse of norm less than $`\gamma ^1`$ w.r.t. $`g_k`$ (we then say that $`s_k`$ is $`\gamma `$-transverse to $`0`$).
It is easy to check that, if the sections $`s_k`$ are asymptotically holomorphic and uniformly transverse to $`0`$, then for large $`k`$ their zero sets are smooth approximately holomorphic symplectic submanifolds.
Also observe that, in the case where the rank of the bundle $`E_k`$ is greater than the dimension of $`X`$, the surjectivity condition imposed by transversality is never satisfied ; $`\gamma `$-transversality to $`0`$ then means that the norm of the section is greater than $`\gamma `$ at every point of $`X`$.
###### Definition 3.2.
A sequence of projective maps $`f_k:X^2`$ determined by asymptotically holomorphic sections $`s_k=(s_k^0,s_k^1,s_k^2)`$ of $`^3L^k`$ for $`k0`$ is quasiholomorphic if there exist constants $`C_j`$, $`\gamma `$, $`\delta `$ independent of $`k`$, almost-complex structures $`\stackrel{~}{J}_k`$ on $`X`$, and finite sets $`𝒞_k,𝒯_k,_kX`$ such that the following properties hold (using $`\stackrel{~}{J}_k`$ to define the $`\overline{}`$ operator) :
$`(0)`$ $`|^j(\stackrel{~}{J}_kJ)|_{g_k}C_jk^{1/2}`$ for every $`j0`$ ; $`\stackrel{~}{J}_k=J`$ outside of the $`2\delta `$-neighborhood of $`𝒞_k𝒯_k_k`$ ; $`\stackrel{~}{J}_k`$ is integrable in the $`\delta `$-neighborhood of $`𝒞_k𝒯_k_k`$ ;
$`(1)`$ the section $`s_k`$ of $`^3L^k`$ is $`\gamma `$-transverse to $`0`$ ;
$`(2)`$ $`|f_k(x)|_{g_k}\gamma `$ at every point $`xX`$ ;
$`(3)`$ the $`(2,0)`$-Jacobian $`\mathrm{Jac}(f_k)=^2f_k`$ is $`\gamma `$-transverse to $`0`$ ; in particular it vanishes transversely along a smooth symplectic curve $`R_kX`$ (the branch curve).
$`(3^{})`$ the restriction of $`\overline{}f_k`$ to $`\mathrm{Ker}f_k`$ vanishes at every point of $`R_k`$ ;
$`(4)`$ the quantity $`(f_{k|R_k})`$, which can be seen as a section of a line bundle over $`R_k`$, is $`\gamma `$-transverse to $`0`$ and vanishes at the finite set $`𝒞_k`$ (the cusp points of $`f_k`$) ; in particular $`f_k(R_k)=D_k`$ is an immersed symplectic curve away from the image of $`𝒞_k`$ ;
$`(4^{})`$ $`f_k`$ is $`\stackrel{~}{J}_k`$-holomorphic over the $`\delta `$-neighborhood of $`𝒞_k`$ ;
$`(5)`$ the section $`(s_k^0,s_k^1)`$ of $`^2L^k`$ is $`\gamma `$-transverse to $`0`$ ; as a consequence $`D_k`$ remains away from the point $`(0:0:1)`$ ;
$`(6)`$ let $`\pi :^2\{(0:0:1)\}^1`$ be the map defined by $`\pi (x:y:z)=(x:y)`$, and let $`\varphi _k=\pi f_k`$. Then the quantity $`(\varphi _{k|R_k})`$ is $`\gamma `$-transverse to $`0`$ over $`R_k`$, and it vanishes over the union of $`𝒞_k`$ with the finite set $`𝒯_k`$ (the tangency points of the branch curve $`D_k`$ with respect to the projection $`\pi `$) ;
$`(6^{})`$ $`f_k`$ is $`\stackrel{~}{J}_k`$-holomorphic over the $`\delta `$-neighborhood of $`𝒯_k`$ ;
$`(7)`$ the projection $`f_k:R_kD_k`$ is injective outside of the singular points of $`D_k`$, and the self-intersections of $`D_k`$ are transverse double points. Moreover, all special points of $`D_k`$ (cusps, nodes, tangencies) lie in different fibers of the projection $`\pi `$, and none of them lies in $`\pi ^1(0:1)`$ ;
$`(8)`$ the section $`s_k^0`$ of $`L^k`$ is $`\gamma `$-transverse to $`0`$ ;
$`(8^{})`$ $`R_k`$ intersects the zero set of $`s_k^0`$ at the points of $`_k`$ ; $`f_k`$ is $`\stackrel{~}{J}_k`$-holomorphic over the $`\delta `$-neighborhood of $`_k`$.
###### Remark 3.1.
Definition 3.2 is slightly stronger than the definition given in . Most notably, property $`(8)`$, which ensures that the fiber of $`\pi f_k`$ above $`(0:1)`$ enjoys suitable genericity properties, has been added for our purposes. Similarly, condition $`(6^{})`$ is significantly stronger than in , where it was only required that $`\overline{}f_k`$ vanish at the points of $`𝒯_k`$. These extra conditions only require minor modifications of the arguments, while allowing the inductive construction described in §5 to be largely simplified.
Observe that, because of property $`(0)`$, the notions of asymptotic holomorphicity with respect to $`J`$ or $`\stackrel{~}{J}_k`$ coincide. Moreover, even though $`\stackrel{~}{J}_k`$ is used implicitly thoughout the definition, the choice of $`J`$ or $`\stackrel{~}{J}_k`$ is irrelevant as far as transversality properties are concerned since they differ by $`O(k^{1/2})`$.
Property $`(1)`$ means that $`s_k`$ is everywhere bounded from below by $`\gamma `$ ; this implies that the projective map $`f_k`$ is well-defined, and that $`|^jf_k|_{g_k}=O(1)`$ and $`|^{j1}\overline{}f_k|_{g_k}=O(k^{1/2})`$ for all $`j`$. The second property can be interpreted in terms of transversality to the codimension 4 submanifold in the space of $`1`$-jets given by the equation $`f=0`$. Properties $`(3)`$ and $`(3^{})`$ yield the correct structure near generic points of the branch curve : the transverse vanishing of $`\mathrm{Jac}(f_k)`$ implies that the branching order is $`2`$, and the compatibility property $`(3^{})`$ ensures that $`\overline{}f_k`$ remains much smaller than $`f_k`$ in all directions, which is needed to obtain the correct local model.
Properties $`(4)`$ and $`(4^{})`$ determine the structure of the covering near the cusp points. More precisely, observe that along $`R_k`$ the tangent plane field $`TR_k`$ and the plane field $`\mathrm{Ker}f_k`$ coincide exactly at the cusp points ; condition $`(4)`$ expresses that these two plane fields are transverse to each other (in and this condition was formulated in terms of a more complicated quantity; the two formulations are easily seen to be equivalent). This implies that cusp points are isolated and non-degenerate. The compatibility condition $`(4^{})`$ then ensures that the expected local model indeed holds.
The remaining conditions are used to ensure the compatibility of the branch curve $`D_k=f_k(R_k)`$ with the projection $`\pi `$ to $`^1`$. In particular, the transversality condition $`(6)`$ and the corresponding compatibility condition $`(6^{})`$ imply that the points where the branch curve $`D_k`$ fails to be transverse to the fibers of $`\pi `$ are isolated non-degenerate tangency points. Moreover, property $`(7)`$ states that the curve $`D_k`$ is transverse to itself. This implies that $`D_k`$ is a braided curve in the following sense :
###### Definition 3.3.
A real $`2`$-dimensional singular submanifold $`D^2`$ is a braided curve if it satisfies the following properties : $`(1)`$ the only singular points of $`D`$ are cusps (with positive orientation) and transverse double points (with either orientation) ; $`(2)`$ the point $`(0:0:1)`$ does not belong to $`D`$ ; $`(3)`$ the fibers of the projection $`\pi :(x:y:z)(x:y)`$ are everywhere transverse to $`D`$, except at a finite set of nondegenerate tangency points where a local model for $`D`$ in orientation-preserving coordinates is $`z_2^2=z_1`$ ; $`(4)`$ the cusps, nodes and tangency points are all distinct and lie in different fibers of $`\pi `$.
We will see in §3.2 that these properties are precisely those needed in order to apply the braid monodromy techniques of Moishezon-Teicher to the branch curve $`D_k`$.
The main result of can be formulated as follows :
###### Theorem 3.1 (,).
For $`k0`$, it is possible to find asymptotically holomorphic sections of $`^3L^k`$ such that the corresponding projective maps $`f_k:X^2`$ are quasiholomorphic branched coverings. Moreover, for large $`k`$ these coverings are canonical up to isotopy and up to cancellations of pairs of nodes in the branch curves $`D_k`$.
The uniqueness statement is to be understood in the same weak sense as for Theorem 2.1 : given two sequences of quasiholomorphic branched coverings (possibly for different choices of almost-complex structures on $`X`$), for large $`k`$ it is possible to find an interpolating one-parameter family of quasiholomorphic coverings, the only possible non-trivial phenomenon being the cancellation or creation of pairs of nodes in the branch curve for certain parameter values.
The proof of Theorem 3.1 follows a standard pattern : in order to construct quasiholomorphic coverings, one starts with any sequence of asymptotically holomorphic sections of $`^3L^k`$ and proceeds by successive perturbations in order to obtain all the required properties, starting with uniform transversality. Since transversality is an open condition, it is preserved by the subsequent perturbations.
So the first part of the proof consists in obtaining, by successive perturbation arguments, the transversality properties $`(1)`$, $`(2)`$, $`(3)`$ and $`(4)`$ of Definition 3.2 as in , $`(5)`$ and $`(6)`$ as in , and also $`(8)`$ by a direct application of the result of . The argument is notably more technical in the case of $`(4)`$ and $`(6)`$ because the transversality conditions involve derivatives along the branch curve, but these can actually all be thought as immediate applications of the general transversality principle mentioned in the Introduction.
The second part of the proof, which is comparatively easier, deals with the compatibility conditions. The idea is to ensure these properties by perturbing the sections $`s_k`$ by quantities bounded by $`O(k^{1/2})`$, which clearly affects neither holomorphicity nor transversality properties. One first chooses suitable almost-complex structures $`\stackrel{~}{J}_k`$ differing from $`J`$ by $`O(k^{1/2})`$ and integrable near the finite set $`𝒞_k𝒯_k_k`$. It is then possible to perturb $`f_k`$ near these points in order to obtain conditions $`(4^{})`$, $`(6^{})`$ and $`(8^{})`$, by the same argument as in §4.1 of . Next, a generic small perturbation yields the self-transversality of $`D`$ (property $`(7)`$). Finally, a suitable perturbation yields property $`(3^{})`$ along the branch curve without modifying $`R_k`$ and $`D_k`$ and without affecting the other compatibility properties.
The uniqueness statement is obtained by showing that, provided that $`k`$ is large enough, all the arguments extend verbatim to one-parameter families of sections. Therefore, given two sequences of quasiholomorphic coverings, one starts with a one-parameter family of sections interpolating between them in a trivial way and perturbs it in such a way that the required properties hold for all parameter values (with the exception of $`(7)`$ when a node cancellation occurs). Since this construction can be performed in such a way that the two end points of the one-parameter family are not affected by the perturbation, the isotopy result follows immediately.
The reader is referred to and for more details (incorporating requirement $`(8)`$ in the arguments is a trivial task).
### 3.2. Braid monodromy invariants
We now describe the monodromy invariants that naturally arise from the quasiholomorphic coverings described in the previous section. This is a relatively direct extension to the symplectic framework of the braid group techniques studied by Moishezon and Teicher in the algebraic case (see , , ).
Recall that the braid group on $`d`$ strings is the fundamental group $`B_d=\pi _1(𝒳_d)`$ of the space $`𝒳_d`$ of unordered configurations of $`d`$ distinct points in the plane $`^2`$. A braid can therefore be thought as a motion of $`d`$ points in the plane. An alternate description involves compactly supported orientation-preserving diffeomorphisms of $`^2`$ which globally preserve a set of $`d`$ given points : $`B_d=\pi _0(\mathrm{Diff}_c^+(^2,\{q_1,\mathrm{},q_d\}))`$. The group $`B_d`$ is generated by half-twists, i.e. braids in which two of the $`d`$ points rotate around each other by 180 degrees while the other points are preserved. For more details see .
Consider a braided curve $`D^2`$ (see Definition 3.3) of fixed degree $`d`$, for example the branch curve of a quasiholomorphic covering as given by Theorem 3.1. Projecting to $`^1`$ via the map $`\pi `$ makes $`D`$ a singular branched covering of $`^1`$. The picture is the following :
Let $`p_1,\mathrm{},p_r`$ be the images by $`\pi `$ of the special points of $`D`$ (nodes, cusps and tangencies). Observing that the fibers of $`\pi `$ are complex lines (or equivalently real planes) which generically intersect $`D`$ in $`d`$ points, we easily get that the monodromy of the map $`\pi _{|D}`$ around the fibers above $`p_1,\mathrm{},p_r`$ takes values in the braid group $`B_d`$.
The monodromy around one of the points $`p_1,\mathrm{},p_r`$ is as follows. In the case of a tangency point, a local model for the curve $`D`$ is $`y^2=x`$ (with projection to the $`x`$ factor), so one easily checks that the monodromy is a half-twist exchanging two sheets of $`\pi _{|D}`$. Since all half-twists in $`B_d`$ are conjugate, it is possible to write this monodromy in the form $`Q^1X_1Q`$, where $`QB_d`$ is any braid and $`X_1`$ is a fixed half-twist (aligning the points $`q_1,\mathrm{},q_d`$ in that order along the real axis, $`X_1`$ is the half-twist exchanging the points $`q_1`$ and $`q_2`$ along a straight line segment). In the case of a transverse double point with positive intersection, the local model $`y^2=x^2`$ implies that the monodromy is the square of a half-twist, which can be written in the form $`Q^1X_1^2Q`$. The monodromy around a double point with negative intersection is the mirror image of the previous case, and can therefore be written as $`Q^1X_1^2Q`$. Finally, the monodromy around a cusp (local model $`y^2=x^3`$) is the cube of a half-twist and can be expressed as $`Q^1X_1^3Q`$.
However, in order to describe the monodromy automorphisms as braids, one needs to identify up to compactly supported diffeomorphisms the fibers of $`\pi `$ with a reference plane $`^2`$. This implicitly requires a trivialization of the fibration $`\pi `$, which is not available over all of $`^1`$. Therefore, as in the case of Lefschetz pencils, it is necessary to restrict oneself to the preimage of an affine subset $`^1`$, by removing the fiber above the point at infinity (which may easily be assumed to be regular). So the monodromy map is only defined as a group homomorphism
$$\rho :\pi _1(\{p_1,\mathrm{},p_r\})B_d.$$
(3)
Since the fibration $`\pi `$ defines a line bundle of degree $`1`$ over $`^1`$, the monodromy around the fiber at infinity is given by the full twist $`\mathrm{\Delta }^2`$, i.e. the braid which corresponds to a rotation of all points by 360 degrees ($`\mathrm{\Delta }^2`$ generates the center of $`B_d`$).
Therefore, choosing as in §2 a system of generating loops in $`\{p_1,\mathrm{},p_r\}`$, we can express the monodromy by a factorization of $`\mathrm{\Delta }^2`$ in the braid group :
$$\mathrm{\Delta }^2=\underset{j=1}{\overset{r}{}}Q_j^1X_1^{r_j}Q_j,$$
(4)
where the elements $`Q_jB_d`$ are arbitrary braids and the degrees $`r_j\{1,\pm 2,3\}`$ depend on the types of the special points lying above $`p_j`$.
As in the case of Lefschetz pencils, this braid factorization, which completely characterizes the braided curve $`D`$ up to isotopy, is only well-defined up to two algebraic operations: simultaneous conjugation of all factors by a given braid in $`B_d`$, and Hurwitz moves. As previously, simultaneous conjugation reflects the different possible choices of an identification diffeomorphism between the fiber of $`\pi `$ above the base point and the standard plane $`(,\{q_1,\mathrm{},q_d\})`$, while Hurwitz moves arise from changes in the choice of a generating system of loops in $`\{p_1,\mathrm{},p_r\}`$.
Starting with any braid factorization of the form (4), it is possible to reconstruct a braided curve $`D`$ in a canonical way up to isotopy (see ; similar statements were also obtained by Moishezon, Teicher and Catanese). Moreover, one easily checks that factorizations which differ only by global conjugations and Hurwitz moves lead to isotopic braided curves (each such operation amounts to a diffeomorphism isotopic to the identity, obtained in the case of a Hurwitz move by lifting by $`\pi `$ a diffeomorphism of $`^1`$, and in the case of a global conjugation by a diffeomorphism in each of the fibers of $`\pi `$).
Moreover, it is important to observe that every braided curve $`D`$ can be made symplectic by a suitable isotopy. In fact, it is sufficient to perform a radial contraction in all the fibers of $`\pi `$, which brings the given curve into an arbitrarily small neighborhood of the zero section of $`\pi `$ (the complex line $`\{z=0\}`$ in $`^2`$). The tangent space to $`D`$ is then very close to that of the complex line (and therefore symplectic) everywhere except near the tangency points; verifying that the property also holds near tangencies by means of the local model, one obtains that $`D`$ is symplectic.
We now briefly describe the structure of the fundamental group $`\pi _1(^2D)`$. Consider a generic fiber of $`\pi `$, intersecting $`D`$ in $`d`$ points $`q_1,\mathrm{},q_d`$. Then the inclusion map $`i:\{q_1,\mathrm{},q_d\}^2D`$ induces a surjective homomorphism on fundamental groups. Therefore, a generating system of loops $`\gamma _1,\mathrm{},\gamma _d`$ in $`\{q_1,\mathrm{},q_d\}`$ provides a set of generators for $`\pi _1(^2D)`$ (geometric generators). Because the fiber of $`\pi `$ can be compactified by adding the pole of the projection, an obvious relation is $`\gamma _1\mathrm{}\gamma _d=1`$. Moreover, each special point of the curve $`D`$, or equivalently every term in the braid factorization, determines a relation in $`\pi _1(^2D)`$ in a very explicit way.
Namely, recall that there exists a natural right action of $`B_d`$ on the free group $`F_d=\pi _1(\{q_1,\mathrm{},q_d\})`$, that we shall denote by $``$, and consider a factor $`Q_j^1X_1^{r_j}Q_j`$ in (4). Then, if $`r_j=1`$, the tangency point above $`p_j`$ yields the relation $`\gamma _1Q_j=\gamma _2Q_j`$ (the two elements $`\gamma _1Q_j`$ and $`\gamma _2Q_j`$ correspond to small loops going around the two sheets of $`\pi _{|D}`$ that merge at the tangency point). Similarly, in the case of a node ($`r_j=\pm 2`$), the relation is $`[\gamma _1Q_j,\gamma _2Q_j]=1`$. Finally, in the case of a cusp ($`r_j=3`$), the relation becomes $`(\gamma _1\gamma _2\gamma _1)Q_j=(\gamma _2\gamma _1\gamma _2)Q_j`$. It is a classical result that $`\pi _1(^2D)`$ is exactly the quotient of $`F_d=\gamma _1,\mathrm{},\gamma _d`$ by the above-listed relations.
Given a branched covering map $`f:X^2`$ with branch curve $`D`$, it is easy to see that the topology of $`X`$ is determined by a group homomorphism from $`\pi _1(^2D)`$ to the symmetric group $`S_n`$ of order $`n=\mathrm{deg}f`$. Considering a generic fiber of $`\pi `$ which intersects $`D`$ in $`d`$ points $`q_1,\mathrm{},q_d`$, the restriction of $`f`$ to its preimage $`\mathrm{\Sigma }`$ is a $`n`$-sheeted branched covering map from $`\mathrm{\Sigma }`$ to $``$ with branch points $`q_1,\mathrm{},q_d`$. This covering is naturally described by a monodromy representation
$$\theta :\pi _1(\{q_1,\mathrm{},q_d\})S_n.$$
(5)
Because the branching index is $`2`$ at a generic point of the branch curve of $`f`$, the group homomorphism $`\theta `$ maps geometric generators to transpositions. Also, $`\theta `$ necessarily factors through the surjective homomorphism $`i_{}:\pi _1(\{q_1,\mathrm{},q_d\})\pi _1(^2D)`$, because the covering $`f`$ is defined everywhere, and the resulting map from $`\pi _1(^2D)`$ to $`S_n`$ is exactly what is needed to recover the 4-manifold $`X`$ from the branch curve $`D`$. The properties of $`\theta `$ are summarized in the following definition due to Moishezon :
###### Definition 3.4.
A geometric monodromy representation associated to a braided curve $`D^2`$ is a surjective group homomorphism $`\theta `$ from the free group $`\pi _1(\{q_1,\mathrm{},q_d\})=F_d`$ to the symmetric group $`S_n`$ of order $`n`$, mapping the geometric generators $`\gamma _i`$ (and thus also the $`\gamma _iQ_j`$) to transpositions, and such that
$`\theta (\gamma _1\mathrm{}\gamma _d)=1,`$
$`\theta (\gamma _1Q_j)=\theta (\gamma _2Q_j)`$ if $`r_j=1`$,
$`\theta (\gamma _1Q_j)`$ and $`\theta (\gamma _2Q_j)`$ are distinct and commute if $`r_j=\pm 2`$,
$`\theta (\gamma _1Q_j)`$ and $`\theta (\gamma _2Q_j)`$ do not commute if $`r_j=3`$.
Observe that, when the braid factorization defining $`D`$ is affected by a Hurwitz move, $`\theta `$ remains unchanged and the compatibility conditions are preserved. On the contrary, when the braid factorization is modified by simultaneously conjugating all factors by a certain braid $`QB_d`$, the system of geometric generators $`\gamma _1,\mathrm{},\gamma _d`$ changes accordingly, and so the geometric monodromy representation $`\theta `$ should be replaced by $`\theta Q_{}`$, where $`Q_{}`$ is the automorphism of $`F_d`$ induced by the braid $`Q`$.
One easily checks that, given a braided curve $`D^2`$ and a compatible monodromy representation $`\theta :F_dS_n`$, it is possible to recover a compact 4-manifold $`X`$ and a branched covering map $`f:X^2`$ in a canonical way. Moreover, as observed above we can assume that the curve $`D`$ is symplectic; in that case, the branched covering map makes it possible to endow $`X`$ with a symplectic structure, canonically up to symplectic isotopy (see , ; a similar result has also been obtained by Catanese).
The above discussion leads naturally to the definition of symplectic invariants arising from the quasiholomorphic coverings constructed in Theorem 3.1. However, things are complicated by the fact that the branch curves of these coverings are only canonical up to cancellations of double points.
On the level of the braid factorization, a pair cancellation amounts to removing two consecutive factors which are the inverse of each other (necessarily one must have degree $`2`$ and the other degree $`2`$); the geometric monodromy representation is not affected. The opposite operation is the creation of a pair of nodes, in which two factors $`(Q^1X_1^2Q).(Q^1X_1^2Q)`$ are added anywhere in the factorization ; it is allowed only if the new factorization remains compatible with the monodromy representation $`\theta `$, i.e. if $`\theta (\gamma _1Q)`$ and $`\theta (\gamma _2Q)`$ are commuting disjoint transpositions.
###### Definition 3.5.
Two braid factorizations (along with the corresponding geometric monodromy representations) are m-equivalent if there exists a sequence of operations which turns one into the other, each operation being either a global conjugation, a Hurwitz move, or a pair cancellation or creation.
In conclusion, we get the following result :
###### Theorem 3.2 ().
The braid factorizations and geometric monodromy representations associated to the quasiholomorphic coverings obtained in Theorem 3.1 are, for $`k0`$, canonical up to m-equivalence, and define symplectic invariants of $`(X^4,\omega )`$.
Conversely, the data consisting of a braid factorization and a geometric monodromy representation, or a m-equivalence class of such data, determines a symplectic 4-manifold in a canonical way up to symplectomorphism.
### 3.3. The braid group and the mapping class group
Let $`f:X^2`$ be a branched covering map, and let $`D^2`$ be its branch curve. It is a simple observation that, if $`D`$ is braided, then the map $`\pi f`$ with values in $`^1`$ obtained by forgetting one of the components of $`f`$ topologically defines a Lefschetz pencil. This pencil is obtained by lifting via the covering $`f`$ the pencil of lines on $`^2`$ defined by $`\pi `$, and its base points are the preimages by $`f`$ of the pole of the projection $`\pi `$.
Moreover, if one starts with the quasiholomorphic coverings given by Theorem 3.1, then the corresponding Lefschetz pencils coincide for $`k0`$ with those obtained by Donaldson in and described in §2.
As a consequence, in the case of a 4-manifold, the invariants described in §3.2 (braid factorization and geometric monodromy representation) completely determine those described in §2 (factorizations in mapping class groups). It is therefore natural to look for a more explicit description of the relation between branched coverings and Lefschetz pencils. This description involves the group of liftable braids, which has been studied in a special case by Birman and Wajnryb in . We recall the following construction from §5 of .
Let $`𝒞_n(q_1,\mathrm{},q_d)`$ be the (finite) set of all surjective group homomorphisms $`F_dS_n`$ which map each of the geometric generators $`\gamma _1,\mathrm{},\gamma _d`$ of $`F_d`$ to a transposition and map their product $`\gamma _1\mathrm{}\gamma _d`$ to the identity element in $`S_n`$. Each element of $`𝒞_n(q_1,\mathrm{},q_d)`$ determines a simple $`n`$-fold covering of $`^1`$ branched at $`q_1,\mathrm{},q_d`$.
Let $`𝒳_d`$ be the space of configurations of $`d`$ distinct points in the plane. The set of all simple $`n`$-fold coverings of $`^1`$ with $`d`$ branch points and such that no branching occurs above the point at infinity can be thought of as a covering $`\stackrel{~}{𝒳}_{d,n}`$ above $`𝒳_d`$, in which the fiber above the configuration $`\{q_1,\mathrm{},q_d\}`$ identifies with $`𝒞_n(q_1,\mathrm{},q_d)`$. Therefore, the braid group $`B_d=\pi _1(𝒳_d)`$ acts on the fiber $`𝒞_n(q_1,\mathrm{},q_d)`$ by deck transformations of the covering $`\stackrel{~}{𝒳}_{d,n}`$. In fact, the action of a braid $`QB_d`$ on $`𝒞_n(q_1,\mathrm{},q_d)`$ is given by $`\theta \theta Q_{}`$, where $`Q_{}\mathrm{Aut}(F_d)`$ is the automorphism induced by $`Q`$ on the fundamental group of $`\{q_1,\mathrm{},q_d\}`$.
Fix a base point $`\{q_1,\mathrm{},q_d\}`$ in $`𝒳_d`$, and consider an element $`\theta `$ of $`𝒞_n(q_1,\mathrm{},q_d)`$ (i.e., a monodromy representation $`\theta :F_dS_n`$). Let $`p_\theta `$ be the corresponding point in $`\stackrel{~}{𝒳}_{d,n}`$.
###### Definition 3.6.
The subgroup $`B_d^0(\theta )`$ of liftable braids is the set of all the loops in $`𝒳_d`$ whose lift at the point $`p_\theta `$ is a closed loop in $`\stackrel{~}{𝒳}_{d,n}`$. Equivalently, $`B_d^0(\theta )`$ is the set of all braids which act on $`F_d=\pi _1(\{q_1,\mathrm{},q_d\})`$ in a manner compatible with the covering structure defined by $`\theta `$.
In other words, $`B_d^0(\theta )`$ is the set of all braids $`Q`$ such that $`\theta Q_{}=\theta `$, i.e. the stabilizer of $`\theta `$ with respect to the action of $`B_d`$ on $`𝒞_n(q_1,\mathrm{},q_d)`$.
There exists a natural bundle $`𝒴_{d,n}`$ over $`\stackrel{~}{𝒳}_{d,n}`$ (the universal curve) whose fiber is a Riemann surface of genus $`g=1n+(d/2)`$ with $`n`$ marked points. Each of these Riemann surfaces naturally carries a structure of branched covering of $`^1`$, and the marked points are the preimages of the point at infinity.
Given an element $`Q`$ of $`B_d^0(\theta )B_d`$, it can be lifted to $`\stackrel{~}{𝒳}_{d,n}`$ as a loop based at the point $`p_\theta `$, and the monodromy of the fibration $`𝒴_{d,n}`$ along this loop defines an element of $`\mathrm{Map}_{g,n}`$ (the mapping class group of a Riemann surface of genus $`g`$ with $`n`$ boundary components), which we call $`\theta _{}(Q)`$. This defines a group homomorphism $`\theta _{}:B_d^0(\theta )\mathrm{Map}_{g,n}`$.
More geometrically, viewing $`Q`$ as a compactly supported diffeomorphism of the plane preserving $`\{q_1,\mathrm{},q_d\}`$, the fact that $`Q`$ belongs to $`B_d^0(\theta )`$ means that it can be lifted via the covering map $`\mathrm{\Sigma }_g^1`$ to a diffeomorphism of $`\mathrm{\Sigma }_g`$ ; the corresponding element in the mapping class group is $`\theta _{}(Q)`$.
It is easy to check that, when the given monodromy representation $`\theta `$ is compatible with a braided curve $`D^2`$, the image of the braid monodromy homomorphism $`\rho :\pi _1(\{p_1,\mathrm{},p_r\})B_d`$ describing $`D`$ is entirely contained in $`B_d^0(\theta )`$ : this is because the geometric monodromy representation $`\theta `$ factors through $`\pi _1(^2D)`$, on which the braids in $`\mathrm{Im}\rho `$ act trivially. Therefore, we can take the image of the braid factorization describing $`D`$ by $`\theta _{}`$ and obtain a factorization in the mapping class group $`\mathrm{Map}_{g,n}`$. One easily checks that $`\theta _{}(\mathrm{\Delta }^2)`$ is, as expected, the twist $`\delta _Z`$ around the $`n`$ marked points.
As observed in , all the factors of degree $`\pm 2`$ or $`3`$ in the braid factorization lie in the kernel of $`\theta _{}`$ ; therefore, the only terms whose contribution to the mapping class group factorization is non-trivial are those arising from the tangency points of the branch curve $`D`$, and each of these is a Dehn twist. More precisely, the image in $`\mathrm{Map}_{g,n}`$ of a half-twist $`QB_d^0(\theta )`$ can be constructed as follows. Call $`\gamma `$ the path joining two of the branch points naturally associated to the half-twist $`Q`$ (i.e. the path along which the twisting occurs). Among the $`n`$ lifts of $`\gamma `$ to $`\mathrm{\Sigma }_g`$, only two hit the branch points of the covering ; these two lifts have common end points, and together they define a loop $`\delta `$ in $`\mathrm{\Sigma }_g`$. Then the element $`\theta _{}(Q)`$ in $`\mathrm{Map}_{g,n}`$ is the positive Dehn twist along the loop $`\delta `$ (see Proposition 4 of ).
In conclusion, the following result holds :
###### Proposition 3.3.
Let $`f:X^2`$ be a branched covering, and assume that its branch curve $`D`$ is braided. Let $`\rho :\pi _1(\{p_1,\mathrm{},p_r\})B_d^0(\theta )`$ and $`\theta :F_dS_n`$ be the corresponding braid monodromy and geometric monodromy representation. Then the monodromy map $`\psi :\pi _1(\{p_1,\mathrm{},p_r\})\mathrm{Map}_{g,n}`$ of the Lefschetz pencil $`\pi f`$ is given by the identity $`\psi =\theta _{}\rho `$.
In particular, for $`k0`$ the symplectic invariants obtained from Theorem 2.1 are obtained in this manner from those given by Theorem 3.2.
###### Remark 3.2.
It is a basic fact that for $`n3`$ the group homomorphism $`\theta _{}:B_d^0(\theta )\mathrm{Map}_{g,n}`$ is surjective, and that for $`n4`$ every Dehn twist is the image by $`\theta _{}`$ of a half-twist. This makes it natural to ask whether every factorization of $`\delta _Z`$ in $`\mathrm{Map}_{g,n}`$ as a product of Dehn twists is the image by $`\theta _{}`$ of a factorization of $`\mathrm{\Delta }^2`$ in $`B_d^0(\theta )`$ compatible with $`\theta `$. This can be reformulated in more geometric terms as the classical problem of determining whether every Lefschetz pencil is topologically a covering of $`^2`$ branched along a curve with node and cusp singularities (a similar question replacing pencils by Lefschetz fibrations and $`^2`$ by ruled surfaces also holds ; presently the answer is only known in the hyperelliptic case, thanks to the results of Fuller, Siebert and Tian).
A natural approach to these problems is to understand the kernel of $`\theta _{}`$. For example, if one can show that this kernel is generated by squares and cubes of half-twists (factors of degree $`2`$ and $`3`$ compatible with $`\theta `$), then the solution naturally follows : given a decomposition of $`\delta _Z`$ as a product of Dehn twists in $`\mathrm{Map}_{g,n}`$, any lift of this word to $`B_d^0(\theta )`$ as a product of half-twists differs from $`\mathrm{\Delta }^2`$ by a product of factors of degree $`2`$ and $`3`$ and their inverses. Adding these factors as needed, one obtains a decomposition of $`\mathrm{\Delta }^2`$ into factors of degrees $`1`$, $`\pm 2`$ and $`\pm 3`$ ; the branch curve constructed in this way may have nodes and cusps with reversed orientation, but it can still be made symplectic.
Even if the kernel of $`\theta _{}`$ is not generated by factors of degree $`2`$ and $`3`$, it remains likely that the result still holds and can be obtained by starting from a suitable lift to $`B_d^0(\theta )`$ of the word in $`\mathrm{Map}_{g,n}`$. A better understanding of the structure of $`\mathrm{Ker}\theta _{}`$ would be extremely useful for this purpose.
## 4. The higher dimensional case
In this section we extend the results of §3 to the case of higher dimensional symplectic manifolds. In §4.1 we prove the existence of quasiholomorphic maps $`X^2`$ given by triples of sections of $`L^k`$ for $`k0`$. The topological invariants arising from these maps are studied in §4.2 and §4.3, and the relation with Lefschetz pencils is described in §4.4.
### 4.1. Quasiholomorphic maps to $`^2`$
Let $`(X^{2n},\omega )`$ be a compact symplectic manifold, endowed with a compatible almost-complex structure $`J`$. Let $`L`$ be the same line bundle as previously (if $`\frac{1}{2\pi }[\omega ]`$ is not integral one works with a perturbed symplectic form as explained in the introduction). Consider three approximately holomorphic sections of $`L^k`$, or equivalently a section of $`^3L^k`$. Then the following result states that exactly the same transversality and compatibility properties can be expected as in the four-dimensional case :
###### Theorem 4.1.
For $`k0`$, it is possible to find asymptotically holomorphic sections of $`^3L^k`$ such that the corresponding $`^2`$ valued projective maps $`f_k`$ are quasiholomorphic (cf. Definition 3.2). Moreover, for large $`k`$ these projective maps are canonical up to isotopy and up to cancellations of pairs of nodes in the critical curves $`D_k`$.
Before sketching a proof of Theorem 4.1, we briefly describe the behavior of quasiholomorphic maps, which will clarify some of the requirements of Definition 3.2.
Condition $`(1)`$ in Definition 3.2 implies that the set $`Z_k`$ of points where the three sections $`s_k^0,s_k^1,s_k^2`$ vanish simultaneously is a smooth codimension 6 symplectic (approximately holomorphic) submanifold. The projective map $`f_k=(s_k^0:s_k^1:s_k^2)`$ with values in $`^2`$ is only defined over the complement of $`Z_k`$. The behavior near the set of base points is similar to what happens for Lefschetz pencils : in suitable local approximately holomorphic coordinates, $`Z_k`$ is given by the equation $`z_1=z_2=z_3=0`$, and $`f_k`$ behaves like the model map $`(z_1,\mathrm{},z_n)(z_1:z_2:z_3)`$. In fact, a map defined everywhere can be obtained by blowing up $`X`$ along the submanifold $`Z_k`$. The behavior near $`Z_k`$ being completely specified by condition $`(1)`$, it is implicit that all the other conditions on $`f_k`$ are only to be imposed outside of a small neighborhood of $`Z_k`$.
The correct statement of condition $`(3)`$ of Definition 3.2 in the case of a manifold of dimension greater than $`4`$ is a bit tricky. Indeed, $`\mathrm{Jac}(f_k)=^2f_k`$ is a priori a section of the vector bundle $`\mathrm{\Lambda }^{2,0}T^{}Xf_k^{}(\mathrm{\Lambda }^{2,0}T^2)`$ of rank $`n(n1)/2`$. However, transversality to $`0`$ in this sense is impossible to obtain, as the expected complex codimension of $`R_k`$ is $`n1`$ instead of $`n(n1)/2`$. Indeed, the section $`\mathrm{Jac}(f_k)`$ takes values in the non-linear subbundle $`\mathrm{Im}(^2)`$, whose fibers are of dimension $`n1`$ at their smooth points (away from the origin). However, transversality to $`0`$ does not have any natural definition in this subbundle, because it is singular along the zero section. The problem is very similar to what happens in the construction of determinantal submanifolds performed in .
In our case, a precise meaning can be given to condition $`(3)`$ by the following observation. Near any point $`xX`$, property $`(2)`$ implies that it is possible to find local approximately holomorphic coordinates on $`X`$ and local complex coordinates on $`^2`$ in which the differential at $`x`$ of the first component of $`f_k`$ is can be written $`f_k^1(x)=\lambda dz_1`$, with $`|\lambda |>\gamma /2`$. This implies that, near $`x`$, the projection of $`^2f_k`$ to its components along $`dz_1dz_2,\mathrm{},dz_1dz_n`$ is a quasi-isometric isomorphism. In other words, the transversality to $`0`$ of $`\mathrm{Jac}(f_k)`$ is to be understood as the transversality to $`0`$ of its orthogonal projection to the linear subbundle of rank $`n1`$ generated by $`dz_1dz_2,\mathrm{},dz_1dz_n`$.
Another equivalent approach is to consider the (non-linear) bundle $`𝒥^1(X,^2)`$ of holomorphic $`1`$-jets of maps from $`X`$ to $`^2`$. Inside this bundle, the $`1`$-jets whose differential is not surjective define a subbundle $`\mathrm{\Sigma }`$ of codimension $`n1`$, smooth away from the stratum $`\{f=0\}`$. Since this last stratum is avoided by the $`1`$-jet of $`f_k`$ (because of condition $`(2)`$), the transversality to $`0`$ of $`\mathrm{Jac}(f_k)`$ can be naturally rephrased in terms of estimated transversality to $`\mathrm{\Sigma }`$ in the bundle of jets (this approach will be developed in ).
With this understood, conditions $`(3)`$ and $`(3^{})`$ imply, as in the four-dimensional case, that the set $`R_k`$ of points where the differential of $`f_k`$ fails to be surjective is a smooth symplectic curve $`R_kX`$, disjoint from $`Z_k`$, and that the differential of $`f_k`$ has rank $`2`$ at every point of $`R_k`$. Also, as before, conditions $`(4)`$ and $`(4^{})`$ imply that $`f_k(R_k)=D_k`$ is a symplectic curve in $`^2`$, immersed outside of the cusp points.
We now describe the proof of Theorem 4.1 ; most of the argument is identical to the 4-dimensional case, and the reader is referred to and for notations and details.
###### Proof of Theorem 4.1.
The strategy of proof is the same as in the 4-dimensional case. One starts with an arbitrary sequence of asymptotically holomorphic sections of $`^3L^k`$ over $`X`$, and perturbs it first to obtain the transversality properties. Provided that $`k`$ is large enough, each transversality property can be obtained over a ball by a small localized perturbation, using the local transversality result of Donaldson (Theorem 12 in ). A globalization argument then makes it possible to combine these local perturbations into a global perturbation that ensures transversality everywhere (Proposition 3 of ). Since transversality properties are open, successive perturbations can be used to obtain all the required properties : once a transversality property is obtained, subsequent perturbations only affect it by at most decreasing the transversality estimate.
Step 1. One first obtains the transversality statements in parts $`(1)`$, $`(5)`$ and $`(8)`$ of Definition 3.2 ; as in the 4-dimensional case, these properties are obtained e.g. simply by applying the main result of . Observe that all required properties now hold near the base locus $`Z_k`$ of $`s_k`$, so we can assume in the rest of the argument that the points of $`X`$ being considered lie away from $`Z_k`$, and therefore that $`f_k`$ is locally well-defined.
One next ensures condition $`(2)`$, for which the argument is an immediate adaptation of that in §2.2 of , the only difference being the larger number of coordinate functions.
Step 2. The next property we want to get is condition $`(3)`$. Here a significant generalization of the argument in §3.1 of is needed. The problem reduces, as usual, to showing that the uniform transversality to $`0`$ of $`\mathrm{Jac}(f_k)`$ can be ensured over a small ball centered at a given point $`xX`$ by a suitable localized perturbation. As in one can assume that $`s_k(x)`$ is of the form $`(s_k^0(x),0,0)`$ and therefore locally trivialize $`^2`$ via the quasi-isometric map $`(x:y:z)(y/x,z/x)`$ ; this reduces the problem to the study of a $`^2`$-valued map $`h_k`$. Because $`|f_k|`$ is bounded from below, we can assume (after a suitable rotation) that $`|h_k^1(x)|`$ is greater than some fixed constant. Also, fixing suitable approximately holomorphic Darboux coordinates $`z_k^1,\mathrm{},z_k^n`$ (using Lemma 3 of , which trivially extends to dimensions larger than 4), we can after a rotation assume that $`h_k^1(x)`$ is of the form $`\lambda dz_k^1`$, where the complex number $`\lambda `$ is bounded from below.
By Lemma 2 of , there exist asymptotically holomorphic sections $`s_{k,x}^{\mathrm{ref}}`$ of $`L^k`$ with exponential decay away from $`x`$. Define the asymptotically holomorphic 2-forms $`\mu _k^j=h_k^1(z_k^js_{k,x}^{\mathrm{ref}}/s_k^0)`$ for $`2jn`$. At $`x`$, the 2-form $`\mu _k^j`$ is proportional to $`dz_k^1dz_k^j`$ ; therefore, over a small neighborhood of $`x`$, the transversality to $`0`$ of $`\mathrm{Jac}(f_k)`$ in the sense explained above is equivalent to the transversality to $`0`$ of the projection of $`\mathrm{Jac}(h_k)`$ onto the subspace generated by $`\mu _k^2,\mathrm{},\mu _k^n`$. In terms of $`1`$-jets, the 2-forms $`\mu _k^j`$ define a local frame in the normal bundle to the stratum of non-regular maps at $`𝒥^1(f_k)`$. Now, express $`\mathrm{Jac}(h_k)`$ in the form $`u_k^2\mu _k^2+\mathrm{}+u_k^n\mu _k^n+\alpha _k`$ over a neighborhood of $`x`$, where $`u_k^2,\mathrm{},u_k^n`$ are complex-valued functions and $`\alpha _k`$ has no component along $`dz_k^1`$. Then, the transversality to $`0`$ of $`\mathrm{Jac}(f_k)`$ is equivalent to that of the $`^{n1}`$-valued function $`u_k=(u_k^2,\mathrm{},u_k^n)`$.
Since the functions $`u_k`$ are asymptotically holomorphic, using suitable Darboux coordinates at $`x`$ we can use Theorem 12 of to obtain, for large enough $`k`$, the existence of constants $`w_k^2,\mathrm{},w_k^n`$ smaller than any given bound $`\delta >0`$ and such that $`(u_k^2w_k^2,\mathrm{},u_k^nw_k^n)`$ is $`\eta `$-transverse to $`0`$ over a small ball centered at $`x`$, where $`\eta =\delta (\mathrm{log}\delta ^1)^p`$ ($`p`$ is a fixed constant). Letting $`\stackrel{~}{s}_k=(s_k^0,s_k^1,s_k^2w_k^jz_k^js_{k,x}^{\mathrm{ref}})`$ and calling $`\stackrel{~}{f}_k`$ and $`\stackrel{~}{h}_k`$ the projective map defined by $`\stackrel{~}{s}_k`$ and the corresponding local $`^2`$-valued map, we get that $`\mathrm{Jac}(\stackrel{~}{h}_k)=\mathrm{Jac}(h_k)w_k^j\mu _k^j`$, and therefore that $`\mathrm{Jac}(\stackrel{~}{f}_k)`$ is transverse to $`0`$ near $`x`$. Since the perturbation of $`s_k`$ has exponential decay away from $`x`$, we can apply the standard globalization argument to obtain property $`(3)`$ everywhere.
Step 3. The next properties that we want to get are $`(4)`$ and $`(6)`$. It is possible to extend the arguments of and to the higher dimensional case ; however this yields a very technical and lengthy argument, so we outline here a more efficient strategy following the ideas of . Thanks to the previously obtained transversality properties $`(1)`$ and $`(5)`$, both $`f_k`$ and $`\varphi _k`$ are well-defined over a neighborhood of $`R_k`$, so the statements of $`(4)`$ and $`(6)`$ are well-defined. Moreover, observe that property $`(6)`$ implies property $`(4)`$, because at any point where $`(f_{k|R_k})`$ vanishes, $`(\varphi _{k|R_k})`$ necessarily vanishes as well, and if it does so transversely then the same is true for $`(f_{k|R_k})`$ as well. So we only focus on $`(6)`$.
This property can be rephrased in terms of transversality to the codimension $`n`$ stratum $`S:\{(\varphi _{|R})=0\}`$ in the bundle $`𝒥^2(X,^2)`$ of holomorphic 2-jets of maps from $`X`$ to $`^2`$. However this stratum is singular, even away from the substratum $`S_{nt}`$ corresponding to the non-transverse vanishing of $`\mathrm{Jac}(f)`$ ; in fact it is reducible and comes as a union $`S_1S_2`$, where $`S_1:\{\mathrm{Jac}(f)=0,(f_{|R})=0\}`$ is the stratum corresponding to non-immersed points of the branch curve, and $`S_2:\{\varphi =0\}`$ is the stratum corresponding to tangency points of the branch curve. Therefore, one first needs to ensure transversality with respect to $`S_0=S_1S_2:\{\varphi =0,(f_{|R})=0\}`$, which is a smooth codimension $`n+1`$ stratum (“vertical cusp points of the branch curve”) away from $`S_{nt}`$.
Step 3a. We first show that a small perturbation can be used to make sure that the quantity $`(\varphi _k,(f_{k|R_k}))`$ remains bounded from below, i.e. that given any point $`xX`$, either $`\varphi _k(x)`$ is larger than a fixed constant, or $`x`$ lies at more than a fixed distance from $`R_k`$, or $`x`$ lies close to a point of $`R_k`$ where $`(f_{k|R_k}))`$ is larger than a fixed constant. Since this transversality property is local and open, we can obtain it by successive small localized perturbations, as for the previous properties.
Fix a point $`xX`$, and assume that $`\varphi _k(x)`$ is small (otherwise no perturbation is needed). By property $`(5)`$, we know that necessarily $`(s_k^0,s_k^1)`$ is bounded away from zero at $`x`$ ; a rotation in the first two coordinates makes it possible to assume that $`s_k^1(x)=0`$ and $`s_k^0`$ is bounded from below near $`x`$. As above, we replace $`f_k`$ by the $`^2`$-valued map $`h_k=(h_k^1,h_k^2)`$, where $`h_k^i=s_k^i/s_k^0`$. By assumption, we get that $`h_k^1(x)`$ is small. This implies in particular that $`\mathrm{Jac}(f_k)`$ is small at $`x`$, and therefore property $`(3)`$ gives a lower bound on its covariant derivative. Moreover, by property $`(2)`$ we also have a lower bound on $`h_k^2(x)`$, which after a suitable rotation can be assumed equal to $`\lambda dz_k^1`$ for some $`\lambda 0`$. So, as above we can express $`^2f_k`$ by looking at its components along $`dz_k^1dz_k^j`$ for $`2jn`$ ; we again define the $`2`$-forms $`\mu _k^j=h_k^2(z_k^js_{k,x}^{\mathrm{ref}}/s_k^0)`$, and the functions $`u_2,\mathrm{},u_n`$ are defined as previously. Define a $`(n,0)`$-form $`\theta `$ over a neighborhood of $`x`$ by $`\theta =u_2\mathrm{}u_nh_k^2`$ : at points of $`R_k`$, the vanishing of $`\theta `$ is equivalent to that of $`h_{k|R_k}^2`$, or equivalently to that of $`f_{k|R_k}`$. So our aim is to show that the quantity $`(h_k^1,\theta )`$, which is a section of a rank $`n+1`$ bundle $`_0`$ near $`x`$, can be made bounded from below by a small perturbation.
For this purpose, we first show the existence of complex-valued polynomials $`(P_j^1,P_j^2)`$ and local sections $`ϵ_j`$ of $`_0`$, $`1jn+1`$, such that :
(a) for any coefficients $`w_j`$, replacing the given sections of $`L^k`$ by $`(s_k^0,s_k^1+w_jP_j^1s_{k,x}^{\mathrm{ref}},s_k^2+w_jP_j^2s_{k,x}^{\mathrm{ref}})`$ affects $`(h_k^1,\theta )`$ by the addition of $`w_jϵ_j+O(w_j^2)`$ ;
(b) the sections $`ϵ_j`$ define a local frame in $`_0`$, and $`ϵ_1\mathrm{}ϵ_{n+1}`$ is bounded from below by a universal constant.
First observe that, by property $`(3)`$, $`u_2\mathrm{}u_n`$ is bounded from below near $`x`$, whereas we may assume that $`\theta =u_2\mathrm{}u_nh_k^2`$ is small (otherwise no perturbation is needed). Therefore, $`h_k^2`$ (which at $`x`$ is colinear to $`dz_k^1`$) lies close to the span of the $`u_j`$. In particular, after a suitable rotation in the $`n1`$ last coordinates on $`X`$, we can assume that $`u_2h_k^2`$ is small at $`x`$. On the other hand, we know that there exists $`j_01`$ such that $`dz_k^{j_0}`$ lies far from the span of the $`u_j(x)`$. We then define $`P_{n+1}^1=z_k^2z_k^{j_0}`$ and $`P_{n+1}^2=0`$. Adding to $`s_k^1`$ a quantity of the form $`wz_k^2z_k^{j_0}s_{k,x}^{\mathrm{ref}}`$ does not affect $`h_k(x)`$, but affects $`u_2(x)`$ by the addition of a non-trivial multiple of $`dz_k^{j_0}`$, and similarly affects $`u_{j_0}(x)`$ by the addition of a non-trivial multiple of $`dz_k^2`$. The other $`u_j(x)`$ are not affected. Therefore, $`\theta (x)`$ changes by an amount of
$$cwdz_k^{j_0}u_3\mathrm{}u_nh_k^2+c^{}wu_2\mathrm{}dz_k^2\mathrm{}u_nh_k^2+O(w^2),$$
where the constants $`c`$ and $`c^{}`$ are bounded from above and below. The first term is bounded from below by construction, while the second term is only present if $`j_02`$ (this requires $`n3`$), and in that case it is small because $`u_2h_k^2`$ is small. Therefore, the local section $`ϵ_{n+1}`$ of $`_0`$ naturally corresponding to such a perturbation is of the form $`(0,ϵ_{n+1}^{})`$ at $`x`$, where $`ϵ_{n+1}^{}`$ is bounded from below.
Next, for $`1jn`$ we define $`P_j^1=z_k^j`$ and $`P_j^2=0`$, and observe that adding $`wz_k^js_{k,x}^{\mathrm{ref}}`$ to $`s_k^1`$ affects $`h_k^1(x)`$ by adding a nontrivial multiple of $`dz_k^j`$. Therefore, the local section of $`_0`$ corresponding to this perturbation is at $`x`$ of the form $`ϵ_j(x)=(c^{\prime \prime }dz_k^j,ϵ_j^{})`$, where $`c^{\prime \prime }`$ is a constant bounded from below.
It follows from this argument that the chosen perturbations $`P_j^1`$ and $`P_j^2`$ for $`1jn+1`$, and the corresponding local sections $`ϵ_j`$ of $`_0`$, satisfy the conditions (a) and (b) expressed above. Observe that, because $`ϵ_j`$ define a local frame at $`x`$ and $`ϵ_1\mathrm{}ϵ_{n+1}`$ is bounded from below at $`x`$, the same properties remain true over a ball of fixed radius around $`x`$.
Now that a local approximately holomorphic frame in $`_0`$ is given, we can write $`(h_k^1,\theta )`$ in the form $`\zeta _jϵ_j`$ for some complex-valued functions $`\zeta _j`$ ; it is easy to check that these functions are asymptotically holomorphic. Therefore, we can again use Theorem 12 of to obtain, if $`k`$ is large enough, the existence of constants $`w_1,\mathrm{},w_{n+1}`$ smaller than any given bound $`\delta >0`$ and such that $`(\zeta _1w_1,\mathrm{},\zeta _{n+1}w_{n+1})`$ is bounded from below by $`\eta =\delta (\mathrm{log}\delta ^1)^p`$ ($`p`$ is a fixed constant) over a small ball centered at $`x`$. Letting $`\stackrel{~}{s}_k=(s_k^0,s_k^1w_jP_j^1s_{k,x}^{\mathrm{ref}},s_k^2w_jP_j^2s_{k,x}^{\mathrm{ref}})`$ and calling $`\stackrel{~}{f}_k`$, $`\stackrel{~}{h}_k`$ and $`\stackrel{~}{\theta }`$ the projective map defined by $`\stackrel{~}{s}_k`$ and the corresponding local maps, we get that $`(\stackrel{~}{h}_k^1,\stackrel{~}{\theta })`$ is by construction bounded from below by $`c_0\eta `$, for a fixed constant $`c_0`$ ; indeed, observe that the non-linear term $`O(w^2)`$ in the perturbation formula does not play any significant role, as it is at most of the order of $`\delta ^2\eta `$. Since the perturbation of $`s_k`$ has exponential decay away from $`x`$, we can apply the standard globalization argument to obtain uniform transversality to the stratum $`S_0𝒥^2(X,^2)`$ everywhere.
Step 3b. We now obtain uniform transversality to the stratum $`S:\{\mathrm{Jac}(f)=0,(\varphi _{|R})=0\}`$. The strategy and notations are the same as above. We again fix a point $`xX`$, and assume that $`x`$ lies close to a point of $`R_k`$ where $`(\varphi _{k|R_k})`$ is small (otherwise, no perturbation is needed). As above, we can assume that $`s_k^0(x)`$ is bounded from below and define a $`^2`$-valued map $`h_k`$. Two cases can occur : either $`h_k^1(x)`$ is bounded away from zero, or it is small and in that case by Step 3a we know that $`(h_{k|R_k}^2)`$ is bounded from below near $`x`$.
We start with the case where $`h_k^1`$ is bounded from below; in other words, we are not dealing with tangency points but only with cusps. In that case, we can use an argument similar to Step 3a, except that the roles of the two components of $`h_k`$ are reversed. Namely, after a rotation we assume that $`h_k^1(x)=\lambda dz_k^1`$ for some nonzero constant $`\lambda `$, and we define components $`u_2,\mathrm{},u_n`$ of $`\mathrm{Jac}(f_k)`$ as previously (using $`h_k^1`$ rather than $`h_k^2`$ to define the $`\mu _k^j`$). Let $`\theta =u_2\mathrm{}u_nh_k^1`$ : along $`R_k`$, the ratio between $`\theta `$ and $`(h_{k|R_k}^1)`$, or equivalently $`(\varphi _{k|R_k})`$, is bounded between two fixed constants, so the transverse vanishing of $`\theta `$ is what we are trying to obtain. More precisely, our aim is to show that the quantity $`(u_2,\mathrm{},u_n,\theta )`$, which is a section of a rank $`n`$ bundle $``$ near $`x`$, can be made uniformly transverse to $`0`$ by a small perturbation.
For this purpose, we first show the existence of complex-valued polynomials $`(P_j^1,P_j^2)`$ and local sections $`ϵ_j`$ of $``$, $`2jn+1`$, such that :
(a) for any coefficients $`w_j`$, replacing the given sections of $`L^k`$ by $`(s_k^0,s_k^1+w_jP_j^1s_{k,x}^{\mathrm{ref}},s_k^2+w_jP_j^2s_{k,x}^{\mathrm{ref}})`$ affects $`(u_2,\mathrm{},u_n,\theta )`$ by the addition of $`w_jϵ_j+O(w_j^2)`$ ;
(b) the sections $`ϵ_j`$ define a local frame in $``$, and $`ϵ_2\mathrm{}ϵ_{n+1}`$ is bounded from below by a universal constant.
By the same argument as in Step 3a, we find after a suitable rotation an index $`j_01`$ such that, letting $`P_{n+1}^1=0`$ and $`P_{n+1}^2=z_k^2z_k^{j_0}`$, the corresponding local section $`ϵ_{n+1}`$ of $``$ is, at $`x`$, of the form $`(0,\mathrm{},0,ϵ_{n+1}^{})`$, with $`ϵ_{n+1}^{}`$ bounded from below by a fixed constant.
Moreover, adding $`wz_k^js_{k,x}^{\mathrm{ref}}`$ to $`s_k^2`$ amounts to adding $`w`$ to $`u_j`$ and does not affect the other $`u_i`$’s, by the argument in Step 2. So, letting $`P_j^1=0`$ and $`P_j^2=z_k^j`$, we get that the corresponding local sections of $``$ are of the form $`ϵ_j=(0,\mathrm{},1,\mathrm{},0,ϵ_j^{})`$, where the coefficient $`1`$ is in $`j`$-th position.
So it is easy to check that both conditions (a) and (b) are satisfied by these perturbations. The rest of the argument is as in Step 3a : expressing $`(u_2,\mathrm{},u_n,\theta )`$ as a linear combination of $`ϵ_2,\mathrm{},ϵ_{n+1}`$, one uses Theorem 12 of to obtain transversality to $`0`$ over a small ball centered at $`x`$.
We now consider the second possibility, namely the case where $`h_k^1(x)`$ is small, which corresponds to tangency points. By property $`(2)`$ we know that $`h_k^2(x)`$ is bounded from below, and we can assume that it is colinear to $`dz_k^1`$. We then define components $`u_2,\mathrm{},u_n`$ of $`\mathrm{Jac}(f_k)`$ as usual (as in Step 3a and unlike the previous case, the $`\mu _k^j`$ are defined using $`h_k^2`$ rather than $`h_k^1`$). Letting $`\theta =u_2\mathrm{}u_nh_k^1`$, we want as before to obtain the transversality to $`0`$ of the quantity $`(u_2,\mathrm{},u_n,\theta )`$, which is a local section of a rank $`n`$ bundle $``$ near $`x`$. For this purpose, as usual we look for polynomials $`P_j^1`$, $`P_j^2`$ and local sections $`ϵ_j`$ satisfying the same properties (a) and (b) as above.
In order to construct $`P_{n+1}^i`$, observe that, by the result of Step 3a, the quantity $`u_2\mathrm{}u_nh_k^2`$ is bounded from below at $`x`$. So, adding to $`s_k^1`$ a small multiple of $`s_k^2`$ does not affect the $`u_j`$’s, but it affects $`\theta `$ non-trivially. However, this perturbation is not localized, so it is not suitable for our purposes (we can’t apply the globalization argument). Instead, let $`P_{n+1}^1`$ be a polynomial of degree $`2`$ in the coordinates $`z_k^j`$ and their complex conjugates, such that $`P_{n+1}^1s_{k,x}^{\mathrm{ref}}`$ coincides with $`s_k^2`$ up to order two at $`x`$. Note that the coefficients of $`P_{n+1}^1`$ are bounded by uniform constants, and that its antiholomorphic part is at most of the order $`O(k^{1/2})`$ (because $`s_k^2`$ and $`s_{k,x}^{\mathrm{ref}}`$ are asymptotically holomorphic); therefore, $`P_{n+1}^1s_{k,x}^{\mathrm{ref}}`$ is an admissible localized asymptotically holomorphic perturbation. Also, define $`P_{n+1}^2=0`$. Then one easily checks that the local section $`ϵ_{n+1}`$ of $``$ corresponding to $`P_{n+1}^1`$ and $`P_{n+1}^2`$ is, at $`x`$, of the form $`(0,\mathrm{},0,ϵ_{n+1}^{})`$, where $`ϵ_{n+1}^{}`$ is bounded from below.
Moreover, let $`P_j^1=z_k^j`$ and $`P_j^2=0`$ : as above, this perturbation affects $`u_j`$ and not the other $`u_i`$’s, and we get that the corresponding local sections of $``$ are of the form $`ϵ_j=(0,\mathrm{},1,\mathrm{},0,ϵ_j^{})`$, where the coefficient $`1`$ is in $`j`$-th position.
Once again, these perturbations satisfy both conditions (a) and (b). Therefore, expressing $`(u_2,\mathrm{},u_n,\theta )`$ as a linear combination of $`ϵ_2,\mathrm{},ϵ_{n+1}`$, Theorem 12 of yields transversality to $`0`$ over a small ball centered at $`x`$ by the usual argument. Now that both possible cases have been handled, we can apply the standard globalization argument to obtain uniform transversality to the stratum $`S𝒥^2(X,^2)`$. This gives properties $`(4)`$ and $`(6)`$ of Definition 3.2.
Step 4. Now that all required transversality properties have been obtained, we perform further perturbations in order to achieve the other conditions in Definition 3.2. These new perturbations are bounded by a fixed multiple of $`k^{1/2}`$, so the transversality properties are not affected. The argument is almost the same as in the case of 4-manifolds (see §4 of and §3.1 of ); the adaptation to the higher-dimensional case is very easy.
One first defines a suitable almost-complex structure $`\stackrel{~}{J}_k`$, by the same argument as in §4.1 of (except that one also considers the points of $`𝒯_k`$ and $`_k`$ besides the cusps). As explained in §4.1 of , a suitable perturbation makes it possible to obtain the local holomorphicity of $`f_k`$ near these points, which yields conditions $`(4^{})`$, $`(6^{})`$ and $`(8^{})`$ ; the argument is the same in all three cases. Next, a generically chosen small perturbation yields the self-transversality of $`D`$ (property $`(7)`$). Finally, as described in §4.2 of , a suitable perturbation yields property $`(3^{})`$ along the branch curve without modifying $`R_k`$ and $`D_k`$ and without affecting the other compatibility properties. This completes the proof of the existence statement in Theorem 4.1.
Uniqueness. The uniqueness statement is obtained by showing that, provided that $`k`$ is large enough, the whole argument extends to the case of families of sections depending continuously on a parameter $`t[0,1]`$. Then, given two sequences of quasiholomorphic maps, one can start with a one-parameter family of sections interpolating between them in a trivial way and perturb it in such a way that the required properties hold for all parameter values (with the exception of $`(7)`$ when a node cancellation occurs). If one moreover checks that the construction can be performed in such a way that the two end points of the one-parameter family are not affected by the perturbation, the isotopy result becomes an immediate corollary. Observe that, in the one-parameter construction, the almost-complex structure is allowed to depend on $`t`$.
Most of the above argument extends to 1-parameter families in a straightforward manner, exactly as in the four-dimensional case ; the key observation is that all the standard building blocks (existence of approximately holomorphic Darboux coordinates $`z_k^j`$ and of localized approximately holomorphic sections $`s_{k,x}^{\mathrm{ref}}`$, local transversality result, globalization principle, …) remain valid in the parametric case, even when the almost-complex structure depends on $`t`$. The only places where the argument differs from the case of 4-manifolds are properties $`(3)`$, $`(4)`$ and $`(6)`$, obtained in Steps 2 and 3 above.
For property $`(3)`$, one easily checks that it is still possible in the parametric case to assume, after composing with suitable rotations depending continuously on the parameter $`t`$, that $`s_k^1(x)=s_k^2(x)=0`$ and that $`h_k^1(x)`$ is bounded from below and directed along $`dz_k^1`$. This makes it possible to define $`\mu _k^j`$ and $`u_k^j`$ as in the non-parametric case, and the parametric version of Theorem 12 of yields a suitable perturbation depending continuously on $`t`$.
The argument of Step 3a also extends to the parametric case, using the following observation. Fix a point $`xX`$, and let $`\rho _k(t)=|\varphi _{k,t}(x)|`$. For all values of $`t`$ such that $`\rho _k(t)`$ is small enough (smaller than a fixed constant $`\alpha >0`$), we can perform the construction as in the non-parametric case, defining $`u_{j,t}`$ and $`\theta _t`$. If $`\rho _k^{}(t)=|\theta _t(x)|`$ is small enough (smaller than $`\alpha `$), then we can apply the same argument as in the non-parametric case to define polynomials $`(P_{j,t}^1,P_{j,t}^2)`$ and local sections $`ϵ_{j,t}`$ of $`_0`$. However the definition of $`P_{n+1}^1`$ needs to be modified as follows. Although it is still possible after a suitable rotation depending continuously on $`t`$ to assume that $`u_2h_k^2(x)`$ is small, the choice of an index $`j_01`$ such that $`dz_k^{j_0}`$ lies far from the span of the $`u_j(x)`$ may depend on $`t`$. Instead, we define $`\nu _{k,t}`$ as a unit vector in $`^{n1}`$ depending continuously on $`t`$ and such that $`_{j=2}^n\nu _{k,t}^jdz_k^j`$ lies far from the span of $`u_j(x)`$, and let $`P_{n+1,t}^1=_{j=2}^n\nu _{k,t}^jz_k^2z_k^j`$. Then the required properties are satisfied, and we can proceed with the argument. So, provided that $`\rho _k(t)`$ and $`\rho _k^{}(t)`$ are both smaller than $`\alpha `$, we can use Theorem 12 of to obtain a localized perturbation $`\tau _{k,t}`$ depending continuously on $`t`$ and such that $`s_{k,t}+\tau _{k,t}`$ satisfies the desired transversality property near $`x`$.
In order to obtain a well-defined perturbation for all values of $`t`$, we introduce a continuous cut-off function $`\beta :_+[0,1]`$ which equals $`1`$ over $`[0,\alpha /2]`$ and vanishes outside of $`[0,\alpha ]`$. Then, we set $`\stackrel{~}{\tau }_{k,t}=\beta (\rho _k(t))\beta (\rho _k^{}(t))\tau _{k,t}`$, which is well-defined for all $`t`$ and depends continuously on $`t`$. Since $`s_{k,t}+\stackrel{~}{\tau }_{k,t}`$ coincides with $`s_{k,t}+\tau _{k,t}`$ when $`\rho _k(t)`$ and $`\rho _k^{}(t)`$ are smaller than $`\alpha /2`$, the required transversality holds for these values of $`t`$ ; moreover, for the other values of $`t`$ we know that the 2-jet of $`s_{k,t}`$ already lies at distance more than $`\alpha /2`$ from the stratum $`S_0`$, and we can safely assume that $`\stackrel{~}{\tau }_{k,t}`$ is much smaller than $`\alpha /2`$, so the perturbation does not affect transversality. Therefore we obtain a well-defined local perturbation for all $`t[0,1]`$, and the one-parameter version of the result of Step 3a follows by the standard globalization argument.
The argument of Step 3b is extended to one-parameter families in the same way : given a point $`xX`$, the same ideas as for Step 3a yield, for all values of the parameter $`t`$ such that the 2-jet of $`s_{k,t}`$ at $`x`$ lies close to the stratum $`S`$, small localized perturbations $`\tau _{k,t}`$ depending continuously on $`t`$ and such that $`s_{k,t}+\tau _{k,t}`$ satisfies the desired property over a small ball centered at $`x`$. As seen above, two different types of formulas for $`\tau _{k,t}`$ arise depending on which component of the stratum $`S`$ is being hit; however, the result of Step 3a implies that, in any interval of parameter values such that the jet of $`s_{k,t}`$ remains close to $`S`$, only one of the two components of $`S`$ has to be considered, so $`\tau _{k,t}`$ indeed depends continuously on $`t`$. The same type of cut-off argument as for Step 3a then makes it possible to extend the definition of $`\tau _{k,t}`$ to all parameter values and complete the proof. ∎
### 4.2. The topology of quasiholomorphic maps
We now describe the topological features of quasiholomorphic maps and the local models which characterize them near the critical points.
###### Proposition 4.2.
Let $`f_k:XZ_k^2`$ be a sequence of quasiholomorphic maps. Then the fibers of $`f_k`$ are codimension $`4`$ symplectic submanifolds, intersecting at the set of base points $`Z_k`$, and smooth away from the critical curve $`R_kX`$. The submanifolds $`R_k`$ and $`Z_k`$ of $`X`$ are smooth and symplectic, and the image $`f_k(R_k)=D_k`$ is a symplectic braided curve in $`^2`$.
Moreover, given any point $`xR_k`$, there exist local approximately holomorphic coordinates on $`X`$ near $`x`$ and on $`^2`$ near $`f_k(x)`$ in which $`f_k`$ is topologically conjugate to one of the two following models :
$`(i)`$ $`(z_1,\mathrm{},z_n)(z_1^2+\mathrm{}+z_{n1}^2,z_n)`$ $`(`$points where $`f_{k|R_k}`$ is an immersion$`)`$ ;
$`(ii)`$ $`(z_1,\mathrm{},z_n)(z_1^3+z_1z_n+z_2^2+\mathrm{}+z_{n1}^2,z_n)`$ $`(`$near the cusp points$`)`$.
###### Proof.
The smoothness and symplecticity properties of the various submanifolds appearing in the statement follow from the observation made by Donaldson in that the zero sets of approximately holomorphic sections satisfying a uniform transversality property are smooth and approximately $`J`$-holomorphic, and therefore symplectic. In particular, the smoothness and symplecticity of the fibers of $`f_k`$ away from $`R_k`$ follow immediately from Definition 3.2 : since $`\mathrm{Jac}(f_k)`$ is bounded from below away from $`R_k`$ (because it satisfies a uniform transversality property), and since the sections $`s_k`$ are asymptotically holomorphic, it is easy to check that the level sets of $`f_k`$ are, away from $`R_k`$, smooth symplectic submanifolds. Symplecticity near the singular points is an immediate consequence of the local models $`(i)`$ and $`(ii)`$ that we will obtain later in the proof.
The corresponding properties of $`Z_k`$ and $`R_k`$ are obtained by the same argument : $`Z_k`$ and $`R_k`$ are the zero sets of asymptotically holomorphic sections, both satisfying a uniform transversality property (by conditions $`(1)`$ and $`(3)`$ of Definition 3.2, respectively), so they are smooth and symplectic.
We now study the local models at critical points of $`f_k`$. We start with the case of a cusp point $`xX`$. By property $`(2)`$ of Definition 3.2, $`f_k`$ has complex rank $`1`$ at $`x`$, so we can find local complex coordinates $`(Z_1,Z_2)`$ on $`^2`$ near $`f_k(x)`$ such that $`\mathrm{Im}f_k(x)`$ is the $`Z_2`$ axis. Pulling back $`Z_2`$ via the map $`f_k`$, we obtain, using property $`(4^{})`$, a $`\stackrel{~}{J}_k`$-holomorphic function whose differential does not vanish near $`x`$ ; therefore, we can find a $`\stackrel{~}{J}_k`$-holomorphic coordinate chart $`(z_1,\mathrm{},z_n)`$ on $`X`$ at $`x`$ such that $`z_n=Z_2f_k`$. In the chosen coordinates, we get $`f_k(z_1,\mathrm{},z_n)=(g(z_1,\mathrm{},z_n),z_n)`$, where $`g`$ is holomorphic and $`g(0)=0`$.
Since $`x`$ is by assumption a cusp point, the tangent direction to $`R_k`$ at $`x`$ lies in the kernel of $`f_k(0)`$, i.e. in the span of the $`n1`$ first coordinate axes ; after a suitable rotation we may assume that $`T_xR_k`$ is the $`z_1`$ axis. Near the origin, $`\mathrm{Jac}(f_k)`$ is characterized by its $`n1`$ components $`(g/z_1,\mathrm{},g/z_{n1})`$, and the critical curve $`R_k`$ is the set of points where these quantities vanish. Therefore, at the origin, $`^2g/z_1^2=^2g/z_1z_2=\mathrm{}=^2g/z_1z_{n1}=0`$. Nevertheless, $`\mathrm{Jac}(f_k)`$ vanishes transversely to $`0`$ at the origin, so the matrix of second derivatives $`M=(^2g/z_iz_j(0))`$, $`2in`$, $`1jn1`$, is non-degenerate (invertible) at the origin. In particular, the first column of $`M`$ (corresponding to $`j=1`$) is non-zero, and therefore $`^2g/z_1z_n(0)`$ is necessarily non-zero ; after a suitable rescaling of the coordinates we may assume that this coefficient is equal to $`1`$. Moreover, the invertibility of $`M`$ implies that the submatrix $`M^{}=(^2g/z_iz_j(0))`$, $`2i,jn1`$ is also invertible, i.e. it represents a non-degenerate quadratic form.
Diagonalizing this quadratic form, we can assume after a suitable linear change of coordinates that the diagonal coefficients of $`M^{}`$ are equal to $`2`$ and the others are zero. Therefore $`g`$ is of the form $`g(z_1,\mathrm{},z_n)=z_1z_n+_{j=2}^{n1}z_j^2+_{j=2}^{n1}\alpha _jz_jz_n+O(z^3)`$. Changing coordinates on $`X`$ to replace $`z_j`$ by $`z_j+\frac{1}{2}\alpha _jz_n`$ for all $`2jn1`$, and on $`^2`$ to replace $`Z_1`$ by $`Z_1+\frac{1}{4}\alpha _j^2Z_2^2`$, we can ensure that $`g(z_1,\mathrm{},z_n)=z_1z_n+_{j=2}^{n1}z_j^2+O(z^3)`$.
Observe that $`R_k`$ is described near the origin by expressing the coordinates $`z_2,\mathrm{},z_n`$ as functions of $`z_1`$. By assumption the expressions of $`z_2,\mathrm{},z_n`$ are all of the form $`O(z_1^2)`$. Substituting into the formula for $`\mathrm{Jac}(f_k)`$, and letting $`g_{ijk}=^3g/z_iz_jz_k(0)`$, we get that local equations of $`R_k`$ near the origin are $`z_j=\frac{3}{2}g_{j11}z_1^2+O(z_1^3)`$ for $`2jn1`$, and $`z_n=3g_{111}z_1^2+O(z_1^3)`$. It follows that $`f_{k|R_k}`$ is locally given in terms of $`z_1`$ by the map $`z_1(2g_{111}z_1^3+O(z_1^4),3g_{111}z_1^2+O(z_1^3))`$. Therefore, the transverse vanishing of $`(f_{k|R_k})`$ at the origin implies that $`g_{111}0`$, so after a suitable rescaling we may assume that the coefficient of $`z_1^3`$ in the power series expansion of $`g`$ is equal to one.
On the other hand, suitable coordinate changes can be used to kill all other degree $`3`$ terms in the expansion of $`g`$ : if $`2in1`$ the coefficient of $`z_iz_jz_k`$ can be made zero by replacing $`z_i`$ by $`z_i+\frac{c}{2}z_jz_k`$ ; similarly for $`z_n^3`$ (replace $`Z_1`$ by $`Z_1+cZ_2^3`$), $`z_1z_n^2`$ and $`z_1^2z_n`$ (replace $`z_1`$ by $`z_1+cz_n^2+c^{}z_1z_n`$). So we get that $`f_k(z_1,\mathrm{},z_n)=(z_1^3+z_1z_n+z_2^2+\mathrm{}+z_{n1}^2+O(z^4),z_n)`$. It is then a standard result of singularity theory that the higher order terms can be absorbed by suitable coordinate changes.
We now turn to the case of where $`x`$ is a point of $`R_k`$ which does not lie close to any of the cusp points. Conditions $`(2)`$ and $`(3^{})`$ imply that the differential of $`f_k`$ at $`x`$ has real rank $`2`$ and that its image lies close to a complex line in the tangent plane to $`^2`$ at $`f_k(x)`$. Therefore, there exist local approximately holomorphic coordinates $`(Z_1,Z_2)`$ on $`^2`$ such that $`\mathrm{Im}f_k(x)`$ is the $`Z_2`$ axis. Moreover, because $`Z_2f_k`$ is an approximately holomorphic function whose derivative at $`x`$ satisfies a uniform lower bound, it remains possible to find local approximately holomorphic coordinates $`z_1,\mathrm{},z_n`$ on $`X`$ such that $`z_n=Z_2f_k`$. As before, we can write $`f_k(z_1,\mathrm{},z_n)=(g(z_1,\mathrm{},z_n),z_n)`$, where $`g`$ is an approximately holomorphic function such that $`g(0)=0`$.
By assumption $`f_k`$ restricts to $`R_k`$ as an immersion at $`x`$, so the projection to the $`z_n`$ axis of $`T_xR_k`$ is non-trivial. In fact, property $`(4)`$ implies that, if $`(f_{k|R_k})`$ is very small at $`x`$, then a cusp point lies nearby ; so we can assume that the $`z_n`$ component of $`T_xR_k`$ is larger than some fixed constant. As a consequence, one can show that $`R_k`$ is locally given by equations of the form $`z_j=h_j(z_n)`$, where the functions $`h_j`$ are approximately holomorphic and have bounded derivatives. Therefore, a suitable change of coordinates on $`X`$ makes it possible to assume that $`R_k`$ is locally given by the equations $`z_1=\mathrm{}=z_{n1}=0`$. Similarly, a suitable approximately holomorphic change of coordinates on $`^2`$ makes it possible to assume that $`f_k(R_k)`$ is locally given by the equation $`Z_1=0`$.
As a consequence, we have that $`g_{|R_k}=0`$ and, since the image of $`f_k`$ at a point of $`R_k`$ coincides with the tangent space to $`f_k(R_k)`$, $`g`$ vanishes at all points of $`R_k`$. In particular this implies that $`^2g/z_jz_n(0)=0`$ for all $`1jn`$. Moreover, property $`(3)`$ implies that $`\mathrm{Jac}(f_k)`$ vanishes transversely at the origin, and therefore that the matrix $`(^2g/z_iz_j(0))`$, $`1i,jn1`$ is invertible, i.e. it represents a non-degenerate quadratic form. This quadratic form can be diagonalized by a suitable change of coordinates ; because the transversality property $`(3)`$ is uniform, the coefficients are bounded between fixed constants. After a suitable rescaling, we can therefore assume that $`^2g/z_iz_j(0)`$ is equal to $`2`$ if $`i=j`$ and $`0`$ otherwise.
In conclusion, we get that $`g(z_1,\mathrm{},z_n)=z_1^2+\mathrm{}+z_{n1}^2+h(z_1,\mathrm{},z_n)`$, where $`h`$ is the sum of a holomorphic function which vanishes up to order $`3`$ at the origin and of a non-holomorphic function which vanishes up to order $`2`$ at the origin and has derivatives bounded by $`O(k^{1/2})`$.
Let $`z`$ be the column vector $`(z_1,\mathrm{},z_{n1})`$, and denote by $`𝐳`$ the vector $`(z_1,\mathrm{},z_n)`$. Using the fact that $`g`$ vanishes up to order $`2`$ along $`R_k`$, we conclude that there exist matrix-valued functions $`\alpha `$, $`\beta `$ and $`\gamma `$ with the following properties :
$`(a)`$ $`g(𝐳)={}_{}{}^{t}z\alpha (𝐳)z+{}_{}{}^{t}\overline{z}\beta (𝐳)z+{}_{}{}^{t}\overline{z}\gamma (𝐳)\overline{z}`$ ; ($`\alpha `$ and $`\gamma `$ are symmetric) ;
$`(b)`$ $`\alpha `$ is approximately holomorphic and has uniformly bounded derivatives ; $`\alpha (0)=I`$ ;
$`(c)`$ $`\beta `$ and $`\gamma `$ and their derivatives are bounded by fixed multiples of $`k^{1/2}`$.
The implicit function theorem then makes it possible to construct a $`C^{\mathrm{}}`$ approximately holomorphic change of coordinates of the form $`z\lambda (𝐳)z+\mu (𝐳)\overline{z}`$ (with $`\lambda (0)`$ orthogonal, $`\lambda `$ approximately holomorphic, $`\mu =O(k^{1/2})`$), such that $`g`$ becomes of the form $`g(𝐳)={}_{}{}^{t}zz+{}_{}{}^{t}\overline{z}\stackrel{~}{\gamma }(𝐳)\overline{z}`$.
Unfortunately, smooth coordinate changes are not sufficient to further simplify this expression; instead, in order to obtain the desired local model one must use as coordinate change an “approximately holomorphic homeomorphism”, which is smooth away from $`R_k`$ but admits only directional derivatives at the points of $`R_k`$. More precisely, starting from $`g={}_{}{}^{t}zz+h`$ and using that $`h/|z|^2`$ is bounded by $`O(k^{1/2})+O(𝐳)`$, we can write
$$g(𝐳)=\underset{j=1}{\overset{n1}{}}\stackrel{~}{z}_j^2,\stackrel{~}{z}_j=z_j\left(1+\frac{\overline{z}_j}{z_j}\frac{h(𝐳)}{|z|^2}\right)^{1/2}.$$
This gives the desired local model and ends the proof. ∎
###### Remark 4.1.
The local model at points of $`R_k`$ only holds topologically (up to an approximately holomorphic homeomorphism), which is not fully satisfactory. However, by replacing $`(3^{})`$ by a stronger condition, it is possible to obtain the same result in smooth approximately holomorphic coordinates. This new condition can be formulated as follows. Away from the cusp points, the complex lines $`(\mathrm{Im}f_k)^{}`$ define a line bundle $`VT_{|D_k}^2`$, everywhere transverse to $`TD_k`$. A neighborhood of the zero section in $`V`$ can be sent via the exponential map of the Fubini-Study metric onto a neighborhood of $`D_k`$ (away from the cusps), in such a way that each fiber $`V_x`$ is mapped holomorphically to a subset $`𝒱_x`$ contained in a complex line in $`^2`$.
Lifting back to a neighborhood of $`R_k`$ in $`X`$, we can define slices $`𝒲_x=f_k^1(𝒱_{f_k(x)})`$ for all $`xR_k`$ lying away from $`𝒞_k`$. It is then possible to identify a neighborhood of $`R_k`$ (away from $`𝒞_k`$) with a neighborhood of the zero section in the vector bundle $`W`$ whose fiber at $`xR_k`$ is $`\mathrm{Ker}f_k(x)`$, in such a way that each fiber $`W_x`$ gets mapped to $`𝒲_x`$. Observe moreover that, since $`W_x`$ is a complex subspace in $`(T_xX,\stackrel{~}{J}_k)`$, $`W`$ is endowed with a natural complex structure induced by $`\stackrel{~}{J}_k`$. It is then possible to ensure that the “exponential map” from $`W_x`$ to $`𝒲_x`$ is approximately $`\stackrel{~}{J}_k`$-holomorphic for every $`x`$, and, using condition $`(4^{})`$, holomorphic when $`x`$ lies at distance less than $`\delta /2`$ from a cusp point.
With this setup understood, and composing on both sides with the exponential maps, $`f_k`$ induces a fiber-preserving map $`\psi _k`$ between the bundles $`W`$ and $`V`$ ; this map is approximately holomorphic everywhere, and holomorphic at distance less than $`\delta /2`$ from $`𝒞_k`$. The condition which we impose as a replacement of $`(3^{})`$ is that $`\psi _k`$ should be fiberwise holomorphic over a neighborhood of the zero section in $`W`$.
The proof of existence of quasiholomorphic maps satisfying this strengthened condition follows a standard argument : trivializing locally $`V`$ and $`W`$ for each value of $`k`$, and given asymptotically holomorphic maps $`\psi _k`$, Lemma 8 of (see also ) implies the existence of a fiberwise holomorphic map $`\stackrel{~}{\psi }_k`$ differing from $`\psi _k`$ by $`O(k^{1/2})`$ over a neighborhood of the zero section. It is moreover easy to check that $`\stackrel{~}{\psi }_k=\psi _k`$ near the cusp points. So, in order to obtained the desired property, we introduce a smooth cut-off function and define a map $`\widehat{\psi }_k`$ which equals $`\stackrel{~}{\psi }_k`$ near the zero section and coincides with $`\psi _k`$ beyond a certain distance. Going back through the exponential maps, we obtain a map $`\widehat{f}_k`$ which differs from $`f_k`$ by $`O(k^{1/2})`$ and coincides with $`f_k`$ outside a small neighborhood of $`R_k`$ and near the cusp points. The corresponding perturbations of the asymptotically holomorphic sections $`s_k\mathrm{\Gamma }(^3L^k)`$ are easy to construct. Moreover, we can always assume that $`\stackrel{~}{\psi }_k`$ and $`\psi _k`$ coincide at order $`1`$ along the zero section, i.e. that $`\widehat{f}_k`$ and $`f_k`$ coincide up to order 1 along the branch curve ; therefore, the branch curve of $`\widehat{f}_k`$ and its image are the same as for $`f_k`$, and so all properties of Definition 3.2 hold for $`\widehat{f}_k`$.
Once this condition is satisfied, getting the correct local model at a point $`xR_k`$ in smooth approximately holomorphic coordinates is an easy task. Namely, we can define, near $`f_k(x)`$, local approximately holomorphic coordinates $`Z_2`$ on $`D_k`$ and $`Z_1`$ on the fibers of $`V`$ ($`Z_1`$ is a complex linear function on each fiber, depending approximately holomorphically on $`Z_2`$). Using the exponential map, we can use $`(Z_1,Z_2)`$ as local coordinates on $`^2`$. Lifting $`Z_2`$ via $`\widehat{f}_k`$ yields a local coordinate $`z_n`$ on $`R_k`$ near $`x`$. Moreover, we can locally define complex linear coordinates $`z_1,\mathrm{},z_{n1}`$ in the fibers of $`W`$, depending approximately holomorphically on $`z_n`$. Using again the exponential map, $`(z_1,\mathrm{},z_n)`$ define local approximately holomorphic coordinates on $`X`$. Then, by construction, local equations are $`z_1=\mathrm{}=z_{n1}=0`$ for $`R_k`$ and $`Z_1=0`$ for $`D_k`$, and $`f_k`$ is given by $`f_k(z_1,\mathrm{},z_n)=(\psi _k(z_1,\mathrm{},z_n),z_n)`$. Moreover, we know that $`\psi _k`$ is, for each value of $`z_n`$, a holomorphic function of $`z_1,\mathrm{},z_{n1}`$, vanishing up to order 2 at the origin. We can then use the argument in the proof of Proposition 4.2 to obtain the expected local model in smooth approximately holomorphic coordinates.
### 4.3. Monodromy invariants of quasiholomorphic maps
We now look at the monodromy invariants naturally arising from quasiholomorphic maps to $`^2`$. Let $`f:XZ^2`$ be one of the maps constructed in Theorem 3.1 for large enough $`k`$. The fibers of $`f`$ are singular along the smooth symplectic curve $`RX`$, whose image in $`^2`$ is a symplectic braided curve. Therefore, we obtain a first interesting invariant by considering the critical curve $`D^2`$.
As in the four-dimensional case, using the projection $`\pi :^2\{(0:0:1)\}^1`$ we can describe the topology of $`D`$ by a braid monodromy map
$$\rho _n:\pi _1(\{p_1,\mathrm{},p_r\})B_d,$$
(6)
where $`p_1,\mathrm{},p_r`$ are the images by $`\pi `$ of the cusps, nodes and tangency points of $`D`$, and $`d=\mathrm{deg}D`$. Alternately, we can also express this monodromy as a braid group factorization
$$\mathrm{\Delta }^2=\underset{j=1}{\overset{r}{}}Q_j^1X_1^{r_j}Q_j.$$
(7)
Like in the four-dimensional case, this braid factorization completely characterizes the curve $`D`$ up to isotopy, but it is only well-defined up to simultaneous conjugation and Hurwitz equivalence.
We now turn to the second part of the problem, namely describing the topology of the map $`f:XZ^2`$ itself. As in the case of Lefschetz pencils, we blow up $`X`$ along $`Z`$ in order to obtain a well-defined map $`\widehat{f}:\widehat{X}^2`$. The fibers of $`\widehat{f}`$ are naturally identified with those of $`f`$, made mutually disjoint by the blow-up process.
Denote by $`\mathrm{\Sigma }^{2n4}`$ the generic fiber, i.e. the fiber above a point of $`^2D`$. The structure of the singular fibers of $`\widehat{f}`$ can be easily understood by looking at the local models obtained in Proposition 4.2. The easiest case is that of the fiber above a smooth point of $`D`$. This fiber intersects $`R`$ transversely in one point, where the local model is $`(z_1,\mathrm{},z_n)(z_1^2+\mathrm{}+z_{n1}^2,z_n)`$, which can be thought of as a one-parameter version of the model map for the singularities of a Lefschetz pencil in dimension $`2n2`$. Therefore, as in that case, the singular fiber is obtained by collapsing a vanishing cycle, namely a Lagrangian sphere $`S^{n2}`$, in the generic fiber $`\mathrm{\Sigma }`$, and the monodromy of $`\widehat{f}`$ maps a small loop around $`D`$ to a positive Dehn twist along the vanishing cycle.
The fiber of $`\widehat{f}`$ above a nodal point of $`D`$ intersects $`R`$ transversely in two points, and is similarly obtained from $`\mathrm{\Sigma }`$ by collapsing two disjoint Lagrangian spheres. In fact, the nodal point does not give rise to any specific local model in $`X`$, as it simply corresponds to the situation where two points of $`R`$ happen to lie in the same fiber.
Finally, in the case of a cusp point of $`D`$, the local model $`(z_1,\mathrm{},z_n)(z_1^3+z_1z_n+z_2^2+\mathrm{}+z_{n1}^2,z_n)`$ can be used to show that the singular fiber is a “fishtail” fiber, obtained by collapsing two Lagrangian spheres which intersect transversely in one point.
With this understood, the topology of $`\widehat{f}`$ is described by its monodromy around the singular fibers. As in the case of Lefschetz fibrations, the monodromy consists of symplectic automorphisms of $`\mathrm{\Sigma }`$ preserving the submanifold $`Z`$. However, as in §2, defining a monodromy map with values in $`\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$ requires a trivialization of the normal bundle of $`Z`$, which is only possible over an affine subset $`^2^2`$. So, the monodromy of $`\widehat{f}`$ is described by a group homomorphism
$$\psi _n:\pi _1(^2D)\mathrm{Map}^\omega (\mathrm{\Sigma },Z).$$
(8)
A simpler description can be obtained by restricting oneself to a generic line $`L^2`$ which intersects $`D`$ transversely in $`d`$ points $`q_1,\mathrm{},q_d`$. In fact, Definition 3.2 implies that we can use the fiber of $`\pi `$ above $`(0:1)`$ for this purpose. As in §3.2, the inclusion $`i:\{q_1,\mathrm{},q_d\}^2D`$ induces a surjective homomorphism on fundamental groups. The relations between the geometric generators $`\gamma _1,\mathrm{},\gamma _d`$ of $`\pi _1(^2D)`$ are again given by the braid factorization (one relation for each factor) in the same manner as in §3.2. Note that the relation $`\gamma _1\mathrm{}\gamma _d=1`$ only holds in $`\pi _1(^2D)`$, not in $`\pi _1(^2D)`$.
It follows from these observations that the monodromy of $`\widehat{f}`$ can be described by the monodromy morphism
$$\theta _{n1}:\pi _1(\{q_1,\mathrm{},q_d\})\mathrm{Map}^\omega (\mathrm{\Sigma },Z)$$
(9)
defined by $`\theta _{n1}=\psi _ni_{}`$. We know from the above discussion on the structure of $`\widehat{f}`$ near its critical points that $`\theta _{n1}`$ maps the geometric generators of $`\pi _1(\{q_1,\mathrm{},q_d\})`$ to positive Dehn twists. Moreover, by considering the normal bundle to the exceptional divisor in $`\widehat{X}`$ one easily checks that the monodromy around infinity is again a twist along $`Z`$ in $`\mathrm{\Sigma }`$, i.e. $`\theta _{n1}(\gamma _1\mathrm{}\gamma _d)=\delta _Z`$.
These properties of $`\theta _{n1}`$ are strikingly similar to those of the monodromy of a symplectic Lefschetz pencil. In fact, let $`W=f^1(L)`$ be the preimage of a complex line $`L=^1^2`$ intersecting $`D`$ transversely. Then the restriction of $`f`$ to the smooth symplectic hypersurface $`WX`$ endows it with a structure of symplectic Lefschetz pencil with generic fiber $`\mathrm{\Sigma }`$ and base set $`Z`$ ; for example, if one chooses $`L=\pi ^1(0:1)`$, then $`W`$ is the zero set of $`s_k^0`$ and the restricted pencil $`f_{|W}:WZ^1`$ is defined by the two sections $`s_k^1`$ and $`s_k^2`$. The monodromy of the restricted pencil is, by construction, given by the map $`\theta _{n1}`$.
The situation is summarized in the following picture :
###### Remark 4.2.
If a cusp point of $`D`$ happens to lie close to the chosen line $`L`$, then two singular points of the restricted pencil $`f_{|W}`$ lie close to each other. This is not a problem here, but in general if we want to avoid this situation we need to impose one additional transversality condition on $`f`$. Namely, we must require the uniform transversality to $`0`$ of $`(f_{|W})`$, which is easily obtained by imitating Donaldson’s argument from . Another situation in which this property naturally becomes satisfied is the one described in §5.
Given a braided curve $`D^2`$ of degree $`d`$ described by a braid factorization as in (7), and given a monodromy map $`\theta _{n1}`$ as in (9), certain compatibility conditions need to hold between them in order to ensure the existence of a $`^2`$-valued map with critical curve $`D`$ and monodromy $`\theta _{n1}`$. Namely, $`\theta _{n1}`$ must factor through $`\pi _1(^2D)`$, and the fibration must behave in accordance with the expected models near the special points of $`D`$. We introduce the following definition summarizing these compatibility properties :
###### Definition 4.1.
A geometric $`(n1)`$-dimensional monodromy representation associated to a braided curve $`D^2`$ is a surjective group homomorphism $`\theta _{n1}`$ from the free group $`\pi _1(\{q_1,\mathrm{},q_d\})=F_d`$ to a symplectic mapping class group $`\mathrm{Map}^\omega (\mathrm{\Sigma }^{2n4},Z^{2n6})`$, mapping the geometric generators $`\gamma _i`$ (and thus also the $`\gamma _iQ_j`$) to positive Dehn twists and such that
$`\theta _{n1}(\gamma _1\mathrm{}\gamma _d)=\delta _Z,`$
$`\theta _{n1}(\gamma _1Q_j)=\theta _{n1}(\gamma _2Q_j)`$ if $`r_j=1`$,
$`\theta _{n1}(\gamma _1Q_j)`$ and $`\theta _{n1}(\gamma _2Q_j)`$ are twists along disjoint Lagrangian spheres if $`r_j=\pm 2`$,
$`\theta _{n1}(\gamma _1Q_j)`$ and $`\theta _{n1}(\gamma _2Q_j)`$ are twists along Lagrangian spheres transversely intersecting in one point if $`r_j=3`$.
As in the four-dimensional case, $`\theta _{n1}`$ remains unchanged and the compatibility conditions are preserved when the braid factorization defining $`D`$ is affected by a Hurwitz move. However, when all factors in the braid factorization are simultaneously conjugated by a certain braid $`QB_d`$, the system of geometric generators $`\gamma _1,\mathrm{},\gamma _d`$ changes accordingly, and so the geometric monodromy representation $`\theta _{n1}`$ should be replaced by $`\theta _{n1}Q_{}`$, where $`Q_{}`$ is the automorphism of $`F_d`$ induced by the braid $`Q`$. For example, conjugating the braid factorization by one of the generating half-twists in $`B_d`$ affects the monodromy $`\theta _{n1}`$ of the restricted pencil by a Hurwitz move.
One easily checks that, given a symplectic braided curve $`D^2`$ and a compatible monodromy representation $`\theta _{n1}:F_d\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$, it is possible to recover a compact $`2n`$-manifold $`X`$ and a map $`f:XZ^2`$ in a canonical way up to smooth isotopy. Moreover, it is actually possible to endow $`X`$ with a symplectic structure, canonically up to symplectic isotopy. Indeed, by first applying Theorem 2.2 to the monodromy map $`\theta _{n1}`$ we can recover a canonical symplectic structure on the total space $`W`$ of the restricted Lefschetz pencil ; furthermore, as will be shown in §4.4 below, the braid monodromy of $`D`$ and the compatible monodromy representation $`\theta _{n1}`$ determine on $`X`$ a structure of Lefschetz pencil with generic fiber $`W`$ and base set $`\mathrm{\Sigma }`$, which implies by a second application of Theorem 2.2 that $`X`$ carries a canonical symplectic structure. The same result can also be obtained more directly, by adapting the statement and proof of Theorem 2.2 to the case of $`^2`$-valued maps.
As in the four-dimensional case, we can naturally define symplectic invariants arising from the quasiholomorphic maps constructed in Theorem 4.1. However, we again need to take into account the possible presence of negative self-intersections in the critical curves of these maps. Therefore, the braid factorizations we obtain are only canonical up to global conjugation, Hurwitz equivalence, and pair cancellations or creations. As in the four-dimensional case, a pair creation operation (inserting two mutually inverse factors anywhere in the braid factorization) is only allowed if the new factorization remains compatible with the monodromy representation $`\theta _{n1}`$, i.e. if $`\theta _{n1}`$ maps the two corresponding geometric generators to Dehn twists along disjoint Lagrangian spheres.
With this understood, we can introduce a notion of m-equivalence as in Definition 3.5. The following result then holds :
###### Theorem 4.3.
The braid factorizations and geometric monodromy representations associated to the quasiholomorphic maps to $`^2`$ obtained in Theorem 4.1 are, for $`k0`$, canonical up to m-equivalence, and define symplectic invariants of $`(X^{2n},\omega )`$.
Conversely, the data consisting of a braid factorization and a geometric $`(n1)`$-dimensional monodromy representation, or a m-equivalence class of such data, determines a symplectic $`2n`$-manifold in a canonical way up to symplectomorphism.
###### Remark 4.3.
The invariants studied in this section are a very natural generalization of those defined in §3.2 for 4-manifolds. Namely, when $`dimX=4`$, we naturally get that $`Z=\mathrm{}`$ and $`dim\mathrm{\Sigma }=0`$, i.e. the generic fiber $`\mathrm{\Sigma }`$ consists of a finite number of points, as expected for a branched covering map. In particular, the mapping class group $`\mathrm{Map}(\mathrm{\Sigma })`$ of the 0-manifold $`\mathrm{\Sigma }`$ is in fact the symmetric group of order $`\mathrm{card}(\mathrm{\Sigma })`$. Finally, a Lagrangian 0-sphere in $`\mathrm{\Sigma }`$ is just a pair of points of $`\mathrm{\Sigma }`$, and the associated Dehn twist is simply the corresponding transposition. With this correspondence, the results of §3 are the exact four-dimensional counterparts of those described here.
### 4.4. Quasiholomorphic maps and symplectic Lefschetz pencils
Consider again a symplectic manifold $`(X^{2n},\omega )`$ and let $`f:XZ^2`$ be a map with the same topological properties as those obtained by Theorem 4.1 from sections of $`L^k`$ for $`k`$ large enough. As in the four-dimensional case, the $`^1`$-valued map $`\pi f`$ defines a Lefschetz pencil structure on $`X`$, obtained by lifting via $`f`$ a pencil of lines on $`^2`$. The base set of this pencil is the fiber of $`f`$ above the pole $`(0:0:1)`$ of the projection $`\pi `$.
In fact, starting from the quasiholomorphic maps $`f_k`$ given by Theorem 4.1, the symplectic Lefschetz pencils $`\pi f_k`$ coincide for $`k0`$ with those obtained by Donaldson in and described in §2 ; calling $`s_k^0,s_k^1,s_k^2`$ the sections of $`L^k`$ defining $`f_k`$, the Lefschetz pencil $`\pi f_k`$ is the one induced by the sections $`s_k^0`$ and $`s_k^1`$.
Therefore, as in the case of a 4-manifold, the invariants described in §4.3 (braid factorization and $`(n1)`$-dimensional geometric monodromy representation) completely determine those discussed in §2 (factorizations in mapping class groups). Once again, the topological description of the relation between quasiholomorphic maps and Lefschetz pencils involves a subgroup of $`\theta _{n1}`$-liftable braids in the braid group, and a group homomorphism from this subgroup to a mapping class group.
Consider a symplectic braided curve $`D^2`$, described by its braid monodromy $`\rho _n:\pi _1(\{p_1,\mathrm{},p_r\})B_d`$, and a compatible $`(n1)`$-dimensional monodromy representation $`\theta _{n1}:F_d=\pi _1(\{q_1,\mathrm{},q_d\})\mathrm{Map}^\omega (\mathrm{\Sigma }^{2n4},Z^{2n6})`$. Then we can make the following definition :
###### Definition 4.2.
The subgroup $`B_d^0(\theta _{n1})`$ of liftable braids is the set of all braids $`QB_d`$ such that $`\theta _{n1}Q_{}=\theta _{n1}`$, where $`Q_{}\mathrm{Aut}(F_d)`$ is the automorphism induced by the braid $`Q`$ on $`\pi _1(\{q_1,\mathrm{},q_d\})`$.
A topological definition of $`B_d^0(\theta _{n1})`$ can also be given in terms of universal fibrations and coverings of configuration spaces, similarly to the description in §3.3.
More importantly, denote by $`W`$ the total space of the symplectic Lefschetz pencil $`LP(\theta _{n1})`$ with generic fiber $`\mathrm{\Sigma }`$ and monodromy $`\theta _{n1}`$. For example, if $`\rho _n`$ and $`\theta _{n1}`$ are the monodromy morphisms associated to a quasiholomorphic map given by sections $`s_k^0,s_k^1,s_k^2`$ of $`L^k`$ over $`X`$, then $`W`$ is the smooth symplectic hypersurface in $`X`$ given by the equation $`s_k^0=0`$ ; indeed, as seen in §4.3, this hypersurface carries a Lefschetz pencil structure with generic fiber $`\mathrm{\Sigma }`$, induced by $`s_k^1`$ and $`s_k^2`$, and the monodromy of this restricted pencil is precisely $`\theta _{n1}`$. A braid $`QB_d`$ can be viewed as a motion of the critical set $`\{q_1,\mathrm{},q_d\}`$ of the Lefschetz pencil $`LP(\theta _{n1})`$ ; after this motion we obtain a new Lefschetz pencil with monodromy $`\theta _{n1}Q_{}`$. So the subgroup $`B_d^0(\theta _{n1})`$ precisely consists of those braids which preserve the monodromy of the Lefschetz pencil $`LP(\theta _{n1})`$.
Viewing braids as compactly supported symplectomorphisms of the plane preserving $`\{q_1,\mathrm{},q_d\}`$, the fact that $`Q`$ belongs to $`B_d^0(\theta _{n1})`$ means that it can be lifted via the Lefschetz pencil map $`WZ^1`$ to a symplectomorphism of $`W`$. Since the monodromy of the pencil $`LP(\theta _{n1})`$ preserves a neighborhood of the base set $`Z`$, the lift to $`W`$ of the braid $`Q`$ coincides with the identity over a neighborhood of $`Z`$. Even better, because $`Q`$ is compactly supported, its lift to $`W`$ coincides with $`\mathrm{Id}`$ near the fiber above the point at infinity in $`^1`$, which can be identified with $`\mathrm{\Sigma }`$. Therefore, the lift of $`Q`$ to $`W`$ is a well-defined element of the mapping class group $`\mathrm{Map}^\omega (W,\mathrm{\Sigma })`$, which we call $`(\theta _{n1})_{}(Q)`$. This construction defines a group homomorphism
$$(\theta _{n1})_{}:B_d^0(\theta _{n1})\mathrm{Map}^\omega (W^{2n2},\mathrm{\Sigma }^{2n4}).$$
Since the geometric monodromy representation $`\theta _{n1}`$ is compatible with the braided curve $`D^2`$, the image of the braid monodromy homomorphism $`\rho _n:\pi _1(\{p_1,\mathrm{},p_r\})B_d`$ describing $`D`$ is entirely contained in $`B_d^0(\theta _{n1})`$. Indeed, it follows from Definition 4.1 that $`\theta _{n1}`$ factors through $`\pi _1(^2D)`$, on which the braids of $`\mathrm{Im}\rho _n`$ act trivially. As a consequence, we can use the group homomorphism $`(\theta _{n1})_{}`$ in order to obtain, from the braid monodromy $`\rho _n`$, a group homomorphism
$$\theta _n=(\theta _{n1})_{}\rho _n:\pi _1(\{p_1,\mathrm{},p_r\})\mathrm{Map}^\omega (W,\mathrm{\Sigma }).$$
If $`\rho _n`$ and $`\theta _{n1}`$ describe the monodromy of a $`^2`$-valued map $`f`$, then $`\theta _n`$ is by construction the monodromy of the corresponding Lefschetz pencil $`\pi f`$. Therefore, the following result holds :
###### Proposition 4.4.
Let $`f:XZ^2`$ be one of the quasiholomorphic maps of Theorem 4.1. Let $`D^2`$ be its critical curve, and denote by $`\rho _n:\pi _1(\{p_1,\mathrm{},p_r\})B_d^0(\theta _{n1})`$ and $`\theta :F_d\mathrm{Map}^\omega (\mathrm{\Sigma },Z)`$ be the corresponding monodromies. Then the monodromy map $`\theta _n:\pi _1(\{p_1,\mathrm{},p_r\})\mathrm{Map}^\omega (W,\mathrm{\Sigma })`$ of the Lefschetz pencil $`\pi f`$ is given by the identity $`\theta _n=(\theta _{n1})_{}\rho _n`$.
In particular, for $`k0`$ the symplectic invariants given by Theorem 2.1 are obtained in this manner from those defined in Theorem 4.3.
As in the four-dimensional case, all the factors of degree $`\pm 2`$ or $`3`$ in the braid monodromy (corresponding to the cusps and nodes of $`D`$) lie in the kernel of $`(\theta _{n1})_{}`$ ; the only terms which contribute non-trivially to the pencil monodromy $`\theta _n`$ are those arising from the tangency points of the branch curve $`D`$, and each of these contributions is a Dehn twist.
More precisely, the image in $`\mathrm{Map}^\omega (W,\mathrm{\Sigma })`$ of a half-twist $`QB_d^0(\theta _{n1})`$ arising as the braid monodromy around a tangency point of $`D`$ can be constructed as follows. Consider the Lefschetz pencil $`LP(\theta _{n1})`$ with total space $`W`$, generic fiber $`\mathrm{\Sigma }`$, critical levels $`q_1,\mathrm{},q_d`$ and monodromy $`\theta _{n1}`$. Call $`\gamma `$ the path joining two of the points $`q_1,\mathrm{},q_d`$ (e.g., $`q_{i_1}`$ and $`q_{i_2}`$) and naturally associated to the half-twist $`Q`$ (the path along which the twisting occurs). By Definition 4.1, the monodromies of $`LP(\theta _{n1})`$ around the two end points $`q_{i_1}`$ and $`q_{i_2}`$ are the same Dehn twists (using $`\gamma `$ to identify the two singular fibers). Even better, in this context one easily shows that the vanishing cycles at the two end points of $`\gamma `$ are isotopic Lagrangian spheres in $`\mathrm{\Sigma }`$. Then it follows from the work of Donaldson and Seidel that, above the path $`\gamma `$, one can find a Lagrangian sphere $`L=S^{n1}W`$, joining the singular points of the fibers above $`q_{i_1}`$ and $`q_{i_2}`$, and intersecting each fiber inbetween in a Lagrangian sphere $`S^{n2}`$. The element $`(\theta _{n1})_{}(Q)`$ in $`\mathrm{Map}^\omega (W,\mathrm{\Sigma })`$ is the positive Dehn twist along the Lagrangian sphere $`L`$.
###### Remark 4.4.
Let $`(X^{2n},\omega )`$ be a compact symplectic manifold, and consider the symplectic Lefschetz pencils given by Donaldson’s result (Theorem 2.1) from pairs of sections of $`L^k`$ for $`k0`$ ; the monodromy of these Lefschetz pencils consists of generalized Dehn twists around Lagrangian $`(n1)`$-spheres in the generic fiber $`W_k`$. It follows from Proposition 4.4 that these Lagrangian spheres are not arbitrary. Indeed, they can all be obtained by endowing $`W_k`$ with a structure of symplectic Lefschetz pencil induced by two sections of $`L^k`$ (the existence of such a structure follows from the results of this section), and by looking for Lagrangian $`(n1)`$-spheres which join two mutually isotopic vanishing cycles of this pencil above a path in the base.
As observed by Seidel, this remarkable structure of vanishing cycles makes it possible to hope for a purely combinatorial description of Lagrangian Floer homology : one can try to use the structure of vanishing cycles in a $`2n`$-dimensional Lefschetz pencil to reduce things first to the $`2n2`$-dimensional case, and then by induction eventually to the case of $`0`$-manifolds, in which the calculations are purely combinatorial.
## 5. Complete linear systems and dimensional induction
We now show how the results of §4 can be used in order to reduce in principle the classification of compact symplectic manifolds to a purely combinatorial problem.
###### Definition 5.1.
Let $`(X^{2n},\omega )`$ be a compact symplectic manifold. We say that asymptotically holomorphic $`(n+1)`$-tuples of sections of $`L^k`$ define 3-good complete linear systems on $`X`$ if, for large values of $`k`$, these sections $`s_0,\mathrm{},s_n\mathrm{\Gamma }(L^k)`$ satisfy the following properties :
$`(a)`$ for $`0rn1`$, the section $`(s_{r+1},\mathrm{},s_n)`$ of $`^{nr}L^k`$ satisfies a uniform transverslity property, and its zero set $`\mathrm{\Sigma }_r=\{s_{r+1}=\mathrm{}=s_n=0\}`$ is a smooth symplectic submanifold of dimension $`2r`$ in $`X`$. We also define $`\mathrm{\Sigma }_n=X`$ and $`\mathrm{\Sigma }_1=\mathrm{}`$ ;
$`(b)`$ for $`1rn`$, the pair of sections $`(s_r,s_{r1})\mathrm{\Gamma }(^2L^k)`$ defines a structure of symplectic Lefschetz pencil on $`\mathrm{\Sigma }_r`$, with generic fiber $`\mathrm{\Sigma }_{r1}`$ and base set $`\mathrm{\Sigma }_{r2}`$ ;
$`(c)`$ for $`2rn`$, the triple of sections $`(s_r,s_{r1},s_{r2})\mathrm{\Gamma }(^3L^k)`$ defines a quasiholomorphic map from $`\mathrm{\Sigma }_r`$ to $`^2`$, with generic fiber $`\mathrm{\Sigma }_{r2}`$ and base set $`\mathrm{\Sigma }_{r3}`$.
One can think of a 3-good complete linear system in the following way. First, the two sections $`s_n`$ and $`s_{n1}`$ define a Lefschetz pencil structure on $`X`$. By adding the section $`s_{n2}`$, this structure is refined into a quasiholomorphic map to $`^2`$. As observed in §4, by restricting to the hypersurface $`\mathrm{\Sigma }_{n1}`$ we get a symplectic Lefschetz pencil defined by $`s_{n1}`$ and $`s_{n2}`$. This structure is in turn refined into a quasiholomorphic map by adding the section $`s_{n3}`$ ; and so on.
Note that, except for the case $`r=1`$, part $`(b)`$ of Definition 5.1 is actually an immediate consequence of part $`(c)`$, because by composing $`^2`$-valued quasiholomorphic maps with the projection $`\pi :^2\{(0:0:1)\}^1`$ one always obtains Lefschetz pencils. Also note that, in order to make sense out of these properties, one implicitly needs to endow the submanifolds $`\mathrm{\Sigma }_r`$ with $`\omega `$-compatible almost-complex structures ; these restricted almost-complex structures can be chosen to differ from the almost-complex structure $`J`$ on $`X`$ by $`O(k^{1/2})`$, so that asymptotic holomorphicity and transversality properties are not affected by this choice.
###### Theorem 5.1.
Let $`(X^{2n},\omega )`$ be a compact symplectic manifold. Then for all large enough values of $`k`$ it is possible to find asymptotically holomorphic sections of $`^{n+1}L^k`$ determining 3-good complete linear systems on $`X`$. Moreover, for large $`k`$ these structures are canonical up to isotopy and up to cancellations of pairs of nodes in the critical curves of the quasiholomorphic $`^2`$-valued maps.
###### Proof.
We only give a sketch of the proof of Theorem 5.1. As usual, we need to obtain two types of properties : uniform transversality conditions, which we ensure in the first part of the argument, and compatibility conditions, which are obtained by a subsequent perturbation. As in previous arguments, the various uniform transversality properties are obtained successively, using the fact that, because transversality is an open condition, it is preserved by any sufficiently small subsequent perturbations.
The first transversality properties to be obtained are those appearing in part $`(a)`$ of Definition 5.1, i.e. the transversality to $`0`$ of $`(s_{r+1},\mathrm{},s_n)`$ for all $`0rn1`$ ; this easy case is e.g. covered by the main result of .
One next turns to the transversality conditions arising from the requirement that the three sections $`(s_n,s_{n1},s_{n2})`$ define quasiholomorphic maps from $`X`$ to $`^2`$ : it follows immediately from the proof of Theorem 4.1 that these properties can be obtained by suitable small perturbations.
Next, we try to modify $`s_{n1}`$, $`s_{n2}`$ and $`s_{n3}`$ in order to ensure that the restrictions to $`\mathrm{\Sigma }_{n1}=s_n^1(0)`$ of these three sections satisfy the transversality properties of Definition 3.2. A general strategy to handle this kind of situation is to use the following remark (Lemma 6 of ) : if $`\varphi `$ is a section of a vector bundle $``$ over $`X`$, satisfying a uniform transversality property, and if $`W=\varphi ^1(0)`$, then the uniform transversality to $`0`$ over $`W`$ of a section $`\xi `$ of a vector bundle $``$ is equivalent to the uniform transversality to $`0`$ over $`X`$ of the section $`\xi \varphi `$ of $``$, up to a change in transversality estimates. This makes it possible to replace all transversality properties to be satisfied over submanifolds of $`X`$ by transversality properties to be satisfied over $`X`$ itself ; each property can then be ensured by the standard type of argument, using the globalization principle to combine suitably chosen local perturbations (see for more details).
However, in our case the situation is significantly simplified by the fact that, no matter how we perturb the sections $`s_{n1}`$, $`s_{n2}`$ and $`s_{n3}`$, the submanifold $`\mathrm{\Sigma }_{n1}`$ itself is not affected. Moreover, the geometry of $`\mathrm{\Sigma }_{n1}`$ is controlled by the transversality properties obtained on $`s_n`$ ; for example, a suitable choice of the constant $`\rho >0`$ (independent of $`k`$) ensures that the intersection of $`\mathrm{\Sigma }_{n1}`$ with any ball of $`g_k`$-radius $`\rho `$ centered at one of its points is topologically a ball (see e.g. Lemma 4 of ). Therefore, we can actually imitate all steps of the argument used to prove Theorem 4.1, working with sections of $`L^k`$ over $`\mathrm{\Sigma }_{n1}`$. The localized reference sections of $`L^k`$ over $`\mathrm{\Sigma }_{n1}`$ that we use in the arguments are now chosen to be the restrictions to $`\mathrm{\Sigma }_{n1}`$ of the localized sections $`s_{k,x}^{\mathrm{ref}}`$ of $`L^k`$ over $`X`$ ; similarly, the approximately holomorphic local coordinates over $`\mathrm{\Sigma }_{n1}`$ in which we work are obtained as the restrictions to $`\mathrm{\Sigma }_{n1}`$ of local coordinate functions on $`X`$. With these two differences understood, we can still construct localized perturbations by the same algorithms as in §4.1 and, using the standard globalization argument, achieve the desired transversality properties over $`\mathrm{\Sigma }_{n1}`$. Moreover, all these local perturbations are obtained as products of the localized reference sections by polynomial functions of the local coordinates. Therefore, they naturally arise as restrictions to $`\mathrm{\Sigma }_{n1}`$ of localized sections of $`L^k`$ over $`X`$, and so we actually obtain well-defined perturbations of the sections $`s_{n1}`$, $`s_{n2}`$ and $`s_{n3}`$ over $`X`$ which yield the desired transversality properties over $`\mathrm{\Sigma }_{n1}`$.
We can continue similarly by induction on the dimension, until we obtain the transversality properties required of $`s_2`$, $`s_1`$ and $`s_0`$ over $`\mathrm{\Sigma }_2`$, and finally the transversality properties required of $`s_1`$ and $`s_0`$ over $`\mathrm{\Sigma }_1`$. Observe that, even though the perturbations performed over each $`\mathrm{\Sigma }_r`$ result in modifications of the submanifolds $`\mathrm{\Sigma }_j`$ ($`j<r`$) lying inside them, these perturbations preserve the transversality properties of $`(s_{j+1},\mathrm{},s_n)`$, and so the submanifolds $`\mathrm{\Sigma }_j`$ retain their smoothness and symplecticity properties.
We now turn to the second part of the argument, i.e. obtaining the desired compatibility conditions. First observe that the proof of Theorem 4.1 shows how, by a perturbation of $`s_n`$, $`s_{n1}`$ and $`s_{n2}`$ smaller than $`O(k^{1/2})`$, we can ensure that the various compatibility properties of Definition 3.2 are satisfied by the $`^2`$-valued map $`f_n`$ defined by these three sections.
Next, we proceed to perturb $`f_{n1}=(s_{n1}:s_{n2}:s_{n3})`$ over a neighborhood of its ramification curve $`R_{n1}\mathrm{\Sigma }_{n1}`$, in order to obtain the required compatibility properties for $`f_{n1}`$, but without losing those previously achieved for $`f_n`$ near its ramification curve $`R_nX`$. For this purpose, we first show that the curve $`R_n`$ satisfies a uniform transversality property with respect to the hypersurface $`\mathrm{\Sigma }_{n1}`$ in $`X`$.
The only way in which $`R_n`$ can fail to be uniformly transverse to $`\mathrm{\Sigma }_{n1}`$ is if $`(\pi f_{n|R_n})`$ becomes small at a point of $`R_n`$ near $`\mathrm{\Sigma }_{n1}`$. Because $`f_n`$ satisfies property $`(6)`$ in Definition 3.2, this can only happen if a cusp point or a tangency point of $`f_n`$ lies close to $`\mathrm{\Sigma }_{n1}`$. However, property $`(7)`$ of Definition 3.2 implies that this point cannot belong to $`\mathrm{\Sigma }_{n1}`$. Therefore, two of the intersection points of $`R_n`$ with $`\mathrm{\Sigma }_{n1}`$ must lie close to each other. Observe that the points of $`R_n\mathrm{\Sigma }_{n1}`$ are precisely the critical points of the Lefschetz pencil induced on $`\mathrm{\Sigma }_{n1}`$ by $`s_{n1}`$ and $`s_{n2}`$, i.e. the tangency points of the map $`f_{n1}`$. The transversality properties already obtained for $`f_{n1}`$ imply that two tangency points cannot lie close to each other ; we get a contradiction, so the cusps and tangencies of $`f_n`$ must lie far away from $`\mathrm{\Sigma }_{n1}`$, and $`R_n`$ and $`\mathrm{\Sigma }_{n1}`$ are mutually transverse.
This implies in particular that a small perturbation of $`s_{n1}`$, $`s_{n2}`$ and $`s_{n3}`$ localized near $`\mathrm{\Sigma }_{n1}`$ cannot affect properties $`(4^{})`$ and $`(6^{})`$ for $`f_n`$, and also that the only place where perturbing $`f_{n1}`$ might affect $`f_n`$ is near the tangency points of $`f_{n1}`$.
We now consider the set $`𝒞_{n1}𝒯_{n1}_{n1}`$ of points where we need to ensure properties $`(4^{})`$, $`(6^{})`$ and $`(8^{})`$ for $`f_{n1}`$. The first step is as usual to perturb $`J`$ into an almost-complex structure which is integrable near these points ; once this is done, we perturb $`f_{n1}`$ to make it locally holomorphic with respect to this almost-complex structure.
We start by considering a point $`x𝒞_{n1}_{n1}`$, where the issue of preserving properties of $`f_n`$ does not arise. We follow the argument in §4.1 of . First, it is possible to perturb the almost-complex structure $`J`$ over a neighborhood of $`x`$ in $`X`$ in order to obtain an almost-complex structure $`\stackrel{~}{J}`$ which differs from $`J`$ by $`O(k^{1/2})`$ and is integrable over a small ball centered at $`x`$. Recall from that $`\stackrel{~}{J}`$ is obtained by choosing approximately holomorphic coordinates on $`X`$ and using them to pull back the standard complex structure of $`^n`$ ; a cut-off function is used to splice $`J`$ with this locally defined integrable structure. Since we can choose the local coordinates in such a way that a local equation of $`\mathrm{\Sigma }_{n1}`$ is $`z_n=0`$, we can easily ensure that $`\mathrm{\Sigma }_{n1}`$ is, over a small neighborhood of $`x`$, a $`\stackrel{~}{J}`$-holomorphic submanifold of $`X`$. Next, we can perturb the sections $`s_{n1},s_{n2},s_{n3}`$ of $`L^k`$ by $`O(k^{1/2})`$ in order to make the projective map defined by them $`\stackrel{~}{J}`$-holomorphic over a neighborhood of $`x`$ in $`X`$ (see ). This holomorphicity property remains true for the restrictions to the locally $`\stackrel{~}{J}`$-holomorphic submanifold $`\mathrm{\Sigma }_{n1}`$. So, we have obtained the desired compatibility property near $`x`$.
We now consider the case of a point $`x𝒯_{n1}`$, where we need to obtain property $`(6^{})`$ for $`f_{n1}`$ while preserving property $`(8^{})`$ for $`f_n`$. We first observe that, by the construction of the previous step (getting property $`(8^{})`$ for $`f_n`$ at $`x`$), we have a readily available almost-complex structure $`\stackrel{~}{J}`$ integrable over a neighborhood of $`x`$ in $`X`$. In particular, by construction $`f_n`$ is locally $`\stackrel{~}{J}`$-holomorphic and $`\mathrm{\Sigma }_{n1}`$ is locally a $`\stackrel{~}{J}`$-holomorphic submanifold of $`X`$. We next try to make the projective map $`f_{n1}`$ holomorphic over a neighborhood of $`x`$, using once again the argument of . The key observation here is that, because one of the sections $`s_{n1}`$ and $`s_{n2}`$ is bounded from below at $`x`$, we can reduce to a $`^2`$-valued map whose first component is already holomorphic. Therefore, the perturbation process described in only affects $`s_{n3}`$, while the two other sections are preserved. This means that we can ensure the local $`\stackrel{~}{J}`$-holomorphicity of $`f_{n1}`$ without affecting $`f_n`$.
It is easy to combine the various localized perturbations performed near each point of $`𝒞_{n1}𝒯_{n1}_{n1}`$ ; this yields properties $`(4^{})`$, $`(6^{})`$ and $`(8^{})`$ of Definition 3.2 for $`f_{n1}`$.
We now use a generically chosen small perturbation of $`s_{n1}`$, $`s_{n2}`$ and $`s_{n3}`$ in order to ensure property $`(7)`$, i.e. the self-transversality of the critical curve of $`f_{n1}`$. It is important to observe that, because $`f_n`$ satisfies property $`(7)`$, the images by the projective map $`(s_{n1}:s_{n2})`$ of the points of $`R_n\mathrm{\Sigma }_{n1}=_n=𝒯_{n1}`$ are all distinct from each other, and because $`f_n`$ satisfies property $`(5)`$ they are also distinct from $`(0:1)`$. Therefore, we can choose a perturbation which vanishes identically over a neighborhood of $`𝒯_{n1}`$ ; this makes it possible to obtain property $`(7)`$ for $`f_{n1}`$ without losing any property of $`f_n`$.
Finally, by the process described in §4.2 of we construct a perturbation yielding property $`(3^{})`$ along the critical curve of $`f_{n1}`$ ; this perturbation is originally defined only for the restrictions to $`\mathrm{\Sigma }_{n1}`$ but it can easily be extended outside of $`\mathrm{\Sigma }_{n1}`$ by using a cut-off function. The two important properties of this perturbation are the following : first, it vanishes identically near the points where $`f_{n1}`$ has already been made $`\stackrel{~}{J}`$-holomorphic, and in particular near the points of $`𝒯_{n1}`$ ; therefore, none of the properties of $`f_n`$ are affected, and properties $`(4^{})`$, $`(6^{})`$ and $`(8^{})`$ of $`f_{n1}`$ are not affected either. Secondly, this perturbation does not modify the critical curve of $`f_{n1}`$ nor its image, so property $`(7)`$ is preserved. We have therefore obtained all desired properties for $`f_{n1}`$.
We can continue similarly by induction on the dimension, until all required compatibility properties are satisfied. Observe that, because the ramification curve of $`f_r`$ remains away from its fiber at infinity $`\mathrm{\Sigma }_{r2}`$, we do not need to worry about the possible effects on $`f_r`$ of perturbations of $`f_{r2}`$. Therefore, the argument remains the same at each step, and we can complete the proof of the existence statement in Theorem 5.1 in this way.
The proof of the uniqueness statement relies, as usual, on the extension of the whole construction to one-parameter families ; this is easily done by following the same ideas as in previous arguments. ∎
The structures of 3-good complete linear systems given by Theorem 5.1 are extremely rich, and lead to interesting invariants of compact symplectic manifolds. Indeed, recall from Definition 5.1 that, for $`1rn`$, the sections $`s_r`$ and $`s_{r1}`$ define a symplectic Lefschetz pencil structure on $`\mathrm{\Sigma }_r`$, with generic fiber $`\mathrm{\Sigma }_{r1}`$ and base set $`\mathrm{\Sigma }_{r2}`$. The monodromy of this pencil is given by a group homomorphism
$$\theta _r:\pi _1(\{p_1,\mathrm{},p_{d_r}\})\mathrm{Map}^\omega (\mathrm{\Sigma }_{r1},\mathrm{\Sigma }_{r2}).$$
(10)
Moreover, for $`2rn`$, the sections $`s_r`$, $`s_{r1}`$ and $`s_{r2}`$ define a quasiholomorphic map from $`\mathrm{\Sigma }_r\mathrm{\Sigma }_{r3}`$ to $`^2`$, with generic fiber $`\mathrm{\Sigma }_{r2}`$. Denote by $`D_r^2`$ the critical curve of this map, and let $`d_{r1}=\mathrm{deg}D_r`$. As shown in §4.3, we obtain two monodromy morphisms : on one hand, the braid monodromy homomorphism characterizing $`D_r`$,
$$\rho _r:\pi _1(\{p_1,\mathrm{},p_{s_r}\})B_{d_{r1}},$$
(11)
and on the other hand, a compatible $`(r1)`$-dimensional monodromy representation, which was shown in §4.3 to be none other than
$$\theta _{r1}:\pi _1(\{p_1,\mathrm{},p_{d_{r1}}\})\mathrm{Map}^\omega (\mathrm{\Sigma }_{r2},\mathrm{\Sigma }_{r3}).$$
Finally, it was shown in §4.4 that $`\mathrm{Im}(\rho _r)B_{d_{r1}}^0(\theta _{r1})`$, and that the various monodromies are related to each other by the identity
$$\theta _r=(\theta _{r1})_{}\rho _r.$$
(12)
In particular, the manifold $`X`$ is completely characterized by the braid monodromies $`\rho _2,\mathrm{},\rho _n`$ and by the map $`\theta _1`$ with values in $`\mathrm{Map}^\omega (\mathrm{\Sigma }_0,\mathrm{})`$, which is a symmetric group ; this data is sufficient to successively reconstruct all morphisms $`\theta _r`$ and all submanifolds $`\mathrm{\Sigma }_r`$ by inductively using equation $`(12)`$.
In other words, a symplectic $`2n`$-manifold is characterized by $`n2`$ braid factorizations and a word in a symmetric group ; or, stopping at $`\theta _2`$, we can also consider $`n3`$ braid factorizations and a word in the mapping class group of a Riemann surface.
These results can be summarized by the following theorem :
###### Theorem 5.2.
The braid monodromies $`\rho _2,\mathrm{},\rho _n`$ and the symmetric group representation $`\theta _1`$ associated to the 3-good complete linear systems obtained in Theorem 5.1 are, for $`k0`$, canonical up to m-equivalence, and define symplectic invariants of $`(X^{2n},\omega )`$.
Conversely, the data consisting of several braid factorizations and a symmetric group representation satisfying suitable compatibility conditions, or a m-equivalence class of such data, determines a symplectic $`2n`$-manifold in a canonical way up to symplectomorphism.
In principle, this result reduces the study of compact symplectic manifolds to purely combinatorial questions about braid groups and symmetric groups ; however, the invariants it introduces are probably quite difficult to compute as soon as one considers examples which are not complex algebraic. Nevertheless, it seems that this construction should be very helpful in improving our understanding of the topology of Lefschetz pencils in dimensions greater than $`4`$.
|
warning/0007/hep-ex0007027.html
|
ar5iv
|
text
|
# 1 The 𝜋⁻𝜋⁺𝜋⁺ effective mass spectrum. The dotted line represents the 𝐷⁰→𝐾⁻𝜋⁺ plus 𝐷_𝑠⁺→𝜂'𝜋⁺ and the dashed line is the total backgound. Events in the hatched areas at the 𝐷_𝑠 mass are used for the 𝐷_𝑠 Dalitz plot analysis in this Letter. The hatched area at the 𝐷⁺ mass is used for the analysis in the following companion paper [4].
Study of the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decay and measurement of $`f_0`$ masses and widths.
E. M. Aitala,<sup>9</sup> S. Amato,<sup>1</sup> J. C. Anjos,<sup>1</sup> J. A. Appel,<sup>5</sup> D. Ashery,<sup>14</sup> S. Banerjee,<sup>5</sup> I. Bediaga,<sup>1</sup> G. Blaylock,<sup>8</sup> S. B. Bracker,<sup>15</sup> P. R. Burchat,<sup>13</sup> R. A. Burnstein,<sup>6</sup> T. Carter,<sup>5</sup> H. S. Carvalho,<sup>1</sup> N. K. Copty,<sup>12</sup> L. M. Cremaldi,<sup>9</sup> C. Darling,<sup>18</sup> K. Denisenko,<sup>5</sup> S. Devmal,<sup>3</sup> A. Fernandez,<sup>11</sup> G. F. Fox,<sup>12</sup> P. Gagnon,<sup>2</sup> C. Gobel,<sup>1</sup> K. Gounder,<sup>9</sup> A. M. Halling,<sup>5</sup> G. Herrera,<sup>4</sup> G. Hurvits,<sup>14</sup> C. James,<sup>5</sup> P. A. Kasper,<sup>6</sup> S. Kwan,<sup>5</sup> D. C. Langs,<sup>12</sup> J. Leslie,<sup>2</sup> B. Lundberg,<sup>5</sup> J. Magnin,<sup>1</sup> A. Massafferri,<sup>1</sup> S. MayTal-Beck,<sup>14</sup> B. Meadows,<sup>3</sup> J. R. T. de Mello Neto,<sup>1</sup> D. Mihalcea,<sup>7</sup> R. H. Milburn,<sup>16</sup> J. M. de Miranda,<sup>1</sup> A. Napier,<sup>16</sup> A. Nguyen,<sup>7</sup> A. B. d’Oliveira,<sup>3,11</sup> K. O’Shaughnessy,<sup>2</sup> K. C. Peng,<sup>6</sup> L. P. Perera,<sup>3</sup> M. V. Purohit,<sup>12</sup> B. Quinn,<sup>9</sup> S. Radeztsky,<sup>17</sup> A. Rafatian,<sup>9</sup> N. W. Reay,<sup>7</sup> J. J. Reidy,<sup>9</sup> A. C. dos Reis,<sup>1</sup> H. A. Rubin,<sup>6</sup> D. A. Sanders,<sup>9</sup> A. K. S. Santha,<sup>3</sup> A. F. S. Santoro,<sup>1</sup> A. J. Schwartz,<sup>3</sup> M. Sheaff,<sup>17</sup> R. A. Sidwell,<sup>7</sup> A. J. Slaughter,<sup>18</sup> M. D. Sokoloff,<sup>3</sup> J. Solano,<sup>1</sup> N. R. Stanton,<sup>7</sup> R. J. Stefanski,<sup>5</sup> K. Stenson,<sup>17</sup> D. J. Summers,<sup>9</sup> S. Takach,<sup>18</sup> K. Thorne,<sup>5</sup> A. K. Tripathi,<sup>7</sup> S. Watanabe,<sup>17</sup> R. Weiss-Babai,<sup>14</sup> J. Wiener,<sup>10</sup> N. Witchey,<sup>7</sup> E. Wolin,<sup>18</sup> S. M. Yang,<sup>7</sup> D. Yi,<sup>9</sup> S. Yoshida,<sup>7</sup> R. Zaliznyak,<sup>13</sup> and C. Zhang<sup>7</sup>
(Fermilab E791 Collaboration)
<sup>1</sup> Centro Brasileiro de Pesquisas Físicas, Rio de Janeiro, Brazil, <sup>2</sup> University of California, Santa Cruz, California 95064, <sup>3</sup> University of Cincinnati, Cincinnati, Ohio 45221, <sup>4</sup> CINVESTAV, Mexico City, Mexico, <sup>5</sup> Fermilab, Batavia, Illinois 60510, <sup>6</sup> Illinois Institute of Technology, Chicago, Illinois 60616, <sup>7</sup> Kansas State University, Manhattan, Kansas 66506, <sup>8</sup> University of Massachusetts, Amherst, Massachusetts 01003, <sup>9</sup> University of Mississippi-Oxford, University, Mississippi 38677, <sup>10</sup> Princeton University, Princeton, New Jersey 08544, <sup>11</sup> Universidad Autonoma de Puebla, Puebla, Mexico, <sup>12</sup> University of South Carolina, Columbia, South Carolina 29208, <sup>13</sup> Stanford University, Stanford, California 94305, <sup>14</sup> Tel Aviv University, Tel Aviv, Israel, <sup>15</sup> Box 1290, Enderby, British Columbia, V0E 1V0, Canada, <sup>16</sup> Tufts University, Medford, Massachusetts 02155, <sup>17</sup> University of Wisconsin, Madison, Wisconsin 53706, <sup>18</sup> Yale University, New Haven, Connecticut 06511
August, 2000
## Abstract
From a sample of 848 $`\pm `$ 44 $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decays, we find $`\mathrm{\Gamma }(D_s^+\pi ^{}\pi ^+\pi ^+)/\mathrm{\Gamma }(D_s^+\varphi \pi ^+)=0.245\pm 0.028_{0.012}^{+0.019}`$. Using a Dalitz plot analysis of this three body decay, we find significant contributions from the channels $`\rho ^0(770)\pi ^+`$, $`\rho ^0(1450)\pi ^+`$, $`f_0(980)\pi ^+`$, $`f_2(1270)\pi ^+`$, and $`f_0(1370)\pi ^+`$. We present also the values obtained for masses and widths of the resonances $`f_0(980)`$ and $`f_0(1370)`$.
The charm meson decay $`D_s^+\pi ^+\pi ^{}\pi ^+`$ and its charge conjugate (implicit throughout this paper) is Cabibbo-favored but has no strange meson in the final state. The decay can proceed via spectator amplitudes, producing intermediate resonant states with hidden strangeness, e.g. $`D_s^+f_0(980)\pi ^+`$, with $`s\overline{s}`$ quarks in the $`f_0(980)`$. Also, the decay can proceed via $`W`$-annihilation amplitudes, producing intermediate states with no strangeness, e.g. $`D_s^+\rho ^0\pi ^+`$. A $`W`$-annihilation amplitude could also produce the intermediate state $`D_s^+f_0(1370)\pi ^+`$, assuming the $`f_0(1370)`$ consists mostly of $`u\overline{u}`$ and $`d\overline{d}`$ quarks as predicted by the simple quark model. To determine the relative importance of these different decay mechanisms, one can use an amplitude analysis. Such an analysis is also able to determine the masses and decay widths of the intermediate states.
In general, scalar resonances have large decay fractions in the 3-body decays of $`D`$-mesons, and such decays provide a relatively clean laboratory in which to study the properties of the scalars. In particular, isoscalar intermediate states are dominant in $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decays. The largest contribution to this final state comes from the decay involving the scalar meson $`f_0(980)`$ whose nature is a long-standing puzzle. It has been described as a $`q\overline{q}`$ state , a $`K\overline{K}`$ molecule, a glueball, and a 4-quark state.
In this paper we extend the reach of previous studies using the larger data sample from Fermilab experiment E791. We present an amplitude analysis which includes a greater number of possible resonant states, and we measure the masses and widths of the scalar resonances $`f_0(980)`$ and $`f_0(1370)`$ with better precision. Taken together with the results of the companion analysis, these results provide new insights into the characteristics of scalar mesons and their importance in charm meson decay.
The data were produced by 500 GeV/$`c`$ $`\pi ^{}`$ interactions in five thin foils (one platinum, four diamond) separated by gaps of 1.34 to 1.39 cm. The detector, the data set, the reconstruction, and the resulting vertex resolutions have been described previously. After reconstruction, events with evidence of well-separated production (primary) and decay (secondary) vertices were retained for further analysis. From the 3-prong secondary vertex candidates, we select a $`\pi ^{}\pi ^+\pi ^+`$ sample with invariant mass ranging from 1.7 to 2.1 GeV/c<sup>2</sup>. For this analysis all charged particles are taken to be pions; i.e., no direct use is made of particle identification.
We require a candidate’s secondary vertex position to be cleanly separated from the event’s primary vertex position and from the closest target material. The sum of the momentum vectors of the three tracks from this secondary vertex must point to the primary vertex. The candidate’s daughter tracks must pass closer to the secondary vertex than to the primary vertex, and must not point back to the primary vertex. The resulting invariant mass spectrum is shown in Figure 1.
We fit the spectrum of Figure 1 as the sum of $`D^+`$ and $`D_s^+`$ signals plus background. To account for the signal’s non Gaussian tails, we model each signal as the sum of two Gaussian distributions with the same centroidbut different widths. We model the background as the sum of four components: a general combinatorial background, the reflection of the $`D^+K^{}\pi ^+\pi ^+`$ decay, reflections of $`D^0K^{}\pi ^+`$ plus one extra track (mostly from the primary vertex), and $`D_s^+\eta ^{}\pi ^+`$ followed by $`\eta ^{}\rho ^0(770)\gamma `$, $`\rho ^0(770)\pi ^+\pi ^{}`$. The $`D^+K^{}\pi ^+\pi ^+`$ reflection is located below 1.85 GeV/$`c^2`$ in the $`\pi ^{}\pi ^+\pi ^+`$ spectrum. The other charm backgrounds populate the whole $`\pi ^{}\pi ^+\pi ^+`$ spectrum. We use Monte Carlo (MC) simulations to determine the shape of each identified charm background in the $`\pi ^{}\pi ^+\pi ^+`$ spectrum. We assume that the combinatorial background falls exponentially with mass. The levels of $`D^0K^{}\pi ^+`$ and $`D_s^+\eta ^{}\pi ^+`$ backgrounds are determined using charm signal rates measured in our total event sample and branching ratios taken from the compilation by the Particle Data Group. The parameters describing the combinatorial background and the level of the $`D^+K^{}\pi ^+\pi ^+`$ reflection are determined from fitting the $`\pi ^{}\pi ^+\pi ^+`$ distribution. The mass (centroid) and both Gaussian widths for each signal float in our fit. The fit finds 1172 $`\pm `$ 61 $`D^+`$ events and 848 $`\pm `$ 44 $`D_s^+`$ events. The $`D_s^+`$ signal is used for the measurement of the decay rate $`\mathrm{\Gamma }(D_s^+\pi ^{}\pi ^+\pi ^+)`$ relative to $`\mathrm{\Gamma }(D_s^+\varphi \pi ^+)`$ and for an amplitude analysis of the $`D_s^+`$ Dalitz plot.
To minimize systematic effects when calculating the ratio of efficiencies, we select the $`D_s^+\varphi \pi ^+`$, $`\varphi K^{}K^+`$ signal using the same track and vertex quality criteria as for the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decay. The number of normalization events is found to be 1038 $`\pm `$ 44.
We use MC simulations to correct the signals for geometrical acceptance and detector efficiency. We measure a ratio of efficiencies $`\epsilon (D_s^+\pi ^{}\pi ^+\pi ^+)/\epsilon (D_s^+\varphi \pi ^+)=1.64\pm 0.15`$ with values $`\epsilon (D_s^+\pi ^{}\pi ^+\pi ^+)`$ and $`\epsilon (D_s^+\varphi \pi ^+)`$ of about 2% and 1%, respectively. The error systematic is dominated by uncertainties in the MC model of $`D`$ production and relative efficiencies of our selection criteria.
The branching fraction for $`D_s^+\pi ^{}\pi ^+\pi ^+`$ relative to that for $`D_s^+\varphi \pi ^+`$ is measured to be:
$$\frac{B(D_s^+\pi ^{}\pi ^+\pi ^+)}{B(D_s^+\varphi \pi ^+)}=0.245\pm 0.028_{0.012}^{+0.019}.$$
(1)
The first error is statistical. The second is systematic, and is dominated by uncertainties related to the signal and background shapes used in the fit, the background levels, and the sample selection criteria. This value is smaller than the other experimental results: E691 found $`0.44\pm 0.10\pm 0.04`$, WA82 quoted $`0.33\pm 0.10\pm 0.04`$ and E687 presented $`0.33\pm 0.058\pm 0.058`$. The PDG presents a value of $`0.28\pm 0.06`$ from a constrained fit.
The symmetrized Dalitz plot of the 937 candidates with invariant mass between 1.95 and 1.99 GeV/c<sup>2</sup> is shown in Fig. 2. The integrated signal to background ratio is $`2`$. The narrow horizontal and vertical bands of $`s_{12}m^2(\pi _1^{}\pi _2^+)`$ and $`s_{13}m^2(\pi _1^{}\pi _3^+)`$ just below 1 GeV$`{}_{}{}^{2}/c^4`$ correspond to the $`f_0(980)\pi ^+`$ state with constructive quantum-mechanical interference evident where the two bands overlap (the event count is four times that of the individual bands). At the upper edge of the diagonal, there is another concentration of events centered at $`s_{12}s_{13}`$1.8 GeV$`{}_{}{}^{2}/c^4`$, corresponding to the $`f_2(1270)\pi ^+`$, $`f_0(1370)\pi ^+`$, and $`\rho ^0(1450)\pi ^+`$.
We fit the distribution shown in Figure 2 to a signal probability distribution function (PDF), which is a coherent sum of amplitudes corresponding to the non-resonant decay plus five different resonant channels, and a background PDF of known shape and magnitude. The resonant channels we include in the fit are $`\rho ^0(770)\pi ^+`$, $`f_0(980)\pi ^+`$, $`f_2(1270)\pi ^+`$, $`f_0(1370)\pi ^+`$, and $`\rho ^0(1450)\pi ^+`$. We weight the signal PDF by the acceptance across the Dalitz plot and use the measured line shape and background to determine the ratio of expected signal to background for each event in the Dalitz plot.
We assume the non-resonant amplitude to be uniform across the Dalitz plot. Each resonant amplitude, except that for the $`f_0(980)`$, is parameterized as a product of form factors, a relativistic Breit-Wigner function, and an angular momentum amplitude which depends on the spin of the resonance,
$$𝒜_n=\frac{F_D^JF_n}{m_{\pi \pi }^2m_0^2+im_0\mathrm{\Gamma }(m_{\pi \pi })}_n^J,$$
(2)
with
$$\mathrm{\Gamma }(m_{\pi \pi })=\mathrm{\Gamma }_0\frac{m_0}{m_{\pi \pi }}\left(\frac{p^{}}{p_0^{}}\right)^{2J+1}\frac{{}_{}{}^{J}F_{n}^{2}(p^{})}{{}_{}{}^{J}F_{n}^{2}(p_0^{})}.$$
(3)
The form factors $`F_D`$ and $`{}_{}{}^{J}F_{n}^{}`$ are the Blatt-Weisskopf damping factors respectively for the $`D`$ and the resonance decays, $`p^{}`$ is the pion momentum in the resonance rest frame at mass $`m_{\pi \pi }(p_0^{}=p^{}(m_0))`$ and $`J`$ is the spin of resonance $`n`$. $`_n^J`$ describes the angular distribution due to the spin. Since we have identical particles in the final state, each signal amplitude is Bose-symmetrized, $`𝒜_n=𝒜_n[(\mathrm{𝟏𝟐})\mathrm{𝟑}]+𝒜_n[(\mathrm{𝟏𝟑})\mathrm{𝟐}]`$.
For the $`f_0(980)\pi ^+`$ we use a coupled-channel Breit-Wigner function, following the parameterization of the WA76 Collaboration,
$$BW_{f_0(980)}=\frac{1}{m_{\pi \pi }^2m_0^2+im_0(\mathrm{\Gamma }_\pi +\mathrm{\Gamma }_K)},$$
(4)
with
$$\mathrm{\Gamma }_\pi =g_\pi \sqrt{m_{\pi \pi }^2/4m_\pi ^2}$$
(5)
and
$$\mathrm{\Gamma }_K=\frac{g_K}{2}\left(\sqrt{m_{\pi \pi }^2/4m_{K^+}^2}+\sqrt{m_{\pi \pi }^2/4m_{K^0}^2}\right).$$
(6)
We multiply each amplitude by a complex coefficient, $`c_j=a_je^{\delta _j}`$. The fit parameters are the magnitudes, $`a_j`$, and the phases, $`\delta _j`$, which accomodate the final state interactions, are fit parameters obtained using the maximum-likelihood method.
Monte Carlo simulations are used to determine the shape and location of the $`D^0K^{}\pi ^+`$ background in the Dalitz plot. The amount the $`D^0`$ background is determined using both MC simulations and data. The contribution of the $`D_s^+\eta ^{}\pi ^+`$ background is negligible. The background proportions are 12 $`\pm `$ 2% for $`D^0K^{}\pi ^+`$ and 88 $`\pm `$ 2% for the combinatoric across the Dalitz plot. Checks of the background model are described in the companion paper.
The parameters of the $`f_0(980)`$ state, $`g_\pi `$, $`g_K`$, and $`m_0`$, as well as the mass and width of the $`f_0(1370)`$, are determined directly from the data, floating them as free parameters in the fit. The other resonance masses and widths are taken from the PDG. The results of the $`D_s^+`$ Dalitz plot fit are shown in Table I. The column labeled Fit A corresponds to our best fit with all six modes. The measured values $`m_0=977\pm 3\pm 2`$ MeV/c<sup>2</sup>, $`g_\pi =`$ 0.09 $`\pm `$ 0.01 $`\pm `$ 0.01 and $`g_K=`$ 0.02 $`\pm `$ 0.04 $`\pm `$ 0.03 have been corrected for small shifts in these parameters due to the $`\pi ^{}\pi ^+`$ mass resolution in this region. This resolution is estimated to be approximately 9 MeV/$`c^2`$. We determine the shifts by folding a 9 MeV/$`c^2`$ resolution together with various $`g_K`$, $`g_\pi `$ combinations and fitting the resulting distributions to the form in Eqns. (4)-(6). The resolution was found to shift $`m_0`$, $`g_K`$, and $`g_\pi `$ by $`1.4`$ MeV, +0.06 and $`0.006`$ respectively. These shifts have been included in the values quoted above and in Table I. Uncertainties in our resolution contribute to the systematic uncertainties. The magnitudes and phases of the resonant amplitudes are relatively insensitive to the value of $`g_K`$. These values are not compatible with WA76 results, $`g_\pi =`$ 0.28 $`\pm `$ 0.04 and $`g_K=`$ 0.56 $`\pm `$ 0.18. For the $`f_0(1370)`$ we find $`m_0=`$$`1434\pm 18`$ MeV/c<sup>2</sup> and $`\mathrm{\Gamma }_0=173\pm 32`$ MeV/c<sup>2</sup>. We have also fit the Dalitz plot using for the $`f_0(980)`$ the same Breit-Wigner function as for the other resonances. The resulting fit is nearly as good as the one using the coupled-channel Breit-Wigner function, and the fractions and phases are indistinguishable. With this parameterization we find $`m_0=`$$`975\pm 3`$ MeV/c<sup>2</sup> and $`\mathrm{\Gamma }_0=44\pm 2\pm 2`$ MeV/c<sup>2</sup>, also shifted by the effect of mass resolution.
Table I shows the magnitudes ($`a_j`$) and phases ($`\delta _j`$) determined from the fit, and corresponding fraction for each decay mode. The fractions are defined as
$$f_j\frac{𝑑s_{12}𝑑s_{13}c_j𝒜_j^2}{𝑑s_{12}𝑑s_{13}_{jk}c_j𝒜_jc_k^{}𝒜_k^{}}.$$
(7)
The first reported error is statistical and the second is systematic, the latter being dominated by the uncertainties in the resonance parameters, in the background parameterization, and in the acceptance correction. The $`f_0(980)\pi ^+`$ is the dominant component, accounting for nearly half of the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decay width, followed by the $`f_0(1370)\pi ^+`$ and $`f_2(1270)\pi ^+`$ components. The contribution of $`\rho ^0(770)\pi ^+`$ and $`\rho ^0(1450)\pi ^+`$ components corresponds to about 10% of the $`\pi ^{}\pi ^+\pi ^+`$ width. We have not found a statistically significant non-resonant component. The $`s_{12}`$ and $`s_{13}`$ projections are nearly independent and the sum of the two is shown in Fig. 3 for Fit A.
To assess the quality of our fit absolutely, and to compare it with other possible fits, we developed a fast-MC algorithm that simulates the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ Dalitz plot from a given signal distribution, background, detector resolution and acceptance. For any given set of input parameters we calculated a $`\chi ^2`$ using the procedure presented in Ref. . From $`\chi ^2`$ and the number of degrees of freedom ($`\nu `$), we calculate a confidence level assuming a Gaussian distribution in $`\chi ^2/\nu `$. The confidence level for the agreement of the projection of Fit A onto the Dalitz plot with the data of Fit A is 35%.
We perform fits excluding amplitudes with small contributions. The fit without the non-resonant amplitude is as good as Fit A, and the resulting parameters are essentially the same. When comparing Fit A with models without the $`\rho ^0\pi ^+`$ amplitudes (Table 1) we calculate $`\mathrm{\Delta }w=2(\mathrm{ln}_i\mathrm{ln}_A)`$, where $`_A`$ is the likelihood of the fit with all modes, and $`_i`$ is the likelihood of the different models, as described in the companion paper. For the model where we exclude the $`\rho ^0(770)\pi ^+`$ amplitude we have, for data, $`\mathrm{\Delta }w=`$20. In the fast-MC with all modes we have $`\mathrm{\Delta }w=37`$. In the fast-MC with no $`\rho ^0(770)\pi ^+`$ we have $`\mathrm{\Delta }w=29`$. We observe a similar behavior in Fit C, where we exclude the $`\rho ^0(1450)\pi ^+`$ amplitude, we have $`\mathrm{\Delta }w=32`$. In the fast-MC with all modes we have $`\mathrm{\Delta }w=30`$ while with the fast-MC with no $`\rho ^0(1450)\pi ^+`$ $`\mathrm{\Delta }w=30`$. In all fast-MC experiments the rms deviation for $`\mathrm{\Delta }w`$ is about 11 units. We conclude that the best description of our data includes both $`\rho ^0(770)\pi ^+`$ and $`\rho ^0(1450)\pi ^+`$ amplitudes.
The contribution of modes having isoscalar mesons completely dominates the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decay. The same isoscalar dominance is observed in the $`D^+\pi ^{}\pi ^+\pi ^+`$ decay. However, there is no evidence in the $`D_s^+`$ decay for a low-mass broad scalar particle as seen in the $`D^+`$ decay. If the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decay is dominated by the Cabibbo-favored spectator mechanism, we would expect final states with a large $`s\overline{s}`$ content. Approximately half of the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ rate is produced via $`f_0(980)\pi ^+`$. The $`f_0(980)`$ is often supposed to have a large $`s\overline{s}`$ component, indicating a large spectator amplitude in this decay. On the other hand, the large contribution from the intermediate state $`f_0(1370)\pi ^+`$ indicates the presence of either $`W`$-annihilation amplitudes or strong rescattering in the final state. In fact this decay is not observed in the $`D_s^+K^+K^{}\pi ^+`$ final state, pointing to the $`f_0(1370)`$ being a non-$`s\overline{s}`$ particle, as suggested by the naive quark model.
In summary, Fermilab experiment E791 has measured the branching ratio of the decay $`D_s^+\pi ^{}\pi ^+\pi ^+`$ relative to $`D_s^+\varphi \pi ^+`$ to be $`0.245\pm 0.028_{0.012}^{+0.019}`$. We measure the mass and width of the $`f_0(980)`$ and $`f_0(1370)`$. Our results for the $`f_0(980)`$ parameters are $`g_\pi =`$ 0.09 $`\pm `$ 0.01 $`\pm `$ 0.01, $`g_K=`$0.02 $`\pm `$ 0.04 $`\pm `$ 0.03, and $`m_0=977\pm 3\pm 2`$ MeV/c<sup>2</sup> Using the same Breit-Wigner function for the $`f_0(980)`$ as for the other resonances, we find $`\mathrm{\Gamma }_0=44\pm 2\pm 2`$ MeV/c<sup>2</sup> and $`m_0=975\pm 3`$ MeV/c<sup>2</sup>. For the $`f_0(1370)`$ we find $`m_0=`$$`1434\pm 18\pm 9`$ MeV/c<sup>2</sup> and $`\mathrm{\Gamma }_0=173\pm 32\pm 6`$ MeV/c<sup>2</sup>. Finally, the fit of the Dalitz plot shows a dominant contribution from the $`f_0(980)\pi ^+`$, significant contributions from the $`f_0(1370)\pi ^+`$ and $`f_2(1270)\pi ^+`$, small contribution from the $`\rho ^0\pi ^+`$ channels, and a negligible non-resonant component. The isoscalar plus $`\pi ^+`$ components correspond to over 90% of the $`D_s^+\pi ^{}\pi ^+\pi ^+`$ decay width.
We gratefully acknowledge the assistance of the staffs of Fermilab and of all the participating institutions. This research was supported by the Brazilian Conselho Nacional de Desenvolvimento Científico e Tecnológico, CONACyT (Mexico), the Israeli Academy of Sciences and Humanities, the U.S. Department of Energy, the U.S.-Israel Binational Science Foundation, and the U.S. National Science Foundation. Fermilab is operated by the Universities Research Association, Inc., under contract with the United States Department of Energy.
|
warning/0007/hep-ph0007326.html
|
ar5iv
|
text
|
# Probe of CP Violation in 𝑒⁺𝑒⁻→𝑡𝑡̄ Near ThresholdTalk given at the Kiken Meeting: “New Perspectives in Elementary Particle Physics”, Kyoto, Japan, July 17 - 20, 2000.
## 1 Introduction
In this article we report our recent theoretical study on how to probe CP violation in the top quark sector at future $`e^+e^{}`$ linear colliders in the $`t\overline{t}`$ threshold region.
Recently studies of various properties of the top quark have been started at Tevatron. The detailed properties will be investigated further in future experiments at LHC and at future $`e^+e^{}`$ linear colliders. Among various interactions of the top quark, testing the CP-violating interactions is particularly interesting due to following reasons:
* Within the Standard Model (SM), CP-violation in the top quark sector is extremely small. \[The electric-dipole-moment (EDM) of a quark is induced first at three-loop level .\] If any CP-violating effect is detected in the top quark sector in a near-future experiment, it immediately signals new physics.
* There can be many sources of CP-violation in models that extend the SM, such as supersymmetric models, Leptoquark models, multi-Higgs-doublet models, Extra-dimensions, etc. Besides, the observed baryon asymmetry in the Universe suggests existence of CP violating mechanisms beyond the SM.
* In relatively wide class of models beyond the SM, CP violation emerges especially sizably in the top quark sector. A typical example is shown in Fig. 1.
Let us state the set-ups of our analysis. We consider CP-violating interactions of top quark with $`\gamma `$, $`Z`$, and $`g`$. In particular, we consider the lowest dimension CP-odd effective operators:
$`_{\text{CP}\text{-odd}}={\displaystyle \frac{ed_{t\gamma }}{2m_t}}(\overline{t}i\sigma ^{\mu \nu }\gamma _5t)_\mu A_\nu {\displaystyle \frac{g_Zd_{tZ}}{2m_t}}(\overline{t}i\sigma ^{\mu \nu }\gamma _5t)_\mu Z_\nu {\displaystyle \frac{g_sd_{tg}}{2m_t}}(\overline{t}i\sigma ^{\mu \nu }\gamma _5T^at)_\mu G_\nu ^a,`$
$`\sigma ^{\mu \nu }\frac{i}{2}[\gamma ^\mu ,\gamma ^\nu ],`$ (1)
where $`e=g_W\mathrm{sin}\theta _W`$ and $`g_Z=g_W/\mathrm{cos}\theta _W`$.
These represent the interactions of $`\gamma `$, $`Z`$, $`g`$ with the EDM, $`Z`$-EDM, chromo-EDM of top quark, respectively.<sup>*</sup><sup>*</sup>* The magnitudes of these EDMs are given by $`ed_{t\gamma }/m_t`$$`g_Zd_{tZ}/m_t`$$`g_sd_{tg}/m_t`$, respectively. $`d_{t\gamma }=1`$ corresponds to $`e/m_t10^{16}e\mathrm{cm}`$, etc. Each of the interactions has $`C=+1`$ and $`P=1`$. As stated, the SM contributions to these couplings are extremely small, $`d_{t\gamma }^{(\mathrm{SM})},d_{tZ}^{(\mathrm{SM})},d_{tg}^{(\mathrm{SM})}10^{14}`$. Since we will not be able to detect them in near-future collider experiments, we neglect the SM contributions below. Our concern is in the anomalous couplings which are induced from some new physics. We assume that generally the couplings $`d_{t\gamma }`$, $`d_{tZ}`$, $`d_{tg}`$ are complex where their imaginary parts may be induced from some absorptive processes.
Many of the readers would be interested in the sensitivities to these couplings expected in future experiments. In Table 1 we summarize the results of the sensitivity studies performed so far, including the results of our present study.
We may compare the sensitivities of experiments in the $`t\overline{t}`$ threshold region at $`e^+e^{}`$ colliders with others. The sensitivities to $`d_{t\gamma }`$ and $`d_{tZ}`$ are comparable to those attainable in the open-top region at $`e^+e^{}`$ colliders. The sensitivity to $`d_{tg}`$ is worse than that expected at a hadron collider but exceeds the sensitivity in the open-top region at $`e^+e^{}`$ colliders.
From this comparison one may find that our study at $`t\overline{t}`$ threshold has little impact on the study of CP violation and is not very interesting. The present author, however, has a slightly different physics interest personally. Although admittedly it is better to have higher sensitivities to the anomalous couplings $`d_{tg}`$, $`d_{t\gamma }`$, $`d_{tZ}`$, at the moment we do not know the sizes of these couplings. Therefore, I am interested more in the following questions than merely in achievable sensitivities: When any of the couplings happens to be sizable enough to be detected in some experiment, through what intriguing phonomena can we detect the anomalous couplings? And how can we extract as much information on the couplings as possible? In these respects, the $`t\overline{t}`$ threshold region has fairly rich physics contents. I would like to describe our investigations from this viewpoint below. So, please imagine a situation where any of the couplings happens to be of order 10% or larger and see what we can learn in that case.
## 2 $`t\overline{t}`$ Threshold
### 2.1 Unique aspects
When studying CP violation of the top quark, unique aspects of the $`t\overline{t}`$ threshold region are:
* The QCD interaction is enhanced in this region, hence the cross section is sensitive to the top-gluon ($`tg`$) couplings. We can study anomalous $`tg`$ couplings in a clean environment in comparison to hadron colliders.
* In certain models (e.g. those in which a neutral Higgs boson is exchanged between $`t`$ and $`\overline{t}`$ ), induced top quark EDM and $`Z`$-EDM are enhanced near the $`t\overline{t}`$ threshold.
* Since top quarks are produced almost at rest, one can reconstruct the spin information of top quarks from distributions of their decay products without solving detailed kinematics.
### 2.2 Time evolution of the $`t\overline{t}`$ system
Let us first review the time evolution of $`t`$ and $`\overline{t}`$, pair-created in $`e^+e^{}`$ collision just below threshold, within the SM (Fig. 2).
They are created close to each other at a relative distance $`r1/m_t`$ and then spread apart non-relativistically. When their relative distance becomes of the order of the Bohr radius, $`r(\alpha _sm_t)^1`$, they start to form a Coulombic boundstate. When the relative distance becomes $`r(m_t\mathrm{\Gamma }_t)^{1/2}`$, where $`\mathrm{\Gamma }_t`$ is the decay width of top quark, either $`t`$ or $`\overline{t}`$ decays via electroweak interaction, and accordingly the boundstate decays. Numerically these two scales have similar magnitudes, $`(\alpha _sm_t)^1(m_t\mathrm{\Gamma }_t)^{1/2}`$, and are much smaller than the hadronization scale $`\mathrm{\Lambda }_{\mathrm{QCD}}^1`$. Since gluons which have wavelengths much longer than the size of the $`t\overline{t}`$ system cannot couple to this color singlet system, the strong interaction participating in the formation of the boundstate is dictated by the perturbative domain of QCD. Due to this reason, we are able to compute the amplitude from the first principles with order 5% accuracy or better, even though the QCD boundstates are involved. At the moment only one exception is the normalization of the total $`t\overline{t}`$ cross section, where we still have 10–15% theoretical uncertainty. The spin and $`PC`$ of the dominantly produced boundstate are $`J^{PC}=1^{}`$. Inside this boundstate: $`t`$ and $`\overline{t}`$ are in the $`S`$-wave state ($`L=0`$); the spins of $`t`$ and $`\overline{t}`$ are aligned to each other and pointing to $`e^{}`$ beam direction $`|`$ or to $`e^+`$ beam direction $`|`$ or they are in a linear combination of the two states ($`S=1`$).
Now let us consider effects of the anomalous interactions eq. (1) on the time evolution of the $`t\overline{t}`$ system (Fig. 3(a)).
CP-violation originating from the $`t\gamma `$ or $`tZ`$ coupling occurs at the stage of the pair creation, i.e. when $`t`$ and $`\overline{t}`$ are very close to each other. The generated boundstate has $`J^{PC}=1^+`$, so $`t`$ and $`\overline{t}`$ are in the $`P`$-wave ($`L=1`$) and spin-0 state $`||`$. On the other hand, CP-violation originating from the $`tg`$ coupling takes place after the boundstate formation when multiple gluons are exchanged between $`t`$ and $`\overline{t}`$, i.e. when $`t`$ and $`\overline{t}`$ are separated at a distance of the Bohr radius. The anomalous top-gluon coupling generates effectively a spin-dependent potential between $`t`$ and $`\overline{t}`$
$`V_{\text{CP}\text{-odd}}={\displaystyle \frac{d_{tg}}{m_t}}(𝐬_t\overline{𝐬}_t)V_\mathrm{C}(r).`$ (2)
Here, $`𝐬_t`$ and $`\overline{𝐬}_t`$ denote the spins of non-relativistic $`t`$ and $`\overline{t}`$, respectively; $`V_\mathrm{C}(r)=C_F\alpha _s/r`$ is the Coulomb potential with the color factor $`C_F=4/3`$. When $`d_{tg}>0`$, the potential $`V_{\text{CP}\text{-odd}}`$ tends to align both chromo-EDMs in the direction of chromo-electric field, or, align $`\overline{𝐬}_t`$ in the direction of $`𝐫=𝐫_t\overline{𝐫}_t`$ and $`𝐬_t`$ in the direction of $`𝐫`$; see Fig. 3(b). Therefore, first the boundstate is formed in $`J^{PC}=1^{}`$ ($`L=0`$ and $`S=1`$) state and after interacting via the potential $`V_{\text{CP}\text{-odd}}`$ it turns into $`J^{PC}=1^+`$ ($`L=1`$ and $`S=0`$) state, i.e. the $`t`$ and $`\overline{t}`$ spins are aligned into antiparallel directions.
We can disentangle the effects of the three couplings, $`d_{t\gamma }`$, $`d_{tZ}`$, $`d_{tg}`$, on the amplitude using the differences in the dependences on the energy and $`e^\pm `$ polarization in the threshold region. Firstly when the c.m. energy is raised the CP-violating effects due to $`d_{t\gamma }`$ and $`d_{tZ}`$ increase proportionally to the velocity of the top quark, since these effects are induced directly by the dimension-five operators. On the other hand, the effect of $`d_{tg}`$ does not increase so rapidly. The enhancement of the $`tg`$-coupling due to multiple exchanges of gluons will be lost when the energy is raised and $`t`$ and $`\overline{t}`$ spread apart quickly without forming boundstates. Secondly, one may vary the relative weight of the photon-induced CP-violating effect and the $`Z`$-induced effect by varying the $`e^\pm `$ longitudinal polarization. This is because $`e_L`$ and $`e_R`$ couple differently to $`\gamma `$ and $`Z`$, and the relative weight of virtual $`\gamma `$ and $`Z`$ changes.
### 2.3 CP-odd observables
Which CP-odd observables are sensitive to the CP-violating couplings $`d_{t\gamma }`$, $`d_{tZ}`$, $`d_{tg}`$? For the process $`e^+e^{}t\overline{t}`$, we may conceive of following expectation values of kinematical variables for CP-odd observables:
$`(𝐩_e\overline{𝐩}_e)(𝐬_t\overline{𝐬}_t),`$
$`(𝐩_t\overline{𝐩}_t)(𝐬_t\overline{𝐬}_t),`$ (3)
$`[(𝐩_e\overline{𝐩}_e)\times (𝐩_t\overline{𝐩}_t)](𝐬_t\overline{𝐬}_t),`$
where the spins and momenta are defined in the c.m. frame. (The initial state is CP-even if we assume the SM interactions of $`e^\pm `$ with $`\gamma `$ and $`Z`$.) The above quantities are the three components of the difference of the $`t`$ and $`\overline{t}`$ spins. One may easily confirm the CP transformations of the above observables: e.g. $`(𝐩_t\overline{𝐩}_t)\stackrel{C}{}(\overline{𝐩}_t𝐩_t)\stackrel{P}{}(\overline{𝐩}_t+𝐩_t)`$. One might say (in a somewhat oversimplified way) that in the SM the $`t`$ and $`\overline{t}`$ spins are parallel to each other, so the SM contributions drop in the difference $`𝐬_t\overline{𝐬}_t`$, whereas the $`t`$ and $`\overline{t}`$ spins become antiparallel to each other by the effects of $`_{\text{CP}\text{-odd}}`$, so they remain in $`𝐬_t\overline{𝐬}_t`$. Thus, we want to measure the difference of the spins of $`t`$ and $`\overline{t}`$. It is equivalent to measuring the difference of the polarization vectors of $`t`$ and $`\overline{t}`$. All other CP-odd observables for $`e^+e^{}t\overline{t}`$ are bilinear in $`𝐬_t`$ and $`\overline{𝐬}_t`$. Since analyses of spin correlations are complicated, we focus on the difference of the polarization vectors.
Practically we can measure the $`t`$ and $`\overline{t}`$ polarization vectors efficiently using $`\mathrm{}^\pm `$ angular distributions. It is known that the angular distribution of the charged lepton $`\mathrm{}^+`$ from the decay of top quark is maximally sensitive to the top quark polarization vector. In the rest frame of top quark, the $`\mathrm{}^+`$ angular distribution is given by
$`{\displaystyle \frac{1}{\mathrm{\Gamma }_t}}{\displaystyle \frac{d\mathrm{\Gamma }(tb\mathrm{}^+\nu )}{d\mathrm{cos}\theta _\mathrm{}^+}}={\displaystyle \frac{1+P\mathrm{cos}\theta _\mathrm{}^+}{2}}`$ (4)
at tree level, where $`P`$ is the top quark polarization and $`\theta _\mathrm{}^+`$ is the angle of $`\mathrm{}^+`$ measured from the direction of the top quark polarization vector. Indeed the $`\mathrm{}^+`$ distribution is ideal for extracting CP-violation in the $`t\overline{t}`$ production process; the above angular distribution is unchanged even if anomalous interactions are included in the $`tbW`$ decay vertex, up to the terms linear in the decay anomalous couplings and within the approximation $`m_b=0`$ . Therefore, if we consider the average of the lepton direction, for instance, we may extract the top quark polarization vector efficiently:
$`𝐧𝐧_{\mathrm{}}_{\mathrm{Lab}}{\displaystyle \frac{1}{3}}𝐧𝐏.`$ (5)
The average is to be taken at the top quark rest frame, but in the threshold region, we may take the average in the laboratory frame barely without loss of sensitivities to the anomalous couplings.
### 2.4 $`t`$ and $`\overline{t}`$ polarization vectors
The polarization vectors of $`t`$ and $`\overline{t}`$ are defined from the production cross section of a $`t\overline{t}`$ pair in the threshold region. The cross section, where ($`t`$,$`\overline{t}`$) have momenta ($`𝐩_t`$,$`𝐩_t`$) and the spins $`+\frac{1}{2}`$ along the quantization axes ($`𝐬_t`$,$`\overline{𝐬}_t`$) in the c.m. frame, is given by
$`{\displaystyle \frac{d\sigma (𝐬_t,\overline{𝐬}_t)}{d^3𝐩_t}}={\displaystyle \frac{d\sigma }{d^3𝐩_t}}{\displaystyle \frac{1+𝐏𝐬_t+\overline{𝐏}\overline{𝐬}_t+(𝐬_t)_i(\overline{𝐬}_t)_j𝐐_{ij}}{4}}.`$ (6)
Here, $`|𝐬_t|=|\overline{𝐬}_t|=1`$. On the right-hand-side, $`d\sigma /d^3𝐩_t`$ represents the production cross section when the spins of $`t`$ and $`\overline{t}`$ are summed over. $`𝐏`$ and $`\overline{𝐏}`$ denote, respectively, the polarization vectors of $`t`$ and $`\overline{t}`$.
According to the above definition we computed the polarization vectors. The SM contributions to the polarization vectors are same for $`t`$ and $`\overline{t}`$, while the contributions from $`_{\text{CP}\text{-odd}}`$ are opposite in sign:
$`𝐏=𝐏_{\mathrm{SM}}+\delta 𝐏,\overline{𝐏}=𝐏_{\mathrm{SM}}\delta 𝐏.`$ (7)
It is convenient to express $`\delta 𝐏`$ in components:
$`\delta 𝐏=\delta \mathrm{P}_{}𝐧_{}+\delta \mathrm{P}_{}𝐧_{}+\delta \mathrm{P}_\mathrm{N}𝐧_\mathrm{N},`$ (8)
where the orthonormal basis is defined from the $`e^{}`$ beam direction and the top quark momentum direction as
$`𝐧_{}={\displaystyle \frac{𝐩_e}{|𝐩_e|}},𝐧_\mathrm{N}={\displaystyle \frac{𝐩_e\times 𝐩_t}{|𝐩_e\times 𝐩_t|}},𝐧_{}=𝐧_\mathrm{N}\times 𝐧_{}.`$ (9)
Then the CP-odd contributions are given by
$`\delta \mathrm{P}_{}=0,`$ (10)
$`\delta \mathrm{P}_{}=\mathrm{Im}\left[d_{tg}B_{}^g\left({\displaystyle \frac{D}{G}}\right)+d_{t\gamma }B_{}^\gamma \left({\displaystyle \frac{F}{G}}\right)+d_{tZ}B_{}^Z\left({\displaystyle \frac{F}{G}}\right)\right]\left({\displaystyle \frac{p_t}{m_t}}\right)\mathrm{sin}\theta _t,`$ (11)
$`\delta \mathrm{P}_\mathrm{N}=\mathrm{Re}\left[d_{tg}B_\mathrm{N}^g\left({\displaystyle \frac{D}{G}}\right)+d_{t\gamma }B_\mathrm{N}^\gamma \left({\displaystyle \frac{F}{G}}\right)+d_{tZ}B_\mathrm{N}^Z\left({\displaystyle \frac{F}{G}}\right)\right]\left({\displaystyle \frac{p_t}{m_t}}\right)\mathrm{sin}\theta _t.`$ (12)
Here, $`B_{}^X`$ and $`B_\mathrm{N}^X`$ denote combinations of the electroweak couplings of $`e^{}`$ and $`t`$ as well as of $`e^\pm `$ beam polarization. $`D`$, $`F`$ and $`G`$ denote the QCD Green functions which incorporate the boundstate effects.
In Fig. 4 we examine the electroweak coefficients $`B_i^X`$’s.
They are given as a function of the polarization parameter of the initial $`e^\pm `$ beams
$`\chi ={\displaystyle \frac{P_{e^+}P_e^{}}{1P_{e^+}P_e^{}}}.`$ (13)
If the positron beam is unpolarized ($`P_{e^+}=0`$), $`\chi =P_e^{}`$. Typical sizes of all the coefficients $`B_i^X`$’s are order one. We also see that their dependences on the beam polarizations are different.
Next we examine the QCD factor in Fig. 5.
The ratios of the Green functions together with the top quark velocity ($`\beta =p_t/m_t`$), $`\beta D/G`$ and $`\beta F/G`$, are shown on a complex plane. These are plotted as a function of the energy $`E=\sqrt{s}2m_t`$ alone by choosing the top momentum to be the typical momentum at a fixed energy. As we raise the energy, the magnitude of the QCD factor associated with the $`t\gamma `$ and $`tZ`$ couplings, $`|\beta F/G|`$, increase proportionally to $`\beta `$. On the other hand, the magnitude of the QCD factor associated with the $`tg`$ coupling, $`|\beta D/G|`$, does not change very much. We see that the size of $`|\beta F/G|`$ is 5–20% while the size of $`|\beta D/G|`$ is 5–10%. Also it can be seen that the strong phases are quite sizable and change rapidly with energy. It is the characteristics of the $`t\overline{t}`$ boundstates that these QCD factors can be computed reliably from the first principles.
## 3 Conclusions
In this work we studied how to probe the anomalous CP-violating couplings of the top quark with $`\gamma `$, $`Z`$ and $`g`$ in the $`t\overline{t}`$ threshold region at future $`e^+e^{}`$ colliders. The sensitivities to the anomalous couplings are given and compared with other future experiments in Table 1. Qualitatively, the characteristics of the $`t\overline{t}`$ region are summarized as follows.
1. We can measure the three couplings $`d_{t\gamma }`$, $`d_{tZ}`$, $`d_{tg}`$ simultaneously and we can disentangle each contribution.
2. We can measure the complex phases of the couplings $`d_{t\gamma }`$, $`d_{tZ}`$, $`d_{tg}`$. Since the strong phases can be modulated at our disposal, a single observable ($`\delta \mathrm{P}_{}`$ or $`\delta \mathrm{P}_\mathrm{N}`$) probes the phases of the couplings.
3. Typical sizes of components of the difference of the $`t`$ and $`\overline{t}`$ polarizations are given by
$$|\delta \mathrm{P}_{}|,|\delta \mathrm{P}_\mathrm{N}|(\text{5–20}\%)\times (d_{t\gamma },d_{tZ},d_{tg}).$$
They can be extracted efficiently from the directions of charged leptons from decays of $`t`$ and $`\overline{t}`$:
$$𝐧(𝐧_{\mathrm{}}+\overline{𝐧}_{\mathrm{}})_{𝐩_t}\frac{2}{3}𝐧\delta 𝐏.$$
We note that if one of the couplings is detected in the future, we would certainly want to measure the others in order to gain deeper understanding of the CP-violating mechanism. This is because one may readily think of various underlying processes which give different contributions to the individual couplings; see Fig. 6.
Regarding (1) and (2) above, QCD interaction is used as a controllable tool for the detection of the anomalous couplings. This would be the first trial to use QCD interaction for such a purpose without requiring any phenomenological inputs.
## Acknowledgements
This work is based on a collaboration with T. Nagano and M. Jeżabek. The author is grateful to the hospitality at the Summer Institute ’99 (August 1999, Yamanashi, Japan) where this work was initiated. Also he thanks K. Fujii, Z. Hioki, K. Ikematsu, T. Takahashi, J.H. Kühn, S. Rindani, M. Tanabashi and M. Yamaguchi for valuable discussions. This work was supported in part by the Japan-German Cooperative Science Promotion Program.
|
warning/0007/nucl-th0007013.html
|
ar5iv
|
text
|
# Spin structure of many-body systems with two-body random interactions
## I Introduction
Random matrix theory has been widely used to model the behavior of complex quantum systems . Of particular importance are two-body random matrix ensembles (TBRE), which were introduced into nuclear physics three decades ago and recently have received a considerable amount of renewed interest in the study of mesoscopic systems and nuclei . This is partly due to the observation of regular structures in the low-lying states of various nuclear models with random two-body interactions. As examples we mention the observed ground state dominance of $`J^P=0^+`$ states, the occurrence of pairing (energy) gaps in random shell models , and the existence of vibrational and rotational bands in random interacting boson models . While these observations concern the low-lying levels of many-body systems, similar regularities also have been reported in the wave function structure. These include signatures of phonon collectivity relating states of different angular momentum , strong pair-transfer amplitudes between ground states of nuclei with different mass numbers , strong electromagnetic transitions in bands of random interacting boson models , deviations from random matrix theory in transition strength distributions for embedded Gaussian ensembles , and regularities in wave functions of spin systems with random interactions . These observations indicate that some typical properties of interacting many-body systems are determined simply by the presence of a rotationally invariant two-body force that allows for transitions between single-particle Fock space states.
In this article we aim at an understanding of some of the reported regularities. We are particularly interested in the quantum numbers of the ground state and in correlations between energies and wave functions of low-lying states. For this purpose we consider a system of $`N`$ spin-$`\frac{1}{2}`$ fermions on $`M`$ orbitals, subject to a spin-conserving random two-body interaction. This system is simpler than the nuclear models studied so far, and allows us to derive analytical results. It may be seen as a model for mesoscopic systems like quantum dots.
This article is divided as follows. In the following Section we focus on the width and shape of spectral densities in sectors of definite spin for two models of spin systems with random interactions, and derive analytical results. Special emphasis is placed on understanding the spin structure of the spectrum close to the ground state. Our analysis extends and complements previous work in nuclei and mesoscopic systems . In Section III we investigate correlations between low-lying states, and consider pairing properties. Finally we give a summary.
## II Spin Structure
### A The model
We begin with a very simple model of random spin-conserving two-body interactions in a many-body system :
$$H=C_0\underset{\alpha ,\alpha ^{}\mathrm{spin0}}{}H_{\alpha \alpha ^{}}A_\alpha ^{}A_\alpha ^{}+C_1\underset{\beta ,\beta ^{}\mathrm{spin1}}{}\stackrel{~}{H}_{\beta \beta ^{}}A_\beta ^{}A_\beta ^{},$$
(1)
where $`\alpha `$ and $`\alpha ^{}`$ each denotes a pair of two spin$`\frac{1}{2}`$ fermions in a $`J=0`$ state, while $`\beta `$ and $`\beta ^{}`$ each denotes a pair of fermions in a $`J=1`$ state. The two-body states can of course be labeled by the individual orbitals occupied by the two particles. Thus, for a system with $`M`$ single-particle orbitals, we have $`M(M1)/2`$ pairs with total spin $`J=1`$, enumerated as $`A_{\beta _{ij}}=(a_ia_j+a_ia_j)/\sqrt{2}`$, where $`a_j`$ annihilates a single particle state, and the single-particle indices are $`1i<jM`$. \[Here we are enumerating two-body $`J=1`$ states with $`J_z=0`$ only; obviously there are equal numbers of $`J_z=\pm 1`$ two-body states which by rotational symmetry must couple in precisely the same way.\] Similarly, there are $`M(M1)/2`$ pairs with total spin $`J=0`$, having the form $`A_{\alpha _{ij}}=(a_ia_ja_ia_j)/\sqrt{2}`$, and an additional $`M`$ pairs where a single site is doubly occupied: $`A_{\alpha _{ii}}=a_ia_i`$. We will assume throughout that the two kinds of $`J=0`$ states interact equally strongly among themselves and with each other; this assumption does not substantially affect any of the results to be presented in this paper. The effect of diagonal spin-spin interactions in a spin system with random wave functions has been investigated in ref. .
Within each two-body spin sector, then, the couplings $`H_{\alpha \alpha ^{}}`$ and $`\stackrel{~}{H}_{\beta \beta ^{}}`$ are assumed to be governed by a GOE random-matrix ensemble,
$`H_{\alpha \alpha ^{}}^2`$ $`=`$ $`1+\delta _{\alpha \alpha ^{}}`$ (2)
$`\stackrel{~}{H}_{\beta \beta ^{}}^2`$ $`=`$ $`1+\delta _{\beta \beta ^{}},`$ (3)
where $`\mathrm{}`$ indicates an ensemble average. We note that the two matrices $`H_{\alpha \alpha ^{}}`$ and $`\stackrel{~}{H}_{\beta \beta ^{}}`$ are each taken to be real and symmetric, consistent with time reversal invariance; the model can easily be generalized to the case where an external magnetic field is present, so that the matrices must be taken from a GUE rather than a GOE ensemble, or from an intermediate ensemble . Finally, we notice that in the model (Eq. 1), couplings in different two-body spin sectors are taken to be completely independent of one another; in Section II J we will discuss an alternative random-coupling model where this is not the case. In Section III we will see what effects on spin spectra may be obtained by adding a one-body term to the two-body random Hamiltonian. Note that the presence of a strong one-body term enhances the likelyhood of finding a spin-0 ground state, due to spin-degeneracy of the one-body spectrum.
### B Preliminary results
We may now consider numerically the spectra of $`N`$fermion systems interacting via the random two-body interaction of the form (Eq. 1). Due to an approximate particle-hole symmetry of this system, maximum density is obtained near half-filling, $`N=M`$, and we may restrict ourselves to the parameter range $`2NM`$. The preponderance of $`S=0`$ many-body ground states, which has been observed previously in shell model calculations, is seen to be an extremely strong effect in this simple random interaction model. For $`N=4`$ fermions on $`M=4`$ sites, with an exclusively spin-$`0`$ interaction ($`C_0=1`$, $`C_1=0`$), we find for an ensemble of $`2000`$ systems that the ground state has total spin $`S=0`$ in $`99.8\%`$ of the cases. For an exclusively spin-$`1`$ interaction ($`C_1=1`$, $`C_0=0`$), the ground state is observed to have total spin $`S=0`$ for $`99.75\%`$ of the systems in the ensemble (a result which is statistically indistinguishable from the previous case). This is despite the fact that in this system $`S=0`$ many-body states comprise only about $`56\%`$ of the total number of states. For $`N=4`$ particles on $`M=5`$ sites, all $`2000`$ computed ground states were observed to have total spin $`S=0`$, independent of whether a pure $`J=0`$ or $`J=1`$ two-body coupling was used. Very similar results are obtained at least until a system size of $`M=12`$, beyond which point the size of the many-body basis makes the diagonalization of a large ensemble of Hamiltonians prohibitively expensive. For $`N=6`$ particles on $`M=6,`$ $`7,`$ and $`8`$ sites, we once again observe an almost $`100\%`$ prevalence of $`S=0`$ ground states.
How may we understand these initially surprising findings within the context of a completely random interaction model? We begin by considering the shape of the many-body density of states $`\rho _S(E)`$ in sectors of total many-body spin $`S`$. It is known that for $`N2`$ particles with a two-body interaction, the moments of the spectral density move away from those associated with a semicircle law, and approach those corresponding to a Gaussian shape . Then, the ensemble-averaged spectra $`\rho _S(E)`$, being centered at $`E=0`$ because of the symmetry of the Hamiltonian, must be described simply by their squared widths, $`\mathrm{Tr}_SH^2`$ (where $`\mathrm{Tr}_S`$ denotes throughout the sum over all basis states in the sector of total spin $`S`$, divided by the number of basis states in that sector). Our naive expectation (assuming that the spectrum can be described by a Gaussian shape well into its tail) is that ground state behavior should be associated with the relative widths of the different spin spectra, a preponderance of $`S=0`$ ground states arising from a large value of $`\mathrm{Tr}_0H^2`$ as compared to widths in the other sectors . Indeed, fixing the number of particles at $`N=4`$, our numerical results show that for $`4M10`$ orbitals the relation $`\mathrm{Tr}_0H^2>\mathrm{Tr}_1H^2>\mathrm{Tr}_2H^2`$ is almost always observed (independent of the relative values of $`C_0`$ and $`C_1`$), consistent with $`S=0`$ ground state dominance. For $`M=11`$ and $`12`$, however, when taking an exclusively $`J=1`$ coupling, this pattern is reversed, with the widths of the higher-spin spectra beginning to exceed those of the $`S=0`$ spectrum. This shows that the width analysis alone is not sufficient to explain numerical results for the ground state behavior. However, these findings do suggest than an interesting transition may be taking place as we approach the dilute regime, which we shall now proceed to understand analytically.
### C Width analysis
The many-body basis states of the system may be classified by the total spin $`S`$ and the number $`D`$ of orbitals that are doubly occupied, $`0DN/2`$. We easily see that in the dilute limit ($`M\mathrm{}`$ for fixed $`N`$), $`D=0`$ states will dominate: a typical $`S=0`$ or $`S=1`$ basis state in this limit is given for even $`N`$ by
$`|\mathrm{\Psi }`$ $`=`$ $`{\displaystyle \frac{1}{2^{N/2}}}\left(a_i^{}a_j^{}\pm a_i^{}a_j^{}\right)`$ (4)
$`\times `$ $`{\displaystyle \underset{z=1}{\overset{N/21}{}}}\left(a_{k_z}^{}a_{l_z}^{}a_{k_z}^{}a_{l_z}^{}\right)|0,`$ (5)
where the plus (minus) sign produces an $`S=1`$ ($`S=0`$) $`N`$body state, the single-particle orbitals $`i`$, $`j`$, $`k_1`$, $`l_1`$, …, $`k_{N/21}`$, $`l_{N/21}`$ are all taken to be distinct, and $`|0`$ denotes the vacuum. However, at finite densities we must consider basis states in subsectors of different $`D`$ for any given value of $`S`$, as these might in general have spectra of different shape and width. It is a straightforward, though somewhat tedious, counting problem to compute the ensemble-averaged squared energy width $`\mathrm{\Psi }|H^2|\mathrm{\Psi }`$, for a typical basis state $`|\mathrm{\Psi }`$ of given quantum numbers $`S`$ and $`D`$, when acted on by the Hamiltonian (Eq. 1). For each two-particle pair $`\alpha ^{}`$ or $`\beta ^{}`$ of the appropriate spin that may be annihilated, one simply has to count contributions from all pairs $`\alpha `$ or $`\beta `$ that may be created in its place, while keeping track of terms that must be added up coherently. So, for example, for a pure $`J=0`$ coupling ($`C_0=1`$, $`C_1=0`$), one obtains
$`\mathrm{Tr}_0H^2`$ $`=`$ $`{\displaystyle \frac{N^2(2MN)^2}{64}}+{\displaystyle \frac{2M^2N+MN^2N^3}{16}}`$ (6)
$`+`$ $`{\displaystyle \frac{6MN+N^2+2N^2\left[\left(\frac{D}{N/2}\right)^2\left(\frac{D}{N/2}\right)\right]}{16}}+{\displaystyle \frac{N}{2}}`$ (7)
$`\mathrm{Tr}_1H^2`$ $`=`$ $`\mathrm{Tr}_0H^2{\displaystyle \frac{2M^22MN+N^2}{4}}`$ (8)
$``$ $`{\displaystyle \frac{3MN}{2}}1.`$ (9)
In Eq. 6, terms are grouped and ordered by decreasing power in the number of orbitals $`M`$ assuming finite density ($`NM`$). In particular, we notice that the leading $`O(M^4)`$ term is invariant under particle-hole symmetry, $`N2MN`$. This symmetry is broken at $`O(1/M)`$ because of new terms that arise in the Hamiltonian (Eq. 1) when one commutes annihilation operators $`A`$ and creation operators $`A^{}`$ corresponding to pairs having at least one orbital in common. We notice also that the effect of double degeneracy ($`D>0`$) is initially to increase the width slightly, and then restore it to the original value when $`D`$ reaches its maximal value of $`N/2`$. However, the effect of double degeneracy appears only at $`O(1/M^2)`$, which we will see below to be small compared with the level spacing near the ground state, in the many-body limit. Thus, even for a non-dilute system ($`NM`$), we are justified in studying the spectra of $`D=0`$ states and taking their widths as proxies for the entire spectrum of that spin, even for the purpose of determining behavior near the edge of the spectrum (where the level spacings are largest). In Section III A we will consider the possible relevance of $`D>0`$ contributions to wave function structure near the ground state. We find there that basis states with high (or low) double occupancy have almost equal overlaps with eigenstates of all energies, and have no special tendency to contribute to the low-lying part of the spectrum. Thus it appears that, to leading order, double occupancy effects are not important to understanding the spin structure of the spectrum for a pure two-body interaction.
### D The four-body system
The width expression of Eq. 6 is rather unwieldy, as it consists of $`14`$ monomial terms, and so we begin our detailed analysis with the special case of $`N=4`$ particles; as we shall see below, all the interesting qualitative behavior of the many-body system is already present in this simple case. In the $`S=0`$ sector we have $`2\left(\genfrac{}{}{0pt}{}{M}{4}\right)`$ independent basis states of type $`D=0`$, having the form
$$\frac{1}{2}\left(a_i^{}a_j^{}a_i^{}a_j^{}\right)\left(a_k^{}a_l^{}a_k^{}a_l^{}\right)|0,$$
(10)
(where $`i`$, $`j`$, $`k`$, and $`l`$ all represent different orbitals), $`M\left(\genfrac{}{}{0pt}{}{M1}{2}\right)`$ basis states of type $`D=1`$, and $`\left(\genfrac{}{}{0pt}{}{M}{2}\right)`$ basis states of type $`D=2`$. Similarly, in the $`S=1`$ sector we have $`3\left(\genfrac{}{}{0pt}{}{M}{4}\right)`$ independent basis states of type $`D=0`$, of the form
$$\frac{1}{2}\left(a_i^{}a_j^{}+a_i^{}a_j^{}\right)\left(a_k^{}a_l^{}a_k^{}a_l^{}\right)|0,$$
(11)
and $`M\left(\genfrac{}{}{0pt}{}{M1}{2}\right)`$ basis states of type $`D=1`$. Finally, in the $`S=2`$ sector we only have the $`\left(\genfrac{}{}{0pt}{}{M}{4}\right)`$ basis states of type $`D=0`$. As the system becomes dilute ($`M\mathrm{}`$), type $`D=0`$ states always dominate, and the total numbers of states in the spin $`0`$, $`1`$, and $`2`$ sectors approach the ratio $`2:3:1`$. \[As we mentioned above, even for finite $`M`$, the inclusion of $`D>0`$ states does not significantly affect the behavior of the spin spectra, leading to corrections that are small compared with the level spacing.\]
What then causes the preponderance of $`S=0`$ ground states in this system? Using $`D=0`$ basis states, one obtains the following expressions for the widths (of course the $`C_0^2`$ part of these results is merely a special case of Eq. 6):
$`\mathrm{Tr}_0H^2`$ $`=`$ $`C_0^2\left({\displaystyle \frac{3}{2}}M^2{\displaystyle \frac{3}{2}}M+3\right)+C_1^2\left({\displaystyle \frac{3}{2}}M^2+{\displaystyle \frac{3}{2}}M+3\right)`$ (12)
$`\mathrm{Tr}_1H^2`$ $`=`$ $`C_0^2\left(M^2M\right)+C_1^2\left(2M^24M+4\right).`$ (13)
The leading $`O(M^2)`$ terms in the above expressions are easiest to understand. We simply need to take the number of spin-$`J`$ pairs in a four-body state of total spin $`S`$ (Eq. 10 or Eq. 11), and multiply by the number of orbital pairs to which the two particles may jump ($`M^2/2`$). So, for example, in a state of total spin $`S=0`$ (Eq. 10), the $`ij`$ and $`kl`$ pairs are in pure $`J=0`$ combinations, while the remaining $`4`$ pairs are uncorrelated, i.e. $`1/4`$ of them have $`J=0`$ and $`3/4`$ have $`J=1`$. This leads to a total of $`3`$ pairs each of $`J=0`$ and $`J=1`$, explaining the two $`\frac{3}{2}M^2`$ terms in the first line of Eq. 13. In a state of total spin $`S=1`$ (Eq. 11), the only difference is that the $`ij`$ pair is now in a $`J=1`$ combination, so there are only $`2`$ pairs with $`J=0`$ and $`4`$ pairs with $`J=1`$, leading to the $`C_0^2M^2`$ and $`2C_1^2M^2`$ terms in the second line of Eq. 13. We see already from this very simple argument that, due to a relatively larger number of $`J=0`$ pairs in a spin-$`0`$ state and a relatively larger number of $`J=1`$ pairs in a spin-$`1`$ state, in the $`M\mathrm{}`$ limit we always have $`\mathrm{Tr}_0H^2>\mathrm{Tr}_1H^2`$ if $`C_0>C_1`$ and $`\mathrm{Tr}_0H^2<\mathrm{Tr}_1H^2`$ if $`C_0<C_1`$. The ratio of the widths in the dilute limit approaches a constant,
$$\frac{\mathrm{Tr}_0H^2}{\mathrm{Tr}_1H^2}\frac{3C_0^2+3C_1^2}{2C_0^2+4C_1^2},$$
(14)
while the ratio of the many-body level spacing near the ground state, $`\rho _S^1(E_{gs})`$, to the width and thus to the width difference is becoming small. Because the spectral shapes $`\rho _{0,1}(E)`$ approach a Gaussian form in the $`M\mathrm{}`$ limit, we expect this width difference to dominate the ground-state behavior, and to see $`100\%`$ $`S=0`$ ground states for a strong $`C_0`$ coupling, and $`0\%`$ for a stronger $`C_1`$ coupling.
### E The $`C_0=C_1`$ special case
The $`M\mathrm{}`$ behavior for $`C_0=C_1`$ is more subtle and requires an analysis of the subleading $`O(M)`$ terms in Eq. 13. To compute these we need to count more carefully the number of places $`\alpha `$ or $`\beta `$ to which the two interacting particles may jump, given that two of the orbitals remain occupied by the two ‘bystanders’ during the interaction. In the case of a $`J=1`$ pair, we also need to treat correctly the interference between $`J_z=0,\pm 1`$ terms. The end result is that overall, after adding together $`J=0`$ and $`J=1`$ couplings, the $`S=1`$ width is suppressed at $`O(M)`$ compared with the $`S=0`$ width, as a simple physical argument shows. In Eq. 10, the $`jk`$ pair (or any one of the other $`3`$ analogous pairs) may interact either in a $`J=0`$ or $`J=1`$ combination, with the final state leaving $`j`$ doubled up with $`i`$ and with $`k`$ ending up on any of the remaining $`M1`$ sites. For an initial state of the form Eq. 11, this process is not allowed (since the $`ij`$ pair start out in a spin-symmetric combination and cannot end up occupying the same orbital), so $`\mathrm{Tr}_1H^2`$ is reduced at the $`O(M)`$ level. Thus, in the large-$`M`$ limit for $`C_0=C_1`$, we obtain
$$\frac{\mathrm{Tr}_0H^2}{\mathrm{Tr}_1H^2}=1+O(1/M).$$
(15)
By itself, however, this width difference should not be sufficient to determine the ground state behavior. That is because for a Gaussian spectral density with $`d`$ states the location of the ground state scales as $`E_{gs}\sqrt{\mathrm{Tr}H^2}\sqrt{\mathrm{log}d}`$, while the level spacing near the ground state goes as $`\rho ^1(E_{gs})\sqrt{\mathrm{Tr}H^2}/\sqrt{\mathrm{log}d}`$. So the ratio of the expected difference (gap) between the lowest $`S=0`$ and $`S=1`$ states to the level spacing scales as
$$\frac{\mathrm{log}d}{M}\frac{4\mathrm{log}M}{M}0,$$
(16)
and thus in a model of random uncorrelated fluctuations under the two Gaussian spectral envelopes $`\rho _0(E)`$ and $`\rho _1(E)`$, we would expect no preference for either $`S=0`$ or $`S=1`$ ground states (i.e. the ratio of $`S=0`$ to $`S=1`$ ground states should approach the $`3:2`$ ratio valid for the total density of states). We note, however, that in our numerical simulations, the ground state always comes out to have spin $`S=0`$, even for the highest values of $`M`$ ($`M=12`$) that we can achieve, where the width difference is very small. Indeed, as we saw above in Section II D, $`S=0`$ ground states can dominate even when the width difference changes sign, as happens for a pure spin-$`1`$ two-body coupling ($`C_0=0`$). A major reason for this we believe to be strong deviations from the Gaussian shape of a four-particle spectrum (though correlations between $`\rho _0(E)`$ and $`\rho _1(E)`$ may also be relevant; we will discuss these correlations in more detail in Section III B).
### F Spectral shape effects
We recall that the approach of an $`N`$body spectrum to a Gaussian shape given only two-body interactions depends crucially on the assumption that repeated applications of the Hamiltonian act on different pairs of particles and therefore may be taken to commute with each other. So, for example,
$`\mathrm{Tr}H^4`$ $`=`$ $`{\displaystyle H_{\alpha \alpha ^{}}H_{\alpha ^{\prime \prime }\alpha ^{\prime \prime \prime }}H_{\alpha ^{\prime \prime \prime }\alpha ^{\prime \prime }}H_{\alpha ^{}\alpha }}`$ (17)
$`+`$ $`{\displaystyle H_{\alpha \alpha ^{}}H_{\alpha ^{}\alpha }H_{\alpha ^{\prime \prime }\alpha ^{\prime \prime \prime }}H_{\alpha ^{\prime \prime \prime }\alpha ^{\prime \prime }}}`$ (18)
$`+`$ $`{\displaystyle H_{\alpha \alpha ^{}}H_{\alpha ^{\prime \prime }\alpha ^{\prime \prime \prime }}H_{\alpha ^{}\alpha }H_{\alpha ^{\prime \prime \prime }\alpha ^{\prime \prime }}}`$ (19)
$`=`$ $`3\mathrm{Tr}H^2^2,`$ (20)
where $`\alpha `$, $`\alpha ^{}`$, $`\alpha ^{\prime \prime }`$, and $`\alpha ^{\prime \prime \prime }`$ label two-body states (either $`J=0`$ or $`J=1`$), as in Eq. 1, and are summed over. Notice that the third term contributes in general only if the two transitions involve distinct pairs of interacting particles. Otherwise (e.g., for a generic sparse matrix with no two-body interaction structure), we would obtain $`\mathrm{Tr}H^4=2\mathrm{Tr}H^2^2`$, the relation appropriate to a semicircular spectral shape rather than to a Gaussian . So the extent to which the spectrum deviates from a Gaussian shape towards a semicircle may be estimated by considering the probability that successive two-body interactions do not involve distinct pairs of particles.
Consider for example a pure $`J=1`$ two-body coupling. In an $`S=0`$ four-body state (Eq. 10), there are a total of three $`J=1`$ pairs, as we have already seen above. After two of the particles interact in a two-body $`J=1`$ combination, the remaining two particles are guaranteed to be also in a $`J=1`$ state. So there is a $`1/3`$ probability that the next interaction will involve the two particles that have not yet interacted, and thus $`\mathrm{Tr}_0H^4(2+\frac{1}{3})\mathrm{Tr}_0H^2^2`$. On the other hand, if the total many-body state is either $`S=1`$ (Eq. 11) or $`S=2`$, similar counting arguments lead to the conclusion that there is only a $`1/6`$ probability that the second interacting $`J=1`$ pair will be disjoint from the first; thus $`\mathrm{Tr}_{1,2}H^4(2+\frac{1}{6})\mathrm{Tr}_{1,2}H^2^2`$. This behavior for the spectral shapes has been numerically confirmed (e.g., for $`M=10`$ orbitals, $`\frac{\mathrm{Tr}H^4}{\mathrm{Tr}H^2^2}=2.42`$, $`2.18`$, and $`2.15`$, for total many-body spin $`S=0`$, $`1`$, and $`2`$, respectively), and is similar for the $`J=0`$ two-body interaction. So in particular for the $`C_0=C_1`$ case of equal couplings, although the widths of the $`S=0`$ and $`S=1`$ spectra approach each other in the $`M\mathrm{}`$ limit (Eq. 15), the fourth moments remain in a finite ratio in this limit and lead to $`\rho _0(E)`$ having a substantially longer tail than $`\rho _1(E)`$. Therefore, despite the almost equal widths, ground states in a $`C_0=C_1`$ system will be predominantly $`S=0`$, even in the dilute $`M\mathrm{}`$ limit.
### G The general many-body problem
The above analysis of the $`N=4`$ particle system generalizes easily to the real $`N`$-body problem that we are interested in. We begin, as before, by addressing the behavior of the spectral widths in the $`S=0`$ and $`S=1`$ spin sectors. Typical basis states in these two sectors were already given above in Eq. 5. Now consider a pure $`J=0`$ coupling. From Eq. 6 we get
$$\frac{\mathrm{Tr}_0H^2}{\mathrm{Tr}_1H^2}=1+O\left(\frac{1}{N^2}\right)$$
(21)
in the dilute $`M\mathrm{}`$ limit. This can be understood as follows. The only difference scaling as $`M^2`$ in the spectral width is that the $`ij`$ pair in Eq. 5 is in a $`J=0`$ combination for a many body-state of total spin $`S=0`$, but not if the many-body state has total spin $`S=1`$. Thus, the $`S=0`$ state has one more $`J=0`$ pair that can be acted on by the Hamiltonian, out of a total of $`N^2/8`$ pairs with $`J=0`$. Similarly, for a pure $`J=1`$ coupling we have one extra pair that can be acted on in an $`S=1`$ many-body state (out of a total of $`3N^2/8`$ pairs with $`J=1`$), so finally
$$\frac{\mathrm{Tr}_0H^2}{\mathrm{Tr}_1H^2}1+\frac{C_0^2C_1^2}{C_0^2+3C_1^2}\times O\left(\frac{1}{N^2}\right).$$
(22)
If we keep the number of particles $`N`$ fixed and take the system size $`M\mathrm{}`$, then the many-body level spacing of course goes to zero, while the width ratio remains finite for $`C_0C_1`$. Once again we see, this time for general values of $`N`$, that assuming the spectral structure is dominated by the spectral width, we expect $`100\%`$ ($`0\%`$) $`S=0`$ ground states if $`C_0>C_1`$ ($`C_1>C_0`$), in the dilute $`M\mathrm{}`$ limit.
However, we saw in our discussion above of the $`N=4`$ special case that spectral cumulants beyond $`\mathrm{Tr}_SH^2`$ need also to be included in a full analysis. The ratio $`\mathrm{Tr}_SH^4/(\mathrm{Tr}_SH^2)^2`$ always favors a longer tail for $`S=0`$, so we only need to consider the case of a $`J=1`$ coupling, where the width effect and the spectral shape effect are competing with one another. Referring back to Eq. 5, we note that if particles on orbitals $`i`$ and $`k_z`$ interact via a two-body $`J=1`$ interaction, then the two particles which remain on sites $`j`$ and $`l_z`$ are also in a $`J=1`$ state and can subsequently interact if and only if the total system is in an $`S=0`$ state (so that the four particles $`i`$, $`j`$, $`k_z`$, $`l_z`$ are in an $`S=0`$ state). Because the $`j`$, $`l_z`$ pair is distinct from the $`i`$, $`k_z`$ pair by construction, this causes it to be more likely that the next $`J=1`$ pair of interacting particles will be distinct from $`i`$, $`k_z`$ if $`S=0`$, and contributes towards an enhancement of the shape ratio $`\mathrm{Tr}_0H^4/(\mathrm{Tr}_0H^2)^2`$ as compared with $`\mathrm{Tr}_1H^4/(\mathrm{Tr}_1H^2)^2`$. However, in a many-body system this effect is suppressed by $`O(1/N^3)`$ (because there are $`O(N^4)`$ ways to choose two pairs, and only $`O(N)`$ of them are involved in this effect). So for $`C_1>C_0`$ the $`O(1/N^2)`$ difference in widths between $`\rho _1(E)`$ and $`\rho _0(E)`$ (Eq. 22) is expected to exceed the difference in spectral shapes, leading to a dominance of $`S>0`$ states near the edge of the spectrum, in the dilute many-body limit (which we define as $`M\mathrm{}`$ first, followed by $`N\mathrm{}`$).
For $`C_0>C_1`$, on the other hand, both the difference in widths and difference in spectral shapes favors an $`S=0`$ ground state, and this behavior is expected to persist into the dilute many-body limit. For $`C_0=C_1`$, the relative width difference vanishes (Eq. 22) in this limit, but the difference in spectral shapes now takes over, again leading to $`S=0`$ dominance near the edge of the spectrum.
### H Transition at finite densities
We have seen that for finite (large enough) number of particles $`N`$, $`S=0`$ ground state dominance is expected in the dilute $`M\mathrm{}`$ limit, as long as the spin-zero two-body coupling is at least as large as the spin-one coupling, $`C_0C_1`$. In the contrary case, the situation is reversed, with higher-spin states dominating behavior near the edge of the spectrum. What happens, though, if the system is not very dilute? For a finite number of single-particle orbitals $`M`$, we must take into account the subleading behavior in $`M`$ of the spectral widths. An argument very similar to the one leading to Eq. 15, but generalized to large $`N`$ and to an arbitrary ratio of couplings $`C_1/C_0`$ shows that the effect discussed there always favors the $`S=0`$ width, at $`O(1/NM)`$, independent of which coupling is stronger. So for $`C_0C_1`$ nothing particularly interesting happens at finite densities; the preference for $`S=0`$ states near the edge is only enhanced by finite density effects. This preference, however, becomes small (compared to the level spacing) in the many-body limit.
On the other hand, for a stronger spin-$`1`$ coupling ($`C_1>C_0`$), we expect the $`O(1/N^2)`$ effect favoring $`S=0`$ ground states and the $`O(1/NM)`$ effect favoring $`S>0`$ ground states to balance each other, leading to a transition in the width ratio at finite densities. In fact a closer analysis of the $`4`$body system (Eq. 13), setting equal the terms proportional to $`C_1^2`$ in the two expressions, shows that for a pure $`J=1`$ coupling we have a transition just below $`M=11`$, with a larger $`S=0`$ width for $`M<11`$ and larger $`S>0`$ widths for $`M11`$. This is consistent with our numerical observations above. For comparable coupling strengths, the transition of course will take place at lower densities; so, for example for $`C_1^2=C_0^2(1+ϵ)`$, the transition in the width ratio occurs near $`M=10/ϵ`$. Although we are unable to observe this transition numerically for larger numbers of particles, the analytical arguments suggest that one should take place in general at densities $`N/Mϵ`$ for a small coupling difference $`ϵ`$ and at an $`O(1)`$ finite density for a pure $`J=1`$ interaction.
### I Numerical results
In Table I we show some results of a numerical calculation for $`N=4`$ particles distributed over $`5M12`$ orbitals, with a pure $`J=1`$ two-body coupling ($`C_1=1`$, $`C_0=0`$). The first row of the Table shows the ratio of widths $`\sqrt{\mathrm{Tr}_SH^2}`$ in sectors of total many-body spin $`S=0`$ and $`S=1`$. We see that $`\mathrm{Tr}_0H^2>\mathrm{Tr}_1H^2`$ holds at low $`M`$ (high density), while in the dilute regime ($`M12`$) the relation is reversed. The crossover occurs near $`M=11`$ for $`N=4`$, in good agreement with our analytical expectations (see Section II H). The same qualitative behavior is predicted and observed for the higher spin sectors; we see from the second row of the Table that the crossover between $`S=1`$ and $`S=2`$ dominance occurs at $`M=9`$.
We proceed to compare the energies of the lowest-lying states in the different spin sectors, and find a trend towards lower-lying states of higher spin as $`M`$ increases, entirely consistent with the results for the spectral widths. However, by comparing rows 1 and 3 (or 2 and 4) of the Table, we see that for any given system size $`M`$, low-lying states with small total spin are preferred more than would be expected based simply on the width ratio. Thus, for $`M=11`$, where the spin-$`0`$ and spin-$`1`$ widths are nearly equal, the ground state of the system is still nearly always $`S=0`$, and on average the ground state is observed to have energy $`12\%`$ lower than that of the lowest $`S=1`$ state. This is due largely to the spectral shape effects discussed above in Section II F, which favor ground states of lower total spin given equal spectral widths in the different spin sectors.
When comparing $`S=0`$ and $`S=1`$ sectors in particular, we recall from Section II F that the ratio $`\mathrm{Tr}_SH^4/\left(\mathrm{Tr}_SH^2\right)^2`$ is significanlty larger for $`S=0`$, for any value of $`M`$. For $`N>4`$ (see Section II G), spectral shape effects are reduced, but unfortunately we are unable to probe the dilute $`MN`$ regime numerically for $`N6`$. Instead we stay at $`N=4`$ particles and compare the $`S=1`$ and $`S=2`$ sectors, which as we saw previously have nearly equal values of the ratio $`\mathrm{Tr}_SH^4/\left(\mathrm{Tr}_SH^2\right)^2`$. Indeed, by $`M=12`$ the mean values of the lowest-lying $`S=1`$ and $`S=2`$ states become almost equal, and already by $`M=9`$ the lowest-lying $`S=2`$ state appears before the lowest-lying $`S=1`$ state in a significant fraction of systems in the ensemble. The last row of the Table supports the analytical arguments presented above for a transition between lower-spin and higher-spin ground states as a function of the particle density.
### J An alternative model
As we have mentioned previously, the model (Eq. 1) assumes that there is no correlation between two-body interactions in $`J=0`$ and $`J=1`$ two-body states. We may instead consider an alternative model, motivated by Ref. and much subsequent work, in which the interaction Hamiltonian is separated into a spin-independent and a spin-dependent part:
$`H`$ $`=`$ $`{\displaystyle \underset{ijkl}{}}{\displaystyle \underset{J_iJ_jJ_kJ_l}{}}\left(C_VV_{ijkl}+C_WW_{ijkl}\stackrel{}{J}_k\stackrel{}{J}_l\right)`$ (23)
$`\times `$ $`a_{iJ_i}^{}a_{jJ_j}^{}a_{kJ_k}a_{lJ_l}\delta (\stackrel{}{J}_i+\stackrel{}{J}_j\stackrel{}{J}_k\stackrel{}{J}_l),`$ (24)
where the indices $`i,`$ $`j,`$ $`k,`$ and $`l`$ label individual orbitals, $`J_i,`$ $`J_j,`$ $`J_k,`$ and $`J_l=\pm \frac{1}{2}`$ are the $`z`$components of spin for each of the four fermions, and total spin conservation has been enforced. The matrix elements $`V_{ijkl}`$ and $`W_{ijkl}`$ are taken to be real Gaussian random and independent variables, with the constraints
$$V_{jilk}=V_{ikjl}=V_{ljki}=V_{ijkl}$$
(25)
(and similarly for the $`W_{ijkl}`$) imposed by the reality of wave functions and the symmetry of the two-body interaction. The constants $`C_V`$ and $`C_W`$ determine, of course, the relative strengths of the spin-independent and spin-coupled interactions.
The behavior near the ground state for the model of Eq. 24 can be understood in terms very similar to those we used for analyzing the original model of Eq. 1. We assume initially that spin-spin interactions cannot be neglected, i.e. $`C_W0`$. We first note that the spin-spin interaction $`\stackrel{}{J}_k\stackrel{}{J}_l=(1/2)((\stackrel{}{J}_k+\stackrel{}{J}_l)^2J_k^2J_l^2)=(1/2)((\stackrel{}{J}_k+\stackrel{}{J}_l)^23/2)`$ is three times stronger when particles $`k`$ and $`l`$ are in a $`J=0`$ two-body state as compared with a $`J=1`$ two-body state. Now we have seen above that an $`S=0`$ $`N`$particle many-body state has $`O(1/N^2)`$ more $`J=0`$ pairs than an $`S=1`$ many body state: this leads to a wider $`\rho _0(E)`$ envelope and $`S=0`$ ground state dominance, just as for the $`C_0>C_1`$ situation considered in the previous model. Thus, a spin-spin interaction (of either sign) is seen immediately to lead to $`S=0`$ ground state dominance ($`100\%`$ dominance in the $`M\mathrm{}`$ many-orbital limit). At finite densities, $`N/M\mathrm{const}`$, we must once again include corrections to the width difference, of order $`1/NM`$, as in Eq. 15 and in Section II H above. We find, however, that the subleading correction, associated with the impossibility of doubling up two particles in a $`J=1`$ two-body state, always favors the $`S=0`$ many-body state. This is a model-independent property, resulting simply from the counting of two-body $`J=0`$ vs. $`J=1`$ pairs in a many-body state. Therefore the $`O(1/N^2)`$ and $`O(1/NM)`$ effects contribute with the same sign, and even at finite density we expect an excess of $`S=0`$ ground states. Finally, for small systems, shape effects (discussed above in Section II F) also contribute, again favoring the $`S=0`$ ground state.
We may also consider the scenario of a spin-free interaction ($`C_W=0`$), which because of the equal strength with which it couples $`S=0`$ and $`S=1`$ pairs may be compared with the $`C_0=C_1`$ special case discussed above for the model of Eq. 1. As in that case, the width difference is now expected to vanish in the $`M\mathrm{}`$ dilute limit; furthermore as before the width difference eventually becomes small compared with the level spacing in the system. However, at moderate values of the particle number $`N`$, spectral shape effects are expected to be very important, leading to $`S=0`$ ground state excess at arbitrarily large values of $`M`$.
## III Correlations and Pairing
In realistic many-body systems there are strong correlations between ground states of different quantum numbers, and the pairing correlation is of particular importance. This correlation has also been observed in various studies of nuclear models with random two-body interactions . In this Section we investigate correlations between energies and wave functions of low lying states with different quantum numbers.
### A Pairing
Pairing is a well known phenomenon in many-body fermion systems. Its spectral signature is an energy gap that separates the ground state from the excited states. In the presence of pairing, the ground state is a condensate of pairs of fermions whose quantum numbers are related to each other by time reversal symmetry. In the case of the spin systems considered in this work, pairing refers to a ground state of the form $`_jb_j^{}b_j^{}|0`$, where $`b_j^{}`$ is a quasi-particle operator that is related to the single-particle operators $`a_i^{}`$ by a unitary transformation, as $`b_j=_{ı=1}^MU_{ji}a_i`$.
Let us first consider pairing in the single-particle basis generated by the $`a_i^{}`$. The results on wave function structure in random spin systems reported in Ref. show that states at high spectral densities have a number of principal components that agrees with standard random matrix theory, while low-lying states have a considerably smaller number of principal components (thus indicating a simpler structure and some degree of regularity ). However, the structure even of the low-lying states was found to be much more complex than a simple Slater determinant. Based on that previous analysis, and on the analytical results obtained above in Eq. 6 showing very weak dependence of spectral width on double occupancy number, we do not expect to find pairing here in the single-particle basis. To compare these arguments with numerical data we measure the number of doubly occupied single-particle orbitals in a given eigenstate $`|E,S`$ with energy $`E`$ and total spin $`S`$ by computing $`P_{E,S}=E,S|_ja_j^{}a_j^{}a_ja_j|E,S`$ (compare the related quantity $`D`$ for basis states in Section II C). For an ensemble of $`100`$ random Hamiltonians with $`N=6`$ spin-$`\frac{1}{2}`$ fermions on $`M=6`$ sites, $`C_0=1,`$ and $`C_1=0`$, $`P_{E,S}`$ shows fluctuations around $`P_{E,0}1.71`$, $`P_{E,1}1.43`$, $`P_{E,2}0.86`$ and does not exhibit any significant energy dependence. Thus, no pairing is present in the single-particle basis $`a_i^{}`$. The numerical results agree well with the analytical expressions $`P_{E,0}=12/7`$, $`P_{E,1}=270/189`$, and $`P_{E,2}=6/7`$.
Pairing in a general quasi-particle basis can most conveniently be investigated by examining its spectral signature. The existence of a pairing gap is closely related to the approximate degeneracy of the first excited spin-0 level $`E_2^{(S=0)}`$ with the lowest-energy spin-one state $`E_1^{(S=1)}`$. Within the two different models considered in this work, we do not observe any degeneracy between $`E_2^{(S=0)}`$ and $`E_1^{(S=1)}`$, and find the ratio of splitting to gap as
$$\mathrm{\Delta }\frac{\left|E_2^{(S=0)}E_1^{(S=1)}\right|}{\mathrm{min}(E_2^{(S=0)},E_1^{(S=1)})E_1^{(S=0)}}2.8$$
(26)
for an ensemble of $`100`$ Hamiltonians with $`C_0=1`$, $`C_1=0`$, and $`M=N=6`$. Similar results hold for other parameter values, and we never see $`\mathrm{\Delta }1`$, which would be indicative of a pairing structure. Thus, the dominance of spin-0 ground states cannot be caused by pairing correlations within our model. Note that this result differs from the observations made for the nuclear shell model with random interactions . As an illustration we show in Fig. 1 the spin dependent spectra for a system of six fermions on six single-particle orbitals. The random two-body interactions only couple pairs of spin $`J=0`$, i.e. we set $`C_1=0`$. The ground state and the low energy states have spin $`S=0`$. The absence of low-lying spin-degenerate states clearly indicates the absence of pairing correlations.
To study further the subject of pairing, we add a one-body part of the form
$$H_1=C\underset{j=1}{\overset{M}{}}j\left(a_j^{}a_j+a_j^{}a_j\right)$$
(27)
to the Hamiltonian (Eq. 1). Fig. 2 shows the pairing observable $`\mathrm{\Delta }`$ as a function of $`cC/(NMC_0)`$ for $`C_1=0`$ and $`N=M=6`$. For $`c1`$ we find $`\mathrm{\Delta }1`$, indicating strong pairing. In this regime the two-body interaction is a small perturbation, and we obtain pairing in the basis of single particle states. Note that this behavior is not caused by the harmonic form of our one-body potential but rather is determined by the applicability of a single-particle picture. For decreasing values of $`c`$ one leaves this perturbative regime, and the increase of $`\mathrm{\Delta }`$ shows that pairing is suppressed. Note that $`\mathrm{\Delta }>1`$ at sufficiently small values of $`c`$. This indicates that there are several spin-0 levels below the spin-1 ground state, as in Fig. 1 above. This finding is consistent with the results of Section II and hints at the dominance of spin-0 low-lying states. As a check we also replaced the spin-1 ground state energy by the highest spin-1 energy (which should not be at all correlated with the spin-0 ground state energy), while correcting for a nonzero mean of the total spectral density, i.e. replacing $`E_1^{(S=1)}\mathrm{Tr}HE_{\mathrm{max}}^{(S=1)}`$, and recomputed $`\mathrm{\Delta }`$. For the system with a pure two-body force we found no difference when compared to the previous results, confirming that the value of $`\mathrm{\Delta }`$ at $`c=0`$ is consistent with random spectral fluctuations.
### B Level and energy correlations
The absence of pairing in the presence of a strong two-body force is peculiar to our model and is not shared by nuclear shell models with random interactions. It is therefore interesting to look for different correlations between low lying states of spin-0 and spin-1. Such correlations have been observed previously in Refs. . We therefore look at the generalized collectivity defined as
$$f\frac{|S=1|B|S=0|^2}{S=0|B^{}B|S=0},$$
(28)
where $`B=_{i<j}c_{ij}B_{ij}`$ and
$$B_{ij}=(a_i^{}a_j^{}+a_i^{}a_j^{})(a_ia_ja_ia_j)$$
(29)
is a spin-flip operator that transforms a pair of spin zero into a pair of spin one. Following Refs. we choose the coefficients as
$$c_{ij}=S=1|B_{ij}|S=0.$$
(30)
This is more convenient than searching for those coefficients $`c_{ij}`$ that maximize $`f`$. To be definite, we take the ground states of spin zero and spin one, i.e. $`|S=0=|E_1^{(S=0)}`$ and $`|S=1=|E_1^{(S=1)}`$, and compare with the corresponding value that is obtained by substituting $`|S=1|E_{\mathrm{max}}^{(S=1)}`$. The states corresponding to the highest and lowest energies in a sector of definite spin are assumed to be uncorrelated. The results are presented in Table II for $`N=4`$ and different values of $`M`$, once again setting $`C_1=0`$. Obviously, there is a strong correlation between the ground states of different spin. It is instructive to let $`|S=1`$ run through the $`n`$ lowest lying spin-$`1`$ states $`|S=1=|E_1^{(S=1)}\mathrm{}|E_n^{(S=1)}`$ while keeping $`|S=0=|E_1^{(S=0)}`$ fixed. Figure 3 shows that the correlation is strongest for the ground states and decreases for the higher lying states. This finding is consistent with the recent observation that low-lying states are weakly localized in Fock space, while high-lying states are delocalized and agree with random matrix predictions.
It is also interesting to consider energy correlations between ground states of spin zero and spin one. The appropriate correlator is
$$G\frac{\left(\delta E_0^{(S=0)}\right)\left(\delta E_0^{(S=1)}\right)}{\left[\left(\delta E_0^{(S=0)}\right)^2\left(\delta E_0^{(S=1)}\right)^2\right]^{1/2}}$$
(31)
where $`\delta E_0^{(S)}E_0^{(S)}E_0^{(S)}`$ and $`\mathrm{}`$ denotes an ensemble average. We find $`G=0.93`$ for $`N=4`$ particles on $`M=9`$ sites (with $`C_1=0`$), indicating a strong correlation. The observable $`G`$, is however, somewhat misleading, since a significant part of the fluctuations $`\delta E_0`$ from one realization of the ensemble to another arises simply from the fluctuation of the centroid and width of the overall spectral envelope, $`\mathrm{Tr}H`$ and $`\mathrm{Tr}H^2`$. The centroid and width for a given system are almost identical in sectors of different total many-body spin. To understand fluctuations under the spectral envelopes, we should therefore remove this trivial effect by considering the correlation $`\stackrel{~}{G}`$ of $`\stackrel{~}{E}_0^{(S=0)}`$ and $`\stackrel{~}{E}_0^{(S=1)}`$, where
$$\stackrel{~}{E}_0=\frac{E_0\mathrm{Tr}H}{\sqrt{\mathrm{Tr}H^2\left(\mathrm{Tr}H\right)^2}}$$
(32)
We then obtain $`\stackrel{~}{G}=0.76`$ for the same set of parameters. The value of $`\stackrel{~}{G}`$ increases as one goes to larger systems; e.g. for $`N=4`$ particles, $`\stackrel{~}{G}`$ increases monotonically from $`0.20`$ to $`0.76`$ as $`M`$ goes from $`4`$ to $`9`$. These observations are consistent with the existence of correlations between the wave functions of the corresponding states.
Note that the observation of strong correlations between low-lying states of spin-0 and spin-1 is in qualitative agreement with results in shell models and interacting boson models. However, in contrast with the findings in these nuclear models, we do not find any indications of a pairing correlation in the spin systems considered in this work. One can only speculate that this is due to the additional isospin degrees of freedom and the more sophisticated angular momentum coupling scheme used in the nuclear models.
## IV Summary
We have investigated the spin structure of low lying states in fermion systems with random two-body interactions. Numerical calculations show that the ground state has spin zero with almost 100% probability. For a dilute system we derive analytical expressions for the shapes and widths of the spectral densities at different spin. These explain our numerical results. For systems with a spin-1 coupling that is stronger than the spin-0 two-body coupling we predict a transition from spin-0 ground states to ground states of higher spin. While it is difficult to access this regime numerically, we do observe one of its precursors – namely a spin-2 ground state that is lower in energy than the spin-1 ground state.
Furthermore, we have shown that there are strong correlations between wave functions and energies of low-lying states of different spin quantum number. The wave function correlations are manifest in expectation values of spin transition operators and indicate some degree of collective behavior. The existence of energy correlations confirm the picture about the dominance of spin-0 ground states. These correlations decrease as one moves away from the ground state region. In the absence of a one-body force we find no indication for any pairing correlation between low lying states, and our model differs herein from the nuclear models with random interactions. Pairing correlations are only observed after a sufficiently strong one-body potential is added to the random Hamiltonian.
## Acknowledgments
We acknowledge useful discussions with R. Bijker, A. Frank, and T. H. Seligman, and we particularly thank G. F. Bertsch for providing the original motivation and making many helpful suggestions. This work was partially supported by Department of Energy.
|
warning/0007/cond-mat0007005.html
|
ar5iv
|
text
|
# A Study of Quantum Many-Body Systems With Nearest And Next-to-Nearest Neighbour Long-Range Interactions
## I introduction
Exactly solvable and quantum integrable many-body systems, with long-range interactions, are one of the most active fields of current research. The Calogero-Sutherland model (CSM), the Sutherland model (SM) and their variants are the most prominent examples among such systems . These models have found application in various branches of physics , ranging from quantum Hall effect , gauge theories , chaos , fractional statistics etc. The CSM and SM type long-range interactions have also manifested in models dealing with pairing interactions and phase transitions. It is known that, the CSM is related to random matrix theory and enables one to capture the universal aspects of various physical phenomena . The Brownian motion model of Dyson connects the random matrix theory (RMT) with exactly solvable models . The role of RMT in the description of the level statistics of chaotic systems is well-known .
Some time back, a short-range Dyson model was introduced to understand the spectral statistics of systems, which are “non-universal with a universal trend” . It is known that, there are dynamical systems which are neither chaotic nor integrable, the so called pseudo-integrable systems which exhibit the above mentioned level statistics . Aharanov-Bohm billiards , three dimensional Anderson model at the metal-insulator transition point and some polygonal billiards fit into the above description. Quite recently, a new class of one dimensional, exactly solvable many-body quantum mechanical models on the line, with nearest and next-to-nearest neighbour interactions, have been introduced , which are related to this short-range Dyson model. Further, using the symmetrized version of this model, it has been shown that there exists an off-diagonal long-range order in the system which indicates the presence of different quantum phases . Apart from possessing a good thermodynamic limit i.e., $`\mathrm{lim}_N\mathrm{}\frac{E_0}{N}`$ is finite, these models are more physical in the sense that, unlike the CSM and SM, where all particles experience pairwise interaction of identical strength, irrespective of their distances, here the interactions are only nearest neighbour and next-to-nearest neighbour. These models are exactly solvable, but not integrable. Hence, it is interesting to enquire as to how many features of the integrable CSM type systems are retained in the present case. For example, the scattering phenomena is quite interesting in the CSM case, since the outgoing waves can be shown to be of the incoming type, with momenta $`k_i^{}=k_{N+1i}(i=1,2,\mathrm{},N)`$, where $`k_i`$’s are the incoming momenta. Remarkably, the phase shifts are energy independent, a result ascribable to the scale invariance of the inverse square interaction. Since, in the present case also, scale invariance holds and the interaction goes to zero as the particle separation increases, it is of deep interest to study the scattering phenomena. Similarly, for the bound state problem, CSM can be exactly made equivalent to a set decoupled oscillators, via a similarity transformation . Hence, it is also of interest to check the same here to understand the precise differences in the Hilbert space structure between integrable and non-integrable Hamiltonians, possessing identical spectra. It should be mentioned that, in the present case, the degeneracy is less, since the lack of quantum integrability i.e., a desired set of mutually commuting operators having common eigenfunctions with the Hamiltonian, reduces the degeneracy. Also, these models, being of recent origin, need to be analyzed thoroughly, in order to unravel their properties, as has been done for the CSM, SM and their generalizations.
The present paper is devoted to an investigation of these models, and deals with, both the scattering and the bound state problems. It is organized as follows: In Sec.II, we study the scattering problem and show that the scattering state is a coherent state. The connection between the scattering Hamiltonian and that of the free particles in this model is then demonstrated. We then study the scattering phase shift and point out its similarities and differences with the Calogero case. In Sec.III, we analyze the relationship of the bound state problem with the decoupled oscillators and find some wavefunctions explicitly in the Cartesian basis. This analysis reveals explicitly that the degeneracy of this model is less as compared to the CSM. Finally, in Sec.IV, the nearest neighbour and next-to-nearest neighbour $`A_{N1}`$ and $`BC_N`$ models on the circle are studied; a part of their excitation spectra is obtained through symmetry arguments.
## II The Scattering Problem
### A Realization of the scattering states as coherent states
The scattering Hamiltonian, with nearest and next-to-nearest neighbour, inverse square interactions, in the units $`\mathrm{}=m=1`$, is given by,
$$H_{sca}=\frac{1}{2}\underset{i=1}{\overset{N}{}}\frac{^2}{x_i^2}+\beta (\beta 1)\underset{i=1}{\overset{N}{}}\frac{1}{(x_ix_{i+1})^2}\beta ^2\underset{i=2}{\overset{N}{}}\frac{1}{(x_{i1}x_i)(x_ix_{i+1})},$$
(1)
here $`x_{N+i}=x_i`$. The corresponding bound state Hamiltonian, contains an additional oscillator potential:
$$H=\frac{1}{2}\underset{i=1}{\overset{N}{}}\frac{^2}{x_i^2}+\frac{1}{2}\underset{1=1}{\overset{N}{}}x_i^2+\beta (\beta 1)\underset{i=1}{\overset{N}{}}\frac{1}{(x_ix_{i+1})^2}\beta ^2\underset{i=1}{\overset{N}{}}\frac{1}{(x_{i1}x_i)(x_ix_{i+1})}.$$
(2)
It is interesting to note that, the scattering eigenstates can be constructed as coherent states of the bound state eigenfunctions like that of the Calogero case . To be precise, we show that the polynomial part of the bound-state wavefunctions enter into the construction of the scattering states. For this purpose, we first identify an $`SU(1,1)`$ algebra containing the bound and scattering Hamiltonians as its elements, after suitable similarity transformations:
$`Z^1(H_{sca})Z`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{^2}{x_i^2}}+\beta {\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{1}{(x_ix_{i+1})}}(_i_{i+1})T_+,`$ (3)
$`\widehat{S}^1(H/2)\widehat{S}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \underset{i}{}}x_i_i+E_0\right)T_0.`$ (4)
Here, $`Z_{i=1}^Nx_ix_{i+1}^\beta `$ and $`\widehat{S}\mathrm{exp}\{\frac{1}{2}_ix_i^2\}Z\mathrm{exp}\{\frac{1}{2}T_+\}`$ .
Defining,
$`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i}{}}x_i^2T_{},`$ (6)
one can easily check that $`T_\pm `$ and $`T_0`$ satisfy the usual $`SU(1,1)`$ algebra:
$$[T_+,T_{}]=2T_0,[T_0,T_\pm ]=\pm T_\pm .$$
The quadratic Casimir for the above algebra is given by,
$`\widehat{C}=T_{}T_+T_0(T_0+1)=T_+T_{}T_0(T_01).`$ (7)
By finding a canonical conjugate of $`T_+`$ :
$$[T_+,\stackrel{~}{T_{}}]=1,$$
(8)
one can construct the coherent state $`<xm,k>`$, the eigenstate of $`T_+`$ , which are nothing but the scattering states,
$$<xm,k>=U^1P_m(x)=e^{\frac{1}{2}k^2\stackrel{~}{T_{}}}P_m(x),$$
(9)
with
$`T_+P_m(x)\left[{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{^2}{x_i^2}}+\beta {\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{1}{(x_ix_{i+1})}}(_i_{i+1})\right]P_m(x)=0.`$ (10)
It is known that this equation admits homogeneous solutions i.e, $`T_0P_m(x)=[(m+E_0)/2]P_m(x)`$. Here $`m`$ refers to the degree of homogeneity of $`P_m(x)`$. It is easy to see that, $`U^1P_m(x)`$ is the eigenstate of $`T_+`$. Starting from $`T_+P_m(x)=0`$, one gets,
$`U^1T_+UU^1P_m(x)=0,`$ (11)
i.e,
$$T_+U^1P_m(x)=\frac{1}{2}k^2U^1P_m(x).$$
(12)
The scattering state is given by, $`\psi _{sca}=ZU^1P_m(x)`$ , since $`H_{sca}=ZT_+Z^1`$, therefore, $`H_{sca}\psi _{sca}=\frac{k^2}{2}\psi _{sca}`$. To find $`<xm,k>`$ explicitly, we have to determine $`\stackrel{~}{T_{}}`$. By choosing $`\stackrel{~}{T_{}}=T_{}F(T_0)`$, Eq. (8) becomes
$`[T_+,T_{}F(T_0)]=F(T_0)T_+T_{}F(T_0+1)T_{}T_+=1,`$ (13)
$`F(T_0)\{\widehat{C}+T_0(T_01)\}F(T_0+1)\{\widehat{C}+T_0(T_0+1)\}=1,`$ (14)
yielding,
$$F(T_0)=\frac{T_0+a}{\widehat{C}+T_0(T_01)}.$$
(15)
Here, $`a`$ is a parameter to be fixed along with the value of the quadratic Casimir $`\widehat{C}`$, by demanding that the above commutator is valid in the eigenspace of $`T_0`$. Eq. (8) when used on $`P_m(x)`$, yields, $`a=1(E_0+m)/2`$. Similarly,
$`\widehat{C}P_m(x)=(T_{}T_+T_0(T_0+1))P_m(x)=CP_m(x),`$ (16)
where, $`C=\frac{1}{2}(m+E_0)(1(m+E_0)/2)`$. One then finds,
$`F(T_0)P_m(x)`$ $`=`$ $`{\displaystyle \frac{T_0+a}{C+T_0(T_01)}}P_m(x)`$ (18)
$`=`$ $`{\displaystyle \frac{1}{T_0(m+E_0)/2}}P_m(x).`$ (19)
Explicitly, we have,
$`<xm,k>`$ $`=`$ $`e^{\frac{1}{2}k^2\stackrel{~}{T_{}}}e^{T_+}P_m(x)`$ (20)
$`=`$ $`e^{\frac{1}{4}k^2}e^{T_+}e^{\frac{1}{2}k^2\stackrel{~}{T_{}}}P_m(x)`$ (21)
$`=`$ $`e^{\frac{1}{4}k^2}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(k^2/2)^n}{(E_0+m+n)!}}L_n^{(E_01+m)}(r^2/2)P_m(x)`$ (22)
$`=`$ $`e^{k^2/4}(k/2)^{(E_01+m)}(r)^{(E_01+m)}J_{E_01+m}(kr)P_m(x),`$ (23)
where, $`r^2=_ix_i^2`$ and $`J_{E_01+m}(kr)`$ is the Bessel function. Note that, there is an additional factor of $`e^{T_+}`$ in the above equation, this has been introduced for calculational convenience and does not alter our results, since $`e^{T_+}P_m=P_m`$. In order to arrive at the above result, we have made use of the following:
$$T_+\left(r^{2n}P_m(x)\right)=2n(E_01+m+n)r^{2(n1)}P_m(x),$$
and also the identity ,
$$J_\alpha (2\sqrt{xz})e^z(xz)^{\alpha /2}=\underset{n=0}{\overset{\mathrm{}}{}}\frac{z^n}{(n+\alpha +1)!}L_n^\alpha (x).$$
Note that, the above wavefunction can also be obtained by solving the scattering Hamiltonian explicitly. However, we have chosen this algebraic method, since it will be of subsequent use.
### B Connection to Free Particles
The fact that as in the Calogero case, the spectrum of the scattering Hamiltonian matches with that of the free particles and that, the phase shift, as will be shown later, are energy independent, suggests a possible connection of this system with free particles. We now show the same by making use of the algebraic structures already introduced.
The following generators,
$`K_+=H_{sca},\stackrel{~}{K_+}=H_{sca}(\beta =0),`$ (24)
$`K_{}={\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}x_i^2=\stackrel{~}{K}_{},`$ (25)
and
$`K_0={\displaystyle \frac{i}{4}}{\displaystyle \underset{i=1}{\overset{N}{}}}(2x_i_i+1)=\stackrel{~}{K}_0,`$ (26)
satisfy $`[K_0,K_\pm (\stackrel{~}{K}_\pm )]=\pm iK_\pm (\stackrel{~}{K}_\pm )`$, $`[K_{}(\stackrel{~}{K}_{})`$, $`K_+(\stackrel{~}{K}_+)]=2iK_0(\stackrel{~}{K}_0)`$. The normalization of the generators have been chosen differently for convenience. It can be verified that, the following operator,
$`U=e^{\frac{i\pi }{2}(\stackrel{~}{K}_++\stackrel{~}{K}_{})}e^{\frac{i\pi }{2}(K_++K_{})},`$ (27)
maps the $`H_{sca}`$, with interactions, to a interaction-free system, $`i.e.`$,
$`UH_{sca}U^{}=H_{sca}(\beta =0)={\displaystyle \frac{1}{2}}{\displaystyle \underset{i}{}}{\displaystyle \frac{^2}{x_i^2}}.`$ (28)
This result motivates one to analyze, explicitly, the precise correspondence of the respective Hilbert spaces of the interacting and non-interacting systems. Care has to be taken to ensure that, the mapped states are members of the Hilbert space. This analysis, for the present case, as well as for the Calogero model has not been carried out so far. We hope to address this problem in the near future.
### C Analysis of the Scattering Phase Shift
As is well known, in order to obtain the scattering phase shifts, one has to analyze the asymptotic behaviour of the eigenfunctions of the Hamiltonian, when all particles are far apart from each other. In the present case, the most general stationary eigenfunction can be written as a superposition of all the $`m`$ dependent states as,
$`\psi =Z{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{q}{}}C_{m,q}(r/2)^{(E_01+m)}J_{E_01+m}(kr)P_{m,q}(x),`$ (29)
where, $`C_{m,q}`$’s are some coefficients and the index $`q`$ refers to the number of independent $`P_m`$’s at the level $`m`$.
The asymptotic limit of Eq. (29) can be obtained from the asymptotic behaviour of the Bessel function $`J_{E_01+m}(kr)`$ and from the fact that $`r^mP_{m,q}(x)`$ does not change in this limit, since $`P_{m,q}`$’s are homogeneous functions of degree m. One obtains,
$`\psi \psi _{\mathrm{in}}+\psi _{\mathrm{out}},`$ (30)
where,
$`\psi _{\mathrm{in}}e^{i(E_01)\pi /2}(2\pi kr)^{\frac{1}{2}}\left(Zr^{(E_01)}{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{q}{}}C_{m,q}(r/2)^me^{im\pi /2}P_{m,q}(x)\right)e^{ikr},`$ (31)
and
$`\psi _{\mathrm{out}}e^{i(E_01)\pi /2}(2\pi kr)^{\frac{1}{2}}\left(Zr^{(E_01)}{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{q}{}}C_{m,q}(r/2)^me^{im\pi /2}P_{m,q}(x)\right)e^{ikr}.`$ (32)
The crucial differences between the Calogero and the present model arise from this point. Denoting the equivalent of $`P_{m,q}(x)`$’s in the Calogero case, as $`\stackrel{~}{P}_{m,q}(x)`$’s, we first describe the Calogero scattering problem, and then compare the results of this model with the same. Though the $`\stackrel{~}{P}_{m,q}(x)`$’s, which are the solutions of the generalized Laplace equation , have not been found explicitly so far, for arbitrary $`m`$, Calogero had shown the existence and completeness of these homogeneous symmetric polynomials. Hence, by choosing suitable values for the coefficients $`C_{m,q}`$’s, one could characterize the incoming wave, for the Calogero case as,
$$\psi _{\mathrm{in}}c\mathrm{exp}\left\{i\underset{i}{}k_ix_i\right\},$$
with $`k_ik_{i+1},i=1,2,\mathrm{},N1`$, as the stationary eigenfunction in the center-of-mass frame, i.e., $`_ik_i=0`$. Further, since, the $`\stackrel{~}{P}_{m,q}`$’s are symmetric under cyclic permutations and are homogeneous, one obtains, $`\stackrel{~}{P}_{m,q}(Tx)=e^{im\pi }\stackrel{~}{P}_{m,q}(x)`$. where, $`T`$ denotes a cyclic permutation of the particle coordinates. Hence, $`\psi _{\mathrm{out}}`$ for the Calogero model can be written as,
$`\psi _{\mathrm{out}}`$ $``$ $`e^{i(E_01)\pi }(2\pi \overline{k}r)^{\frac{1}{2}}\left(Zr^{(E_01)}{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{q}{}}C_{m,q}(r/2)^me^{im\pi /2}\stackrel{~}{P}_{m,q}(Tx)\right)e^{i\overline{k}r},`$ (33)
where, $`\overline{k}=k`$, and $`Tx_i=x_{N+1i}`$. The action of $`(T)`$ takes a given particle ordering $`x_ix_{i+1}`$ to $`x_{N+1i}x_{Ni}`$, and hence preserves the order i.e., the sector of the configuration space. This is the reason, invariance of $`\stackrel{~}{P}_{m,q}(x)`$’s under cyclic permutation is enough to compute the phase shifts. Comparison with $`\psi _{\mathrm{in}}`$, yields,
$`\psi _{\mathrm{out}}`$ $`=`$ $`ce^{i(E_01)\pi }\mathrm{exp}\left\{i{\displaystyle \underset{i}{}}\overline{k}_iTx_i\right\}`$ (34)
$`=`$ $`ce^{i(E_01)\pi }\mathrm{exp}\left\{i{\displaystyle \underset{i}{}}k_{N+1i}x_i\right\},`$ (35)
where, we have used $`\overline{k}_i=k_i`$ and the cyclic permutation has been carried on the momentum variables. From above, it is clear that, the initial scattering situation characterized by the initial momenta, $`k_i(i=1,2,\mathrm{},N)`$, goes over to the final configuration, characterized by the final momenta, $`k_i^{}=k_{N+1i}`$ and the phase shifts are energy independent. In the present model, $`P_{m,q}(x)`$’s are the solutions of Eq. (10), which is symmetric only under cyclic permutations. Hence, one needs to check if the same steps as narrated above for Calogero’s case also applies here. First of all, we should find out if the number of $`P_{m,q}`$’s are same here. For the sake of clarity, we first consider the four particle case and concentrate on the homogeneous solutions of degree four. Since the monomial symmetric functions, provide a linearly independent basis set, we can expand $`P_4(x_i)`$ as,
$`P_4=a{\displaystyle \underset{i=1}{\overset{4}{}}}x_i^4+b{\displaystyle \underset{ij}{\overset{4}{}}}x_i^3x_j+c{\displaystyle \underset{i<j}{\overset{4}{}}}x_i^2x_j^2+d{\displaystyle \underset{ijl}{\overset{4}{}}}x_i^2x_jx_l+e{\displaystyle \underset{i<j<l<p}{\overset{4}{}}}x_ix_jx_lx_p.`$ (36)
The operation of $`T_+`$ on $`P_4`$ gives the following three sets of conditions,
$`[2(3+4\beta )a2\beta b+(3+4\beta )c2\beta d]{\displaystyle \underset{i=1}{\overset{4}{}}}x_i^2=0,`$ (37)
$`[4\beta a+2(3+4\beta )b2\beta c+2(1\beta )d\beta e]{\displaystyle \underset{i=1}{\overset{4}{}}}x_ix_{i+1}=0,`$ (38)
$`[6(1+2\beta )b+2(12\beta )d]{\displaystyle \underset{i=1}{\overset{4}{}}}x_ix_{i+2}=0,`$ (39)
where a, b, c, d and e are unknown constants to be determined from the above equations. For the purpose of comparison, in the Calogero case,
$`\stackrel{~}{T}_+\stackrel{~}{P}_4(x)\left[{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{4}{}}}{\displaystyle \frac{^2}{x_i^2}}+\beta {\displaystyle \underset{i<j}{\overset{4}{}}}{\displaystyle \frac{1}{(x_ix_j)}}(_i_j)\right]\stackrel{~}{P}_4(x)=0,`$ (40)
yields,
$`[6(1+2\beta )a3\beta b+3(1+2\beta )c3\beta d]{\displaystyle \underset{i=1}{\overset{4}{}}}x_i^2=0,`$ (41)
$`[4\beta a+2(3+2\beta )b2\beta c+2(1\beta )d\beta e]{\displaystyle \underset{i<j}{\overset{4}{}}}x_ix_j=0.`$ (42)
Hence, the number of solutions are less in the present model as compared to the Calogero case. This reduction in the number of solutions takes place because, the interaction term contained in the $`T_+`$ operator, leads to the split of the monomial symmetric functions of degree $`(m2)`$, giving additional conditions unlike the Calogero case. Explicit calculations for a number of few body examples yields similar results. Since the number of $`P_{m,q}`$’s also represent degeneracy, the same is less here. Hence, it is clear that, $`P_{m,q}`$’s do not form a complete set. All the solutions obtained so far are symmetric and belongs to a subset of the Calogero case. The reason of the symmetric nature of the polynomials lies in the interaction term in $`T_+`$. In order that the action of the interaction term, $`\beta _{i=1}^N\frac{1}{x_ix_{i+1}}(_i_{i+1})`$, on the polynomial $`P_m(x)`$ results in a polynomial of degree $`m2`$, the denominator needs to be cancelled. This results in the symmetrization of the polynomial, since the nearest neighbour couplings arising from $`\frac{1}{x_ix_{i+1}}`$ connects each particle with every other member of the set. The reduction in the number of symmetric polynomials can also be understood from the point of view of integrability. Since Calogero model is fully integrable, there are desired number of operators, commuting with the Hamiltonian, which can be used for connecting the members of a given set of $`P_{m,q}`$’s akin to the angular momentum raising and lowering operators in the central force problem. The fact that, the number of $`P_{m,q}`$’s are less here indicates that, the corresponding commuting constants of motion are less here. This point will be further elaborated in the next section.
As has been mentioned earlier, the completeness of the solutions of the generalized Laplace equation in the Calogero model enables one to relate all the partial waves of the incoming state with those of the outgoing state, with constant energy independent phase shifts. As is clear from Eqs. (31), (32) and (33), in the present case, the outgoing wavefront can be made to look like the incoming wavefront. Hovever, only those partial waves, which are solutions of Eq. (10) will acquire energy independent phase shifts $`e^{i(E_01)\pi }`$, the rest will be unaffected by the interaction. This difference between the Calogero and the present model arises, because of the reduction of the number of $`P_{m,q}`$’s here.
## III THE BOUND STATE PROBLEM
### A Mapping of the model to decoupled oscillators
The bound state Hamiltonian is given by,
$$H=\frac{1}{2}\underset{i=1}{\overset{N}{}}_i^2+\frac{1}{2}\underset{1=1}{\overset{N}{}}x_i^2+\beta (\beta 1)\underset{i=1}{\overset{N}{}}\frac{1}{(x_ix_{i+1})^2}\beta ^2\underset{i=1}{\overset{N}{}}\frac{1}{(x_{i1}x_i)(x_ix_{i+1})},$$
(43)
where, $`_i\frac{}{x_i}`$ and $`x_{N+i}x_i`$. It is worth pointing out that, for three particles, this model is equivalent to the CSM, since the three body term vanishes in this case. The ground-state wavefunction and the energy of this system are respectively given by, $`\psi _0=GZ`$ and $`E_0=(N/2+N\beta )`$, where $`G`$ and $`Z`$ have been defined earlier. The first term of $`E_o`$ is the ground state energy of $`N`$ oscillators and the second one comes from the interaction.
In the following, we make use of a method developed in Ref.() to show the equivalence of this model to a set of decoupled oscillators. For that purpose, we perform a similarity transformation on the Hamiltonian, by its ground-state wavefunction, to yield
$`H^{}\psi _0^1H\psi _0={\displaystyle \underset{i}{}}x_i_i+E_0T_+,`$ (44)
where, $`T_+\frac{1}{2}_{i=1}^N\frac{^2}{x_i^2}+\beta _{i=1}^N\frac{1}{(x_ix_{i+1})}(_i_{i+1})`$. Here, we confine ourselves to a sector of the configuration space given by $`x_1x_2\mathrm{}x_{N1}x_N`$. Using the identity,
$$[\underset{i}{}x_i_i,e^{T_+/2}]=T_+e^{T_+/2},$$
it is easy to see that
$`\overline{H}e^{T_+/2}H^{}e^{T_+/2}={\displaystyle \underset{i}{}}x_i_i+E_0.`$ (45)
\>From the above diagonalized form, it is evident that, the spectrum of $`H`$ is like that of $`N`$ uncoupled oscillators and is linear in the coupling parameter $`\beta `$. Explicitly, $`\overline{H}`$ can be made equivalent to the decoupled oscillators:
$`Ge^{T_+(\beta =0)/2}\overline{H}e^{T_+(\beta =0)/2}G^1={\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}_i^2+{\displaystyle \frac{1}{2}}{\displaystyle \underset{1=1}{\overset{N}{}}}x_i^2+E_0N/2.`$ (46)
Making inverse similarity transformations, one can write down the raising and lowering operators for $`H`$, akin to the CSM. However, the eigenfunctions of $`H`$ can be constructed straightforwardly by making use of Eq. (45); since the eigenfunctions of $`_ix_i_i`$ are homogeneous polynomails of degree $`n`$ in the particle coordinates, $`n`$ being any integer. Although, the similarity transformation formally maps the Hilbert space of the interacting problem to that of the free oscillators, it needs a careful study. The singular terms present in $`T_+`$ may yield states, which are not members of the Hilbert space. Below, we clarify this point.
### B Eigenfunctions in the Cartesian Basis
In the following, we present some eigenfunctions for the $`N`$-particle case, computed using the power-sum basis, $`P_l(x)=_{i=1}^Nx_i^l`$, ı.e., $`\psi _l=\psi _0S_l`$; here, $`S_le^{T_+/2}P_l`$, and the corresponding energy eigenvalue is $`E_l=(l+E_0)`$, $`l`$ being an integer.
The wavefunction (unnormalized) corresponding to the center-of-mass degree of freedom, $`R=\frac{1}{N}_{i=1}^Nx_i`$, is found to be (we use the notation $`\psi _{n_1,n_2,\mathrm{},n_N}=\psi _0\mathrm{exp}\{\frac{1}{2}T_+\}_lP_l^{n_l}`$),
$$\psi _{n_1,0,0,\mathrm{}}=\psi _0\mathrm{exp}\{\frac{1}{2}T_+\}R^{n_1}=\psi _0\mathrm{exp}\{\frac{1}{4}\underset{i=1}{\overset{N}{}}_i^2\}R^{n_1}.$$
(47)
This can be cast in the form ,
$$\psi _{n_1,0,0,\mathrm{}}=c\psi _0\underset{_{i=1}^Nm_i=n_1}{}\underset{i=1}{\overset{N}{}}\frac{H_{m_i}(x_i)}{m_i!},$$
(48)
where, $`H_{m_i}(x_i)`$’s are the Hermite polynomials, and c is a constant. Similarly, the eigenfunction for the radial degree of freedom, $`r^2=_ix_i^2`$, can be obtained from,
$$\psi _{0,n_2,0,\mathrm{}}=\psi _0\mathrm{exp}\{\frac{1}{2}T_+\}(r^2)^{n_2}=\psi _0e^{\frac{1}{2}T_+}P_2^{n_2}.$$
(49)
For the sake of clarity, we give below the explicit derivation for $`\psi _{0,1,0,0}`$ for the four particle case and then generalize the result for arbitrary number of particles and levels. For four particles,
$`\psi _{0,1,0,0}=\psi _0e^{\frac{1}{2}T_+}P_2=\psi _0e^{\frac{1}{2}T_+}(x_1^2+x_2^2+x_3^2+x_4^2).`$ (50)
We note that, $`T_+P_2=4(2\beta +1)`$ , $`T_+^2P_2=0`$, and hence, $`\psi _{0,1,0,0}=\psi _0(P_22(2\beta +1))`$. For $`N`$ particles, it can be verified that, $`T_+r^{2n}=2n(E_01+n)r^{2(n1)}`$ and this gives $`\psi _{0,n_2,0,\mathrm{}}`$ as ,
$`\psi _{0,n_2,0,\mathrm{}}`$ $`=`$ $`\psi _0{\displaystyle \underset{m=0}{\overset{n_2}{}}}{\displaystyle \frac{(1)^m}{m!(n_2m)!}}{\displaystyle \frac{(E_01+n_2)!}{(E_01+m)!}}(r^2)^m,`$ (51)
$`=`$ $`c\psi _0L_{n_2}^{E_01}(r^2),`$ (52)
where, $`L_{n_2}^{E_01}(r^2)`$ is the Lagurre polynomial.
Further, for four particles, $`S_3=P_3\frac{3}{2}(2\beta +1)P_1`$. However, it can be checked that, $`S_4=e^{T_+/2}P_4`$, does not terminate as a polynomial and results in a function with negative powers of the particle coordinates, which is not normalizable with respect to the ground-state wavefunction as a measure. This indicates that, there is less degeneracy in the present model, as compared to the symmetrized states of the decoupled oscillators. Finding the other wavefunctions explicitly, for an arbitrary number of particles, and also the exact degeneracy structure of this model remains an open problem. In the above, we have concentrated in finding the wavefunctions in the Cartesian basis; the interested readers are refererred to Ref.() for some wavefunctions in the angular basis.
Below, we list a few eigenfunctions constructed by using the elementary symmetric functions (we follow the notations of Ref.() for the symmetric polynomials).
$`e_1={\displaystyle \underset{1iN}{}}x_i,e_2={\displaystyle \underset{1i<jN}{}}x_ix_j,e_3={\displaystyle \underset{1i<j<kN}{}}x_ix_jx_k,\mathrm{},e_N={\displaystyle \underset{i=1}{\overset{N}{}}}x_i.`$ (53)
In this case, $`\psi _{\{m_i\}}=\psi _0B_{\{m_i\}},`$ $`B_{\{m_i\}}=e^{T_+/2}_i(e_i)^{m_i}`$, and the corresponding eigenvalues are, $`E_{\{m_i\}}=_iim_i+E_0`$. Some of the $`B_{\{m_i\}}`$’s for the four particle case are listed below:
$`B_{2,0,0,0}`$ $`=`$ $`e_1^22,B_{1,1,0,0}=e_1e_2{\displaystyle \frac{1}{2}}(34\beta )e_1,B_{3,0,0,0}=e_1^36e_1,`$ (54)
$`B_{2,1,0,0}`$ $`=`$ $`e_1^2e_2(32\beta )e_1^22e_2+(34\beta ),B_{4,0,0,0}=e_1^412e_1^2+12,`$ (55)
$`B_{3,1,0,0}`$ $`=`$ $`e_1^3e_2{\displaystyle \frac{1}{2}}(94\beta )e_1^36e_1e_2+6(32\beta )e_1,B_{5,0,0,0}=e_1^520e_1^3+60e_1,`$ (56)
$`B_{4,1,0,0}`$ $`=`$ $`e_1^4e_22(3\beta )e_1^412e_1^2e_2+12e_2+6(94\beta )e_1^212(32\beta ).`$ (57)
At this point, it is worth recollecting the Stanley-Macdonald conjecture , which states that, the coefficients of the interaction parameter $`\beta `$ are positive integers, when the Jack polynomials are expressed in terms of the monomial symmetric functions with a suitable normalization. This conjecture was later proved by Sahi . Similar feature appears in the case of the Hi-Jack polynomials , which are the polynomial part of the wavefunctions of the CSM, but with an exception that the coefficient $`\beta `$ can also be negative. Remarkably, from the above explicit computations of the polynomials, we also find that the coefficients of the interacting parameter $`\beta `$ are integers (both positive and negative), though we have used elementary symmetric functions. It will be interesting to check whether the modified Stanley-Macdonald conjecture also holds in the present case for $`N`$ particles.
## IV Model on a Circle
### A $`A_{N1}`$ Model
Recently, Jain et al. have studied a model with nearest and next to nearest neighbour interactions and with periodic boundary conditions as given by
$`H=`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{j}{}}_j^2+\beta (\beta 1){\displaystyle \frac{\pi ^2}{L^2}}{\displaystyle \underset{j}{}}{\displaystyle \frac{1}{\mathrm{sin}^2[\frac{\pi }{L}(x_jx_{j+1})]}}`$ (58)
$``$ $`\beta ^2{\displaystyle \frac{\pi ^2}{L^2}}{\displaystyle \underset{j}{}}\mathrm{cot}[{\displaystyle \frac{\pi }{L}}(x_{j1}x_j)]\mathrm{cot}[{\displaystyle \frac{\pi }{L}}(x_jx_{j+1})],`$ (59)
with $`x_{i+N}=x_i`$. They have shown that the ground state energy eigenvalue and the eigenfunction for this model are given by
$`\psi _0={\displaystyle \underset{j}{\overset{N}{}}}[\mathrm{sin}{\displaystyle \frac{\pi }{L}}(x_jx_{j+1})]^\beta ,E_0=N\beta ^2{\displaystyle \frac{\pi ^2}{L^2}}.`$ (60)
The purpose of this section is to obtain a part of the excitation spectrum of this model. To that end, we substitute
$`\psi =\psi _0\varphi ,`$ (61)
in the eigenvalue equation for the Hamiltonian. It is then easily shown that $`\varphi `$ satisfies the equation
$`\left({\displaystyle \frac{1}{2}}{\displaystyle \underset{j=1}{\overset{N}{}}}_j^2\beta {\displaystyle \frac{\pi }{L}}{\displaystyle \underset{j=1}{\overset{N}{}}}\left[\mathrm{cot}{\displaystyle \frac{\pi }{L}}(x_jx_{j+1})\mathrm{cot}{\displaystyle \frac{\pi }{L}}(x_{j1}x_j)\right]_j+E_0E\right)\varphi =0.`$ (62)
Introducing, $`z_j=\mathrm{exp}(2i\pi x_j/L)`$, Eq. (62) reduces to
$`H_1\varphi =(ϵϵ_0)\varphi `$ (63)
where
$`H_1={\displaystyle \underset{j=1}{\overset{N}{}}}D_j^2+\beta {\displaystyle \underset{j=1}{\overset{N}{}}}\left[{\displaystyle \frac{z_j+z_{j+1}}{z_jz_{j+1}}}\right](D_jD_{j+1}),`$ (64)
with $`D_jz_j\frac{}{z_j}`$ and $`ϵϵ_0=(EE_0)\frac{L^2}{2\pi ^2}`$. It is worth pointing out that the Eqs. (63) and (64) are structurally similar to those in the SM except $`z_k`$ is replaced by $`z_{j+1}`$ in our case.
It may be noted that $`H_1`$ commutes with the momentum operator $`P=\frac{2\pi }{L}_{i=1}^Nz_i\frac{}{z_i}`$. Hence $`\varphi `$ is also an eigenstate of the momentum operator, i.e.,
$`P\varphi =\kappa \varphi .`$ (65)
Further, if $`\varphi `$ is an eigenstate of $`H_1`$ and $`P`$ then
$`\varphi ^{}=G^q\varphi ,G=\mathrm{\Pi }_{i=1}^Nz_i,`$ (66)
is also an eigenstate of $`H^{}`$ and P with eigenvalues $`ϵϵ_0+Nq^2+2q\kappa `$ and $`\kappa +Nq`$ respectively. Here $`q`$ is any integer (both positive and negative). Note that the multiplication by $`G`$ implements Galilei boost.
It may be noted that the Hamiltonian and hence the $`\varphi `$ equation is invariant under $`z_jz_j^1`$. Since, $`z_j=e^{2i\pi x_j/L}`$, hence $`z_j^1=e^{2i\pi x_j/L}`$ thereby indicating the presence of left and right moving modes with momentum $`\kappa `$ and -$`\kappa `$. Hence it follows that, if one obtains a solution with momentum $`\kappa `$, then by changing $`z_jz_j^1`$, one can get another solution with the same energy but with the opposite momentum (-$`\kappa `$). Thus all the excited states with nonzero momentum are (at least) doubly degenerate.
Finally, let us discuss the solutions to the $`\varphi `$ equation. So far we have been able to obtain the following four solutions.
$`(i)\varphi `$ $`=`$ $`e_1,ϵϵ_0=1+2\beta ,`$ (67)
$`(ii)\varphi `$ $`=`$ $`e_{N1},ϵϵ_1=N1+2\beta ,`$ (68)
$`(iii)\varphi `$ $`=`$ $`e_1e_{N1}{\displaystyle \frac{N}{1+2\beta }}e_N,ϵϵ_0=N+2+4\beta ,`$ (69)
$`(iv)\varphi `$ $`=`$ $`e_N,ϵϵ_0=N.`$ (70)
Here $`e_j`$ (j=1,2,…,N) denotes the elementary symmetric functions as defined by Eq. (53) (defined in terms of $`z_j`$). For example, $`e_2=z_1z_2+\mathrm{}+z_{N1}z_N`$ and it has N( N - 1)/2 number of terms.
As mentioned above, each of these solution is doubly degenerate. For example solutions $`e_1`$ and $`e_{N1}/e_N`$ are degenerate. By taking the linear combinations of these two complex solutions, it is easily seen that the two degenerate real solutions are
$`\varphi ={\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{cos}u_i,\varphi ={\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{sin}u_i,`$ (71)
where $`u_i=2\pi x_i/L`$. Similarly, all other doubly excited state solutions can be rewritten as two independent real solutions. It would appear from this discussion that all the excited states are doubly degenerate. However, this is not so. In particular, consider
$`\varphi ={\displaystyle \frac{e_1e_{N1}}{e_N}}{\displaystyle \frac{N}{1+2\beta }}.`$ (72)
It is easily shown that it is an exact solution to Eq. (63) with $`ϵϵ_0=2+4\beta `$ but with momentum eigenvalue $`\kappa =0`$. This is a nondegenerate solution as it remains invariant under $`z_iz_i^1`$. In terms of the trignometric functions it can be rewritten as
$`\varphi ={\displaystyle \underset{i<j}{\overset{N}{}}}\mathrm{cos}(u_iu_j)+{\displaystyle \frac{N\beta }{1+2\beta }}.`$ (73)
At first sight it appears somewhat surprising that whereas in the Sutherland model, there are so many excited state solutions, in our case one is able to obtain so few solutions. In this context it may be noted that whereas the Hamiltonian in the Sutherland case is invariant under the full permutation group $`S_N`$, in our case for $`N>3`$ the Hamiltonian $`H_1`$ as given by Eq. (64) has only cyclic symmetry. If one looks at the solutions to Eq. (64), then one finds that the condition of no pole in the $`\beta `$-dependent term almost forces $`\varphi `$ to ‘be invariant under $`S_N`$. Now out of the various $`e_i`$ (i=1,2,…,N), the only ones in which the demand of cyclic invariance necessarily ensures invariance under full permutation group are precisely $`e_1,e_{N1},e_N`$, in terms of which we have obtained the four solutions.
### B $`BC_N`$ Model
Recently Auberson et al. have studied a $`BC_N`$ model with nearest and next-to-nearest neighbour interactions and with periodic boundary conditions as given by
$`H`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{^2}{x_{i}^{}{}_{}{}^{2}}}+\beta (\beta 1){\displaystyle \frac{\pi ^2}{L^2}}{\displaystyle \underset{i=1}{\overset{N}{}}}\left[{\displaystyle \frac{1}{\mathrm{sin}^2}}{\displaystyle \frac{\pi }{L}}(x_ix_{i+1})+{\displaystyle \frac{1}{\mathrm{sin}^2}}{\displaystyle \frac{\pi }{L}}(x_i+x_{i+1})\right]`$ (74)
$``$ $`\beta ^2{\displaystyle \frac{\pi ^2}{L^2}}{\displaystyle \underset{i=1}{\overset{N}{}}}\left[\mathrm{cot}{\displaystyle \frac{\pi }{L}}(x_{i1}x_i)\mathrm{cot}{\displaystyle \frac{\pi }{L}}(x_{i1}+x_i)\right]`$ (76)
$`\left[\mathrm{cot}{\displaystyle \frac{\pi }{L}}(x_ix_{i+1})+\mathrm{cot}{\displaystyle \frac{\pi }{L}}(x_i+x_{i+1})\right]+g_1{\displaystyle \frac{\pi ^2}{L^2}}{\displaystyle \underset{i}{}}{\displaystyle \frac{1}{\mathrm{sin}^2}}{\displaystyle \frac{\pi }{L}}x_i+g_2{\displaystyle \frac{\pi ^2}{L^2}}{\displaystyle \underset{i}{}}{\displaystyle \frac{1}{\mathrm{sin}^2}}{\displaystyle \frac{2\pi }{L}}x_i.`$
Following them, we restrict the coordinates $`x_i`$ to the sector $`Lx_1x_2\mathrm{}.x_N0`$. As shown by Auberson et al. , the ground state eigenfunction is given by
$`\psi _0={\displaystyle \underset{i=1}{\overset{N}{}}}\mathrm{sin}^\gamma \theta _i{\displaystyle \underset{i=1}{\overset{N}{}}}(\mathrm{sin}^22\theta _i)^{\gamma _1/2}{\displaystyle \underset{i=1}{\overset{N}{}}}[\mathrm{sin}^2(\theta _i\theta _{i+1})]^{\beta /2}{\displaystyle \underset{i=1}{\overset{N}{}}}[\mathrm{sin}^2(\theta _i+\theta _{i+1})]^{\beta /2},`$ (77)
where $`g_1,g_2`$ are related to $`\gamma ,\gamma _1`$ by
$`g_1={\displaystyle \frac{\gamma }{2}}[\gamma +2\gamma _11],g_2=2\gamma _1(\gamma _11).`$ (78)
The corresponding ground state energy turns out to be
$`E_0={\displaystyle \frac{N\pi ^2}{2L^2}}(\gamma +2\gamma _1+2\beta )^2.`$ (79)
By setting one or both of the coupling constants $`\gamma ,\gamma _1`$ to zero we get the other root systems i.e.
$`B_N:\gamma _1=0,C_N:\gamma =0,D_N:\gamma =\gamma _1=0.`$ (80)
The purpose of this subsection is to obtain a part of the excitation spectrum of the $`BC_N,B_N,C_N,D_N`$ models. To that end, we substitute
$`\psi =\psi _0\varphi ,`$ (81)
in the eigenvalue equation for the above Hamiltonian where $`\psi _0`$ is as given by Eq. (77). It is easy to show that in that case $`\varphi `$ satisfies the equation
$`[{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{^2}{\theta _j^2}}+2\beta {\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{cot}(\theta _j\theta _{j+1})({\displaystyle \frac{}{\theta _j}}{\displaystyle \frac{}{\theta _{j+1}}})`$ (82)
$`+2\beta {\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{cot}(\theta _j+\theta _{j+1})({\displaystyle \frac{}{\theta _j}}+{\displaystyle \frac{}{\theta _{j+1}}})+2\gamma {\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{cot}\theta _j{\displaystyle \frac{}{\theta _j}}`$ (83)
$`+4\gamma _1{\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{cot}2\theta _j{\displaystyle \frac{}{\theta _j}}+(EE_0){\displaystyle \frac{2L^2}{\pi ^2}}]\varphi =0,`$ (84)
where $`\theta _j=\frac{\pi x_j}{L}`$. Introducing, $`z_j=\mathrm{exp}(2i\theta _j)`$, Eq. (82) reduces to
$`H_1\varphi =(ϵϵ_0)\varphi `$ (85)
where
$`H_1={\displaystyle \underset{j=1}{\overset{N}{}}}D_j^2+\gamma {\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{z_j+1}{z_j1}}D_j+2\gamma _1{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{z_j^2+1}{z_j^21}}D_j`$ (86)
$`+\beta {\displaystyle \underset{j=1}{\overset{N}{}}}\left[{\displaystyle \frac{z_j+z_{j+1}}{z_jz_{j+1}}}\right](D_jD_{j+1})+\beta {\displaystyle \underset{j=1}{\overset{N}{}}}\left[{\displaystyle \frac{z_jz_{j+1}+1}{z_jz_{j+1}1}}\right](D_j+D_{j+1}).`$ (87)
Here $`D_jz_j\frac{}{z_j}`$ while $`ϵϵ_0=(EE_0)\frac{L^2}{2\pi ^2}`$. It is worth pointing out that the above eqation is structurally similar to that of the $`BC_N`$ of SM, except $`z_{j+1}`$ is replaced by $`z_k`$ in our case.
Note that apart from the cyclic symmetry, the Hamiltonian and hence the $`\varphi `$ equation is also invariant under $`z_jz_j^1`$. As a consequence, as in the $`BC_N`$ Sutherland model , it turns out that even in our case the polynomial eigenfunctions of $`H_1`$ with $`BC_N`$ symmetry as given by Eq. (86) are symmetric polynomials in $`(z_j+\frac{1}{z_j})`$ i.e. in $`\mathrm{cos}(2\pi x_j/L)`$.
Finally, let us discuss the solutions to the $`\varphi `$ equation. So far we have been able to obtain only one solution in the $`BC_N`$ case but are able to obtain several solutions in the $`D_N`$ case and also a few in the $`B_N`$ and $`C_N`$ cases.
### C Exact Solution for the $`BC_N`$ Model
It is easily checked that the exact solution is
$`\varphi \varphi _{BC_N}=\varphi _1+\alpha ,ϵϵ_0=1+\gamma +2\gamma _1+4\beta ,`$ (88)
where
$`\varphi _1={\displaystyle \underset{j=1}{\overset{N}{}}}(z_j+{\displaystyle \frac{1}{z_j}}),\alpha ={\displaystyle \frac{2N\gamma }{1+\gamma +2\gamma _1+4\beta }}.`$ (89)
We will see that this solution (and in fact other solutions, if any, in the $`BC_N`$ case) will play important roles in our construction of solutions for other systems like $`B_N,C_N`$ and $`D_N`$.
### D Exact Solutions For the $`B_N`$ Model ($`\gamma _1=0`$)
Apart from the obvious solution $`\varphi (z_j;\beta ,\gamma ,\gamma _1=0)`$ as given by Eq. (88) it turns out that there are other solutions corresponding to the spinorial representation for the $`B_N`$ model. These are
$`\varphi \varphi ^+=\mathrm{\Pi }_{j=1}^N(\sqrt{z_j}+{\displaystyle \frac{1}{\sqrt{z_j}}}),ϵ^+ϵ_0={\displaystyle \frac{N}{4}}[1+2\gamma +4\beta ].`$ (90)
In order to obtain the other solution, we start with the ansatz
$`\varphi =\varphi ^+\psi ^+`$ (91)
where $`\varphi ^+`$ is as given by Eq. (90) and consider the equation $`H_1\varphi =(ϵϵ_0)\varphi `$. It is easily seen that in that case $`\psi ^+`$ satisfies
$`H_1^+(z_j;\beta ,\gamma ,\gamma _1=0)\psi ^+=(ϵϵ^+)\psi ^+,`$ (92)
where $`ϵ^+`$ is as given by Eq. (90) while
$`H_1^+`$ $`=`$ $`H_1(z_j;\beta ,\gamma ,\gamma _1=0)+{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{z_j1}{z_j+1}}D_j`$ (93)
$`=`$ $`H_1(z_j;\beta ,\gamma 1,\gamma _1=1).`$ (94)
We thus see that $`\psi ^+`$ essentially satisfies the same equation as satisfied by the $`BC_N`$ Hamiltonian but with the value of $`\gamma `$ and $`\gamma _1`$ shifted to $`\gamma 1`$ and 1 respectively. Thus it follows that once we obtain solutions of the $`BC_N`$ problem, all these will give us new solutions of the spinorial type for the $`B_N`$ model as given by Eqs. (91) to (93). Unfortunately, so far we have been able to obtain only one solution in the $`BC_N`$ case as given by Eqs. (88) and (89). Using that solution, it then follows that the new spinorial solution for the $`B_N`$ model is as given by Eq. (91) with energy
$`ϵϵ_0=2+\gamma +4\beta +{\displaystyle \frac{N}{4}}[1+2\gamma +4\beta ],`$ (95)
while $`\psi ^+\varphi _{BC_N}(z_j;\beta ,\gamma 1,\gamma _1=1)`$ with $`\varphi _{BC_N}`$ being given by Eqs. (88) and (89).
### E Exact Solutions For the $`D_N`$ Model ($`\gamma =\gamma _1=0`$)
Apart from the above solutions (88), (90) and (91) (with $`\gamma =\gamma _1=0`$), we have found several other solutions in the $`D_N`$ case. This is related to the fact that unlike $`B_N`$, there are two distinct classes of spinor representations for $`D_N`$. Besides, there are also some additional solutions in this case. It may be noted that in this case the Hamiltonian $`H_1`$ acting on $`\varphi `$ is
$`H_1(z_j;\beta )=H_1(z_j;\beta ,\gamma =0,\gamma _1=0).`$ (96)
The first new solution that we have is given by
$`\varphi \varphi ^{}=\mathrm{\Pi }_{j=1}^N(\sqrt{z_j}{\displaystyle \frac{1}{\sqrt{z_j}}}),ϵ^{}ϵ_0={\displaystyle \frac{N}{4}}[1+4\beta ].`$ (97)
Note that the two solutions (90) (with $`\gamma =0`$) and (97) which correspond to the two different spinorial representations are degenerate in energy.
In order to obtain the other solution, we start with the ansatz
$`\varphi =\varphi ^{}\psi ^{}`$ (98)
where $`\varphi ^{}`$ is as given by Eq. (97) and consider the equation $`H_1\varphi =(ϵϵ_0)\varphi `$. It is easily seen that in that case $`\psi ^{}`$ satisfies
$`H_1^{}(z_j;\beta ,\gamma =0,\gamma _1=0)\psi ^{}=(ϵϵ^{})\psi ^{},`$ (99)
where $`ϵ^{}`$ is as given by Eq. (97) while
$`H_1^{}`$ $`=`$ $`H_1(z_j;\beta ,\gamma =0,\gamma _1=0)+{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{z_j+1}{z_j1}}D_j`$ (100)
$`=`$ $`H_1(z_j;\beta ,\gamma =1,\gamma _1=0).`$ (101)
We thus see that $`\psi ^{}`$ essentially satisfies the same equation as satisfied by the $`B_N`$ Hamiltonian but with the value of $`\gamma `$ being fixed at 1. Using the solution for the $`B_N`$ case as given by Eqs. (88) and (89) (with $`\gamma =1,\gamma _1=0`$) it then follows that the new spinorial solution for the $`D_N`$ model is as given by Eq. (98) with energy
$`ϵϵ_0=2+4\beta +{\displaystyle \frac{N}{4}}[1+4\beta ],`$ (102)
while $`\psi ^{}\varphi _{BC_N}(z_j;\beta ,\gamma =1,\gamma _1=0)`$ with $`\varphi _{BC_N}`$ being given by Eqs. (88) and (89). Notice that this solution is degenerate in energy with the solution (91) (with $`\gamma =0`$).
In addition, we find that the product of the two “spinorial solutions” is also a solution of the $`D_N`$ model, i.e.
$`\varphi \varphi ^+\varphi ^{}=\mathrm{\Pi }_{j=1}^N(z_j{\displaystyle \frac{1}{z_j}}),ϵ^+ϵ_0=N[1+2\beta ].`$ (103)
In order to obtain another solution, as above we start with the ansatz
$`\varphi =\varphi ^+\varphi ^{}\psi ^+`$ (104)
where $`\varphi ^+,\varphi ^{}`$ are as given by Eqs. (90) and (96) respectively and consider the equation $`H_1\varphi =(ϵϵ_0)\varphi `$. It is easily seen that in that case $`\psi ^+`$ satisfies
$`H_1^+(z_j;\beta ,\gamma =0,\gamma _1=0)\psi ^+=(ϵϵ^+)\psi ^+,`$ (105)
where $`ϵ^+`$ is as given by Eq. (103) while
$`H_1^+`$ $`=`$ $`H_1(z_j;\beta ,\gamma =0,\gamma _1=0)+2{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{z_j^2+1}{z_j^21}}D_j`$ (106)
$`=`$ $`H_1(z_j;\beta ,\gamma =0,\gamma _1=1).`$ (107)
We thus see that $`\psi ^+`$ essentially satisfies the same equation as satisfied by the $`C_N`$ Hamiltonian but with the value of $`\gamma _1`$ being fixed at 1. Using the solution for the $`BC_N`$ case as given by Eqs. (88) and (89) (with $`\gamma =0,\gamma _1=1`$), it then follows that the new solution for the $`D_N`$ model is as given by Eq. (104) with energy
$`ϵϵ_0=3+4\beta +N[1+2\beta ],`$ (108)
while $`\psi ^+\varphi _{BC_N}(z_j;\beta ,\gamma =0,\gamma _1=1)`$ with $`\varphi _{BC_N}`$ being given by eqs. (88) and (89).
### F Exact Solution For the $`C_{N=4}`$ Model ($`\gamma =0`$)
So far we have discussed all the solutions which are valid for any N. In addition, in the special case of $`N=4`$, we have been able to obtain a solution in the $`C_{N=4}`$ (and two solutions in the $`D_{N=4}`$) case. The solution for the $`C_{N=4}`$ case is given by
$`\varphi \varphi ^{31}=A[(z_1+{\displaystyle \frac{1}{z_1}})(z_2+{\displaystyle \frac{1}{z_2}})(z_3+{\displaystyle \frac{1}{z_3}})+C.P.]+B\varphi _1,`$ (109)
where
$`ϵϵ_0=3+6\gamma _1+8\beta ,B={\displaystyle \frac{8A\beta }{1+2\gamma _1+2\beta }}.`$ (110)
Here, by C.P. one means cyclic permutations and $`\varphi _1`$ is as given by Eq. (88).
### G Exact Solutions For the $`D_{N=4}`$ Model ($`\gamma =\gamma _1=0`$)
Clearly, one obvious solution in the $`D_{N=4}`$ case is obtained from the solution (109) by putting $`\gamma _1=0`$. The other solution is obtained by making use of the ansatz as given by Eq. (104), i.e. let
$`\varphi =\varphi ^{31}\varphi ^+,`$ (111)
with $`\varphi ^{31}`$ being given by Eq. (109). Using Eqs. (104) to (106) it then follows that this is a solution for the $`D_{N=4}`$ model with
$`ϵϵ_0=13+16\beta ,B={\displaystyle \frac{8A\beta }{3+2\beta }}.`$ (112)
## V Conclusions
In conclusion, we have carried out a systematic study of the many-body Hamiltonian, related to the short range Dyson model. The scattering state of this model is obtained and is shown to be a coherent state. Akin to the CSM, the connection of the scattering Hamiltonian to a free system is established. Unlike the Calogero model, analysis of the scattering process for the present model reveals that, only a part of the partial waves acquire energy independent phase shifts. We then showed the mapping of the bound system to decoupled oscillators and by explicitly computing some of the bound-state eigenfunctions, we find that, the present model has less degeneracy as compared to the Calogero-Sutherland model. Finally, we have studied the $`A_{N1}`$, $`BC_N`$, $`B_N`$, $`C_N`$ and $`D_N`$ models on a circle and obtained a part of their excitation spectrum. A number of open problems, like finding the $`P_{m,q}(x)`$’s explicitly for the N-body problem, characterizing the degeneracy structure, finding the complete eigenspectra and conserved quantities for these type of models still remains to be tackled. We hope to come back to some of these issues in future.
M.E, N.G and P.K.P would like to thank Prof. V. Srinivasan and Prof. S. Chaturvedi for useful discussions.
|
warning/0007/hep-ph0007006.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Once a light CP-even Higgs boson is discovered, it will be crucial to measure as many of its couplings as possible with the highest accuracy possible. By measuring the Higgs couplings to gauge bosons, one can learn whether only one Higgs boson is involved in electroweak symmetry breaking dynamics. Moreover, the Higgs couplings to vector bosons are sensitive to the possible existence of non-doublet isospin structure in the Higgs sector. By measuring the Higgs couplings to fermion pairs, one can learn whether the Higgs boson is responsible for fermion mass generation. Knowledge of the trilinear and quartic Higgs self-couplings, although extremely difficult to obtain, would allow one to reconstruct the Higgs potential and directly test the mechanism of electroweak symmetry breaking. Finally, if the couplings can be measured at the level of the radiative corrections, one could then derive significant constraints on new physics beyond the reach of the present accelerators. A detailed study of radiative corrections to the Higgs couplings would be especially important if a light Higgs boson were discovered in the mass range predicted by the minimal supersymmetric extension of the Standard Model (MSSM), but supersymmetric (SUSY) particles were not found. In this case, the precise experimental determination of Higgs couplings could provide indirect information about the preferred region of SUSY parameter space. For example, one could predict (in the context of the MSSM) whether the data favored a SUSY spectrum below the 1 TeV energy scale.
It is well known that the tree-level couplings of the lightest MSSM Higgs boson ($`h^0`$) to fermion pairs and gauge bosons tend to their Standard Model (SM) values in the decoupling limit, $`M_AM_Z`$ , where $`M_A`$ is the mass of the CP-odd neutral Higgs boson ($`A^0`$) of the MSSM. As a consequence of this decoupling, distinguishing the lightest MSSM Higgs boson in the large $`M_A`$ limit from the Higgs boson of the Standard Model (SM) will be very difficult. Formally, the decoupling of all SUSY particles (including the radiative corrections) implies that in the effective low-energy theory, all observables tend to their SM values in the limit of large SUSY masses and large $`M_A`$. It has been shown that all of the SUSY particles in the MSSM, including the heavy Higgs bosons $`H^0`$, $`A^0`$ and $`H^\pm `$, decouple at one-loop order from the low-energy electroweak gauge boson physics . In particular, the contributions of the SUSY particles to low-energy processes either fall as inverse powers of the SUSY mass parameters or can be absorbed into counterterms for the tree-level couplings of the low-energy theory . As a result, the radiative corrections involving SUSY particles go to zero in the asymptotic large SUSY mass limit. Our aim is to determine the nature of the decoupling limit at one-loop for the couplings of $`h^0`$ to SM particles.
In this paper, we focus on the $`h^0`$ coupling to $`b\overline{b}`$. This coupling determines the partial width of $`h^0b\overline{b}`$, which is by far the dominant decay mode of $`h^0`$ in most of the MSSM parameter space. Because this decay is dominant, accurate knowledge of the $`h^0b\overline{b}`$ coupling is very important for Higgs boson searches. At LEP and the Tevatron, the primary Higgs search channel is $`h^0b\overline{b}`$. The experimental reach of Higgs boson searches at the upcoming Tevatron Run 2 depends critically on the $`h^0b\overline{b}`$ branching ratio . In contrast, the Higgs boson searches at LEP do not depend as critically on the $`h^0b\overline{b}`$ decay. At LEP, there is sufficient cross-section to detect the Higgs boson in multiple channels. Moreover, even without observing the Higgs decay products, the Higgs boson mass can be reconstructed by detecting the recoiling $`Z`$ boson in $`e^+e^{}Zh^0`$. At the CERN LHC, the primary Higgs search channel in the mass region below 130 GeV is the rare decay $`h^0\gamma \gamma `$. The Higgs event rate in this channel is affected strongly if the total width of $`h^0`$ is modified due to corrections to the dominant $`b\overline{b}`$ decay mode . The same holds true for other search channels at the LHC such as $`h^0\tau ^+\tau ^{}`$ .
In this paper we study the MSSM radiative corrections to the $`h^0b\overline{b}`$ coupling at one loop, to leading order in $`\alpha _s`$, and we analyze in detail their behavior in the decoupling limit.<sup>1</sup><sup>1</sup>1The radiatively-corrected $`h^0b\overline{b}`$ coupling in the decoupling limit has also been considered previously in refs. . These corrections are due to the SUSY-QCD (SQCD) sector and arise from gluino and bottom-squark (sbottom) exchange. Because of the dependence on the strong coupling constant, these are expected to be the most significant one-loop MSSM contributions over much of the MSSM parameter space. Potentially significant contributions can also arise from the SUSY-electroweak sector (the most significant of which are proportional to the Higgs-top quark Yukawa coupling); we will address these corrections elsewhere and do not consider them here.
The SQCD corrections to the $`h^0b\overline{b}`$ coupling were first calculated using a diagrammatic approach in ref. (which also contains results for the SUSY-electroweak corrections). These corrections have also been obtained in refs. . The SQCD corrections were also computed in an effective Lagrangian approach in ref. , using the SUSY contributions to the $`b`$–quark self energy obtained in refs. and neglecting terms suppressed by inverse powers of SUSY masses.
The radiatively-corrected $`h^0b\overline{b}`$ coupling depends on the CP-even Higgs mixing angle $`\alpha `$. At tree-level, this mixing angle is determined by fixing $`\mathrm{tan}\beta `$ and $`M_A`$. At one-loop order, there are no $`𝒪(\alpha _s)`$ corrections to this mixing angle. As a result, working to leading order in $`\alpha _s`$, we may employ tree-level relations for $`\alpha `$ in our computation of the $`h^0b\overline{b}`$ coupling. This procedure is no longer adequate once one-loop SUSY-electroweak effects are included. In the latter case, the one-loop radiative corrections to $`\alpha `$ must be taken into account, as described in refs. . These papers show that the interplay between the radiative corrections to the mixing angle and to the $`h^0b\overline{b}`$ coupling can be very important for Higgs collider phenomenology, particularly in the case when the branching ratio for $`h^0b\overline{b}`$ is suppressed.<sup>2</sup><sup>2</sup>2The relevance of the suppressed $`h^0b\overline{b}`$ coupling for phenomenology has also been emphasized in refs. . This is most easily seen as follows. When radiative corrections to the mixing angle $`\alpha `$ are included, it becomes possible to tune this angle to zero independent of the value of $`\mathrm{tan}\beta `$ by varying the SUSY parameters. At $`\alpha =0`$, the tree-level couplings of $`h^0`$ to $`b\overline{b}`$ and $`\tau ^+\tau ^{}`$ vanish, as do the ordinary QCD corrections to the $`h^0b\overline{b}`$ coupling. However, because the SQCD corrections to the $`h^0b\overline{b}`$ coupling include contributions from diagrams involving the $`h^0`$ coupling to sbottoms, they remain nonzero at $`\alpha =0`$. As a result, the $`h^0\tau ^+\tau ^{}`$ coupling goes to zero at a different point in SUSY parameter space than the $`h^0b\overline{b}`$ coupling does . We will come back to these issues and study the approach to decoupling of the SUSY-electroweak corrections in a later paper.
In some regions of the MSSM parameter space, the SQCD corrections to the $`h^0b\overline{b}`$ coupling become so large that it is important to take into account higher-order corrections. This has been carried out in refs. by resumming the leading $`\mathrm{tan}\beta `$ contributions to all orders of perturbation theory using an effective Lagrangian approach. This resummation is not important in our present work because we are interested in the decoupling limit in which the one-loop corrections to the $`h^0b\overline{b}`$ coupling are small.
In this paper we analyze the full diagrammatic formulae for the on-shell one-loop SQCD corrections to the $`h^0b\overline{b}`$ coupling. We perform expansions in inverse powers of SUSY masses in order to examine the decoupling behavior when the SUSY masses are large compared to $`M_Z`$. The SQCD corrections depend on a number of different SUSY mass parameters, and the relative sizes of these masses affect the manifestation of the decoupling. To remain as model-independent as possible, we make no assumptions about relations among the SUSY parameters that may arise from grand unification or specific SUSY-breaking scenarios. We consider the soft-SUSY-breaking parameters and the $`\mu `$ parameter to be independent parameters whose magnitudes are all of order 1 TeV.
In this paper, we demonstrate that in the limit of large $`M_A`$ (in this limit one also has $`M_{H^0},M_{H^\pm }M_Z`$) and large sbottom and gluino masses ($`M_{\stackrel{~}{b}_i},M_{\stackrel{~}{g}}M_Z`$), the SM expression for the $`h^0b\overline{b}`$ one-loop coupling is recovered. That is, the SQCD corrections to the $`h^0b\overline{b}`$ coupling decouple in the limit of large SUSY masses and large $`M_A`$. In particular, we examine the case of large $`\mathrm{tan}\beta `$, for which the SQCD corrections are enhanced. This enhancement can delay the onset of decoupling and give rise to a significant one-loop correction, even for moderate to large values of the SUSY masses.
This paper is organized as follows. In Section 2 we define our notation and briefly review the Higgs and sbottom sectors of the MSSM. In Section 3 we give the exact one loop formula for the SQCD corrections to the $`h^0b\overline{b}`$ coupling. In Section 4 we derive analytic expressions for the SQCD corrections in the limit of large SUSY masses. We analyze the decoupling of the SQCD corrections for various hierarchies of mass parameters, and numerically compare the analytic approximations to the exact one-loop result. In Section 5 we summarize our conclusions. Finally, the Appendix contains expansions of the one-loop integrals used in our calculations.
## 2 Higgs and sbottom masses in the MSSM
In the MSSM, the parameters of the Higgs sector are constrained at tree-level in such a way that the Higgs masses and mixing angles depend on only two unknown parameters. These are commonly chosen to be the mass of the CP-odd neutral Higgs boson $`A^0`$ and the ratio of the vacuum expectation values (vevs) of the two Higgs doublets, $`\mathrm{tan}\beta =v_2/v_1`$. (For a review of the MSSM Higgs sector, see .) In terms of these parameters, the mass of the charged Higgs boson $`H^\pm `$ at tree level is $`M_{H^\pm }^2=M_A^2+M_W^2`$ , and the masses of the CP-even neutral Higgs bosons $`h^0`$ and $`H^0`$ are obtained by diagonalizing the tree-level mass-squared matrix,
$`^2`$ $`=`$ $`\left(\begin{array}{cc}M_A^2\mathrm{sin}^2\beta +M_Z^2\mathrm{cos}^2\beta & (M_A^2+M_Z^2)\mathrm{sin}\beta \mathrm{cos}\beta \\ (M_A^2+M_Z^2)\mathrm{sin}\beta \mathrm{cos}\beta & M_A^2\mathrm{cos}^2\beta +M_Z^2\mathrm{sin}^2\beta \end{array}\right).`$ (2.3)
The eigenvalues of this matrix are,
$$M_{H^0,h^0}^2=\frac{1}{2}\left[M_A^2+M_Z^2\pm \sqrt{\left(M_A^2+M_Z^2\right)^24M_A^2M_Z^2\mathrm{cos}^22\beta }\right],$$
(2.4)
with $`M_{h^0}<M_{H^0}`$. At tree-level, $`M_{h^0}M_Z|\mathrm{cos}2\beta |`$; this bound is saturated at large $`M_A`$. We choose a convention where the vevs are positive so that $`0<\beta <\pi /2`$. The mixing angle that diagonalizes $`^2`$ is given at tree-level by
$$\mathrm{tan}2\alpha =\mathrm{tan}2\beta \frac{M_A^2+M_Z^2}{M_A^2M_Z^2}.$$
(2.5)
In the conventions employed here, $`\pi /2<\alpha <0`$ (see ref. for further details). From the above results it is easy to obtain:
$$\mathrm{cos}^2(\beta \alpha )=\frac{M_{h^0}^2(M_Z^2M_{h^0}^2)}{M_A^2(M_{H^0}^2M_{h^0}^2)}.$$
(2.6)
In the limit of $`M_AM_Z`$, the expressions for the Higgs masses and mixing angle simplify and one finds
$`M_{h^0}^2`$ $``$ $`M_Z^2\mathrm{cos}^22\beta ,`$
$`M_{H^0}^2`$ $``$ $`M_A^2+M_Z^2\mathrm{sin}^22\beta ,`$
$`\mathrm{cos}^2(\beta \alpha )`$ $``$ $`{\displaystyle \frac{M_Z^4\mathrm{sin}^24\beta }{4M_A^4}}.`$ (2.7)
Two consequences are immediately apparent. First, $`M_AM_{H^0}M_{H^\pm }`$, up to corrections of $`𝒪(M_Z^2/M_A)`$. Second, $`\mathrm{cos}(\beta \alpha )=0`$ up to corrections of $`𝒪(M_Z^2/M_A^2)`$. This limit is known as the decoupling limit because when $`M_A`$ is large, one can define an effective low-energy theory below the scale of $`M_A`$ in which the effective Higgs sector consists only of one light CP-even Higgs boson, $`h^0`$, whose couplings to Standard Model particles are indistinguishable from those of the SM Higgs boson . From eq. 2.7, one can easily derive:
$$\mathrm{cot}\alpha =\mathrm{tan}\beta \frac{2M_Z^2}{M_A^2}\mathrm{tan}\beta \mathrm{cos}2\beta +𝒪\left(\frac{M_Z^4}{M_A^4}\right).$$
(2.8)
When radiative corrections to the CP-even Higgs mass-squared matrix are taken into account, the upper bound on $`M_{h^0}`$ increases substantially to $`M_{h^0}<135`$ GeV (assuming all supersymmetric particles are no heavier than about 1 TeV), and corrections to $`\alpha `$ become substantial for low $`M_A`$. These corrections are well known and the leading contributions have been computed up to two-loop order. In this paper we consider only the contributions to the $`h^0b\overline{b}`$ coupling of order $`\alpha _s`$ at one loop. Because the $`𝒪(\alpha _s)`$ contributions to the CP-even Higgs mass-squared matrix only first arise at the two-loop level, the radiative corrections to this matrix are irrelevant to our present work. (In contrast, they do contribute to the one-loop SUSY-electroweak corrections to the $`h^0b\overline{b}`$ coupling.)
From direct searches at LEP the MSSM $`h^0`$ and $`A^0`$ masses are constrained to be $`M_{h^0}>88.3`$ GeV and $`M_A>88.4`$ GeV . For a range of values of $`\mathrm{tan}\beta `$ close to one, the theoretical upper bound on $`M_{h^0}`$ is lower than the experimental lower bound, so the corresponding region of $`\mathrm{tan}\beta `$ can be ruled out. Because of the radiative corrections, the variation of the upper bound depends primarily on the precise value of the top quark mass and the mixing in the stop sector. For the conservative choice of $`m_t<179.4`$ GeV and mixing in the stop sector that maximizes the upper bound on $`M_{h^0}`$, values of $`\mathrm{tan}\beta `$ between 0.8 and 1.5 are excluded .
We now discuss the parameters of the sbottom sector. The tree-level sbottom squared-mass matrix is:
$$_{\stackrel{~}{b}}^2=\left(\begin{array}{cc}M_L^2& m_bX_b\\ m_bX_b& M_R^2\end{array}\right),$$
(2.9)
where we use the notation,
$`X_b`$ $`=`$ $`A_b\mu \mathrm{tan}\beta ,`$
$`M_L^2`$ $`=`$ $`M_{\stackrel{~}{Q}}^2+m_b^2+M_Z^2(I_3^bQ_bs_W^2)\mathrm{cos}2\beta ,`$
$`M_R^2`$ $`=`$ $`M_{\stackrel{~}{D}}^2+m_b^2+M_Z^2Q_bs_W^2\mathrm{cos}2\beta .`$ (2.10)
Here $`I_3^b=1/2`$ and $`Q_b=1/3`$ are the isospin and electric charge of the $`b`$-quark, respectively and $`s_W\mathrm{sin}\theta _W`$. The parameters $`M_{\stackrel{~}{Q}}`$ and $`M_{\stackrel{~}{D}}`$ are the soft-SUSY-breaking masses for the third-generation SU(2) squark doublet $`\stackrel{~}{Q}=(\stackrel{~}{t}_L,\stackrel{~}{b}_L)`$ and the singlet $`\stackrel{~}{D}=\stackrel{~}{b}_R^{}`$, respectively. $`A_b`$ is a soft-SUSY-breaking trilinear coupling and $`\mu `$ is the bilinear coupling of the two Higgs doublet superfields. The sbottom mass eigenstates are
$$\stackrel{~}{b}_1=\mathrm{cos}\theta _{\stackrel{~}{b}}\stackrel{~}{b}_L+\mathrm{sin}\theta _{\stackrel{~}{b}}\stackrel{~}{b}_R;\stackrel{~}{b}_2=\mathrm{sin}\theta _{\stackrel{~}{b}}\stackrel{~}{b}_L+\mathrm{cos}\theta _{\stackrel{~}{b}}\stackrel{~}{b}_R,$$
(2.11)
where $`\pi /4\theta _{\stackrel{~}{b}}\pi /4`$ is defined so that $`\stackrel{~}{b}_1`$ ($`\stackrel{~}{b}_2`$) is predominantly $`\stackrel{~}{b}_L`$ ($`\stackrel{~}{b}_R`$). The sbottom mass eigenvalues are then given by
$$M_{\stackrel{~}{b}_{1,2}}^2=\frac{1}{2}\left[M_L^2+M_R^2\pm \sigma _{LR}\sqrt{(M_L^2M_R^2)^2+4m_b^2X_b^2}\right],$$
(2.12)
where<sup>3</sup><sup>3</sup>3If $`M_L=M_R`$, then $`\sigma _{LR}`$ is not well-defined. In the present context, a useful convention is to set $`\sigma _{LR}=\sigma _X`$ \[where $`\sigma _X\mathrm{sgn}(X_b)`$\] if $`M_L=M_R`$. Nevertheless, one can check that our final expressions for the radiative corrections in Section 4 are independent of this choice.
$$\sigma _{LR}\mathrm{sgn}(M_L^2M_R^2),$$
(2.13)
and the mixing angle $`\theta _{\stackrel{~}{b}}`$ is given by
$`\mathrm{cos}2\theta _{\stackrel{~}{b}}`$ $`=`$ $`{\displaystyle \frac{M_L^2M_R^2}{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}},`$
$`\mathrm{sin}2\theta _{\stackrel{~}{b}}`$ $`=`$ $`{\displaystyle \frac{2m_bX_b}{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}.`$ (2.14)
Note that in our conventions, $`M_{\stackrel{~}{b}_1}>M_{\stackrel{~}{b}_2}`$ if $`\sigma _{LR}>0`$, whereas the order of the sbottom masses switches if $`\sigma _{LR}<0`$.
From direct searches at the Tevatron , the sbottoms must be heavier than about 140 GeV, assuming that the mass of the lightest neutralino $`\stackrel{~}{\chi }_1^0`$ is less than half the mass of the lighter sbottom. For heavier neutralinos, the Tevatron searches lose efficiency. In this region the direct searches at LEP place a lower bound on the sbottom masses of about 85 GeV. The LEP bounds are valid only for $`\stackrel{~}{b}\stackrel{~}{\chi }_1^0`$ mass splittings larger than about 5 GeV, so that the decay mode $`\stackrel{~}{b}b\stackrel{~}{\chi }_1^0`$ is kinematically accessible.
The limits on the gluino mass $`M_{\stackrel{~}{g}}`$ are more model-dependent. If one assumes relations between the gaugino masses such that they unify at the GUT scale, then $`M_{\stackrel{~}{g}}`$ is constrained from direct searches at the Tevatron to be greater than 173 GeV, independent of the squark masses .
## 3 SQCD corrections to $`h^0b\overline{b}`$
The $`h^0b\overline{b}`$ coupling is given at one-loop level to order $`\alpha _s`$ by
$$\overline{g}_{hbb}=g_{hbb}+\delta g_{hbb}^{QCD}+\delta g_{hbb}^{SQCD}g_{hbb}\left(1+\mathrm{\Delta }_{QCD}+\mathrm{\Delta }_{SQCD}\right),$$
(3.1)
where $`\overline{g}_{hbb}`$ is the one-loop coupling, $`g_{hbb}`$ is the tree-level coupling, $`\delta g_{hbb}^{QCD}`$ is the radiative correction from pure QCD , and $`\delta g_{hbb}^{SQCD}`$ is the one-loop SQCD contribution.
The tree-level $`h^0b\overline{b}`$ coupling is given by
$$g_{hbb}=\frac{gm_b\mathrm{sin}\alpha }{2M_W\mathrm{cos}\beta }.$$
(3.2)
Note that in the limit of large $`M_A`$, $`\mathrm{sin}\alpha \mathrm{cos}\beta `$ and $`g_{hbb}`$ tends to the SM coupling, $`g_{hbb}^{SM}=gm_b/(2M_W)`$. The one-loop corrections to the $`h^0b\overline{b}`$ coupling modify the $`h^0b\overline{b}`$ decay width as follows, keeping only correction terms of $`𝒪(\alpha _s)`$:
$$\overline{\mathrm{\Gamma }}(h^0b\overline{b})=\mathrm{\Gamma }(h^0b\overline{b})(1+2\mathrm{\Delta }_{QCD}+2\mathrm{\Delta }_{SQCD}),$$
(3.3)
where $`\overline{\mathrm{\Gamma }}`$ is the one-loop partial width and $`\mathrm{\Gamma }`$ is the tree-level partial width.
The SQCD contribution to the $`h^0b\overline{b}`$ coupling comes from diagrams involving the exchange of virtual gluinos ($`\stackrel{~}{g}`$) and sbottoms ($`\stackrel{~}{b_i}`$), as shown in fig. 1.
We have
$$\delta g_{hbb}^{SQCD}=\left(\delta g_{hbb}\right)_3^{SQCD}+\left(\delta g_{hbb}\right)_2^{SQCD}+\left(\delta g_{hbb}\right)_\mathrm{X}^{SQCD},$$
(3.4)
consisting of contributions from the vertex correction, the $`b`$-quark wave function renormalization, and the counterterm from the renormalization of the $`b`$-quark Yukawa coupling, respectively. To compute the one-loop Yukawa counterterm contribution, we note that the Higgs wave function, the vevs (and hence $`\mathrm{tan}\beta `$) and the parameters $`g`$, $`M_W`$ and $`\alpha `$ receive no $`𝒪(\alpha _s)`$ corrections at one-loop. Thus, to leading order in $`\alpha _s`$, $`\left(\delta g_{hbb}\right)_\mathrm{X}^{SQCD}`$ can be easily obtained from eq. 3.2 and depends only on the $`b`$-quark mass counterterm as follows:
$$\left(\delta g_{hbb}\right)_\mathrm{X}^{SQCD}=g_{hbb}\frac{(\delta m_b)^{SQCD}}{m_b}.$$
(3.5)
In eq. 3.5, $`(\delta m_b)^{SQCD}`$ is the SQCD contribution to the $`b`$-quark mass counterterm, which is fixed by defining $`m_b`$ to be the pole of the one-loop $`𝒪(\alpha _s)`$ $`b`$-quark propagator. This is the on-shell renormalization scheme.
We have computed the various contributions to $`\delta g_{hbb}^{SQCD}`$ \[see eq. 3.4\]. The contribution of the one-loop vertex is given by:
$`{\displaystyle \frac{\left(\delta g_{hbb}\right)_3^{SQCD}}{g_{hbb}}}={\displaystyle \frac{\alpha _s}{3\pi }}\{[{\displaystyle \frac{2M_Z^2}{m_b}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}(I_3^bc_b^2Q_bs_W^2c_{2b})+2m_b+Y_bs_{2b}]`$
$`\times \left[m_bC_{11}+M_{\stackrel{~}{g}}s_{2b}C_0\right](m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_1}^2)`$
$`+\left[{\displaystyle \frac{2M_Z^2}{m_b}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}(I_3^bs_b^2+Q_bs_W^2c_{2b})+2m_bY_bs_{2b}\right]`$
$`\times \left[m_bC_{11}M_{\stackrel{~}{g}}s_{2b}C_0\right](m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2,M_{\stackrel{~}{b}_2}^2)`$
$`+\left[{\displaystyle \frac{M_Z^2}{m_b}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}(I_3^b2Q_bs_W^2)s_{2b}+Y_bc_{2b}\right]`$
$`\times \left[2M_{\stackrel{~}{g}}c_{2b}C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_2}^2)\right]\},`$ (3.6)
where $`c_b\mathrm{cos}\theta _{\stackrel{~}{b}}`$, $`c_{2b}\mathrm{cos}2\theta _{\stackrel{~}{b}}`$, $`s_b\mathrm{sin}\theta _{\stackrel{~}{b}}`$, etc., and $`Y_b`$ arises in the Higgs coupling to sbottoms:
$$Y_bA_b+\mu \mathrm{cot}\alpha .$$
(3.7)
The contribution from the $`b`$-quark self-energy and the $`h^0b\overline{b}`$ vertex counterterm is given by
$`{\displaystyle \frac{\left(\delta g_{hbb}\right)_2^{SQCD}+\left(\delta g_{hbb}\right)_\mathrm{X}^{SQCD}}{g_{hbb}}}=`$
$`{\displaystyle \frac{\alpha _s}{3\pi }}\{{\displaystyle \frac{M_{\stackrel{~}{g}}}{m_b}}s_{2b}[B_0(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)B_0(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2)]`$
$`2m_b^2\left[B_1^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)+B_1^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2)\right]`$
$`2m_bM_{\stackrel{~}{g}}s_{2b}[B_0^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)B_0^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2)]\}.`$ (3.8)
Our notation for the loop integrals $`B_0`$, $`B_0^{}`$, $`B_1^{}`$, $`C_0`$ and $`C_{11}`$ is defined in the Appendix. We have checked that our results are in agreement with the calculations of ref. .
## 4 Analytic and numerical results
We now analyze the decoupling behavior of the SQCD corrections to the $`h^0b\overline{b}`$ coupling. We derive approximate analytic expressions for the SQCD corrections in the limit of large SUSY mass parameters and explore the nature of the decoupling limit.
We define our expansion parameters as follows. Since we are interested in the limit of large SUSY mass parameters, we consider all the soft-SUSY-breaking mass parameters and the $`\mu `$ parameter to be of the same order (collectively denoted by $`M_{SUSY}`$) and much heavier than the electroweak scale. That is,
$$M_{SUSY}M_LM_RM_{\stackrel{~}{g}}\mu A_bM_{EW},$$
(4.1)
where $`M_L`$ and $`M_R`$ are defined in eq. 2.10. We give expansions of the SQCD corrections to the $`h^0b\overline{b}`$ coupling in inverse powers of the SUSY mass parameters, up to order $`M_{EW}^2/M_{SUSY}^2`$. We consider $`M_Z`$, $`M_{h^0}`$, $`m_b\mathrm{tan}\beta `$, and $`m_b\mathrm{cot}\alpha `$ to all be of order $`M_{EW}`$. We neglect small contributions of order $`m_b^2/M_{SUSY}^2`$ and $`m_bM_{EW}/M_{SUSY}^2`$ that are not enhanced by $`\mathrm{tan}\beta `$ or $`\mathrm{cot}\alpha `$. The expansions of the loop integrals are given in the Appendix. There are two possible extreme configurations of the sbottom mass-squared matrix that we must separately consider: maximal and near-zero mixing.
Maximal mixing ($`\theta _{\stackrel{~}{b}}\pm \pi /4`$) between $`\stackrel{~}{b}_L`$ and $`\stackrel{~}{b}_R`$ arises when the splitting between the diagonal elements of the mass-squared matrix is small compared to the off-diagonal elements: $`|M_L^2M_R^2|m_b|X_b|`$. Because of the $`\mathrm{tan}\beta `$ enhancement in $`X_b`$, $`m_bX_b`$ is of order $`M_{EW}M_{SUSY}`$. In this case we consider $`|M_L^2M_R^2|`$ to be of order $`M_{EW}^2`$, so that the mass splitting between the two sbottoms is small compared to their masses and we must take care to treat it properly in the expansions. We consider this case in Section 4.1.
Near-zero mixing between $`\stackrel{~}{b}_L`$ and $`\stackrel{~}{b}_R`$ arises when the splitting between the diagonal elements of the mass-squared matrix is large compared to the off-diagonal elements, $`|M_L^2M_R^2|m_b|X_b|`$. This is the case usually considered in the literature, because $`M_L`$ and $`M_R`$ depend on two different soft-SUSY-breaking parameters $`M_{\stackrel{~}{Q}}`$ and $`M_{\stackrel{~}{D}}`$, respectively, and the $`b`$-quark mass in the off-diagonal elements is small. In this case the mass splitting between the two sbottoms will be of the same order as their masses (i.e., $`|M_L^2M_R^2|`$ is of order $`M_{SUSY}^2`$) and this has to be treated properly in the expansions. We consider this case in Section 4.2.
### 4.1 Maximal $`\stackrel{~}{b}_L\stackrel{~}{b}_R`$ mixing
Maximal mixing in the sbottom sector arises when $`|M_L^2M_R^2|m_b|X_b|`$. In this limit, we can expand the sbottom mass-squared eigenvalues in powers of the small parameter $`(M_L^2M_R^2)/m_bX_b`$ (which is of order $`M_{EW}/M_{SUSY}`$) as follows:
$$M_{\stackrel{~}{b}_{1,2}}^2M_S^2\pm \mathrm{\Delta }^2,$$
(4.2)
where we have defined
$`M_S^2`$ $`=`$ $`\frac{1}{2}(M_L^2+M_R^2)=\frac{1}{2}(M_{\stackrel{~}{b}_1}^2+M_{\stackrel{~}{b}_2}^2)`$
$`\mathrm{\Delta }^2`$ $`=`$ $`\sigma _{LR}m_b|X_b|\left[1+{\displaystyle \frac{(M_L^2M_R^2)^2}{8m_b^2X_b^2}}\right].`$ (4.3)
Here $`M_S^2`$ is of order $`M_{SUSY}^2`$ while $`\mathrm{\Delta }^2`$ is of order $`M_{EW}M_{SUSY}`$. Expanding the expressions for the mixing angle in terms of the same small parameter, we obtain
$`\mathrm{cos}2\theta _{\stackrel{~}{b}}`$ $``$ $`\left|{\displaystyle \frac{M_L^2M_R^2}{2m_bX_b}}\right|,`$
$`\mathrm{sin}2\theta _{\stackrel{~}{b}}`$ $``$ $`\sigma _{LR}\sigma _X\left[1{\displaystyle \frac{(M_L^2M_R^2)^2}{8m_b^2X_b^2}}\right],`$ (4.4)
where $`\sigma _X\mathrm{sgn}(X_b)`$. Expanding eqs. 3.6 and 3.8 to order $`M_{EW}^2/M_{SUSY}^2`$, we find
$`\mathrm{\Delta }_{SQCD}={\displaystyle \frac{\alpha _s}{3\pi }}\{{\displaystyle \frac{\mu M_{\stackrel{~}{g}}}{M_S^2}}(\mathrm{tan}\beta +\mathrm{cot}\alpha )f_1(R){\displaystyle \frac{Y_bM_{\stackrel{~}{g}}M_{h^0}^2}{12M_S^4}}f_4(R)`$
$`+{\displaystyle \frac{\mu ^2m_b^2\mathrm{tan}^2\beta }{2M_S^4}}\left[{\displaystyle \frac{\mathrm{cot}\alpha }{\mathrm{tan}\beta }}f_4(R){\displaystyle \frac{M_{\stackrel{~}{g}}}{M_S^2}}\left(Y_b2A_b{\displaystyle \frac{\mathrm{cot}\alpha }{\mathrm{tan}\beta }}\right)f_3(R)\right]`$
$`+{\displaystyle \frac{2}{3}}{\displaystyle \frac{M_Z^2}{M_S^2}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}I_3^b(f_5(R)+{\displaystyle \frac{M_{\stackrel{~}{g}}X_b}{M_S^2}}f_2(R))+𝒪\left({\displaystyle \frac{m_bM_{EW}}{M_{SUSY}^2}}\right)\},`$ (4.5)
where $`RM_{\stackrel{~}{g}}/M_S`$. The functions $`f_i(R)`$ arise from the expansions of the loop integrals and are given in the Appendix. They are normalized so that $`f_i(1)=1`$. Note that terms of order $`(M_L^2M_R^2)^2/(m_b^2X_b^2)`$ cancel exactly in the leading order of the large $`M_{SUSY}`$ expansion \[eq. 4.5\].
The first term in eq. 4.5 is zeroth order in $`M_{SUSY}`$. That is, if the ratios between the SUSY parameters are fixed and the SUSY mass scale is taken arbitrarily heavy, this term remains constant. This non-decoupling behavior has been pointed out previously in refs. . If the SUSY mass scale is much larger than $`M_A`$, then one may define a low-energy effective theory by integrating out the SUSY particles. This low-energy effective theory contains two Higgs doublets, whose couplings to fermions are unrestricted (i.e., each Higgs doublet couples to both up-type and down-type quarks), characteristic of the so-called general type-III model .
The remaining terms are of order $`M_{EW}^2/M_{SUSY}^2`$. In contrast to the first term, they depend on $`A_b`$ (through $`X_b`$ and $`Y_b`$). However, the contribution proportional to $`A_b`$ is not enhanced when $`\mathrm{tan}\beta `$ (or $`\mathrm{cot}\alpha `$) is large, and so is less significant at large $`\mathrm{tan}\beta `$ than the contribution proportional to $`\mu `$. Neglecting all terms that are not enhanced by large $`\mathrm{tan}\beta `$ or $`\mathrm{cot}\alpha `$, we find that $`\mathrm{\Delta }_{SQCD}`$ is proportional to the product $`\mu M_{\stackrel{~}{g}}`$. Because of this, for large $`\mathrm{tan}\beta `$ the sign of $`\mathrm{\Delta }_{SQCD}`$ can be used as a test of the anomaly-mediated SUSY breaking scenario , which predicts a negative $`M_{\stackrel{~}{g}}`$ . Of course, the sign of $`\mu `$ must be determined from another SUSY process for the sign of $`M_{\stackrel{~}{g}}`$ to be extracted.
### 4.2 Near-zero $`\stackrel{~}{b}_L\stackrel{~}{b}_R`$ mixing
Near-zero mixing in the sbottom sector arises when $`|M_L^2M_R^2|m_b|X_b|`$. This corresponds to taking the difference between the physical sbottom masses to be of the same order as the masses themselves. In this case we write our results in terms of the physical sbottom masses and expand in powers of the small parameter $`m_bX_b/(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)`$, which we take to be of order $`M_{EW}/M_{SUSY}`$. The mixing angle is then given by eq. 2, from which one easily derives the expansion
$$\mathrm{cos}2\theta _{\stackrel{~}{b}}1\frac{2m_b^2X_b^2}{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)^2}.$$
(4.6)
Expanding eqs. 3.6 and 3.8 to order $`M_{EW}^2/M_{SUSY}^2`$, and writing the result in terms of the physical sbottom masses, we find:
$`\mathrm{\Delta }_{SQCD}={\displaystyle \frac{\alpha _s}{3\pi }}\{{\displaystyle \frac{2\mu M_{\stackrel{~}{g}}}{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}(\mathrm{tan}\beta +\mathrm{cot}\alpha )h_1(R_1,R_2)+2M_{h^0}^2{\displaystyle \frac{M_{\stackrel{~}{g}}Y_bh_2(R_1,R_2)}{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)^2}}`$
$`+2M_Z^2{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}[(I_3^bQ_bs_W^2)({\displaystyle \frac{f_5(R_1)}{3M_{\stackrel{~}{b}_1}^2}}{\displaystyle \frac{M_{\stackrel{~}{g}}X_b}{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}{\displaystyle \frac{f_1(R_1)}{M_{\stackrel{~}{b}_1}^2}}`$
$`+{\displaystyle \frac{2M_{\stackrel{~}{g}}X_bh_1(R_1,R_2)}{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)^2}})`$
$`+Q_bs_W^2({\displaystyle \frac{f_5(R_2)}{3M_{\stackrel{~}{b}_2}^2}}+{\displaystyle \frac{M_{\stackrel{~}{g}}X_b}{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}{\displaystyle \frac{f_1(R_2)}{M_{\stackrel{~}{b}_2}^2}}{\displaystyle \frac{2M_{\stackrel{~}{g}}X_bh_1(R_1,R_2)}{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)^2}})]`$
$`{\displaystyle \frac{2\mu ^2M_{\stackrel{~}{g}}m_b^2\mathrm{tan}^2\beta }{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)^2}}\left(Y_b2A_b{\displaystyle \frac{\mathrm{cot}\alpha }{\mathrm{tan}\beta }}\right)\left({\displaystyle \frac{f_1(R_1)}{M_{\stackrel{~}{b}_1}^2}}+{\displaystyle \frac{f_1(R_2)}{M_{\stackrel{~}{b}_2}^2}}{\displaystyle \frac{4h_1(R_1,R_2)}{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}\right)`$
$`{\displaystyle \frac{2\mu ^2m_b^2\mathrm{tan}\beta \mathrm{cot}\alpha }{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}({\displaystyle \frac{f_5(R_1)}{3M_{\stackrel{~}{b}_1}^2}}{\displaystyle \frac{f_5(R_2)}{3M_{\stackrel{~}{b}_2}^2}})+𝒪\left({\displaystyle \frac{m_bM_{EW}}{M_{SUSY}^2}}\right)\},`$ (4.7)
where $`R_iM_{\stackrel{~}{g}}/M_{\stackrel{~}{b}_i}`$ ($`i=1,2`$). The functions $`h_i(R_1,R_2)`$ and $`f_{1,5}(R_i)`$ arise from the expansions of the loop integrals and are given in the Appendix.
As in the case of maximal sbottom mixing, the first term in eq. 4.7 is zeroth order in $`M_{SUSY}`$. The remaining terms are of order $`M_{EW}^2/M_{SUSY}^2`$. As in the previous section, if we neglect all terms that are not enhanced by large $`\mathrm{tan}\beta `$ or $`\mathrm{cot}\alpha `$, we find that the dependence on $`A_b`$ drops out and $`\mathrm{\Delta }_{SQCD}`$ is again proportional to the product $`\mu M_{\stackrel{~}{g}}`$.
### 4.3 The approach to the decoupling limit
If we take all SUSY mass parameters large at fixed $`M_A`$ in eqs. 4.5 and 4.7, then $`\mathrm{\Delta }_{SQCD}`$ tends to a nonzero constant; i.e., the SQCD corrections do not decouple. However, we are especially interested in the case where both $`M_{SUSY}`$ and $`M_A`$ are large. In eqs. 4.5 and 4.7, the terms of zeroth order in $`M_{SUSY}`$ are proportional to $`\mathrm{tan}\beta +\mathrm{cot}\alpha `$. From eq. 2.8,
$$\mathrm{tan}\beta +\mathrm{cot}\alpha =\frac{2M_Z^2}{M_A^2}\mathrm{tan}\beta \mathrm{cos}2\beta +𝒪\left(\frac{M_{EW}^4}{M_A^4}\right).$$
(4.8)
Thus, the first term in eq. 4.5 and in eq. 4.7 is of order $`M_{EW}^2\mathrm{tan}\beta /M_A^2`$, and therefore decouples in the limit of large $`M_A`$. However, the approach to decoupling is delayed in the large $`\mathrm{tan}\beta `$ regime.<sup>4</sup><sup>4</sup>4The enhancement of the radiatively-corrected $`h^0b\overline{b}`$ coupling at large $`\mathrm{tan}\beta `$ has also been emphasized in refs. . Specifically, for values of $`M_A^2M_Z^2\mathrm{tan}\beta `$, we see that $`\mathrm{tan}\beta +\mathrm{cot}\alpha 𝒪(1)`$. For example, if $`\mathrm{tan}\beta 50`$, then even for values of $`M_A1`$ TeV, decoupling has not yet set in.
Other terms in eqs. 4.5 and 4.7 also exhibit delayed decoupling. In particular, eq. 4.8 implies that
$$Y_b=X_b+𝒪\left(\frac{M_{SUSY}M_{EW}^2\mathrm{tan}\beta }{M_A^2}\right),$$
(4.9)
so that $`Y_b`$ is also enhanced at large $`\mathrm{tan}\beta `$. Hence, all terms in eqs. 4.5 and 4.7 that are proportional to either $`X_b`$ or $`Y_b`$ are of order $`M_{EW}^2\mathrm{tan}\beta /M_{SUSY}^2`$. Again, if $`\mathrm{tan}\beta 50`$ and $`M_{SUSY}1`$ TeV, decoupling has not yet set in.
The remaining terms in eqs. 4.5 and 4.7 exhibit the expected decoupling in the usual sense (with no delay). In particular, we may set $`\alpha =\beta \pi /2`$ in the decoupling limit to obtain
$$\frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }=\mathrm{cos}2\beta +𝒪\left(\frac{M_{EW}^2}{M_A^2}\right),$$
(4.10)
which exhibits no $`\mathrm{tan}\beta `$ enhancement. All remaining factors of $`\mathrm{tan}\beta `$ are multiplied by the appropriate power of $`m_b`$, and since $`m_b\mathrm{tan}\beta M_{EW}`$, no delayed decoupling results from these terms.
We have thus shown analytically that the one-loop SQCD corrections to the $`h^0b\overline{b}`$ coupling decouple in the limit of large $`M_{SUSY}`$ and large $`M_A`$. The decoupling takes the generic form:
$$\mathrm{\Delta }_{SQCD}C_1\frac{M_{EW}^2}{M_A^2}+C_2\frac{M_{EW}^2}{M_{SUSY}^2}.$$
(4.11)
In general $`C_1`$ approaches a non-zero constant as $`M_{SUSY}\mathrm{}`$, while $`C_2`$ approaches a (different) non-zero constant as $`M_A\mathrm{}`$. Thus, the decoupling limit requires both $`M_A`$ and $`M_{SUSY}`$ to become simultaneously large (as compared to $`M_{EW}`$). However, we will demonstrate that in some cases the SQCD radiative corrections vanish in the limit where some SUSY particle masses are large, independent of the value of $`M_A`$.
This decoupling is shown numerically <sup>5</sup><sup>5</sup>5In our numerical analysis we have taken the $`b`$-quark pole-mass to be $`4.75`$ GeV and $`\alpha _s=0.119`$. Because of the experimental constraints on the sbottom masses, we consider only regions of parameter space in which both sbottoms are heavier than 100 GeV. in figs. 2 and 3.
In fig. 2, we plot the exact one-loop expression for $`\mathrm{\Delta }_{SQCD}`$ (solid lines) and the expansion of eq. 4.5 (dashed lines) for $`\mathrm{tan}\beta =50`$ and $`M_{SUSY}=M_L=M_R=M_S=M_{\stackrel{~}{g}}=\mu =A_b`$. The lines labeled (a) show $`\mathrm{\Delta }_{SQCD}`$ as a function of $`M_{SUSY}`$. We have fixed $`M_A`$ very large, $`M_A=3000`$ GeV, in order to eliminate the contribution to $`\mathrm{\Delta }_{SQCD}`$ that decouples at large $`M_A`$. We use the exact tree-level formula for $`\mathrm{cot}\alpha `$ as a function of $`M_A`$ and $`\mathrm{tan}\beta `$. The lines labeled (b) show $`\mathrm{\Delta }_{SQCD}`$ as a function of $`M_A`$, where now we have fixed $`M_{SUSY}`$ to be very large, $`M_{SUSY}=3000`$ GeV, in order to examine only the contribution to $`\mathrm{\Delta }_{SQCD}`$ that does not decouple at large $`M_{SUSY}`$. We note that for very large $`M_{SUSY}`$ and $`M_A=1`$ TeV, $`\mathrm{\Delta }_{SQCD}`$ is of order $`1\%`$ for $`\mathrm{tan}\beta =50`$. We have plotted our results for $`\mu M_{\stackrel{~}{g}}`$ positive. In the approximation of neglecting terms not enhanced by large $`\mathrm{tan}\beta `$ or $`\mathrm{cot}\alpha `$, changing the sign of $`\mu M_{\stackrel{~}{g}}`$ simply flips the sign of $`\mathrm{\Delta }_{SQCD}`$.
In fig. 3 we again plot the exact one-loop expression for $`\mathrm{\Delta }_{SQCD}`$ (solid lines) and the expansion of eq. 4.5 (dashed lines) for all the SUSY mass parameters equal, $`M_{SUSY}=M_L=M_R=M_S=M_{\stackrel{~}{g}}=\mu =A_b`$, and three values of $`\mathrm{tan}\beta `$<sup>6</sup><sup>6</sup>6Although we have chosen $`M_L=M_R`$ for simplicity, our results are not particularly sensitive to this choice as long as $`|M_L^2M_R^2|m_b|X_b|`$ (c.f. the remarks below eq. 4.5). Note the change in the vertical scale for the plots with different values of $`\mathrm{tan}\beta `$. We show the dependence of $`\mathrm{\Delta }_{SQCD}`$ on $`M_{SUSY}`$ (left-hand panels) and $`M_A`$ (right-hand panels). Clearly, in the limit of large $`M_{SUSY}`$, $`\mathrm{\Delta }_{SQCD}`$ tends to a non-vanishing constant, and this constant tends to zero in the large $`M_A`$ limit. Similarly, in the limit of large $`M_A`$, $`\mathrm{\Delta }_{SQCD}`$ tends to a non-vanishing constant, and this constant tends to zero in the large $`M_{SUSY}`$ limit.
Notice that from the numerical comparison between the exact and analytic formulae in fig. 3, we can conclude that our expansion is a good approximation for large enough SUSY mass parameters. In particular, it is reasonably accurate for $`M_{SUSY}`$ larger than 300 GeV. Also, it is clear that as $`\mathrm{tan}\beta `$ grows, not only does $`\mathrm{\Delta }_{SQCD}`$ increase in magnitude, but the agreement between the exact and analytic formulae becomes worse at low $`M_{SUSY}`$. This is due to the fact that the splitting between the squared masses of the two sbottoms in the maximal mixing case is proportional to $`m_b\mathrm{tan}\beta `$, which we have taken to be of order $`M_{EW}`$ in our expansion. As $`\mathrm{tan}\beta `$ increases, the mass of the lighter sbottom decreases, and the higher order terms that we have neglected in our expansion become more important.
All numerical results presented so far correspond to $`\mu M_{\stackrel{~}{g}}>0`$. In the case of $`\mu M_{\stackrel{~}{g}}<0`$, the qualitative features of $`|\mathrm{\Delta }_{SQCD}|`$ remain unchanged. From the analytic formulae derived in this section, one can see that at large $`\mathrm{tan}\beta `$ the dominant effect of changing the sign of $`\mu M_{\stackrel{~}{g}}`$ is to change the overall sign of $`\mathrm{\Delta }_{SQCD}`$. We can illustrate this point in the simple limiting case in which all SUSY mass parameters and $`M_A`$ are equal. Simplifying eq. 4.5 in this limit, we end up with a simple formula for the case of $`\mu M_{\stackrel{~}{g}}>0`$:
$`\mathrm{\Delta }_{SQCD}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{3\pi }}\{{\displaystyle \frac{M_Z^2}{3M_{SUSY}^2}}\mathrm{cos}2\beta (7\mathrm{tan}\beta 2)+{\displaystyle \frac{M_{h^0}^2}{12M_{SUSY}^2}}(\mathrm{tan}\beta 1)`$ (4.12)
$`+{\displaystyle \frac{m_b^2\mathrm{tan}^2\beta }{2M_{SUSY}^2}}(\mathrm{tan}\beta 4)+𝒪\left({\displaystyle \frac{m_bM_{EW}}{M_{SUSY}^2}}\right)\},`$
where $`M_{SUSY}=M_S=M_{\stackrel{~}{g}}=\mu =A_b=M_A`$. To obtain the result for $`\mu M_{\stackrel{~}{g}}<0`$, one replaces $`\mathrm{tan}\beta `$ with $`\mathrm{tan}\beta `$ in eq. 4.12. The formula of eq. 4.12 is plotted in fig. 4 for three values of $`\mathrm{tan}\beta `$ and both signs of $`\mu `$ (taking $`M_{\stackrel{~}{g}}`$ to be positive, by convention).
Clearly, $`\mathrm{\Delta }_{SQCD}`$ decouples like $`(M_{EW}^2/M_{SUSY}^2)`$, but this decoupling is delayed at large $`\mathrm{tan}\beta `$. For example, even at $`M_{SUSY}=1`$ TeV, $`|\mathrm{\Delta }_{SQCD}|1`$% for $`\mathrm{tan}\beta 30`$. Note that as expected, changing the sign of $`\mu `$ simply changes the sign of the dominant contribution to $`\mathrm{\Delta }_{SQCD}`$. In the remainder of our analysis, we will display results only for $`\mu >0`$.
Next, we consider the decoupling of the SQCD corrections to the $`h^0b\overline{b}`$ coupling as individual SUSY particles become heavy compared to the common SUSY mass scale. We examine three cases: large $`M_S`$ with maximal sbottom mixing, large $`M_{\stackrel{~}{g}}`$ with maximal sbottom mixing, and one heavy sbottom state with near-zero sbottom mixing.
We first consider the case of large $`M_S`$ with maximal sbottom mixing, with $`M_SM_{\stackrel{~}{g}}\mu A_bM_{EW}`$. If $`M_S`$ is taken large while the rest of the SUSY mass parameters remain fixed, then we may expand the functions $`f_i(R)`$ in eq. 4.5 for $`M_SM_{\stackrel{~}{g}}`$, or $`R1`$. The result is:
$$\mathrm{\Delta }_{SQCD}=\frac{\alpha _s}{3\pi }\left\{\frac{2\mu M_{\stackrel{~}{g}}}{M_S^2}(\mathrm{tan}\beta +\mathrm{cot}\alpha )+\frac{M_Z^2}{M_S^2}\frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }I_3^b+𝒪\left(\frac{M^4}{M_S^4}\right)\right\},$$
(4.13)
where $`M`$ is one of the lighter SUSY particle masses. Note that in this limit, the SQCD corrections decouple like $`M^2/M_S^2`$ even for light $`M_A`$. Thus it is only in the case of large $`M_{\stackrel{~}{g}}`$ and $`\mu `$, of the same order as $`M_S`$, that large $`M_A`$ is required for decoupling.
In fig. 5 we plot the exact one-loop expression for $`\mathrm{\Delta }_{SQCD}`$ and the expansions of eqs. 4.5 and 4.13 as a function of $`M_S`$, for fixed $`M_{\stackrel{~}{g}}=\mu =A_b=M_A=200`$ GeV and three different values of $`\mathrm{tan}\beta `$. This figure shows the decoupling of $`\mathrm{\Delta }_{SQCD}`$ with $`M_S`$ as discussed above.
Similarly we examine the case of a very heavy gluino compared to the rest of the SUSY spectrum. We still focus on the case of maximal sbottom mixing. Expanding the functions $`f_i(R)`$ in eq. 4.5 for $`M_{\stackrel{~}{g}}M_S`$, or $`R1`$, we see that in this case the SQCD corrections decouple with the gluino mass like $`M/M_{\stackrel{~}{g}}`$, where again $`M`$ is one of the other light SUSY masses:
$`\mathrm{\Delta }_{SQCD}={\displaystyle \frac{\alpha _s}{3\pi }}\{{\displaystyle \frac{2\mu }{M_{\stackrel{~}{g}}}}(\mathrm{tan}\beta +\mathrm{cot}\alpha )[1\mathrm{log}\left({\displaystyle \frac{M_{\stackrel{~}{g}}^2}{M_S^2}}\right)]{\displaystyle \frac{Y_b}{3M_{\stackrel{~}{g}}}}{\displaystyle \frac{M_{h^0}^2}{M_S^2}}`$ (4.14)
$`+{\displaystyle \frac{2X_b}{M_{\stackrel{~}{g}}}}{\displaystyle \frac{M_Z^2}{M_S^2}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}I_3^b{\displaystyle \frac{\mu ^2m_b^2\mathrm{tan}^2\beta }{M_{\stackrel{~}{g}}M_S^4}}(Y_b2A_b{\displaystyle \frac{\mathrm{cot}\alpha }{\mathrm{tan}\beta }})+𝒪\left({\displaystyle \frac{M^2}{M_{\stackrel{~}{g}}^2}}\right)\}.`$
Note that the decoupling of the SQCD corrections at large $`M_{\stackrel{~}{g}}`$ (with all other SUSY mass parameters held fixed) is very slow: $`\mathrm{\Delta }_{SQCD}`$ falls off only as one power of $`M_{\stackrel{~}{g}}`$. This is due to the factor of $`M_{\stackrel{~}{g}}`$ in the numerator of eqs. 3.6 and 3.8, which arises from the gluino propagator. $`\mathrm{\Delta }_{SQCD}`$ is also enhanced by the factor $`\mathrm{log}(M_{\stackrel{~}{g}}^2/M_S^2)`$. We again see the phenomenon of delayed decoupling at large $`\mathrm{tan}\beta `$ due to the terms in eq. 4.14 proportional to either $`X_b`$ or $`Y_b`$.
In fig. 6 we plot the exact one-loop expression for $`\mathrm{\Delta }_{SQCD}`$ and the expansions of eqs. 4.5 and 4.14 as a function of $`M_{\stackrel{~}{g}}`$, for $`M_S=\mu =A_b=M_A=300`$ GeV and three different values of $`\mathrm{tan}\beta `$. This figure shows the slow decoupling of $`\mathrm{\Delta }_{SQCD}`$ with $`M_{\stackrel{~}{g}}`$. For example, for $`M_{\stackrel{~}{g}}=500`$ GeV and $`\mathrm{tan}\beta =30`$, $`\mathrm{\Delta }_{SQCD}6\%`$ for $`M_S=\mu =A_b=M_A=300`$ GeV. If the latter masses are reduced to 200 GeV, one finds $`\mathrm{\Delta }_{SQCD}13\%`$, which is a significant correction. Fig. 6 also illustrates the validity of the large gluino mass expansion. This expansion is particularly poor for large values of $`\mathrm{tan}\beta `$ out to a very large gluino mass of about 2000 GeV.
Finally we study the case in which one of the sbottoms becomes heavy while the other sbottom mass and the rest of the SUSY mass parameters are fixed. We choose $`M_RM_LM_{\stackrel{~}{g}}\mu A_bM_{EW}`$, so that $`M_{\stackrel{~}{b}_2}M_{\stackrel{~}{b}_1}`$. This is necessarily the case of near-zero sbottom mixing. Expanding eq. 4.7 in inverse powers of $`M_{\stackrel{~}{b}_2}`$, we find:
$`\mathrm{\Delta }_{SQCD}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{3\pi }}\{{\displaystyle \frac{2}{3}}{\displaystyle \frac{M_Z^2}{M_{\stackrel{~}{b}_1}^2}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}(I_3^bQ_bs_W^2)f_5(R_1)`$ (4.15)
$`+`$ $`{\displaystyle \frac{2\mu M_{\stackrel{~}{g}}}{M_{\stackrel{~}{b}_2}^2}}(\mathrm{tan}\beta +\mathrm{cot}\alpha )\left[h(R_1)+\mathrm{log}\left({\displaystyle \frac{M_{\stackrel{~}{g}}^2}{M_{\stackrel{~}{b}_2}^2}}\right)\right]`$
$`+`$ $`{\displaystyle \frac{M_Z^2}{M_{\stackrel{~}{b}_2}^2}}{\displaystyle \frac{\mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )}{\mathrm{sin}\alpha }}\left[(I_3^bQ_bs_W^2){\displaystyle \frac{2M_{\stackrel{~}{g}}X_b}{M_{\stackrel{~}{b}_1}^2}}f_1(R_1)+Q_bs_W^2\right]`$
$`+`$ $`{\displaystyle \frac{2}{3}}{\displaystyle \frac{\mu ^2m_b^2\mathrm{tan}\beta \mathrm{cot}\alpha }{M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2}}f_5(R_1)+𝒪\left({\displaystyle \frac{M^4}{M_{\stackrel{~}{b}_2}^4}}\right)\},`$
where again $`M`$ is one of the other light SUSY masses and the function $`h(R_1)`$ is given in the Appendix. Note that the first term does not decouple as $`M_{\stackrel{~}{b}_2}`$ is taken large. This behavior is independent of the value of $`M_A`$ (and therefore holds even if $`M_A\mathrm{}`$). However, this first term is not enhanced by large $`\mathrm{tan}\beta `$ and is numerically negligible as can be seen in fig. 7. The terms that are enhanced by large $`\mathrm{tan}\beta `$ decouple like $`M^2/M_{\stackrel{~}{b}_2}^2`$. In fig. 7 we plot the exact one-loop expression for $`\mathrm{\Delta }_{SQCD}`$ and the expansions of eqs. 4.7 and 4.15, as a function of $`M_{\stackrel{~}{b}_2}`$, for $`M_{\stackrel{~}{b}_1}=M_{\stackrel{~}{g}}=\mu =A_b=M_A=200`$ GeV and three different values of $`\mathrm{tan}\beta `$. Clearly, in order for $`\mathrm{\Delta }_{\mathrm{SQCD}}`$ to be large in the case of near-zero sbottom mixing, both of the sbottoms must be reasonably light. Note however that, due to the enhancement in $`\mathrm{tan}\beta `$, the $`1/M_{\stackrel{~}{b}_2}^2`$ suppression is not so small. As an example, for $`\mathrm{tan}\beta =50`$, $`M_{\stackrel{~}{b}_1}=M_{\stackrel{~}{g}}=\mu =A_b=M_A=200`$ GeV and $`M_{\stackrel{~}{b}_2}=500`$ GeV \[1000 GeV\], one obtains $`\mathrm{\Delta }_{\mathrm{SQCD}}10\%`$ \[$`5\%`$\].
The various cases examined in this section can be summarized by specifying the behavior of $`C_1`$ and $`C_2`$ of eq. 4.11 on the model parameters. In Table 1, four cases are shown. In all cases, $`M_{SUSY}`$ is identified with the largest supersymmetry-breaking mass, while $`M`$ refers to a possible intermediate supersymmetric mass scale (with $`M_{EW}MM_{SUSY}`$). The presence of a factor of $`\mathrm{tan}\beta `$ (unless multiplied by $`M/M_{SUSY}`$) indicates delayed decoupling. In the case of $`M_{\stackrel{~}{g}}=M_{SUSY}`$, $`C_2(M_{SUSY}/M)\mathrm{tan}\beta `$ implies a delayed decoupling that vanishes only as a single power of $`1/M_{SUSY}`$. Finally, in the case of large $`M_{\stackrel{~}{b}_2}`$, $`C_2M_{SUSY}^2/M^2`$ implies no decoupling as $`M_{SUSY}\mathrm{}`$ with $`M`$ held fixed. This is not a violation of the usual decoupling theorem , since in the latter case, only part of the supersymmetric spectrum has been removed from the low-energy effective theory. Decoupling is recovered in the limit of $`M\mathrm{}`$, as expected.
## 5 Conclusions
In this paper we have studied the one loop SQCD corrections to the $`h^0b\overline{b}`$ coupling in the limit of large SUSY masses. In order to understand analytically the behavior of the corrections in this limit, we have performed expansions for the SUSY mass parameters large compared to the electroweak scale. We have shown that for the SUSY mass parameters and $`M_A`$ large and all of the same order, the SQCD corrections decouple like the inverse square of these mass parameters. However, if the mass parameters are not all of the same size, then this behavior can be modified. If $`M_A`$ is light, then the SQCD corrections to the $`h^0b\overline{b}`$ coupling generically do not decouple in the limit of large SUSY mass parameters. In this case, the low-energy theory at the electroweak scale contains two full Higgs doublets with Higgs-fermion couplings of the general type-III model.
We have also examined three cases in which there is a hierarchy among the SUSY mass parameters. In the case of maximal sbottom mixing with $`M_S`$ large and the other SUSY mass parameters and $`M_A`$ of order a common mass scale $`M`$ (chosen such that $`M_{EW}MM_S`$), the SQCD corrections decouple like $`M^2/M_S^2`$. Second, we examined the case of a large gluino mass with the other SUSY mass parameters of order a common mass scale $`M`$ (chosen such that $`M_{EW}MM_{\stackrel{~}{g}}`$). In this case we found that the SQCD corrections decouple more slowly, like $`(M/M_{\stackrel{~}{g}})\mathrm{log}(M_{\stackrel{~}{g}}^2/M_S^2)`$. Finally, we examined the case in which one sbottom is much heavier than the other SUSY mass parameters, which are fixed at a scale $`M`$. In this case the mixing angle in the sbottom sector is near zero. We found that the piece of the SQCD corrections that is enhanced at large $`\mathrm{tan}\beta `$ decouples like $`M^2/M_{\stackrel{~}{b}_2}^2`$. There is also a piece of the SQCD corrections that does not decouple as $`M_{\stackrel{~}{b}_2}`$ is taken large, but it is not enhanced by $`\mathrm{tan}\beta `$ and is numerically negligible compared to the decoupling piece, up to a very high value of the heavier sbottom mass.
The decoupling behavior of the SQCD corrections to the $`h^0b\overline{b}`$ coupling implies that distinguishing the lightest MSSM Higgs boson from the SM Higgs boson will be very difficult if $`A^0`$ and the SUSY spectrum are heavy, even after one-loop SUSY corrections are taken into account. However, because of the enhancement at large $`\mathrm{tan}\beta `$, the onset of decoupling is delayed, and the corrections can still be at the percent level for $`\mathrm{tan}\beta 50`$ and all SUSY mass parameters and $`M_A`$ of order 1 TeV. If one or both of the sbottoms, the gluino, and/or $`A^0`$ lie below the TeV scale, then the SQCD corrections will be larger still. The decoupling limit provides a challenge for Higgs searches at future colliders. Even if the light CP-even Higgs boson is found, the direct discovery of supersymmetric particles may be essential for unraveling the origin of electroweak symmetry breaking.
Acknowledgments
We wish to thank M. Carena, K. Matchev, C. E. M. Wagner and G. Weiglein for interesting discussions. M.H., S.P., S.R. and D.T. kindly thank T. Hahn for providing the code of LoopTools (which was used in the calculation) and for many helpful suggestions. H.L. is grateful to G. Kribs for a discussion on models of SUSY breaking.
This work has been supported in part by the Spanish Ministerio de Educacion y Cultura under project CICYT AEN97-1678. S.R. has been partially supported by the European Union through contract ERBFMBICT972474. H.E.H. is supported in part by the U.S. Department of Energy under contract DE-FG03-92ER40689. Fermilab is operated by Universities Research Association Inc. under contract no. DE-AC02-76CH03000 with the U.S. Department of Energy.
Appendix
## Appendix A Expansions of loop functions
In this Appendix we define our notation for the two- and three-point integrals that appear in eqs. 3.6 and 3.8 and give formulae for their expansions in powers of the SUSY mass parameters.
We follow the definitions and conventions of . The two-point integrals are given by:
$$\mu ^{4D}\frac{d^Dk}{(2\pi )^D}\frac{\{1;k^\mu \}}{[k^2m_1^2][(k+q)^2m_2^2]}=\frac{i}{16\pi ^2}\{B_0;q^\mu B_1\}(q^2;m_1^2,m_2^2).$$
(A.1)
The derivatives of the two-point functions are defined as follows:
$$B_{0,1}^{}(p^2;m_1^2,m_2^2)=\frac{}{q^2}B_{0,1}(q^2;m_1^2,m_2^2)|_{q^2=p^2}.$$
(A.2)
Finally, the three-point integrals are given by:
$`\mu ^{4D}`$ $`{\displaystyle \frac{d^Dk}{(2\pi )^D}\frac{\{1;k^\mu \}}{[k^2m_1^2][(k+p_1)^2m_2^2][(k+p_1+p_2)^2m_3^2]}}`$ (A.3)
$`={\displaystyle \frac{i}{16\pi ^2}}\{C_0;p_1^\mu C_{11}+p_2^\mu C_{12}\}(p_1^2,p_2^2,p^2;m_1^2,m_2^2,m_3^2),`$
where $`p=p_1p_2`$.
We now give the large $`M_{SUSY}`$ expansions of the loop integrals.
### A.1 Maximal $`\stackrel{~}{b}_L\stackrel{~}{b}_R`$ mixing
The loop integrals are expanded as follows, where $`M_S^2`$ and $`\mathrm{\Delta }^2`$ are defined in eq. 4.3.
$`C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_1}^2)`$
$`{\displaystyle \frac{1}{2M_S^2}}f_1(R)+{\displaystyle \frac{\mathrm{\Delta }^2}{3M_S^4}}f_2(R){\displaystyle \frac{\mathrm{\Delta }^4}{4M_S^6}}f_3(R){\displaystyle \frac{M_{h^0}^2}{24M_S^4}}f_4(R)`$
$`C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2,M_{\stackrel{~}{b}_2}^2)`$
$`{\displaystyle \frac{1}{2M_S^2}}f_1(R){\displaystyle \frac{\mathrm{\Delta }^2}{3M_S^4}}f_2(R){\displaystyle \frac{\mathrm{\Delta }^4}{4M_S^6}}f_3(R){\displaystyle \frac{M_{h^0}^2}{24M_S^4}}f_4(R)`$
$`C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_2}^2)`$
$`{\displaystyle \frac{1}{2M_S^2}}f_1(R){\displaystyle \frac{\mathrm{\Delta }^4}{12M_S^6}}f_3(R){\displaystyle \frac{M_{h^0}^2}{24M_S^4}}f_4(R)`$
$`C_{11}(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_1}^2)`$
$`{\displaystyle \frac{1}{3M_S^2}}f_5(R){\displaystyle \frac{\mathrm{\Delta }^2}{4M_S^4}}f_4(R)+{\displaystyle \frac{\mathrm{\Delta }^4}{5M_S^6}}f_6(R)+{\displaystyle \frac{M_{h^0}^2}{30M_S^4}}f_7(R)`$
$`C_{11}(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2,M_{\stackrel{~}{b}_2}^2)`$
$`{\displaystyle \frac{1}{3M_S^2}}f_5(R)+{\displaystyle \frac{\mathrm{\Delta }^2}{4M_S^4}}f_4(R)+{\displaystyle \frac{\mathrm{\Delta }^4}{5M_S^6}}f_6(R)+{\displaystyle \frac{M_{h^0}^2}{30M_S^4}}f_7(R)`$
$`B_0(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)B_0(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{\mathrm{\Delta }^2}{M_S^2}}f_1(R)`$
$`B_0^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)B_0^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{\mathrm{\Delta }^2}{6M_S^4}}f_8(R)`$
$`B_1^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)+B_1^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{1}{6M_S^2}}f_4(R){\displaystyle \frac{\mathrm{\Delta }^4}{15M_S^6}}f_9(R).`$ (A.4)
The functions $`f_i(R)`$ are given in terms of the ratio $`RM_{\stackrel{~}{g}}/M_S`$:
$`f_1(R)`$ $`=`$ $`{\displaystyle \frac{2}{(1R^2)^2}}\left[1R^2+R^2\mathrm{log}R^2\right]`$
$`f_2(R)`$ $`=`$ $`{\displaystyle \frac{3}{(1R^2)^3}}\left[1R^4+2R^2\mathrm{log}R^2\right]`$
$`f_3(R)`$ $`=`$ $`{\displaystyle \frac{4}{(1R^2)^4}}\left[1+\frac{3}{2}R^23R^4+\frac{1}{2}R^6+3R^2\mathrm{log}R^2\right]`$
$`f_4(R)`$ $`=`$ $`{\displaystyle \frac{4}{(1R^2)^4}}\left[\frac{1}{2}3R^2+\frac{3}{2}R^4+R^63R^4\mathrm{log}R^2\right]`$
$`f_5(R)`$ $`=`$ $`{\displaystyle \frac{3}{(1R^2)^3}}\left[\frac{1}{2}2R^2+\frac{3}{2}R^4R^4\mathrm{log}R^2\right]`$
$`f_6(R)`$ $`=`$ $`{\displaystyle \frac{5}{(1R^2)^5}}\left[\frac{1}{2}4R^2+4R^6\frac{1}{2}R^86R^4\mathrm{log}R^2\right]`$
$`f_7(R)`$ $`=`$ $`{\displaystyle \frac{5}{(1R^2)^5}}\left[\frac{1}{3}2R^2+6R^4\frac{10}{3}R^6R^8+4R^6\mathrm{log}R^2\right]`$
$`f_8(R)`$ $`=`$ $`{\displaystyle \frac{12}{(1R^2)^4}}\left[\frac{1}{2}+2R^2\frac{5}{2}R^4+2R^2\mathrm{log}R^2+R^4\mathrm{log}R^2\right]`$
$`f_9(R)`$ $`=`$ $`{\displaystyle \frac{5}{(1R^2)^6}}[112R^236R^4+44R^6+3R^8`$ (A.5)
$`24R^6\mathrm{log}R^236R^4\mathrm{log}R^2].`$
Note that in the special case $`M_{\stackrel{~}{g}}=M_S`$, $`R=1`$ and $`f_i(1)=1`$.
### A.2 Near-zero $`\stackrel{~}{b}_L\stackrel{~}{b}_R`$ mixing
The loop integrals are expanded as follows:
$`C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_1}^2){\displaystyle \frac{1}{2M_{\stackrel{~}{b}_1}^2}}f_1(R_1){\displaystyle \frac{M_{h^0}^2}{24M_{\stackrel{~}{b}_1}^4}}f_4(R_1)`$
$`C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{1}{2M_{\stackrel{~}{b}_2}^2}}f_1(R_2){\displaystyle \frac{M_{h^0}^2}{24M_{\stackrel{~}{b}_2}^4}}f_4(R_2)`$
$`C_0(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{h_1(R_1,R_2)}{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)}}+{\displaystyle \frac{M_{h^0}^2h_2(R_1,R_2)}{(M_{\stackrel{~}{b}_1}^2M_{\stackrel{~}{b}_2}^2)^2}}`$
$`C_{11}(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2,M_{\stackrel{~}{b}_1}^2){\displaystyle \frac{1}{3M_{\stackrel{~}{b}_1}^2}}f_5(R_1)+{\displaystyle \frac{M_{h^0}^2}{30M_{\stackrel{~}{b}_1}^4}}f_7(R_1)`$
$`C_{11}(m_b^2,M_{h^0}^2,m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{1}{3M_{\stackrel{~}{b}_2}^2}}f_5(R_2)+{\displaystyle \frac{M_{h^0}^2}{30M_{\stackrel{~}{b}_2}^4}}f_7(R_1)`$
$`B_0(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)B_0(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2)h_1(R_1,R_2)`$
$`B_0^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)B_0^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{1}{6M_{\stackrel{~}{b}_1}^2}}f_2(R_1){\displaystyle \frac{1}{6M_{\stackrel{~}{b}_2}^2}}f_2(R_2)`$
$`B_1^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_1}^2)+B_1^{}(m_b^2;M_{\stackrel{~}{g}}^2,M_{\stackrel{~}{b}_2}^2){\displaystyle \frac{1}{12M_{\stackrel{~}{b}_1}^2}}f_4(R_1){\displaystyle \frac{1}{12M_{\stackrel{~}{b}_2}^2}}f_4(R_2),`$ (A.6)
where $`R_iM_{\stackrel{~}{g}}/M_{\stackrel{~}{b}_i}`$ ($`i=1,2`$). The functions $`f_i(R)`$ were given in eq. A.1. The functions $`h_1(R_1,R_2)`$ and $`h_2(R_1,R_2)`$ are defined as follows:
$`h_1(R_1,R_2)`$ $`=`$ $`h(R_1)h(R_2),\mathrm{with}h(R)={\displaystyle \frac{\mathrm{log}R^2}{1R^2}},`$
$`h_2(R_1,R_2)`$ $`=`$ $`1+{\displaystyle \frac{R_1^2+R_2^22R_1^2R_2^2}{2(1R_1^2)(1R_2^2)}}`$ (A.7)
$``$ $`{\displaystyle \frac{1}{2(R_1^2R_2^2)}}[{\displaystyle \frac{\mathrm{log}R_1^2}{(1R_1^2)^2}}(R_1^2+R_2^22R_1^4)`$
$`{\displaystyle \frac{\mathrm{log}R_2^2}{(1R_2^2)^2}}(R_1^2+R_2^22R_2^4)].`$
The functions $`h`$ and $`h_2`$ have the following properties:
$`h(1)`$ $`=`$ $`1,`$
$`h_2(R_1,R_2)`$ $`=`$ $`h_2(R_2,R_1),`$
$`h_2(1,R_2)`$ $`=`$ $`{\displaystyle \frac{1}{(1R_2^2)^2}}\left[\frac{5}{4}R_2^2\frac{1}{4}R_2^4+\left(\frac{1}{2}+R_2^2\right)\mathrm{log}R_2^2\right].`$ (A.8)
|
warning/0007/hep-th0007192.html
|
ar5iv
|
text
|
# UFIFT-HEP-00-18 hep-th/0007192 A Note on Non-Locality and Ostrogradski’s Construction
## 1 Motivations
Woodard’s construction has boundaries in the non-local integration range at $`x=0`$ and $`x=\mathrm{\Delta }t`$. If caution is not applied, a breakdown of the Jacobi identity and a loss of differentiability of functionals may lead to inconsistencies. As mentioned in , additional boundary conditions must be imposed. Intuitively, the boundaries of the non-local integration range should not have any physical significance. We shall devise a slightly modified construction that avoids boundaries and hence steers clear of such hazards.
In addition, we are interested in the close relationship between Ostrogradski’s derivatives and the higher Euler-Lagrange derivatives used among other things to construct boundary Poisson brackets, cf. .
Compared with earlier work, we do not have second class constraints. We obtain a clear separation of the equations of motion from the constraints, which is preferable at the conceptual level.
We shall have nothing to say about the soundness of non-local theories and higher derivative theories in general. Instead we refer to the ongoing discussion in the Literature .
## 2 Lagrangian Variables
For simplicity, we shall consider a $`0+1`$ dimensional systems with one dynamical variable $`q(t)`$. The construction can be straightforwardly extended to include more variables and to field theories. We shall assume that the time coordinate $`t`$ has no temporal boundaries, $`i.e.`$$`tIR`$ can take any real value.
We are free to think of the dynamical variable $`q(t)`$ as a $`1+1`$ dimensional field $`Q(x,t)`$ that satisfies a chirality condition
$$\frac{d}{dt}Q(x,t)=_xQ(x,t).$$
(2.1)
Explicitly, the one-to-one correspondence between $`q(t)`$ and the left-mover $`Q(x,t)`$ is given by $`Q(x,t)=q(x+t)`$. Keep in mind for later that eq. (2.1) implies
$$\left(\frac{d}{dt}\right)^nQ(x,t)=\left(_x\right)^nQ(x,t),nIN_0.$$
(2.2)
## 3 Lagrangian
By non-locality we mean that the Lagrangian $`L[q](t)`$ depends on the dynamical variable $`q`$ at other times than $`t`$. To deal with this in a systematic manner we shall assume that the Lagrangian can be written as a $`d`$-dimensional integral
$$L[Q](t)=_{\mathrm{}}^{\mathrm{}}𝑑x^1\mathrm{}_{\mathrm{}}^{\mathrm{}}𝑑x^d(x^1,\mathrm{},x^d,t)$$
(3.1)
over a density function $``$. To be precise, besides the explicit dependence of $`x^1,\mathrm{},x^d`$ and $`t`$, the density function $`(x^1,\mathrm{},x^d,t)`$ is assumed to be a function of a finite number of the following variables:
$$(_x)^kQ(x^i,t),i=1,\mathrm{},d,,kIN_0.$$
(3.2)
The replacement of $`q(t)`$ with $`Q(x,t)`$ has several advantages:
First of all, $`L[Q](t)`$ does only depend on $`Q(x,t)`$’s of the very same $`t`$. The non-locality is encoded in the new variable $`x`$. Negative and positive values of $`x`$ correspond to interactions with the past and the future, respectively. (We should stress that $`x`$ has nothing to do with space; we have merely named the new variable $`x`$ because the formulas fit the framework of field theory.)
Secondly, in the $`Q`$-formulation we have removed all derivatives wrt. $`t`$ which appeared in the original Lagrangian $`L[q](t)`$ by using the chirality condition (2.2). This will prepare us for a very smooth transition into its Hamiltonian counterpart, $`i.e.`$the Lagrangian does not depend on the velocities, on the accelerations, etc. and hence on the momenta. (Note that in the Hamiltonian formulation our starting point will be $`L[Q](t)`$ without assuming chirality of $`Q(x,t)`$. We shall later see that in the Hamiltonian formulation the chirality condition (2.1) becomes the equation of motion for $`Q(x,t)`$.)
## 4 Local Field Theory
A functional (3.1) with the assumption that its density $`(x^1,\mathrm{},x^d,t)`$ depends on a finite number of variables from the list (3.2) is commonly known as the very definition of a local functional. We may say that the original non-local theory has become a local field theory.
The case of discrete non-locality has been studied extensively in the Literature . We define it as the case where there exists a discrete family of curves $`\{tx_i(t)\}_{iI}`$ so that $`L[Q](t)`$ only depends on a finite number of the following variable: $`t`$, $`x_i(t)`$ and
$$(_x)^kQ(x_i(t),t),iI,kIN_0.$$
(4.1)
For technical reasons we shall assume the function $`(x^1,\mathrm{},x^d,t)`$ is $`C^{\mathrm{}}`$-differentiable. This smoothness assumption unfortunately rules out the case of discrete non-locality. However, we shall investigate the discrete case in Section 13.
## 5 Compact Support
We assume that the density function $``$ (and other physically meaningful objects) has a compact support in the $`x`$-directions. As we shall see this assumption has radical implications for the theory.
## 6 Momenta
In a non-local Lagrangian theory the usual definition of momenta as the derivatives of the Lagrangian wrt. the velocities is not useful. Instead, we shall seek a new and better definition. We take as our initial guess the partial differential equation
$$\frac{d}{dt}P(x,t)=_xP(x,t)+\frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)},$$
(6.1)
subjected to the following boundary condition:
$$P(x,t)\mathrm{has}\mathrm{compact}\mathrm{support}\mathrm{in}\mathrm{the}x\mathrm{direction}.$$
(6.2)
The $`(0)`$ appearing on the symbol for the functional derivatives denotes that we use the algebraic Euler-Lagrange definition rather than a variational definition of the functional derivatives, cf. . We shall see below that our initial guess (6.1) needs off-shell modifications.
Both the partial differential equation (6.1) and the boundary condition (6.2) appear naturally in the Hamiltonian treatment to be given below. Purely from a Lagrangian perspective, the boundary condition (6.2) arises as a natural consequence of requiring the density function $``$ itself to have compact support.
In general the above boundary value problem (6.1) and (6.2) may not have continuous solutions. Let us allow for a potential discontinuity along a curve $`x=x_0(t)`$. The unique solution is then
$`P(x,t)`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x^{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑t^{}\delta _{IR}(x^{}+t^{}xt)\left[\theta (x^{}x)\theta (xx_0(t))\theta (x_0(t)x)\theta (xx^{})\right]{\displaystyle \frac{\delta L[Q](t^{})}{\delta Q^{(0)}(x^{},t^{})}}`$ (6.3)
$`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑u\left[\theta (xx_0(t))\theta (u)\theta (x_0(t)x)\theta (u)\right]{\displaystyle \frac{\delta L[Q](tu)}{\delta Q^{(0)}(x+u,tu)}}.`$ (6.5)
In a Lagrangian treatment we promote this formula to be the very definition of momenta (see \[1, formula (20)\] and \[2, formula (6)\] ). Needless to say, formula (6.5) may have other discontinuities if $`\delta L[Q](t)/\delta Q^{(0)}(x,t)`$ is singular, cf. Section 13.
## 7 Lagrangian Equation of Motion
Let $`S[q]=_{\mathrm{}}^{\mathrm{}}𝑑tL[q](t)`$ denote the action. Recall that we cannot vary $`Q`$ freely because of the chirality condition eq. (2.1). Therefore, the Lagrangian Equation of Motion for $`Q`$ does not provide us with the relevant physical information. Instead, the pertinent equation of motion is given by the Lagrangian Equation of Motion for $`q`$:
$`0`$ $`=`$ $`{\displaystyle \frac{\delta S[q]}{\delta q^{(0)}(x+t)}}\stackrel{(\text{2.1})}{=}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑u{\displaystyle \frac{\delta L[Q](tu)}{\delta Q^{(0)}(x+u,tu)}}`$ (7.1)
$``$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x^{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑t^{}\delta _{IR}(x^{}+t^{}xt){\displaystyle \frac{\delta L[Q](t^{})}{\delta Q^{(0)}(x^{},t^{})}}.`$ (7.3)
In a $`Q`$-formulation, with the chirality condition eq. (2.1) not imposed manifestly, the Equation of Motion for $`q`$ does strictly speaking not make sense (at least not a priori). However, even in that case, it is natural to call the last two expressions in eq. (7.3) the Lagrangian Equation of Motion for $`q`$.
Note that eq. (7.3) is precisely the condition that the momentum formula (6.5) does not have an extra discontinuity at $`x=x_0(t)`$:
$$\underset{xx_0(t)^{}}{lim}P(x,t)=\underset{xx_0(t)^+}{lim}P(x,t).$$
(7.4)
From a Hamiltonian point of view, where the momentum is a fundamental rather than a derived quantity, this is of course trivially guaranteed by restricting ourselves to continuous fields. (In fact we shall only allow $`C^{\mathrm{}}`$-fields.)
Closely related to this fact is the following: If we include singular terms at the kink curve, the above momentum formula (6.5) satisfies the following partial differential equation:
$$\frac{d}{dt}P(x,t)=_xP(x,t)+\frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}+\left(\dot{x}_0(t)1\right)\delta _{IR}(xx_0(t))\frac{\delta S[q]}{\delta q^{(0)}(x+t)}.$$
(7.5)
This differs from the original eq. (6.1) with a delta function contribution along the curve that is proportional to the Lagrangian Equation of Motion. The extra term also vanishes along equal-time curves $`x_0(t)+t=\mathrm{constant}`$.
## 8 Gauge Symmetry
We will now discuss the Hamiltonian formulation. As mention earlier, our starting point is the Lagrangian $`L[Q](t)`$ without assuming chirality of $`Q(x,t)`$. The newly gained freedom of the $`Q(x,t)`$-fields introduces a gauge symmetry for the the Lagrangian $`L[Q](t)`$ in the following way; for a given time $`t`$, let (for the time being) $`\mathrm{\Sigma }_t`$ denote the support
$$\mathrm{supp}\left(x\frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}\right)\overline{\left\{xIR\right|\frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}0\}}$$
(8.1)
of the Euler-Lagrange function
$$x\frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}.$$
(8.2)
It follows from previously made assumptions that $`\mathrm{\Sigma }_t`$ is compact. The Lagrangian $`L[Q](t)`$ will be invariant under all transformations $`\delta Q(x,t)`$ that leaves $`\mathrm{\Sigma }_t`$ invariant:
$$x\mathrm{\Sigma }_t:\delta Q(x,t)=0.$$
(8.3)
The value of the field $`Q(x,t)`$ for $`x\mathrm{\Sigma }_t`$ has no physical content. It represents a gauge degree of freedom for the system.
## 9 Boundary Poisson Bracket
We take the boundary Poisson bracket for local functionals $`F(t)`$ and $`G(t)`$ to be given by the following ultra-local ansatz :
$$\{F(t),G(t)\}=\underset{k,\mathrm{}=0}{\overset{\mathrm{}}{}}c_{k,\mathrm{}}_{\mathrm{}}^{\mathrm{}}dx(_x)^{k+\mathrm{}}\left[\frac{\delta F(t)}{\delta Q^{(k)}(x,t)}\frac{\delta G(t)}{\delta P^{(\mathrm{})}(x,t)}\right](FG).$$
(9.1)
Here $`\delta /\delta Q^{(k)}(x,t)`$ are the higher Euler-Lagrange derivatives, cf. , and the coefficients $`c_{k,\mathrm{}}`$ are constants. They are normalized such that $`c_{0,0}=1`$ and such that the Jacobi identity is satisfied. In particular, one may show that
$$c_{k,0}=1=c_{0,\mathrm{}}.$$
(9.2)
We can extract the usual canonical equal-$`t`$ relation from (9.1):
$$\{Q(x,t),P(x^{},t)\}=\delta _{IR}(xx^{}).$$
(9.3)
## 10 First Class Constraints
The gauge symmetry is generated by the following first class contraints
$$x\mathrm{\Sigma }_t:P(x,t)0,$$
(10.1)
which is the Hamiltonian version of the boundary condition (6.2). ( The wavy double line $``$ is a notation first introduced by Dirac to denote equality modulo first class constraints, so-called weak equality.) To check in detail that the first class constraint (10.1) generates the gauge transformations (8.3), consider the smeared first class constraint
$$T[\xi ](t)_{\mathrm{}}^{\mathrm{}}𝑑x\xi (x,t)P(x,t),$$
(10.2)
where $`\xi (x,t)`$ is a test function that vanishes on $`\mathrm{\Sigma }_t`$ :
$$x\mathrm{\Sigma }_t:\xi (x,t)=0.$$
(10.3)
Here $`\xi (x,t)`$ does not depend on the dynamical variables $`P(x,t)`$ and $`Q(x,t)`$. Let $`\delta Q=\xi `$ be a infinitesimal gauge transformation. The gauge variation of the local functional $`F[Q,P](t)`$ is given by
$`\delta _\xi F[Q,P](t)`$ $``$ $`F[Q+\xi ,P](t)F[Q,P](t)`$ (10.4)
$`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x(_x)^k\left[{\displaystyle \frac{\delta F[Q,P](t)}{\delta Q^{(k)}(x,t)}}\xi (x,t)\right]`$ (10.6)
$`=`$ $`\{F[Q,P](t),T[\xi ](t)\}.`$ (10.8)
In the last equality, we used (9.2).
## 11 Hamiltonian
The bare action $`S`$ and the bare Hamiltonian $`H(t)`$ is given by
$`S`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑t\left[{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑xP(x,t)\dot{Q}(x,t)H(t)\right]`$ (11.1)
$`H(t)`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑xP(x,t)_xQ(x,t)L[Q](t).`$ (11.3)
The Hamiltonian Equation of Motion
$$\dot{F}(t)\{F(t),H(t)\}$$
(11.4)
for a local functional $`F(t)`$ becomes
$`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x(_x)^k\left[{\displaystyle \frac{\delta F(t)}{\delta Q^{(k)}(x,t)}}\dot{Q}(x,t)+{\displaystyle \frac{\delta F(t)}{\delta P^{(k)}(x,t)}}\dot{P}(x,t)\right]`$ (11.5)
$``$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x(_x)^k\left[{\displaystyle \frac{\delta F(t)}{\delta Q^{(k)}(x,t)}}_xQ(x,t)+{\displaystyle \frac{\delta F(t)}{\delta P^{(k)}(x,t)}}\left(_xP(x,t)+{\displaystyle \frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}}\right)\right]`$ (11.9)
$`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}c_{1,k}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x(_x)^{k+1}\left[{\displaystyle \frac{\delta F(t)}{\delta P^{(k)}(x,t)}}P(x,t)\right],`$
where we have used (9.2) and the fact that $``$ has compact support. If we furthermore apply the first class constraints (10.1), we can get rid of the last term on the right hand side. The remaining equation is clearly equivalent to the chirality condition eq. (2.1) for $`Q(x,t)`$ and eq. (6.1) for $`P(x,t)`$.
One may indeed check that the first class constraint (10.1) is preserved under the Hamiltonian flow, as it should be:
$`\{T[\xi ](t),H(t)\}`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x\xi (x,t)\left(_xP(x,t)+{\displaystyle \frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}}\right)0,`$ (11.10)
$`\delta _\xi S`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑t\left[{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x\xi (x,t)\dot{P}(x,t)+\{T[\xi ](t),H(t)\}\right]0.`$ (11.12)
Finally, an important remark: We could generalize the previous discussion by letting $`\mathrm{\Sigma }_t`$ be any compact set which contains the support (8.1), thereby deliberately choosing a smaller gauge symmetry. In fact, a transformation $`\delta Q(x,t)`$ which leaves the support (8.1) – but not the larger set $`\mathrm{\Sigma }_t`$ – invariant, would no longer be a symmetry for the action $`S`$ nor the Hamiltonian $`H(t)`$. The first class constraint (10.1) is creating its own justification! This is because the original Lagrangian theory – build out of rigid left-movers – does not possess any gauge symmetry at all. We see that the theory changes with the choice of $`\mathrm{\Sigma }_t`$. The smaller we choose $`\mathrm{\Sigma }_t`$, the more the system is prohibited in sending left-moving momenta between different connected components of the support (8.1). On the other hand, we do not want to completely eliminate the gauge symmetry by choosing $`\mathrm{\Sigma }_t`$ as large as possible, $`i.e.`$$`\mathrm{\Sigma }_t=IR`$. That would complicate matters by activating the extra boundary terms appearing in the equations of motion (11.9). In conclusion, to ensure that our Hamiltonian system corresponds to the original Lagrangian theory, we let $`\mathrm{\Sigma }_t`$ be a compact set bigger than the convex hull of the support (8.1).
## 12 Ostrogradski’s Framework
Let us recall how the non-local formulation translates into Ostrogradski’s formulation of infinite order. For other treatments, see \[5, Appendix A\], \[4, Section VI A\] and \[1, Section 5\]. We assume for simplicity that the discontinuity curve $`x_0(t)=x_0`$ is constant. Ostrogradski’s coordinates $`Q^{(n)}(t)`$, $`nIN_0`$, are defined as
$$Q^{(n)}(t)=(_x)^nQ(x,t)|_{x=x_0}.$$
(12.1)
The inverse relation is given by the Taylor expansion around $`x=x_0`$:
$$Q(x,t)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{(xx_0)^n}{n!}Q^{(n)}(t).$$
(12.2)
The chirality condition eq. (2.1) translates into
$$\dot{Q}^{(n)}(t)=Q^{(n+1)}(t).$$
(12.3)
To ensure that the boundary Poisson bracket eq. (9.1) corresponds to the discrete analogue given by
$$\{Q^{(n)}(t),P_{(m)}(t)\}=\delta _m^n,$$
(12.4)
we define Ostrogradski’s momenta $`P_{(n)}(t)`$ as
$$P_{(n)}(t)=_{\mathrm{}}^{\mathrm{}}𝑑x\frac{(xx_0)^n}{n!}P(x,t).$$
(12.5)
The integral is well-defined because the momenta $`P(x,t)`$ have compact support, cf. eq. (6.2). The inverse relation reads
$$P(x,t)=\underset{n=0}{\overset{\mathrm{}}{}}P_{(n)}(t)(_x)^n\delta _{IR}(xx_0).$$
(12.6)
Alternatively, the formulas for the momenta follows from the Schrödinger representation
$$\frac{\delta }{\delta Q^{(0)}(x,t)}=\underset{n=0}{\overset{\mathrm{}}{}}(_x)^n\delta _{IR}(xx_0)\frac{}{Q^{(n)}(t)},$$
(12.7)
and equivalently
$$\frac{}{Q^{(n)}(t)}=_{\mathrm{}}^{\mathrm{}}𝑑x\frac{(xx_0)^n}{n!}\frac{\delta }{\delta Q^{(0)}(x,t)}.$$
(12.8)
The equations (6.5) and (6.1) translate into
$$P_{(n)}(t)=\underset{m=n}{\overset{\mathrm{}}{}}(_t)^{mn}\frac{L[Q](t)}{Q^{(m+1)}(t)}$$
(12.9)
and
$$\{\begin{array}{ccc}\hfill \dot{P}_{(n)}(t)+P_{(n1)}(t)& =& \frac{L[Q](t)}{Q^{(n)}(t)},nIN,\hfill \\ & & \\ \hfill \dot{P}_{(0)}(t)& =& \frac{L[Q](t)}{Q^{(0)}(t)},\hfill \end{array}$$
(12.10)
respectively. The very last equation is the Lagrangian Equation of Motion. The Hamiltonian (11.3) translates into
$$H(t)=\underset{n=0}{\overset{\mathrm{}}{}}P_{(n)}(t)Q^{(n+1)}(t)L[Q](t).$$
(12.11)
The Hamiltonian Equations of Motion are (12.3) and (12.10).
## 13 Discrete Case
Finally, let us consider the discrete case, cf. Section 4, with constant curves $`x_i(t)=x_i`$. The Euler-Lagrange equation for $`Q`$ reads
$`{\displaystyle \frac{\delta L[Q](t)}{\delta Q^{(0)}(x,t)}}`$ $`=`$ $`{\displaystyle \underset{iI}{}}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{L[Q](t)}{Q^{(k)}(x_i,t)}}(_x)^k\delta _{IR}(xx_i)`$ (13.1)
$`=`$ $`{\displaystyle \underset{iI}{}}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}E_{(k)}(x_i,t+xx_i)(_x)^k\delta _{IR}(xx_i).`$ (13.3)
Here we have introduced the higher Euler-Lagrange derivatives, cf. :
$$E_{(k)}(x,t)=\underset{m=k}{\overset{\mathrm{}}{}}\left(\begin{array}{c}m\\ k\end{array}\right)(_t)^{mk}\frac{L[Q](t)}{Q^{(m)}(x,t)},x=x_i,kZZ.$$
(13.4)
An alternative basis is provided by the Ostrogradski derivatives:
$$O_{(k)}(x,t)=\underset{m=k}{\overset{\mathrm{}}{}}(_t)^{mk}\frac{L[Q](t)}{Q^{(m)}(x,t)},x=x_i,kZZ.$$
(13.5)
The partial derivatives can be recovered via the inverse relation
$$\frac{L[Q](t)}{Q^{(k)}(x,t)}=O_{(k)}(x,t)+_tO_{(k+1)}(x,t),x=x_i,kZZ.$$
(13.6)
We assume that these are regular $`C^{\mathrm{}}`$-functions. Note the peculiar fact, that although the Lagrangian $`L[Q](t)`$ contains no temporal derivatives $`_t`$ – only spatial derivatives – it is the temporal derivatives $`_t`$ that is used in the construction of the above functions. Of course, this is the same on-shell, $`i.e.`$when (2.1) holds.
The momentum formula (6.5) becomes:
$`P(x,t)`$ $`=`$ $`\theta (xx_0(t)){\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}(_x)^k{\displaystyle \underset{iI}{}}\left[\theta (x_ix){\displaystyle \frac{L[Q](t+xx_i)}{Q^{(k)}(x_i,t+xx_i)}}\right]`$ (13.9)
$`\theta (x_0(t)x){\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}(_x)^k{\displaystyle \underset{iI}{}}\left[\theta (xx_i){\displaystyle \frac{L[Q](t+xx_i)}{Q^{(k)}(x_i,t+xx_i)}}\right]`$
$`=`$ $`{\displaystyle \underset{iI}{}}{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}E_{(k)}(x_i,t+xx_i)\left[\theta (xx_0(t))(_x)^k\theta (x_ix)\theta (x_0(t)x)(_x)^k\theta (xx_i)\right]`$ (13.11)
$`=`$ $`{\displaystyle \underset{iI}{}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}E_{(k)}(x_i,t+xx_i)(_x)^{k1}\delta _{IR}(xx_i)`$ (13.15)
$`+{\displaystyle \underset{iI}{}}E_{(0)}(x_i,t+xx_i)\left[\theta (x_ix)\theta (xx_0(t))\theta (x_0(t)x)\theta (xx_i)\right]`$
$`=`$ $`{\displaystyle \underset{iI}{}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}O_{(k)}(x_i,t)(_x)^{k1}\delta _{IR}(xx_i)`$ (13.19)
$`+{\displaystyle \underset{iI}{}}O_{(0)}(x_i,t+xx_i)\left[\theta (x_ix)\theta (xx_0(t))\theta (x_0(t)x)\theta (xx_i)\right].`$
We learn that the typical momenta for a discrete system will be distributional in nature with non-smooth behavior at the discrete points $`x=x_i`$. Note that the support of the momenta $`xP(x,t)`$ is inside the convex hull of the support (8.1) (= $`\{x_i|iI\}`$), if $`x_0(t)`$ is.
The Lagrangian Equation of Motion for $`q`$, cf. eq. (7.3), reads:
$$0=\frac{\delta S[q]}{\delta q^{(0)}(x+t)}\stackrel{(\text{2.1})}{=}\underset{iI}{}E_{(0)}(x_i,t+xx_i).$$
(13.20)
## 14 Discrete Local Formulation
We continue to study the discrete case. We have already given a local field formulation above. But after all, it is awkward to use field theory to quantize a discrete system! It is natural to ask if one can get a local formulation without introducing fields? Ostrogradski’s formal approach, cf. Section 12, is cumbersome in practice if the index set $`I`$ contains more than one element.
If we wish to address the original problem – and hence choosing $`\mathrm{\Sigma }_t`$ to be the convex hull of $`\{x_i|iI\}`$ – there seems to be no easy/useful/natural local discrete formulation in general.
However, if we let $`\mathrm{\Sigma }_t`$ be the smallest set possible, $`i.e.`$$`\{x_i|iI\}`$, we can do better. The first class constraints (10.1) along with the equations of motion (6.1) would then exclude the last term involving the Heaviside step function $`\theta `$ in the last two expressions of (13.19). It implies an Euler-Lagrange equation for each individual $`iI`$:
$$O_{(0)}(x_i,t)E_{(0)}(x_i,t)=0.$$
(14.1)
This corresponds to a theory without the chirality condition (2.1). We can fit this inside a generalized Ostrogradski framework with new coordinates $`Q^{i(n)}(t)`$, $`iI`$, $`nIN_0`$, defined as
$$Q^{i(n)}(t)=(_x)^nQ(x,t)|_{x=x_i}.$$
(14.2)
and momenta $`P_{i(n)}(t)`$, $`iI`$, $`nIN_0`$,
$$P_{i(n)}(t)=\underset{ϵ0^+}{lim}_{x_iϵ}^{x_i+ϵ}𝑑x\frac{(xx_i)^n}{n!}P(x,t),$$
(14.3)
canonical Poisson bracket
$$\{Q^{i(n)}(t),P_{j(m)}(t)\}=\delta _m^n\delta _j^i,$$
(14.4)
and Hamiltonian
$$H(t)=\underset{iI}{}\underset{n=0}{\overset{\mathrm{}}{}}P_{i(n)}(t)Q^{i(n+1)}(t)L[Q](t).$$
(14.5)
More relations can be read off from the Section 12 with the obvious minor modifications. In this way we have obtained an unconstrained discrete Hamiltonian system with slightly generalized Ostrogradski coordinates. For instance, the classical momentum formula (13.19) turns into
$$P_{i(n)}(t)=O_{(n+1)}(x_i,t),iI,nIN_0.$$
(14.6)
Acknowledgements. We would like to thank R.P. Woodard and B.D. Baker for comments. The research is supported by DoE grant no. DE-FG02-97ER-41029.
|
warning/0007/hep-ex0007045.html
|
ar5iv
|
text
|
# Studies of the Response of the Prototype CMS Hadron Calorimeter, Including Magnetic Field Effects, to Pion, Electron, and Muon Beams
## 1 The CMS Scintillator Tile-Fiber Hadron Calorimeter
The hadron calorimeter (HCAL) for the Compact Muon Solenoid (CMS) Detector will be used in the determination of quark, gluon, and neutrino final state momenta by measuring the energies and directions of particle jets and the missing transverse energy flow. The determination of the missing energy flow is crucial in searches for new particles and phenomena, such as possible supersymmetric partners of quarks and gluons. Adequate granularity, resolution, and containment of particle showers are essential in attaining these performance goals and provide one of the benchmarks in the design of the CMS hadron calorimeter. In this communication we report on test beam results used in the optimization of the design of HCAL, including choice of total absorber depth, sampling frequency, and longitudinal readout segmentation.
The central CMS hadron calorimeter is located inside the 4 Tesla coil of the CMS solenoid magnet (inner diameter of 5.9 m). The central pseudorapidity range ( $`|\eta |`$ $`<3.0`$ ) is covered by the barrel and endcap calorimeter systems. Segments of the crystal PbWO<sub>4</sub> electromagnetic calorimeter with a silicon pre-shower detector (in the endcap region only) are supported by the barrel and endcap hadron calorimeters. The combined response of the PbWO<sub>4</sub> crystal electromagnetic calorimeter and the hadron calorimeter is used in the reconstruction of particle jets and missing energy in the central pseudorapidity range.
The barrel and endcap sampling hadron calorimeters are located inside the 4 Tesla field of the CMS solenoid. Therefore, the calorimeters are constructed with non-magnetic material. The absorber and structural elements are made out of cartridge brass (70% Cu/ 30% Zn), and stainless steel plates, respectively. Cartridge brass is easy to machine and its hadronic interaction length is approximately 10% shorter than iron. The active sampling elements are 3.7 mm thick plastic scintillator tiles with wavelength-shifting (WLS) fiber readout. The barrel calorimeter inside the solenoid is relatively thin (about 7 interaction lengths at $`\eta =0`$). To ensure adequate sampling depth for $`|\eta |`$ $`<1.4`$, a hadron outer calorimeter is installed outside the solenoid. The outer calorimeter utilizes the solenoid magnet as an additional absorber equal to 1.4/sin($`\theta `$) interaction lengths and is used to identify and quantify the contribution from late starting showers.
### 1.1 Design Requirements
The design of the central hadron calorimeter requires good hermeticity, good transverse granularity, moderate energy resolution, and sufficient depth for hadron shower containment. We have chosen a lateral granularity of $`\mathrm{\Delta }\eta \times \mathrm{\Delta }\varphi =0.087\times 0.087`$ for $`|\eta |<2.0`$. This granularity is sufficient to insure good di-jet separation and mass resolution. The calorimeter readout is required to have a dynamic range from 20 MeV to 3 TeV. The sensitivity at the low end allows for the observation of single muons in a calorimeter tower for calibration and trigger redundancy. The scale at the high end is set by the maximum energy expected to be deposited by a jet in a single calorimeter tower.
Initial simulations of the CMS calorimeter indicate that a resolution of $`\sigma _E/E=120\%/\sqrt{E}`$ $``$ 5% for single incident hadrons is sufficient. In this case, the energy resolution for a jet of particles between 50 GeV and 3 TeV is not degraded by the measurement in the calorimeter, when other fluctuations which are inherent in jets are also considered.
### 1.2 Magnetic Field Effects
Particular attention must be paid to the effects of the high magnetic field on the response of the calorimeter. The intrinsic shower energy development and the containment of hadron showers have been shown not to be affected by the high magnetic field. However, the scintillator itself exhibits an increased response in a magnetic field. The increased response originates from two sources: (1) an intrinsic brightening of the scintillator ($``$ 5%) for B fields above 0.3 Tesla, and (2) a geometric effect due to the increased path lengths of soft electrons in the scintillator. The latter effect depends on the specific orientation of the magnetic field relative to the calorimeter absorber plates.
### 1.3 The Test Beam Program
A test beam program was initiated in 1994 and was continued in 1995 and 1996. A test module of moveable brass absorber plates and scintillator tile-fiber sampling was exposed to negative hadrons, electrons, and muons in the CERN test beam over a large energy range. During this period the effects of the following sources on the response of the calorimeter module were investigated:
* Magnetic field (scintillator “brightening” and geometric effects)
* Absorber thickness (optimization of resolution versus containment)
* Absorber depth (energy containment)
* Absorber non-uniformity (presence of the magnet coil)
* Crystal electromagnetic calorimeter (e/$`\pi `$ effects)
Data were taken both with and without a prototype lead tungstate crystal electromagnetic calorimeter (ECAL) placed upstream of the HCAL module. These data have been used in the optimization of the HCAL design, including the choice of total absorber depth, sampling frequency, and longitudinal readout segmentation.
The test beam data show that the presence of the crystal electromagnetic calorimeter in front of HCAL degrades the linearity and resolution of the combined calorimeter system. However, a large fraction of this degradation can be corrected for by applying constant energy-independent weighting factors to the various longitudinal readout segments of HCAL. The response of the combined calorimeter system using these optimized weights meets the design requirements for resolution and containment of hadron showers.
The experimental setup, various types of studies, and data sets are described in section 2. The measurement of the effects of a magnetic field on the response of HCAL to muons, electrons, and pions is presented in section 3. The performance of HCAL and the combined ECAL+HCAL system is discussed in sections 4 and 5. Extraction of the intrinsic ratio of the response of HCAL to the electromagnetic and hadronic components of the shower, $`e/h`$, is presented in section 6. The optimization studies leading to the choices for the total absorber depth, absorber sampling frequency, and longitudinal readout segmentation is presented in section 7. A comparison of the test beam results with Monte Carlo simulations is discussed in section 8. Section 9 gives a summary of the results and conclusions.
## 2 Experimental Setup
The combined ECAL+HCAL calorimetric system for CMS has been tested in 1995 and 1996 at the H2 and H4 beamlines at CERN. Data were taken using beams of muons, electrons, and hadrons, ranging in momenta from 15 to 375 GeV/c. These test beam prototypes of HCAL are based on the hanging file structure, in which copper alloy absorber plates ($`\lambda _{INT}`$(Cu)=15.06 cm) varying in thickness from 2 cm to 10 cm are interspersed with scintillator tiles read out with WLS fibers . Each scintillator tile is read out independently with a photomultiplier tube (PMT) located at the end of 10 m long optical cable. The total interaction depth of the HCAL prototypes corresponds to 10.3 $`\lambda _{INT}`$ (H2, 1995 setup), 8.5 $`\lambda _{INT}`$ (H4, 1995 setup), and 10.1 $`\lambda _{INT}`$ (H2, 1996 setup). The transverse size of the prototype scintillator tiles is 64 cm$`\times `$64 cm. Details of the H2(1995), H4(1995), and H2(1996) experimental setups are shown in Figure 1. The longitudinal segmentation of each of the modules is described in Tables 1 through 3.
In the H2(1995) run, the HCAL prototypes have been tested with the detector placed inside a large 3 Tesla magnet . The orientation of the magnetic field, with B field lines perpendicular to the scintillator planes (magnetic field parallel to the beam) corresponds to the endcap configuration of a typical collider detector. The H2(1995) tests include an ECAL module consisting of a lead-scintillator sampling calorimeter, with 10 layers of 1.6 cm Pb interspersed with 6 mm thick scintillator plates. The transverse size of the lead-scintillator sampling ECAL calorimeter for the H2(1995) is 32 cm$`\times `$32 cm.
During the H4(1995) run, the combined lead tungstate crystal ECAL+HCAL CMS prototype calorimeter has been tested with no magnetic field. The prototype H4(1995) ECAL detector consists of a matrix of 7$`\times `$7 PbWO<sub>4</sub> crystals, each 23 cm long (25.8 X<sub>0</sub>, 1.1 $`\lambda _{INT}`$), and 2 cm$`\times `$2 cm in transverse dimensions. The total transverse size of the ECAL module matrix is 14 cm$`\times `$14 cm. The crystal ECAL module is placed approximately 50 cm upstream of the front face of the HCAL prototype. The PbWO<sub>4</sub> crystals are read out by avalanche photodiodes (APD) with a gain approximately equal to 50.
The combined crystal ECAL+HCAL calorimetric system has been tested again in 1996 at the CERN H2 beamline with a magnetic field. The 3 Tesla magnet is oriented in such a way that the lines of the B field are parallel to the scintillator planes, corresponding to the HCAL barrel configuration. For the first 5 interaction lengths of the calorimeter, the magnetic field is uniform to within 10%. Figure 2 shows the relative value of the magnetic field, as a function of depth in HCAL.
### 2.1 The Relative Calibration of HCAL Layers
For the H2(1995) and H4(1995) tests, each scintillator tile is read out by 20 parallel WLS fibers. Figure 3 shows the details of the design of a single scintillator tile. Twenty parallel grooves, spaced every 3.2 cm, are machined with a ball-groove tool bit in 4 mm thick SCSN-81 scintillator plates. The tiles are painted at the edges using Bicron BC-620 white reflective paint, wrapped with white reflective $`Tyvek`$ , and sealed for light leaks with a black $`Tedlar`$ . The WLS fibers are mirrored at one end, placed in the grooves, and epoxied in groups of ten into two optical connectors placed at the edge of each scintillator tile, as shown in Figure 3. The stainless steel tubes are installed inside a 2 mm plastic cover and used to guide a radioactive Cs<sup>137</sup> source for calibration purposes. The entire package is held together with a set of brass rivets.
For the H2(1996) tests, a different set of scintillator tiles is used. These tiles are closer to the tower geometry of the final CMS design. Each tile is segmented into a 3$`\times `$3 matrix read out by nine WLS fibers inserted into $`\sigma `$ pattern grooves and is connected to a single PMT. The scintillator tile layout is shown in Figure 4. The tiles are made from 4 mm thick SCSN-81 scintillator produced by Kuraray. Multiclad Y-11 WLS fibers, made by Kuraray, 0.83 mm in diameter, are used. The cross-sectional view of a scintillator tile is shown in Figure 5. Scintillator tiles are packaged between 2 mm thick and 1 mm thick plastic covers. The transverse uniformity of response of the scintillator plates, measured using a radioactive photon source, has an rms of 4%.
A schematic view of the CERN H2 test beam setup is shown in Figure 6. The calorimeter is placed inside a superconducting magnet, which provides a magnetic field of up to 3 Tesla. Each scintillator tile is read out independently by a 10 stage, EMI 9902 KA photomultiplier tube. Optical cables carry light from the scintillator tiles to the PMT box. In order to avoid large PMT gain variations from the fringe fields of the magnet, the PMTs are located approximately 5 m away from the magnet, behind a 1 m thick iron wall. At this location, the fringe field for the magnet at 3 Tesla is approximately 50 Gauss. Therefore, each PMT can be adequately shielded by a regular inner $`\mu `$-metal shield and an outer, 3 mm thick, soft iron pipe.
The phototube calibration system is a crucial part of the setup. In order to understand the effects of the B field on the phototube gains, several redundant calibration schemes to monitor the PMT gain variation are used. The relative calibration of the scintillator tiles is established with an accuracy of $``$ 3% by equalizing the average response of each layer to muons. Figure 7 shows an ADC spectrum for 225 GeV/c muons traversing a single scintillator layer. The absolute light yield of each scintillator tile, determined using muons, is approximately 1.5 photoelectrons per minimum ionizing particle. This light yield is sufficiently high such as to not contribute significantly to the hadronic energy resolution of HCAL.
An independent method of monitoring the gain of each tile is provided by a system of radioactive source guide tubes . A radioactive Cs<sup>137</sup> source located at the tip of a long stainless steel wire is periodically inserted into guide tubes embedded in the plastic cover sheet for each tile. The source is moved using a computer controlled source driver.
The muon calibration and the wire source calibration schemes monitor variations in the product of the light yield of the scintillator tiles and the gain of the photomultipliers. In order to separately monitor variations in phototube gains (from temperature dependence and B fringe field effects), each PMT is also connected to a special scintillator block via a set of separate Y11 calibration fibers. The scintillator block is excited using either a laser beam, or a Cs<sup>137</sup> source. In addition, the Y11 calibration fibers can be excited by a light from a blue light emitting diode (LED). The scintillator block is located in a region with a low (approximately 50 Gauss) magnetic field, and thus is in a region where scintillator brightening from the fringe magnetic field is below 1% .
## 3 Calorimeter Performance in a Magnetic Field
The CMS hadron calorimeter, which uses plastic scintillators, will operate in a high magnetic field (4 Tesla). As has been previously observed , the light yield of plastic scintillators placed in high magnetic fields increases by 5 to 8%. This increase is due to polymer excitation that increases the energy of the short wavelength primary light in the ultra-violet (UV) region.
In addition, the presence of a magnetic field can affect the energy deposited in active parts of a sampling calorimeter. To understand this, the response of HCAL to muons, electrons, and pions has been studied as a function of magnetic field strength and orientation for several different sampling configurations. For all magnetic field studies the ECAL module was moved out of the beamline. During the H2(1995) tests, the B-field was oriented perpendicular to the scintillator planes, corresponding to the endcap HCAL configuration, while during the H2(1996) tests, the B-field was parallel to scintillator planes, corresponding to the barrel HCAL configuration.
### 3.1 B field perpendicular to the scintillator planes (endcap configuration)
When the magnetic field lines are perpendicular to the scintillator planes, the intrinsic light yield of the scintillator is higher than in the case where there is no magnetic field. This scintillator brightening effect leads to the same overall increased response of the calorimeter to muons, electrons, and pions. Figure 8 shows the response of HCAL to pions and electrons as a function of B field relative to B=0 Tesla. The observed $``$ 5% increase in the response of the calorimeter to particles is consistent with the increased light yield of the scintillator plastic, as measured by the calibration system using the radioactive Cs<sup>137</sup> photon source.
Figure 9 shows a comparison of the average shower profile for 300 GeV/c pions for two magnetic field settings: B=0 Tesla and B= 3 Tesla. The average energy deposition per sampling layer, in minimum ionizing particles (mips), is plotted as a function of HCAL absorber depth. Since the gain of each scintillator layer is established by equalizing its response to muons independently for the 0 T and 3 T data sets, the effect of scintillator brightening is already removed. As can be seen in the figure, the shape of the pion shower profiles remains unchanged in the presence of a 3 Tesla magnetic field perpendicular to the scintillator planes.
In summary, we conclude that a B field perpendicular to the scintillator planes results in an overall increase in the scintillator light yield. This scintillator brightening effect can be effectively measured by the calibration system using a radioactive $`\gamma `$ source.
### 3.2 B field parallel to the scintillator planes (barrel configuration)
In contrast, when the magnetic field lines are parallel to the scintillator planes, the observed average hadron shower profiles are altered. A comparison of the average observed shower profiles for 300/c GeV pions as a function of the HCAL absorber depth, for B = 0 and B = 3 Tesla, is shown in Figure 10. Energy deposited by hadron showers in mips shows an additional increased response, in the B = 3 Tesla field, relative to the B= 0 Tesla case.
The relative average response of HCAL to 100 GeV/c electrons and pions as a function of magnetic field is shown in Figure 11. As indicated by the data, the HCAL response to electrons in mips increases by approximately 20% at B=3 Tesla, relative to B=0 Tesla. However, the response for pions in mips increases by only $``$ 8%. The results are compared to simulations using the GEANT Monte Carlo program (10 GeV/c electrons and 50 GeV/c pions). As shown in the figure, the simulations reproduce data well.
Studies using a GEANT Monte Carlo simulation indicate that the increased response is due to a geometric effect resulting from a change in the path length of low energy electrons (between 1 and 10 MeV) exiting the absorber plates and traversing a circular path in the scintillator layer (because of the strong magnetic field). The radius of curvature of a 1 MeV/c electron in a 3 Tesla magnetic field is approximately 1 mm. The size of this geometric effect is expected to be proportional to the strength of the B field and depends on the detailed structure and composition of the absorber/scintillator package.
A GEANT study models the calorimeter response to 10 GeV electrons as a function of the air gap between the scintillator package and the upstream absorber (see Figure 12) in the presence of a strong magnetic field. The results are summarized in Figure 13. The increase in the electron response is reduced as the distance between scintillator package and upstream absorber is increased.
The results presented so far have been taken with the scintillator package placed in the absorber gap in the following way: 1 mm plastic + 4 mm scintillator + 2 mm plastic, with respect to the beam direction. In this configuration (configuration A), the scintillator is separated from the upstream-most absorber plate by 1 mm $`\pm `$ 0.5 mm of air and 1 mm of plastic. The Monte Carlo studies indicated that the effect of a parallel magnetic field depends on the distance between the scintillator and the most upstream absorber. Therefore, we have investigated the calorimeter response to pions, electrons, and muons with the reversed orientation of the scintillator package in the absorber gap (configuration B). In configuration B, the position of the scintillator package relative to the beam direction is: 2 mm plastic + 4 mm scintillator + 1 mm plastic. In the case of configuration B, the scintillator is separated from the upstream most absorber plate by 1 mm $`\pm `$ 0.5 mm of air and 2 mm of plastic.
Figure 14 shows the average electron and pion response of HCAL in mips as a function of B field, relative to the response at B= 0 Tesla, measured with the configuration B scintillator package. Here, we measure the response of electrons at B = 3 Tesla to be only 14 $`\pm `$ 1% higher (relative to B =0 Tesla), compared to 20% $`\pm `$ 1% effect for configuration A. GEANT simulation reproduces the electron data well. The additional 1 mm plastic upstream of the scintillator in configuration B (vs configuration A) helps range out more low energy electrons coming out of the upstream absorber plate.
In order to minimize the effect of the magnetic field on the response of the CMS hadron barrel calorimeter, one would like to maximize the distance between the scintillator and the upstream absorber. Therefore the configuration B of scintillator placement in the gaps between absorber plates has been chosen for the CMS HCAL design. A system of thin brass (Venetian Blind type) springs pushes the scintillator tiles radially outwards, such that the scintillator is always positioned closest to the downstream absorber plate. The use of springs ensures that the distance between inner absorber and the scintillator plate will consist of 2 mm $`\pm `$ 0.5 mm of air and 2 mm of plastic.
Figure 15 shows the ratio of HCAL response to pions as a function of pion energy, at 3 Tesla and 1.5 Tesla, relative to its response at 0 Tesla for the two configurations of the scintillator package inside the absorber gap. The lines shown on the Figure 15 indicate the values for $`\pi (3T)/\pi (0T)`$ and $`\pi (1.5T)/\pi (0T)`$ as a function of particle momentum. These functions were calculated using the Wigmans and Groom parameterizations<sup>1</sup><sup>1</sup>1Detailed study of the average fraction of the electromagnetic component in pion showers, $`F(\pi _0)`$, and extraction of $`e/h`$ is presented in section 6. of the average fraction of the electromagnetic component in pion showers. The average fraction of the electromagnetic component in pion showers increases as a function of pion energy. With $`e/h`$ of HCAL greater than unity, this increase results in non-linearity of HCAL response to pions. For example, the relative response of the calorimeter to 300 GeV/c pions is higher by $``$ 5%, with respect to 50 GeV/c pions. However, a 3 Tesla magnetic field increases the $`e/h`$ ratio for HCAL by $``$ 20 % for configuration A and 14% for configuration B, see Figs. 11 and 14. Therefore, a 3 Tesla magnetic field is expected to increase the non-linearity of HCAL’s response to pions between 50 GeV/c and 300 GeV/c by an additional $``$ 2-3%.
The above estimate is relevant only for the subset of single pions which interact in HCAL (minimum ionizing in ECAL). However, in the actual CMS detector, the common situation is that of a jet of particles depositing a large fraction of energy in the ECAL calorimeter. Since the ECAL calorimeter is not sensitive to magnetic field effects, the additional non-linearity from the 3 Tesla B field is greatly reduced.
In summary, a B field parallel to the scintillator planes results in an increased response due to two effects: (a) the scintillator brightening effect and (b) an additional increase in the response to the electromagnetic component of hadron showers (geometrical effect). The latter effect is strongly dependent on the placement of the scintillator packages with respect to the absorber plates of the calorimeter.
The results reported in this paper are consistent with earlier reports on the behavior of plastic scintillators in magnetic fields and on the influence of magnetic fields on response of scintillator based calorimeter . Note however, that in the latter publication the conclusion that electron response follows the dependence of the light yield (”brightening effect”) is true only for magnetic fields below 0.3 Tesla. As shown in this report, in the presense of parallel magnetic fields above 0.3 Tesla, the geometric effect leads to additional increase in response of the sampling calorimeters to electrons.
We also note that the origin of the latter effect is due to the relative geometry of the absorber with respect to the active elements of the calorimeter and is not sensitive to any special property of the scintillator as the active sampling medium. Therefore, similar effects are expected for any type of sampling calorimeters, with similar magnetic field/absorber configurations, independent of the choice of material for the active sampling medium.
## 4 Performance the HCAL calorimeter with a Pb-Scintillator ECAL
During the H2(1995) running period, a Pb-scintillator ECAL calorimeter was placed in front of the HCAL calorimeter. Linearity of the response and energy resolution of the combined system was studied for electrons, pions, and muons with momenta from 50 to 300 GeV/c.
In the following discussion we consider two different samples of pion data. The first, which we refer to as ”mip-in-ECAL” pions, includes only pions which are minimum ionizing in ECAL, i.e. do not interact in ECAL and begin interacting in HCAL. This sample is selected by requiring that the energy deposition in ECAL is consistent with minimum ionizing deposition, i.e. is less than 2.5 GeV. The second sample, which we refer to as the ”full pion sample”, includes all pions, i.e. interacting in either ECAL or HCAL.
The absolute energy scale of the HCAL calorimeter is set using 300 GeV/c ”mip-in-ECAL” pions. The average pulse height deposited by muons in HCAL corresponds to approximately 3 GeV of equivalent hadronic energy. Figure 16 shows the energy deposited by 225 GeV/c muons in HCAL. A pedestal trigger is a random trigger during the beam. This pedestal distribution is shown as a dashed line in Fig. 16. The width of the energy distribution for random pedestal triggers is equivalent to 80 MeV. Hence, the contribution of electronic noise to the energy resolution of HCAL is negligible.
The absolute energy scale of the ECAL calorimeter is set using 50 GeV/c electrons. The energy response of the Pb-scintillator sampling ECAL calorimeter ($``$ 2.9 X<sub>0</sub> sampling) for electrons with momenta in range 50 to 150 GeV/c is linear to within 1%. The fractional energy resolution for 50 and 100 GeV/c electrons is 7.2% and 6.2%, respectively.
The energy response of the combined Pb-scintillator ECAL+HCAL calorimeter system is calculated as follows:
$$E_{TOT}=\underset{i}{\overset{ECAL}{}}\frac{ADC_iw_i}{\mu _i}/Escale+\underset{j}{\overset{HCAL}{}}\frac{ADC_jw_j}{\mu _j}/Hscale$$
(1)
Here, $`Escale`$ and $`Hscale`$ are constants which define the absolute energy scales of ECAL and HCAL, and $`ADC_i`$ are the ADC values of each readout layer of ECAL and HCAL. The relative calibration constants of each layer, $`\mu _i`$, correspond to the average muon response of a layer expressed in units of ADC counts. The coefficients $`w_i`$ are proportional to the arithmetic mean of the interaction lengths of the absorber plates upstream and downstream of each scintillator layer (the ”Simpson” approximation formula), $`w=0.5\times (t_{upstream}+t_{downstream})`$.
Scatter plots of the energies in HCAL vs ECAL for 50, 100, 150, and 300 GeV/c particles in the hadron beam are shown in Figure 17. As indicated in the figure, the nominal hadron beam, especially at the 50 GeV/c and 100 GeV/c beam tunes, contains a large fraction of electrons. Pion induced events are selected by requiring that energy deposited in HCAL is at least 0.5 GeV. In order to remove particles that interact upstream of ECAL, the energy deposition in the first ECAL layer is required to be consistent with that of a minimum ionizing particle.
The total (ECAL+HCAL) energy distributions for ”mip-in-ECAL” pions are shown in Figure 18. The total energy distributions for the ”full pion sample” are shown in Figure 19. The reconstructed energy distributions are well described by Gaussian fits for both sets of data. Figure 20 shows the linearity versus energy for the calorimeter response for ”mip-in-ECAL” pions, and for the ”full pion sample” using the Pb-scintillator ECAL+HCAL calorimetric system. The average response of the combined Pb-scintillator ECAL+HCAL system to the ”full pion sample” is approximately 1% lower than the response the calorimeter for the subset of ”mip-in-ECAL” pions.
A comparison of the relative energy resolutions of the calorimeter for ”mip-in-ECAL” pions and for the ”full pion sample” is shown in Figure 21. The energy resolutions for these two cases are comparable. At low energies, the energy resolution of the ”full pion sample” is slightly improved because of the finer sampling of the Pb-scintillator ECAL.
The fractional energy resolutions are parameterized by the following function:
$$\sigma _E/E=(stoch.term)/\sqrt{E}(const.term)$$
(2)
Here $`E`$ is the particle energy in GeV, and the symbol $``$ implies that the stochastic and constant terms in the resolution are combined in quadrature.
The following values for the stochastic term and the constant term are extracted from the data:
$$\sigma _E^{pionsmipinECAL}/E=(93.8\pm 0.9)\%/\sqrt{E}(4.4\pm 0.1)\%$$
(3)
for ”mip-in-ECAL” pions, and
$$\sigma _E^{allpions}/E=(82.6\pm 0.6)\%/\sqrt{E}(4.5\pm 0.1)\%$$
(4)
for the ”full pion sample”.
## 5 Performance of the HCAL calorimeter with a PbWO<sub>4</sub> crystal ECAL
In this section the linearity of response and energy resolution of the HCAL calorimeter with a PbWO<sub>4</sub> crystal matrix ECAL are discussed. The performance of the combined PbWO<sub>4</sub> ECAL+HCAL calorimetric system was first investigated during 1995 in the H4 beamline.
The calibration of the ECAL crystals uses 50 GeV/c electrons directed into the center of each crystal. Figure 22 shows the energy response of ECAL for 25, 50, 100, and 150 GeV/c electrons using the 7 $`\times `$ 7 ECAL matrix sum. The energy resolutions of the crystal ECAL at these energies are found to be 0.6 GeV, 0.7 GeV, 1.0 GeV, and 1.4 GeV, corresponding to fractional energy resolutions of 2.4%, 1.4%, 1.0%, and 0.9%, respectively. During these tests, the electronic noise of the 7$`\times `$7 crystal matrix energy sum had a rms width equivalent to 440 MeV. Much better performance of the crystal ECAL calorimeter has been attained in other tests which were dedicated to special ECAL studies.
The energy response of the combined crystal ECAL+HCAL calorimeter system is calculated as follows:
$$E_{TOT}=\underset{i}{\overset{ECAL}{}}\frac{ADC_i}{e_i}/Escale+\underset{j}{\overset{HCAL}{}}\frac{ADC_jw_j}{\mu _j}/Hscale$$
(5)
Here, $`Escale`$ and $`Hscale`$ are constants which define the absolute energy scales of ECAL and HCAL, $`ADC_i`$ are the ADC values of each readout layer (tower) of the HCAL (ECAL) calorimeter. The relative calibration constants of each tower of ECAL, $`e_i`$, correspond to the average 50 GeV/c electron response of a tower expressed in units of ADC counts. The coefficients $`w_j`$ and $`\mu _j`$ are defined as before (Eqn. 1).
Figure 23 shows a scatter plot of the energy deposition in HCAL versus ECAL taken during the H4(1995) test beam run with pions with momenta ranging between 15 and 375 GeV/c. Unlike in the case with the Pb-scintillator ECAL, the data points are not centered on a straight line corresponding to $`E_{ECAL}+E_{HCAL}`$ = $`p_{beam}`$. The energy response of the HCAL calorimeter to ”mip-in-ECAL” pions is shown in Figure 24. The response of the combined crystal ECAL+HCAL calorimeter to the ”full pion sample” is shown in Figure 25. The distributions for the combined crystal ECAL+HCAL system deviate from a Gaussian shape, especially at low energies.
As demonstrated earlier, the H2(1995) data with a Pb-scintillator ECAL show good linearity for the combined ECAL+HCAL calorimeter system. However, as can be seen in Figure 26, the H4(1995) data with a PbWO<sub>4</sub> crystal ECAL is quite non-linear. In addition, the average response of the combined crystal ECAL+HCAL calorimeter to ”full pion sample” is approximately 10% lower than the response to pions which interact in HCAL only.
Figure 27 shows the fractional energy resolutions of the calorimeter for the ”mip-in-ECAL” pions for the ”full pion sample”. The energy resolutions of the combined crystal ECAL+HCAL calorimeter system are significantly degraded, especially at high energies. These results are attributed to the large difference between the response of the crystal ECAL to electrons and hadrons, i.e. the large $`e/h`$ (the ratio of the response to the electromagnetic and hadronic components of the shower) of the PbWO<sub>4</sub> crystals. This large $`e/h`$ results in an overall response of the calorimeter which is very sensitive to fluctuations in the initial electromagnetic component of hadronic showers.
A fit to data for the fractional energy resolution yields the following values for the stochastic term and constant term parameters:
$$\sigma _E^{pionsmipinECAL}/E=(101\pm 0.1)\%/\sqrt{E}(4.0\pm 0.1)\%$$
(6)
for the case of ”mip-in-ECAL” pions , and
$$\sigma _E^{allpions}/E=(127\pm 0.7)\%/\sqrt{E}(6.5\pm 0.1)\%$$
(7)
for the case of the ”full pion sample”. Note that the first fit (”mip-in-ECAL” data , eqn. 6) is consistent with the H2(1995) ”mip-in-ECAL” data (eqn. 3).
## 6 Response of HCAL to electrons and the $`e/h`$ ratio for HCAL
The linearity and energy resolution of hadron calorimeter depends on the intrinsic $`e/h`$. The average fraction of the energy in the electromagnetic component of hadron induced showers, F($`\pi ^0`$), increases as a function of incident energy. For non-compensating calorimeters ( $`e/h`$ $``$1), this implies a non-linear hadron energy response. The event-by-event fluctuations in F($`\pi ^0`$) contribute to variations in the reconstructed shower energy and dominate the energy resolution for hadrons at high energies.
In order to study the relative response of HCAL to electron and pions, data were taken with ECAL removed from the beamline. In order to reject electrons present in the hadron tunes from the sample, a minimum ionizing signal was required in the first three sampling layers of HCAL.
The studies of the energy response of HCAL to pions indicate that the HCAL calorimeter is somewhat non-linear. There is $``$ 9% increase in the relative response of HCAL to pions between 15 and 375 GeV/c, as shown in Figure 28. The total interaction length <sup>2</sup><sup>2</sup>2Taking into account the requirement that energy deposition in first three HCAL counters is consistent with minimum ionizing particles. of the H4(1995) HCAL module is approximately 8 $`\lambda _{INT}`$. At very high energies, an 8 $`\lambda _{INT}`$ calorimeter is not expected to fully contain the hadronic shower. The correction for longitudinal leakage for an 8 $`\lambda _{INT}`$ calorimeter is estimated to be 1.5% at 120 GeV/c and 3.5% at 375 GeV/c, using CCFR and NuTeV test beam data. Therefore, the intrinsic non-linearity of HCAL is somewhat larger if the corrections for leakage at high energies are applied.
The response of the H4(1995) HCAL module to electrons is linear at a level of 2% and is approximately 15% higher than the average response to pions as shown in Figure 28. The absolute energy scale of HCAL is set (as before) by the 375 GeV/c pion data point.
The electron and pion data can be used to extract $`e/h`$ by applying the following expressions:
$$\pi =F(\pi ^0)\times e+(1F(\pi ^0))\times h$$
(8)
$$\frac{\pi }{e}=\frac{1+(\frac{e}{h}1)\times F(\pi ^0)}{\frac{e}{h}}$$
(9)
Here $`e`$ is the response of the calorimeter to the electromagnetic components of hadron showers and $`h`$ is the response to the hadronic component of pion showers. $`\pi `$ is the response to pions and F($`\pi ^0`$) is the fraction of the energy in electromagnetic component of pion showers.
The extracted value of $`e/h`$ depends on the assumed parameterization for F($`\pi ^0`$) as a function of energy. The Wigmans formula for F($`\pi ^0`$) (eqn. 10) increases with energy ($``$ log(E)) and becomes non-physical at very high energies. The Groom parameterization (eqn. 11) is extracted from a CALOR Monte Carlo simulation and uses a power law dependence.
$$F_W(\pi ^0)=0.11\times ln(E)$$
(10)
$$F_G(\pi ^0)=1(E/0.96)^{0.8161}$$
(11)
Using the above two parameterizations, we extract values of $`e/h`$ of 1.41 $`\pm `$ 0.01 (Wigmans) and 1.51 $`\pm `$ 0.01 (Groom). The best fit to our data is shown in Figure 29.
The above results can be compared with similar analyses for other iron-scintillator sampling calorimeters. Extraction of $`e/h`$ for an iron-scintillator sampling calorimeter has also been done by CDF. The CDF endcap hadron calorimeter consists of 22 layers of 5 cm iron absorber plates and 6 mm scintillators. The CDF results yield $`e/h`$ 1.34 $`\pm `$ 0.01 (Wigmans) and 1.42 $`\pm `$ 0.015 (Groom). ATLAS measures $`e/h`$= 1.34 $`\pm `$ 0.03 assuming the Wigmans parameterization for $`F(\pi ^0)`$. The ATLAS hadron calorimeter uses iron as an absorber with the tiles are oriented in the radial direction.
The NuTeV Collaboration reports a $`e/h`$ ratio for 10 cm iron sampling calorimeter of 1.08 $`\pm `$ 0.01 using the Groom parameterization. Note that $`e/h`$ is sensitive to the sampling fraction and other geometrical effects. The CCFR/NuTeV calorimeter uses 2.5 cm thick liquid scintillator counters, which are clad with acrylic and water bags. The cladding and the thick scintillators in CCFR/NuTeV greatly reduce the response to electrons relative to the response to pions. Also, the water and thick scintillators tend to increases the rate of conversion of low energy neutrons to protons in the active medium.
## 7 Optimization of the design of HCAL
The conceptual design of the barrel HCAL was a 5.3 $`\lambda _{INT}`$ thick calorimeter. The inner half of the calorimeter had 3 cm sampling (first readout segment, H1) while the second readout segment consisted of the outer half of the calorimeter with 6 cm sampling.
For the H2(1996) tests, the prototype HCAL module was segmented into 27 readout layers. Using these data we simulated various sampling configurations and studied the performance of HCAL as a function of total interaction length and sampling frequency. These data allowed us to optimize the calorimeter response to pions and jets, while taking into account the existing design (e.g. geometrical and readout) constraints imposed on HCAL within the CMS detector.
Important design choices that were made based on these data are:
* the absorber sampling thickness;
* the depth of the HCAL inside the magnet (in interaction lengths);
* the longitudinal segmentation of the inner HCAL into H1 and H2 readout segments;
* possibility of adding scintillator layers outside of the magnet to create an outer calorimeter.
### 7.1 Choice of the longitudinal segmentation for HCAL
Studies performed prior to the H2(1996) beam tests resulted in no compelling argument to set the optimal partition between the two readout segments of inner HCAL. However, the 1996 test beam data indicate that the proper partition is that which is most useful in correcting for large $`e/h`$ response of the ECAL crystals. To compensate for a large $`e/h`$ of the combined ECAL + HCAL system, we chose to make a novel longitudinal segmentation, with the first depth segment, H1, reading out only the first scintillator layer. The remaining scintillator layers inside the magnet are combined into the second readout, H2. The reason for this choice is the following. A large H1 signal indicates that a significant amount of hadronic energy has been deposited in ECAL and is underestimated because of the large $`e/h`$ of the crystals. The information from H1 can be used to make an effective correction for the hadronic non-linearity of the ECAL crystals.
### 7.2 Response of the combined crystal ECAL+HCAL calorimeters
Figure 31 and 32 show the energy response and fractional energy resolutions of the combined PbWO<sub>4</sub> crystal ECAL+HCAL calorimeters to pions. The inner HCAL(5.3 $`\lambda _{INT}`$) consists of two independent readouts: H1 (following a 2 cm Cu plate) and H2 (thirteen 6 cm Cu samplings). A single readout outer calorimeter (HO) consists of three samplings: (1) the first sample is immediately after the magnetic coil (which is mimicked by 18 cm of Cu in the test beam ),(2) a second 22 cm Cu sampling layer, and (3) a third 16 cm Cu sampling layer. The combined coil and HO samplings correspond to a total of 2.5$`\lambda _{INT}`$.
We have investigated two possible approaches to correct for the degradation of the performance of the combined crystal ECAL+HCAL calorimeters. Both of these methods make use of the segmented readout of HCAL. The first approach, called passive weighting, increases the weight ($`\alpha `$) of the first (H1) HCAL readout segment, where $`\alpha `$ is an energy independent constant. Note that $`\alpha `$ =1 corresponds to using the standard ”absorber Simpson’s weighting” used elsewhere in the calorimeter.
$$E_{TOT}=E_{ECAL}+\alpha \times E_{H1}+E_{H2}+E_{HO}$$
(12)
The overall linearity and fractional energy resolution of the combined crystal ECAL+HCAL system for 300 GeV/c pions as a function of the parameter $`\alpha `$ are shown in Figure 30. Both the linearity and energy resolution of the combined crystal ECAL+HCAL system are improved for the value of $`\alpha `$ = 1.4.
The second approach, called dynamic weighting is an event-by-event correction. It depends on the fraction of the energy deposited in the first readout segment of HCAL immediately downstream of the crystal ECAL. The dynamic weighting effectively allows one to have an energy dependent correction (for single particles) for the low response to pions which interact in ECAL.
$$E_{TOT}=(1+2\times f(H1))\times E_{ECAL}+E_{H1}+E_{H2}+E_{HO},$$
(13)
$$f(H1)=E(H1)/(E(H1)+E(H2)+E(HO)),f(H1)0.1$$
(14)
With either the passive or dynamic weighting, the nonlinearity and resolutions for pions with energies between 30 and 300 GeV/c are improved. With the passive weighing, the fractional energy resolution of of the combined ECAL+HCAL calorimetric system can be described by the function $`\sigma _E/E`$ = 122%/$`\sqrt{E}`$ $``$ 5%. Note that while the passive weighting can be applied to single particles and jets, the dynamic weighting may introduce high energy tails in the case of particle jets.
Techniques to overcome the problem of the different non-compensation properties (i.e. $`e/h`$) of the electromagnetic and hadronic compartments for combined calorimetry systems have been studied by the ATLAS calorimetry group. The algorithms proposed by the ATLAS group depend on the energy fraction deposited in the electromagnetic compartment and therefore can be only applied to single particle energy reconstruction.
### 7.3 Studies of the total HCAL absorber depth
We use the 27 sampling layers of the H2(1996) test beam module to study the calorimeter performance as a function of total depth. Figure 33 shows the average pion shower profiles as a function of total absorber depth for 50, 100, 150, and 300 GeV/c pions. As shown in the figure, the average pion shower profiles extend significantly beyond 7 $`\lambda _{INT}`$, in particular at high pion energies. Fluctuations in the leakage for high energy pions also become large.
To collect this leakage energy, the barrel HCAL has been augmented with an additional outer calorimeter (HO) consisting of a single layer of scintillators beyond the solenoid magnet (the solenoid thickness is 1.4 $`\lambda _{INT}`$). For pseudorapidity $`\eta `$ less than 0.4, a layer of iron (thickness = 18 cm = 1.1 $`\lambda _{INT}`$) has been instrumented with an additional scintillator layer. For either case (one or two layers of scintillator), the scintillators cover the same solid angle as the interior calorimeter and are read out as a single depth segment. The combined solenoid + iron of the HO corresponds to an additional 2.5 $`\lambda _{INT}`$.
The fraction of 300 GeV/c pions with reconstructed energy less than 200 GeV (approximately 3 $`\sigma `$ below the mean, or 100 GeV of missing energy) is shown in Figure 34. The four points correspond to: 5.9 $`\lambda _{INT}`$, the HCAL alone; 7.0 $`\lambda _{INT}`$, HCAL + ECAL; 9.5 $`\lambda _{INT}`$, HCAL \+ ECAL + HO; and 11 $`\lambda _{INT}`$ which is the total thickness of the 27 layer test beam module. We see that for the CMS barrel design, at 9.5 $`\lambda _{INT}`$, less than 2% of the pions are catastrophically mismeasured. To increase the interaction length of the calorimeter to 9.5 $`\lambda _{INT}`$ we added two additional absorber plates inside the magnet with respect to the conceptual design. This was achieved by reducing the inner radius of barrel HCAL.
### 7.4 Optimization of Absorber Sampling Thickness
The fractional energy resolution has been investigated for the following three choices for the inner HCAL absorber samplings: a) 3 cm Cu sampling for the first eight layers followed by 6 cm Cu sampling, b) 6 cm Cu uniform sampling, and c) 12 cm Cu uniform sampling. The data, shown in Figure 35 indicate that the energy resolution is not dominated by the sampling fluctuations in HCAL. A factor of two decrease in the sampling frequency ( 3 cm/6 cm to 6 cm uniform sampling) for HCAL does not result in a noticeable degradation of the energy resolution of the combined detector system. In the case of 12 cm Cu uniform sampling, the degradation in energy resolution is noticeable, but does not scale with $`\sqrt{t}`$, where $`t`$ is the thickness of the absorber plates. For the final design, we chose uniform 5 cm sampling which does not degrade the energy resolution and puts more absorber inside the magnet, relative to the conceptual design.
### 7.5 Final Design of CMS Barrel HCAL
The final design of the CMS barrel HCAL consists of 17 absorber/scintillator samples with 5 cm absorber thickness. The total thickness of absorber at 90 degrees is 5.9 $`\lambda _{INT}`$. With the ECAL included, the total thickness inside the magnet is 7.0 $`\lambda _{INT}`$. Exterior to the magnet, there is a 2.5 $`\lambda _{INT}`$ outer calorimeter. Each projective tower has 3 readouts in depth: the very first scintillator layer (H1); the remaining 16 scintillator layers inside the magnet (H2); and the outer calorimeter (HO).
## 8 Monte Carlo simulation of the test beam results
Several hadron shower generators are available within the GEANT framework. These include GHEISHA , GFLUKA (which is an implementation of FLUKA within GEANT), and GCALOR (which is an implementation of CALOR within GEANT). Although GHEISHA is native to the GEANT program, GFLUKA and GCALOR are imperfect implementations of the original FLUKA and CALOR programs. Both of these programs have been known to produce somewhat different results than the original generators.
GEANT is used in various studies for evaluation of the calorimeter design for CMS. In order to verify those simulations and to understand their limitations, GCALOR is used to simulate the H2(1996) test beam data. It also serves as a reference for comparison with the other generators. Our aim is first to anchor the Monte Carlo model to the ensemble of test beam data. Once it is so constrained, it is assumed that it can be used to make small extrapolations to model the final CMS calorimeter system.
Details of the crystal ECAL and HCAL test beam geometry are implemented in the GCALOR simulation. For ECAL, this includes a 7 x 7 matrix of individual crystals surrounded by copper blocks as well as mechanical and cooling structures. The HCAL geometry includes the layer structure of copper plates, scintillator, plastic cover plates, and air gaps on both sides of each scintillator. In order to take into account all experimental effects, the transverse beam profiles are simulated using information from the test beam tracking chamber. Electronic noise and photo-statistics effects are simulated based on the measured distributions of pedestals, electron, and muon signals in ECAL and HCAL. The longitudinal light collection efficiencies in the central nine crystals are included in the simulation. The energy cut values in the GEANT simulation are set to the GEANT default values, 1 MeV for electrons and 10 MeV for hadrons.
In order to compare response of pions interacting in HCAL, we have removed ECAL from the beam. For this dataset, the test beam data are in good agreement with the results of the GCALOR Monte Carlo simulations. Good agreement is observed in the longitudinal shower profile (Fig. 36), the pion response versus energy (Fig.37), and in the fractional pion energy resolution (Fig.38).
A comparison of the linearity and fractional energy resolutions for the ”full pion sample” using the combined crystal ECAL+HCAL system for test beam data and the GEANT simulations are shown Figures 39 and 40. The GEANT simulations, which include all experimental detector effects such as electronics noise, are in good agreement with the test beam data. Also shown in Figure 40 are the results of a MC simulation of the crystal ECAL+HCAL energy resolution excluding test beam detector effects, such as the ECAL electronic noise, and the energy leakage from the small prototype ECAL. As can be seen in Figures 40, the detector effects present during the H2(1996) test beam data taking significantly degrade the fractional pion energy resolution, especially at the low energies. These test beam detector effects are not expected to be present in the final CMS configuration.
Comparisons of the fractional energy resolution simulated by GCALOR and other GEANT hadron simulators are shown in Figure 41. The simulations based on GHEISHA predict somewhat worse resolutions, while simulations based on GFLUKA+MICAP predict much better resolutions than predicted with GCALOR and observed in the data.
## 9 Conclusions
Comprehensive tests of the performance of prototype CMS central hadron calorimeters have been done in the H2 and H4 beamlines at CERN. Data were taken with both a stand-alone HCAL calorimeter and with HCAL in combination with an upstream ECAL calorimeter.
One of the primary objectives of the HCAL test beam studies was to investigate the calorimeter performance in the presence of perpendicular and parallel magnetic fields. A high magnetic field changes the response of the calorimeter in two different ways: (1) the overall light yield of the scintillator is increased in a high magnetic field and (2) the field affects the observed energy deposition of electromagnetic components of showers when it is parallel to the scintillator plates (i.e. perpendicular to the particle direction). For a collider experiment with a solenoid magnet, the magnetic field is parallel to the calorimeter plates in the central part of the detector (barrel configuration) and is perpendicular to calorimeter plates in the large $`\eta `$ region ( endcap configuration).
When the magnetic field is perpendicular to the scintillator planes (endcap configuration), only an intrinsic increase of the light yield of the scintillator of approximately 5-8% is observed relative to the case with no magnetic field. This effect leads to the same overall increased response of the calorimeter to muons, electrons, pions, and $`\gamma `$ rays from a radioactive source. Therefore a calibration source can be used to track and correct for this effect.
An additional geometric effect leads to an increased response of the calorimeter to the electromagnetic component of showers. This effect occurs in the case in which the magnetic field lines are parallel to the scintillator planes (barrel configuration) and originates from an increase in the geometrical path length of low energy electrons in a magnetic field. Since this effect is not present for the calibration source, $`\mathrm{𝑖𝑛}\mathrm{𝑠𝑖𝑡𝑢}`$ (B field on) calibration is required for hadron barrel calorimeter.
The size of this effect is approximately proportional to the strength of the B field and depends on the detailed structure and composition of the absorber and scintillator planes. However, the effect is expected to be present in any sampling calorimeter (independent of readout technology) situated in a magnetic field which is parallel to the readout planes (barrel configuration).
Due to the non-compensating nature of the lead tungstate crystal ECAL the linearity and energy resolution of the combined crystal ECAL+HCAL is worse than for HCAL alone. Therefore, improvements in the linearity and resolution using weighting of the two longitudinal readouts (H1 and H2) of HCAL have been investigated. Using a passive weighting, the fractional energy resolution of the combined crystal ECAL+HCAL calorimetric system is described by the function $`\sigma _E/E`$=122%/$`\sqrt{E}`$ $``$ 5%. This performance is achieved by a novel longitudinal segmentation of HCAL with H1 being a single layer immediately behind ECAL. We conclude that for combined ECAL+HCAL calorimeter systems with an ECAL with a large e/h, a very thin first depth segment in HCAL can be used to largely correct for the resultant non-linearity and degradation in energy resolution. Monte Carlo studies of the CMS detector in a collider enviornment indicate that with the above performance, the energy resolution for jets is not dominated by the energy measurement in the hadron calorimeter, but by other fluctuations which are inherent in jets .
We find that the average longitudinal hadron shower profiles extend past the inner HCAL located inside the magnetic coil. Therefore, the CMS central calorimeter design includes instrumenting an outer calorimeter (HO) outside the coil to measure the energy in the tail of high energy hadronic showers. The addition of HO leads to a total ECAL+HCAL+HO depth of at least 9.5 $`\lambda _{INT}`$ for almost the entire $`\eta `$ range spanned by the CMS hadron calorimeters.
Various GEANT based simulations predict the impact of the performance of HCAL on a variety of potential physics searches in the CMS detector. These simulations extrapolate the performance of HCAL from the test beam configuration to the final configuration chosen for CMS HCAL design. The Monte Carlo programs have been shown to successfully simulate various test beam setups, including hadron shower calorimetry in a setting with both crystal and copper-scintillator detectors in a strong magnetic field. The best description of the data is provided by GEANT with the GCALOR module for the generation of hadron showers.
## 10 Acknowledgments
We thank the CERN staff and the staff of the participating institutions for their vital contributions to this experiment. We particularly acknowledge efforts of Maurice Haguenauer and Jean Bourotte (H4 beamline), Pascal Petiot, and Gerhard Waurick (H2 beamline), Brian Powell (3T magnet), Albert Ito (PMTs), as well as Anatoly Zarubin (DAQ).
|
warning/0007/math0007161.html
|
ar5iv
|
text
|
# Morse theory on graphs
## 1. Introduction
As Richard Stanley so astutely observes in \[St\], “The number of systems of terminology presently used in graph theory is equal, to a close approximation, to the number of graph theorists.” Our terminological conventions in this paper will be the following: Given a finite $`d`$-valent graph, $`\mathrm{\Gamma }`$, we will denote by $`V_\mathrm{\Gamma }`$ the vertices of $`\mathrm{\Gamma }`$ and by $`E_\mathrm{\Gamma }`$ the oriented edges of $`\mathrm{\Gamma }`$. For each oriented edge, $`e`$, we will denote by $`\overline{e}`$ the edge obtained by reversing the orientation of $`e`$ and by $`i(e)`$ and $`t(e)`$ the initial and terminal vertices of $`e`$. (Thus “$`d`$-valent” means that for every vertex, $`p`$, there are exactly $`d`$ edges with $`i(e)=p`$.)
Let $`G`$ be a torus of dimension $`n>1`$, and let $`\varrho `$ be a function which associates to every oriented edge, $`e`$, of $`\mathrm{\Gamma }`$, a one-dimensional representation, $`\varrho _e`$, of $`G`$, with character $`\chi _e:GS^1`$, and $`\tau `$ a function which associates to every vertex, $`p`$, of $`\mathrm{\Gamma }`$, a $`d`$-dimensional representation, $`\tau _p`$, of $`G`$. We will say that $`\varrho `$ and $`\tau `$ define *an action of* $`G`$ *on* $`\mathrm{\Gamma }`$ if they satisfy the three axioms below:
Axiom A1:
$$\tau _p\underset{i(e)=p}{}\varrho _e.$$
Axiom A2:
$$\varrho _{\overline{e}}\varrho _e^{}.$$
Axiom A3:
Let $`G_e`$ be the kernel of $`\chi _e:GS^1`$ and let $`p=i(e)`$ and $`q=t(e)`$. Then
$$\tau _p|_{G_e}\tau _q|_{G_e}.$$
The data $`(\mathrm{\Gamma },\rho ,\tau )`$ can be viewed as the schematic description of a genuine $`G`$-action, namely an action of $`G`$ on a $`d`$-dimensional complex manifold, $`M`$, having the following properties:
1. : The fixed point set of $`G`$ is $`V_\mathrm{\Gamma }`$, and, for every $`pV_\mathrm{\Gamma }`$, $`\tau _p`$ is the isotropy representation of $`G`$ on the tangent space to $`M`$ at $`p`$.
2. : Each edge, $`e`$, of $`G`$, corresponds to a $`G`$-invariant imbedded projective line, $`\gamma :P^1M_eM`$. The points $`p=i(e)=\gamma ([1:0])`$ and $`q=t(e)=\gamma ([0:1])`$ are the two fixed points for the action of $`G`$ on $`M_e`$, and $`\varrho _e`$ is the isotropy representation of $`G`$ on the tangent space to $`M_e`$ at $`p`$.
These properties are exhibited by a class of manifolds, called *GKM manifolds*, which were first introduced in \[GKM\] by Goresky, Kottwitz and MacPherson and have subsequently been studied in a number of recent papers (\[GZ1\], \[GZ2\], \[KR\], \[LLY\], \[TW\]) by ourselves and by others. By definition, a complex $`G`$-manifold, $`M`$, is *GKM* if, for every fixed point, $`p`$, of $`G`$, the weights, $`\alpha _i`$, $`i=1,..,d`$, of the isotropy representation of $`G`$ on the tangent space to $`M`$ at $`p`$ are pairwise linearly independent: *i.e.* $`\alpha _i`$ and $`\alpha _j`$ are linearly independent if $`ij`$. For such a manifold we define its *one-skeleton* to be the set of points, $`pM`$, for which the stabilizer group of $`p`$ is of dimension at least $`n1`$. What Goresky, Kottwitz and MacPherson observe is that this one-skeleton consists of closed submanifolds of $`M`$ on which $`G`$ acts in a fixed point free fashion and imbedded $`P^1`$’s satisfying P1-P2.
In \[GZ2\] we proved a converse to this result. Let $`(\mathrm{\Gamma },\varrho ,\tau )`$ be a graph-theoretical $`G`$-space in the sense described above. For $`eE_\mathrm{\Gamma }`$ let $`\alpha _e`$ be the weight of the representation, $`\varrho _e`$, and assume that for every vertex, $`p`$, of $`\mathrm{\Gamma }`$, the weights $`\alpha _e`$, $`i(e)=p`$, of the representation, $`\tau _p`$, are pairwise linearly independent. Then there exists a complex $`G`$-manifold, $`M_\mathrm{\Gamma }`$, of GKM type, whose one-skeleton has the properties P1-P2. (In fact, $`M_\mathrm{\Gamma }`$ is canonically constructed from the data $`(\mathrm{\Gamma },\rho ,\tau )`$ and is basically just a tubular neighborhood of this one-skeleton.)
From Axiom A1, it is clear that the function, $`\rho `$, determines the function, $`\tau `$; and $`\rho `$, in turn, is determined by the function
$$\alpha :E_\mathrm{\Gamma }𝔤^{},e\alpha _e.$$
In the geometric realization of $`\mathrm{\Gamma }`$ which we described above this function tells us how each of the two-spheres in the one-skeleton is rotated about its axis of symmetry; so we will call this function the *axial function* of the $`G`$-action on $`\mathrm{\Gamma }`$. The axioms, A1-A3, can easily be translated into axioms on this axial function (as we will do in Section 2).
By realizing $`(\mathrm{\Gamma },\rho ,\tau )`$ as a $`G`$-manifold, $`M_\mathrm{\Gamma }`$, we can attach to it a cohomology ring: the equivariant cohomology ring, $`H_G(M_\mathrm{\Gamma })`$. The constant map, $`M_\mathrm{\Gamma }pt`$, makes this into a module over the ring, $`H_G(pt)=\mathrm{SS}(𝔤^{})`$; and we will denote by $`H_G(\mathrm{\Gamma })`$ the torsion-free part of this module. The questions we will be concerned with in this paper are:
1. Is this module a free $`\mathrm{SS}(𝔤^{})`$-module ?
2. If so, how many generators does it have in each dimension ?
3. Is there a nice combinatorial description of these generators ?
We will now formulate these questions more precisely. We will henceforth abandon any pretense of being concerned with $`G`$-actions on manifolds and deal only with $`G`$-actions on graphs. Fortunately, there is a very beautiful description in \[GKM\] of $`H_G(\mathrm{\Gamma })`$ in terms of $`\mathrm{\Gamma }`$ and $`\alpha `$ which allows us to do this.
Fix a vector, $`\xi 𝔤`$, with the property that
$$\alpha _e(\xi )0$$
(1.1)
for all $`eE_\mathrm{\Gamma }`$. We will call such a vector *polarizing*, and denote by $`𝒫`$ the set of polarizing vectors. By axiom A2
$$\alpha _{\overline{e}}(\xi )=\alpha _e(\xi ),$$
(1.2)
so we can orient $`\mathrm{\Gamma }`$ by assigning to each unoriented edge, the orientation which makes $`\alpha _e(\xi )`$ positive. We will denote this orientation by $`o_\xi `$. To be able to do Morse theory on $`\mathrm{\Gamma }`$ we make the following essential assumption.
###### Acyclicity axiom.
For some $`\xi `$ satisfying (1.1), the orientation, $`o_\xi `$, of $`\mathrm{\Gamma }`$ that we have just described has no oriented cycles.
We will say that $`(\mathrm{\Gamma },\alpha )`$ is *$`\xi `$-acyclic* if this axiom is satisfied.
Let $`V=V_\mathrm{\Gamma }`$. The acyclicity assumption implies that one can find a function $`\varphi :V`$ with the property that, for every oriented edge, $`e`$, of $`\mathrm{\Gamma }`$, with initial vertex $`p`$ and terminal vertex $`q`$, $`\varphi (p)<\varphi (q)`$. We will say that such a function is $`\xi `$-*compatible* and call any function with this property a *Morse function*. A canonical choice of such a $`\varphi `$ is the following: For every vertex, $`p`$, consider all oriented paths in ($`\mathrm{\Gamma }`$, $`o_\xi `$) with terminal point $`p`$ and let $`\varphi (p)`$ be the length of the longest of these paths. It is clear that for this function to be unambiguously defined, $`\mathrm{\Gamma }`$ has to satisfy the acyclicity assumption. By a small perturbation if necessary, we may assume, without loss of generality, that $`\varphi `$ is injective.
Following \[GKM\] we define the *cohomology ring*, $`H_G(\mathrm{\Gamma })`$, of $`(\mathrm{\Gamma },\alpha )`$, to be the subring of the graded ring
$$\text{Maps}(V,\mathrm{SS}(𝔤^{}))=\underset{k0}{}\text{Maps}(V,\mathrm{SS}^k(𝔤^{}))$$
(1.3)
consisting of all maps, $`f:V\mathrm{SS}(𝔤^{})`$, which satisfy
$$f_pf_q(mod\alpha _e)$$
(1.4)
for every edge, $`e`$, of $`\mathrm{\Gamma }`$, $`f_p`$ and $`f_q`$ being the values of $`f`$ at the endpoints, $`p`$ and $`q`$, of $`e`$. Since $`f_p`$ and $`f_q`$ are elements of $`\mathrm{SS}(𝔤^{})`$, they can be thought of as polynomial functions on $`𝔤`$ and (1.4) asserts that the restriction of $`f_p`$ to the hyperplane $`\alpha _e=0`$ is equal to the restriction of $`f_q`$. More precisely, if $`𝔤_e^{}=𝔤^{}/\alpha _e`$, then the projection $`𝔤^{}𝔤_e^{}`$ induces an epimorphism of rings
$$r_e:\mathrm{SS}(𝔤^{})\mathrm{SS}(𝔤_e^{})$$
and condition (1.4) can be expressed as
$$r_e(f_{i(e)})=r_e(f_{t(e)}).$$
(1.5)
Since the constant maps of $`V`$ into $`\mathrm{SS}(𝔤^{})`$ satisfy (1.4), $`\mathrm{SS}(𝔤^{})`$ is a subring of $`H_G(\mathrm{\Gamma })`$; so, in particular, $`H_G(\mathrm{\Gamma })`$ is an $`\mathrm{SS}(𝔤^{})`$-module. Moreover, since it sits inside the free $`\mathrm{SS}(𝔤^{})`$-module (1.3), it is a torsion-free module. It is not, however, *a priori* clear that it is itself a free module; this “freeness” issue is the first of the three questions we will take up below.
The second question concerns the number of generators of $`H_G(\mathrm{\Gamma })`$. Let $`\varphi :V`$ be a Morse function, and, for every vertex, $`p`$, let $`ind_p\varphi `$ be the number of vertices, $`q`$, adjacent to $`p`$, with $`\varphi (q)<\varphi (p)`$. This number, called *the index of $`p`$* (and also denoted by $`\sigma _p`$), is the same as the number of oriented edges, $`e`$, with $`i(e)=p`$ and $`\alpha _e(\xi )<0`$. We define *the $`k`$-th Betti number*, $`b_k(\mathrm{\Gamma })`$, to be the number of vertices with $`ind_p\varphi =k`$ <sup>3</sup><sup>3</sup>3In \[GZ1\] and \[GZ2\], we used $`b_{2k}(\mathrm{\Gamma })`$ to denote this number; also, the $`k`$-th homogeneous component of $`H_G(\mathrm{\Gamma })`$ is denoted there by $`H^{2k}(\mathrm{\Gamma },\alpha )`$ and here by $`H_G^k(\mathrm{\Gamma })`$. One can show (\[GZ1, Theorem 2.6\]) that these numbers don’t depend on $`\xi `$ or $`\varphi `$, *i.e.* are graph-theoretic invariants. Our question about the number of generators of $`H_G(\mathrm{\Gamma })`$ can be formulated as a conjecture:
###### Conjecture 1.
The dimension of the $`k`$-th graded component of the ring
$$H_G(\mathrm{\Gamma })_{\mathrm{SS}(𝔤^{})}$$
(1.6)
is equal to $`b_k(\mathrm{\Gamma })`$.
###### Remark.
It is clear that the orientation $`o_\xi `$ depends only on the connected component of $`𝒫`$ in which $`\xi `$ sits. On the other hand it is clear that different components will give rise to different orientations. For instance, replacing $`\xi `$ with $`\xi `$ reverses the orientation and changes a vertex of index $`k`$ into a vertex of index $`dk`$. Therefore
$$b_k(\mathrm{\Gamma })=b_{dk}(\mathrm{\Gamma }).$$
(1.7)
Finally, what do the generators of the ring $`H_G(\mathrm{\Gamma })`$ actually look like ? For a vertex $`p`$ of $`\mathrm{\Gamma }`$, let $`F_p`$ be the flow-up of $`p`$, i.e. the set of vertices of the oriented graph $`(\mathrm{\Gamma },o_\xi )`$ that can be reached by a path along which the Morse function $`\varphi `$ is strictly increasing. ($`F_p`$ is the graph theoretic analogue of the “unstable manifold” at a critical point, $`p`$, in ordinary Morse theory.) We also define the flow-down of $`p`$ to be the set of vertices of the oriented graph $`(\mathrm{\Gamma },o_\xi )`$ from where we can reach $`p`$ by a path along which the Morse function $`\varphi `$ is strictly increasing. Note that the flow-down of $`p`$ with respect to $`o_\xi `$ is the flow-up of $`p`$ with respect to $`o_\xi `$. Our conjecture is that, just as in ordinary Morse theory, $`H_G(\mathrm{\Gamma })`$ is generated by cohomology classes dual to these “unstable manifolds”. More explicitly:
###### Conjecture 2.
There exists a set of independent generators
$$\tau ^{(p)}H_G^k(\mathrm{\Gamma }),k=ind_p\varphi $$
(1.8)
with $`supp\tau ^{(p)}F_p`$.
Unfortunately, neither of these conjectures is true without some extra assumptions on $`(\mathrm{\Gamma },\alpha )`$. (An example of a pair $`(\mathrm{\Gamma },\alpha )`$ for which both conjectures fail to hold is described in \[GZ1, Section 2.1\]). Therefore, making a virtue of necessity, we will declare, by definition, that $`(\mathrm{\Gamma },\alpha )`$ satisfies *the Morse package* if $`H_G(\mathrm{\Gamma })`$ is a free $`\mathrm{SS}(𝔤^{})`$-module and if the two conjectures above are true.
We can now formulate the main result of this paper. Let $`𝔨`$ be a linear subspace of $`𝔤`$, $`𝔥=𝔤/𝔨`$ and $`𝔥^{}`$ its dual, which we can identify with
$$𝔨^{}=\{\gamma 𝔤^{};\gamma (\eta )=0\text{ for all }\eta 𝔨\}𝔤^{}.$$
(1.9)
Define $`\mathrm{\Gamma }_𝔥^{}`$ to be the subgraph of $`\mathrm{\Gamma }`$ whose edges are all edges, $`e`$, of $`\mathrm{\Gamma }`$ for which $`\alpha _e𝔥^{}`$.
###### Theorem.
If $`(\mathrm{\Gamma },\alpha )`$ satisfies the Morse package, then, for every $`𝔥^{}`$ as above, $`(\mathrm{\Gamma }_𝔥^{},\alpha )`$ satisfies the Morse package. Conversely, if $`(\mathrm{\Gamma }_𝔥^{},\alpha )`$ satisfies the Morse package for all $`𝔥^{}𝔤^{}`$ of dimension 2, then $`(\mathrm{\Gamma },\alpha )`$ satisfies the Morse package.
###### Example 1.
The flag variety, $`G/B`$, $`G=SL(n,)`$. The graph $`\mathrm{\Gamma }`$ is a Cayley graph associated with the permutation group, $`S_n`$. For this graph, $`V=S_n`$, and two permutations, $`\sigma _k`$, $`k=1,2`$, are adjacent in $`\mathrm{\Gamma }`$ if and only if there exists a transposition
$$\tau _{ij}:ij,1i<jn$$
with $`\sigma _2=\sigma _1\tau _{ij}`$. Moreover, if $`e`$ is the edge joining $`\sigma `$ to $`\sigma \tau _{ij}`$, the weight labeling $`e`$ is
$$\alpha _e=\{\begin{array}{cc}ϵ_jϵ_i,\hfill & \text{ if }\sigma (j)>\sigma (i)\hfill \\ ϵ_iϵ_j,\hfill & \text{ if }\sigma (j)<\sigma (i),\hfill \end{array}$$
(1.10)
$`ϵ_1,..,ϵ_n`$ being the standard basis vectors of the lattice $`^n`$. The function
$$\varphi :V,\varphi (\sigma )=\text{length}(\sigma )$$
(1.11)
is a self-indexing Morse function on $`V`$; our theorem says that for $`(\mathrm{\Gamma },\alpha )`$ to satisfy the Morse package, it suffices to check that the subgraphs associated with the subgroups $`S_2`$ and $`S_3`$ of $`S_n`$ satisfy the Morse package. This can be done more or less by inspection. For instance the graph associated with $`S_3`$ is the permutahedron (see Figure 1) and its “Thom classes” (1.8) are given by simple monomials of degree 1, 2 and 3 in $`\alpha _1,\alpha _2`$ and $`\alpha _1+\alpha _2`$, with $`\alpha _1=ϵ_2ϵ_1`$ and $`\alpha _2=ϵ_3ϵ_2`$ (see the table below).
| | $`\tau ^1`$ | $`\tau ^{(12)}`$ | $`\tau ^{(23)}`$ | $`\tau ^{(231)}`$ | $`\tau ^{(312)}`$ | $`\tau ^{(13)}`$ |
| --- | --- | --- | --- | --- | --- | --- |
| 1 | 1 | 0 | 0 | 0 | 0 | 0 |
| (12) | 1 | $`\alpha _1`$ | 0 | 0 | 0 | 0 |
| (23) | 1 | 0 | $`\alpha _2`$ | 0 | 0 | 0 |
| (231) | 1 | $`\alpha _1\alpha _2`$ | $`\alpha _2`$ | $`\alpha _2(\alpha _1+\alpha _2)`$ | 0 | 0 |
| (312) | 1 | $`\alpha _1`$ | $`\alpha _1\alpha _2`$ | 0 | $`\alpha _1(\alpha _1+\alpha _2)`$ | 0 |
| (13) | 1 | $`\alpha _1\alpha _2`$ | $`\alpha _1\alpha _2`$ | $`\alpha _2(\alpha _1+\alpha _2)`$ | $`\alpha _1(\alpha _1+\alpha _2)`$ | $`\alpha _1\alpha _2(\alpha _1+\alpha _2)`$ |
###### Example 2.
The Grassmannian $`Gr^k(^n)`$. Here the graph in question is the *Johnson graph* (see \[BCN\]). Its vertices are the $`k`$-element subsets, $`S`$, of the $`n`$-element set $`\{1,..,n\}`$, and two vertices, $`S_k`$, $`k=1,2`$, are adjacent if $`S_1S_2`$ is a $`(k1)`$-element set, or, in other words, if $`S_1S_2`$ and $`S_2S_1`$ are one-element sets. If $`e`$ is the oriented set joining $`S_1`$ to $`S_2`$ the weight labeling this edge is
$$\alpha _e=\alpha _i\alpha _j,$$
(1.12)
where $`\{i\}=S_1S_2`$, $`\{j\}=S_2S_1`$ and $`\alpha _1,..,\alpha _n`$ are the standard basis vectors of $`^n`$. To define a Morse function on $`V`$ we first observe that there is a one-to-one correspondence between the vertices of $`\mathrm{\Gamma }`$ and Young diagrams (\[Fu\]). Namely, let $`S`$ be a $`k`$-element subset of $`\{1,..,n\}`$, with elements $`i_1<i_2<..<i_k`$. Then we make correspond to $`S`$ the Young diagram, $`\sigma _S`$, with rows of length
$$i_rr,r=k,k1,\mathrm{},1.$$
(1.13)
By means of this correspondence we can define a self-indexing Morse function, $`\varphi `$, on $`V`$, by setting
$$\varphi (S)=\text{ number of boxes in the diagram }\sigma _S.$$
(1.14)
For the Johnson graph, our theorem says that to verify the Morse package for $`\mathrm{\Gamma }`$, it suffices to verify the Morse package for the Johnson graph associated with the two-element subsets of $`\{1,2,3\}`$, and just as in the previous example, this can be done more or less by inspection.
A few words about the contents of this paper: We mentioned above that the axioms A1-A3 can be translated into axioms on the axial function, $`\alpha `$. In Section 2 we will do this and also give more precise definitions of some of the concepts that we have mentioned in this introduction. In Section 3 we will prove the easy part of our main theorem, the “only if” part. The hard part of the theorem, the “if” part, requires some algebraic results about Vandermonde matrices which we will discuss in the Appendix. Another ingredient in our proof is an *integration* operation on graphs. Given an element, $`f`$, of the ring (1.3), we define its *integral* over the graph to be the sum
$$_\mathrm{\Gamma }f=\underset{pV}{}\frac{f_p}{\alpha _e},$$
(1.15)
the denominator in the $`p^{th}`$ summand being the product over all oriented edges, $`e`$, with $`i(e)=p`$. The expression on the right appears to belong to the quotient field of $`\mathrm{SS}(𝔤^{})`$; however, if $`fH_G(\mathrm{\Gamma })`$, one can show that this sum is in $`\mathrm{SS}(𝔤^{})`$. Moreover, $`H_G(\mathrm{\Gamma })`$ has the following duality property:
###### Proposition 1.1.
If $`f\text{Maps}(V,\mathrm{SS}(𝔤^{}))`$ and
$$_\mathrm{\Gamma }fh\mathrm{SS}(𝔤^{})$$
for all $`hH_G(\mathrm{\Gamma })`$, then $`fH_G(\mathrm{\Gamma })`$.
These results are old results of ours from \[GZ1\], but to prove the main theorem, we will need to generalize them to *hypergraphs*. (A hypergraph is, like a graph, an object consisting of vertices, hyperedges and an incidence relation which tells one when a vertex is incident to a hyperedge. For hypergraphs, however, more than two vertices can be incident to a hyperedge.)
A third ingredient in our proof is a “cross-section” construction for graphs which we already made extensive use of in \[GZ2\]. A real number, $`c`$, is a *critical value* of the Morse function $`\varphi `$ if $`c=\varphi (p)`$ for some $`pV`$; otherwise, it is a *regular value*. Let $`c`$ be a regular value and let $`V_c`$ be the set of oriented edges, $`e`$, of $`\mathrm{\Gamma }`$, which intersect the level set $`\varphi =c`$, in the sense that $`\varphi (p)<c<\varphi (q),`$ $`p=i(e)`$ and $`q=t(e)`$ being the vertices of $`e`$. Then $`V_c`$ is the vertex set of a hypergraph, $`\mathrm{\Gamma }_c`$, whose edges are the intersections of the level set $`\varphi =c`$ with the connected components of the graphs $`\mathrm{\Gamma }_𝔥^{}`$, $`𝔥^{}`$ being a dimension 2 subspace of $`𝔤^{}`$.
The trickiest and most subtle part of our proof is figuring out how to define the cohomology ring of this hypergraph. One can equip $`\mathrm{\Gamma }_c`$ with an analogue of the axial function, $`\alpha `$, (see Section 5) and mimic the definition (1.4) of $`H_G(\mathrm{\Gamma })`$; however, this turns out *not* to be the right approach. Trial and error show that the right approach is to mimic the definition of $`H_G(\mathrm{\Gamma })`$ “by duality” in Proposition 1.1. (See Section 4 for details). This approach, however, has one apparent short-coming: The equivariant cohomology theory of graphs that we described above is a *local* theory in the sense that a prospective cohomology class, $`f\text{Maps }(V,\mathrm{SS}(𝔤^{}))`$, belongs to $`H_G(\mathrm{\Gamma })`$ if and only if it satisfies condition (1.4) at each edge. If we define hypergraph cohomology by duality it is not at all clear that it will be a local theory in this sense. The localization theorem which we prove in Section 5 shows, however, that it is.
Denoting the cohomology ring of $`\mathrm{\Gamma }_c`$ by $`H(\mathrm{\Gamma }_c,\alpha _c)`$, we show in Section 7 how the proof of the main theorem can be reduced to the following problem: How does the structure of $`H(\mathrm{\Gamma }_c,\alpha _c)`$ change as one passes through a critical value of $`\varphi `$ ? If $`c`$ and $`c^{}`$ are regular values of $`\varphi `$ and there is just one critical point, $`pV_\mathrm{\Gamma }`$, with $`c<\varphi (p)<c^{}`$, how are $`H(\mathrm{\Gamma }_c,\alpha _c)`$ and $`H(\mathrm{\Gamma }_c^{},\alpha _c^{})`$ related ? The main technical result of this paper answers this question by showing that the elements of $`H(\mathrm{\Gamma }_c^{},\alpha _c^{})`$ are expressible in terms of the elements of $`H(\mathrm{\Gamma }_c,\alpha _c)`$ via an “integral” equation of Vandermonde type.
## 2. The Main Result
Let $`\mathrm{\Gamma }`$ be a regular $`d`$-valent graph, $`V`$ its set of vertices, and, for each vertex $`pV`$, let $`E_p`$ be the set of oriented edges with initial vertex $`p`$. Define the “tangent bundle” of $`\mathrm{\Gamma }`$ to be ($`E_\mathrm{\Gamma }`$, $`\pi `$), with projection $`\pi :E_\mathrm{\Gamma }V_\mathrm{\Gamma }`$, $`\pi (e)=i(e)`$. Let $`𝔤^{}`$ be an $`n`$-dimensional real vector space.
###### Definition 2.1.
An *abstract one-skeleton* is a pair $`(\mathrm{\Gamma },\alpha )`$ consisting of a $`d`$-valent graph, $`\mathrm{\Gamma }`$, and a function
$$\alpha :E_\mathrm{\Gamma }𝔤^{}$$
(called axial function), satisfying the axioms:
1. : For every $`pV`$, the vectors $`\{\alpha _e;eE_p\}`$ are pairwise linearly independent.
2. : If $`eE_\mathrm{\Gamma }`$ then
$$\alpha _{\overline{e}}=\alpha _e.$$
3. : Let $`eE_\mathrm{\Gamma }`$ with $`p=i(e)`$ and $`p^{}=t(e)`$. Let $`E_p=\{e_i;i=1,..,d\}`$ and $`E_p^{}=\{e_i^{};i=1,..,d\}`$. Then we can order $`E_p`$ and $`E_p^{}`$ such that $`e_d=e`$, $`e_d^{}=\overline{e}`$ and
$$\alpha _{e_i^{}}=\alpha _{e_i}+c_{i,e}\alpha _e,\text{with }c_{i,e}.$$
If $`𝔥^{}`$ is a linear subspace of $`𝔤^{}`$, we define $`\mathrm{\Gamma }_𝔥^{}`$ to be the subgraph of $`\mathrm{\Gamma }`$ whose edges are all edges $`e`$ for which $`\alpha _e𝔥^{}`$. Then $`(\mathrm{\Gamma }_𝔥^{},\alpha )`$ is an abstract one-skeleton, whose axial function takes values in $`𝔥^{}`$. Note that if $`𝔨`$ is a subspace of $`𝔤`$ and $`𝔥=𝔤/𝔨`$ then $`𝔥^{}`$ is canonically a subspace of $`𝔤^{}`$ and
$$\mathrm{\Gamma }_𝔥^{}\mathrm{\Gamma }_{(𝔤/𝔨)^{}}.$$
###### Definition 2.2.
A vector $`\xi 𝔤`$ is called generic if for every vertex $`pV`$ and every four edges $`e_1,e_2,e_3`$ and $`e_4`$ in $`E_p`$ such that $`e_1e_2`$, $`e_3e_4`$ and $`(e_1,e_2)(e_3,e_4)`$, we have
$$\frac{\alpha _{e_1}}{\alpha _{e_1}(\xi )}\frac{\alpha _{e_2}}{\alpha _{e_2}(\xi )}\frac{\alpha _{e_3}}{\alpha _{e_3}(\xi )}\frac{\alpha _{e_4}}{\alpha _{e_4}(\xi )}.$$
The set of generic elements, $`𝒫_0`$, is dense in $`𝔤`$.
From now on, we assume that $`(\mathrm{\Gamma },\alpha )`$ is a $`\xi `$-acyclic one-skeleton for some generic and polarizing $`\xi `$ and that $`\varphi :V`$ is $`\xi `$-compatible. Without loss of generality, we can assume that $`\varphi `$ is injective. Let $`H(\mathrm{\Gamma },\alpha )`$ be the $`\mathrm{SS}(𝔤^{})`$-module defined by (1.4).
###### Definition 2.3.
An element $`\tau ^{(p)}`$ of $`H(\mathrm{\Gamma },\alpha )`$ is called *a generating class at* $`p`$ if it satisfies the conditions:
1. $`\tau ^{(p)}`$ is supported on the flow-up at $`p`$, $`F_p`$
2. $`\tau ^{(p)}(p)=^{}\alpha _e`$
the product in the second condition being over the edges $`eE_p`$ with $`\alpha _e(\xi )<0`$. A set $`\{\tau ^{(p)}\}_{pV}`$, containing one generating class for each vertex, is called *a generating family*.
###### Theorem 2.1.
The one skeleton $`(\mathrm{\Gamma },\alpha )`$ admits a generating family if and only if for every $`m0`$,
$$dimH^m(\mathrm{\Gamma },\alpha )=\underset{k=0}{\overset{d}{}}b_k(\mathrm{\Gamma })\lambda _{mk,n},$$
(2.1)
where
$$\lambda _j=\lambda _{j,n}=\{\begin{array}{cc}dim\mathrm{SS}^j(𝔤^{})\hfill & ,\text{ if }j0\hfill \\ 0\hfill & ,\text{ if }j<0.\hfill \end{array}$$
###### Proof.
The equivalence follows from Theorems 2.4.2 and 2.4.4 of \[GZ2\]. ∎
###### Remark.
A generating family is a basis of $`H(\mathrm{\Gamma },\alpha )`$ as an $`\mathrm{SS}(𝔤^{})`$-module. So, if a generating family exists, $`H(\mathrm{\Gamma },\alpha )`$ is a free $`\mathrm{SS}(𝔤^{})`$-module.
The main result of this paper is the following theorem:
###### Theorem 2.2.
The one-skeleton $`(\mathrm{\Gamma },\alpha )`$ admits a generating family if and only if for every two-dimensional subspace $`𝔥^{}𝔤^{}`$, every connected component of the induced one-skeleton $`(\mathrm{\Gamma }_𝔥^{},\alpha )`$ admits a generating family.
One also has the following sharpening of the “if” part of this theorem:
###### Theorem 2.3.
Suppose that for every $`qF_p\{p\}`$, the index of $`q`$ is strictly greater that the index of $`p`$. Then the class $`\tau ^{(p)}`$ is unique.
In view of Theorem 2.1, to prove Theorem 2.2, it suffices to show that for every $`m`$, the dimension of $`H^m(\mathrm{\Gamma },\alpha )`$ is given by (2.1). To compute this dimension, we first realize $`(\mathrm{\Gamma },\alpha )`$ as a cross-section of a product one-skeleton $`\mathrm{\Gamma }^{\mathrm{}}=\mathrm{\Gamma }\times L`$ using a process akin to “symplectic cutting” (\[Le\]). To each cross-section we attach a module and we show that the module we associate to a cross-section “isomorphic” to $`(\mathrm{\Gamma },\alpha )`$ is the same as the cohomology ring of $`(\mathrm{\Gamma },\alpha )`$. When the level is just above the lowest vertex, the cross-section is very simple and we determine the associated module directly. Then we determine how this module changes when we pass from one cross-section to another, going over a vertex. Finally, we compute the dimension of $`H^m(\mathrm{\Gamma },\alpha )`$ by starting with the dimension of the first cross-section and counting the changes that occur until we have reached the cross-section of $`\mathrm{\Gamma }^{\mathrm{}}`$ that is the same as $`(\mathrm{\Gamma },\alpha )`$.
## 3. The “only if” part of Theorem 2.2
Let $`𝔥^{}`$ be a two-dimensional subspace of $`𝔤^{}`$, $`𝔤`$ be the dual of $`𝔤^{}`$ and $`𝔨`$ be the annihilator of $`𝔥^{}`$ in $`𝔤`$. Choose $`𝔥`$ to be a complementary subspace to $`𝔨`$ in $`𝔤`$; then we can identify $`𝔥^{}`$ with the vector space dual to $`𝔥`$.
Now let $`p`$ be a vertex of $`(\mathrm{\Gamma }_𝔥^{},\alpha )`$ and let $`\tau `$ be the Thom class, in $`H(\mathrm{\Gamma },\alpha )`$, associated with the flow-up of $`p`$. Thus
$$\tau _p=\underset{i=1}{\overset{r}{}}\alpha _{e_i},$$
(3.1)
where $`r`$ is the index of $`p`$ and $`e_1`$,…,$`e_r`$ are the “downward-pointing” edges at $`p`$, *i.e.* $`\alpha _{e_i}(\xi )<0`$. Suppose that $`\alpha _{e_i}𝔥^{}`$ for $`i=1,..,s`$ and $`\alpha _{e_i}𝔥^{}`$ for $`i=s+1,..,r`$. Choose $`\xi 𝔨`$ such that $`\alpha _{e_i}(\xi )0`$ for $`i=s+1,..,r`$. Given $`f\mathrm{SS}(𝔤^{})`$, let $`f^\mathrm{\#}`$ be the restriction of this polynomial function on $`𝔤`$ to the two-dimensional affine subspace
$$\{\xi \}\times 𝔥𝔨\times 𝔥=𝔨𝔥=𝔤.$$
Then we can think of $`f^\mathrm{\#}`$ as an element of $`\mathrm{SS}(𝔥^{})`$.
###### Lemma 3.1.
Let $`V_𝔥^{}`$ be the set of vertices of $`\mathrm{\Gamma }_𝔥^{}`$. For $`hH(\mathrm{\Gamma },\alpha )`$, let $`h^\mathrm{\#}:V_𝔥^{}\mathrm{SS}(𝔥^{})`$ be defined by
$$(h^\mathrm{\#})(p)=(h(p))^\mathrm{\#}.$$
Then $`h^\mathrm{\#}H(\mathrm{\Gamma }_𝔥^{},\alpha )`$.
###### Proof.
If $`p`$ and $`q`$ are in $`V_𝔥^{}`$ and are joined by an edge $`e`$ of $`\mathrm{\Gamma }_𝔥^{}`$, then
$$h(p)h(q)=f\alpha _e$$
for some $`f\mathrm{SS}(𝔤^{})`$, so that
$$h^\mathrm{\#}(p)h^\mathrm{\#}(q)=f^\mathrm{\#}(\alpha _e)^\mathrm{\#}.$$
However, since $`\alpha _e𝔥^{}`$, we have $`(\alpha _e)^\mathrm{\#}=\alpha _e`$. ∎
Consider, in particular, $`\tau ^\mathrm{\#}`$. By (3.1),
$$\tau _p^\mathrm{\#}=\left(\underset{i=s+1}{\overset{r}{}}\alpha _{e_i}^\mathrm{\#}\right)\underset{i=1}{\overset{s}{}}\alpha _{e_i}.$$
Moreover, for $`s+1ir`$,
$$\alpha _{e_i}^\mathrm{\#}=c_i+\beta _{e_i},$$
where $`c_i=\alpha _{e_i}(\xi )0`$ and $`\beta _{e_i}𝔥^{}`$. Thus, if $`c`$ is the product of the $`c_i`$’s,
$$\tau _p^\mathrm{\#}=c\underset{i=1}{\overset{s}{}}\alpha _{e_i}+g,\text{ with }g\underset{js+1}{}\mathrm{SS}^j(𝔥^{}).$$
Let $`\tau _𝔥^{}^{(p)}`$ be the homogeneous component of degree $`s`$ of the cohomology class $`c^1\tau ^\mathrm{\#}`$. Then
$$\tau _𝔥^{}^{(p)}(p)=\underset{i=1}{\overset{s}{}}\alpha _{e_i}$$
and, since $`\tau ^\mathrm{\#}`$ is supported on the flow-up of $`p`$ in $`\mathrm{\Gamma }_𝔥^{}`$, so is $`\tau _𝔥^{}^{(p)}`$.
## 4. Cross-sections
We will say that an axial function, $`\alpha `$, is 3-independent if, for every $`pV`$ and every triple of edges, $`e_iE_p`$, $`i=1,2,3`$, the vectors $`\alpha _{e_i}`$ are linearly independent. In \[GZ2\] the “if” part of Theorem 2.2 was proved modulo this 3-independence assumption. We will *not* make this assumption here, however, we will make use of a key ingredient in our earlier proof: Fixing a regular value of $`\varphi `$, $`c\varphi (V)`$, we will define the $`c`$-cross-section, $`V_c`$, of $`\mathrm{\Gamma }`$, to be the set of oriented edges, $`e`$, of $`\mathrm{\Gamma }`$, with end points $`p=i(e)`$ and $`q=t(e)`$, for which
$$\varphi (p)<c<\varphi (q).$$
(4.1)
In \[GZ2\] we showed that $`V_c`$ is the set of vertices of a graph, $`\mathrm{\Gamma }_c`$, whose edges are defined as follows: For every two dimensional subspace, $`𝔥^{}`$, of $`𝔤^{}`$, and every connected component, $`\mathrm{\Gamma }_𝔥^{}^0`$, of $`\mathrm{\Gamma }_𝔥^{}`$, let $`V_c(\mathrm{\Gamma }_𝔥^{}^0)`$ be the edges of $`\mathrm{\Gamma }_𝔥^{}`$ satisfying (4.1). If $`\alpha `$ is 3-independent and if the “only if” part of Theorem 2.2 holds, these subsets of $`V_c`$ are of cardinality 2 and are by definition the edges of $`\mathrm{\Gamma }_c`$.
If we drop the 3-independence assumption we can still define these sets to be the *hyperedges* of $`\mathrm{\Gamma }_c`$, but the object we get will no longer be a graph but, instead, a *hypergraph*. Nonetheless many of the results of \[GZ2\] will continue to hold. We will describe some of these results in this section.
Let $`𝔤_\xi ^{}`$ be the annihilator of $`\xi `$ in $`𝔤^{}`$, $`\mathrm{SS}(𝔤_\xi ^{})`$ be the symmetric algebra of $`𝔤_\xi ^{}`$ and $`Q(𝔤_\xi ^{})`$ be the quotient field of $`\mathrm{SS}(𝔤_\xi ^{})`$. We define a morphism of graded rings
$$𝒦_c:H(\mathrm{\Gamma },\alpha )\text{Maps}(V_c,\mathrm{SS}(𝔤_\xi ^{}))$$
(4.2)
as follows: For $`eV_c`$ let $`p`$ and $`q`$ be the vertices of $`e`$. Since $`\alpha _e(\xi )0`$, the projection $`𝔤^{}𝔤_e^{}`$ maps $`𝔤_\xi ^{}`$ bijectively onto $`𝔤_e^{}`$, so one has a composite map
$$𝔤^{}𝔤_e^{}𝔤_\xi ^{}$$
and hence an induced morphism of graded rings:
$$\mathrm{SS}(𝔤^{})\mathrm{SS}(𝔤_e^{})\mathrm{SS}(𝔤_\xi ^{}).$$
If $`f`$ is in $`H(\mathrm{\Gamma },\alpha )`$, the images of $`f_p`$ and $`f_q`$ in $`\mathrm{SS}(𝔤_e^{})`$ are the same by (1.5) and hence so are their images in $`\mathrm{SS}(𝔤_\xi ^{})`$. We define $`𝒦_c(f)(e)`$ to be this common image and call the map $`𝒦_c`$ the Kirwan map.
Let $`\{x,y_1,\mathrm{},y_{n1}\}`$ be a basis of $`𝔤^{}`$ such that $`x(\xi )=1`$ and $`\{y_1,\mathrm{},y_{n1}\}`$ is a basis of $`𝔤_\xi ^{}`$. Let $`\alpha =\alpha (\xi )(x\beta (y))𝔤^{}`$ such that $`\alpha (\xi )0`$. Then the map
$$\rho _\alpha :\mathrm{SS}(𝔤^{})\mathrm{SS}(𝔤_\xi ^{}),$$
given via the identification $`𝔤_\xi ^{}𝔤^{}/\alpha `$, will send $`x`$ to $`x\alpha /\alpha (\xi )𝔤_\xi ^{}`$ and $`y_j`$ to $`y_j`$. Therefore $`\rho _\alpha `$ will send a polynomial $`P(x,y)\mathrm{SS}(𝔤^{})`$ to the polynomial $`P(x\alpha /\alpha (\xi ),y)=P(\beta (y),y)\mathrm{SS}(𝔤_\xi ^{})`$.
Hence, if $`\alpha _e=m_e(x\beta _e(y))`$, then
$$𝒦_c(f)(e)=f_p(\beta _e(y),y)\mathrm{SS}(𝔤_\xi ^{}).$$
Let $`eV_c`$, with endpoints $`p=i(e)`$ and $`q=t(e)`$, such that $`\varphi (p)<c<\varphi (q)`$. If $`e_i`$, $`i=1,..,d1`$ are the other edges issuing from $`p`$ and $`e_i^{}`$, $`i=1,..,d1`$ are the other edges issuing from $`q`$, we define the Thom class of $`e`$ in $`\mathrm{\Gamma }`$, $`\tau _eH(\mathrm{\Gamma },\alpha )`$, by
$$\tau _e(v)=\{\begin{array}{cc}\alpha _{e_i}\hfill & \text{ if }v=p\hfill \\ \alpha _{e_i^{}}\hfill & \text{ if }v=q\hfill \\ 0\hfill & \text{ otherwise}.\hfill \end{array}$$
Let
$$\alpha _i^\mathrm{\#}=𝒦_c(\alpha _i)=\alpha _i\frac{\alpha _i(\xi )}{\alpha _e(\xi )}\alpha _e$$
and
$$\delta _e=(m_e𝒦_c(\tau _e))^1=(m_e\underset{i=1}{\overset{d1}{}}\alpha _i^\mathrm{\#})^1Q(𝔤_\xi ^{})$$
(4.3)
where $`m_e=\alpha _e(\xi )>0`$.
We now define the integral operation
$$_{\mathrm{\Gamma }_c}:\text{Maps}(V_c,Q(𝔤_\xi ^{}))Q(𝔤_\xi ^{})$$
as integration with respect to the discrete measure on $`\mathrm{\Gamma }_c`$ with density $`\delta _e`$ at $`e`$, i.e.
$$_{\mathrm{\Gamma }_c}f=\underset{eV_c}{}\delta _ef(e)=\underset{eV_c}{}\frac{f(e)}{m_e_{i=1}^{d1}\alpha _i^\mathrm{\#}}.$$
(4.4)
One of the main results in \[GZ1\] (Theorem 2.5) is :
###### Theorem 4.1.
If $`fH(\mathrm{\Gamma },\alpha )`$ then
$$_{\mathrm{\Gamma }_c}𝒦_c(f)\mathrm{SS}(𝔤_\xi ^{}).$$
We will use this property to associate an $`\mathrm{SS}(𝔤_\xi ^{})`$-module to the cross-section.
###### Definition 4.1.
We denote by $`H(\mathrm{\Gamma }_c,\alpha _c)`$ the set of all maps $`f:V_cQ(𝔤_\xi ^{})`$ with the property that
$$_{\mathrm{\Gamma }_c}f𝒦_c(h)\mathrm{SS}(𝔤_\xi ^{})$$
for all $`hH(\mathrm{\Gamma },\alpha )`$.
Remarks:
1. Let $`fH(\mathrm{\Gamma }_c,\alpha _c)`$. Then for every edge $`e\mathrm{\Gamma }_c`$,
$$f(e)=m_e_{\mathrm{\Gamma }_c}f𝒦_c(\tau _e)\mathrm{SS}(𝔤_\xi ^{}).$$
Therefore $`H(\mathrm{\Gamma }_c,\alpha _c)\text{Maps}(V_c,\mathrm{SS}(𝔤_\xi ^{}))`$.
2. Since the Kirwan map (4.2) is a morphism of rings, from Theorem 4.1 follows that the image of the Kirwan map is included in $`H(\mathrm{\Gamma }_c,\alpha _c)`$.
###### Example 3.
Let $`p`$ be the vertex of $`\mathrm{\Gamma }`$ on which $`\varphi `$ attains its minimum and let $`c`$ be a regular value such that $`p`$ is the only vertex below the $`c`$-level. Then the $`c`$-cross section consists of the oriented edges with initial vertex at $`p`$.
For such an edge, $`e_i`$, let
$$\alpha _{e_i}=m_i(x\beta _i(y)),$$
with $`m_i=\alpha _{e_i}(\xi )`$ and $`\beta _i(y)\mathrm{SS}^1(𝔤_\xi ^{})`$. Consider $`\tau :V_c𝔤_\xi ^{}`$, $`\tau (e_i)=\beta _i(y)`$.
Let $`g:V_c\mathrm{SS}(𝔤_\xi ^{})`$. For $`hH(\mathrm{\Gamma },\alpha )`$, let $`P=h(p)\mathrm{SS}(𝔤^{})\mathrm{SS}(𝔤_\xi ^{})[x]`$. Then
$$_{\mathrm{\Gamma }_c}g𝒦_c(h)=(\underset{i=1}{\overset{d}{}}m_i)^1_{V_c}gP(\tau ),$$
where the integral on the right hand side is defined as in (8.12) in the Appendix. Therefore, by Lemma 8.4 in the Appendix, $`g`$ is an element of $`H(\mathrm{\Gamma }_c,\alpha _c)`$ if and only if there exist $`g_0,\mathrm{},g_{d1}\mathrm{SS}(𝔤_\xi ^{})`$ such that
$$g=\underset{k=0}{\overset{d1}{}}g_k\tau ^k.$$
(4.5)
We conclude that
$$\text{dim}H^m(\mathrm{\Gamma }_c,\alpha _c)=\underset{k=0}{\overset{d1}{}}\lambda _{mk,n}.$$
Moreover, let $`gH(\mathrm{\Gamma }_c,\alpha _c)`$ be given by (4.5) and consider
$$h=\underset{k=0}{\overset{d1}{}}g_kx^k\mathrm{SS}(𝔤_\xi ^{})[x]\mathrm{SS}(𝔤^{})H(\mathrm{\Gamma },\alpha ).$$
Then $`g=𝒦_c(h)`$; this proves that for this cross-section, the Kirwan map $`𝒦_c`$ is surjective.
## 5. The localization theorem for $`H(\mathrm{\Gamma }_c,\alpha _c)`$
At the beginning of Section 4 we pointed out that the set $`V_c`$ can be made into the set of vertices of a hypergraph, $`\mathrm{\Gamma }_c`$. In this section we will describe the structure of this hypergraph in more detail and give an alternative description of $`H(\mathrm{\Gamma }_c,\alpha _c)`$, which is more in the spirit of the description (1.4) of $`H(\mathrm{\Gamma },\alpha )`$.
Recall that an edge of $`\mathrm{\Gamma }_c`$ is a subset of $`V_c`$ of the form
$$E=V_c(\mathrm{\Gamma }_𝔥^{}^0),$$
(5.1)
where $`𝔥^{}`$ is a 2-dimensional subspace of $`𝔤^{}`$, spanned by axial vectors $`\alpha _{e_1}`$ and $`\alpha _{e_2}`$, $`e_iE_p`$, and $`\mathrm{\Gamma }_𝔥^{}^0`$ is the connected component of $`\mathrm{\Gamma }_𝔥^{}`$ containing $`p`$. If one assigns to the edge (5.1) the multiplicity
$$\mu _E=(\text{ valence of }\mathrm{\Gamma }_𝔥^{}^0)1,$$
then $`\mathrm{\Gamma }_c`$ becomes a $`(d1)`$-valent hypergraph: each vertex is the point of intersection of $`d1`$ edges, counting multiplicities. Let us label each of the edges, $`E`$, by a non-zero vector, $`\alpha _E`$, in the one-dimensional vector space $`𝔥^{}𝔤_\xi ^{}`$. This labeling is not unique since $`\alpha _E`$ is only defined up to a scalar; however, one can show that it satisfies axioms similar to the axioms A1-A3 of definition 2.1 (see \[GZ2, Section 2.1\]). Now fix a vector $`\gamma 𝔤_\xi ^{}0`$ and let $`\mathrm{\Gamma }_c^\gamma `$ be the subgraph of $`\mathrm{\Gamma }_c`$ consisting of the edges, $`E`$, of $`\mathrm{\Gamma }_c`$, with $`\alpha _E\gamma `$.
###### Lemma 5.1.
The hypergraph $`\mathrm{\Gamma }_c^\gamma `$ is totally disconnected: no two distinct edges, $`E_1`$ and $`E_2`$, of $`\mathrm{\Gamma }_c^\gamma `$, contain a common vertex.
###### Proof.
Let $`v`$ be a vertex of $`\mathrm{\Gamma }_c^\gamma `$, corresponding to an edge $`e`$ of $`\mathrm{\Gamma }`$. Let $`E`$ be an edge of $`\mathrm{\Gamma }_c^\gamma `$ incident to $`v`$ and let $`𝔥^{}`$ be the two-dimensional subspace of $`𝔤^{}`$ such that $`E=V_c(\mathrm{\Gamma }_𝔥^{}^0)`$. Then $`𝔥^{}`$ is generated by $`\gamma `$ and $`\alpha _e`$, therefore is uniquely determined. ∎
As in (5.1), let $`𝔥^{}`$ be the 2-dimensional subspace of $`𝔤^{}`$ spanned by vectors $`\alpha _{e_1},\alpha _{e_2}`$, $`e_1,e_2E_p`$ and let $`\mathrm{\Gamma }_𝔥^{}^0`$ be the connected component of $`\mathrm{\Gamma }_𝔥^{}`$ containing $`p`$.
###### Definition 5.1.
The *Thom class* of $`\mathrm{\Gamma }_𝔥^{}^0`$,
$$\tau =\tau (\mathrm{\Gamma }_𝔥^{}^0)H(\mathrm{\Gamma },\alpha ),$$
is the class which assigns to each vertex $`q`$ of $`\mathrm{\Gamma }_𝔥^{}^0`$ the product
$$\underset{e_iE_qE_q(\mathrm{\Gamma }_𝔥^{}^0)}{}\alpha _{e_i}$$
and assigns zero to every other vertex of $`\mathrm{\Gamma }`$.
The existence of such a class enables one to define an injection
$$H(\mathrm{\Gamma }_𝔥^{},\alpha )H(\mathrm{\Gamma },\alpha ),ff\tau .$$
Moreover, from $`\tau `$ one gets an important identity relating integration over $`\mathrm{\Gamma }_c`$ and integration over the edges (5.1) of $`\mathrm{\Gamma }_c`$:
###### Lemma 5.2.
Given $`gH(\mathrm{\Gamma }_𝔥^{}^0,\alpha )`$ and $`f:V_c\mathrm{SS}(𝔤_\xi ^{})`$, one has
$$_Ef_E𝒦_c(g)=_{\mathrm{\Gamma }_c}f𝒦_c(\tau g),$$
(5.2)
where $`f_E`$ is the restriction of $`f`$ to $`E`$ and the Kirwan map $`𝒦_c`$ on the left hand side is the map
$$𝒦_c:H(\mathrm{\Gamma }_𝔥^{}^0,\alpha )H((\mathrm{\Gamma }_𝔥^{}^0)_c,\alpha _c).$$
We will use Lemma 5.2 to prove:
###### Theorem 5.1.
A map $`f:V_c\mathrm{SS}(𝔤_\xi ^{})`$ is in $`H(\mathrm{\Gamma }_c,\alpha _c)`$ if and only if its restriction to every edge $`E`$ of the hypergraph $`\mathrm{\Gamma }_c`$ is in $`H(E,\alpha _c)`$.
###### Proof.
Assume first that $`fH(\mathrm{\Gamma }_c,\alpha _c)`$ and let $`E`$ be an edge of $`\mathrm{\Gamma }_c`$, given by (5.1). Denote by $`f_E`$ the restriction of $`f`$ to $`E`$. For every $`gH(\mathrm{\Gamma }_𝔥^{}^0,\alpha )`$ we have
$$_Ef_E𝒦_c(g)=_{\mathrm{\Gamma }_c}f𝒦_c(\tau g)\mathrm{SS}(𝔤_\xi ^{}),$$
hence $`f_EH(E,\alpha _c)`$.
Conversely, assume that $`f_EH(E,\alpha _c)`$ for every edge $`E`$ of $`\mathrm{\Gamma }_c`$. Let $`gH(\mathrm{\Gamma },\alpha )`$. We need to show that
$$_{\mathrm{\Gamma }_c}f𝒦_c(g)\mathrm{SS}(𝔤_\xi ^{}).$$
(5.3)
where, according to (4.4),
$$_{\mathrm{\Gamma }_c}f𝒦_c(g)=\underset{eV_c}{}\delta (e)f(e)𝒦_c(g)(e).$$
(5.4)
Let $`𝔤_\xi ^{}`$ be the set of all linear factors that appear in the denominators of $`\delta (e)`$ (see (4.3)), for all edges $`e`$ that intersect the $`c`$-level. For $`\gamma `$, let $`_\gamma `$ be the multiplicative system in $`\mathrm{SS}(𝔤_\xi ^{})`$ generated by elements of $``$ that are not multiples of $`\gamma `$ and let $`\mathrm{SS}(𝔤_\xi ^{})_\gamma `$ be the corresponding localized ring. To show (5.3) it suffices to show that
$$_{\mathrm{\Gamma }_c}f𝒦_c(g)\mathrm{SS}(𝔤_\xi ^{})_\gamma $$
(5.5)
for every $`\gamma `$.
Fix $`\gamma `$ and let $`\mathrm{\Gamma }_c^\gamma `$ be the corresponding subgraph of $`\mathrm{\Gamma }_c`$. Let $`e`$ be a vertex of $`\mathrm{\Gamma }_c^\gamma `$ corresponding to an edge of $`\mathrm{\Gamma }`$, and let $`E=V_c(\mathrm{\Gamma }_𝔥^{}^0)`$ be an edge of $`\mathrm{\Gamma }_c^\gamma `$ incident to $`e`$, where $`𝔥^{}`$ is a two-dimensional subspace of $`𝔤^{}`$. Then $`e`$ is an edge in both $`\mathrm{\Gamma }`$ and $`\mathrm{\Gamma }_𝔥^{}^0`$; the density associated to $`e`$ when we integrate over $`\mathrm{\Gamma }_c`$ is given by (4.3) and the density associated to $`e`$ when we integrate over $`E`$ is
$$\delta _e^{}=(m_e𝒦_c(\tau _e^{}))^1,$$
(5.6)
where $`\tau _e^{}`$ is the Thom class of $`e`$ in $`\mathrm{\Gamma }_𝔥^{}^0`$. If $`\tau H(\mathrm{\Gamma },\alpha )`$ is the Thom class of $`\mathrm{\Gamma }_𝔥^{}^0`$ and
$$𝔫_E(e)=𝒦_c(\tau )(e)_\gamma ,\text{ for all }eE$$
(5.7)
then the two densities are related by
$$\delta _e=\frac{\delta _e^{}}{𝔫_E(e)}.$$
(5.8)
Then (5.4) differs from
$$\underset{E}{}\underset{eE}{}\delta _ef(e)𝒦_c(g)(e)=\underset{E}{}\underset{eE}{}\frac{\delta _e^{}f(e)𝒦_c(g)(e)}{𝔫_E(e)}$$
(5.9)
by a term in $`\mathrm{SS}(𝔤_\xi ^{})_\gamma `$ (the first sums on each side are over all edges of $`\mathrm{\Gamma }_c^\gamma `$).
But
$$\underset{eE}{}\frac{\delta _e^{}f(e)𝒦_c(g)(e)}{𝔫_E(e)}=\frac{1}{_{eE}𝔫_E(e)}\underset{eE}{}\left(\delta _e^{}f(e)𝒦_c(g)(e)\underset{e^{}e}{}𝔫_E(e^{})\right),$$
(5.10)
and, according to Lemma 8.2 of the Appendix, there exists a polynomial $`P_E\mathrm{SS}(𝔤_\xi ^{})[X]`$ such that
$$\underset{e^{}e}{}𝔫_E(e^{})=P_E(𝔫_E(e));$$
(5.11)
then, since $`𝒦_c`$ is a morphism of rings, (5.9) becomes
$$\underset{E}{}\frac{1}{_{eE}𝔫_E(e)}_Ef_E𝒦_c(gP_E(\tau ))\mathrm{SS}(𝔤_\xi ^{})_\gamma ,$$
(5.12)
which concludes the proof. ∎
## 6. Cutting
Let $`L`$ be a one-valent graph, with two vertices, labeled 0 and 1, and one edge connecting them. Consider the product graph $`\mathrm{\Gamma }^{\mathrm{}}=\mathrm{\Gamma }\times L`$, with the set of vertices $`V^{\mathrm{}}`$; we define an axial function
$$\alpha ^{\mathrm{}}:E_\mathrm{\Gamma }^{\mathrm{}}(𝔤)^{}𝔤^{}^{},$$
by
$`\alpha _{(p,0)(q,0)}^{\mathrm{}}`$ $`=\alpha _{pq},`$ $`\text{if }pqE_\mathrm{\Gamma }`$
$`\alpha _{(p,1)(q,1))}^{\mathrm{}}`$ $`=\alpha _{pq},`$ $`\text{if }pqE_\mathrm{\Gamma }`$
$`\alpha _{(p,0)(p,1)}^{\mathrm{}}`$ $`=\mathrm{𝟏}^{};`$
(this would correspond to the product action of $`G\times S^1`$ on $`\mathrm{\Gamma }^{\mathrm{}}=\mathrm{\Gamma }\times L`$).
Then the pair $`(\mathrm{\Gamma }^{\mathrm{}},\alpha ^{\mathrm{}})`$ is a one-skeleton, acyclic with respect to the generic vector $`(\xi ,1)𝔤`$. Choose $`a>\varphi _{max}\varphi _{min}>0`$ and define $`\mathrm{\Phi }:V^{\mathrm{}}`$ by
$$\mathrm{\Phi }(p,t)=\varphi (p)+at$$
Then $`\mathrm{\Phi }`$ is $`(\xi ,1)`$-compatible.
Let $`c(\varphi _{max},\varphi _{min}+a)`$. The cross section $`V_c^{\mathrm{}}`$ consists of oriented edges of type $`(p,0)(p,1)`$ and we can naturally identify it with $`V`$, the set of vertices of $`\mathrm{\Gamma }`$. We can make $`V_c^{\mathrm{}}`$ into a graph, $`\mathrm{\Gamma }_c^{\mathrm{}}`$, by joining the points corresponding to $`(p,0)(p,1)`$ and $`(q,0)(q,1)`$ if and only if $`p`$ and $`q`$ are joined by an edge of $`\mathrm{\Gamma }`$.
###### Theorem 6.1.
If $`c(\varphi _{max},\varphi _{min}+a)`$ then
$$H(\mathrm{\Gamma }_c^{\mathrm{}},\alpha _c^{\mathrm{}})H(\mathrm{\Gamma },\alpha )$$
###### Proof.
There is a natural isomorphism of linear spaces $`𝔤^{}(𝔤)_{(\xi ,1)}^{}𝔤^{}^{}`$, given by
$$\sigma (\sigma ,\sigma (\xi )\mathrm{𝟏}).$$
This induces an isomorphism of rings
$$\rho :\mathrm{SS}((𝔤)_{(\xi ,1)}^{})\mathrm{SS}(𝔤^{}).$$
and, together with the identification $`V_c^{\mathrm{}}V`$, an isomorphism
$$\rho _{}:\text{Maps}(V_c^{\mathrm{}},\mathrm{SS}((𝔤)_{(\xi ,1)}^{}))\text{Maps}(V,\mathrm{SS}(𝔤^{})).$$
(6.1)
We will show that (6.1) restricts to a bijection
$$\rho _{}:H(\mathrm{\Gamma }_c^{\mathrm{}},\alpha _c^{\mathrm{}})H(\mathrm{\Gamma },\alpha ).$$
First, let $`gH(\mathrm{\Gamma }_c^{\mathrm{}},\alpha _c^{\mathrm{}})`$. Let $`e`$ be an edge of $`\mathrm{\Gamma }`$, with endpoints $`p=i(e)`$ and $`q=t(e)`$ and let $`\mathrm{\Gamma }_e`$ be the subgraph of $`\mathrm{\Gamma }^{\mathrm{}}`$ with vertices $`(p,0),(p,1),(q,0)`$ and $`(q,1)`$. Consider the Thom class of $`\mathrm{\Gamma }_e`$ in $`\mathrm{\Gamma }^{\mathrm{}}`$. This, by definition, is the map
$$h_e:V^{\mathrm{}}\mathrm{SS}((𝔤)^{})$$
that is 0 at vertices not in $`\mathrm{\Gamma }_e`$ and, at each vertex, $`v`$, of $`\mathrm{\Gamma }_e`$, is equal to the product of the values of $`\alpha ^{\mathrm{}}`$ on edges at $`v`$ which are not edges of $`\mathrm{\Gamma }_e`$. Then
$$\frac{\rho _{}(g)(p)\rho _{}(g)(q)}{\alpha _e}=\rho \left(_{\mathrm{\Gamma }_c^{\mathrm{}}}g𝒦_c(h_e)\right)\mathrm{SS}(𝔤^{}).$$
Since this is true for all edges of $`\mathrm{\Gamma }`$, it follows that $`\rho _{}(g)H(\mathrm{\Gamma },\alpha )`$.
Conversely, let $`gH(\mathrm{\Gamma },\alpha )`$ and $`hH(\mathrm{\Gamma }^{\mathrm{}},\alpha ^{\mathrm{}})`$. The projection on the first factor $`𝔤^{}^{}𝔤^{}`$ induces a map $`\pi _1:\mathrm{SS}((𝔤)^{})\mathrm{SS}(𝔤^{})`$. Let
$$h_0=\pi _1h|_{\mathrm{\Gamma }\times \{0\}};$$
then $`h_0H(\mathrm{\Gamma },\alpha )`$ and therefore $`gh_0H(\mathrm{\Gamma },\alpha )`$.
A direct computation shows that
$$\rho \left(_{\mathrm{\Gamma }_c^{\mathrm{}}}\rho _{}^1(g)𝒦_c(h)\right)=_\mathrm{\Gamma }gh_0.$$
Theorem 2.2 in \[GZ1\] implies that the right hand side is an element of $`\mathrm{SS}(𝔤^{})`$ and hence
$$_{\mathrm{\Gamma }_c^{\mathrm{}}}\rho _{}^1(g)𝒦_c(h)\mathrm{SS}((𝔤)_{(\xi ,1)}^{}),$$
for all $`hH(\mathrm{\Gamma },\alpha )`$. Therefore
$$\rho _{}^1(g)H(\mathrm{\Gamma }_c^{\mathrm{}},\alpha _c^{\mathrm{}}),$$
which proves the theorem. ∎
## 7. The changes in cohomology
Let $`c`$ and $`c^{}`$ be regular values of $`\varphi `$ such that $`c<c^{}`$ and such that there is exactly one vertex, $`p`$, with $`c<\varphi (p)<c^{}`$. Let $`r`$ be the index of $`p`$, and let $`s=dr`$. Let $`e_i`$, $`i=1,..,r`$ be the edges, $`e`$, issuing from $`p`$, for which $`\alpha _{p,e}(\xi )<0`$ and let $`e_a`$, $`a=r+1,..,d`$ be the other edges issuing from $`p`$. We assume that $`1rd1`$, so that $`d1s1`$. Let $`\mathrm{\Delta }_c=\{e_i;i=1,..,r\}V_c`$ and $`\mathrm{\Delta }_c^{}=\{e_a;a=r+1,..,d\}V_c^{}`$. Define
$$V^\mathrm{\#}=(V_c\mathrm{\Delta }_c)(\mathrm{\Delta }_c\times \mathrm{\Delta }_c^{})$$
and a map
$$\pi _c:V^\mathrm{\#}V_c$$
which is the identity on $`V_c\mathrm{\Delta }_c`$ and the projection on the first factor on $`\mathrm{\Delta }_c\times \mathrm{\Delta }_c^{}`$. Note that
$$V^\mathrm{\#}=(V_c^{}\mathrm{\Delta }_c^{})(\mathrm{\Delta }_c\times \mathrm{\Delta }_c^{})$$
and that there exists a similar map
$$\pi _c^{}:V^\mathrm{\#}V_c^{}.$$
We define the pull-back maps
$$\pi _c^{}:\text{Maps}(V_c,\mathrm{SS}(𝔤_\xi ^{}))\text{Maps}(V^\mathrm{\#},\mathrm{SS}(𝔤_\xi ^{}))$$
and
$$\pi _c^{}^{}:\text{Maps}(V_c^{},\mathrm{SS}(𝔤_\xi ^{}))\text{Maps}(V^\mathrm{\#},\mathrm{SS}(𝔤_\xi ^{})).$$
The projections of $`\mathrm{\Delta }_c\times \mathrm{\Delta }_c^{}`$ onto factors define pull-back maps
$$\text{pr}_1^{}:\text{Maps}(\mathrm{\Delta }_c,\mathrm{SS}(𝔤_\xi ^{}))\text{Maps}(V^\mathrm{\#},\mathrm{SS}(𝔤_\xi ^{}))$$
and
$$\text{pr}_2^{}:\text{Maps}(\mathrm{\Delta }_c^{},\mathrm{SS}(𝔤_\xi ^{}))\text{Maps}(V^\mathrm{\#},\mathrm{SS}(𝔤_\xi ^{})),$$
by extending with 0 outside $`\mathrm{\Delta }_c\times \mathrm{\Delta }_c^{}`$. Since the maps $`\pi _c^{},\pi _c^{}^{},\text{pr}_1^{}`$ and $`\text{pr}_2^{}`$ are injective, we can regard the sets, $`\text{Maps}(V_c,\mathrm{SS}(𝔤_\xi ^{}))`$, $`\text{Maps}(V_c^{},\mathrm{SS}(𝔤_\xi ^{}))`$, $`\text{Maps}(\mathrm{\Delta }_c,\mathrm{SS}(𝔤_\xi ^{}))`$ and $`\text{Maps}(\mathrm{\Delta }_c^{},\mathrm{SS}(𝔤_\xi ^{}))`$ as subsets of $`\text{Maps}(V^\mathrm{\#},\mathrm{SS}(𝔤_\xi ^{}))`$.
Let
$`\alpha _{e_i}`$ $`=m_i(x\beta _i(y))`$ and
$`\alpha _{e_a}`$ $`=m_a(x\beta _a(y)),`$
with $`m_i<0<m_a`$ and $`\beta _i(y),\beta _a(y)𝔤_\xi ^{}`$, for $`i=1,..,r`$ and $`a=r+1,..,d`$. Consider the maps
$`\tau _c:`$ $`\mathrm{\Delta }_c𝔤_\xi ^{},`$ $`\tau _c(e_i)`$ $`=\beta _i(y);`$
$`\tau _c^{}:`$ $`\mathrm{\Delta }_c^{}𝔤_\xi ^{},`$ $`\tau _c^{}(e_a)`$ $`=\beta _a(y),\text{and}`$
$`\tau ^\mathrm{\#}:`$ $`\mathrm{\Delta }_c\times \mathrm{\Delta }_c^{}𝔤_\xi ^{},`$ $`\tau ^\mathrm{\#}(e_i,e_a)`$ $`=\beta _i(y)\beta _a(y).`$
###### Remark.
The fact that $`\xi `$ is generic implies that $`\tau ^\mathrm{\#}(e_i,e_a)0`$.
Let $`H(\mathrm{\Delta }_c,\tau _c)`$ be the ring associated to the finite set $`\mathrm{\Delta }_c`$ and the injective function $`\tau _c:\mathrm{\Delta }_c𝔤_\xi ^{}`$ (see the Appendix, Definition 8.1).
###### Lemma 7.1.
If $`fH(\mathrm{\Gamma },\alpha )`$ then $`f|_{\mathrm{\Delta }_c}H(\mathrm{\Delta }_c,\tau _c)`$.
###### Proof.
The proof of this “localization” theorem is similar to that of the localization theorem in Section 5. Let $`f_0=f|_{\mathrm{\Delta }_c}`$ be the restriction of $`f`$ to $`\mathrm{\Delta }_c`$. To show that $`f_0H(\mathrm{\Delta }_c,\tau _c)`$ we need to show that
$$_{\mathrm{\Delta }_c}f_0h(\tau _c)\mathrm{SS}(𝔤_\xi ^{})$$
(7.1)
for all $`h\mathrm{SS}(𝔤_\xi ^{})[Y]`$, and just as in the proof of Theorem 5.1, we will show that this can be “localized” to analogous statements about the integrals of $`f_0h(\tau _c)`$ over the hyperedges of the hypergraph $`\mathrm{\Delta }_c`$. Here are the details.
Let $`=\{\beta _i\beta _j;ij\}𝔤_\xi ^{}`$. For $`\gamma `$, let $`_\gamma `$ be the multiplicative system in $`\mathrm{SS}(𝔤_\xi ^{})`$ generated by elements of $``$ which are not collinear with $`\gamma `$ and let $`\mathrm{SS}(𝔤_\xi ^{})_\gamma `$ be the corresponding localized ring. We will show that
$$_{\mathrm{\Delta }_c}f_0h(\tau _c)\mathrm{SS}(𝔤_\xi ^{})_\gamma $$
(7.2)
for each $`\gamma `$; this will imply (7.1).
Let $`\gamma =\beta _i\beta _j`$. We join two points $`v_k`$ and $`v_l`$ of $`\mathrm{\Delta }_c`$ if $`\gamma `$ and $`\beta _k\beta _l`$ are collinear.
Let $`(\mathrm{\Delta }_c)_\gamma ^0=\{v_{j_1},..,v_{j_k}\}`$ be a non-trivial connected component of this new graph. Then $`(\mathrm{\Delta }_c)_\gamma ^0`$ is a complete graph. Let $`𝔥^{}`$ be the two-dimensional subspace of $`𝔤^{}`$ generated by $`\alpha _{j_1}`$ and $`\alpha _{j_2}`$ and let $`\mathrm{\Gamma }_𝔥^{}^0`$ be the connected component of $`\mathrm{\Gamma }_𝔥^{}`$ that contains $`p`$. Then
$$(V_𝔥^{}^0)_c\mathrm{\Delta }_c=(\mathrm{\Delta }_c)_\gamma ^0.$$
We now use our hypothesis : $`\mathrm{\Gamma }_𝔥^{}^0`$ admits a generating family. Then the same is true if we change $`\xi `$ to $`\xi `$; let $`\tau _𝔥^{}^{(p)}H(\mathrm{\Gamma }_𝔥^{},\alpha )`$ be a generating class at $`p`$ with respect to $`o_\xi `$ and define
$$\psi =\psi _𝔥^{}\tau _𝔥^{}^{(p)}H(\mathrm{\Gamma },\alpha ),$$
where $`\psi _𝔥^{}`$ is the Thom class of $`\mathrm{\Gamma }_𝔥^{}^0`$ in $`\mathrm{\Gamma }`$. Then $`\psi `$ is supported on the flow-down of $`p`$ in $`\mathrm{\Gamma }_𝔥^{}^0`$ and
$$\psi (p)=\underset{t\{j_1,..,j_k\}}{}\alpha _{e_t}.$$
For every $`P\mathrm{SS}(𝔤^{})\mathrm{SS}(𝔤_\xi ^{})[Y]`$ we have
$$_{(\mathrm{\Delta }_c)_\gamma ^0}fP(\tau _c)=(\underset{t\{j_1,\mathrm{},j_k\}}{}m_t)_{\mathrm{\Gamma }_c}f𝒦_c(P\psi );$$
(7.3)
since $`fH(\mathrm{\Gamma }_c,\alpha _c)`$, (7.3) implies that $`f|_{(\mathrm{\Delta }_c)_\gamma ^0}H((\mathrm{\Delta }_c)_\gamma ^0,\tau _c)`$.
Let $`R\mathrm{SS}(𝔤_\xi ^{})[X]`$ be the polynomial
$$R(X)=\underset{t\{j_1,..,j_k\}}{}(X\beta _t)$$
and let $`T_0\mathrm{SS}(𝔤_\xi ^{})[X_1,\mathrm{},X_{k1}]`$ be given by
$$T_0=\underset{l=1}{\overset{k1}{}}R(X_l).$$
Then $`T_0(\mathrm{SS}(𝔤_\xi ^{})[X_1,\mathrm{},X_{k1}])^{\mathrm{\Sigma }_{k1}}`$, hence, according to Lemma 8.2 of the Appendix, there exists $`T(\mathrm{SS}(𝔤_\xi ^{})[X_1,\mathrm{},X_k])^{\mathrm{\Sigma }_k}[Y]`$ such that $`T_0=T(X_k)`$. In other words, by inserting $`X_k`$ for $`Y`$ in this polynomial of $`k+1`$ variables we get back our original polynomial in $`k1`$ variables. Using this we deduce that
$$\underset{l=1}{\overset{k}{}}\frac{f(e_{j_l})h(\beta _{j_l})}{_{tj_l}(\beta _{j_l}\beta _t)}=\frac{1}{_{l=1}^kR(\beta _{j_l})}_{(\mathrm{\Delta }_c)_\gamma ^0}(\mathrm{\Psi }h)(\tau _c)f,$$
(7.4)
where $`\mathrm{\Psi }\mathrm{SS}(𝔤_\xi ^{})[Y]\mathrm{SS}(𝔤^{})`$ is the polynomial obtained by replacing $`X_l`$ with $`\beta _{j_l}`$, for all $`l=1,.,k`$.
Since $`_{l=1}^kR(\beta _{j_l})_\gamma `$, using (7.3) we conclude that the left hand side of (7.4) is an element of $`\mathrm{SS}(𝔤_\xi ^{})_\gamma `$. Now (7.2) follows from the fact that the integral of $`f_0h(\tau _c)`$ is a sum of terms of this form. ∎
###### Lemma 7.2.
For every $`fH(\mathrm{\Gamma }_c,\tau _c)`$ and every $`f_iH(\mathrm{\Delta }_c,\alpha _c)`$, $`i=1,..,s1`$, there exist unique $`f^{}H(\mathrm{\Gamma }_c^{},\alpha _c^{})`$ and $`f_j^{}H(\mathrm{\Delta }_c^{},\tau _c^{})`$, $`j=1,..,r1`$, such that
$$f^{}+\underset{j=1}{\overset{r1}{}}(\tau ^\mathrm{\#})^jf_j^{}=f+\underset{i=1}{\overset{s1}{}}(\tau ^\mathrm{\#})^if_i.$$
(7.5)
###### Proof.
Uniqueness: It is clear that $`f^{}`$ should be equal to $`f`$ on $`\mathrm{\Gamma }_c\mathrm{\Delta }_c=\mathrm{\Gamma }_c^{}\mathrm{\Delta }_c^{}`$. If we restrict (7.5) to each set of the form $`\mathrm{\Delta }_c\times \{e_a\}`$, we get a linear system whose matrix is a non-singular Vandermonde matrix; therefore, the solution is unique. Hence there exist unique $`f^{}\text{Maps}(\mathrm{\Gamma }_c^{},Q(𝔤_\xi ^{}))`$ and $`f_j^{}\text{Maps}(\mathrm{\Delta }_c^{},Q(𝔤_\xi ^{}))`$ which satisfy (7.5).
Existence: We need to show that $`f^{}H(\mathrm{\Gamma }_c^{},\alpha _c^{})`$ and $`f_j^{}H(\mathrm{\Delta }_c^{},\tau _c^{})`$, where $`f`$ and $`f_j^{}`$ are the ones obtained above.
Using Lemma 8.3 and Corollary 8.1 of the Appendix, we deduce that
$$f_j^{}=\underset{m^{}}{}\left(\underset{m,k}{}P_{m,m^{},k}(y)(_{\mathrm{\Delta }_c}\tau _c^mf_k)\right)(\tau _c^{})^m^{},$$
where $`P_{m,m^{},k}(y)\mathrm{SS}(𝔤_\xi ^{})`$; since $`f_kH(\mathrm{\Delta }_c,\tau _c)`$ for all $`k=0,..,s1`$, we can use Lemma 8.4 to conclude that $`f_j^{}H(\mathrm{\Delta }_c^{},\tau _c^{})`$.
Let $`hH(\mathrm{\Gamma },\alpha )`$. Again, we use Lemma 8.3 and Corollary 8.1 to obtain that
$$_{\mathrm{\Gamma }_c^{}}f^{}𝒦_c^{}(h)=_{\mathrm{\Gamma }_c}f𝒦_c(h)+\underset{i=0}{\overset{s1}{}}_{\mathrm{\Delta }_c}P_i(\tau _c)f_i,$$
where $`P_i\mathrm{SS}(𝔤_\xi ^{})[X]`$. We now use that $`fH(\mathrm{\Gamma }_c,\alpha _c)`$, $`f_iH(\mathrm{\Delta }_c,\tau _c)`$ for all $`i=0,..,s1`$, and Lemma 8.4 to conclude that
$$_{\mathrm{\Gamma }_c^{}}f^{}𝒦_c^{}(h)\mathrm{SS}(𝔤_\xi ^{})$$
for all $`hH(\mathrm{\Gamma },\alpha )`$, i.e. that $`f^{}H(\mathrm{\Gamma }_c^{},\alpha _c^{})`$. ∎
We now use this lemma to determine the change in the cohomology of the cross-section as we pass over the vertex $`p`$.
###### Corollary 7.1.
For every $`m0`$ we have
$$dimH^m(\mathrm{\Gamma }_c^{},\alpha _c^{})=dimH^m(\mathrm{\Gamma }_c,\alpha _c)+\underset{k=0}{\overset{s1}{}}\lambda _{mk,n1}\underset{k=0}{\overset{r1}{}}\lambda _{mk,n1}.$$
(7.6)
###### Proof.
When $`0<\sigma (p)<d`$, this follows immediately from (7.5) and (8.14). When $`\sigma (p)=0`$ it follows from (8.14) and
$$H(\mathrm{\Gamma }_c^{},\alpha _c^{})=H(\mathrm{\Gamma }_c,\alpha _c)H(\mathrm{\Delta }_c^{},\alpha _c^{})$$
and when $`\sigma (p)=d`$ it follows from (8.14) and
$$H(\mathrm{\Gamma }_c,\alpha _c)=H(\mathrm{\Gamma }_c^{},\alpha _c^{})H(\mathrm{\Delta }_c,\alpha _c).\mathit{}$$
## 8. Adding-up dimensions
To compute the dimension of $`H^m(\mathrm{\Gamma },\alpha )`$, we will apply Corollary 7.1 several times to the cross-sections of $`\mathrm{\Gamma }^{\mathrm{}}=\mathrm{\Gamma }\times L`$.
Let $`c_0>\varphi _{min}`$ such that there is only one vertex, $`(p,t)`$, of $`\mathrm{\Gamma }^{\mathrm{}}=\mathrm{\Gamma }\times L`$, with $`\mathrm{\Phi }(p,t)<c_0`$. Then
$$dimH^m(\mathrm{\Gamma }_{c_0}^{\mathrm{}},\alpha _{c_0}^{\mathrm{}})=\underset{k=0}{\overset{d}{}}\lambda _{mk,n}.$$
If $`c_1(\varphi _{max},\varphi _{min}+a)`$ then, according to Theorem 6.1, we get
$$dimH^m(\mathrm{\Gamma }_{c_1}^{\mathrm{}},\alpha _{c_1}^{\mathrm{}})=dimH^m(\mathrm{\Gamma },\alpha ).$$
Now let $`c_0<a<b<c_1`$ such that there is exactly one vertex of $`\mathrm{\Gamma }`$, $`p`$, such that $`a<\mathrm{\Phi }(p,0)<b`$. If the index of $`p`$ in $`\mathrm{\Gamma }`$ is $`\sigma (p)=r`$ then the index of $`(p,0)`$ in $`\mathrm{\Gamma }^{\mathrm{}}`$ is also $`r`$. Thus we can apply (7.6) to obtain
$$dimH^m(\mathrm{\Gamma }_b^{\mathrm{}},\alpha _b^{\mathrm{}})=dimH^m(\mathrm{\Gamma }_a^{\mathrm{}},\alpha _a^{\mathrm{}})+\underset{k=1}{\overset{dr}{}}\lambda _{mk,n}\underset{k=1}{\overset{r1}{}}\lambda _{mk,n}.$$
Adding together these changes we get
$$dimH^m(\mathrm{\Gamma },\alpha )=\underset{pV}{}\left(\underset{k=0}{\overset{d\sigma (p)}{}}\lambda _{mk}\underset{k=0}{\overset{\sigma (p)1}{}}\lambda _{mk}\right)$$
The minimum value for $`k`$ is 0 and the maximum is $`d`$; $`\lambda _{mk}`$ appears in the first sum when $`\sigma (p)dk`$ and in the second one when $`\sigma (p)k+1`$. Therefore
$$dimH^m(\mathrm{\Gamma },\alpha )=\underset{k=0}{\overset{d}{}}\left(\underset{l=0}{\overset{dk}{}}b_l(\mathrm{\Gamma })\underset{l=k+1}{\overset{d}{}}b_l(\mathrm{\Gamma })\right)\lambda _{mk}.$$
Because $`b_{dl}(\mathrm{\Gamma })=b_l(\mathrm{\Gamma })`$ (see (1.7)), the expression in bracket reduces to $`b_k(\mathrm{\Gamma })`$ and therefore
$$dimH^m(\mathrm{\Gamma },\alpha )=\underset{k=0}{\overset{d}{}}b_k(\mathrm{\Gamma })\lambda _{mk,n}.$$
This concludes the proof of Theorem 2.2.
###### Remark.
The results of \[GZ2, Section 2.5.2\] are valid in this case as well. In particular, the dimension of $`H^m(\mathrm{\Gamma }_c,\alpha _c)`$ is the same as the dimension of the image of the Kirwan map
$$𝒦_c:H^m(\mathrm{\Gamma },\alpha )\text{Maps}(V_c,\mathrm{SS}^m(𝔤_\xi ^{})).$$
Since $`\text{im}(𝒦_c)H(\mathrm{\Gamma }_c,\alpha _c)`$, it follows that the Kirwan map is surjective. Thus, in particular, $`H(\mathrm{\Gamma }_c,\alpha _c)`$ is not only an $`\mathrm{SS}(𝔤_\xi ^{})`$-module, but also a ring.
## Appendix
###### Lemma 8.1.
$$\underset{k=1}{\overset{m}{}}\frac{X_k^N}{_{jk}(X_kX_j)}=\underset{\begin{array}{c}i_1+..+i_m=Nm+1\\ i_1,..,i_m0\end{array}}{}X_1^{i_1}\mathrm{}X_m^{i_m}.$$
(8.1)
###### Proof.
Consider the decomposition in partial fractions
$$\frac{Z^N}{(ZX_1)\mathrm{}(ZX_m)}=Q(Z)+\underset{k=1}{\overset{m}{}}\frac{X_k^N}{_{jk}(X_kX_j)}\frac{1}{ZX_k},$$
where $`Q(Z)`$ is a polynomial in $`Z`$. We use the expansion
$$\frac{1}{ZX_k}=\underset{i_k=0}{\overset{\mathrm{}}{}}X_k^{i_k}Z^{1i_k},$$
(8.2)
to obtain that:
$$Z^{Nm}\underset{k=1}{\overset{m}{}}\left(\underset{i_k=0}{\overset{\mathrm{}}{}}X_k^{i_k}Z^{i_k}\right)=Q(Z)+\frac{1}{Z}\underset{k=1}{\overset{m}{}}\left(\frac{X_k^N}{_{jk}(X_kX_j)}\underset{l=0}{\overset{\mathrm{}}{}}X_k^lZ^l\right).$$
Now the formula (8.1) follows by comparing coefficients of $`Z^1`$ on both sides. ∎
###### Corollary 8.1.
Let R be a ring and $`PR[X_1,..,X_m][Y]`$. Then
$$\underset{k=1}{\overset{m}{}}\frac{P(X_k)}{_{jk}(X_kX_j)}R[X_1,..,X_m].$$
###### Lemma 8.2.
Let $`P_0([X_1,..,X_{m1}])^{\mathrm{\Sigma }_{m1}}`$ be a symmetric polynomial. Then there exists $`P([X_1,..,X_m])^{\mathrm{\Sigma }_m}[Y]`$ such that $`P_0=P(X_m)`$.
###### Proof.
It suffices to show that the lemma is true if $`P_0`$ is a fundamental symmetric polynomial in $`X_1,..,X_{m1}`$ and for this we can use an inductive argument.
We can also start with the identity
$$(Z+X_1)\mathrm{}(Z+X_{m1})=\frac{(Z+X_1)\mathrm{}(Z+X_m)}{Z}\frac{1}{1+X_mZ^1}$$
and we use (8.2) to deduce that
$$(Z+X_1)\mathrm{}(Z+X_{m1})=(\underset{k=0}{\overset{m}{}}\sigma _kZ^{mk})(\underset{j=0}{\overset{\mathrm{}}{}}(1)^jX_m^jZ^{j1}),$$
where $`\sigma _k([X_1,..,X_m])^{\mathrm{\Sigma }_m}`$ is the fundamental symmetric polynomial of degree $`k`$. The lemma follows by comparing coefficients of $`Z^k`$ on both sides. ∎
###### Lemma 8.3.
Let $`R=[X_1,\mathrm{},X_m]`$ and
$$A=\left(\begin{array}{cccc}1& X_1& \mathrm{}& X_1^{m1}\\ 1& X_2& \mathrm{}& X_2^{m1}\\ 3\\ 1& X_m& \mathrm{}& X_m^{m1}\end{array}\right)$$
be a Vandermonde matrix. Then $`A`$ is invertible in the space of matrices with entries in the quotient field of $`R`$ and for each $`i=1,..,m`$, there exists $`P_iR[Y]`$ such that
$$(A^1)_{ij}=\frac{P_i(X_j)}{_{kj}(X_jX_k)}.$$
Moreover, for $`i=1`$ we have:
$$(A^1)_{1j}=\underset{kj}{}\left(\frac{X_k}{X_jX_k}\right).$$
(8.3)
###### Proof.
Since
$$detA=\underset{1k<lm}{}(X_lX_k)0,$$
(8.4)
the matrix $`A`$ is invertible; the entries of the inverse are
$$(A^1)_{ij}=(1)^{i+j}\frac{detA_{ji}}{detA},$$
(8.5)
where $`A_{ji}`$ is obtained from $`A`$ by removing the $`j^{th}`$ row and $`i^{th}`$ column.
We start with the identity
$$(ZX_1)\mathrm{}(ZX_{j1})(ZX_{j+1})\mathrm{}(ZX_m)=\underset{k=0}{\overset{m1}{}}(1)^{m1k}\sigma _{m1k}^jZ^k,$$
(8.6)
where $`\sigma _{m1k}^j`$ is the fundamental symmetric polynomial of degree $`m1k`$ in variables $`X_1,..,X_{j1},X_{j+1},..,X_m`$. Since the left hand side of (8.6) is 0 for $`Z=X_l`$, $`lj`$, we get:
$$X_l^{m1}=\underset{k=0}{\overset{m2}{}}(1)^{mk}\sigma _{mk1}^jX_l^k.$$
(8.7)
The left hand side of (8.7) is the last column of $`A_{ji}`$; using (8.7) and basic properties of determinants we deduce that
$$detA_{ji}=\sigma _{mi}^j\underset{\begin{array}{c}1k<lm\\ kjl\end{array}}{}(X_lX_k).$$
(8.8)
Using (8.4) and (8.8) in (8.5) we obtain that
$$(A^1)_{ij}=\frac{(1)^{n+i}\sigma _{ni}^j}{_{kj}(X_jX_k)}.$$
(8.9)
We now use Lemma 8.2 to finish the proof. Since $`\sigma _{n1}^j=X_1\mathrm{}X_{j1}X_{j+1}\mathrm{}X_m`$, for $`i=1`$, (8.9) becomes (8.3). ∎
###### Remark.
The result of Lemma 8.2 can be made more precise. By induction on $`k`$ we get that
$$\sigma _k^j=\underset{l=0}{\overset{k}{}}(1)^l\sigma _{kl}X_j^l.$$
(8.10)
Then (8.9) becomes:
$$(A^1)_{ij}=\frac{_{k=0}^{ni}(1)^{n+i+k}\sigma _{nik}X_j^k}{_{kj}(X_jX_k)}.$$
(8.11)
Let $`W`$ be an $`n`$-dimensional vector space, $`\mathrm{\Delta }=\{v_1,\mathrm{},v_d\}`$ be a finite set and $`\tau :\mathrm{\Delta }W`$ be an injective function. Let $`\mathrm{SS}(W)`$ be the symmetric algebra and $`Q(W)`$ be the quotient field of $`\mathrm{SS}(W)`$.
We define an integral operation $`_\mathrm{\Delta }:\text{Maps}(\mathrm{\Delta },Q(W))Q(W)`$ by
$$_\mathrm{\Delta }g=\underset{k=1}{\overset{d}{}}\frac{g(v_k)}{_{jk}(\tau (v_k)\tau (v_j))}.$$
(8.12)
###### Definition 8.1.
We define $`H(\mathrm{\Delta },\tau )`$ to be the space of all maps $`g`$ satisfying the condition
$$_\mathrm{\Delta }gP(\tau )\mathrm{SS}(W),\text{ for all }P\mathrm{SS}(W)[Y].$$
###### Lemma 8.4.
A map $`g:\mathrm{\Delta }Q(W)`$ is in $`H(\mathrm{\Delta },\tau )`$ if and only if there exist $`g_0,\mathrm{},g_{d1}\mathrm{SS}(W)`$ such that
$$g=\underset{k=0}{\overset{d1}{}}g_k\tau ^k.$$
(8.13)
###### Proof.
The fact that every map of the form (8.13) is in $`H(\mathrm{\Delta },\tau )`$ is a direct consequence of Corollary 8.1.
To show that every element of $`H(\mathrm{\Delta },\tau )`$ can be written as in (8.13) we proceed as follows: We can regard (8.13) as a linear system whose matrix is a Vandermonde matrix; hence there are $`g_0,..,g_{d1}Q(W)`$ such that (8.13) is true. Moreover, we can use Lemma 8.3 to deduce that
$$g_i=_\mathrm{\Delta }gP_{i+1}(\tau ).$$
Since $`gH(\mathrm{\Delta },\tau )`$, we conclude that $`g_i\mathrm{SS}(W)`$. ∎
Remarks:
1. $`H(\mathrm{\Delta },\tau )`$ is a graded ring and an $`\mathrm{SS}(W)`$-module.
2. If $`gH(\mathrm{\Delta },\tau )`$ then $`g\text{Maps}(\mathrm{\Delta },\mathrm{SS}(W))`$.
3. If $`gH(\mathrm{\Delta },\tau )`$ then the polynomials $`g_0,..,g_{d1}`$ are unique.
4. Let $`H^m(\mathrm{\Delta },\tau )=H(\mathrm{\Delta },\tau )\text{Maps}(\mathrm{\Delta },\mathrm{SS}^m(W))`$. Then
$$\text{dim}H^m(\mathrm{\Delta },\tau )=\underset{k=0}{\overset{d1}{}}\lambda _{mk,n}.$$
(8.14)
|
warning/0007/gr-qc0007051.html
|
ar5iv
|
text
|
# General relativistic dynamics of compact binaries at the third post-Newtonian order
## I Introduction
The present work is a contribution to the problem of the dynamics of two compact objects at the so-called third post-Newtonian (3PN) approximation of general relativity. By 3PN we mean the relativistic corrections in the binary’s equations of motion corresponding to the order $`1/c^6`$ relatively to the Newtonian acceleration, when the speed of light $`c`$ tends to infinity. Why studying the equations of motion to such a frightful post-Newtonian order? A side reason is the strange beauty of the post-Newtonian expansion, which becomes quite intricate at the 3PN order, where it requires some interesting mathematical methods. The main reason, however, is that inspiralling compact binaries, namely systems of two neutron stars or black holes (or one of each) moving on a relativistic orbit prior to their final merger, should be routinely observed by the gravitational-wave detectors LIGO, VIRGO and their fellows. Several analysis show that the post-Newtonian templates required for the detection and parameter extraction of inspiralling compact binaries should include the relativistic corrections in the binary’s orbital phase at approximately the level of the 3PN order .
Lorentz and Droste were the first to obtain the correct equations of motion of two non-spinning particles at the 1PN approximation (see for reviews). An important work by Einstein, Infeld and Hoffmann showed that the 1PN acceleration can in fact be deduced from the vacuum gravitational field outside the masses. This result is interesting because, in their approach, the bodies are allowed to carry a strong internal gravity. Unfortunately, the computation of the surface integrals surrounding the masses is very difficult even at the 1PN order (see for a recent derivation of the Einstein-Infeld-Hoffmann equations). The same equations were also obtained by Fock and followers for the motion of the centers of mass of bodies with finite size. The next approximation, 2PN, has been tackled by Otha, Okamura, Kimura and Hiida with a direct post-Newtonian computation of the Hamiltonian of $`N`$ point-particles; however the first complete 2-particle case in their framework is only given by Damour and Schäfer , and the fully explicit 3-particle case is due to Schäfer . Up to the 2PN level, the equations of motion are conservative (existence of ten conserved quantities, including a conserved energy). The non-conservative effect, which is associated to the radiation reaction force, arises at the 2.5PN order. The first correct equations of motion of two masses at the 2.5PN order were obtained by Damour, Deruelle and collaborators in harmonic coordinates. These equations are applicable to systems of strongly self-gravitating bodies such as neutron stars (see Damour for the proof). Moreover, Kopejkin and Grishchuk and Kopejkin obtained the same equations in the case of weakly self-gravitating extended bodies. The corresponding result at 2.5PN order was also derived by Schäfer using the ADM Hamiltonian approach. Later, the harmonic-coordinates equations of motion were re-computed by Blanchet, Faye and Ponsot following a direct post-Newtonian iteration of the field equations. Some of the latter derivations opt for a formal description of the compact objects by point-particles. There is a nice agreement between all these different methods at the 2.5PN order. In addition, the complete 2.5PN gravitational field generated by point-particles in harmonic coordinates was derived in .
At the 3PN order, the equations of motion have been obtained using a Hamiltonian and formal delta functions by Jaranowski and Schäfer in the center-of-mass frame, and by Damour, Jaranowski and Schäfer in an arbitrary frame. These authors found an irreducible ambiguity linked probably with an incompleteness in the regularization of the infinite self-field of the particles. In this paper, following the method initiated in , we address the problem of the 3PN dynamics of point-particles in harmonic coordinates. Earlier in , our result has been already discussed and reported in the case of circular orbits. We find the presence of one (and only one) undetermined coefficient in the 3PN equations of motion, in agreement with . Recently, the physical equivalence between our result in harmonic coordinates and the result given by the ADM-Hamiltonian approach has been established .
Another line of research, initiated by Chandrasekhar and collaborators , consists of working with continuous hydrodynamical fluids from the start, and derived the metric and equations of motion of an isolated fluid ball up to the 2.5PN order (the derivation in the case of two fluid balls, in the limit of zero size of the bodies, being due to ). Our iteration of the gravitational field and equations of motion in the previous paper is close to the latter line of work in the sense that it is based on the reduction of some general expressions of the post-Newtonian metric, initially valid for continuous fluids, to point-like particles. The choice of point-particles, adopted here as well, is motivated by the efficiency of the delta-functions in performing some complicated non-linear integrations. The price we have to pay is the necessity of a self-field regularization. We apply systematically in this paper the regularization of Hadamard, based on the concept of “partie finie” of singular functions and divergent integrals . This technique is indeed extensively used in this field . More precisely, we apply a variant of the Hadamard regularization, together with a theory of pseudo-functions and distributional derivatives, that is compatible with the Lorentzian structure of the gravitational field. All the details about this regularization can be found in . We use notably a specific form of distributional stress-energy tensor based on “delta-pseudo-functions” (with support limited to the world lines of the particles). In a sense, these delta-pseudo-functions constitute some mathematically well-defined versions of the so-called “good delta functions” introduced long ago by Infeld (see also an appendix in the book of Infeld and Plebanski ).
Thus, we are using a formal regularization method, based on a clear mathematical framework , but that we cannot justify physically (why should the compact objects be described by such delta-pseudo-function singularities?). Definitely, our main justification is that this method permits the derivation of a result in a consistent and well-defined way (i.e. all the difficult non-linear integrals at the 3PN order are computed unambiguously). Furthermore, we shall check that some different regularization prescriptions yield equations of motion that are physically the same, in the sense that they differ from each other by merely a coordinate transformation. Moreover, a justification a posteriori is that the end result owns all the physical properties that we expect the true equations of motion of compact objects to obey. In particular, there is agreement with the known results at the previous post-Newtonian orders, we get the correct geodesic limit for the motion of a test particle in a Schwarzschild background, find that the 3PN equations of motion stay invariant under global Lorentz transformations, and obtain a conserved energy at 3PN (neglecting the radiation reaction). The investigation of the Lagrangian formulation of the equations is dealt with in a separate work .
Ideally, one should perform, instead of a computation valid only for point particles (and necessitating a regularization), a complete calculation in the case of extended bodies, i.e. taking into account the details of the internal structure of the bodies. By considering the limit where the radius of the two objects tend to zero, one should recover the same result as obtained by means of the point-mass regularization. This would demonstrate the suitability of the regularization. In fact, this program has been achieved at the 2PN order by Grishchuk and Kopejkin , who proved that the compactness parameters associated with each object disappear from the equations of motion, and obtained the same equations as in the case of point particles. At the 3PN order there is no such a proof that the method with extended bodies would give the same result as with point particles.
The main problem is that from the 3PN level one cannot compute the most difficult of the non-linear integrals in closed form for two extended fluid bodies of finite radius (though these integrals could perhaps be obtained as power series valid when the two radius tend to zero). Presently the only approach which is able to overcome this problem is the one followed in this paper: namely, to model the source by delta functions and to use a regularization. The price we have to pay is the appearance of one physical undetermined coefficient at the 3PN order. As a consequence, this method should be completed (hopefully in a future work) by the study of the limit relation of the point-particle result with the physical result valid for extended bodies in the limit of zero size. This study should probably give the value of the undetermined parameter left out by the regularization.
The plan of this paper is the following. In Section II, we review some necessary tools concerning the regularization and the definition of the point-particle model. In Section III, we perform the post-Newtonian iteration of the field equations and write the 3PN metric in terms of some convenient non-linear potentials. Section IV is devoted to the computation of the compact-support and quadratically non-linear parts of the potentials. The most difficult potentials, involving notably some non-compact cubic non-linearities at 1PN, are obtained in Section V. The so-called Leibniz and non-distributivity contributions to the equations of motion are derived in Section VI. Finally, we present in Section VII the result for the compact binary’s 3PN acceleration (in the case of general orbits) and the associated 3PN energy.
## II Hadamard regularization
In this section we present a short account about the regularization of Hadamard , the associated generalized or pseudo-functions, and the choice of stress-energy tensor for point-particles. We follow (and refer to) the detailed investigations in . Consider the class $``$ of functions $`F(𝐱)`$ which are smooth ($`C^{\mathrm{}}`$) on $`^3`$ deprived from two singular points $`𝐲_1`$ and $`𝐲_2`$, around which they admit a power-like singular expansion of the type
$$n,F(𝐱)=\underset{a_0an}{}r_1^a\underset{1}{f_a}(𝐧_1)+o(r_1^n),$$
(1)
and similarly for the other point 2. Here $`r_1=|𝐱𝐲_1|0`$, and the coefficients $`{}_{1}{}^{}f_{a}^{}`$ of the various powers of $`r_1`$ depend on the unit direction $`𝐧_1=(𝐱𝐲_1)/r_1`$ of approach to the singular point. The powers $`a`$ of $`r_1`$ are real, range in discrete steps (i.e. $`a(a_i)_i`$) and are bounded from below ($`a_0a`$). The coefficients $`{}_{1}{}^{}f_{a}^{}`$ (and $`{}_{2}{}^{}f_{a}^{}`$) for which $`a<0`$ are referred to as the singular coefficients of $`F`$. If $`F`$ and $`G`$ belong to $``$ so does the ordinary pointwise product $`FG`$, as well as the ordinary gradient $`_iF`$. We define the Hadamard “partie finie” of $`F`$ at the location of the singular point 1 as
$$(F)__1=\frac{d\mathrm{\Omega }_1}{4\pi }\underset{1}{f_0}(𝐧_1),$$
(2)
where $`d\mathrm{\Omega }_1=d\mathrm{\Omega }(𝐧_1)`$ denotes the solid angle element centered on $`𝐲_1`$ and of direction $`𝐧_1`$. Furthermore, the Hadamard partie finie ($`\mathrm{Pf}`$) of the integral $`d^3𝐱F`$, which is in general divergent at the two singular points $`𝐲_1`$ and $`𝐲_2`$ (we assume no divergence at infinity), is defined by
$`\mathrm{Pf}_{s_1,s_2}{\displaystyle d^3𝐱F}`$ $`=`$ $`\underset{s0}{lim}\{{\displaystyle _{𝒟(s)}}d^3𝐱F`$ (4)
$`+4\pi {\displaystyle \underset{a+3<0}{}}{\displaystyle \frac{s^{a+3}}{a+3}}\left({\displaystyle \frac{F}{r_1^a}}\right)__1+4\pi \mathrm{ln}\left({\displaystyle \frac{s}{s_1}}\right)\left(r_1^3F\right)__1+12\}.`$
The first term integrates over a domain $`𝒟(s)`$ defined as $`^3`$ to which the two spherical balls $`r_1s`$ and $`r_2s`$ of radius $`s`$ and centered on the singularities are removed. The other terms, in which the value of a function at 1 takes the meaning (2), are such that they cancel out the divergent part of the first term in the limit where $`s0`$ (the symbol $`12`$ means the same terms but corresponding to the other point 2). Note that the Hadamard partie finie depends on two strictly positive constants $`s_1`$ and $`s_2`$, associated with the logarithms in (4). See and Section V below for alternative expressions of the partie-finie integral.
To any $`F`$ we associate a partie finie pseudo-function $`\mathrm{Pf}F`$ defined as the linear form on $``$ given by the duality bracket,
$$G,<\mathrm{Pf}F,G>=\mathrm{Pf}d^3𝐱FG.$$
(5)
The pseudo-function $`\mathrm{Pf}F`$, when restricted to the set of smooth functions with compact support, is a distribution in the sense of Schwartz . The product of pseudo-functions coincides with the ordinary pointwise product, namely $`\mathrm{Pf}F.\mathrm{Pf}G=\mathrm{Pf}(FG)`$. A particularly interesting pseudo-function, constructed in on the basis of the Riesz delta function , is the delta-pseudo-function $`\mathrm{Pf}\delta _1`$, which plays the same role as the Dirac measure in distribution theory, in the sense that
$$F,<\mathrm{Pf}\delta _1,F>=\mathrm{Pf}d^3𝐱\delta _1F=(F)__1,$$
(6)
where $`(F)__1`$ is the partie finie of $`F`$ as defined by (2). From the product of $`\mathrm{Pf}\delta _1`$ with any $`\mathrm{Pf}F`$ we obtain the new pseudo-function $`\mathrm{Pf}(F\delta _1)`$ which is such that
$$G,<\mathrm{Pf}(F\delta _1),G>=(FG)__1.$$
(7)
Next, the spatial derivative of a pseudo-function of the type $`\mathrm{Pf}F`$, namely $`_i(\mathrm{Pf}F)`$, is treated as follows. Essentially, we require in the so-called rule of integration by parts, namely that we are allowed to freely operate by parts any duality bracket, with the all-integrated (“surface”) terms always zero exactly like in the case of non-singular functions. This requirement is motivated by our will that a computation involving singular functions be as much as possible the same as a computation valid for regular functions. Thus,
$$F,G,<_i(\mathrm{Pf}F),G>=<_i(\mathrm{Pf}G),F>.$$
(8)
Furthermore, we assume that when all the singular coefficients of $`F`$ vanish, the derivative of $`\mathrm{Pf}F`$ reduces to the ordinary derivative, i.e. $`_i(\mathrm{Pf}F)=\mathrm{Pf}(_iF)`$. As a particular case, we see from these assumptions that the integral of a gradient is always zero: $`<_i(\mathrm{Pf}F),1>=0`$. Certainly this should be the case if we want to apply to the case of singular sources a formula which is defined modulo a total divergence for continuous sources. We have also at our disposal a distributional time derivative and the associated partial derivatives with respect to the points 1 and 2 (see Section IX in ). The difference between the distributional derivative and the ordinary one gives the distributional terms $`\mathrm{D}_i[F]`$ present in the derivative of $`F`$,
$$_i(\mathrm{Pf}F)=\mathrm{Pf}(_iF)+\mathrm{D}_i[F].$$
(9)
A simple solution of our basic relation (8), denoted $`\mathrm{D}_i^{\mathrm{part}}[F]`$ standing for the “particular” solution, was obtained in as the following functional of the singular coefficients of $`F`$,
$$\mathrm{D}_i^{\mathrm{part}}[F]=4\pi \mathrm{Pf}(n_1^i[\frac{1}{2}r_1\underset{1}{f_1}+\underset{k0}{}\frac{1}{r_1^k}\underset{1}{f_{2k}}\left]\delta _1\right)+12,$$
(10)
where we assume for simplicity that the powers $`a`$ in the expansion of $`F`$ are relative integers, $`a`$. (The sum over $`k`$ is always finite.) The distributional term (10) is of the form $`\mathrm{Pf}(G\delta _1)`$ (plus $`12`$). However, the particular solution (10) does not represent the most satisfying derivative operator acting on pseudo-functions. It is shown in that one can require also the rule of commutation of successive derivatives, which is not satisfied in general by (10). Still we are motivated when asking for the commutation of derivatives that the properties of our distributional derivative be the closest possible to those of the ordinary derivative. The most general derivative operator satisfying the same properties as (10) and, in addition, the commutation of derivatives (Schwarz lemma) is given by
$$\mathrm{D}_i[F]=4\pi \underset{l=0}{\overset{+\mathrm{}}{}}\mathrm{Pf}\left(C_l\left[n_1^{iL}\underset{1}{\widehat{f}_1^L}n_1^L\underset{1}{\widehat{f}_1^{iL}}\right]r_1\delta _1+\underset{k0}{}\frac{n_1^{iL}}{r_1^k}\underset{1}{\widehat{f}_{2k}^L}\delta _1\right)+12,$$
(11)
where we denote by $`{}_{1}{}^{}\widehat{f}_{a}^{L}`$ the STF-harmonics of the expansion coefficient $`{}_{1}{}^{}f_{a}^{}`$, which is such that $`{}_{1}{}^{}f_{a}^{}=_{l0}n_1^L{}_{1}{}^{}\widehat{f}_{a}^{L}`$ (see for details). A particularity of this derivative is that it depends on an arbitrary constant $`K`$ through the $`l`$-dependent coefficient
$$C_l=(l+1)\left[K+\underset{j=1}{\overset{l}{}}\frac{1}{j+1}\right].$$
(12)
Both the derivative operators (10) and (11)-(12) represent some generalizations of the Schwartz distributional derivative , that are appropriate to the singular functions of the class $``$. In was shown in Section VIII in that the distributional terms associated with the $`l`$th distributional derivative, i.e. $`\mathrm{D}_L[F]=_L\mathrm{Pf}F\mathrm{Pf}_LF`$, where $`L=i_1i_2\mathrm{}i_l`$ denotes a multi-index composed of $`l`$ indices, is given by
$$\mathrm{D}_L[F]=\underset{k=1}{\overset{l}{}}_{i_1\mathrm{}i_{k1}}\mathrm{D}_{i_k}[_{i_{k+1}\mathrm{}i_l}F].$$
(13)
Though this is not manifest on this formula, $`\mathrm{D}_L[F]`$ in the case of the “correct” derivative (11)-(12) is fully symmetric in the $`l`$ indices forming $`L`$. Note that neither of the derivatives (10) and (11) satisfy the Leibniz rule for the derivation of a product. Rather, the investigation in has suggested that, in order to construct a consistent theory (using the “ordinary” product for pseudo-functions), the Leibniz rule should in a sense be weakened, and replaced by the rule of integration by part (8), which is in fact nothing but an “integrated” version of the Leibniz rule. In this paper, we shall be careful about taking into account the violation of the Leibniz rule by the distributional derivative. We shall also investigate the fate of the constant $`K`$ appearing in (12) when deriving the 3PN equations of motion.
The Hadamard regularization $`(F)__1`$ is defined by (2) in a preferred spatial hypersurface $`t=`$const of a coordinate system, and consequently is not a priori compatible with the global Lorentz invariance of special relativity. If we restrict the coordinates to satisfy the usual harmonic gauge conditions, we introduce a preferred Minkowski metric, and thus we can view the gravitational field as a relativistic Lorentz tensor field in special relativity, that we certainly want to regularize in a Lorentz-invariant way. To achieve this we defined in a new regularization, denoted $`[F]__1`$, by performing the Hadamard regularization within the spatial hypersurface which is geometrically orthogonal (in the Minkowskian sense) to the four-velocity of the particle. In a sense, the regularization $`[F]__1`$ permits us to get rid of the anisotropic Lorentz contraction due to their motion when defining the point-masses. The Lorentzian regularization $`[F]__1`$ differs from the old one $`(F)__1`$ by relativistic corrections of order $`1/c^2`$ at least. All the formulas for its computation are given in in the form of some infinite expansion series in the relativistic parameter $`1/c^2`$. The regularization $`[F]__1`$ plays a crucial role in the present computation, as it will be seen that the breakdown of the Lorentz invariance due to the old regularization $`(F)__1`$ occurs precisely at the 3PN order in the equations of motion. Associated with the new regularization in we can define, exactly like in (6), a “Lorentzian” delta-pseudo-function $`\mathrm{Pf}\mathrm{\Delta }_1`$ which when applied on any $`F`$ gives $`[F]__1`$. More generally we have, similarly to (7),
$$G,<\mathrm{Pf}(F\mathrm{\Delta }_1),G>=[FG]__1.$$
(14)
Notice that as a general rule we are not allowed to replace $`F`$ in the pseudo-function $`\mathrm{Pf}(F\mathrm{\Delta }_1)`$ by its regularized value, i.e. $`\mathrm{Pf}(F\mathrm{\Delta }_1)[F]__1\mathrm{Pf}\mathrm{\Delta }_1`$. This is a consequence of the “non-distributivity” of the Hadamard partie finie with respect to the multiplication, i.e. $`[FG]__1[F]__1[G]__1`$. In this paper, we shall (heuristically) model the compact objects by point-particles, and in order to describe those point-particles we shall use a particular representation of the stress-energy tensor which has been derived in Section V of on the basis of an action principle compatible with the Lorentzian regularization $`[F]__1`$. The proposal made in is that
$$T^{\mu \nu }=m_1c\frac{v_1^\mu v_1^\nu }{\sqrt{[g_{\rho \sigma }]__1v_1^\rho v_1^\sigma }}\mathrm{Pf}\left(\frac{\mathrm{\Delta }_1}{\sqrt{g}}\right)+12.$$
(15)
Most importantly about this expression are the facts that (i) $`[g_{\rho \sigma }]__1`$ within the first factor means the Lorentzian regularization of the metric in the previous sense, (ii) the pseudo-function $`\mathrm{Pf}\left(\frac{1}{\sqrt{g}}\mathrm{\Delta }_1\right)`$ is of the type $`\mathrm{Pf}(F\mathrm{\Delta }_1)`$ which is defined by (14). We denote by $`m_1`$ the (constant) mass of the particle 1, by $`𝐲_1(t)`$ its trajectory parametrized by the harmonic-coordinate time $`t`$, and by $`𝐯_1(t)=d𝐲_1/dt`$ the coordinate velocity \[with $`v_1^\mu =(c,𝐯_1)`$\]. In the next section, we look for solutions in the form of post-Newtonian expansions of the Einstein field equations having the latter stress-energy tensor as a matter source.
## III The third post-Newtonian metric
### A The Einstein field equations
We base our investigation on a system of harmonic coordinates $`x^0=ct`$, $`(x^i)=𝐱`$, since such coordinates are especially well-suited to a post-Newtonian (or post-Minkowskian) iteration of the field equations. We define the gravitational perturbation $`h^{\mu \nu }`$ associated with the “gothic metric” as
$$h^{\mu \nu }=\sqrt{g}g^{\mu \nu }\eta ^{\mu \nu },$$
(16)
with $`g^{\mu \nu }`$ and $`g`$ being the inverse and the determinant of the covariant metric $`g_{\mu \nu }`$, and where $`\eta ^{\mu \nu }=\mathrm{diag}(1,1,1,1)`$ denotes an auxiliary Minkowski metric. Under the condition of harmonic coordinates,
$$_\nu h^{\mu \nu }=0,$$
(17)
the Einstein field equations take the form
$$\mathrm{}h^{\mu \nu }=\frac{16\pi G}{c^4}|g|T^{\mu \nu }+\mathrm{\Lambda }^{\mu \nu },$$
(18)
where $`\mathrm{}=\eta ^{\mu \nu }_\mu _\nu `$ denotes the flat d’Alembertian operator, where $`T^{\mu \nu }`$ is the matter stress-energy tensor defined in our case of point-particle binaries by (15), and where $`\mathrm{\Lambda }^{\mu \nu }`$ is the gravitational source term. Using the integral of the retarded potentials given by
$$\mathrm{}_{}^1\tau (𝐱,t)=\frac{d^3𝐱^{}}{4\pi }\frac{\tau (𝐱^{},t|𝐱𝐱^{}|/c)}{|𝐱𝐱^{}|},$$
(19)
we can also re-write the solution of the field equations (18), under a condition of no-incoming radiation, under the form
$$h^{\mu \nu }=\mathrm{}_{}^1\left[\frac{16\pi G}{c^4}|g|T^{\mu \nu }+\mathrm{\Lambda }^{\mu \nu }\right].$$
(20)
The gravitational source term $`\mathrm{\Lambda }^{\mu \nu }`$ is related to the Landau-Lifchitz pseudo-tensor $`t_{\mathrm{LL}}^{\mu \nu }`$ by
$$\mathrm{\Lambda }^{\mu \nu }=\frac{16\pi G}{c^4}|g|t_{\mathrm{LL}}^{\mu \nu }+_\rho h^{\mu \sigma }_\sigma h^{\nu \rho }h^{\rho \sigma }_{\rho \sigma }h^{\mu \nu },$$
(21)
and can be expanded as an infinite non-linear series in $`h`$ and its first and second space-time derivatives; in this paper we need only the non-linear terms up to the quartic ($`h^4`$ or $`G^4`$) level, viz
$$\mathrm{\Lambda }^{\mu \nu }=N^{\mu \nu }(h,h)+M^{\mu \nu }(h,h,h)+L^{\mu \nu }(h,h,h,h)+𝒪(h^5),$$
(22)
where the quadratic non-linearity $`N^{\mu \nu }`$, the cubic one $`M^{\mu \nu }`$ and the quartic $`L^{\mu \nu }`$ are explicitly given by
$`N^{\mu \nu }=`$ $``$ $`h^{\rho \sigma }_{\rho \sigma }h^{\mu \nu }+{\displaystyle \frac{1}{2}}^\mu h_{\rho \sigma }^\nu h^{\rho \sigma }{\displaystyle \frac{1}{4}}^\mu h^\nu h+_\sigma h^{\mu \rho }(^\sigma h_\rho ^\nu +_\rho h^{\nu \sigma })`$ (24)
$``$ $`2^{(\mu }h_{\rho \sigma }^\rho h^{\nu )\sigma }+\eta ^{\mu \nu }\left[{\displaystyle \frac{1}{4}}_\tau h_{\rho \sigma }^\tau h^{\rho \sigma }+{\displaystyle \frac{1}{8}}_\rho h^\rho h+{\displaystyle \frac{1}{2}}_\rho h_{\sigma \tau }^\sigma h^{\rho \tau }\right],`$ (25)
$`M^{\mu \nu }=`$ $``$ $`h^{\rho \sigma }\left(^\mu h_{\rho \tau }^\nu h_\sigma ^\tau +_\tau h_\rho ^\mu ^\tau h_\sigma ^\nu _\rho h_\tau ^\mu _\sigma h^{\nu \tau }\right)`$ (26)
$`+`$ $`h^{\mu \nu }\left[{\displaystyle \frac{1}{4}}_\tau h_{\rho \sigma }^\tau h^{\rho \sigma }+{\displaystyle \frac{1}{8}}_\rho h^\rho h+{\displaystyle \frac{1}{2}}_\rho h_{\sigma \tau }^\sigma h^{\rho \tau }\right]+{\displaystyle \frac{1}{2}}h^{\rho \sigma }^{(\mu }h_{\rho \sigma }^{\nu )}h`$ (27)
$`+`$ $`2h^{\rho \sigma }_\tau h_\rho ^{(\mu }^{\nu )}h_\sigma ^\tau +h^{\rho (\mu }\left[^{\nu )}h_{\sigma \tau }_\rho h^{\sigma \tau }2_\sigma h_\tau ^{\nu )}_\rho h^{\sigma \tau }{\displaystyle \frac{1}{2}}^{\nu )}h_\rho h\right]`$ (28)
$`+`$ $`\eta ^{\mu \nu }[{\displaystyle \frac{1}{8}}h^{\rho \sigma }_\rho h_\sigma h{\displaystyle \frac{1}{4}}h^{\rho \sigma }_\tau h_{\rho \sigma }^\tau h{\displaystyle \frac{1}{4}}h^{\tau \lambda }_\tau h_{\rho \sigma }_\lambda h^{\rho \sigma }`$ (29)
$``$ $`{\displaystyle \frac{1}{2}}h^{\tau \lambda }_\rho h_{\tau \sigma }^\sigma h_\lambda ^\rho +{\displaystyle \frac{1}{2}}h^{\tau \lambda }_\rho h_\tau ^\sigma ^\rho h_{\lambda \sigma }],`$ (30)
$`L^{\mu \nu }=`$ $``$ $`{\displaystyle \frac{1}{2}}h^{\mu \nu }h_{\rho \sigma }_\tau h^{\rho \lambda }_\lambda h^{\sigma \tau }{\displaystyle \frac{1}{4}}h^{\mu \nu }h_{\rho \sigma }^\rho h_{\tau \lambda }^\sigma h^{\tau \lambda }+{\displaystyle \frac{1}{8}}h^{\mu \nu }h_{\rho \sigma }^\rho h^\sigma h`$ (31)
$`+`$ $`{\displaystyle \frac{1}{2}}h^{\mu \nu }h_{\rho \sigma }_\tau h_\lambda ^\rho ^\tau h^{\sigma \lambda }{\displaystyle \frac{1}{4}}h^{\mu \nu }h_{\rho \sigma }_\tau h^{\rho \sigma }^\tau h+h_{\rho \lambda }h_\sigma ^\lambda _\tau h^{\mu \rho }^\tau h^{\nu \sigma }`$ (32)
$``$ $`2h_{\rho \lambda }h_\sigma ^\lambda _\tau h^{\rho (\mu }^{\nu )}h^{\sigma \tau }+h_{\rho \lambda }h_\sigma ^\lambda ^\mu h_\tau ^\rho ^\nu h^{\sigma \tau }{\displaystyle \frac{1}{2}}h_{\rho \lambda }h_\sigma ^\lambda ^{(\mu }h^{\rho \sigma }^{\nu )}h`$ (33)
$``$ $`h_{\rho \sigma }h_{\tau \lambda }^\tau h^{\mu \rho }^\lambda h^{\nu \sigma }+{\displaystyle \frac{1}{2}}h_{\rho \sigma }h_{\tau \lambda }^\mu h^{\rho \tau }^\nu h^{\sigma \lambda }{\displaystyle \frac{1}{4}}h_{\rho \sigma }h_{\tau \lambda }^\mu h^{\rho \sigma }^\nu h^{\tau \lambda }`$ (34)
$`+`$ $`2h_{\sigma \tau }h^{\rho (\mu }_\lambda h^{\nu )\tau }_\rho h^{\sigma \lambda }2h_{\sigma \tau }h^{\rho (\mu }^{\nu )}h_\lambda ^\sigma _\rho h^{\tau \lambda }+{\displaystyle \frac{1}{2}}h_{\sigma \tau }h^{\rho (\mu }^{\nu )}h^{\sigma \tau }_\rho h`$ (35)
$`+`$ $`{\displaystyle \frac{1}{2}}h_{\sigma \tau }h^{\rho (\mu }^{\nu )}h_\rho h^{\sigma \tau }+{\displaystyle \frac{1}{2}}h^{\mu \rho }h^{\nu \sigma }_\rho h_{\tau \lambda }_\sigma h^{\tau \lambda }{\displaystyle \frac{1}{4}}h^{\mu \rho }h^{\nu \sigma }_\rho h_\sigma h`$ (36)
$`+`$ $`\eta ^{\mu \nu }[{\displaystyle \frac{1}{2}}h_{\rho \pi }h_\sigma ^\pi _\tau h^{\rho \lambda }_\lambda h^{\sigma \tau }{\displaystyle \frac{1}{2}}h_{\rho \pi }h_\sigma ^\pi _\lambda h_\tau ^\rho ^\lambda h^{\sigma \tau }+{\displaystyle \frac{1}{4}}h_{\rho \pi }h_\sigma ^\pi _\tau h^{\rho \sigma }^\tau h`$ (37)
$``$ $`{\displaystyle \frac{1}{4}}h_{\rho \sigma }h_{\tau \lambda }_\pi h^{\rho \tau }^\pi h^{\sigma \lambda }+{\displaystyle \frac{1}{8}}h_{\rho \sigma }h_{\tau \lambda }_\pi h^{\rho \sigma }^\pi h^{\tau \lambda }+{\displaystyle \frac{1}{2}}h_{\rho \sigma }h_{\tau \lambda }^\rho h_\pi ^\tau ^\sigma h^{\lambda \pi }`$ (38)
$``$ $`{\displaystyle \frac{1}{4}}h_{\rho \sigma }h_{\tau \lambda }^\rho h^{\tau \lambda }^\sigma h].`$ (39)
All indices are lowered and raised with the Minkowski metric $`\eta _{\mu \nu }`$; $`h=\eta ^{\mu \nu }h_{\mu \nu }`$; the parenthesis around indices indicate the symmetrization.
To describe the matter source we find convenient to introduce the density of mass $`\sigma `$, of current $`\sigma _i`$ and of stress $`\sigma _{ij}`$ defined by
$`\sigma c^2=T^{00}+T^{ii},`$ (41)
$`\sigma _ic=T^{0i},`$ (42)
$`\sigma _{ij}=T^{ij},`$ (43)
(where $`T^{ii}=\delta _{ij}T^{ij}`$). These definitions are such that $`\sigma `$, $`\sigma _i`$ and $`\sigma _{ij}`$ admit a finite non-zero limit when $`c+\mathrm{}`$ (since $`T^{\mu \nu }`$ has the dimension of an energy density). In the case of our model of point-particles \[stress-energy tensor given by (15)\], we obtain
$`\sigma (𝐱,t)`$ $`=`$ $`\mathrm{Pf}(\stackrel{~}{\mu }_1\mathrm{\Delta }_1)+12,`$ (45)
$`\sigma _i(𝐱,t)`$ $`=`$ $`\mathrm{Pf}(\mu _1v_1^i\mathrm{\Delta }_1)+12,`$ (46)
$`\sigma _{ij}(𝐱,t)`$ $`=`$ $`\mathrm{Pf}(\mu _1v_1^iv_1^j\mathrm{\Delta }_1)+12,`$ (47)
where $`\mathrm{\Delta }_1\mathrm{\Delta }[𝐱𝐲_1(t)]`$, and where $`\mu _1`$ and $`\stackrel{~}{\mu }_1`$ represent some effective masses defined by
$`\mu _1(𝐱,t)`$ $`=`$ $`{\displaystyle \frac{m_1c}{\sqrt{[g_{\rho \sigma }]__1v_1^\rho v_1^\sigma }}}.{\displaystyle \frac{1}{\sqrt{g(𝐱,t)}}},`$ (49)
$`\stackrel{~}{\mu }_1(𝐱,t)`$ $`=`$ $`\mu _1(𝐱,t)\left[1+{\displaystyle \frac{𝐯_1^2}{c^2}}\right].`$ (50)
Note that $`\mu _1`$ and $`\stackrel{~}{\mu }_1`$ depend on time and space. Indeed, while the first factor in (49) is clearly a mere function of time through the values of the positions and velocities of the particles at the instant $`t`$, the second factor $`(g)^{1/2}`$ is evaluated at the field point $`t,𝐱`$ instead of the source point $`t,𝐲_1`$. From the non-distributivity of the Hadamard regularization, one is not allowed to replace $`(g)^{1/2}`$ by its (regularized) value at the point 1, even though it is multiplied by a delta-pseudo-function at 1.
### B The 3PN iteration of the metric
In what follows we sketch the main steps of our iteration of the Einstein field equations (18)-(III A) generated by two particles at the 3PN order. For more clarity in the presentation, we reason by induction over the post-Newtonian order $`n`$. However, we do not have proved the validity of this method to any order $`n`$; simply we applied the method outlined below to construct the metric at the 3PN order.
(I) Suppose by induction over $`n`$ that we have succeeded in obtaining some approximate post-Newtonian metric coefficients $`h_{[2n2]}^{\mu \nu }`$, as well as the previous coefficients $`h_{[m]}^{\mu \nu }`$ for any $`m`$ such that $`2m2n2`$, which approach the true metric modulo a small post-Newtonian remainder,
$$h^{\mu \nu }=h_{[2n2]}^{\mu \nu }+𝒪(2n1),$$
(51)
with the notation $`𝒪(2n1)=𝒪(1/c^{2n1})`$, here and elsewhere, for the post-Newtonian error terms. We assume that the $`h_{[m]}^{\mu \nu }`$’s are at once some explicit functions of the field point $`𝐱`$ and functionals of the two trajectories $`𝐲_1(t)`$, $`𝐲_2(t)`$ and velocities $`𝐯_1(t)`$, $`𝐯_2(t)`$. Since the matter source of the field equations is made of delta-pseudo-functions, the metric coefficients become singular at the location of the particles (indeed, this is already true at the Newtonian order). As a matter of fact, we assume for the present iteration that,
$$m2n2,h_{[m]}^{\mu \nu },$$
(52)
where $``$ is the class of functions considered in and Section II. This is not a completely rigorous assumption because of the presence of logarithms in the expansions around the singularities; but we shall see that this assumption is justified at the 3PN order where one can consider these logarithms as mere constants.
(II) Consider for simplicity the combination $`\frac{1}{2}(h^{00}+h^{ii})`$ only, for which we need the maximal post-Newtonian precision since it is directly connected to $`g_{00}`$. The structure of the Einstein field equations (18), containing notably the gravitational source term (22)-(III A), reads as
$$\mathrm{}\left(\frac{h^{00}+h^{ii}}{2}\right)=\frac{8\pi G}{c^2}|g|\sigma +h\mathrm{}hhh,$$
(53)
where $`\sigma `$ is given by (41), where the sum runs over non-linearities and the two partial derivatives $``$ have to be distributed among the $`h`$’s (with double derivatives allowed in the quadratic term). In order to obtain an equation valid at the next post-Newtonian order $`n`$, we replace the approximate metric (51) into the right-hand side of (53). Furthermore, we replace the partial derivatives $``$ in (53) by the distributional derivatives (9) \[we shall discuss the effect of using either the particular derivative (10) of the more correct one (11)\]. Using also the density of particles in the form (45), we get
$`\mathrm{}\left({\displaystyle \frac{h^{00}+h^{ii}}{2}}\right)`$ $`=`$ $`\{{\displaystyle \frac{8\pi G}{c^2}}(\mathrm{Pf}|g|\stackrel{~}{\mu }_1\mathrm{\Delta }_1+\mathrm{Pf}|g|\stackrel{~}{\mu }_2\mathrm{\Delta }_2)`$ (54)
$`+`$ $`{\displaystyle \underset{m_1,\mathrm{},m_p2n2}{}}h_{[m_1]}\mathrm{}h_{[m_{p2}]}\left(\mathrm{Pf}h_{[m_{p1}]}\right)\left(\mathrm{Pf}h_{[m_p]}\right)\}_{[2n]}`$ (55)
$`+`$ $`𝒪(2n+1),`$ (56)
where the $`h_{[m_1]}`$, $`\mathrm{}`$, $`h_{[m_p]}`$ (with $`2pn`$) denote the metric coefficients known from the previous iterations, and where as indicated by the label $`[2n]`$ a truncation up to the post-Newtonian order $`1/c^{2n}`$ is understood. At this stage, any subsequent transformation of the right-hand side must be done using the rules for handling the pseudo-functions and their derivatives .
(III) We integrate the latter equation by means of the retarded integral given by (19):
$`{\displaystyle \frac{h^{00}+h^{ii}}{2}}`$ $`=`$ $`\mathrm{}_{}^1\{{\displaystyle \frac{8\pi G}{c^2}}(\mathrm{Pf}|g|\stackrel{~}{\mu }_1\mathrm{\Delta }_1+\mathrm{Pf}|g|\stackrel{~}{\mu }_2\mathrm{\Delta }_2)`$ (57)
$`+`$ $`{\displaystyle \underset{m_1,\mathrm{},m_p2n2}{}}h_{[m_1]}\mathrm{}h_{[m_{p2}]}\left(\mathrm{Pf}h_{[m_{p1}]}\right)\left(\mathrm{Pf}h_{[m_p]}\right)\}_{[2n]}`$ (58)
$`+`$ $`𝒪(2n+1).`$ (59)
This defines the solution to the $`n`$th order, and so, by recursion, to any order (in principle). The partie-finie symbols $`\mathrm{Pf}`$ take care of the divergences of the retarded integral at the locations of the particles; that is, the retarded integral is considered as a partie-finie integral in the sense of . More precisely, the retardations in (57) are expanded to the $`n`$PN order and the resulting Poisson-like integrals computed using the duality brackets in the way specified by Section V in . Actually, the Poisson-like integrals, which have a non-compact support, become rapidly divergent at infinity when $`n`$ increases, and the correct solution we use is not the Poisson-like integral but is obtained by a matching of the inner metric to the multipole expansion of the exterior field. So, in fact,
$`\left(\mathrm{}_{}^1\mathrm{Pf}F\right)(𝐱^{},t)`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{k=0}{\overset{2n}{}}}{\displaystyle \frac{()^k}{k!c^k}}<\left({\displaystyle \frac{}{t}}\right)^k\left[\mathrm{Pf}F(𝐱,t)\right],|𝐱𝐱^{}|^{k1}>_{\mathrm{match}}`$ (60)
$`+`$ $`𝒪(2n+1),`$ (61)
where the subscript “match” refers to the matching process that is described in Section IV in the case of the 3PN order. Notice that the time-derivatives $`(/t)^k`$ resulting from the Taylor expansion of the retardations are distributional derivatives and therefore can be put outside the duality bracket (see Section IX in ). Thus, equivalently,
$`\left(\mathrm{}_{}^1\mathrm{Pf}F\right)(𝐱^{},t)`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{k=0}{\overset{2n}{}}}{\displaystyle \frac{()^k}{k!c^k}}\left({\displaystyle \frac{}{t}}\right)^k[<\mathrm{Pf}F(𝐱,t),|𝐱𝐱^{}|^{k1}>_{\mathrm{match}}]`$ (62)
$`+`$ $`𝒪(2n+1).`$ (63)
(IV) Once the solution (57) to the $`n`$th post-Newtonian order is in hands we perform many simplifications of the expression, following the rules of application of the distributional derivative. In particular, we find very useful to use the fact that a double gradient can be re-expressed in terms of d’Alembertians as
$$_\mu F^\mu G=\frac{1}{2}\left[\mathrm{}\left(FG\right)F\mathrm{}GG\mathrm{}F\right],$$
(64)
which implies that the retarded integral reads as
$$\mathrm{}_{}^1[_\mu F^\mu G]=\frac{1}{2}FG\frac{1}{2}\mathrm{}_{}^1\left[F\mathrm{}G+G\mathrm{}F\right].$$
(65)
The first term is “all-integrated”, while the second term, in which one can replace the d’Alembertians by their corresponding sources, brings in general many interesting cancellations with other terms. Unfortunately, the formula (64) is valid only in an ordinary sense but not in the distributional sense, because the distributional derivative does not satisfy, in general, the Leibniz rule. Thus, in general,
$$_\mu (\mathrm{Pf}F)^\mu (\mathrm{Pf}G)\frac{1}{2}\left[\mathrm{}\left(\mathrm{Pf}FG\right)F\mathrm{}(\mathrm{Pf}G)G\mathrm{}(\mathrm{Pf}F)\right].$$
(66)
Nevertheless, the strategy we have chosen to follow in this paper is to take advantage of the many simplifications brought about by the latter process, at the price of introducing some extra terms (named “Leibniz”) accounting for the violation of the Leibniz rule. This means that we shall write, similarly to (65),
$$\mathrm{}_{}^1[_\mu (\mathrm{Pf}F)^\mu (\mathrm{Pf}G)]=\frac{1}{2}\mathrm{Pf}FG\frac{1}{2}\mathrm{}_{}^1\left[F\mathrm{}(\mathrm{Pf}G)+G\mathrm{}(\mathrm{Pf}F)\right]+\delta _{\mathrm{Leibniz}}T,$$
(67)
where the Leibniz term is given by
$$\delta _{\mathrm{Leibniz}}T=\mathrm{}_{}^1\left[_\mu (\mathrm{Pf}F)^\mu (\mathrm{Pf}G)\frac{1}{2}\mathrm{}(\mathrm{Pf}FG)+\frac{1}{2}F\mathrm{}(\mathrm{Pf}G)+\frac{1}{2}G\mathrm{}(\mathrm{Pf}F)\right].$$
(68)
Obviously the Leibniz term depends only on the purely distributional part of the derivative. See the Appendix A for the complete list of the Leibniz terms. As it will turn out these terms are not too difficult to compute, and, of course, arise precisely at the 3PN order. They give a contribution to the metric and the equations of motion that we shall be able to check from the requirement of Lorentz invariance (see Section VI).
### C The 3PN non-linear potentials
The post-Newtonian iteration sketched in the previous subsection is implemented to the 3PN order. The computation is long but straightforward. After the simplification process described above we find that the metric is parametrized by certain non-linear potentials, which do not carry a physical signification by themselves, but turn out to be useful in the present computation. The 3PN metric reads as
$`g_{00}`$ $`=`$ $`1+{\displaystyle \frac{2}{c^2}}V{\displaystyle \frac{2}{c^4}}V^2+{\displaystyle \frac{8}{c^6}}\left(\widehat{X}+V_iV_i+{\displaystyle \frac{V^3}{6}}\right)`$ (71)
$`+{\displaystyle \frac{32}{c^8}}\left(\widehat{T}{\displaystyle \frac{1}{2}}V\widehat{X}+\widehat{R}_iV_i{\displaystyle \frac{1}{2}}VV_iV_i{\displaystyle \frac{1}{48}}V^4\right)+𝒪(10),`$
$`g_{0i}`$ $`=`$ $`{\displaystyle \frac{4}{c^3}}V_i{\displaystyle \frac{8}{c^5}}\widehat{R}_i{\displaystyle \frac{16}{c^7}}\left(\widehat{Y}_i+{\displaystyle \frac{1}{2}}\widehat{W}_{ij}V_j+{\displaystyle \frac{1}{2}}V^2V_i\right)+𝒪(9),`$ (72)
$`g_{ij}`$ $`=`$ $`\delta _{ij}\left[1+{\displaystyle \frac{2}{c^2}}V+{\displaystyle \frac{2}{c^4}}V^2+{\displaystyle \frac{8}{c^6}}\left(\widehat{X}+V_kV_k+{\displaystyle \frac{V^3}{6}}\right)\right]+{\displaystyle \frac{4}{c^4}}\widehat{W}_{ij}`$ (74)
$`+{\displaystyle \frac{16}{c^6}}\left(\widehat{Z}_{ij}+{\displaystyle \frac{1}{2}}V\widehat{W}_{ij}V_iV_j\right)+𝒪(8).`$
We recall our notation for the small post-Newtonian remainders: $`𝒪(n)=𝒪(1/c^n)`$. The various post-Newtonian orders are parametrized by some potentials which are defined by means of the retarded integral (19). At the “Newtonian” and 1PN orders we pose
$`V`$ $`=`$ $`\mathrm{}_{}^1[4\pi G\sigma ],`$ (76)
$`V_i`$ $`=`$ $`\mathrm{}_{}^1[4\pi G\sigma _i],`$ (77)
in which the source densities were defined by (III A). Next, at the 2PN order, we define
$`\widehat{X}`$ $`=`$ $`\mathrm{}_{}^1[4\pi GV\sigma _{ii}+\widehat{W}_{ij}_{ij}V+2V_i_t_iV+V_t^2V+{\displaystyle \frac{3}{2}}(_tV)^22_iV_j_jV_i],`$ (79)
$`\widehat{R}_i`$ $`=`$ $`\mathrm{}_{}^1[4\pi G(V\sigma _iV_i\sigma )2_kV_iV_k{\displaystyle \frac{3}{2}}_tV_iV],`$ (80)
$`\widehat{W}_{ij}`$ $`=`$ $`\mathrm{}_{}^1[4\pi G(\sigma _{ij}\delta _{ij}\sigma _{kk})_iV_jV].`$ (81)
Finally, at the 3PN order, we have
$`\widehat{T}`$ $`=`$ $`\mathrm{}_{}^1[4\pi G({\displaystyle \frac{1}{4}}\sigma _{ij}\widehat{W}_{ij}+{\displaystyle \frac{1}{2}}V^2\sigma _{ii}+\sigma V_iV_i)+\widehat{Z}_{ij}_{ij}V+\widehat{R}_i_t_iV2_iV_j_j\widehat{R}_i`$ (85)
$`_iV_j_t\widehat{W}_{ij}+VV_i_t_iV+2V_i_jV_i_jV+{\displaystyle \frac{3}{2}}V_i_tV_iV+{\displaystyle \frac{1}{2}}V^2_t^2V`$
$`+{\displaystyle \frac{3}{2}}V(_tV)^2{\displaystyle \frac{1}{2}}(_tV_i)^2]+\delta _{\mathrm{Leibniz}}\widehat{T},`$
$`\widehat{Y}_i`$ $`=`$ $`\mathrm{}_{}^1[4\pi G(\sigma \widehat{R}_i\sigma VV_i+{\displaystyle \frac{1}{2}}\sigma _k\widehat{W}_{ik}+{\displaystyle \frac{1}{2}}\sigma _{ik}V_k+{\displaystyle \frac{1}{2}}\sigma _{kk}V_i)+\widehat{W}_{kl}_{kl}V_i_t\widehat{W}_{ik}_kV`$ (88)
$`+_i\widehat{W}_{kl}_kV_l_k\widehat{W}_{il}_lV_k2_kV_i\widehat{R}_k{\displaystyle \frac{3}{2}}V_k_iV_kV{\displaystyle \frac{3}{2}}V_tV_iV`$
$`2V_kV_kV_i+V_t^2V_i+2V_k_k_tV_i]+\delta _{\mathrm{Leibniz}}\widehat{Y}_i,`$
$`\widehat{Z}_{ij}`$ $`=`$ $`\mathrm{}_{}^1[4\pi GV(\sigma _{ij}\delta _{ij}\sigma _{kk})2_{(i}V_tV_{j)}+_iV_k_jV_k+_kV_i_kV_j2_{(i}V_k_kV_{j)}`$ (90)
$`\delta _{ij}_kV_m(_kV_m_mV_k){\displaystyle \frac{3}{4}}\delta _{ij}(_tV)^2]+\delta _{\mathrm{Leibniz}}\widehat{Z}_{ij}.`$
Note the presence in the 3PN potentials of the Leibniz contributions described in the previous subsection, which are due to the simplifications we did to arrive at these relatively simple expressions (with respect to what could be expected at the high 3PN order). The Leibniz contributions will be computed in Section VI. Of course, in the case where the matter source is continuous — an hydrodynamical fluid for instance —, the 3PN metric (III C) and all the expressions of non-linear potentials are valid with simply the Leibniz contributions set to zero.
The potentials (III C)-(III C) are connected by the following approximate post-Newtonian differential identities (equivalent to the condition of harmonic coordinates at the 3PN order):
$`_t\left\{V+{\displaystyle \frac{1}{c^2}}\left[{\displaystyle \frac{1}{2}}\widehat{W}_{kk}+2V^2\right]+{\displaystyle \frac{4}{c^4}}\left[\widehat{X}+{\displaystyle \frac{1}{2}}\widehat{Z}_{kk}+{\displaystyle \frac{1}{2}}V\widehat{W}_{kk}+{\displaystyle \frac{2}{3}}V^3\right]\right\}`$ (92)
$`+_i\left\{V_i+{\displaystyle \frac{2}{c^2}}\left[\widehat{R}_i+VV_i\right]+{\displaystyle \frac{4}{c^4}}\left[\widehat{Y}_i{\displaystyle \frac{1}{2}}\widehat{W}_{ij}V_j+{\displaystyle \frac{1}{2}}\widehat{W}_{kk}V_i+V\widehat{R}_i+V^2V_i\right]\right\}=𝒪(6),`$ (93)
$`_t\left\{V_i+{\displaystyle \frac{2}{c^2}}\left[\widehat{R}_i+VV_i\right]\right\}+_j\left\{\widehat{W}_{ij}{\displaystyle \frac{1}{2}}\widehat{W}_{kk}\delta _{ij}+{\displaystyle \frac{4}{c^2}}\left[\widehat{Z}_{ij}{\displaystyle \frac{1}{2}}\widehat{Z}_{kk}\delta _{ij}\right]\right\}=𝒪(4).`$ (94)
We shall check that the (regularized) potentials we compute satisfy these identities. They are in turn respectively equivalent to the equation of continuity at the 2PN order and the equation of motion at the 1PN order:
$`_t\left[\sigma \left(1+{\displaystyle \frac{2\widehat{W}_{ii}}{c^4}}\right)\right]`$ $`+`$ $`_j\left[\sigma _j\left(1+{\displaystyle \frac{2\widehat{W}_{ii}}{c^4}}\right)\right]`$ (96)
$`=`$ $`{\displaystyle \frac{1}{c^2}}\left(_t\sigma _{jj}\sigma _tV\right){\displaystyle \frac{4}{c^4}}\left(\sigma V_j_jV+\sigma _{jk}_jV_k\right)+𝒪(6),`$ (97)
$`_t\left[\sigma _i\left(1+{\displaystyle \frac{4V}{c^2}}\right)\right]`$ $`+`$ $`_j\left[\sigma _{ij}\left(1+{\displaystyle \frac{4V}{c^2}}\right)\right]`$ (98)
$`=`$ $`\sigma _iV+{\displaystyle \frac{4}{c^2}}\left[\sigma _tV_i+\sigma _j\left(_jV_i_iV_j\right)\right]+𝒪(4).`$ (99)
### D Computing the equations of motion
The equations of motion of the particle 1 are deduced from the covariant conservation of the stress-energy tensor of the particles,
$$_\nu T^{\mu \nu }=0,$$
(100)
where $`T^{\mu \nu }`$ is given by the definite expression (15) made of the delta-pseudo-functions defined in . It is shown in Section V of that by integrating (100) over a volume surrounding the particle 1 (and only 1), i.e. by constructing the duality bracket of (100) with the characteristic function of that volume, we obtain the equations of motion of the particle 1 in the form
$$\frac{d}{dt}\left(\frac{[g_{\lambda \mu }]__1v_1^\mu }{\sqrt{[g_{\rho \sigma }]__1\frac{v_1^\rho v_1^\sigma }{c^2}}}\right)=\frac{1}{2}\frac{[_\lambda g_{\mu \nu }]__1v_1^\mu v_1^\nu }{\sqrt{[g_{\rho \sigma }]__1\frac{v_1^\rho v_1^\sigma }{c^2}}}.$$
(101)
These equations of motion take the same form as the geodesic equations for a test particle moving on a smooth background, but with the role of the background metric played by the true metric generated by the two bodies and regularized according to the Lorentzian prescription .
In this paper we compute the spatial acceleration of body 1, which corresponds to the equation with spatial index $`\lambda =i`$ in (101); we do not consider the energy which would be given by the equation with time index $`\lambda =0`$. Indeed, the energy of the binary system will be determined directly from the (fully order-reduced) acceleration. From (101) we can write the equations into the form
$$\frac{dP_1^i}{dt}=F_1^i,$$
(102)
where the “linear momentum density” $`P_1^i`$ and “force density” $`F_1^i`$ are given by
$`P_1^i`$ $`=`$ $`{\displaystyle \frac{[g_{i\mu }]__1v_1^\mu }{\sqrt{[g_{\rho \sigma }]__1\frac{v_1^\rho v_1^\sigma }{c^2}}}},`$ (104)
$`F_1^i`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{[_ig_{\mu \nu }]__1v_1^\mu v_1^\nu }{\sqrt{[g_{\rho \sigma }]__1\frac{v_1^\rho v_1^\sigma }{c^2}}}}.`$ (105)
The expressions of both $`P_1^i`$ and $`F_1^i`$ in terms of the non-linear potentials follow from insertion of the 3PN metric coefficients (III C). We obtain some complicated sums of products of potentials which are regularized at the point 1 following the prescription $`[F]__1`$. Since the computation will turn out to be quite involved, we decide to adopt the following “step-by-step” strategy:
(A) We compute, in Sections IV and V, all the needed individual potentials and their gradients at the point 1 following the non-Lorentzian regularization $`(F)__1`$; for instance we obtain $`(_iV)__1`$ at the 3PN order, $`(V)__1`$ at the 2PN order, $`(_i\widehat{X})__1`$ at the 1PN order, $`(_i\widehat{T})__1`$ at the Newtonian order, and so on. (Because of the length of the formulas, and since the results for each of these individual regularized potentials are only intermediate, we shall not give them in this paper; see the appendices of for complete expressions.)
(B) We add up the corrections brought about by the Lorentzian regularization $`[F]__1`$ with respect to $`(F)__1`$. We find, at the end of Section V, that the only effect of the new regularization at the 3PN order, when computing the values of potentials at 1 (but the new regularization affects also the corrections due to the non-distributivity), is a crucial 1PN correction arising from the so-called “cubic non-compact” part of $`\widehat{X}`$; that is, we find $`[_i\widehat{X}^{(\mathrm{CNC})}]__1(_i\widehat{X}^{(\mathrm{CNC})})__10`$.
(C) We replace all the individually regularized potentials $`[F]__1`$ and their gradients into the equations of motion (102)-(III D) which would be obtained while supposing that the Hadamard regularization is “distributive” with respect to the multiplication, i.e. supposing incorrectly that we are allowed to write everywhere $`[FG]__1=[F]__1[G]__1`$. In doing this we obtain what we call the “distributive” parts of the linear momentum and force densities (III D), namely $`(P_1^i)_{_{\mathrm{distr}}}`$ and $`(F_1^i)_{_{\mathrm{distr}}}`$. (Other types of non-distributivity arising in the potentials themselves are discussed in Section IV.)
(D) Finally, we compute separately, in Section VI, the corrections due to the non-distributivity, i.e. the differences $`P_1^i(P_1^i)_{_{\mathrm{distr}}}`$ and $`F_1^i(F_1^i)_{_{\mathrm{distr}}}`$. Note that these corrections reflect quantitatively the specific form that we have adopted for the stress-energy tensor of point-particles (15). Had we used another stress-energy tensor, for instance by replacing incorrectly $`\mathrm{Pf}(\frac{1}{\sqrt{g}}\mathrm{\Delta }_1)`$ by $`[\frac{1}{\sqrt{g}}]__1\mathrm{Pf}\mathrm{\Delta }_1`$ inside (15), we would have obtained a different non-distributivity, and thereby some different equations of motion. Note also that thanks to the new regularization $`[F]__1`$ the corrections due to the non-distributivity do not alter the Lorentz invariance of the equations of motion. At last, we find the 3PN acceleration of body 1 as
$$a_1^i=F_1^i\frac{d}{dt}\left(P_1^iv_1^i\right).$$
(106)
We report now the expressions of the distributive parts of the linear momentum and force densities as straightforwardly obtained by substitution of the 3PN metric (III C). The expressions of the correcting terms due to the non-distributivity (i.e. $`[FG]__1[F]__1[G]__1`$) are relegated to Section VI, where it is seen that they contribute only at the 3PN order.
$`(P_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`v_1^i+{\displaystyle \frac{1}{c^2}}\left({\displaystyle \frac{1}{2}}v_1^2v_1^i+3[V]__1v_1^i4[V_i]__1\right)`$ (108)
$`+`$ $`{\displaystyle \frac{1}{c^4}}({\displaystyle \frac{3}{8}}v_1^4v_1^i+{\displaystyle \frac{7}{2}}[V]__1v_1^2v_1^i4[V_j]__1v_1^iv_1^j2[V_i]__1v_1^2`$ (110)
$`+{\displaystyle \frac{9}{2}}[V]__1^2v_1^i4[V]__1[V_i]__1+4[\widehat{W}_{ij}]__1v_1^j8[\widehat{R}_i]__1)`$
$`+`$ $`{\displaystyle \frac{1}{c^6}}({\displaystyle \frac{5}{16}}v_1^6v_1^i+{\displaystyle \frac{33}{8}}[V]__1v_1^4v_1^i{\displaystyle \frac{3}{2}}[V_i]__1v_1^46[V_j]__1v_1^iv_1^jv_1^2+{\displaystyle \frac{49}{4}}[V]__1^2v_1^2v_1^i`$ (115)
$`+2[\widehat{W}_{ij}]__1v_1^jv_1^2+2[\widehat{W}_{jk}]__1v_1^iv_1^jv_1^k10[V]__1[V_i]__1v_1^220[V]__1[V_j]__1v_1^iv_1^j`$
$`4[\widehat{R}_i]__1v_1^28[\widehat{R}_j]__1v_1^iv_1^j+{\displaystyle \frac{9}{2}}[V]__1^3v_1^i+12[V_j]__1[V_j]__1v_1^i`$
$`+12[\widehat{W}_{ij}]__1[V]__1v_1^j+12[\widehat{X}]__1v_1^i+16[\widehat{Z}_{ij}]__1v_1^j10[V]__1^2[V_i]__1`$
$`8[\widehat{W}_{ij}]__1[V_j]__18[V]__1[\widehat{R}_i]__116[\widehat{Y}_i]__1)+𝒪(8),`$
$`(F_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`[_iV]__1+{\displaystyle \frac{1}{c^2}}\left([V]__1[_iV]__1+{\displaystyle \frac{3}{2}}[_iV]__1v_1^24[_iV_j]__1v_1^j\right)`$ (116)
$`+`$ $`{\displaystyle \frac{1}{c^4}}({\displaystyle \frac{7}{8}}[_iV]__1v_1^42[_iV_j]__1v_1^jv_1^2+{\displaystyle \frac{9}{2}}[V]__1[_iV]__1v_1^2`$ (119)
$`+2[_i\widehat{W}_{jk}]__1v_1^jv_1^k4[V_j]__1[_iV]__1v_1^j4[V]__1[V_j]__1v_1^j`$
$`8[_i\widehat{R}_j]__1v_1^j+{\displaystyle \frac{1}{2}}[V]__1^2[_iV]__1+8[V_j]__1[_iV_j]__1+4[_i\widehat{X}]__1)`$
$`+`$ $`{\displaystyle \frac{1}{c^6}}({\displaystyle \frac{11}{16}}v_1^6[_iV]__1{\displaystyle \frac{3}{2}}[_iV_j]__1v_1^jv_1^4+{\displaystyle \frac{49}{8}}[V]__1[_iV]__1v_1^4+[_i\widehat{W}_{jk}]__1v_1^2v_1^jv_1^k`$ (127)
$`10[V_j]__1[_iV]__1v_1^2v_1^j10[V]__1[_iV_j]__1v_1^2v_1^j4[_i\widehat{R}_j]__1v_1^2v_1^j`$
$`+{\displaystyle \frac{27}{4}}[V]__1^2[_iV]__1v_1^2+12[V_j]__1[_iV_j]__1v_1^2+6[\widehat{W}_{jk}]__1[_iV]__1v_1^jv_1^k`$
$`+6[V]__1[_i\widehat{W}_{jk}]__1v_1^jv_1^k+6[_i\widehat{X}]__1v_1^2+8[_i\widehat{Z}_{jk}]__1v_1^jv_1^k20[V_j]__1[V]__1[_iV]__1v_1^j`$
$`10[V]__1^2[_iV_j]__1v_1^j8[V_k]__1[_i\widehat{W}_{jk}]__1v_1^j8[\widehat{W}_{jk}]__1[_iV_k]__1v_1^j`$
$`8[\widehat{R}_j]__1[_iV]__1v_1^j8[V]__1[_i\widehat{R}_j]__1v_1^j16[_i\widehat{Y}_j]__1v_1^j{\displaystyle \frac{1}{6}}[V]__1^3[_iV]__1`$
$`4[V_j]__1[V_j]__1[_iV]__1+16[\widehat{R}_j]__1[_iV_j]__1+16[V_j]__1[_i\widehat{R}_j]__18[V]__1[V_j]__1[_iV_j]__1`$
$`4[\widehat{X}]__1[_iV]__14[V]__1[_i\widehat{X}]__1+16[_i\widehat{T}]__1)+𝒪(8).`$
Recall that it is supposed that all the accelerations appearing in the potentials have been order-reduced by means of the equations of motion. Notably, during the reduction of the “Newtonian” term $`[_iV]__1`$ in Section IV, we shall need the equations of motion to the 2PN order. Furthermore, we see from (115) that when computing the time-derivative of $`P_i`$ we meet an acceleration at 1PN which is thus also to be replaced by the 2PN equations of motion. We recall here that the latter 2PN (or, rather, 2.5PN) equations in harmonic coordinates are
$`{\displaystyle \frac{dv_1^i}{dt}}=`$ $``$ $`{\displaystyle \frac{Gm_2}{r_{12}^2}}n_{12}^i+{\displaystyle \frac{Gm_2}{r_{12}^2c^2}}\{v_{12}^i[4(n_{12}v_1)3(n_{12}v_2)]`$ (129)
$`+n_{12}^i[v_1^22v_2^2+4(v_1v_2)+{\displaystyle \frac{3}{2}}(n_{12}v_2)^2+5{\displaystyle \frac{Gm_1}{r_{12}}}+4{\displaystyle \frac{Gm_2}{r_{12}}}]\}`$
$`+`$ $`{\displaystyle \frac{Gm_2}{r_{12}^2c^4}}n_{12}^i\{[2v_2^4+4v_2^2(v_1v_2)2(v_1v_2)^2+{\displaystyle \frac{3}{2}}v_1^2(n_{12}v_2)^2+{\displaystyle \frac{9}{2}}v_2^2(n_{12}v_2)^2`$ (135)
$`6(v_1v_2)(n_{12}v_2)^2{\displaystyle \frac{15}{8}}(n_{12}v_2)^4]`$
$`+{\displaystyle \frac{Gm_1}{r_{12}}}[{\displaystyle \frac{15}{4}}v_1^2+{\displaystyle \frac{5}{4}}v_2^2{\displaystyle \frac{5}{2}}(v_1v_2)`$
$`+{\displaystyle \frac{39}{2}}(n_{12}v_1)^239(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{17}{2}}(n_{12}v_2)^2]`$
$`+{\displaystyle \frac{Gm_2}{r_{12}}}\left[4v_2^28(v_1v_2)+2(n_{12}v_1)^24(n_{12}v_1)(n_{12}v_2)6(n_{12}v_2)^2\right]`$
$`+{\displaystyle \frac{G^2}{r_{12}^2}}[{\displaystyle \frac{57}{4}}m_1^29m_2^2{\displaystyle \frac{69}{2}}m_1m_2]\}`$
$`+`$ $`{\displaystyle \frac{Gm_2}{r_{12}^2c^4}}v_{12}^i\{v_1^2(n_{12}v_2)+4v_2^2(n_{12}v_1)5v_2^2(n_{12}v_2)4(v_1v_2)(n_{12}v_1)`$ (138)
$`+4(v_1v_2)(n_{12}v_2)6(n_{12}v_1)(n_{12}v_2)^2+{\displaystyle \frac{9}{2}}(n_{12}v_2)^3`$
$`+{\displaystyle \frac{Gm_1}{r_{12}}}\left[{\displaystyle \frac{63}{4}}(n_{12}v_1)+{\displaystyle \frac{55}{4}}(n_{12}v_2)\right]+{\displaystyle \frac{Gm_2}{r_{12}}}\left[2(n_{12}v_1)2(n_{12}v_2)\right]`$
$`+`$ $`{\displaystyle \frac{4G^2m_1m_2}{5c^5r_{12}^3}}\{n_{12}^i(n_{12}v_{12})[6{\displaystyle \frac{Gm_1}{r_{12}}}+{\displaystyle \frac{52}{3}}{\displaystyle \frac{Gm_2}{r_{12}}}+3v_{12}^2]`$ (140)
$`+v_{12}^i[2{\displaystyle \frac{Gm_1}{r_{12}}}8{\displaystyle \frac{Gm_2}{r_{12}}}v_{12}^2]\}+𝒪(6).`$
Unavoidably, because of the proliferation of possible terms, the equations of motion at the next 3PN order are even much longer \[see (423) below\].
## IV Compact support and quadratic potentials
All the potentials that enter the linear momentum $`(P_1^i)_{_{\mathrm{distr}}}`$ and the force density $`(F_1^i)_{_{\mathrm{distr}}}`$ are computed at the point 1 by means of the Lorentzian regularization $`[F]__1`$. However, we shall first determine their Hadamard partie finie in the usual sense $`(F)__1`$, i.e. by approaching the singularity in the spatial slice $`t=\mathrm{const}`$. The difference between the two regularization processes does not affect any compact or quadratic potentials.
### A Iterative computation of compact support potentials
In this paragraph, we are interested in the compact terms involved in the equation of motion \[see (115) and (127)\]. According to our previous remark, it is sufficient to evaluate them with the classical Hadamard prescription. We need $`(_iV)__1`$ up to 3PN order; $`(V)__1`$, $`(V_i)__1`$, and $`(_iV_j)__1`$ at 2PN; $`(\widehat{W}_{ij}^{(\mathrm{C})})__1`$, $`(_i\widehat{W}_{jk}^{(\mathrm{C})})__1`$, $`(\widehat{R}_i^{(\mathrm{C})})__1`$, $`(_i\widehat{R}_j^{(\mathrm{C})})__1`$ and $`(_i\widehat{X}^{(\mathrm{C})})__1`$ at 1PN. The remaining contributions are Newtonian: $`(\widehat{X}^{(\mathrm{C})})__1`$, $`(_i\widehat{T}^{(\mathrm{C})})__1`$, $`(\widehat{Y}_i^{(\mathrm{C})})__1`$ and $`(_i\widehat{Y}_j^{(\mathrm{C})})__1`$. We follow the same classification and nomenclature concerning the various parts of potentials — compact, non-compact, etc. — as in Section II of . The compact (C) potentials are generated by sources with (spatially compact) support limited to the particles; for instance, $`V^{(\mathrm{C})}=V`$, and, from (85),
$$\widehat{T}^{(\mathrm{C})}=\mathrm{}_{}^1\left[4\pi G\left(\frac{1}{4}\sigma _{ij}\widehat{W}_{ij}+\frac{1}{2}V^2\sigma _{ii}+\sigma V_iV_i\right)\right].$$
(141)
Thus, by definition, the source $`S(𝐱,t)`$ of each compact potential $`P^{(\mathrm{C})}`$ is made of Dirac pseudo-functions, multiplied by some functions of the class $``$:
$$S(𝐱,t)=\mathrm{Pf}(F\mathrm{\Delta }_1)+\mathrm{Pf}(G\mathrm{\Delta }_2),$$
with $`F`$, $`G`$. As a result, it is in general possible to find an explicit expression of $`P^{(\mathrm{C})}`$ over the whole space (for any $`𝐱`$). Besides, the expansion under the integration symbol of the retardation of $`S(𝐱^{},t|\mathrm{x}\mathrm{x}^{}|/c)`$ as $`c`$ goes to infinity is perfectly licit, because the integrand has a compact support:
$`P^{(\mathrm{C})}`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{n=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{()^n}{n!c^n}}_t^n{\displaystyle d^3𝐱^{}|𝐱𝐱^{}|^{n1}S(𝐱^{},t)}`$ (142)
$`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{n=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{()^n}{n!c^n}}_t^n\left([F|𝐱𝐱^{}|^{n1}]__1+[G|𝐱𝐱^{}|^{n1}]__2\right).`$ (143)
The sources $`S(𝐱,t)`$ are supposed to be known at the current order. This implies to proceed iteratively as explained in . The reader is referred to this paper for more details. In short, we start from the $`V`$ and $`V_i`$ potentials, whose sources do not depend on any other ones at the lowest order. Indeed, we have $`\mathrm{}V=4\pi G\mathrm{Pf}(\stackrel{~}{\mu }_1\mathrm{\Delta }_1)+12`$ (and similarly for $`V_i`$), where $`\stackrel{~}{\mu }_1=m_1+𝒪(2)`$, as it follows from insertion of the “Newtonian” metric into the definition (III A) of the effective mass. Hence,
$`V`$ $`=`$ $`G{\displaystyle \frac{d^3𝐱^{}}{|𝐱𝐱^{}|}\left\{\mathrm{Pf}\left[\stackrel{~}{\mu }_1\mathrm{\Delta }_1\right]\frac{1}{c}_t\mathrm{Pf}\left[|𝐱𝐱^{}|\stackrel{~}{\mu }_1\mathrm{\Delta }_1\right]\right\}}+12+𝒪(2)`$ (144)
$`=`$ $`{\displaystyle \frac{Gm_1}{r_1}}{\displaystyle \frac{G}{c}}_tm_1+12+𝒪(2)={\displaystyle \frac{Gm_1}{r_1}}+{\displaystyle \frac{Gm_2}{r_2}}+𝒪(2),`$ (145)
To obtain the regularized metric (at the location of the first body, say), we need the partie finie of the potential $`V`$ at point 1, $`(V)__1`$. Since we use Hadamard regularization, it is simply given by the value of its non-singular part when $`𝐱=𝐲_1`$. Here, we find $`(V)__1=Gm_2/r_{12}+𝒪(2)`$, with the notation $`r_{12}=|𝐲_1𝐲_2|`$.
The computation of more complicated compact terms necessitates the knowledge of the effective masses $`\mu _1`$ and $`\stackrel{~}{\mu }_1`$ beyond the Newtonian approximation. By substituting to $`g_{\mu \nu }`$ the explicit 3PN expression (III C) for the metric in the equations (III A), we get the general forms of both effective masses. As an example, $`\stackrel{~}{\mu }_1`$ at 2PN order reads
$`{\displaystyle \frac{\stackrel{~}{\mu }_1}{m_1}}`$ $`=`$ $`1+{\displaystyle \frac{1}{c^2}}\left[2V+[V]__1+{\displaystyle \frac{3}{2}}v_1^2\right]`$ (146)
$`+`$ $`{\displaystyle \frac{1}{c^4}}\left[2\widehat{W}_{ii}+2V^22V[V]__1+{\displaystyle \frac{3}{2}}[V]__1^2[V^2]__13Vv_1^2+{\displaystyle \frac{7}{2}}[V]__1v_1^24[V_i]__1v_1^i+{\displaystyle \frac{7}{8}}v_1^4\right]`$ (147)
$`+`$ $`𝒪(6),`$ (148)
where we are careful at distinguishing the potentials computed at the field point $`𝐱`$ from those computed at the source point $`𝐲_1`$, and where we take into account the non-distributivity of the regularization ($`\mu _1`$ differs only by some numerical coefficients). Thus, as emphasized in Section III, $`\mu _1`$ and $`\stackrel{~}{\mu }_1`$ are functions of time *and* space. Replacing them by the regularized quantities $`(\mu _1)__1`$, $`(\stackrel{~}{\mu }_1)__1`$ (and $`12`$) is definitely forbidden because, on one side, the partie finie is not distributive, and, on the other side, the usual Hadamard regularization does not coincide with the Lorentzian one. However, this replacement does not modify any compact potentials, with the notable exception of the 3PN contributions in $`V`$ (see below). It is thus convenient to pose:
$$V_{\mathrm{distr}}=\mathrm{}_{}^1\left[4\pi G(\stackrel{~}{\mu }_1)__1\mathrm{Pf}\delta _14\pi G(\stackrel{~}{\mu }_2)__2\mathrm{Pf}\delta _2\right],$$
and to calculate $`V_{\mathrm{distr}}`$ and $`VV_{\mathrm{distr}}`$ separately. In the other compact sources, we shall employ $`(\mu _1)__1`$, $`(\stackrel{~}{\mu }_1)__1`$, etc. instead of $`\mu _1`$ and $`\stackrel{~}{\mu }_1`$ for practical calculations at the 3PN approximation. Furthermore, in all the compact terms, the action of the Lorentzian delta-pseudo-functions $`\mathrm{Pf}\mathrm{\Delta }_1`$ and $`\mathrm{Pf}\mathrm{\Delta }_2`$, remarkably, reduces to the one of $`\mathrm{Pf}\delta _1`$ and $`\mathrm{Pf}\delta _2`$. From what precedes, it becomes obvious that, after the evaluation of $`(\mu _1)__1`$ or $`(\stackrel{~}{\mu }_1)__1`$ at a given post-Newtonian order $`n`$, we can determine all the potentials to the precision $`1/c^{2n}`$. As all the terms involving the retarded potentials in $`\stackrel{~}{\mu }_1`$ appear already with a factor $`1/c^2`$ at least, we are then in a position to compute the right-hand-side of the equation (148). The process is initiated by the computation of the Newtonian value of $`V`$ as presented above. Most of the quantities needed to get $`(\stackrel{~}{\mu }_1)__1`$ at the 3PN order are obtained in . Finally, the regularized value of $`\stackrel{~}{\mu }_1`$ at point 1 is
$`{\displaystyle \frac{(\stackrel{~}{\mu }_1)__1}{m_1}}`$ $`=1+{\displaystyle \frac{1}{c^2}}[{\displaystyle \frac{Gm_2}{r_{12}}}+{\displaystyle \frac{3}{2}}v_1^2]+{\displaystyle \frac{1}{c^4}}[{\displaystyle \frac{Gm_2}{r_{12}}}({\displaystyle \frac{1}{2}}v_1^24(v_1v_2)+2v_2^2+{\displaystyle \frac{1}{2}}(n_{12}v_2)^2{\displaystyle \frac{1}{2}}{\displaystyle \frac{Gm_1}{r_{12}}}`$ (149)
$`+{\displaystyle \frac{3}{2}}{\displaystyle \frac{Gm_2}{r_{12}}})+{\displaystyle \frac{7}{8}}v_1^4]+{\displaystyle \frac{8G^2m_1m_2}{3c^5r_{12}^2}}((n_{12}v_1)(n_{12}v_2))`$ (150)
$`+{\displaystyle \frac{1}{c^6}}[{\displaystyle \frac{G^2m_1m_2}{r_{12}^2}}({\displaystyle \frac{3}{2}}{\displaystyle \frac{Gm_1}{r_{12}}}{\displaystyle \frac{39}{4}}{\displaystyle \frac{Gm_2}{r_{12}}}+{\displaystyle \frac{35}{8}}v_1^2{\displaystyle \frac{41}{4}}(v_1v_2)+{\displaystyle \frac{41}{8}}v_2^2{\displaystyle \frac{9}{8}}(n_{12}v_1)^2`$ (151)
$`+{\displaystyle \frac{25}{4}}(n_{12}v_1)(n_{12}v_2){\displaystyle \frac{41}{8}}(n_{12}v_2)^2)`$ (152)
$`+{\displaystyle \frac{G^2m_2^2}{r_{12}^2}}\left({\displaystyle \frac{3}{2}}v_2^2(n_{12}v_2)^23(v_1v_2)(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{1}{2}}(n_{12}v_1)^2{\displaystyle \frac{3}{2}}{\displaystyle \frac{Gm_2}{r_{12}}}+{\displaystyle \frac{15}{4}}v_1^2\right)`$ (153)
$`+{\displaystyle \frac{Gm_2}{r_{12}}}(2(v_1v_2)^2+5v_1^2v_2^210(v_1v_2)v_1^2+{\displaystyle \frac{33}{8}}v_1^4+2v_2^4{\displaystyle \frac{1}{2}}(n_{12}v_2)^2v_2^2{\displaystyle \frac{3}{8}}(n_{12}v_2)^4`$ (154)
$`4v_2^2(v_1v_2)+2(n_{12}v_2)^2(v_1v_2){\displaystyle \frac{1}{4}}v_1^2(n_{12}v_2)^2)+{\displaystyle \frac{11}{16}}v_1^6]+𝒪(7).`$ (155)
In our notation, two vectors $`𝐯_1`$, $`𝐯_2`$ between brackets represent the scalar product: $`(v_1v_2)=v_1^iv_2^i`$; $`v_1^2=v_1^iv_1^i`$. We recall that it is important to keep the grouping of factors imposed by the regularization in products of potentials. For instance: $`(V\widehat{W}_{ij})__1(V)__1(\widehat{W}_{ij})__1`$.
Among the compact potentials, the 3PN value of $`V`$ is certainly the most difficult one to obtain, since the other quantities require only lower orders in powers of $`1/c`$. We shall focus on $`V_{\mathrm{distr}}`$ to illustrate the method we have followed. The difference $`VV_{\mathrm{dist}}`$ will be handled in the next subsection. We begin with specializing the general formula for $`V`$ to the case of $`V_{\mathrm{distr}}`$:
$`V_{\mathrm{distr}}`$ $`=`$ $`{\displaystyle \frac{G(\stackrel{~}{\mu }_1)__1}{r_1}}{\displaystyle \frac{G}{c}}_t(\stackrel{~}{\mu }_1)__1+{\displaystyle \frac{G}{2c^2}}_t^2\left[(\stackrel{~}{\mu }_1)__1r_1\right]{\displaystyle \frac{G}{6c^3}}_t^3\left[(\stackrel{~}{\mu }_1)__1r_1^2\right]+{\displaystyle \frac{G}{24c^4}}_t^4\left[(\stackrel{~}{\mu }_1)__1r_1^3\right]`$ (157)
$`{\displaystyle \frac{G}{120c^5}}_t^5\left[(\stackrel{~}{\mu }_1)__1r_1^4\right]+{\displaystyle \frac{G}{720c^6}}_t^6\left[(\stackrel{~}{\mu }_1)__1r_1^5\right]+12+𝒪(7).`$
Since the Schwarzschild mass $`m_1`$ is constant, $`_t\stackrel{~}{\mu }_1/c`$ is of order $`𝒪(3)`$ and does not contribute at the 1PN level. For convenience, we shall introduce some special notation for the terms that occur at this approximation; we pose:
$$U=\frac{G(\stackrel{~}{\mu }_1)__1}{r_1}+\frac{G(\stackrel{~}{\mu }_2)__2}{r_2}\mathrm{and}X=G(\stackrel{~}{\mu }_1)__1r_1+G(\stackrel{~}{\mu }_2)__2r_2.$$
Actually the potentials are to be considered as pseudo-functions and it is understood that there is a symbol $`\mathrm{Pf}`$ in front of them. Notably, the time derivatives appearing in (157) are distributional. The regularized effective mass $`(\stackrel{~}{\mu }_1)__1`$ as well as the distance to the first body $`r_1`$ depend on time through the trajectories $`𝐲_{1,2}(t)`$ and velocities $`𝐯_{1,2}(t)`$. We explicit the time-differentiations and obtain, at 1PN,
$`V_{\mathrm{distr}}`$ $`=`$ $`U+{\displaystyle \frac{1}{2c^2}}_t^2X+𝒪(3)`$ (158)
$`=`$ $`{\displaystyle \frac{Gm_1}{r_1}}\left[1+{\displaystyle \frac{1}{c^2}}\left({\displaystyle \frac{Gm_2}{r_{12}}}+{\displaystyle \frac{3}{2}}v_1^2\right)\right]+{\displaystyle \frac{Gm_1}{2c^2}}\left(a_1^i_ir_1+v_1^iv_1^j_{ij}^2r_1\right)+12+𝒪(3).`$ (159)
The accelerations are order-reduced by means of the equations of motion at previous orders. Notably, for computing the 1PN term $`\frac{G}{2c^2}_t^2\mathrm{Pf}[(\stackrel{~}{\mu }_1)__1r_1]`$ at relative order 3PN, we need the 2PN acceleration given by (129). Once we have got $`V_{\mathrm{distr}}`$ all over the space, the last stage consists of regularizing it, as well as its gradient, at $`𝐱=𝐲_1`$ using the Hadamard partie finie. Now, $`V_{\mathrm{distr}}`$ can be divided into two parts, $`V_{\mathrm{distr}𝐫_1}`$ and $`V_{\mathrm{distr}𝐫_2}`$ corresponding to the sources $`4\pi G(\stackrel{~}{\mu }_1)__1\mathrm{Pf}\delta _1`$ and $`4\pi G(\stackrel{~}{\mu }_2)__2\mathrm{Pf}\delta _2`$ respectively. The first part, $`V_{\mathrm{distr}𝐫_1}`$, depends on $`𝐱`$ through $`𝐫_1`$ only, and contains many terms that are either singular or vanish when $`𝐱𝐲_1`$, giving no contribution to the partie finie; on the opposite, the smooth terms with odd $`1/c`$-power factors in (157) generally contribute. The part $`V_{\mathrm{distr}𝐫_2}`$ does not necessitate any regularization since it is already regular in the neighbourhood of $`𝐱=𝐲_1`$.
The remaining potentials are determined in the same way. However, we have to apply properly the formalism developed in . In particular:
(1) The regularized value of some potential $`P^{(\mathrm{C})}`$ is the partie finie of $`P^{(\mathrm{C})}`$ computed initially outside the singularity. In the case where $`P^{(\mathrm{C})}`$ is the Poisson integral of a compact source $`\mathrm{Pf}(F\delta _1)`$, with $`F`$, we must take care that
$$P_1^{(\mathrm{C})}=\left(\mathrm{Pf}\frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}F\delta _1\right)__1\mathrm{Pf}\frac{d^3𝐱}{4\pi }\frac{1}{r_1}F\delta _1.$$
(160)
\[We generally do not write the dependence of the integrand on the integration variable, as it is evident from the context; thus, $`F\delta _1`$ is computed at point $`𝐱^{}`$ in the intermediate expression of (160) and at point $`𝐱`$ in the last one.\]
(2) If $`F`$ is not regular at point 1, we generally have $`\mathrm{Pf}(F\delta _1)(F)__1\mathrm{Pf}\delta _1`$, even when both members act, in the sense of pseudo-functions, on smooth test functions. This distinction is crucial, for instance, in the determination of $`\widehat{T}^{(\mathrm{C})}`$ at Newtonian order. Indeed, one of its contributions \[first term in (141)\], denoted by $`\widehat{T}^{(\mathrm{C1})}`$, reads as:
$`\widehat{T}^{(\mathrm{C1})}`$ $`=`$ $`{\displaystyle \frac{1}{4}}G\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{|𝐱𝐱^{}|}\sigma _{ij}\widehat{W}_{ij}}+𝒪(1)`$
$`=`$ $`{\displaystyle \frac{1}{4}}Gm_1v_1^iv_1^j\left({\displaystyle \frac{\widehat{W}_{ij}}{|𝐱𝐱^{}|}}\right)__1+12+𝒪(1),`$
which is different from $`Gm_1v_1^iv_1^j(\widehat{W}_{ij})__1/4r_1+12+𝒪(1).`$ Had we used the latter expression instead of $`\widehat{T}^{(\mathrm{C1})}`$, we would have obtained a different potential $`\widehat{T}_{\mathrm{distr}}`$; this would have been correct if the partie finie operation had been “distributive” (see Section I), but we have actually
$$\widehat{T}\widehat{T}_{\mathrm{distr}}=\frac{G^3m_1^3}{r_1^3}\left[\frac{1}{240}v_1^2+\frac{1}{80}(n_1v_1)^2\right]+12.$$
Notice that the latter expression is not Galilean-invariant by itself, and therefore will be checked later when verifying that the final equations of motion stay invariant under Lorentz transformations.
### B Non-distributivity in the potential $`V`$
We call non-distributivity in the potential $`V`$ that contribution which arises because the coefficient of the delta-pseudo-function $`\mathrm{Pf}\mathrm{\Delta }_1`$ in the matter stress-energy tensor (15) is a function not only on time but also on space through the factor $`1/\sqrt{g}`$. It will turn out that this contribution is purely of order 3PN. A related contribution, due to the non-distributivity in $`\widehat{T}`$, has just been computed in the previous subsection. The potential $`V`$ is generated by the source density $`\sigma (𝐱,t)=\stackrel{~}{\mu }_1\mathrm{\Delta }[𝐱𝐲_1(t)]+12`$, where $`\stackrel{~}{\mu }_1`$ is a function of space-time given explicitly by
$$\stackrel{~}{\mu }_1(𝐱,t)=\frac{m_1c\left(1+\frac{𝐯_1^2}{c^2}\right)}{\sqrt{[g_{\rho \sigma }]__1v_1^\rho v_1^\sigma }}.\frac{1}{\sqrt{g(𝐱,t)}}.$$
(161)
The first factor is a function of time, and the second one depends on both time and space (non-distributivity). The potential $`V`$ is given by the retarded integral (76), whose retardations we expand up to any post-Newtonian order:
$$V(𝐱,t)=G\underset{n=0}{\overset{+\mathrm{}}{}}\frac{()^n}{n!c^n}_t^nd^3𝐱^{}|𝐱𝐱^{}|^{n1}\sigma (𝐱^{},t).$$
(162)
(Actually we shall see that an expansion to the 1PN order is sufficient for our purpose.) We insert into that expression the source density $`\sigma `$, use the definition of the delta-pseudo-function $`\mathrm{Pf}(F\mathrm{\Delta }_1)`$ given by (14), and arrive at
$$V(𝐱,t)=G\underset{n=0}{\overset{+\mathrm{}}{}}\frac{()^n}{n!c^n}_t^n\left[\stackrel{~}{\mu }_1(𝐱^{},t)|𝐱𝐱^{}|^{n1}\right]__1+12.$$
(163)
Here, the square brackets refer to the Lorentzian Hadamard regularization when $`𝐱^{}𝐲_1`$. Using a multipolar expansion, we obtain immediately these brackets as
$$\left[\stackrel{~}{\mu }_1(𝐱^{},t)|𝐱𝐱^{}|^{n1}\right]__1=\underset{l=0}{\overset{+\mathrm{}}{}}\frac{()^l}{l!}_L\left(r_1^{n1}\right)\left[r_{}^{}{}_{1}{}^{l}n_{}^{}{}_{1}{}^{L}\stackrel{~}{\mu }_1^{}\right]__1,$$
(164)
where $`\stackrel{~}{\mu }_1^{}\stackrel{~}{\mu }_1(𝐱,t)`$. If $`\stackrel{~}{\mu }_1`$ were a function of time only, then we see that all the multipolar contributions in the right-hand side of (164) but the scalar $`l=0`$ one would be zero, because of the factor $`r_{}^{}{}_{1}{}^{l}`$ with $`l1`$ (this is clear with the old regularization, and easily checked to be true with the Lorentzian regularization as well). We defined $`V_{\mathrm{distr}}`$ as being $`V`$ but computed with the function of time $`[\stackrel{~}{\mu }_1]__1`$ instead of the true $`\stackrel{~}{\mu }_1(𝐱,t)`$. This $`V_{\mathrm{distr}}`$ is exactly the one which has been computed in the subsection IV A. \[It can be checked that up to the 3PN order $`[\stackrel{~}{\mu }_1]__1=(\stackrel{~}{\mu }_1)__1`$.\] Therefore, by the previous argument, $`V_{\mathrm{distr}}`$ is produced entirely by the scalar part $`l=0`$ in the latter multipolar expansion, so that its complementary to the true potential $`V`$ reads as
$$VV_{\mathrm{distr}}=G\underset{n=0}{\overset{+\mathrm{}}{}}\frac{()^n}{n!c^n}\frac{^n}{t^n}\left\{\underset{l=1}{\overset{+\mathrm{}}{}}\frac{()^l}{l!}_L\left(r_1^{n1}\right)\left[r_{}^{}{}_{1}{}^{l}n_{}^{}{}_{1}{}^{L}\stackrel{~}{\mu }_1^{}\right]__1\right\}+12,$$
(165)
where the sum over $`l`$ starts with $`l=1`$. Thus, the problem reduces to the computation of each regularization terms $`[r_{}^{}{}_{1}{}^{l}n_{}^{}{}_{1}{}^{L}\stackrel{~}{\mu }_1^{}]__1`$. Obviously, at a given post-Newtonian order, these terms will all become zero for $`l`$ larger than a certain value. We find that, up to the 3PN order, all the regularizations are zero starting at $`l=3`$, namely $`[r_{}^{}{}_{l}{}^{}n_{}^{}{}_{1}{}^{L}\stackrel{~}{\mu }_1^{}]__1=𝒪(8)`$ for any $`l3`$, while the non-zero values for $`l=1,2`$ are given by
$`\left[r_{}^{}{}_{1}{}^{}n_{}^{}{}_{1}{}^{i}\stackrel{~}{\mu }_1^{}\right]__1=3{\displaystyle \frac{G^3m_1^3m_2}{c^6r_{12}^2}}n_{12}^i+𝒪(8),`$ (167)
$`\left[r_{}^{}{}_{2}{}^{}n_{}^{}{}_{1}{}^{ij}\stackrel{~}{\mu }_1^{}\right]__1={\displaystyle \frac{G^2m_1^3}{c^4}}\delta ^{ij}3{\displaystyle \frac{G^3m_1^3m_2}{c^6r_{12}}}\delta ^{ij}+{\displaystyle \frac{3}{2}}{\displaystyle \frac{G^2m_1^3}{c^6}}v_1^2\delta ^{ij}{\displaystyle \frac{G^2m_1^3}{c^6}}v_1^iv_1^j+𝒪(8).`$ (168)
Replacing these results back into (165), and using the fact that $`\delta ^{ij}_{ij}\frac{1}{r_1}=0`$, leads to the intermediate form
$`VV_{\mathrm{distr}}`$ $`=`$ $`3{\displaystyle \frac{G^4m_1^3m_2}{c^6r_{12}^2}}n_{12}^i_i\left({\displaystyle \frac{1}{r_1}}\right){\displaystyle \frac{G^3m_1^3}{2c^6}}v_1^iv_1^j_{ij}\left({\displaystyle \frac{1}{r_1}}\right)+{\displaystyle \frac{G^3m_1^3}{2c^6}}{\displaystyle \frac{^2}{t^2}}\left({\displaystyle \frac{1}{r_1}}\right)`$ (169)
$`+`$ $`𝒪(8)+12.`$ (170)
As we see, the non-distributivity of the potential $`V`$ is a 3PN effect. Expanding the time derivative in the last term we find that the dependence on the velocity $`v_1^i`$ cancels out, which is normal because a velocity-dependent term would violate the Lorentz invariance, in contradiction with our use of the Lorentzian regularization $`[F]__1`$. The final expression is simple:
$$VV_{\mathrm{distr}}=\frac{5}{2}\frac{G^4m_1^3m_2}{c^6r_{12}^2}n_{12}^i_i\left(\frac{1}{r_1}\right)+𝒪(8)+12.$$
(171)
The contribution of the non-distributivity in the acceleration of 1 is given by the gradient at 1 as
$$[_iV]__1[_iV_{\mathrm{distr}}]__1=5\frac{G^4m_1m_2^3}{c^6r_{12}^5}n_{12}^i+𝒪(8).$$
(172)
### C Computation of quadratic potentials $`(VV)`$
By definition, the quadratic potentials are those whose sources are made of products of two compact factors, like $`V`$, $`V_i`$, $`W_{ij}^{(\mathrm{C})}`$, etc. (or their derivatives, in most of the time). A typical source term for them is of the type “$`VV`$”, hence their denomination; for instance
$$\widehat{W}_{ij}^{(VV)}=\mathrm{}_{}^1[_iV_jV].$$
(173)
But the quadratic source terms may also involve other quantities of the same structure, as it is the case for $`_t\widehat{W}_{ik}^{(\mathrm{C})}_kV`$ appearing in the source of the potential $`\widehat{Y}_i^{(VV)}`$ \[cf (88)\].
#### 1 Matching to the external field
The retardation of the compact potentials defining the metric of an isolated fluid can be expanded in powers of $`1/c`$ only in the “near zone” $`D_{\mathrm{near}}`$ of the source, at a distance much smaller than the typical wave length of the emitted radiation. The question then is how to incorporate in the post-Newtonian metric the no-incoming radiation conditions at past null infinity. We achieve this by performing a matching between the post-Newtonian expansion of the metric, adequate in the near zone, and its multipole expansion, valid in the region $`D_{\mathrm{ext}}`$ exterior to the compact support of the source. Recall that for slowly moving sources, one can always choose $`D_{\mathrm{near}}`$ and $`D_{\mathrm{ext}}`$ in such a way that their intersection is not empty: $`D_{\mathrm{near}}D_{\mathrm{ext}}\mathrm{}`$ (see e.g. ). The field $`h^{\mu \nu }`$ admits a multipole-type expansion $`(h^{\mu \nu })`$, in the sense of , at every spatial point $`𝐱D_{\mathrm{ext}}`$. As a matter of fact, it is shown in that the multipole expansion of the exterior field (a vacuum solution of the field equations) that matches, according to the theory of matched asymptotic expansions, to the post-Newtonian expansion in the interior of the source, is given by
$$(h^{\mu \nu })=\mathrm{FP}_{B0}\mathrm{}_{}^1\left[\left(\frac{r}{r_0}\right)^B(\mathrm{\Lambda }^{\mu \nu })\right]\frac{4G}{c^4}\underset{l=0}{\overset{+\mathrm{}}{}}\frac{()^l}{l!}_L\left\{\frac{1}{r}_L^{\mu \nu }(tr/c)\right\}$$
(174)
(with $`L`$ a multi-index of order $`l`$). The multipole moments $`_L^{\mu \nu }`$ entering the right-hand-side read as
$$_L^{\mu \nu }(u)=\mathrm{FP}_{B0}d^3𝐲\left(\frac{|𝐲|}{r_0}\right)^By_L\overline{\tau }^{\mu \nu }(𝐲,u),$$
where $`\overline{\tau }^{\mu \nu }`$ represents the (formal) post-Newtonian expansion of the complete source term $`\tau ^{\mu \nu }=|g|T^{\mu \nu }+\frac{c^4}{16\pi G}\mathrm{\Lambda }^{\mu \nu }`$ of the field equations (18). These expressions are defined by analytic continuation in $`B`$, and the symbol $`\mathrm{FP}_{B0}`$ denotes the finite part when $`B`$ goes to zero of the Laurent expansion of the analytic continuation (we refer to for more details about this finite part).
Let us show how we find the “matched” solution of the equation $`\mathrm{}P=S`$ at the relative 1PN order (this is all we shall need in this paper). We neglect all higher-order post-Newtonian contributions in the source term $`S`$, and look for the solution of
$$\mathrm{}P=S^{1\mathrm{P}\mathrm{N}}+𝒪(3).$$
(175)
Since the formula (174) results from the properties of the d’Alembertian operator (and is not specific to the field variable $`h^{\mu \nu }`$), we can use it with the replacements of $`(\mathrm{\Lambda })`$ by $`(S^{1\mathrm{P}\mathrm{N}})`$ and of $`\overline{\tau }`$ by $`\frac{c^4}{16\pi G}S^{1\mathrm{P}\mathrm{N}}`$. Thus, the multipole expansion of the solution must satisfy
$`(P)`$ $`=`$ $`\mathrm{FP}_{B0}\mathrm{}_{}^1\left[\left({\displaystyle \frac{r}{r_0}}\right)^B(S^{1\mathrm{P}\mathrm{N}})\right]{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{l=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{()^l}{l!}}_L\left\{{\displaystyle \frac{1}{r}}𝒫_L(tr/c)\right\}`$ (176)
$`+`$ $`𝒪(3),`$ (177)
$$\mathrm{with}𝒫_L[S^{1\mathrm{P}\mathrm{N}}](u)=\mathrm{FP}_{B0}d^3𝐲\left(\frac{|𝐲|}{r_0}\right)^By_LS^{1\mathrm{P}\mathrm{N}}(𝐲,u).$$
The partie-finie retarded integral of the multipole source $`(S^{1\mathrm{P}\mathrm{N}})`$ has to be handled with care. It is not licit to develop when $`c+\mathrm{}`$ the integrand under the integration symbol because the source is not compact supported. The correct formula was shown in to be:
$`\mathrm{FP}_{B0}\mathrm{}_{}^1\left[\left({\displaystyle \frac{r}{r_0}}\right)^B\left(S^{1\mathrm{P}\mathrm{N}}\right)\right]`$ (178)
$`=\mathrm{FP}_{B0}{\displaystyle \underset{k=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{1}{(2k)!}}\left({\displaystyle \frac{}{ct}}\right)^{2k}{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}|𝐱𝐱^{}|^{2k1}\left({\displaystyle \frac{r^{}}{r_0}}\right)^B\left(S^{1\mathrm{P}\mathrm{N}}\right)`$ (179)
$`{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{l0}{}}{\displaystyle \frac{()^l}{l!}}\widehat{}_L\left\{{\displaystyle \frac{_L(tr/c)_L(t+r/c)}{2r}}\right\}.`$ (180)
The hat on the partial derivatives $`\widehat{}_L`$ indicates that the trace has been removed, i.e. $`\widehat{}_L=\mathrm{STF}(_L)`$. The $`_L`$ functions parametrize the general solution of d’Alembert equations that are smooth near the origin: “antisymmetric” solution as given by the last term in (180). We have, more precisely,
$`_L(u)`$ $`=`$ $`\mathrm{FP}_{B0}{\displaystyle d^3𝐲\widehat{y}^L\left(\frac{|𝐲|}{r_0}\right)^BT_l(𝐲,u)},`$ (181)
$`\mathrm{with}:T_l(𝐲,u)`$ $`=`$ $`()^{l+1}{\displaystyle \frac{(2l+1)!!}{2^ll!}}{\displaystyle _1^+\mathrm{}}𝑑z(z^21)^l\left(S^{1\mathrm{P}\mathrm{N}}\right)(𝐲,uz|𝐲|/c).`$ (182)
Here, $`\widehat{y}^L`$ denotes the symmetric trace-free tensor associated with $`y^{i_1}\mathrm{}y^{i_l}`$, for $`l`$. With (177) and (180) we can write the multipole expansion $`(P)`$ at the 1PN order as
$`(P)`$ $`=`$ $`\mathrm{FP}_{B0}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\left(\frac{r^{}}{r_0}\right)^B\left(S^{1\mathrm{P}\mathrm{N}}\right)}{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{l=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{()^l}{l!}}_L\left({\displaystyle \frac{1}{r}}\right)𝒫_L(t)`$ (183)
$`+`$ $`{\displaystyle \frac{1}{2c^2}}\left({\displaystyle \frac{}{t}}\right)^2\left\{\mathrm{FP}_{B0}{\displaystyle \frac{d^3𝐱^{}}{4\pi }}\right|𝐱𝐱^{}|\left({\displaystyle \frac{r^{}}{r_0}}\right)^B\left(S^{1\mathrm{P}\mathrm{N}}\right){\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{l=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{()^l}{l!}}_L\left(r\right)𝒫_L(t)\}`$ (184)
$`+`$ $`{\displaystyle \frac{1}{4\pi c}}\left[\dot{}(t)+\dot{𝒫}(t)\right]+𝒪(3)`$ (185)
where the last term, of order $`1/c`$, is a simple function of time made of the functions $`(t)`$ and $`𝒫(t)`$ defined as being $`_L(t)`$ and $`𝒫_L(t)`$ with $`l=0`$ (the dot indicates the time-derivative). Now, it can be shown that the latter multipole expansion can be re-written under the new form
$$(P)=(P^{(\mathrm{I})})+\frac{1}{4\pi c}\left[\dot{}(t)+\dot{𝒫}(t)\right]+\frac{1}{2c^2}\left(\frac{}{t}\right)^2\left[(P^{(\mathrm{II})})\right]+𝒪(3),$$
or, equivalently (indeed the second term is a mere function of time, and the multipole expansion obviously commutes with the time derivative),
$$(P)=\left(P^{(\mathrm{I})}+\frac{1}{4\pi c}\left[\dot{}(t)+\dot{𝒫}(t)\right]+\frac{1}{2c^2}\left(\frac{}{t}\right)^2\left[P^{(\mathrm{II})}\right]\right)+𝒪(3).$$
(186)
In these equations, $`P^{(\mathrm{I})}`$ and $`P^{(\mathrm{II})}`$ denote the matched solutions of the following Poisson equations
$`\mathrm{\Delta }P^{(\mathrm{I})}`$ $`=`$ $`S^{1\mathrm{P}\mathrm{N}},`$ (188)
$`\mathrm{\Delta }P^{(\mathrm{II})}`$ $`=`$ $`2P^{(\mathrm{I})}.`$ (189)
Therefore, we have reduced the problem of finding the matched solution of the d’Alembertian equation (175) to that of solving and matching the two successive Poisson equations (188) and (189). Now, from the equation (186), it is evident that the correct matched solution of (175) reads in terms of the matched solutions of (IV C 1) as
$$P=P^{(\mathrm{I})}+\frac{1}{4\pi c}\left[\dot{}(t)+\dot{𝒫}(t)\right]+\frac{1}{2c^2}\left(\frac{}{t}\right)^2\left[P^{(\mathrm{II})}\right]+𝒪(3).$$
(190)
To recall the meaning of this solution we shall often denote it as $`P=P_{\mathrm{match}}`$ below; similarly for $`P=P_{\mathrm{match}}^{(\mathrm{I})}`$ (for instance $`g_{\mathrm{match}}`$ computed below) and $`P=P_{\mathrm{match}}^{(\mathrm{II})}`$ (e.g. $`f_{\mathrm{match}}`$). Actually, we shall find that the function $`(t)`$ appearing in the $`1/c`$ term of our solution (190) is in fact always either zero or cancelled out by a spatial gradient in the case of the applications made in the present paper. Thus, it will not be considered in this paper, whereas the function $`𝒫(t)`$ plays a role and is given by
$$𝒫(t)=\mathrm{Pf}\left\{\mathrm{FP}_{B0}d^3𝐱\left(\frac{r}{r_0}\right)^BS^{1\mathrm{P}\mathrm{N}}(𝐱,t)\right\}$$
(191)
(of course $`S^{1\mathrm{P}\mathrm{N}}`$ there could be replaced with this approximation by $`S^{0.5\mathrm{PN}}`$). See (217) below for an example of computation of this function.
In practice, in order to find the matched solution of a Poisson equation, $`P^{(\mathrm{I})}`$ for instance, we proceed as follows. Suppose that we know a particular solution of the equation, say $`P_{\mathrm{part}}^{(\mathrm{I})}`$. Then the correct solution is necessarily of the type $`P^{(\mathrm{I})}=P_{\mathrm{part}}^{(\mathrm{I})}+h_{\mathrm{hom}}^{(\mathrm{I})}`$, where $`h_{\mathrm{hom}}^{(\mathrm{I})}`$ denotes an homogeneous solution of the Laplace equation (harmonic function): $`\mathrm{\Delta }h_{\mathrm{hom}}^{(\mathrm{I})}=0`$, which is moreover regular at the location of the source points. Note that its multipole expansion coincides with itself, $`(h_{\mathrm{hom}}^{(\mathrm{I})})=h_{\mathrm{hom}}^{(\mathrm{I})}`$. Now, the latter homogeneous solution is determined by the matching equation as
$`h_{\mathrm{hom}}^{(\mathrm{I})}`$ $`=`$ $`\mathrm{FP}_{B0}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\left(\frac{r^{}}{r_0}\right)^B\left(S^{1\mathrm{P}\mathrm{N}}\right)}{\displaystyle \frac{1}{4\pi }}{\displaystyle \underset{l=0}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{()^l}{l!}}_L\left({\displaystyle \frac{1}{r}}\right)𝒫_L(t)`$ (192)
$``$ $`(P_{\mathrm{part}}^{(\mathrm{I})}).`$ (193)
It is not evident on that expression that the right-hand side is an harmonic function; but it really is, as can be verified explicitly in practice. We compute the multipole expansion of the source term $`S^{1\mathrm{P}\mathrm{N}}`$ as well as of our particular solution $`P_{\mathrm{part}}^{(\mathrm{I})}`$. In our case this means computing the formal expansions of $`S^{1\mathrm{P}\mathrm{N}}`$ and $`P_{\mathrm{part}}^{(\mathrm{I})}`$ when $`r`$ tends to infinity or equivalently when the two source points $`𝐲_{1,2}`$ tend to zero. The computation is greatly simplified if one considers the dimensionality of the source. Suppose for instance that $`[S^{1\mathrm{P}\mathrm{N}}]=(\mathrm{length})^d`$ which means $`[h_{\mathrm{hom}}^{(\mathrm{I})}]=(\mathrm{length})^{d+2}`$. Then, using the fact that this function is harmonic, its structure is necessarily of the type $`h_{\mathrm{hom}}^{(\mathrm{I})}\widehat{x}_Ly_1^{L_1}y_2^{L_2}`$ with $`l+l_1+l_2=d+2`$. This shows that in order to obtain $`h_{\mathrm{hom}}^{(\mathrm{I})}`$ completely it is sufficient to develop the right side of (192) when $`𝐲_{1,2}0`$ up to the order $`d+2`$ included (i.e. to control all the terms $`y_1^{L_1}y_2^{L_2}`$ in the expansions which have $`l_1+l_2d+2`$). All the higher-order terms, having $`l_1+l_2d+3`$ in the right side of (192) must manage to give zero. The same method is used to compute the homogeneous solution $`h_{\mathrm{hom}}^{(\mathrm{II})}`$ contained in $`P^{(\mathrm{II})}`$. We shall implement this method in practice below.
#### 2 Structure of the quadratic sources
In the context of the present paper, we will not need to compute the quadratic sources beyond the 1PN order. As a consequence, we will deal with only a few kinds of elementary sources. By equation (158), we already know the structure of $`V`$ at the 1PN approximation. The other compact retarded potentials have a very similar form. After expansion of the retardation of any of them, say $`P^{(\mathrm{C})}`$, we get:
$`P^{(\mathrm{C})}`$ $`=`$ $`\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}F\delta _1}{\displaystyle \frac{1}{c}}_t\mathrm{Pf}{\displaystyle \frac{d^3𝐱}{4\pi }F\delta _1}`$ (194)
$`+`$ $`{\displaystyle \frac{1}{2c^2}}_t^2\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|F\delta _1}+12+𝒪(3),`$ (195)
with $`F`$. The first contribution has been calculated in \[see equation (6.18) there\]. What is interesting for us is that the result writes as a sum of space derivative of $`1/r_1`$ (or $`1/r_2`$), i.e. $`_L(1/r_1)`$ (with the convention that $`L`$ designates a multi-index of length $`l`$). Similarly, it is easy to convince oneself that the third contribution is composed of terms $`_Lr_1`$ (or $`_Lr_2`$). Moreover, the action of time derivatives in front of the integral leaves the latter structure unchanged, in accordance with formulas such as $`_t_Lr_1=v_1^i_i_Lr_1`$. The second contribution in (195) is a mere constant with respect to $`𝐱`$. In fact, as they appear in the quadratic sources, the compact potentials are preceded by some space or time derivatives. Now, these derivations have to be performed in the sense of pseudo-functions . From these considerations, we are now in a position to tell what is the precise structure of the sources of quadratic potentials. They read as a sum of what we shall call elementary terms. As we are interested here in their spatial behaviour only, we shall omit purely time dependent factors, though they are normally included. Newtonian elementary terms are themselves products of two pseudo-function derivatives of contributions coming from the first integral in the generic expression (195): $`_J\mathrm{Pf}_K(1/r_1)\times _L\mathrm{Pf}_M(1/r_1)`$, or $`_J\mathrm{Pf}_K(1/r_1)\times _L\mathrm{Pf}_M(1/r_2)`$ (and similarly with $`12`$), where $`J`$, $`K`$, $`L`$, $`M`$ are multi-indices of respective length $`j`$, $`k`$, $`l`$, $`m`$. In the same manner, the 1PN terms result from products of pseudo-function derivatives of Newtonian and post-Newtonian integrals as the first and the third ones in (195): $`_J\mathrm{Pf}_K(1/r_1)\times _L\mathrm{Pf}_Mr_1`$, or $`_J\mathrm{Pf}_K(1/r_1)\times _L\mathrm{Pf}_Mr_2`$, and $`12`$. As for the 0.5PN terms, they are simply the pseudo-function derivatives of $`_L(1/r_1)`$ or $`_L(1/r_2)`$ (times a mere function of time). It is also natural to distinguish between the “self” elementary terms on one side, which depend on one body only, e.g. $`_i\mathrm{Pf}(1/r_1)\times _j\mathrm{Pf}(1/r_1)`$, and always admit pre-factors $`\mu _1^2`$, $`\stackrel{~}{\mu }_1^2`$, $`\stackrel{~}{\mu }_1\mu _1`$, and the “interaction” terms on the other side, involving both objects, e.g. $`_i\mathrm{Pf}(1/r_1)\times _j\mathrm{Pf}(1/r_2)`$. The 0.5PN terms $`_L(1/r_{1,2})`$ are considered separately.
To be more explicit, we shall provide as an example the 1PN source of $`\widehat{W}_{ij}^{(VV)}`$ defined by (173):
$`\mathrm{}\widehat{W}_{ij}^{(VV)}=`$ $``$ $`G^2\stackrel{~}{\mu }_1^2_i\mathrm{Pf}{\displaystyle \frac{1}{r_1}}_j\mathrm{Pf}{\displaystyle \frac{1}{r_1}}+{\displaystyle \frac{G^2m_1^2}{c^2}}\left(a_1^k_{(i}\mathrm{Pf}{\displaystyle \frac{1}{r_1}}_{j)k}\mathrm{Pf}r_1v_1^kv_1^l_{(i}\mathrm{Pf}{\displaystyle \frac{1}{r_1}}_{j)kl}\mathrm{Pf}r_1\right)`$ (196)
$``$ $`G^2\stackrel{~}{\mu }_1\stackrel{~}{\mu }_2_i\mathrm{Pf}{\displaystyle \frac{1}{r_1}}_j\mathrm{Pf}{\displaystyle \frac{1}{r_2}}`$ (197)
$`+`$ $`{\displaystyle \frac{G^2m_1m_2}{c^2}}\left(a_1^k_{(i}\mathrm{Pf}{\displaystyle \frac{1}{r_2}}_{j)k}\mathrm{Pf}r_1v_1^kv_1^l_{(i}\mathrm{Pf}{\displaystyle \frac{1}{r_2}}_{j)kl}\mathrm{Pf}r_1\right)+12+𝒪(3).`$ (198)
Here, we have used the fact that $`\mathrm{Pf}_ir_1=_i\mathrm{Pf}r_1`$ and $`\mathrm{Pf}_{ij}r_1=_{ij}\mathrm{Pf}r_1`$.
The sum of the retarded integral of the elementary terms then gives us the complete quadratic potentials after expansion in $`1/c`$ and matching. Therefore, these potentials are generated by the sources through some partie-finie integrals, which can be regarded as the result of the action of the elementary terms, considered as pseudo-functions, on smooth quantities in the field point. By inspection, it can be shown that the distributional part of the self terms never contributes to the previous integrals, whereas the partie-finie derivatives applied to the interaction terms coincide with those of the Schwartz distribution theory.
#### 3 Integration of the elementary sources
We now come to the solving of $`\mathrm{}P=S`$ at the 1PN order for each of the elementary terms composing the quadratic sources. We proceed following the method we exposed at the end of paragraph IV C 1. For this purpose, we first need to find a particular solution of the following Poisson equations:
$`\mathrm{\Delta }P_{\mathrm{part}}^{(\mathrm{I})}`$ $`=`$ $`_L\mathrm{Pf}{\displaystyle \frac{1}{r_1}}_K\mathrm{Pf}{\displaystyle \frac{1}{r_1}},\mathrm{\Delta }P_{\mathrm{part}}^{(\mathrm{I})}=_L\mathrm{Pf}r_1_K\mathrm{Pf}{\displaystyle \frac{1}{r_1}},`$ (200)
$`\mathrm{and}\mathrm{\Delta }P_{\mathrm{part}}^{(\mathrm{I})}`$ $`=`$ $`_L\mathrm{Pf}{\displaystyle \frac{1}{r_1}}_K\mathrm{Pf}{\displaystyle \frac{1}{r_2}},\mathrm{\Delta }P_{\mathrm{part}}^{(\mathrm{I})}=_L\mathrm{Pf}r_1_K\mathrm{Pf}{\displaystyle \frac{1}{r_2}},`$ (201)
with $`L=i_1\mathrm{}i_l`$ and $`K=j_1\mathrm{}j_k`$. From $`P_{\mathrm{part}}^{(\mathrm{I})}`$, we deduce the matched value $`P^{(\mathrm{I})}`$ by computing $`h_{\mathrm{hom}}^{(\mathrm{I})}`$ according to the relation (192) adapted for each elementary terms.
The equation (200) involves only the vector variable $`𝐫_1`$, so that it is simple enough to be integrable in a systematic way. To put the sources into a more suitable form, we start by applying the derivative operator that enters the self terms in the sense of functions, since the purely distributional part of the derivative does not contribute. The result is an adequate power of $`r_1`$ times a finite sum of partial terms $`\delta ^{i_1i_2}\mathrm{}\delta ^{i_{2k1}i_{2k}}n_1^{i_{2k+1}}\mathrm{}n_1^{i_l}`$, we shall denote more compactly as $`\delta ^{2K}n_1^{L2K}`$. The solving of the Poisson equations rests then on the well-known identities (easily checked by direct calculation):
$`r_1^a\widehat{n}_1^L`$ $`=`$ $`\mathrm{\Delta }\left\{{\displaystyle \frac{r_1^{a+2}\widehat{n}_1^L}{(al+2)(a+l+3)}}\right\}\mathrm{for}a\{l2,l3\},`$ (202)
$`r_1^{l2}\widehat{n}_1^L`$ $`=`$ $`\mathrm{\Delta }\left\{{\displaystyle \frac{1}{2l+1}}\left[\mathrm{ln}\left({\displaystyle \frac{r_1}{r_{\mathrm{1\hspace{0.17em}0}}}}\right){\displaystyle \frac{1}{2l+1}}\right]r_1^l\widehat{n}_1^L\right\},`$ (203)
$`r_1^{l3}\widehat{n}_1^L`$ $`=`$ $`\mathrm{\Delta }\left\{{\displaystyle \frac{1}{2l+1}}\left[\mathrm{ln}\left({\displaystyle \frac{r_1}{r_{\mathrm{1\hspace{0.17em}0}}}}\right)+{\displaystyle \frac{1}{2l+1}}\right]{\displaystyle \frac{\widehat{n}_1^L}{r_1^{l+1}}}\right\},`$ (204)
where $`\widehat{n}_1^L`$ is the trace-free part of $`n_1^L`$, and $`r_{\mathrm{1\hspace{0.17em}0}}`$ a strictly positive constant. The quantities between braces are particular solutions $`P_{\mathrm{part}}^{(\mathrm{I})}`$ of $`\mathrm{\Delta }P^{(\mathrm{I})}=r_1^a\widehat{n}_1^L`$, and we must in general add to them some harmonic functions to be evaluated by matching to the external field. Equations (201) are *a priori* the most difficult ones, because of the mixing of the sources 1 and 2. As a matter of fact, determining $`P_{\mathrm{part}}^{(\mathrm{I})}`$ amounts to solving:
$$\mathrm{\Delta }g=\frac{1}{r_1r_2}\mathrm{\Delta }f^{12}=\frac{r_1}{r_2},$$
(205)
in the sense of distributions, on account of the fact that, for instance,
$$_L\mathrm{Pf}\frac{1}{r_1}_K\mathrm{Pf}\frac{1}{r_2}=()^{k+l}_{1L}_{2K}\mathrm{Pf}\frac{1}{r_1r_2},$$
where $`_{1L}`$ and $`_{2K}`$ denote the partial derivatives with respect to $`𝐲_1`$ and $`𝐲_2`$ \[the same transformation applies to $`_L\mathrm{Pf}r_1\times _K\mathrm{Pf}(1/r_2)`$\]. As a consequence, $`{}_{L}{}^{}g_{K}^{}_{1L}_{2K}g`$ and $`{}_{L}{}^{}f_{K}^{12}_{1L}_{2K}f^{12}`$ clearly verify:
$$\mathrm{\Delta }\left[()^{k+l}{}_{L}{}^{}g_{K}^{}\right]=_L\mathrm{Pf}\frac{1}{r_1}_K\mathrm{Pf}\frac{1}{r_2}\mathrm{and}\mathrm{\Delta }\left[()^{k+l}{}_{L}{}^{}f_{K}^{12}\right]=_L\mathrm{Pf}r_1_K\mathrm{Pf}\frac{1}{r_2}.$$
(206)
Note that the derivatives above should be understood as mere (Schwartz) distributional derivatives. Luckily, particular solutions of equations (205) in the whole space can be exhibited . We may take:
$`g`$ $`=`$ $`\mathrm{ln}S,\mathrm{with}S=r_1+r_2+r_{12},`$ (208)
$`f^{12}`$ $`=`$ $`{\displaystyle \frac{1}{3}}r_1r_{12}(n_1n_{12})\left(g{\displaystyle \frac{1}{3}}\right)+{\displaystyle \frac{1}{6}}(r_2r_{12}+r_1r_2r_1r_{12}),`$ (209)
where $`(n_1n_{12})`$ denotes the scalar product of Euclidean vectors. The function $`g`$ is symmetric in its three variables $`𝐱`$, $`𝐲_1`$ and $`𝐲_2`$, so that
$$\mathrm{\Delta }_1g=\frac{1}{r_1r_{12}}\mathrm{and}\mathrm{\Delta }_2g=\frac{1}{r_2r_{12}},$$
in the sense of distributions. For a more complete list of useful formulas, see . We have also the two identities:
$$\mathrm{\Delta }_1f^{12}=2g\mathrm{and}\mathrm{\Delta }_2f^{12}=\frac{r_{12}}{r_2}.$$
Their proof is straightforward and call for some simple relations permitting to express the scalar products $`(n_1n_2)`$, $`(n_1n_{12})`$ and $`(n_2n_{12})`$ by means of some fractions involving $`r_1`$, $`r_2`$ and $`r_{12}`$. These relations, given by (5.14) in , are very convenient in most of our computations. The function $`f^{12}`$ is obtained by exchanging $`𝐱`$ and $`𝐲_1`$ in the function $`f`$ which was introduced in the appendix B of ,
$$f=\frac{1}{3}r_1r_2(n_1n_2)\left(g\frac{1}{3}\right)+\frac{1}{6}(r_1r_{12}+r_2r_{12}r_1r_2),$$
(210)
and which satisfies, in the sense of distributions, the equation
$$\mathrm{\Delta }f=2g.$$
Once Poisson equations with 1PN source are integrated, it remains to find the homogeneous solutions to be added to get the full matched solution. Most of the self terms are already correct, namely those that go to zero when $`r+\mathrm{}`$. The other ones are determined from the interaction terms by taking the limit $`𝐲_2𝐲_1`$, which happens to be always possible. The matching formula (192) provides the function $`h_{\mathrm{hom}}^{(\mathrm{I})}`$ associated to $`P_{\mathrm{part}}^{(\mathrm{I})}=g`$. The computation is very easy because the dimension of the source is $`[1/(r_1r_2)]=(\mathrm{length})^2`$ (i.e. $`d=2`$ in the notation of the end of paragraph IV C 1), therefore one needs to control only the constant term in the right side of (192) when $`𝐲_{1,2}0`$. We arrive at $`h_{\mathrm{hom}}^{(\mathrm{I})}=\mathrm{ln}(2r_0)1`$ hence the correct $`g_{\mathrm{match}}`$ solution of $`\mathrm{\Delta }g_{\mathrm{match}}=1/r_1r_2`$:
$$g_{\mathrm{match}}=\mathrm{ln}\left(\frac{S}{2r_0}\right)1,$$
where $`r_0`$ is the positive constant occurring in (174). Similarly, but with a little more work because the dimensionality of the source is now $`d=0`$ so we must expand to second order in $`𝐲_{1,2}`$, we obtain the matched value corresponding to $`f^{12}`$ as
$$f_{\mathrm{match}}^{12}=\frac{1}{3}r_1r_{12}(n_1n_{12})\left(g_{\mathrm{match}}\frac{1}{3}\right)+\frac{1}{6}(r_2r_{12}+r_1r_2r_1r_{12})\frac{1}{6}r(ny_1)\frac{1}{6}(y_1y_2)+\frac{1}{2}r(ny_2)$$
(where $`(ny_1)`$ for instance denotes the scalar product of $`𝐧=𝐱/r`$ with $`𝐲_1`$). As a consequence, the potentials $`P^{(\mathrm{I})}`$ satisfying (201) are respectively $`()^{l+k}{}_{L}{}^{}g_{\mathrm{match}K}^{}`$ and $`()^{l+k}{}_{L}{}^{}f_{\mathrm{match}K}^{12}`$. With this result in hands, we are able to deduce very simply all the self terms that do not match properly yet. We shall content ourselves with examining how this works on an example. Let us suppose we want to solve:
$$\mathrm{\Delta }P^{(\mathrm{I})}=r_1_{ij}\mathrm{Pf}\frac{1}{r_1}.$$
(211)
We make the correspondence with the equation $`\mathrm{\Delta }f_{\mathrm{match}ij}^{12}=r_1_{ij}\mathrm{Pf}(1/r_2)`$, whose source coincides with $`r_1_{ij}\mathrm{Pf}(1/r_1)`$ for $`𝐲_1=𝐲_2`$ (we recall our notation $`f_{\mathrm{match}ij}^{12}=_{2ij}f_{\mathrm{match}}^{12}`$). The distributional part of the derivative yields a compact supported contribution to $`f_{\mathrm{match}ij}^{12}`$ given by
$$<\mathrm{Pf}r_1\left(\frac{4\pi }{3}\right)\delta ^{ij}\delta _2,\frac{1}{4\pi }\frac{1}{|𝐱𝐱^{}|}>=\frac{r_{12}}{3r_2}\delta ^{ij},$$
which is zero in the limit $`𝐲_2𝐲_1`$, while the ordinary part yields a Poisson integral which is well-defined and easily evaluated for $`𝐲_1=𝐲_2`$. We conclude that the value of $`f_{\mathrm{match}ij}^{12}`$ when $`𝐲_2𝐲_1`$,
$$P^{(\mathrm{I})}=(f_{\mathrm{match}ij}^{12})_{𝐲_2𝐲_1}=\frac{1}{6}\delta ^{ij}\frac{1}{2}n_1^in_1^j,$$
*is* precisely the matched solution of the Poisson equation (211).
Let us complete now the program presented at the end of the paragraph IV C 1. Because of the presence of time derivatives at the 1PN order, we restore in the elementary terms all the coefficients depending only on time, either through the trajectories $`𝐲_{1,2}`$ or the velocities $`𝐯_{1,2}`$ \[we shall generically call $`\alpha (t)`$ this time-dependent coefficient; for instance $`\alpha =(v_1v_2)`$\]. It is worth noting that the potentials assimilated to $`P^{(\mathrm{II})}`$ in (189) are needed only at the Newtonian order, as they come with a $`1/c^2`$ factor. Consequently, the sources of the $`P^{(\mathrm{II})}`$’s are simply (two times) the matched solutions of the Poisson equations $`\mathrm{\Delta }P^{(\mathrm{I})}=S^\mathrm{N}+𝒪(1)`$, where $`S^\mathrm{N}`$ are the Newtonian-type sources $`\alpha (t)_L\mathrm{Pf}(1/r_{1,2})\times _K\mathrm{Pf}(1/r_{1,2})`$; so, all we have to solve is:
$`\mathrm{\Delta }P^{(\mathrm{II})}`$ $`=`$ $`2\alpha (t)r_1^p\delta ^{2K}n_1^{L2K},`$ (213)
$`\text{or}\mathrm{\Delta }P^{(\mathrm{II})}`$ $`=`$ $`2\alpha (t){}_{L}{}^{}g_{\mathrm{match}K}^{},`$ (214)
with $`p`$, and $`L`$, $`K`$ some multi-indices. The elementary self potentials obeying (213) are evaluated by application of the identities (204), before matching the full term (i.e. included the $`\alpha `$ coefficients) to the external field. The latter stage will be dropped here, because, on one hand, the general procedure has been explained before, and, on the other hand, powerful methods permitting to deal with the trickiest integrals one could encounter here will be expounded in Section V. Let us look next at the second equation (214). To get $`P^{(\mathrm{II})}`$ from some particular $`P_{\mathrm{part}}^{(\mathrm{II})}`$ one must perform the complete matching including all the time-dependent factors $`\alpha (t)`$. Here, for simplicity’s sake, we give the result in the case where the (Newtonian) source is $`1/(r_1r_2)`$ hence $`P^{(\mathrm{I})}=g_{\mathrm{match}}`$ as we have seen before. Then, we have to find the matched solution of
$$\mathrm{\Delta }P^{(\mathrm{II})}=2g_{\mathrm{match}}.$$
A particular solution $`P_{\mathrm{part}}^{(\mathrm{II})}`$ of this equation is easily obtained with the help of the function $`f`$ defined by (210) \[indeed, the Laplacian of $`P_{\mathrm{part}}^{(\mathrm{II})}f`$ is a mere constant\]. The corresponding homogeneous solution $`h_{\mathrm{hom}}^{(\mathrm{II})}`$ is computed using the same equation as (192) but using the source $`2P^{(\mathrm{I})}2g_{\mathrm{match}}`$. The result, that we naturally call $`f_{\mathrm{match}}P^{(\mathrm{II})}`$, reads as
$`f_{\mathrm{match}}`$ $`=`$ $`{\displaystyle \frac{1}{3}}r_1r_2(n_1n_2)\left(g_{\mathrm{match}}{\displaystyle \frac{1}{3}}\right){\displaystyle \frac{1}{6}}(r_1r_{12}+r_2r_{12}r_1r_2)`$ (215)
$``$ $`{\displaystyle \frac{1}{6}}r(ny_1){\displaystyle \frac{1}{6}}r(ny_2)+{\displaystyle \frac{1}{2}}(y_1y_2).`$ (216)
Notice that in the case of the source $`1/(r_1r_2)`$ the only “odd” contribution $`1/c`$ in the formula (190) is that given by the function $`𝒫(t)`$ defined by (191); the contribution due to $`(t)`$ is of higher order in this case. We readily find
$$𝒫(t)=\mathrm{FP}_{B0}d^3𝐱\left(\frac{r}{r_0}\right)^B\frac{1}{r_1r_2}=2\pi r_{12}$$
(217)
(no need of the symbol $`\mathrm{Pf}`$). This calculation is also done in equation (5.8) of . In Section V, we shall see more generally how such integrals can be obtained. Thus, our definitions of $`g_{\mathrm{match}}`$ and $`f_{\mathrm{match}}`$ are such that
$$\left(\mathrm{}_{}^1\frac{1}{r_1r_2}\right)_{\mathrm{match}}=g_{\mathrm{match}}\frac{1}{2c}\dot{r}_{12}+\frac{1}{2c^2}_t^2f_{\mathrm{match}}+𝒪(3).$$
(218)
Finally, we have all the material to integrate the individual post-Newtonian terms in such a way that the inner metric matches to the external field at the 3PN order. Let us remark however that, in fact, the work we have done on the matching is, as seen a posteriori, unnecessary. Indeed, summing up all the contributions in the potentials, we find that, had we made use of some “un-matched” elementary functions, e.g. $`g`$ and $`f^{12}`$ defined by (IV C 3), to compute the interaction terms instead of the corresponding matched quantities, and had we deduced jointly the corresponding self terms from the limit $`𝐲_2𝐲_1`$, we would have arrived at the same potentials up to the 3PN order. This means that the new contributions brought about by the matching to the external field actually cancel out in the final 3PN equations of motion. In particular, the constant $`r_0`$ which enters into the matched quantities $`g_{\mathrm{match}}`$, $`f_{\mathrm{match}}^{12}`$ and $`f_{\mathrm{match}}`$ disappears from the final result. Though we have verified this, we stick to our presentation and use systematically all the matched functions determined previously.
To end this section, we shall achieve the example of the potential $`\widehat{W}_{ij}^{(VV)}`$ defined by (173). We indeed already know its source from (196). We split the potential itself into:
$$\widehat{W}_{ij}^{(VV)}=U_{ij}\frac{1}{c^2}K_{ij}+\frac{1}{c}L_{ij}\frac{1}{2c^2}_t^2X_{ij}.$$
(219)
The first two contributions are respectively the matched solutions of the Poisson equations
$$\mathrm{\Delta }U_{ij}=_iU_jU\mathrm{and}\mathrm{\Delta }K_{ij}=_{(i}U_{j)}_t^2X,$$
which come from $`_iU_jU+\frac{1}{c^2}_{(i}U_{j)}_t^2X+𝒪(3)=`$ $`_iV_jV`$. Recall that the potentials $`U`$, $`X`$ and so on have to be viewed as pseudo-functions (for instance $`U=\mathrm{Pf}G(\stackrel{~}{\mu }_1)__1/r_1+12`$), so the derivatives entering the source terms are distributional derivatives. The self terms can be determined with the help of the relations (204) and matching. To get the interaction part, we change the spatial derivatives to partial derivatives with respect to the source points $`𝐲_{1,2}`$, and next, we make the replacement $`1/(r_1r_2)g_{\mathrm{match}}`$, $`r_1/r_2f_{\mathrm{match}}^{12}`$, and $`r_2/r_1f_{\mathrm{match}}^{21}`$. The “odd” term $`L_{ij}`$ is a pure function of time given by
$$L_{ij}=_t\frac{d^3𝐱}{4\pi }_iU_jU,$$
which is already known from equation (5.9) in ; it can also be computed with the methods of Section V. The contribution $`X_{ij}`$ is the matched solution of the double-Poisson equation
$$\mathrm{\Delta }^2X_{ij}=2_iU_jU.$$
whose source is to be considered at the Newtonian order only. The iterative application of (204) plus matching yields the self terms; for interaction terms, we replace $`_L(1/r_1)_K(1/r_2)`$ by $`()^{l+k}{}_{L}{}^{}f_{\mathrm{match}K}^{}`$. The results are
$`U_{ij}`$ $`=`$ $`{\displaystyle \frac{G^2\stackrel{~}{\mu }_1^2}{8}}\left(_{ij}^2\mathrm{ln}r_1+{\displaystyle \frac{\delta ^{ij}}{r_1^2}}\right)+G^2\stackrel{~}{\mu }_1\stackrel{~}{\mu }_2{}_{i}{}^{}g_{\mathrm{match}j}^{},`$ (221)
$`K_{ij}`$ $`=`$ $`G^2\stackrel{~}{\mu }_1^2[{\displaystyle \frac{a_1^{(i}}{4}}_{j)}\mathrm{ln}r_1+{\displaystyle \frac{a_1^k}{8}}\delta ^{ij}_k\mathrm{ln}r_1{\displaystyle \frac{a_1^k}{48}}_{ijk}\left(r_1^2\mathrm{ln}r_1\right)+{\displaystyle \frac{v_1^2\delta ^{ij}}{16r_1^2}}+{\displaystyle \frac{v_1^iv_1^i}{8r_1^2}}{\displaystyle \frac{v_1^kv_1^l}{16}}\delta ^{ij}_{kl}^2\mathrm{ln}r_1`$ (223)
$`+{\displaystyle \frac{v_1^2}{16}}_{ij}\mathrm{ln}r_1+{\displaystyle \frac{v_1^kv_1^l}{96}}_{ijkl}\left(r_1^2\mathrm{ln}r_1\right)]`$
$`+`$ $`G^2\stackrel{~}{\mu }_1\stackrel{~}{\mu }_2\left[a_1^k{}_{k(i}{}^{}f_{\mathrm{match}j)}^{12}+v_1^kv_1^l{}_{kl(i}{}^{}f_{\mathrm{match}j)}^{12}\right]+12,`$ (224)
$`L_{ij}`$ $`=`$ $`G^2_t\left[\stackrel{~}{\mu }_1^2\mathrm{Pf}{\displaystyle \frac{d^3𝐱}{4\pi }_i\frac{1}{r_1}_j\frac{1}{r_1}}\right]+G^2_t\left[\stackrel{~}{\mu }_1\stackrel{~}{\mu }_2_{1i}_{2j}\mathrm{Pf}{\displaystyle \frac{d^3𝐱}{4\pi }\frac{1}{r_1r_2}}\right]+12+𝒪(2)`$ (225)
$`=`$ $`G^2\stackrel{~}{\mu }_1\stackrel{~}{\mu }_2_t_{1i}_{2j}{\displaystyle \frac{r_{12}}{2}}+12+𝒪(2),`$ (226)
$`X_{ij}`$ $`=`$ $`{\displaystyle \frac{G^2\stackrel{~}{\mu }_1^2}{4}}\left[{\displaystyle \frac{1}{6}}_{ij}\left(r_1^2\mathrm{ln}r_1\right)+\delta ^{ij}\mathrm{ln}r_1\right]+G^2\stackrel{~}{\mu }_1\stackrel{~}{\mu }_2{}_{i}{}^{}f_{\mathrm{match}j}^{}+12+\mathrm{const}.`$ (227)
In the last equation we do not write for simplicity a constant (associated with a function of type $``$) which is cancelled out by the time derivative $`_t^2`$ in front of that term. The self terms have been written in the form of some (ordinary) space derivatives in order to prepare the computation of the cubic sources.
## V Cubic potentials
### A methodological scheme
For methodological reasons, it is convenient to express all the cubic sources in a similar way, with the help of the same set of elementary integrals. The so-called “cubic-non-compact” term
$$\widehat{X}^{(\mathrm{CNC})}=\mathrm{}_{}^1\left\{\widehat{W}_{ij}^{(VV)}_{ij}V\right\},$$
(228)
which is part of the $`\widehat{X}`$-potential \[see (79)\], is a good example to understand the successive transformation operations we perform in practice. Furthermore, this cubic-non-compact term is the only one we need to compute at the relative 1PN order; all the other ones, which enter into $`\widehat{T}`$ and $`\widehat{Y}_i`$, are merely Newtonian. So the practical computation of (228) is the most difficult one we face at the 3PN approximation. Recall that $`\widehat{W}_{ij}^{(VV)}`$ was defined by (173). We start from the expression of the source of $`\widehat{X}^{(\mathrm{CNC})}`$ obtained by insertion of (158) and (219) into (228). We get:
$`\mathrm{}\widehat{X}^{(\mathrm{CNC})}=`$ $``$ $`U_{ij}_{ij}U+{\displaystyle \frac{1}{c}}L_{ij}_{ij}U{\displaystyle \frac{1}{c^2}}K_{ij}_{ij}U`$ (229)
$``$ $`{\displaystyle \frac{1}{2c^2}}\left[_t^2X_{ij}_{ij}U+U_{ij}_{ij}_t^2X\right]+𝒪(3),`$ (230)
using the notation introduced in (219). In the right side, the potentials are seen as pseudo-functions (involving a $`\mathrm{Pf}`$) and the derivatives are distributional. After carrying on the expansion of retardations up to the 1PN approximation, we find:
$`\widehat{X}^{(\mathrm{CNC})}`$ $`=`$ $`{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}U_{ij}_{ij}U}{\displaystyle \frac{1}{c}}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}L_{ij}_{ij}U}`$ (231)
$``$ $`{\displaystyle \frac{1}{c}}_t{\displaystyle \frac{d^3𝐱^{}}{4\pi }U_{ij}_{ij}U}+{\displaystyle \frac{1}{c^2}}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}K_{ij}_{ij}U}`$ (232)
$`+`$ $`{\displaystyle \frac{1}{2c^2}}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}_t^2X_{ij}_{ij}U}+{\displaystyle \frac{1}{2c^2}}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}U_{ij}_{ij}_t^2X}`$ (233)
$`+`$ $`{\displaystyle \frac{1}{2c^2}}_t^2{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|U_{ij}_{ij}U}+{\displaystyle \frac{1}{c^2}}_t{\displaystyle \frac{d^3𝐱^{}}{4\pi }L_{ij}_{ij}U}+𝒪(3).`$ (234)
We have checked explicitly that the sum of the integrals occuring in this formula yields an integral convergent at infinity when considering the regularized value of the gradient $`(_i\widehat{X}^{(\mathrm{CNC})})__1`$ which is the only thing required; thus we do not need to introduce a finite part at infinity (but of course the regularization $`\mathrm{Pf}`$ is needed to cure the point-particle singularities). The next step consists of replacing the potentials $`U`$, $`X`$, $`L_{ij}`$, $`K_{ij}`$ and $`X_{ij}`$ given by (IV C 3) above by their values at the field point $`𝐱𝐲_1`$ and $`𝐱𝐲_2`$. The spatial and time derivations appearing in each of the integrals of (234) are to be understood in the sense of pseudo-functions (see Section II). Consider, as an example, the term
$$\frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}K_{ij}_{ij}U.$$
Remind that $`K_{ij}`$ is given by (224). Let us multiply (224) by $`_{ij}U=_{ij}\mathrm{Pf}(G\stackrel{~}{\mu }_1/r_1+G\stackrel{~}{\mu }_2/r_2)`$ and develop the product. The result is made of a sum of terms of the type $`(1/r_1^2)\times _{ij}\mathrm{Pf}(1/r_1)`$, $`_k\mathrm{ln}r_1\times _{ij}\mathrm{Pf}(1/r_1)`$, $`_{ijkl}(r_2^2\mathrm{ln}r_2)\times _{ij}\mathrm{Pf}(1/r_1)`$, $`{}_{ikl}{}^{}(f_{\mathrm{match}}^{12})_{j}^{}\times _{ij}\mathrm{Pf}(1/r_1)`$, etc. Some of them are functions of $`𝐫_1`$ only; we call them “self” terms, whereas those depending on both $`𝐫_1`$ and $`𝐫_2`$ are called “interaction” terms.
### B Self terms
We agree on considering only the self terms (i) that are proportional to $`m_1^3`$ rather than $`m_2^3`$ (they are the same modulo the replacement $`12`$), (ii) that do contribute to the 1PN order at most. We leave aside the terms that are generated by $`L_{ij}`$, since their structure is especially simple and they are evaluated at the end of the section. By explicitly writing down all the sources, as done previously for $`K_{ij}_{ij}U`$, we can draw the complete list of intervening terms. There are three types of terms: the $`VVV`$-type concerns one kind of term only, i.e. $`1/r_1\times _i\mathrm{Pf}(1/r_1)\times _j\mathrm{Pf}(1/r_1)`$; the so-called $`𝒴`$-type refers to $`_L\mathrm{Pf}r_1^p\times _K\mathrm{Pf}r_1^q`$ terms, where $`p`$ and $`q`$ are positive or negative integers; the $`𝒩`$-type terms come as $`_L\mathrm{Pf}_M(r_1^p\mathrm{ln}r_1)\times _J\mathrm{Pf}_Kr_1^q`$ (the terms $`𝒴`$ and $`𝒩`$ are named after some integrals introduced below). There may exist contracted indices among the set of multi-indices. In particular, some terms involve a factor $`\mathrm{\Delta }\mathrm{Pf}(1/r_1)`$ and are thus purely compact supported. In fact all the terms can be split into compact and non-compact parts. The latter part is an ordinary function that we are able to calculate explicitly. The former is determined from the results of Sections VI-VIII in and depends on the pseudo-function derivative we use. We shall refer to it as the self partie-finie-derivative contribution to the potentials. If we take the term $`_{ijkl}(r_1^2\mathrm{ln}r_1)\times _{ij}\mathrm{Pf}(1/r_1)`$ for instance, it reads:
$$_{ijkl}(r_1^2\mathrm{ln}r_1)_{ij}\mathrm{Pf}\frac{1}{r_1}=_{ijkl}(r_1^2\mathrm{ln}r_1)_{ij}\frac{1}{r_1}+_{ijkl}(r_1^2\mathrm{ln}r_1)\mathrm{D}_{ij}\left[\frac{1}{r_1}\right].$$
In the case of the “particular” derivative defined by (10), we have $`\mathrm{D}_{ij}^{\mathrm{part}}[1/r_1]=2\pi \mathrm{Pf}(\delta ^{ij}5n_1^in_1^j)\delta _1`$; so that
$$_{ijkl}(r_1^2\mathrm{ln}r_1)_{ij}\mathrm{Pf}\frac{1}{r_1}=12\frac{n_1^kn_1^l\delta ^{kl}}{r_1^5}+8\pi \mathrm{Pf}\frac{3n_1^kn_1^l+4\delta ^{kl}}{r_1^2}\delta _1$$
(235)
(like in the first term of the right side we sometimes do not write the $`\mathrm{Pf}`$ when there is no possible confusion). In most of this section, we shall use the particular derivative $`\mathrm{D}_{ij}^{\mathrm{part}}[F]`$ given by (10) instead of the more “correct” derivative $`\mathrm{D}_{ij}[F]`$ defined by (11)-(12); in Section VI we shall discuss the effect on the final 3PN equations of motion of using the derivative $`\mathrm{D}_{ij}[F]`$. In order to obtain the self cubic potentials, all we have to do now is to apply the operator $`\frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|^1`$ to the various sources we are focusing on, and $`_t^2\frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|`$ to the Newtonian source of $`\widehat{X}^{(\mathrm{CNC})}`$. As a matter of fact, the resulting integrals can be viewed as partie finie pseudo-functions like (235) acting on $`1/(4\pi |𝐱𝐱^{}|)`$ or $`|𝐱𝐱^{}|/(4\pi )`$; both quantities are smooth at point 1, so the pseudo-functions associated with the non-compact part reduce to Schwartz distributions in that case (but, in order to construct the pseudo-functions themselves, we used the generalized distributions of ). Each integral is indeed a sum of terms of the form $`r_1^p\delta ^{2K}n_1^{L2K}`$, where $`p`$ belongs to $``$. It is convenient to write them as sums of pseudo-function (or, equivalently here, distributional) derivatives of quantities without indices (“scalars”), times some possible Kronecker symbols. We have for example:
$`{\displaystyle \frac{n_1^in_1^j}{r_1^5}}`$ $`=`$ $`{\displaystyle \frac{1}{15}}_{ij}{\displaystyle \frac{1}{r_1^3}}+{\displaystyle \frac{1}{5}}{\displaystyle \frac{\delta ^{ij}}{r_1^5}}`$ (236)
$`=`$ $`{\displaystyle \frac{1}{15}}_{ij}\mathrm{Pf}{\displaystyle \frac{1}{r_1^3}}+{\displaystyle \frac{1}{5}}{\displaystyle \frac{\delta ^{ij}}{r_1^5}}+{\displaystyle \frac{4\pi }{15}}(8n_1^in_1^j\delta ^{ij}){\displaystyle \frac{\delta _1}{r_1^2}},`$ (237)
and other similar formulas for $`n_1^i/r_1^4`$, $`n_1^in_1^jn_1^k/r_1^6`$, etc.
As an illustration of our handling of the sources, here are the effects of these transformations on $`_{ijkl}(r_1^2\mathrm{ln}r_1)_{ij}\mathrm{Pf}(1/r_1)`$. Starting from equation (235) and using (237), we find
$`{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}_{ijkl}^{^{}}(r_1^{}_{}{}^{}2\mathrm{ln}r_1^{})_{ij}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}`$
$`={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\left[\frac{4}{5}_{kl}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}_{}{}^{}3}\frac{48}{5}\mathrm{Pf}\frac{\delta ^{kl}}{r_1^{}_{}{}^{}5}+\frac{8\pi }{5}\mathrm{Pf}\frac{n_1^{}_{}{}^{}kn_1^{}_{}{}^{}l18\delta ^{kl}}{r_1^{}_{}{}^{}2}\delta _1^{}\right]},`$
where the first term is generated by the specific derivative (10). In this term the derivative can be changed to a partial derivative with respect to the point 1, and since we employ a pseudo-function derivative, we are allowed to permute integration and derivation symbols. This yields
$`\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}_{ijkl}^{^{}}(r_1^{}_{}{}^{}2\mathrm{ln}r_1^{})_{ij}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}`$
$`={\displaystyle \frac{4}{5}}_{1kl}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}_{}{}^{}3}}{\displaystyle \frac{48}{5}}\delta ^{kl}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}_{}{}^{}5}}`$
$`+\left({\displaystyle \frac{2}{5}}{\displaystyle \frac{n_1^{}_{}{}^{}kn_1^{}_{}{}^{}l}{r_1^{}_{}{}^{}2|𝐱𝐱^{}|}}+{\displaystyle \frac{36}{5}}{\displaystyle \frac{\delta ^{kl}}{r_1^{}_{}{}^{}2|𝐱𝐱^{}|}}\right)_1.`$
The first two terms are left in this form for the time being. On the other hand the last term is computed following the procedure explained by the equations (6.17)-(6.18) in ; see also (272) below. By implementing the previous procedure for all the self terms entering $`K_{ij}_{ij}U`$, we finally arrive at:
$`\left({\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}K_{ij}_{ij}U\right)_{\mathrm{self}}=G^3m_1^3\{{\displaystyle \frac{a_1^i}{6}}_{1i}\mathrm{Pf}{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}{\displaystyle \frac{1}{r_1^{}_{}{}^{}3}}`$ (238)
$`+{\displaystyle \frac{v_1^iv_1^j}{30}}_{1ij}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}_{}{}^{}3}}{\displaystyle \frac{2v_1^2}{5}}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}_{}{}^{}5}}`$ (239)
$`{\displaystyle \frac{(n_1a_1)}{12r_1^2}}+{\displaystyle \frac{7(n_1v_1)^2}{150r_1^3}}{\displaystyle \frac{7v_1^2}{450r_1^3}}\}+12+𝒪(1).`$ (240)
We follow the same way to treat the self parts of the other cubic potentials of interest here: $`\widehat{T}^{(\mathrm{CNC})}`$, $`\widehat{Y}_i^{(\mathrm{CNC})}`$, or the remaining of $`\widehat{X}^{(\mathrm{CNC})}`$.
We find that there are definite contributions, coming at the 3PN order, due specifically to the pseudo-function derivative introduced in . Indeed, the distributional part of the derivative gives some well-defined non-zero contributions, while for instance the Schwartz derivative yields some terms which are ill-defined in this case. These contributions of the pseudo-function derivative actually take part in the values of $`\widehat{T}^{(\mathrm{CNC})}`$ and $`\widehat{X}^{(\mathrm{CNC})}`$ only. Denoting them by $`\delta _{\mathrm{self}}\widehat{T}`$ and $`\delta _{\mathrm{self}}\widehat{X}`$ in the case of the particular derivative (10) we find
$`\delta _{\mathrm{self}}\widehat{T}`$ $`=`$ $`{\displaystyle \frac{7}{12}}{\displaystyle \frac{G^4m_1^3m_2}{r_1^2r_{12}^2}}(n_1n_{12})+12`$ (242)
$`\delta _{\mathrm{self}}\widehat{X}`$ $`=`$ $`{\displaystyle \frac{G^3m_1^3}{c^2r_1^3}}\left({\displaystyle \frac{17}{72}}{\displaystyle \frac{Gm_2}{r_{12}^2}}r_1(n_1n_{12})+{\displaystyle \frac{1}{40}}(n_1v_1)^2{\displaystyle \frac{1}{120}}v_1^2\right)+12`$ (243)
### C Interaction terms
We consider exclusively the interaction terms (i) that are proportional to $`m_1^2m_2`$ rather than $`m_1m_2^2`$, (ii) that contribute at relative 1PN order, leaving aside those which are generated by $`L_{ij}`$. Depending on whether they come from “simple” or “composite” cubic parts as shown respectively below, the elementary terms composing the sources read schematically
$`\mathrm{Pf}[F(𝐫_1)]\mathrm{Pf}[G(𝐫_1)]\mathrm{Pf}[H(𝐫_2)]`$ (245)
$`\mathrm{and}`$ $`\mathrm{Pf}[F(𝐫_1)]_1_2\mathrm{Pf}[G(𝐫_1,𝐫_2)]=_2\mathrm{Pf}\{\mathrm{Pf}[F(𝐫_1)]_1\mathrm{Pf}[G(𝐫_1,𝐫_2)]\};`$ (246)
the functions $`G`$, $`H`$ belong to $``$, and it is also the case of $`F`$ in general. However, there exist some composite terms for which $`F=\mathrm{ln}r_1`$, but this is not a trouble since $`F`$ is still in $``$. In the cases needed in this problem, $`G`$ is always one of the four functions: $`g_{\mathrm{match}}`$, $`f_{\mathrm{match}}^{12}`$, $`f_{\mathrm{match}}^{21}`$, or $`f_{\mathrm{match}}`$. Then, $`G`$ is “regular enough” so that $`\mathrm{Pf}G`$ and $`\mathrm{Pf}G`$ coincide in any cases, and further simplifications of the sources do not seem to be possible at this level. All we need, thus, is to transform the simple cubic contributions (245) similarly to the self terms (see paragraph V B). A typical example of elementary source we have to handle is $`1/(r_1r_2)\times _L(1/r_1)`$, where $`L`$ represents at most two non-contracted indices. We can check that this term can be computed to Newtonian order only, hence it is given simply by a Poisson integral. In the language of pseudo-functions, this means that we have to evaluate:
$$<\frac{1}{r_1^{}r_2^{}}_L^{^{}}\mathrm{Pf}\frac{1}{r_1^{}},\frac{1}{4\pi |𝐱𝐱^{}|}>=<\frac{1}{r_1^{}}_L^{^{}}\mathrm{Pf}\frac{1}{r_1^{}},\frac{1}{4\pi |𝐱𝐱^{}|r_2^{}}>.$$
The compact part of the dual bracket, which is associated with the distributional part of the derivative, when acting on $`1/(4\pi |𝐱𝐱^{}|)`$, i.e.
$$<\frac{1}{r_1^{}}\mathrm{D}_L\left[\frac{1}{r_1^{}}\right],\frac{1}{4\pi |𝐱𝐱^{}|r_2^{}}>$$
($`l`$=1 or 2), leads *a priori* to a non-zero result. Here, $`\mathrm{D}_L`$ denotes the distributional part of the multi-derivative, obtained in Section VIII of and recalled by the equation (13) above. However, the left side of the bracket, which is homogeneous to the $`(l2)`$th power of a length, is necessarily of the type $`r_1^{1l}\mathrm{Pf}\delta _1`$, times some dimensionless angular function whose multipolarity differs from $`l`$ by an even integer (because of the index structure of the operator $`\mathrm{D}_L`$). Now, the previous compact part is equal to the angular integral of the $`r_1^{l1}`$ Taylor coefficient of $`1/(4\pi |𝐱𝐱^{}|)`$, times the angular dependence of $`\mathrm{D}_L[1/r_1]`$. The integrand then appears as a sum of terms whose multipolarity differs from $`l+(l1)=2l1`$ by an even integer and so, is always odd; thereby the angular integral gives zero. By similar arguments, we can prove that the other compact sources associated with the distributional derivatives will never contribute to the Poisson integrals constituting the potentials we are considering here. Actually, it is possible to put together all the various kinds of simple-type \[see (245)\] cubic terms into a unique one, which is $`n_1^L/r_1^{l+2}\times \mathrm{Pf}(1/r_2)`$. We express at last the first factor (if $`l0`$) as a sum of derivatives of “scalars” thanks to identities such as (237). Since the pseudo-function derivatives will give here the same results as the Schwartz distributional ones, and by virtue of $`_L\mathrm{Pf}(1/r_1^2)|_𝒟=\mathrm{Pf}_L(1/r_1^2)|_𝒟`$ \[where $`𝒟`$ is the set of smooth functions with compact support\], the last transformation can be done in the sense of functions. Note however that the multiple derivatives of $`1/r_2`$ are indeed distributional and play an important role in the sources.
To sum up what precedes, all the interaction terms have the general structure:
$$\frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|^p_{2L}\mathrm{Pf}[_J^{^{}}\mathrm{Pf}F(𝐫_1^{})_{1K}\mathrm{Pf}G(𝐫_1^{},𝐫_2^{})]$$
(247)
($`p=1`$ or 1, $`F`$ and $`G`$ functions of $``$). After commuting the integral and the derivative $`_{2L}`$, which is always licit for integrals converging at infinity (since $`_{2L}`$ is followed by a $`\mathrm{Pf}`$), the general cubic term (247) becomes:
$$_{2L}\frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|^p_J^{^{}}\mathrm{Pf}F(𝐫_1^{})_{1K}\mathrm{Pf}G(𝐫_1^{},𝐫_2^{}).$$
In the case where $`G=G(𝐫_2)`$ (and $`k=0`$), this rearranges as
$`_{2L}{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|^p()^j_{1J}[\mathrm{Pf}F(𝐫_1^{})\mathrm{Pf}G(𝐫_2^{})]}`$ (248)
$`=()^j_{1J}_{2L}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|^pF(𝐫_1^{})G(𝐫_2^{})}.`$ (249)
We shall end with implementing concretely our treatment of the source on two typical terms:
$`{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}}_i^{}\mathrm{Pf}\frac{1}{r_1^{}}_j^{}\mathrm{Pf}\frac{1}{r_2^{}}}={\displaystyle \frac{1}{2}}_{1i}_{2j}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}_{}{}^{}2r_2^{}}},`$ (251)
$`{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{ki}{}^{}f_{\mathrm{match}j}^{12}_{ij}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}=_{2j}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{ki}{}^{}f_{\mathrm{match}}^{12}_{ij}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}},`$ (252)
and by providing the complete interaction component corresponding to the Poisson integral of $`K_{ij}_{ij}U`$, which completes the self part obtained previously in (240). The compact support terms have been explicitly determined, while the other ones are left un-evaluated for the moment:
$`\left({\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}K_{ij}_{ij}U\right)_{\mathrm{int}}=G^3m_1^2m_2\{{\displaystyle \frac{a_1^i}{4}}_{2i}D\mathrm{Pf}{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}{\displaystyle \frac{\mathrm{ln}r_1^{}}{r_2^{}}}`$ (253)
$`+{\displaystyle \frac{a_1^i}{48}}_{1i}D^2\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{r_1^{}_{}{}^{}2\mathrm{ln}r_1^{}}{r_2^{}}}+a_1^j_{2i}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{jk}{}^{}f_{\mathrm{match}}^{12}_{ik}^{^{}}\frac{1}{r_1^{}}}`$ (254)
$`+a_2^j_{2ij}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{k}{}^{}f_{\mathrm{match}}^{21}_{ik}^{^{}}\frac{1}{r_1^{}}}{\displaystyle \frac{(n_{12}a_1)}{8r_{12}r_2}}`$ (255)
$`+{\displaystyle \frac{v_1^iv_1^j}{8}}_{2ij}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{1}{r_1^{}_{}{}^{}2r_2^{}}}+{\displaystyle \frac{v_1^2}{16}}D^2\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{\mathrm{ln}r_1^{}}{r_2^{}}}`$ (256)
$`+{\displaystyle \frac{v_1^iv_1^j}{96}}_{1ij}D^2\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}\frac{r_1^{}_{}{}^{}2\mathrm{ln}r_1^{}}{r_2^{}}}`$ (257)
$`+v_1^jv_1^k_{2i}\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{jkl}{}^{}f_{\mathrm{match}}^{12}_{il}^{^{}}\frac{1}{r_1^{}}}`$ (258)
$`+v_2^jv_2^k_{2ijk}\mathrm{Pf}{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}{}_{l}{}^{}f_{\mathrm{match}}^{21}_{il}^{^{}}{\displaystyle \frac{1}{r_1^{}}}+{\displaystyle \frac{(n_{12}v_1)^2}{8r_{12}^2r_2}}\}+12+𝒪(1),`$ (259)
where $`D`$ denotes the operator $`_{1i}_{2i}`$.
### D Elementary integrals
#### 1 Nomenclature
The inspection of the formula (259) for interaction terms issued from $`K_{ij}_{ij}U`$ suggests that we should re-express this potential, as well as all the other ones, by means of a restricted number of elementary integrals; basically one for each source type. Hence the proposal for a useful systematic nomenclature, which reflects their structure. We shall introduce the following notations (and ditto $`12`$):
$`\stackrel{(n,p)}{𝒴}=\mathrm{Pf}{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}r_1^{}_{}{}^{}nr_2^{}_{}{}^{}p\underset{1}{}_{(P,Q)}^{12}={\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}{}_{iQ}{}^{}f_{\mathrm{match}}^{12}_{iP}^{^{}}\mathrm{Pf}{\displaystyle \frac{1}{r_1^{}}}`$ (260)
$`\underset{1}{\overset{(n,p)}{𝒩}}=\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}r_1^{}_{}{}^{}nr_2^{}_{}{}^{}p\mathrm{ln}r_1^{}}\underset{1}{}_{(P,Q)}^{21}={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{iQ}{}^{}f_{\mathrm{match}}^{21}_{iP}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}`$ (261)
$`\underset{1}{𝒰}_P={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{k}{}^{}g_{\mathrm{match}\mathrm{k}}^{}_P^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}\underset{1}{}_{(P,Q)}={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{iQ}{}^{}f_{\mathrm{match}}^{}_{iP}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}`$ (262)
$`\underset{1}{𝒦}_P={\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}g_{\mathrm{match}}_P\mathrm{Pf}{\displaystyle \frac{1}{r_1^{}}}\stackrel{(n,p)}{𝒮}=\mathrm{Pf}{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}|𝐱𝐱^{}|r_1^{}_{}{}^{}nr_2^{}_{}{}^{}p`$ (263)
$`\underset{1}{𝒢}_{(P,Q)}={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{iQ}{}^{}g_{\mathrm{match}}^{}_{iP}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}\underset{1}{\overset{(n,p)}{}}=\mathrm{Pf}{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|r_1^{}_{}{}^{}nr_2^{}_{}{}^{}p\mathrm{ln}r_1^{}}`$ (264)
$`{}_{i}{}^{}\underset{1}{𝒢}_{(P,Q)}^{}={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{iQ}{}^{}g_{\mathrm{match}}^{}_P^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}\underset{1}{𝒬}_{(P,Q)}={\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|{}_{iQ}{}^{}g_{\mathrm{match}}^{}_{iP}^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}}`$ (265)
$`\underset{1}{}_{(P,Q)}={\displaystyle \frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{iQ}{}^{}g_{\mathrm{match}}^{}_{iP}^{^{}}\mathrm{Pf}r_1^{}}.`$ (266)
The value of the previous integrals is not generally known at any space location, except in some special cases. The reason is that their sources involve three points, in addition to the integration variable $`𝐱^{}`$: the point $`𝐱`$ where the field is calculated, and the two source points $`𝐲_1`$, $`𝐲_2`$. The few ones that are computable for any $`𝐱`$ include notably some self integrals like $`𝒴^{(n,0)}`$ and $`𝒮^{(n,0)}`$, and the two integrals $`D^2𝒩_1^{(0,1)}`$ and $`_{2i}𝒢_{1(i,0)}`$ entering the interaction part of the $`\widehat{X}^{(\mathrm{CNC})}`$ potential at the Newtonian order. Actually, there are no other cubic contributions up to the 2.5PN order, and that is why we were able in to get the complete expression of the metric in the near zone at this order. This property of the integrals $`D^2𝒩_1^{(0,1)}`$ and $`_{2i}𝒢_{1(i,0)}`$ is linked to the specific form of the integrands, which are made of products of two second derivatives applied on appropriate functions, such that the indices of the first derivative are contracted with those of the second one: $`_{ij}\mathrm{Pf}\mathrm{ln}r_1\times _{ij}\mathrm{Pf}(1/r_2)`$ and $`{}_{i}{}^{}g_{j}^{}\times _{ij}\mathrm{Pf}(1/r_1)`$. In both cases, particular solutions in the sense of distributions of the corresponding Poisson equations
$$\mathrm{\Delta }K_1=2_{ij}\mathrm{Pf}\mathrm{ln}r_1_{ij}\mathrm{Pf}\frac{1}{r_2},\mathrm{\Delta }H_1=2{}_{i}{}^{}g_{j}^{}_{ij}\mathrm{Pf}\frac{1}{r_1}.$$
(267)
can be exhibited . The solutions $`K_1`$ and $`H_1`$ of (267) that go to zero as $`r\mathrm{}`$ read:
$`K_1`$ $`=`$ $`\left({\displaystyle \frac{1}{2}}\mathrm{\Delta }\mathrm{\Delta }_1\right)\left[{\displaystyle \frac{\mathrm{ln}r_1}{r_2}}\right]+{\displaystyle \frac{1}{2}}\mathrm{\Delta }_2\left[{\displaystyle \frac{\mathrm{ln}r_{12}}{r_2}}\right]+{\displaystyle \frac{r_2}{2r_{12}^2r_1^2}}+{\displaystyle \frac{1}{r_{12}^2r_2}},`$ (269)
$`H_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Delta }_1\left[{\displaystyle \frac{g}{r_1}}+{\displaystyle \frac{\mathrm{ln}r_1}{r_{12}}}\mathrm{\Delta }_1\left({\displaystyle \frac{r_1+r_{12}}{2}}g\right)\right]`$ (270)
$`+`$ $`_i_{2i}\left[{\displaystyle \frac{\mathrm{ln}r_{12}}{r_1}}+{\displaystyle \frac{\mathrm{ln}r_1}{2r_{12}}}\right]{\displaystyle \frac{1}{r_1}}_{2i}[(_ig)__1]{\displaystyle \frac{r_2}{2r_1^2r_{12}^2}},`$ (271)
with $`\mathrm{\Delta }_1=_{1ii}`$, $`\mathrm{\Delta }_2=_{2ii}`$; see also (V E) for the expanded forms of these solutions. Thus, we have
$$D^2𝒩_1^{(0,1)}=\frac{1}{2}K_1\text{and}_{2i}𝒢_{1(i,0)}=\frac{1}{2}H_1.$$
In the list (260) above, we note the appearance of iterated Poisson integrals such as $`𝒮^{(5,0)}`$, $`^{(0,1)}`$ and $`_{2i}𝒬_{1(i,0)}`$ which come from the 1PN contribution to the retardation expansion of $`\widehat{X}^{(\mathrm{CNC})}`$, and thus enter this potential through their second time derivative. What we shall have to compute for our purpose is their Hadamard regularized value. However, $`_{2i}𝒬_{1(i,0)}`$ is not available on the whole space (for any $`𝐱`$), so that we cannot deduce $`_t^2_{2i}𝒬_{1(i,0)}`$ from it directly. We shall adopt then a different approach. In a first stage, we express the operator $`_t^2`$ with the help of the partial derivatives $`_{1i}`$ and $`_{2i}`$: if $`F(𝐫_1,𝐫_2)=_{2i}𝒬_{1(i,0)}`$, then
$$_t^2F=_t\left[v_1^i_{1i}F\right]=v_1^iv_1^j_{1ij}F+a_1^i_{1i}F+12.$$
Next, we commute the partial derivatives $`_{1L}`$ with the integration sign, so that the derivatives act on the source of the Poisson integral. This operation is legitimate only if the derivatives $`_{1L}`$ when acting on the integrands are viewed as distributional (in the sense of Section IX in ). The new integrands write then as a partie finie derivative of a product, e.g. $`_{1L}({}_{iQ}{}^{}g_{iP}^{^{}}\frac{1}{r_1^{}})`$; but remember that we are not *a priori* allowed to develop them according to the Leibniz rule in our formalism. In fact, these specific non Leibniz corrections happen to give zero contribution to the 3PN potentials. This can be seen by applying successively the formula (7.23) in (which is indeed sufficient since the “test” function $`|𝐱𝐱^{}|`$ is smooth at the points 1 and 2) to all the sources we are dealing with. Therefore, we can employ the usual rule for derivatives of products to perform our final transformation. In summary, we will have for instance
$`_{1kl}{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|{}_{i}{}^{}g_{j}^{}_{ij}\mathrm{Pf}\frac{1}{r_1^{}}}={\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|{}_{ikl}{}^{}g_{j}^{}_{ij}\mathrm{Pf}\frac{1}{r_1^{}}}`$
$`+2{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|{}_{i(k}{}^{}g_{\underset{¯}{j}}^{}_{l)ij}\mathrm{Pf}\frac{1}{r_1^{}}}+{\displaystyle \frac{d^3𝐱^{}}{4\pi }|𝐱𝐱^{}|{}_{i}{}^{}g_{j}^{}_{ijkl}\mathrm{Pf}\frac{1}{r_1^{}}},`$
(the index $`\underset{¯}{j}`$ means that $`j`$ is excluded from the symmetrization operation). On this form we can apply the partie finie at 1 while staying in the same class of elementary integrals (260). We conclude by going back to the example of the cubic term generated by $`K_{ij}_{ij}U`$; we have, after appropriate reshaping,
$`{\displaystyle }{\displaystyle \frac{d^3𝐱^{}}{4\pi }}{\displaystyle \frac{1}{|𝐱𝐱^{}|}}K_{ij}_{ij}U=G^3m_1^3\{{\displaystyle \frac{a_1^i}{6}}_{1i}\stackrel{(3,0)}{𝒴}+{\displaystyle \frac{v_1^iv_1^j}{30}}_{1ij}\stackrel{(3,0)}{𝒴}{\displaystyle \frac{2v_1^2}{5}}\stackrel{(5,0)}{𝒴}`$
$`{\displaystyle \frac{(n_1a_1)}{12r_1^2}}+{\displaystyle \frac{7(n_1v_1)^2}{150r_1^3}}{\displaystyle \frac{7v_1^2}{450r_1^3}}\}`$
$`+G^3m_1^2m_2\{{\displaystyle \frac{a_1^i}{4}}_{2i}D\underset{1}{\overset{(0,1)}{𝒩}}+{\displaystyle \frac{a_1^i}{48}}_{1i}D^2\underset{1}{\overset{(2,1)}{𝒩}}+a_1^j_{2i}\underset{1}{}_{(i,j)}^{12}+a_2^j_{2ij}\underset{1}{}_{(i,0)}^{21}`$
$`{\displaystyle \frac{(n_{12}a_1)}{8r_{12}r_2}}+{\displaystyle \frac{v_1^iv_1^j}{8}}_{2ij}\stackrel{(2,1)}{𝒴}+{\displaystyle \frac{v_1^2}{16}}D^2\underset{1}{\overset{(0,1)}{𝒩}}+{\displaystyle \frac{v_1^iv_1^j}{96}}_{1ij}D^2\underset{1}{\overset{(2,1)}{𝒩}}`$
$`+v_1^jv_1^k_{2i}\underset{1}{}_{(i,jk)}^{12}+v_2^jv_2^k_{2ijk}\underset{1}{}_{(i,0)}^{21}+{\displaystyle \frac{(n_{12}v_1)^2}{8r_{12}^2r_2}}\}+12+𝒪(1).`$
We refer to for the expressions of the other non-linear potentials expressed in this manner by means of the same nomenclature. The problem is now to evaluate all the elementary integrals from which the 3PN cubic potentials have been built.
#### 2 Parties finies of the elementary integrals
As mentioned before, in most cases, we do not have at our disposal the explicit values of the elementary integrals in all space. This does not matter since all we need is their Hadamard partie finie at point 1 (or 2). Notice that the partial derivative with respect to $`𝐲_2`$ is the only one that commutes with the partie finie operation at 1; to be more explicit:
$`[_{2i}F(𝐱,𝐲_1,𝐲_2)]__1=_{2i}[F(𝐱,𝐲_1,𝐲_2)]__1,`$
$`\mathrm{but}`$ $`[_{1i}F(𝐱,𝐲_1,𝐲_2)]__1_{1i}[F(𝐱,𝐲_1,𝐲_2)]__1(\mathrm{with}F,G).`$
Thus, for each elementary integral, we shall determine first the partie finie of the quantity figuring under the derivation symbol $`_2`$ and, only then, apply the latter operator. On the contrary, we cannot bring the derivatives with respect to $`𝐲_1`$ out of the partie finie at 1, so we are led to incorporate them to the sources, in the sense of pseudo-functions, by permutation with the integration sign. As a consequence, the integrals we are interested in are of the type:
$`{\displaystyle \frac{d^3𝐱}{4\pi }|𝐱𝐱^{}|^p_1[\mathrm{Pf}F(𝐫_1)]_1[\mathrm{Pf}G(𝐫_1,𝐫_2)]}`$
$`={\displaystyle \frac{1}{4\pi }}<_1[\mathrm{Pf}F(𝐫_1)]_1[\mathrm{Pf}G(𝐫_1,𝐫_2)],|𝐱𝐱^{}|^p>.`$
They involve both a compact part (C) and a non-compact part (NC). The compact part is produced by the purely distributional contributions of derivatives in the integrands:
$$\frac{1}{4\pi }<\mathrm{D}_1[F(𝐫_1)]\mathrm{Pf}_1G(𝐫_1,𝐫_2),|𝐱𝐱^{}|^p>\frac{1}{4\pi }<\mathrm{D}_1[G(𝐫_1,𝐫_2)]\mathrm{Pf}_1F(𝐫_1),|𝐱𝐱^{}|^p>.$$
As mentioned at the beginning of the paragraph V C, the partie finie derivatives reduce here, in our case, to those of the Schwartz theory. This is obvious when $`G=G(𝐫_2)`$, because then the source $`F(𝐫_1)`$, regarded as a linear functional $`\mathrm{Pf}F(𝐫_1)`$, acts on a function which is smooth in a neighbourhood of $`𝐱=𝐲_1`$; in the other cases, the result follows from explicit calculations. Note that the number $`l`$ of derivatives in front of $`G(𝐫_1,𝐫_2)=g_{\mathrm{match}}`$, $`f_{\mathrm{match}}^{12}`$, $`f_{\mathrm{match}}^{21}`$, or $`f_{\mathrm{match}}`$ is always small enough so that $`_{1L1}G(𝐫_1,𝐫_2)`$ is bounded, hence $`_{1L}\mathrm{Pf}G(𝐫_1,𝐫_2)=\mathrm{Pf}_{1L}G(𝐫_1,𝐫_2)`$. Once we have in hands the compact part, it remains to obtain its regularized value at point 1. As a matter of fact, if the source is of the type $`\mathrm{Pf}F\delta _2`$, where $`F`$, then
$$<\mathrm{Pf}(F\delta _2)(𝐱^{}),\frac{1}{|𝐱𝐱^{}|}>=\frac{1}{4\pi }\underset{l0}{}\frac{()^l}{l!}\left(r_2^{}_{}{}^{}ln_2^{}_{}{}^{}LF\right)_2_L\left(\frac{1}{r_2}\right)$$
(272)
is smooth at point 1 and we need not call for the Hadamard regularization. Therefore, we are allowed to replace directly $`|𝐱𝐱^{}|`$ by $`r_1^{}`$ in the left-hand-side of (272). When the source is of the generic type $`\mathrm{Pf}F\delta _1`$, the same identity as (272) holds, but with $`21`$. This shows that the integral $`<\mathrm{Pf}(F\delta _1),|𝐱𝐱^{}|^1>`$ is purely singular as $`r_10`$, which means that it has no partie finie at $`𝐱=𝐲_1`$: the $`\delta _1`$ type sources do not contribute to the potentials computed at body 1. Summarizing, we have
$`(<\mathrm{Pf}(F\delta _2),{\displaystyle \frac{1}{|𝐱𝐱^{}|}}>)_1`$ $`=`$ $`<\mathrm{Pf}(F\delta _2),{\displaystyle \frac{1}{r_1}}>,`$ (274)
$`(<\mathrm{Pf}(F\delta _1),{\displaystyle \frac{1}{|𝐱𝐱^{}|}}>)_1`$ $`=`$ $`0.`$ (275)
We refer the reader to equations (6.17)-(6.20) in for more details.
Let us now focus our attention on the non-compact parts of the elementary integrals, whose integrands are made, by definition, of ordinary functions. The problem is to get the Hadamard partie finie of the Poisson integral $`P`$ of $`F`$:
$$P(𝐱^{})=\mathrm{Pf}\frac{d^3𝐱}{4\pi }\frac{1}{|𝐱𝐱^{}|}F(𝐱),$$
as well as the one of the iterated Poisson integral $`Q`$, given by:
$$Q(𝐱^{})=\mathrm{Pf}\frac{d^3𝐱}{4\pi }|𝐱𝐱^{}|F(𝐱).$$
\[Each of these functions depends also on the source points $`𝐲_{1,2}`$ and velocities $`𝐯_{1,2}`$.\] Now, the partie finie prescription applies only to functions admitting power-like expansions near their singularities \[see (1)\], whereas $`P`$ or $`Q`$ may contain logarithmic coefficients in their development: if we take for instance $`F=1/r_1^3`$, we shall have $`P=[1+\mathrm{ln}(r_1^{}/s_1)]/r_1^{}`$, where $`s_1`$ is the constant appearing in (4). Following , we shall simply include the possible logarithms (i.e. $`\mathrm{ln}r_1^{}`$) appearing in the zeroth power coefficient of the expansion of $`P`$ or $`Q`$ in the definition of the partie finie: see the equation (5.4) in . With this generalized notion of “partie finie”, we can give a sense to $`(P)__1`$ and $`(Q)__1`$, as well as their gradients $`(_iP)__1`$ and $`(_iQ)__1`$. We make then the following statements (see Section V of for proofs and discussion):
$`\left(P\right)__1`$ $`=`$ $`\mathrm{Pf}_{s_1,s_2}{\displaystyle \frac{d^3𝐱}{4\pi }\frac{1}{r_1}F}+\left[\mathrm{ln}\left({\displaystyle \frac{r_1^{}}{s_1}}\right)1\right](r_1^2F)__1,`$ (277)
$`\left(Q\right)__1`$ $`=`$ $`\mathrm{Pf}_{s_1,s_2}{\displaystyle \frac{d^3𝐱}{4\pi }r_1F}+\left[\mathrm{ln}\left({\displaystyle \frac{r_1^{}}{s_1}}\right)+{\displaystyle \frac{1}{2}}\right](r_1^4F)__1,`$ (278)
$`\left(_iP\right)__1`$ $`=`$ $`\mathrm{Pf}_{s_1,s_2}{\displaystyle \frac{d^3𝐱}{4\pi }\frac{n_1^i}{r_1^2}F}+\mathrm{ln}\left({\displaystyle \frac{r_1^{}}{s_1}}\right)(n_1^ir_1F)__1,`$ (279)
$`\left(_iQ\right)__1`$ $`=`$ $`\mathrm{Pf}_{s_1,s_2}{\displaystyle \frac{d^3𝐱}{4\pi }n_1^iF}+\left[\mathrm{ln}\left({\displaystyle \frac{r_1^{}}{s_1}}\right)+{\displaystyle \frac{1}{2}}\right](n_1^ir_1^3F)__1,`$ (280)
where the first terms in the right sides are made of some partie-finie integrals in the sense of the definition (4). The “constant” $`r_1^{}`$ is the variable which tends toward zero when evaluating the partie finie. It is easy to show that the constant $`s_1`$ cancels out between the two terms in each of the second members of the identities (V D 2). Indeed, using the general dependence of the partie-finie integral on the constants $`s_1`$, $`s_2`$ as given by (4.20) in , we easily see that $`(P)__1`$ given by (277), for instance, depends on the constants $`r_1^{}`$ and $`s_2`$ through the formula
$$\left(P\right)__1=\mathrm{ln}\left(\frac{r_{12}}{r_1^{}}\right)(r_1^2F)__1\mathrm{ln}\left(\frac{r_{12}}{s_2}\right)\left(\frac{r_2^3}{r_1}F\right)_2+\mathrm{}$$
(281)
where the dots indicate the terms that are independent of any constants. As we see, the constant $`s_1`$ has been so to speak “replaced” by $`r_1^{}`$. This makes clearer why it is convenient to keep the $`\mathrm{ln}r_1^{}`$ in the definition of the Hadamard partie finie; if we had decided to exclude this logarithm from it, we would have found some bare $`\mathrm{ln}r_{12}`$ in the first term of (281) instead of a nicer logarithm of a dimensionless quantity: $`\mathrm{ln}(r_{12}/r_1^{})`$; but this is simply a matter of convenience, because we shall see that in fact the “constants” $`\mathrm{ln}r_1^{}`$ and $`\mathrm{ln}r_2^{}`$ can be gauged away from the 3PN equations of motion. The same argument is valid for all cases in (V D 2). As a consequence, the acceleration of the first body will depend only on two unspecified constants: $`\mathrm{ln}r_1^{}`$, and of course $`\mathrm{ln}s_2`$ (and ditto for the acceleration of the second body). See Section VII for further discussion of these constants.
The relations (V D 2) answer the problem of evaluating the elementary integrals at the location of particle 1 without knowing their values at an arbitrary field point. The subsequent task consists of calculating the partie-finie integrals in the right-hand-sides, which will turn out to be always possible.
#### 3 Integration methods
The non compact parts of the regularized elementary integrals consist essentially of some integrals $`\mathrm{Pf}d^3𝐱F(𝐫_1,𝐫_2)`$, with $`F`$. It is worth noting that the sources depend on $`𝐫_1`$, $`𝐫_2`$ exclusively, and not on the separate variables $`𝐱`$, $`𝐲_1`$, and $`𝐲_2`$, because scalar products like $`(xy_1)`$, $`(xy_2)`$ or $`(y_1y_2)`$ occurring in $`f_{\mathrm{match}}^{12}`$, $`f_{\mathrm{match}}^{21}`$, and $`f_{\mathrm{match}}`$ are killed by the derivatives that precede them in the integrand. In this work, we make use of two different integration methods: (1) the angular method, in which we determine successively angular and radial integrals in spherical coordinates, (2) the analytic continuation method, based on the so-called Riesz formula.
##### a Angular method.
Let $`F=F(𝐫_1,𝐫_2)`$ be a function in the class $``$. We assume for a moment that $`F`$ is locally integrable at point 2; so we are allowed to compute the integral over the whole space, deprived from a small ball $`_1(s)`$ of center $`𝐲_1`$ and radius $`s>0`$. Thus, we start with the well-defined quantity $`_{^3_1(s)}d^3𝐱F(𝐫_1,𝐫_2)`$, but for convenience we write it in spherical coordinates $`(r_1,\theta _1,\varphi _1)`$, such that the azimuthal angle $`\theta _1`$ coincides with the separation angle between $`𝐲_{12}`$ and $`𝐫_1`$,
$$_{^3_1(s)}d^3𝐱F(𝐫_1,𝐫_2)=_{r_1}^+\mathrm{}𝑑r_1r_1^2𝑑\mathrm{\Omega }_1F(𝐫_1,𝐫_2).$$
Actually the function $`F`$ may be a tensor with many indices, but the only unit vectors in the problem are $`𝐧_1`$, $`𝐧_2`$, $`𝐧_{12}`$, and only two of them, say $`𝐧_1`$ and $`𝐧_{12}`$, are independent, by virtue of the relation $`r_1𝐧_1+r_{12}𝐧_{12}=r_2𝐧_2`$; and therefore, $`F`$ can be expressed under the form of a finite sum of tensorial products of type $`n_1^Ln_{12}^K`$ ($`l`$, $`k`$); moreover, each factor admits a symmetric trace-free (STF) decomposition on the basis $`\widehat{n}_1^L=\mathrm{STF}n_1^L`$. Hence we have:
$$F=\underset{l=0}{\overset{l_0}{}}\underset{k=0}{\overset{k_0}{}}\widehat{n}_1^Ln_{12}^KG_{(lk)}(r_1,r_2),l_0,k_0.$$
Here, the $`G_{(lk)}`$’s are some scalar functions of $`r_1`$ and $`r_2`$. Now, $`r_2`$ is related to $`r_1`$ and the scalar product $`(n_1n_{12})`$ by $`r_2=\sqrt{r_1^2+r_{12}^2+2r_1r_{12}(n_1n_{12})}`$. So that the angular integral of $`F`$ can be obtained by means of the “mean formula”; see, e.g., the formula (A26) in . We get
$$𝑑\mathrm{\Omega }_1F=2\pi \underset{l=0}{\overset{l_0}{}}\underset{k=0}{\overset{k_0}{}}\widehat{n}_{12}^Ln_{12}^K_1^1𝑑zG_{(lk)}(r_1,\sqrt{r_1^2+r_{12}^2+2r_1r_{12}z})P_l(z),$$
(282)
where $`P_l`$ is the Legendre polynomial of order $`l`$. In the cases of interest here, $`G_{(lk)}`$ is always a sum of rational fractions with general structure $`r_1^pr_2^q/(r_1+r_2+r_{12})^n`$, with $`p`$, $`q`$ positive or negative integers and $`n`$. It is easy to check that the result of the angular integration depends on the relative positions of $`r_1`$ and $`r_{12}`$. Therefore we must split the radial integral into two contributions according to the integration domains $`r_1]0,r_{12}]`$ or $`r_1]r_{12},+\mathrm{}[`$. Typical terms coming from the angular integration are $`r_1^p/(r_1+r_{12})^n`$, as well as some more complicated logarithmic terms such as $`r_1^p\mathrm{ln}(1+r_{12}/r_1)`$. Most of the corresponding radial integrals are obtained straightforwardly using some integrations by part, applying the partie finie at the bound $`r_1=s`$ \[i.e., removing the poles $`1/s^k`$, with $`k1`$, and replacing $`\mathrm{ln}s`$ by $`\mathrm{ln}s_1`$\]. In the case of the latter logarithmic terms with $`p=1`$, integrating by parts does not lead to anything, but the radial integrals can be found in standard mathematical textbooks:
$`{\displaystyle _{r_{12}}^+\mathrm{}}{\displaystyle \frac{dr_1}{r_1}}\mathrm{ln}\left(1+{\displaystyle \frac{r_{12}}{r_1}}\right)={\displaystyle _0^{r_{12}}}{\displaystyle \frac{dr_1}{r_1}}\mathrm{ln}\left(1+{\displaystyle \frac{r_1}{r_{12}}}\right)={\displaystyle \frac{\pi ^2}{12}},`$ (283)
$`{\displaystyle _{r_{12}}^+\mathrm{}}{\displaystyle \frac{dr_1}{r_1}}\mathrm{ln}\left({\displaystyle \frac{1+{\displaystyle \frac{r_{12}}{r_1}}}{1{\displaystyle \frac{r_{12}}{r_1}}}}\right)={\displaystyle _0^{r_{12}}}{\displaystyle \frac{dr_1}{r_1}}\mathrm{ln}\left({\displaystyle \frac{1+{\displaystyle \frac{r_1}{r_{12}}}}{1{\displaystyle \frac{r_1}{r_{12}}}}}\right)={\displaystyle \frac{\pi ^2}{4}}.`$ (284)
It can be shown that the integrals diverging at $`𝐲_1`$ involve in general a logarithm $`\mathrm{ln}(r_{12}/s_1)`$ but never any $`\pi ^2`$ terms. The procedure we have just described indeed permits calculating most of the elementary integrals. Consider for instance the integral
$${}_{i}{}^{}\underset{1}{𝒢}_{(j,0)}^{}=\frac{d^3𝐱^{}}{4\pi }\frac{1}{|𝐱𝐱^{}|}{}_{i}{}^{}g_j^{^{}}\mathrm{Pf}\frac{1}{r_1^{}}$$
(285)
(in which we replaced $`g_{\mathrm{match}}`$ by $`g=\mathrm{ln}S`$ since they merely differ by a constant). We are interested in the value of this integral at point 1, following the regularization. From (275) we know that the distributional part of the derivative will not contribute. Then, using (277), we readily find
$$\left({}_{i}{}^{}\underset{1}{𝒢}_{(j,0)}^{}\right)_1=\mathrm{Pf}\frac{d^3𝐱}{4\pi }\frac{1}{r_1}{}_{i}{}^{}g_j\frac{1}{r_1}+\frac{\delta ^{ij}}{6r_{12}}\left[\mathrm{ln}\left(\frac{r_1^{}}{s_1}\right)1\right].$$
The non-compact integral has a sole divergence at point 1, so that we can apply the previous method without any change, and get
$$\mathrm{Pf}\frac{d^3𝐱}{4\pi }\frac{1}{r_1}{}_{i}{}^{}g_j\frac{1}{r_1}=\frac{\delta ^{ij}}{18r_{12}}\frac{n_{12}^in_{12}^j}{12r_{12}}\frac{\delta ^{ij}}{6r_{12}}\mathrm{ln}\left(\frac{r_{12}}{s_1}\right).$$
(286)
As expected the constant $`\mathrm{ln}s_1`$ cancels out and we arrive at:
$$\left({}_{i}{}^{}\underset{1}{𝒢}_{(j,0)}^{}\right)__1=\frac{2\delta ^{ij}}{9r_{12}}\frac{n_{12}^in_{12}^j}{12r_{12}}\frac{\delta ^{ij}}{6r_{12}}\mathrm{ln}\left(\frac{r_{12}}{r_1^{}}\right).$$
(287)
A few elementary integrals diverge at the locations $`𝐲_1`$ and $`𝐲_2`$ of both particles. In this occurrence, the integral $`_{^3_1(s)}d^3𝐱F`$ has no meaning, and the previous integration process is not adequate anymore. However, Proposition 2 in Section IV B of allows us to extend it to this case. We introduce the auxiliary source
$$\stackrel{~}{F}_2=F\underset{b+30}{}r_2^b\underset{2}{f_b},$$
which is locally integrable near $`𝐲_2`$ but does not converge at infinity. As before, the angular integral of $`\stackrel{~}{F}_2`$ around $`𝐲_1`$ takes a different expression depending on whether $`r_1r_{12}`$ or $`r_1>r_{12}`$, so we must split the radial integration into the two domains $`]0,r_{12}[`$ and $`]r_{12},+\mathrm{}[`$. Then, with full generality, the partie-finie integral of the source $`F`$ is given by
$`\mathrm{Pf}_{s_1,s_2}{\displaystyle d^3𝐱F}`$ $`=`$ $`\mathrm{Pf}_{s_1}{\displaystyle _0^{r_{12}}}𝑑r_1r_1^2{\displaystyle 𝑑\mathrm{\Omega }_1\stackrel{~}{F}_2}+{\displaystyle _{r_{12}}^+\mathrm{}}𝑑r_1r_1^2{\displaystyle 𝑑\mathrm{\Omega }_1\left[\stackrel{~}{F}_2+\frac{1}{r_1^3}\left(r_2^3F\right)__2\right]}`$ (288)
$`+`$ $`4\pi \left(r_2^3F\right)__2\mathrm{ln}\left({\displaystyle \frac{r_{12}}{s_2}}\right).`$ (289)
If this integral comes from a Poisson integral evaluated at 1, the constant $`s_1`$ will be cancelled out as we have seen previously and replaced by $`r_1^{}`$; but there will remain in general a constant $`s_2`$ coming from the singularity at the other point. With the angular method we were able to obtain all the elementary integrals (and their gradients) at point 1. See an appendix of for the complete list of those results.
##### b Analytical continuation method.
The equivalence between the Hadamard partie finie prescription for integrals and the analytic continuation regularization has long been known (see e.g. ), and we have recovered it in the three-dimensional case by the Theorem 2 of . More precisely, for any $`F`$ that behaves like $`o(1/r^3)`$ when $`r+\mathrm{}`$, the integral $`d^3𝐱(r_1/s_1)^\alpha (r_2/s_2)^\beta F`$ of two complex variables $`\alpha `$ and $`\beta `$ admits an analytic continuation in the neighbourhood of $`\alpha =\beta =0`$, and we have
$$\mathrm{Pf}_{s_1,s_2}d^3𝐱F=\mathrm{FP}_{\genfrac{}{}{0.0pt}{}{\alpha 0}{\beta 0}}d^3𝐱\left(\frac{r_1}{s_1}\right)^\alpha \left(\frac{r_2}{s_2}\right)^\beta F=\mathrm{FP}_{\genfrac{}{}{0.0pt}{}{\beta 0}{\alpha 0}}d^3𝐱\left(\frac{r_1}{s_1}\right)^\alpha \left(\frac{r_2}{s_2}\right)^\beta F,$$
(290)
where $`\mathrm{FP}_{\genfrac{}{}{0.0pt}{}{\alpha 0}{\beta 0}}`$ means taking the finite part of the Laurent expansion of the (analytic continuation of the) integral when $`\alpha 0`$ and $`\beta 0`$ successively. This result is particularly useful in the case where $`F`$ is of the type $`r_1^pr_2^q`$, with $`p`$ and $`q`$ relative integers, since the integral is directly computable thanks to the Riesz formula :
$$d^3𝐱r_1^{\alpha +p}r_2^{\beta +q}=\pi ^{\frac{3}{2}}\frac{\mathrm{\Gamma }\left(\frac{\alpha +p+3}{2}\right)\mathrm{\Gamma }\left(\frac{\beta +q+3}{2}\right)\mathrm{\Gamma }\left(\frac{\alpha +\beta +p+q+3}{2}\right)}{\mathrm{\Gamma }\left(\frac{\alpha +p}{2}\right)\mathrm{\Gamma }\left(\frac{\beta +q}{2}\right)\mathrm{\Gamma }\left(\frac{\alpha +\beta +p+q+6}{2}\right)}r_{12}^{\alpha +\beta +p+q+3}.$$
(291)
One may consult for an example of practical computation. Most of the time, the structure of the sources of elementary integrals is more complicated than a simple $`r_1^pr_2^q`$, notably it involves many “free” tensorial indices which imply that generally the sources of elementary integrals, when fully developped, involve numerous inverse powers of $`S=r_1+r_2+r_{12}`$. However, by considering all the possible contractions of these free indices with vectors $`n_{12}^i`$ and Kronecker symbols $`\delta ^{ij}`$, it happens that we reduce the computation to that of several scalar integrals we can obtain thanks to the Riesz formula (291) (i.e., when performing the contractions and after simplification of the result, we are always led to the simple structure $`r_1^pr_2^q`$ without $`1/S`$ powers). It should be noted that the set of scalar functions that we compute contains the complete information about the complicated tensorial integral, i.e. it permits to reconstitute it exactly.
Let us illustrate the method with the computation of the integral (286). It involves two free indices $`ij`$ and is necessarily of the type
$$\mathrm{Pf}\frac{d^3𝐱}{4\pi }\frac{1}{r_1}{}_{i}{}^{}g_j\frac{1}{r_1}=\varphi (r_{12})n_{12}^in_{12}^j+\psi (r_{12})\delta ^{ij},$$
(292)
where $`\varphi `$ and $`\psi `$ are some unknown “scalars” depending on $`r_{12}`$. By contracting successively the integrand with $`n_{12}^in_{12}^j`$ and $`\delta ^{ij}`$, and simplifying, we find:
$`n_{12}^in_{12}^j{}_{i}{}^{}g_j{\displaystyle \frac{1}{r_1}}`$ $`=`$ $`{\displaystyle \frac{1}{4r_1^3}}{\displaystyle \frac{r_2}{4r_1^4}}+{\displaystyle \frac{1}{4r_1r_{12}^2}}{\displaystyle \frac{r_2}{4r_1^2r_{12}^2}}{\displaystyle \frac{r_2^2}{4r_1^3r_{12}^2}}`$
$`+`$ $`{\displaystyle \frac{r_2^3}{4r_1^4r_{12}^2}}+{\displaystyle \frac{1}{4r_1^2r_{12}}}{\displaystyle \frac{r_2^2}{4r_1^4r_{12}}}+{\displaystyle \frac{r_{12}}{4r_1^4}},`$
$`\delta ^{ij}{}_{i}{}^{}g_j{\displaystyle \frac{1}{r_1}}`$ $`=`$ $`{\displaystyle \frac{1}{2r_1^3}}+{\displaystyle \frac{1}{2r_1^2r_{12}}}{\displaystyle \frac{r_2}{2r_1^3r_{12}}},`$
thus obtaining a sum of terms of the type $`r_1^pr_2^q`$ which can all be integrated with the help of the formula (291). This computation yields a system of equations for the scalars $`\varphi `$ and $`\psi `$:
$`\varphi +\psi ={\displaystyle \frac{5}{36r_{12}}}{\displaystyle \frac{1}{6r_{12}}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{s_1}}\right),`$
$`\varphi +3\psi ={\displaystyle \frac{1}{4r_{12}}}{\displaystyle \frac{1}{2r_{12}}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{s_1}}\right).`$
By solving the previous system, and inserting the results into (292), we recover exactly the value given by (286). We have checked in this manner all the elementary integrals previously computed with the angular method.
#### 4 Finite part of integrals diverging at infinity
For the moment, we have left aside the case where our integrals are defined by means of a finite part dealing with a divergence occuring at infinity. Such a study is needed to compute some 0.5PN integrals encountered in Section IV. From now on, we suppose that the source $`F`$ admits an expansion when $`r+\mathrm{}`$ which is made of simple powers $`r^{n3}`$, with $`nn_{\mathrm{max}}`$ (so the integral a priori diverges at infinity when $`n_{\mathrm{max}}0`$), and we consider the quantity:
$$\mathrm{Pf}\left\{\mathrm{FP}_{B0}d^3𝐱\left(\frac{r}{r_0}\right)^BF\right\}.$$
(293)
We split the integral above into two integrals $`I_{\mathrm{int}}`$ and $`I_{\mathrm{ext}}`$ extending respectively over a domain $`D_{\mathrm{int}}`$ including the two local singularities $`𝐲_{1,2}`$, and the complementary domain $`D_{\mathrm{ext}}`$ comprising the regions at infinity. Using the integration variable $`𝐫_1=𝐱𝐲_1`$, the external integral (on which the regularization $`\mathrm{Pf}`$ can be removed) reads as
$$I_{\mathrm{ext}}=\mathrm{FP}_{B0}_{D_{\mathrm{ext}}}d^3𝐫_\mathrm{𝟏}\left(\frac{|𝐫_1+𝐲_1|}{r_0}\right)^BF(𝐫_\mathrm{𝟏}+𝐲_1).$$
If we assume that the original integral can generate only simple poles $`1/B`$ at infinity (which will be always the case here), we can replace it by
$`I_{\mathrm{ext}}`$ $`=`$ $`\mathrm{FP}_{B0}{\displaystyle _{D_{\mathrm{ext}}}}d^3𝐫_\mathrm{𝟏}\left({\displaystyle \frac{r_1}{r_0}}\right)^BF`$ (294)
$`+`$ $`\mathrm{FP}_{B0}{\displaystyle \frac{B}{2}}{\displaystyle _{D_{\mathrm{ext}}}}d^3𝐫_\mathrm{𝟏}\left({\displaystyle \frac{r_1}{r_0}}\right)^B\mathrm{ln}\left(1+2(n_1y_1){\displaystyle \frac{1}{r_1}}+{\displaystyle \frac{y_1^2}{r_1^2}}\right)F`$ (295)
Indeed, in the case of simple poles, the other terms, involving at least a factor $`B^2`$, will always give zero. The second term is non-zero only if the corresponding integral does admit a pole, and can be calculated in a simple way by picking up the term of order $`1/r_1^3`$ in the expansion of the integrand when $`r+\mathrm{}`$. Notably, in the important case where the function $`F`$ goes to zero like $`1/r^3`$ at infinity, its product with the log-term behaves at least like $`1/r^4`$ and therefore the second term in (295) gives no contribution. For instance, an integral divergent at infinity that we encounter in the problem is
$$\mathrm{FP}_{B0}d^3𝐱\left(\frac{r}{r_0}\right)^B_{ij}\left(\mathrm{Pf}\frac{1}{r_1}\right)=\frac{4\pi }{3}\delta _{ij}.$$
### E Lorentzian regularization of potentials
All the potentials and their gradients (compact-support, quadratic and non-compact potentials) which have been computed in this section and the previous one were obtained at the point 1 using the standard Hadamard regularization $`(F)__1`$. However, this regularization, being defined within the hypersurface $`t=`$const of the harmonic coordinates, must break at some point the Lorentz-invariance properties of the potentials. That is, if a potential defined for a smooth “fluid” behaves in a certain way under a Lorentz transformation, we expect that its regularized value at the point 1 in the sense of $`(F)__1`$ will generically not behave in the same way. Nevertheless, the equations of motion in harmonic coordinates, as computed with the regularization $`(F)__1`$, are known to be Lorentz invariant up to the 2.5PN order . Perhaps not surprisingly because of this fact, it turns out that the Lorentzian regularization $`[F]__1`$ (defined in ) yields no difference with respect to the old regularization $`(F)__1`$ for all the 3PN potentials but for one, namely the cubic-non-compact potential $`\widehat{X}^{(\mathrm{CNC})}`$ defined by (228) and which had to be computed at the relative 1PN order. (Evidently we have $`[F]__1=(F)__1`$ for all the potentials which are to be computed with Newtonian accuracy.) Thus all the results obtained so far but the one for $`(_i\widehat{X}^{(\mathrm{CNC})})__1`$ \[the relevant quantity for the equations of motion\] are valid in the case of the new regularization. In this subsection we compute the remaining part $`[_i\widehat{X}^{(\mathrm{CNC})}]__1(_i\widehat{X}^{(\mathrm{CNC})})__1`$.
Since the regularization $`[F]__1`$ brings some new terms with respect to the old one $`(F)__1`$ starting at the relative 1PN order, and since $`\widehat{X}^{(\mathrm{CNC})}`$ is to be computed at the 1PN order only, it is sufficient for this calculation to use the lowest-order, Newtonian value of $`\widehat{X}^{(\mathrm{CNC})}`$. From the computation in we know the analytic closed-form expression of $`\widehat{X}^{(\mathrm{CNC})}`$ at Newtonian order for any field point $`(t,𝐱)`$,
$$\widehat{X}^{(\mathrm{CNC})}=\frac{G^3m_1^3}{12r_1^3}G^3m_1^2m_2\left\{\frac{1}{8r_2r_{12}^2}+\frac{1}{16}K_1+H_1\right\}+𝒪(1)+12,$$
(296)
where the functions $`K_1`$ and $`H_1`$, which are solutions of certain Poisson equations, are given explicitly by (V D 1). The complete developed forms of these functions are
$`K_1`$ $`=`$ $`{\displaystyle \frac{1}{r_2^3}}+{\displaystyle \frac{1}{r_2r_{12}^2}}{\displaystyle \frac{1}{r_1^2r_2}}+{\displaystyle \frac{r_2}{2r_1^2r_{12}^2}}+{\displaystyle \frac{r_{12}^2}{2r_1^2r_2^3}}+{\displaystyle \frac{r_1^2}{2r_2^3r_{12}^2}},`$ (298)
$`H_1`$ $`=`$ $`{\displaystyle \frac{1}{2r_1^3}}{\displaystyle \frac{1}{4r_{12}^3}}{\displaystyle \frac{1}{4r_1^2r_{12}}}{\displaystyle \frac{r_2}{2r_1^2r_{12}^2}}+{\displaystyle \frac{r_2}{2r_1^3r_{12}}}+{\displaystyle \frac{3r_2^2}{4r_1^2r_{12}^3}}+{\displaystyle \frac{r_2^2}{2r_1^3r_{12}^2}}{\displaystyle \frac{r_2^3}{2r_1^3r_{12}^3}}.`$ (299)
We replace these expressions into (296), and we implement all the rules for the new regularization $`[F]__1`$ defined in Section III of . Equivalently, since the order of the computation is limited to 1PN, we can use the closed form formula
$$[F]__1(F)__1=\frac{1}{c^2}\left((𝐫_1.𝐯_1)[_tF+\frac{1}{2}v_1^i_iF]\right)__1+𝒪\left(4\right),$$
(300)
derived in Section IV of . As a result, we obtain, for the potential itself,
$`[\widehat{X}^{(\mathrm{CNC})}]__1`$ $``$ $`(\widehat{X}^{(\mathrm{CNC})})__1`$ (301)
$`=`$ $`{\displaystyle \frac{G^3m_1^2m_2}{c^2r_{12}^3}}\left\{{\displaystyle \frac{43}{40}}(n_{12}v_1)^2(n_{12}v_1)(n_{12}v_2){\displaystyle \frac{43}{120}}v_1^2+{\displaystyle \frac{1}{3}}(v_1v_2)\right\}+𝒪(4).`$ (302)
In the case of the gradient needed for the equations of motion, we get
$`[_i\widehat{X}^{(\mathrm{CNC})}]__1`$ $``$ $`(_i\widehat{X}^{(\mathrm{CNC})})__1`$ (303)
$`=`$ $`{\displaystyle \frac{G^3m_1^2m_2}{c^2r_{12}^4}}\{[{\displaystyle \frac{27}{56}}(n_{12}v_1)^2{\displaystyle \frac{3}{4}}(n_{12}v_1)(n_{12}v_2){\displaystyle \frac{27}{280}}v_1^2+{\displaystyle \frac{3}{20}}(v_1v_2)]n_{12}^i`$ (304)
$``$ $`{\displaystyle \frac{27}{140}}(n_{12}v_1)v_1^i+{\displaystyle \frac{3}{20}}(n_{12}v_2)v_1^i+{\displaystyle \frac{3}{20}}(n_{12}v_1)v_2^i\}+𝒪(4).`$ (305)
As we see, the new regularization brings some definite non-zero contributions at the 1PN order in the case of this potential, which will constitute a crucial contribution to the 3PN equations of motion. The right-hand side of (303) is not invariant by itself under a Lorentz transformation — it cannot be —, but will ensure finally the Lorentz-invariance of the 3PN equations of motion.
## VI Leibniz terms and non-distributivity
### A Effect of a gauge transformation
In this subsection we study the effect of a gauge transformation on the 3PN equations of motion as well as energy of the two particles. Let $`\{x^\mu \}`$ denote the harmonic coordinate system and $`g_{\mu \nu }(x)`$ be the harmonic-coordinate metric, generated by the two particles, that we have iterated in previous sections up to the 3PN order. The metric depends on the position $`𝐱`$ of the field point, and on the coordinate time $`t=x^0/c`$ through the trajectories $`𝐲_{1,2}(t)`$ and velocities $`𝐯_{1,2}(t)`$ of the particles, i.e.
$$g_{\mu \nu }(x)=g_{\mu \nu }[𝐱;𝐲_1(t),𝐲_2(t);𝐯_1(t),𝐯_2(t)].$$
(306)
We know that the dependence of the metric over the velocities arises at the 1PN order (see e.g. the equations (7.2) in ), namely the order $`𝒪(4,3,4)`$, where this notation is a shorthand for saying $`𝒪(4)=𝒪(1/c^4)`$ in $`g_{00}`$, $`𝒪(3)`$ in $`g_{0i}`$ and $`𝒪(4)`$ in $`g_{ij}`$. Consider an infinitesimal coordinate transformation of the type
$`x_{}^{}{}_{}{}^{\mu }=x^\mu +\xi ^\mu (x),`$ (308)
$`\xi ^\mu (x)=\xi ^\mu [𝐱;𝐲_1(t),𝐲_2(t)],`$ (309)
where, in order to simplify the presentation, we assume that the gauge vector $`\xi ^\mu `$ depends on the positions $`𝐲_{1,2}`$ of the particles, but not on their velocities. Furthermore, we suppose that this gauge transformation is at the level of the 3PN order, which means that $`\xi ^0=𝒪(7)`$ and $`\xi ^i=𝒪(6)`$, or equivalently $`\xi ^\mu =𝒪(7,6)`$. In addition, in a first stage, we suppose that the vector $`\xi ^\mu (x)`$ is a smooth function of the coordinates even at the positions of the particles. The new metric in the new coordinate system $`\{x_{}^{}{}_{}{}^{\mu }\}`$ is
$$g_{\mu \nu }^{}(x^{})=g_{\mu \nu }^{}[𝐱^{};𝐲_1^{}(t^{}),𝐲_2^{}(t^{});𝐯_1^{}(t^{}),𝐯_2^{}(t^{})],$$
(310)
where the new trajectories and velocities $`𝐲_{1,2}^{}`$, $`𝐯_{1,2}^{}`$ are parametrized by the new coordinate time $`t^{}=x_{}^{}{}_{}{}^{0}/c`$. The coordinate change (VI A), when applied at the location of each of the particles, yields the relations between the new and old trajectories, which, when retaining only the terms up to the order $`𝒪(6)`$, read as
$`y_{}^{}{}_{1}{}^{i}(t^{})`$ $`=`$ $`y_1^i(t)+\xi ^i(y_1)+𝒪(8),`$ (312)
$`y_{}^{}{}_{2}{}^{i}(t^{})`$ $`=`$ $`y_2^i(t)+\xi ^i(y_2)+𝒪(8),`$ (313)
where the $`\xi ^\mu (y_{1,2})`$’s denote the gauge vector at the position of the particles, for instance $`\xi ^\mu (y_1)=\xi ^\mu [𝐲_1(t);𝐲_1(t),𝐲_2(t)]`$. The new metric (310), when expressed in terms of the old variables, follows from this as
$$g_{\mu \nu }^{}(x^{})=g_{\mu \nu }^{}(x)+\xi ^i(x)_ig_{\mu \nu }+\xi ^i(y_1){}_{1}{}^{}_{i}^{}g_{\mu \nu }+\xi ^i(y_2){}_{2}{}^{}_{i}^{}g_{\mu \nu }+𝒪(10,9,10),$$
(314)
where we have used the fact that the dependence of the metric on the velocities starts at the 1PN order, so the terms due to the modification of the velocities do not contribute to (314). Since the 3PN metric depends on space $`𝐱`$ only through the two distances $`𝐱𝐲_1`$ and $`𝐱𝐲_2`$, we have $`_ig_{\mu \nu }+{}_{1}{}^{}_{i}^{}g_{\mu \nu }+{}_{2}{}^{}_{i}^{}g_{\mu \nu }=0`$, and so an equivalent form of (314) is
$$g_{\mu \nu }^{}(x^{})=g_{\mu \nu }^{}(x)+[\xi ^i(y_1)\xi ^i(x)]{}_{1}{}^{}_{i}^{}g_{\mu \nu }+[\xi ^i(y_2)\xi ^i(x)]{}_{2}{}^{}_{i}^{}g_{\mu \nu }+𝒪(10,9,10).$$
(315)
The equation (315), when combined with the law of transformation of tensors, i.e. in the present case
$$g_{\mu \nu }(x)=g_{\mu \nu }^{}(x^{})+_\mu \xi _\nu +_\nu \xi _\mu +𝒪(10,9,8),$$
(316)
where $`\xi _\mu =\eta _{\mu \nu }\xi ^\nu `$, gives the metric variation or Lie derivative $`\delta _\xi g_{\mu \nu }=g_{\mu \nu }^{}(x)g_{\mu \nu }(x)`$ (where the same variable $`x`$ is used for both the transformed and original metrics) as
$$\delta _\xi g_{\mu \nu }=_\mu \xi _\nu _\nu \xi _\mu +[\xi ^i(x)\xi ^i(y_1)]{}_{1}{}^{}_{i}^{}g_{\mu \nu }+[\xi ^i(x)\xi ^i(y_2)]{}_{2}{}^{}_{i}^{}g_{\mu \nu }+𝒪(10,9,8).$$
(317)
In fact, up to this order, only the $`00`$ component of the metric includes a “non-linear” correction term; and, within that non-linear term, the metric can be approximated by its Newtonian part, so
$`\delta _\xi g_{00}`$ $`=`$ $`2_0\xi _0+{\displaystyle \frac{2}{c^2}}\left([\xi ^i(x)\xi ^i(y_1)]{}_{1}{}^{}_{i}^{}U+[\xi ^i(x)\xi ^i(y_2)]{}_{2}{}^{}_{i}^{}U\right)+𝒪(10),`$ (319)
$`\delta _\xi g_{0i}`$ $`=`$ $`_0\xi _i_i\xi _0+𝒪(9),`$ (320)
$`\delta _\xi g_{ij}`$ $`=`$ $`_i\xi _j_j\xi _i+𝒪(8),`$ (321)
where $`U=\frac{Gm_1}{r_1}+\frac{Gm_2}{r_2}`$ is the Newtonian potential (with a small inconsistency of notation with respect to previous sections). Now, it is easy to check that, in the sense of distributions,
$$\mathrm{\Delta }\left([\xi ^i(x)\xi ^i(y_1)]{}_{1}{}^{}_{i}^{}U+[\xi ^i(x)\xi ^i(y_2)]{}_{2}{}^{}_{i}^{}U\right)=2_i\xi _j_{ij}U\mathrm{\Delta }\xi _i_iU.$$
Indeed, the delta-functions at the points 1 and 2, which come from the Laplacian of $`U`$, are killed respectively by the factors $`\xi ^i(x)\xi ^i(y_1)`$ and $`\xi ^i(x)\xi ^i(y_2)`$, which vanish respectively at these two points, in front of them. So, we can write for $`\delta _\xi g_{00}`$ the simpler but equivalent expression
$$\delta _\xi g_{00}=2_0\xi _0\frac{2}{c^2}\mathrm{\Delta }^1[2_i\xi _j_{ij}U+\mathrm{\Delta }\xi _i_iU]+𝒪(10),$$
(322)
where $`\mathrm{\Delta }^1`$ denotes the usual Poisson integral.
The latter result can be generalized to our framework of singular metrics by allowing the gauge vector $`\xi ^\mu `$ to become singular at the positions of the particles (in the sense that $`\xi ^\mu `$), provided that the integral appearing in (322) is treated as the Hadamard partie finie of a Poisson integral in the way which is investigated in Section V of . Let us consider, for example, the 3PN gauge vector given by
$$\xi _\mu =\frac{G^3m^3}{c^6}_\mu \left(\frac{ϵ_1}{r_1}+\frac{ϵ_2}{r_2}\right),$$
(323)
where $`ϵ_1`$ and $`ϵ_2`$ denote two dimensionless constants or possibly functions of time $`t`$ (and where $`m=m_1+m_2`$). Note that with this choice of gauge vector the new coordinates satisfy the condition of harmonic coordinates outside the singularities (i.e. in the sense of functions) at the 3PN order: indeed $`\mathrm{}\xi ^\mu =𝒪(9,8)`$. When inserting (323) into (322), we must be careful about evaluating the last term of (322) in the sense of distributions, taking into account the fact that $`\mathrm{\Delta }\xi _i`$ is distributional. For this term we obtain
$$\mathrm{\Delta }^1[\mathrm{\Delta }\xi _i_iU]=\frac{G^3m^3}{c^6}\left[\gamma _1^i_i\left(\frac{ϵ_1}{r_1}\right)+\gamma _2^i_i\left(\frac{ϵ_2}{r_2}\right)\right],$$
where $`\gamma _1^i`$ and $`\gamma _2^i`$ are the Newtonian accelerations of 1 and 2. Therefore, we find
$$\delta _\xi g_{00}=\frac{2G^3m^3}{c^8}\left\{\left[_t^2+\gamma _1^i_i\right]\left(\frac{ϵ_1}{r_1}\right)+2\mathrm{\Delta }^1\left[_{ij}\left(\frac{ϵ_1}{r_1}\right)_{ij}U\right]\right\}+𝒪(10)+12.$$
(324)
In the case where $`ϵ_1`$ and $`ϵ_2`$ are some pure constants (independent on time) we can somewhat simplify the latter expression by using the fact that the accelerations cancel out in the first term. In this case, we obtain the full metric transformations as
$`\delta _\xi g_{00}`$ $`=`$ $`{\displaystyle \frac{2G^3m^3}{c^8}}\left\{v_1^{ij}_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)+2\mathrm{\Delta }^1\left[_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)_{ij}U\right]\right\}+𝒪(10)+12,`$ (326)
$`\delta _\xi g_{0i}`$ $`=`$ $`{\displaystyle \frac{2G^3m^3}{c^7}}v_1^j_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)+𝒪(9)+12,`$ (327)
$`\delta _\xi g_{ij}`$ $`=`$ $`{\displaystyle \frac{2G^3m^3}{c^6}}_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)+𝒪(8)+12.`$ (328)
By comparing this with the 3PN metric (III C), we see that the gauge transformation induces the following changes in the 3PN potentials $`\widehat{T}`$, $`\widehat{Y}_i`$ and $`\widehat{Z}_{ij}`$:
$`\delta _\xi \widehat{T}`$ $`=`$ $`{\displaystyle \frac{G^3m^3}{16}}\left\{v_1^{ij}_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)+2\mathrm{\Delta }^1\left[_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)_{ij}U\right]\right\}+12,`$ (330)
$`\delta _\xi \widehat{Y}_i`$ $`=`$ $`{\displaystyle \frac{G^3m^3}{8}}v_1^j_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)+12,`$ (331)
$`\delta _\xi \widehat{Z}_{ij}`$ $`=`$ $`{\displaystyle \frac{G^3m^3}{8}}_{ij}\left({\displaystyle \frac{ϵ_1}{r_1}}\right)+12.`$ (332)
The computation of the non-linearity term in (330) is straightforward, and we get
$$\mathrm{\Delta }^1\left[_{ij}\left(\frac{1}{r_1}\right)_{ij}U\right]=\frac{Gm_1}{2r_1^4}+Gm_2{}_{ij}{}^{}g_{ij}^{},$$
where $`g=\mathrm{ln}(r_1+r_2+r_{12})`$ is a kernel satisfying $`\mathrm{\Delta }g=\frac{1}{r_1r_2}`$ \[see also (IV C 3)\], and where we denote $`{}_{ij}{}^{}g_{ij}^{}=_{1ij}_{2ij}g=D^2g`$. At last, we insert the latter changes of the 3PN potentials into the (regularized) equations of motion (102) with (III D) (there is no need to include a correction due to the non-distributivity), and obtain the corresponding change in the acceleration of the particle 1 as
$`\delta _\xi a_1^i`$ $`=`$ $`{\displaystyle \frac{2G^4m^3}{c^6r_{12}^5}}\left(ϵ_1m_2ϵ_2m_1\right)n_{12}^i`$ (333)
$`+`$ $`{\displaystyle \frac{G^3m^3ϵ_2}{c^6r_{12}^4}}\left[15(n_{12}v_{12})^2n_{12}^i+3v_{12}^2n_{12}^i+6(n_{12}v_{12})v_{12}^i\right].`$ (334)
In the case where $`ϵ_1`$ and $`ϵ_2`$ depend on time, there are some extra contributions proportional to $`\dot{ϵ}_2`$ and $`\ddot{ϵ}_2`$. A good check of (333) is the fact that to the change in the acceleration (333) always corresponds a change in the associated energy; that is, the gauge transformation does not modify the existence of a conserved energy (see Section VII for the computation of the 3PN energy). Namely, we find that the combination $`m_1\delta _\xi a_1^iv_1^i+m_2\delta _\xi a_2^iv_2^i`$ is a total time-derivative, and from this we obtain the gauge transformation of the energy as
$$\delta _\xi E=ϵ_2\frac{G^3m^3m_1}{c^6r_{12}^3}\left[\frac{Gm_2}{r_{12}}3(n_{12}v_1)(n_{12}v_{12})+(v_1v_{12})\right]+12.$$
(335)
### B Leibniz contributions
An important ingredient of the present computation is the novel distributional derivative associated with the Hadamard regularization which has been introduced in (see also Section II). This derivative permits us to derive in a systematic and consistent way all the integrals encountered in the problem; however it represents merely a mathematical tool, which is maybe not connected to any relevant Physics. Therefore, it is important to know exactly the role played by this derivative in the 3PN equations of motion, with respect to say the Schwartz distributional derivative . We know that our distributional derivative affects the computation of two types of terms: (i) the “self” terms entering a priori in the non-linear potentials $`\widehat{X}`$, $`\widehat{T}`$ and $`\widehat{Y}_i`$ and which are ill-defined in the case of the Schwartz derivative (see Section V), (ii) the “Leibniz” terms which account for the violation of the Leibniz rule during the 3PN iteration of the metric as discussed in Section III. In the present subsection, we compute the Leibniz terms and combine the result with the one of Section V concerning the self terms. The conclusion is that the terms coming from the use of the distributional derivative are necessary for keeping track of the Lorentz invariance of the equations of motion, and that no other Physics is involved with them in the present formalism (see also Section VII). We do the computation for both the “particular” derivative defined by (10) and the more correct one given by (11)-(12).
The Leibniz terms discussed in Section III consist of those contributions of the type (68) and alike which arise in the process of simplification of the 3PN metric $`h^{\mu \nu }`$ by means of the Leibniz rule. These terms depend only on the distributional part of the derivative, $`\mathrm{D}_i^{\mathrm{part}}[F]`$ or $`\mathrm{D}_i[F]`$. The formulas giving the complete Leibniz terms in $`h^{\mu \nu }`$, not being very attractive, are relegated to Appendix A. When reducing explicitly these formulas we find that all the terms take the same simple structure and, not surprisingly, arise only at the 3PN order. As already announced in (III C), the Leibniz terms imply a priori a net contribution to the 3PN potentials $`\widehat{T}`$, $`\widehat{Y}_i`$ and $`\widehat{Z}_{ij}`$. Actually, in the case of the particular derivative, their contributions to the vector and tensor potentials $`\widehat{Y}_i`$ and $`\widehat{Z}_{ij}`$ turn out to be zero,
$`\delta _{\mathrm{Leibniz}}\widehat{Y}_i=0,`$ (337)
$`\delta _{\mathrm{Leibniz}}\widehat{Z}_{ij}=0,`$ (338)
while the contribution to the scalar potential $`\widehat{T}`$, also in the case of $`\mathrm{D}_i^{\mathrm{part}}[F]`$, is found to be
$$\delta _{\mathrm{Leibniz}}\widehat{T}=\frac{G^3m_1^3}{96}v_1^iv_1^j_{ij}\left(\frac{1}{r_1}\right)+\frac{11}{36}\frac{G^4m_1^3m_2}{r_{12}^2}n_{12}^i_i\left(\frac{1}{r_1}\right)+12.$$
(339)
The modification of the acceleration of body 1 which is generated by the latter Leibniz terms reads as
$$\delta _{\mathrm{Leibniz}}a_1^i=\frac{G^3m_2^3}{c^6r_{12}^4}\left[\frac{5}{2}(n_{12}v_2)^2n_{12}^i\frac{1}{2}v_2^2n_{12}^i(n_{12}v_2)v_2^i\right]\frac{88}{9}\frac{G^4m_1m_2^3}{c^6r_{12}^5}n_{12}^i.$$
(340)
On the other hand, we computed in Sections IV and V many distributional terms associated with the derivative of the non-linear potentials in the right-hand side of the field equations \[see (54)\]. Most of these terms are simply given by the Schwartz distributional derivative. The only terms which require the new distributional derivative of come from the computation of the “self” parts of the non-compact potentials (see Section V). In this case, the modifications of the potentials have been found to be given by (V B), from which we obtain the following modification of the acceleration:
$$\delta _{\mathrm{self}}a_1^i=\frac{G^3m_2^3}{c^6r_{12}^4}\left[\frac{1}{2}(n_{12}v_2)^2n_{12}^i+\frac{1}{10}v_2^2n_{12}^i+\frac{1}{5}(n_{12}v_2)v_2^i\right]+\frac{151}{9}\frac{G^4m_1m_2^3}{c^6r_{12}^5}n_{12}^i.$$
(341)
Adding up (340) and (341) we therefore obtain the total effect of the (particular) distributional derivative as
$$\delta _{\mathrm{distribution}}a_1^i=\frac{G^3m_2^3}{c^6r_{12}^4}\left[2(n_{12}v_2)^2n_{12}^i\frac{2}{5}v_2^2n_{12}^i\frac{4}{5}(n_{12}v_2)v_2^i\right]+7\frac{G^4m_1m_2^3}{c^6r_{12}^5}n_{12}^i.$$
(342)
Interestingly, this quite simple piece of the acceleration of particle 1 involves the velocity of particle 2 alone, and therefore does not stay by itself invariant under a Lorentz transformations, or, rather, at this order, a Galilean transformation (indeed, for this to be true the term should depend on the relative velocity $`𝐯_{12}=𝐯_1𝐯_2`$). Therefore, if we are correct, since we are using harmonic coordinates and have employed a Lorentzian regularization, the result (342) has to combine with other pieces in the acceleration so as to maintain the Lorentz invariance of the equations. We have found that this is exactly what happens: the dependence of (342) over the velocity $`𝐯_2`$ is mandatory for the Lorentz invariance of the final 3PN equations to work. This constitutes, in our opinion, an important check of the relevance of the distributional derivative introduced in . It shows also that this derivative is merely a tool for preventing a breakdown of the Lorentz invariance when performing integrations by parts of complicated divergent integrals \[the last term in (342), which is not checked by the Lorentz invariance, will turn out to be absorbed into the adjustement of a certain constant, see Section VII\].
The previous check has been done with the “particular” distributional derivative (10), and it is interesting to redo the computation in the case of the distributional derivative defined by (11)-(12), that we recall is more satisfying than the particular one because it obeys the rule of commutation of successive derivatives. (But note in passing that we have verified that the particular derivative does not yield any ambiguity at the 3PN order which would be due to the non-commutation of derivatives; however, such ambiguities could arise at higher orders, in which case the “correct” derivative would be more appropriate.) In particular, while the particular derivative is entirely deterministic, the derivative (11)-(12) depends on a constant $`K`$, and it is important to know the fate of this constant in the final equations of motion, and how the test of the Lorentz invariance will manage to be satisfied in fine. Like for the particular derivative we find that the incidence of this derivative is through two distinct contributions, Leibniz and self. Consider the Leibniz contribution: we perform exactly the same computation as before, i.e. based on the formulas in the appendix A, and find that in the case of the new derivative (11)-(12) the terms in the potentials $`\widehat{Y}_i`$ and $`\widehat{Z}_{ij}`$ are no longer zero, but are given by
$`\delta _{\mathrm{Leibniz}}\widehat{Y}_{i}^{}{}_{|_K}{}^{}=\left({\displaystyle \frac{1}{15}}+{\displaystyle \frac{2}{15}}K\right)G^3m_1^3v_1^j_{ij}\left({\displaystyle \frac{1}{r_1}}\right)+12,`$ (344)
$`\delta _{\mathrm{Leibniz}}\widehat{Z}_{ij}^{}{}_{|_K}{}^{}=\left({\displaystyle \frac{1}{15}}+{\displaystyle \frac{2}{15}}K\right)G^3m_1^3_{ij}\left({\displaystyle \frac{1}{r_1}}\right)+12.`$ (345)
Our convention is that the explicit indication in the left-hand side of the dependence over $`K`$ means that the computation is performed using the “correct” distributional derivative. In the case of the modification of the potential $`\widehat{T}`$ the things are a little more complicated because we have to take into account, in addition to a “linear” contribution similar to those of (VI B), the “non-linear” term which is generated by the modification of the tensor potential $`\widehat{Z}_{ij}`$ shown in (345); cf the source term $`\widehat{Z}_{ij}_{ij}V`$ in the definition (85) of $`\widehat{T}`$. We obtain
$`\delta _{\mathrm{Leibniz}}\widehat{T}_{|_K}`$ $`=`$ $`\left({\displaystyle \frac{53}{480}}+{\displaystyle \frac{2}{5}}K\right)G^3m_1^3v_1^iv_1^j_{ij}\left({\displaystyle \frac{1}{r_1}}\right)`$ (346)
$`+`$ $`\left({\displaystyle \frac{19}{288}}+{\displaystyle \frac{47}{24}}K\right){\displaystyle \frac{G^4m_1^3m_2}{r_{12}^2}}n_{12}^i_i\left({\displaystyle \frac{1}{r_1}}\right)`$ (347)
$`+`$ $`\mathrm{\Delta }^1\left[\delta _{\mathrm{Leibniz}}\widehat{Z}_{ij}^{}{}_{|_K}{}^{}_{ij}U\right]+12,`$ (348)
where $`U=\frac{Gm_1}{r_1}+\frac{Gm_2}{r_2}`$. Now, using the results of the previous subsection, we see that many of these terms are in the form of a gauge transformation corresponding to a gauge vector $`\xi ^\mu `$ of the type (323). Indeed, we pose
$`ϵ_{1}^{}{}_{|_K}{}^{}`$ $`=`$ $`{\displaystyle \frac{8}{15}}\left(12K\right)\left({\displaystyle \frac{m_1}{m}}\right)^3,`$ (350)
$`ϵ_{2}^{}{}_{|_K}{}^{}`$ $`=`$ $`{\displaystyle \frac{8}{15}}\left(12K\right)\left({\displaystyle \frac{m_2}{m}}\right)^3.`$ (351)
With this choice the Leibniz corrections in $`\widehat{Y}_i`$ and $`\widehat{Z}_{ij}`$ become pure gauge,
$`\delta _{\mathrm{Leibniz}}\widehat{Y}_{i}^{}{}_{|_K}{}^{}`$ $`=`$ $`\delta _\xi \widehat{Y}_{i}^{}{}_{|_K}{}^{},`$ (353)
$`\delta _{\mathrm{Leibniz}}\widehat{Z}_{ij}^{}{}_{|_K}{}^{}`$ $`=`$ $`\delta _\xi \widehat{Z}_{ij}^{}{}_{|_K}{}^{},`$ (354)
while we can re-write (346) in the simplified form
$`\delta _{\mathrm{Leibniz}}\widehat{T}_{|_K}`$ $`=`$ $`\left({\displaystyle \frac{37}{480}}+{\displaystyle \frac{1}{3}}K\right)G^3m_1^3v_1^iv_1^j_{ij}\left({\displaystyle \frac{1}{r_1}}\right)`$ (355)
$`+`$ $`\left({\displaystyle \frac{19}{288}}+{\displaystyle \frac{47}{24}}K\right){\displaystyle \frac{G^4m_1^3m_2}{r_{12}^2}}n_{12}^i_i\left({\displaystyle \frac{1}{r_1}}\right)+12+\delta _\xi \widehat{T}_{|_K}.`$ (356)
The corresponding modification of the acceleration of particle 1 is found to be
$`\delta _{\mathrm{Leibniz}}a_{1}^{i}{}_{|_K}{}^{}`$ $`=`$ $`\left({\displaystyle \frac{37}{2}}80K\right){\displaystyle \frac{G^3m_2^3}{c^6r_{12}^4}}\left[(n_{12}v_2)^2n_{12}^i{\displaystyle \frac{1}{5}}v_2^2n_{12}^i{\displaystyle \frac{2}{5}}(n_{12}v_2)v_2^i\right]`$ (357)
$`+`$ $`\left({\displaystyle \frac{19}{9}}{\displaystyle \frac{188}{3}}K\right){\displaystyle \frac{G^4m_1m_2^3}{c^6r_{12}^5}}n_{12}^i+\delta _\xi a_{1}^{i}{}_{|_K}{}^{},`$ (358)
where the last term represents the gauge term (333) but computed with (VI B). Therefore, modulo a change of gauge, we see that the Leibniz modification of the acceleration brought about by the correct derivative has exactly the same form as that, given by (340), due to the particular one. However, we must also include the contribution of the self terms. We have redone the computation of the self terms as in Section V but using the $`K`$-dependent derivative and compared the corresponding acceleration with the previous result (341). We get
$`\delta _{\mathrm{self}}a_{1}^{i}{}_{|_K}{}^{}\delta _{\mathrm{self}}a_1^i`$ $`=`$ $`{\displaystyle \frac{G^3m_2^3}{c^6r_{12}^4}}\left[{\displaystyle \frac{9}{2}}(n_{12}v_2)^2n_{12}^i{\displaystyle \frac{9}{10}}v_2^2n_{12}^i{\displaystyle \frac{9}{5}}(n_{12}v_2)v_2^i\right]`$ (359)
$`+`$ $`\left({\displaystyle \frac{20}{3}}+{\displaystyle \frac{44}{3}}K\right){\displaystyle \frac{G^4m_1m_2^3}{c^6r_{12}^5}}n_{12}^i.`$ (360)
Subtracting (340) and (357) for the Leibniz terms, and adding up the difference of self terms given by (359), we thereby obtain the difference between the total effects of the two distributional derivatives in the acceleration as
$`\delta _{\mathrm{distribution}}a_{1}^{i}{}_{|_K}{}^{}\delta _{\mathrm{distribution}}a_1^i`$ $`=`$ $`\left({\displaystyle \frac{41}{2}}80K\right){\displaystyle \frac{G^3m_2^3}{c^6r_{12}^4}}\left[(n_{12}v_2)^2n_{12}^i{\displaystyle \frac{1}{5}}v_2^2n_{12}^i{\displaystyle \frac{2}{5}}(n_{12}v_2)v_2^i\right]`$ (361)
$`+`$ $`\left(148K\right){\displaystyle \frac{G^4m_1m_2^3}{c^6r_{12}^5}}n_{12}^i+\delta _\xi a_{1}^{i}{}_{|_K}{}^{}.`$ (362)
As we see, there is a dependence on the individual velocity $`𝐯_2`$ which is left out. Anticipating the result that $`a_1^i`$, computed with the particular derivative, is invariant under Lorentz transformations, this means that the $`K`$-dependent derivative breaks down the Lorentz invariance for general values of $`K`$ (indeed the gauge term cannot modify the behaviour under Lorentz transformations). Fortunately, we are now able to fine tune the constant $`K`$ so that the velocity-dependent terms in (361) vanish. Therefore, we obtain a unique value,
$$K=\frac{41}{160},$$
(363)
for which the equations of motion computed with the help of the correct derivative (11)-(12) are Lorentz invariant, as they are with the particular derivative.
Thus, in the case of the correct derivative, the equations of motion are not in general Lorentz-invariant, despite the use of the Lorentzian regularization. The likely reason is that the distributional derivatives were not defined in a “Lorentzian” way (their distributional terms involve the delta-pseudo-function $`\mathrm{Pf}\delta _1`$ and not the Lorentzian one $`\mathrm{Pf}\mathrm{\Delta }_1`$). Recall that the Lorentzian regularization permitted to add some crucial contributions, proportional to $`m_1^2m_2`$ in the acceleration of particle 1 \[see for instance (303)\], which are mandatory for satisfying the Lorentz invariance. In the case of the correct derivative, we find that there is still a limited class of terms, proportional to $`m_2^3`$, which do not obey the Lorentz invariance, unless $`K`$ is adjusted to the unique value (363). Finally we obtain
$$\delta _{\mathrm{distribution}}a_{1}^{i}{}_{|_{{\scriptscriptstyle \frac{41}{160}}}}{}^{}\delta _{\mathrm{distribution}}a_1^i=\frac{113}{10}\frac{G^4m_1m_2^3}{c^6r_{12}^5}n_{12}^i+\delta _\xi a_{1}^{i}{}_{|_{{\scriptscriptstyle \frac{41}{160}}}}{}^{}.$$
(364)
We shall see in Section VII that the effect of the first term is simply to modify a logarithmic constant $`\mathrm{ln}(r_2^{}/s_2)`$ that we shall adjust when we look for a conserved 3PN energy. After adjustement of this constant we find that the 3PN equations of motion computed with the two derivatives are physically the same since they merely differ by the gauge transformation appearing in (364).
### C Non-distributivity contributions
The distributive parts of the linear momentum $`P_1^i`$ and force $`F_1^i`$ densities have been written down in (III D). They were obtained under the uncorrect hypothesis of distributivity, that is $`[FG]__1=[F]__1[G]__1`$, and we must now correct for this. (As explained in Section III, our strategy has been to delineate as much as possible the problems, by concentrating our attention first on the computation of the regularized values of the potentials when taken individually, second on the corrections due to the non-distributivity, i.e. $`[FG]__1[F]__1[G]__1`$.) Again, we find that such a subtlety as the non-distributivity makes a difference starting precisely at the 3PN order. We get for the required corrections in $`P_1^i`$ and $`F_1^i`$:
$`P_1^i(P_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{1}{c^4}}\left(v_1^i[V^2]__1v_1^i[V]__1^2\right)`$ (366)
$`+`$ $`{\displaystyle \frac{1}{c^6}}(12v_1^i[V_jV_j]__112v_1^i[V_j]__1[V_j]__1+2v_1^i[V^3]__1`$ (371)
$`3v_1^i[V]__1[V^2]__1+v_1^i[V]__1^3`$
$`8[V_j\widehat{W}_{ij}]__1+8[V_j]__1[\widehat{W}_{ij}]__1+8v_1^j[V\widehat{W}_{ij}]__18v_1^j[V]__1[\widehat{W}_{ij}]__1`$
$`8[V^2V_i]__1+4[V^2]__1[V_i]__1+4[V]__1^2[V_i]__1`$
$`+{\displaystyle \frac{1}{2}}v_1^2v_1^i[V^2]__1{\displaystyle \frac{1}{2}}v_1^2v_1^i[V]__1^216v_1^j[V_iV_j]__1+16v_1^j[V_i]__1[V_j]__1),`$
$`F_1^i(F_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{1}{c^2}}\left(2[V_iV]__1+2[V]__1[_iV]__1\right)`$ (372)
$`+`$ $`{\displaystyle \frac{1}{c^4}}(v_1^2[V]__1[_iV]__1+v_1^2[V_iV]__18[V_j]__1[_iV_j]__1+8[V_j_iV_j]__1`$ (374)
$`+[V]__1^2[_iV]__1[V^2]__1[_iV]__1+2[V^2_iV]__12[V]__1[V_iV]__1)`$
$`+`$ $`{\displaystyle \frac{1}{c^6}}({\displaystyle \frac{1}{4}}v_1^4[V]__1[_iV]__1+{\displaystyle \frac{1}{4}}v_1^4[V_iV]__1`$ (389)
$`+{\displaystyle \frac{3}{2}}v_1^2[V]__1^2[_iV]__1{\displaystyle \frac{3}{2}}v_1^2[V^2]__1[_iV]__1+3v_1^2[V^2_iV]__13v_1^2[V]__1[V_iV]__1`$
$`+{\displaystyle \frac{8}{3}}[V]__1^3[_iV]__13[V]__1[V^2]__1[_iV]__1+{\displaystyle \frac{2}{3}}[V^3]__1[_iV]__1`$
$`3[V]__1^2[V_iV]__1+2[V^2]__1[V_iV]__1+2[V]__1[V^2_iV]__1{\displaystyle \frac{4}{3}}[V^3_iV]__1`$
$`16[_iV_j]__1[\widehat{R}_j]__1+16[_iV_j\widehat{R}_j]__116[V_j]__1[_i\widehat{R}_j]__1+16[V_j_i\widehat{R}_j]__1`$
$`+8[V]__1[_iV_j]__1[V_j]__1+8[V]__1[_iV_jV_j]__116[V_iV_jV_j]__1`$
$`+4[_iV]__1[V_j]__1[V_j]__1+4[_iV]__1[V_jV_j]__18[_iVV_jV_j]__1`$
$`12v_1^2[_iV_j]__1[V_j]__1+12v_1^2[_iV_jV_j]__1`$
$`+4v_1^j[V]__1^2[_iV_j]__1+4v_1^j[V^2]__1[_iV_j]__18v_1^j[V^2_iV_j]__1`$
$`+8v_1^j[V]__1[_iV]__1[V_j]__1+8v_1^j[V_iV]__1[V_j]__116v_1^j[V_iVV_j]__1`$
$`+8v_1^j[V_k]__1[_i\widehat{W}_{jk}]__18v_1^j[V_k_i\widehat{W}_{jk}]__1+8v_1^j[_iV_k]__1[\widehat{W}_{jk}]__18v_1^j[_iV_k\widehat{W}_{jk}]__1`$
$`4v_1^jv_1^k[V]__1[_i\widehat{W}_{jk}]__1+4v_1^jv_1^k[V_i\widehat{W}_{jk}]__1`$
$`4v_1^jv_1^k[_iV]__1[\widehat{W}_{jk}]__1+4v_1^jv_1^k[_iV\widehat{W}_{jk}]__1`$
$`+16v_1^jv_1^k[_iV_k]__1[V_j]__116v_1^jv_1^k[_iV_kV_j]__1`$
$`+8[\widehat{X}]__1[_iV]__18[\widehat{X}_iV]__1+8[V]__1[_i\widehat{X}]__18[V_i\widehat{X}]__1).`$
These formulas look complicated but are in fact rather simple to evaluate because they require only some lower-order post-Newtonian precision in the potentials, with notably all the difficult non-compact potentials needed at the Newtonian order only (hence the interest of separating out the problems as we did). Note that it is crucial here to employ the Lorentzian regularization $`[F]__1`$. The net result of this computation is
$`P_1^i(P_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{G^3m_1^2m_2}{c^6r_{12}^3}}\left({\displaystyle \frac{2}{5}}(n_{12}v_{12})n_{12}^i{\displaystyle \frac{2}{15}}v_{12}^i\right),`$ (391)
$`F_1^i(F_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{G^3m_1^2m_2}{c^6r_{12}^4}}([{\displaystyle \frac{241}{70}}{\displaystyle \frac{Gm_1}{r_{12}}}{\displaystyle \frac{51}{70}}{\displaystyle \frac{Gm_2}{r_{12}}}]n_{12}^i`$ (392)
$`+`$ $`{\displaystyle \frac{723}{28}}(n_{12}v_{12})^2n_{12}^i{\displaystyle \frac{723}{140}}v_{12}^2n_{12}^i{\displaystyle \frac{723}{70}}(n_{12}v_{12})v_{12}^i).`$ (393)
Therefore the supplement of acceleration linked to the non-distributivity is
$`a_1^i(a_1^i)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{G^3m_1^2m_2}{c^6r_{12}^4}}([{\displaystyle \frac{779}{210}}{\displaystyle \frac{Gm_1}{r_{12}}}{\displaystyle \frac{97}{210}}{\displaystyle \frac{Gm_2}{r_{12}}}]n_{12}^i`$ (394)
$`+`$ $`{\displaystyle \frac{779}{28}}(n_{12}v_{12})^2n_{12}^i{\displaystyle \frac{779}{140}}v_{12}^2n_{12}^i{\displaystyle \frac{779}{70}}(n_{12}v_{12})v_{12}^i).`$ (395)
Since only the relative velocity $`𝐯_{12}`$ is involved this part of the acceleration is Galilean invariant. Furthermore, it can be expressed in a simpler way by introducing an infinitesimal gauge transformation of the type (323). We pose
$`(ϵ_1)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{779}{420}}{\displaystyle \frac{m_1m_2^2}{m^3}},`$ (397)
$`(ϵ_1)_{_{\mathrm{distr}}}`$ $`=`$ $`{\displaystyle \frac{779}{420}}{\displaystyle \frac{m_1^2m_2}{m^3}},`$ (398)
and easily obtain
$$a_1^i(a_1^i)_{_{\mathrm{distr}}}=(\delta _\xi a_1^i)_{_{\mathrm{distr}}}+\frac{G^4m_1m_2^2}{c^6r_{12}^5}\left[\frac{97}{210}m_1+\frac{779}{210}m_2\right]n_{12}^i.$$
(399)
Thus, the only Physics brought about by the non-distributivity (i.e. which is not affected by a gauge transformation) is constituted by the quartic ($`G^4`$) term displayed in the right-hand side of (399).
## VII The 3PN equations of motion
### A Existence of the conserved energy
At present, the equations of motion are complete. We want now to look for the conserved energy associated with these equations at the 3PN order (considering of course only the conservative part of the equations, i.e. excluding the radiation reaction acceleration at the 2.5PN order). We shall see that the existence of an energy is not immediate, but requires the adjustment of a certain constant.
We proved in Section V that the equations of motion of body 1 depend on two arbitrary constants, which are the constant $`r_1^{}`$, tending to zero as we approach the particle 1 (but considered here as taking some finite non-zero value), and the constant $`s_2`$ associated with the Hadamard regularization near the other particle 2 \[see (4)\]. Similarly, the equations of body 2 depend on the constants $`r_2^{}`$ and $`s_1`$. All these constants appear inside the logarithms entering the equations of motion in harmonic coordinates. Gathering the results for the “logarithmic” part of the equations of body 1, we obtain
$`a_1^i`$ $`=`$ $`{\displaystyle \frac{44}{3}}{\displaystyle \frac{G^4m_1^3m_2}{c^6r_{12}^5}}n_{12}^i\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right){\displaystyle \frac{44}{3}}{\displaystyle \frac{G^4m_1m_2^3}{c^6r_{12}^5}}n_{12}^i\mathrm{ln}\left({\displaystyle \frac{r_{12}}{s_2}}\right)`$ (400)
$`+`$ $`{\displaystyle \frac{G^3m_1^2m_2}{c^6r_{12}^4}}\left[110(n_{12}v_{12})^2n_{12}^i22v_{12}^2n_{12}^i44(n_{12}v_{12})v_{12}^i\right]\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right)+\mathrm{},`$ (401)
where the dots indicate the terms which do not contain any logarithms. The terms shown in (400) contain the whole dependence of the acceleration of 1 over $`r_1^{}`$ and $`s_2`$; there are no other constants elsewhere. Notice that $`s_2`$ enters a single quartic-order term proportional to $`G^4m_1m_2^3`$. Now, most of the terms in (400) can in fact be gauged away. To see this, we apply the formula (333) with the particular choice
$`(ϵ_1)_{\mathrm{ln}}`$ $`=`$ $`{\displaystyle \frac{22}{3}}{\displaystyle \frac{m_1m_2^2}{m^3}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_2^{}}}\right),`$ (403)
$`(ϵ_2)_{\mathrm{ln}}`$ $`=`$ $`{\displaystyle \frac{22}{3}}{\displaystyle \frac{m_1^2m_2}{m^3}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right).`$ (404)
The corresponding transformation of the acceleration is (333), except that $`(ϵ_1)_{\mathrm{ln}}`$ and $`(ϵ_2)_{\mathrm{ln}}`$ depend on time through the orbital separation $`r_{12}`$, so in fact this formula should contain also some terms proportional to the time-derivatives of $`(ϵ_1)_{\mathrm{ln}}`$ and $`(ϵ_2)_{\mathrm{ln}}`$; but the point for us is that these extra terms are free of any logarithms. Therefore, modulo the dots indicating the logarithmic-free terms, we can write
$$a_1^i=(\delta _\xi a_1^i)_{\mathrm{ln}}\frac{44}{3}\frac{G^4m_1m_2^3}{c^6r_{12}^5}n_{12}^i\mathrm{ln}\left(\frac{r_2^{}}{s_2}\right)+\mathrm{},$$
(405)
where $`(\delta _\xi a_1^i)_{\mathrm{ln}}`$ denotes the coordinate change of the acceleration.
The term in (405) which is left out after this coordinate change depends only on the ratio between $`r_2^{}`$ and $`s_2`$ (similarly, in the equations of motion of body 2, we would find the ratio of $`r_1^{}`$ and $`s_1`$). This term is of the same type as the one in (364) giving the difference of accelerations, modulo a change of gauge, when different distributional derivatives are used. Notice that the constant $`r_2^{}`$ was originally absent from the equations of motion of 1, but has to be introduced in order to “remove” these logarithms by the coordinate transformation. Therefore, the only physical freedom remaining in the equations of motion is the yet unspecified constant $`\mathrm{ln}\left(\frac{r_2^{}}{s_2}\right)`$. Now we use this freedom to find a conserved energy associated with the equations of motion, which means a local-in-time functional $`E`$ of the trajectories and velocities of the two particles, i.e.
$$E=E[𝐲_1(t),𝐲_2(t);𝐯_1(t),𝐯_2(t)],$$
(406)
which is constant as a consequence of the 3PN equations of motion, i.e.
$$\frac{dE}{dt}v_1^i\frac{E}{y_1^i}+v_2^i\frac{E}{y_2^i}+a_1^i\frac{E}{v_1^i}+a_2^i\frac{E}{v_2^i}=\overline{𝒪}(7).$$
(407)
The accelerations $`𝐚_1`$ and $`𝐚_1`$ are to be replaced by the functionals of the positions and velocities given by the 3PN equations of motion. Our special notation for the remainder means a radiation-reaction term which is purely of order 2.5PN plus the neglected terms at 3.5PN; schematically $`\overline{𝒪}(7)=\frac{1}{c^5}F_5+𝒪(7)`$. See (489) below for the expression of the term $`\frac{1}{c^5}F_5`$. If an energy exists, the quantity $`m_1a_1^iv_1^i+m_2a_2^iv_2^i`$ must be a total time derivative. In practive, we look for a local-in-time functional $`D[𝐲_1,𝐲_2;𝐯_1,𝐯_2]`$ such that
$$m_1a_1^iv_1^i+m_2a_2^iv_2^i+\frac{dD}{dt}=\overline{𝒪}(7),$$
(408)
and we obtain the energy as $`E=\frac{1}{2}m_1𝐯_1^2+\frac{1}{2}m_2𝐯_2^2+D`$. Now, the computation with our 3PN equations of motion (obtained by means of, say, the particular derivative) shows that the quantity $`D`$ does not exist for any values of the constants $`\mathrm{ln}\left(r_2^{}/s_2\right)`$ and $`\mathrm{ln}\left(r_1^{}/s_1\right)`$. However, we find that this “nearly” works, because we can determine some $`\widehat{D}`$ such that
$$m_1a_1^iv_1^i+m_2a_2^iv_2^i+\frac{d\widehat{D}}{dt}=\frac{44}{3}\frac{G^4m_1^2m_2^2}{c^6r_{12}^5}\{m_2(n_{12}v_1)[\mathrm{ln}\left(\frac{r_2^{}}{s_2}\right)\frac{159}{308}]+12\}+\overline{𝒪}(7).$$
(409)
From the computation we obtain $`\widehat{D}`$ as a well-defined local functional of the positions and velocities of the particules \[containing in particular some logarithms $`\mathrm{ln}\left(r_{12}/r_1^{}\right)`$ and $`\mathrm{ln}\left(r_{12}/r_2^{}\right)`$\]. The right-hand side of (409) cannot be written, for generic values of $`\mathrm{ln}\left(r_2^{}/s_2\right)`$ and $`\mathrm{ln}\left(r_1^{}/s_1\right)`$, in the form of a total time-derivative. It would be possible, for this to be the case, to adopt the simplest choice that both these constants are numerically equal to $`\frac{159}{308}`$. However, this choice does not represent the most general solution for obtaining a total time-derivative. Indeed, nothing prevents $`\mathrm{ln}\left(r_1^{}/s_1\right)`$ and $`\mathrm{ln}\left(r_2^{}/s_2\right)`$ to depend also on the masses $`m_1`$ and $`m_2`$, and therefore such a dependence on the masses should in fact be mandatory (totalitarian principle). Since the regularization procedure followed in this paper is more mathematical than physical, we can be confident that no Physics will be overlooked only if at each step we obtain the most general solution allowed by the process. Unfortunately, the most general solution in this case contains an arbitrary parameter.
The necessary and sufficient condition for the right side of (409) to be a total time-derivative is that the factor of $`(n_{12}v_1)`$ in (409) be invariant by exchanging the particle’s labels 2 and 1, i.e.
$$m_2\left[\mathrm{ln}\left(\frac{r_2^{}}{s_2}\right)\frac{159}{308}\right]=m_1\left[\mathrm{ln}\left(\frac{r_1^{}}{s_1}\right)\frac{159}{308}\right].$$
(410)
Denoting by $`\lambda m`$ the common value of both sides of (410), where $`\lambda `$ is a constant and $`m=m_1+m_2`$, we obtain the most general solution as
$`\mathrm{ln}\left({\displaystyle \frac{r_2^{}}{s_2}}\right)`$ $`=`$ $`{\displaystyle \frac{159}{308}}+\lambda {\displaystyle \frac{m}{m_2}},`$ (412)
$`\mathrm{ln}\left({\displaystyle \frac{r_1^{}}{s_1}}\right)`$ $`=`$ $`{\displaystyle \frac{159}{308}}+\lambda {\displaystyle \frac{m}{m_1}}.`$ (413)
This $`\lambda `$ is a dimensionless quantity which is the same for the two particles 1 and 2. We now prove that $`\lambda `$ is necessarily a pure numerical constant, independent of the masses. Notice that the $`\lambda `$-term in (412) will yield a contribution to the acceleration of 1 which is, as concerns the dependence over the masses, of the type $`m_1m_2^2m\lambda `$ \[see (405)\]. If $`\lambda `$ depends on the masses, it must be a symmetric function of $`m_1`$ and $`m_2`$, and therefore it can be expressed solely in terms of the symmetric mass ratio $`\nu =\frac{m_1m_2}{m^2}`$. Suppose that $`\lambda =_{\mathrm{}}^+\mathrm{}\lambda _i\nu ^i`$, where the $`\lambda _i`$’s are numerical constants, so the $`\lambda `$-term in the acceleration of 1 is of the type $`m_1m_2^2m\lambda _i\nu ^i`$. First, we see that all the cases $`i1`$ are excluded because the equations of motion would not have the correct perturbative limit when $`\nu 0`$; for instance, in the case $`i=1`$, we get a term of the type $`m^3m_2`$ which tends to $`m_2^4`$ in this limit, and therefore modifies the geodesic motion of a test particle around a Schwarzschild black hole, which is of course excluded. Second, the cases $`i1`$, though they pass the simplest physical requirements, imply that the individual particle accelerations are no longer polynomials in the two individual masses $`m_1`$ and $`m_2`$, because of the appearance of inverse powers of the total mass $`m=m_1+m_2`$. For instance, the case $`i=1`$ leads to a term of the type $`\frac{m_1^2m_2^3}{m}`$. But we know that when doing a diagrammatic expansion of the $`N`$ body problem based on the post-Minkowskian expansion (see for the details of the method) that each successive diagram is a polynomial of the $`N`$ masses. Therefore, we exclude the possibility that some inverse powers of the total mass appear, and find, in conclusion, that $`\lambda `$ is a pure constant ($`\lambda =\lambda _0`$).
At last, we have succeeded in finding a conserved energy at the 3PN order by specifying an unknown logarithmic ratio, but at the price of having introduced an arbitrary purely numerical constant $`\lambda `$. The constant $`\lambda `$ will be left undetermined in the present work. So the final 3PN equations of motion we obtain in this paper, as well as the final 3PN energy, depend on the unknown parameter $`\lambda `$. The appearance of $`\lambda `$ suggests that the present formalism, based on a point-mass regularization, is physically incomplete. The resulting ambiguity is equivalent to the “static” ambiguity found by Jaranowski and Schäfer . It is probably linked to the fact that one can write the Einstein field equations into many different forms, which are all equivalent in the case of regular sources, but which are in general not equivalent in the case of point-particles because the distributional derivative does not obey the Leibniz rule. If we had chosen initially a different form of the field equations, the Leibniz terms we computed in Section VI, could have been different. More precisely, only that part of the Leibniz terms which is Galilean-invariant and consequently is not required by the Lorentz-invariance symmetry could change. But we have seen in (340) and (364) that the Galilean-invariant part of the Leibniz terms is precisely made of only one term, which is of the same type (proportional to $`G^4m_1m_2^3`$) as the term containing the constant $`\mathrm{ln}(r_2^{}/s_2)`$ \[see (405)\] that we have adjusted in the equations (412), resulting in the appearance of the constant $`\lambda `$. Thus, in agreement with Jaranowski and Schäfer , we might say that $`\lambda `$ encodes an ambiguity associated with the violation of the Leibniz rule by the distributional derivative. At a deeper level, this would mean that the ambiguity is a consequence of a theorem of Schwartz according to which it is impossible to define at once a multiplication of distributions which agrees with the ordinary product for continuous functions, and a derivation of distributions which satisfies the Leibniz rule and reduces to the ordinary derivative in the case of $`C^1`$ functions. If this explanation is correct, it is unlikely that the constant $`\lambda `$ could be determined within the present formalism.
We find by combining (VII A) and (409) that the dependence of the 3PN energy $`E`$ on $`\lambda `$ is
$$E=\widehat{E}\frac{11}{3}\lambda \frac{G^4m_1^2m_2^2m}{c^6r_{12}^4},$$
(414)
where $`\widehat{E}`$ does not depend on $`\lambda `$, while we obtain, using (405), that the acceleration writes
$$a_1^i=\widehat{a}_1^i\frac{44}{3}\lambda \frac{G^4m_1m_2^2m}{c^6r_{12}^5}n_{12}^i,$$
(415)
where similarly $`\widehat{a}_1^i`$ is independent of $`\lambda `$. On the other hand, the acceleration and energy depend also on the two constants $`r_1^{}`$ and $`r_2^{}`$, but from the previous discussion this is not a problem because $`r_1^{}`$ and $`r_2^{}`$ are associated with an arbitrariness in the choice of coordinates: the 3PN equations of motion contain the logarithms $`\mathrm{ln}\left(r_{12}/r_1^{}\right)`$ and $`\mathrm{ln}\left(r_{12}/r_2^{}\right)`$ which have been shown in (400) to be in the form of the gauge transformation associated with (VII A). In particular, the “constants” $`\mathrm{ln}r_1^{}`$ and $`\mathrm{ln}r_2^{}`$, which might be said to be formally infinite because $`r_1^{}`$ and $`r_2^{}`$ were tending to zero \[recall the discussion after (V D 2)\], will never appear in any physical result. Similarly, the dependence of the energy on the logarithms $`\mathrm{ln}\left(r_{12}/r_1^{}\right)`$ and $`\mathrm{ln}\left(r_{12}/r_2^{}\right)`$ is pure gauge. From (335) we get
$$E=\frac{22}{3}\frac{G^3m_1^3m_2}{c^6r_{12}^3}\left[\frac{Gm_2}{r_{12}}+3(n_{12}v_1)(n_{12}v_{12})(v_1v_{12})\right]\mathrm{ln}\left(\frac{r_{12}}{r_1^{}}\right)+12+\mathrm{},$$
(416)
where the dots denote the terms independent of logarithms \[this result can also be checked directly using (400)\].
Finally, to be more specific about the influence of the distributional derivative, notice that the solution we have obtained in (VII A) corresponds to the “particular” distributional derivative defined by (10). If one uses the “correct” derivative (11)-(12) instead, with the value $`K=\frac{41}{160}`$ we have obtained in (363) from the Lorentz invariance, we obtain the same equation to be solved as (409) but with the rational fraction $`+\frac{783}{3080}`$ instead of $`\frac{159}{308}`$. This is easily seen thanks to (364). So, the solution becomes in this case
$`\mathrm{ln}\left({\displaystyle \frac{r_2^{}}{s_2}}\right)`$ $`=`$ $`{\displaystyle \frac{783}{3080}}+\lambda {\displaystyle \frac{m}{m_2}},`$ (418)
$`\mathrm{ln}\left({\displaystyle \frac{r_1^{}}{s_1}}\right)`$ $`=`$ $`{\displaystyle \frac{783}{3080}}+\lambda {\displaystyle \frac{m}{m_1}}.`$ (419)
Replacing this into the equations of motion (and associated energy), it is then clear that they are physically the same as those computed with the other derivative, because they differ by the mere change of gauge,
$$a_{1}^{i}{}_{|_{{\scriptscriptstyle \frac{41}{160}}}}{}^{}a_1^i=\delta _\xi a_{1}^{i}{}_{|_{{\scriptscriptstyle \frac{41}{160}}}}{}^{},$$
(420)
that we obtained in (364). We give it here thoroughtly for completeness:
$`\delta _\xi a_{1}^{i}{}_{|_{{\scriptscriptstyle \frac{41}{160}}}}{}^{}`$ $`=`$ $`{\displaystyle \frac{13}{25}}{\displaystyle \frac{G^4m_1m_2}{c^6r_{12}^5}}\left(m_1^2m_2^2\right)n_{12}^i`$ (421)
$`+`$ $`{\displaystyle \frac{13}{50}}{\displaystyle \frac{G^3m_2^3}{c^6r_{12}^4}}\left[15(n_{12}v_{12})^2n_{12}^i+3v_{12}^2n_{12}^i+6(n_{12}v_{12})v_{12}^i\right].`$ (422)
### B End results
We present the 3PN equations of motion of the particle 1 in harmonic coordinates, which are obtained by summing up all the contributions of the potentials computed in Sections IV and V, as well as the pieces due to the non-distributivity and the Leibniz terms (see Section VI). The equations depend on two gauge constants $`r_1^{}`$ and $`r_2^{}`$ through some logarithms, and on one unknown purely numerical coefficient $`\lambda `$. The equations of the particle 2 are obtained by exchanging all the labels $`12`$.
$`a_1^i`$ $`=`$ $`{\displaystyle \frac{Gm_2n_{12}^i}{r_{12}^2}}`$ (423)
$`+`$ $`{\displaystyle \frac{1}{c^2}}\{[{\displaystyle \frac{5G^2m_1m_2}{r_{12}^3}}+{\displaystyle \frac{4G^2m_2^2}{r_{12}^3}}+{\displaystyle \frac{Gm_2}{r_{12}^2}}({\displaystyle \frac{3}{2}}(n_{12}v_2)^2v_1^2+4(v_1v_2)2v_2^2)]n_{12}^i`$ (425)
$`+{\displaystyle \frac{Gm_2}{r_{12}^2}}(4(n_{12}v_1)3(n_{12}v_2))(v_1^iv_2^i)\}`$
$`+`$ $`{\displaystyle \frac{1}{c^4}}\{[{\displaystyle \frac{57G^3m_1^2m_2}{4r_{12}^4}}{\displaystyle \frac{69G^3m_1m_2^2}{2r_{12}^4}}{\displaystyle \frac{9G^3m_2^3}{r_{12}^4}}+{\displaystyle \frac{Gm_2}{r_{12}^2}}({\displaystyle \frac{15}{8}}(n_{12}v_2)^4+{\displaystyle \frac{3}{2}}(n_{12}v_2)^2v_1^2`$ (432)
$`6(n_{12}v_2)^2(v_1v_2)2(v_1v_2)^2+{\displaystyle \frac{9}{2}}(n_{12}v_2)^2v_2^2+4(v_1v_2)v_2^22v_2^4)`$
$`+{\displaystyle \frac{G^2m_1m_2}{r_{12}^3}}({\displaystyle \frac{39}{2}}(n_{12}v_1)^239(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{17}{2}}(n_{12}v_2)^2{\displaystyle \frac{15}{4}}v_1^2{\displaystyle \frac{5}{2}}(v_1v_2)`$
$`+{\displaystyle \frac{5}{4}}v_2^2)+{\displaystyle \frac{G^2m_2^2}{r_{12}^3}}(2(n_{12}v_1)^24(n_{12}v_1)(n_{12}v_2)6(n_{12}v_2)^28(v_1v_2)+4v_2^2)]n_{12}^i`$
$`+[{\displaystyle \frac{G^2m_2^2}{r_{12}^3}}(2(n_{12}v_1)2(n_{12}v_2))+{\displaystyle \frac{G^2m_1m_2}{r_{12}^3}}({\displaystyle \frac{63}{4}}(n_{12}v_1)+{\displaystyle \frac{55}{4}}(n_{12}v_2))`$
$`+{\displaystyle \frac{Gm_2}{r_{12}^2}}(6(n_{12}v_1)(n_{12}v_2)^2+{\displaystyle \frac{9}{2}}(n_{12}v_2)^3+(n_{12}v_2)v_1^24(n_{12}v_1)(v_1v_2)`$
$`+4(n_{12}v_2)(v_1v_2)+4(n_{12}v_1)v_2^25(n_{12}v_2)v_2^2)](v_1^iv_2^i)\}`$
$`+`$ $`{\displaystyle \frac{1}{c^5}}\{[{\displaystyle \frac{208G^3m_1m_2^2}{15r_{12}^4}}((n_{12}v_1)(n_{12}v_2)){\displaystyle \frac{24G^3m_1^2m_2}{5r_{12}^4}}((n_{12}v_1)(n_{12}v_2))`$ (435)
$`+{\displaystyle \frac{12G^2m_1m_2}{5r_{12}^3}}((n_{12}v_1)(n_{12}v_2))[v_1^22(v_1v_2)+v_2^2]]n_{12}^i+[{\displaystyle \frac{8G^3m_1^2m_2}{5r_{12}^4}}`$
$`{\displaystyle \frac{32G^3m_1m_2^2}{5r_{12}^4}}{\displaystyle \frac{4G^2m_1m_2}{5r_{12}^3}}[v_1^22(v_1v_2)+v_2^2]](v_1^iv_2^i)\}`$
$`+`$ $`{\displaystyle \frac{1}{c^6}}\{[{\displaystyle \frac{Gm_2}{r_{12}^2}}({\displaystyle \frac{35}{16}}(n_{12}v_2)^6{\displaystyle \frac{15}{8}}(n_{12}v_2)^4v_1^2+{\displaystyle \frac{15}{2}}(n_{12}v_2)^4(v_1v_2)+3(n_{12}v_2)^2(v_1v_2)^2`$ (464)
$`{\displaystyle \frac{15}{2}}(n_{12}v_2)^4v_2^2+{\displaystyle \frac{3}{2}}(n_{12}v_2)^2v_1^2v_2^212(n_{12}v_2)^2(v_1v_2)v_2^22(v_1v_2)^2v_2^2`$
$`+{\displaystyle \frac{15}{2}}(n_{12}v_2)^2v_2^4+4(v_1v_2)v_2^42v_2^6)+{\displaystyle \frac{G^2m_1m_2}{r_{12}^3}}({\displaystyle \frac{171}{8}}(n_{12}v_1)^4`$
$`+{\displaystyle \frac{171}{2}}(n_{12}v_1)^3(n_{12}v_2){\displaystyle \frac{723}{4}}(n_{12}v_1)^2(n_{12}v_2)^2+{\displaystyle \frac{383}{2}}(n_{12}v_1)(n_{12}v_2)^3`$
$`{\displaystyle \frac{455}{8}}(n_{12}v_2)^4+{\displaystyle \frac{229}{4}}(n_{12}v_1)^2v_1^2{\displaystyle \frac{205}{2}}(n_{12}v_1)(n_{12}v_2)v_1^2+{\displaystyle \frac{191}{4}}(n_{12}v_2)^2v_1^2{\displaystyle \frac{91}{8}}v_1^4`$
$`{\displaystyle \frac{229}{2}}(n_{12}v_1)^2(v_1v_2)+244(n_{12}v_1)(n_{12}v_2)(v_1v_2){\displaystyle \frac{225}{2}}(n_{12}v_2)^2(v_1v_2)`$
$`+{\displaystyle \frac{91}{2}}v_1^2(v_1v_2){\displaystyle \frac{177}{4}}(v_1v_2)^2+{\displaystyle \frac{229}{4}}(n_{12}v_1)^2v_2^2{\displaystyle \frac{283}{2}}(n_{12}v_1)(n_{12}v_2)v_2^2`$
$`+{\displaystyle \frac{259}{4}}(n_{12}v_2)^2v_2^2{\displaystyle \frac{91}{4}}v_1^2v_2^2+43(v_1v_2)v_2^2{\displaystyle \frac{81}{8}}v_2^4)+{\displaystyle \frac{G^2m_2^2}{r_{12}^3}}(6(n_{12}v_1)^2(n_{12}v_2)^2`$
$`+12(n_{12}v_1)(n_{12}v_2)^3+6(n_{12}v_2)^4+4(n_{12}v_1)(n_{12}v_2)(v_1v_2)+12(n_{12}v_2)^2(v_1v_2)`$
$`+4(v_1v_2)^24(n_{12}v_1)(n_{12}v_2)v_2^212(n_{12}v_2)^2v_2^28(v_1v_2)v_2^2+4v_2^4)`$
$`+{\displaystyle \frac{G^3m_2^3}{r_{12}^4}}\left((n_{12}v_1)^2+2(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{43}{2}}(n_{12}v_2)^2+18(v_1v_2)9v_2^2\right)`$
$`+{\displaystyle \frac{G^3m_1m_2^2}{r_{12}^4}}({\displaystyle \frac{415}{8}}(n_{12}v_1)^2{\displaystyle \frac{375}{4}}(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{1113}{8}}(n_{12}v_2)^2`$
$`{\displaystyle \frac{615}{64}}((n_{12}v_1)(n_{12}v_2))^2\pi ^2+18v_1^2+{\displaystyle \frac{123}{64}}\pi ^2(v_1v_2)^2+33(v_1v_2){\displaystyle \frac{33}{2}}v_2^2)`$
$`+{\displaystyle \frac{G^3m_1^2m_2}{r_{12}^4}}({\displaystyle \frac{45887}{168}}(n_{12}v_1)^2+{\displaystyle \frac{24025}{42}}(n_{12}v_1)(n_{12}v_2){\displaystyle \frac{10469}{42}}(n_{12}v_2)^2`$
$`+{\displaystyle \frac{48197}{840}}v_1^2{\displaystyle \frac{36227}{420}}(v_1v_2)+{\displaystyle \frac{36227}{840}}v_2^2+110\left((n_{12}v_1)(n_{12}v_2)\right)^2\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right)`$
$`22(v_1v_2)^2\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right))+{\displaystyle \frac{16G^4m_2^4}{r_{12}^5}}+{\displaystyle \frac{G^4m_1^2m_2^2}{r_{12}^5}}({\displaystyle \frac{34763}{210}}{\displaystyle \frac{44\lambda }{3}}{\displaystyle \frac{41}{16}}\pi ^2)`$
$`+{\displaystyle \frac{G^4m_1^3m_2}{r_{12}^5}}({\displaystyle \frac{3187}{1260}}+{\displaystyle \frac{44}{3}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right))+{\displaystyle \frac{G^4m_1m_2^3}{r_{12}^5}}({\displaystyle \frac{10478}{63}}{\displaystyle \frac{44\lambda }{3}}{\displaystyle \frac{41}{16}}\pi ^2`$
$`{\displaystyle \frac{44}{3}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_2^{}}}\right))]n_{12}^i+[{\displaystyle \frac{Gm_2}{r_{12}^2}}({\displaystyle \frac{15}{2}}(n_{12}v_1)(n_{12}v_2)^4{\displaystyle \frac{45}{8}}(n_{12}v_2)^5{\displaystyle \frac{3}{2}}(n_{12}v_2)^3v_1^2`$
$`+6(n_{12}v_1)(n_{12}v_2)^2(v_1v_2)6(n_{12}v_2)^3(v_1v_2)2(n_{12}v_2)(v_1v_2)^2`$
$`12(n_{12}v_1)(n_{12}v_2)^2v_2^2+12(n_{12}v_2)^3v_2^2+(n_{12}v_2)v_1^2v_2^24(n_{12}v_1)(v_1v_2)v_2^2`$
$`+8(n_{12}v_2)(v_1v_2)v_2^2+4(n_{12}v_1)v_2^47(n_{12}v_2)v_2^4)`$
$`+{\displaystyle \frac{G^2m_2^2}{r_{12}^3}}(2(n_{12}v_1)^2(n_{12}v_2)+8(n_{12}v_1)(n_{12}v_2)^2`$
$`+2(n_{12}v_2)^3+2(n_{12}v_1)(v_1v_2)+4(n_{12}v_2)(v_1v_2)2(n_{12}v_1)v_2^24(n_{12}v_2)v_2^2)`$
$`+{\displaystyle \frac{G^2m_1m_2}{r_{12}^3}}({\displaystyle \frac{243}{4}}(n_{12}v_1)^3+{\displaystyle \frac{565}{4}}(n_{12}v_1)^2(n_{12}v_2){\displaystyle \frac{269}{4}}(n_{12}v_1)(n_{12}v_2)^2`$
$`{\displaystyle \frac{95}{12}}(n_{12}v_2)^3+{\displaystyle \frac{207}{8}}(n_{12}v_1)v_1^2{\displaystyle \frac{137}{8}}(n_{12}v_2)v_1^236(n_{12}v_1)(v_1v_2)`$
$`+{\displaystyle \frac{27}{4}}(n_{12}v_2)(v_1v_2)+{\displaystyle \frac{81}{8}}(n_{12}v_1)v_2^2+{\displaystyle \frac{83}{8}}(n_{12}v_2)v_2^2)`$
$`+{\displaystyle \frac{G^3m_2^3}{r_{12}^4}}(4(n_{12}v_1)+5(n_{12}v_2))+{\displaystyle \frac{G^3m_1m_2^2}{r_{12}^4}}({\displaystyle \frac{307}{8}}(n_{12}v_1)+{\displaystyle \frac{479}{8}}(n_{12}v_2)`$
$`+{\displaystyle \frac{123}{32}}((n_{12}v_1)(n_{12}v_2))\pi ^2)+{\displaystyle \frac{G^3m_1^2m_2}{r_{12}^4}}({\displaystyle \frac{31397}{420}}(n_{12}v_1){\displaystyle \frac{36227}{420}}(n_{12}v_2)`$
$`44((n_{12}v_1)(n_{12}v_2))\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right))](v_1^iv_2^i)\}+𝒪(7).`$
These equations are in full agreement with the known results valid up to the 2.5PN order . They have the correct perturbative limit given by the geodesics of the Schwarzschild metric at the 3PN order. Most importantly, the equations are invariant under Lorentz transformations (developed to 3PN order); this can be checked using for instance the formulas developed in . Finally, as we have seen previously, the equations of motion admit a conserved energy at the 3PN order. The study of the Lagrangian (and Hamiltonian) formulation of these equations is reported in a separate work . The energy is given by
$`E`$ $`=`$ $`{\displaystyle \frac{m_1v_1^2}{2}}{\displaystyle \frac{Gm_1m_2}{2r_{12}}}`$ (465)
$`+`$ $`{\displaystyle \frac{1}{c^2}}\left\{{\displaystyle \frac{G^2m_1^2m_2}{2r_{12}^2}}+{\displaystyle \frac{3m_1v_1^4}{8}}+{\displaystyle \frac{Gm_1m_2}{r_{12}}}\left({\displaystyle \frac{1}{4}}(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{3}{2}}v_1^2{\displaystyle \frac{7}{4}}(v_1v_2)\right)\right\}`$ (466)
$`+`$ $`{\displaystyle \frac{1}{c^4}}\{{\displaystyle \frac{G^3m_1^3m_2}{2r_{12}^3}}{\displaystyle \frac{19G^3m_1^2m_2^2}{8r_{12}^3}}+{\displaystyle \frac{5m_1v_1^6}{16}}+{\displaystyle \frac{Gm_1m_2}{r_{12}}}({\displaystyle \frac{3}{8}}(n_{12}v_1)^3(n_{12}v_2)`$ (470)
$`+{\displaystyle \frac{3}{16}}(n_{12}v_1)^2(n_{12}v_2)^2{\displaystyle \frac{9}{8}}(n_{12}v_1)(n_{12}v_2)v_1^2{\displaystyle \frac{13}{8}}(n_{12}v_2)^2v_1^2+{\displaystyle \frac{21}{8}}v_1^4`$
$`+{\displaystyle \frac{13}{8}}(n_{12}v_1)^2(v_1v_2)+{\displaystyle \frac{3}{4}}(n_{12}v_1)(n_{12}v_2)(v_1v_2){\displaystyle \frac{55}{8}}v_1^2(v_1v_2)+{\displaystyle \frac{17}{8}}(v_1v_2)^2+{\displaystyle \frac{31}{16}}v_1^2v_2^2)`$
$`+{\displaystyle \frac{G^2m_1^2m_2}{r_{12}^2}}({\displaystyle \frac{29}{4}}(n_{12}v_1)^2{\displaystyle \frac{13}{4}}(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{1}{2}}(n_{12}v_2)^2{\displaystyle \frac{3}{2}}v_1^2+{\displaystyle \frac{7}{4}}v_2^2)\}`$
$`+`$ $`{\displaystyle \frac{1}{c^6}}\{{\displaystyle \frac{35m_1v_1^8}{128}}+{\displaystyle \frac{Gm_1m_2}{r_{12}}}({\displaystyle \frac{5}{16}}(n_{12}v_1)^5(n_{12}v_2){\displaystyle \frac{5}{16}}(n_{12}v_1)^4(n_{12}v_2)^2`$ (487)
$`{\displaystyle \frac{5}{32}}(n_{12}v_1)^3(n_{12}v_2)^3+{\displaystyle \frac{19}{16}}(n_{12}v_1)^3(n_{12}v_2)v_1^2+{\displaystyle \frac{15}{16}}(n_{12}v_1)^2(n_{12}v_2)^2v_1^2`$
$`+{\displaystyle \frac{3}{4}}(n_{12}v_1)(n_{12}v_2)^3v_1^2+{\displaystyle \frac{19}{16}}(n_{12}v_2)^4v_1^2{\displaystyle \frac{21}{16}}(n_{12}v_1)(n_{12}v_2)v_1^42(n_{12}v_2)^2v_1^4+{\displaystyle \frac{55}{16}}v_1^6`$
$`{\displaystyle \frac{19}{16}}(n_{12}v_1)^4(v_1v_2)(n_{12}v_1)^3(n_{12}v_2)(v_1v_2){\displaystyle \frac{15}{32}}(n_{12}v_1)^2(n_{12}v_2)^2(v_1v_2)`$
$`+{\displaystyle \frac{45}{16}}(n_{12}v_1)^2v_1^2(v_1v_2)+{\displaystyle \frac{5}{4}}(n_{12}v_1)(n_{12}v_2)v_1^2(v_1v_2)+{\displaystyle \frac{11}{4}}(n_{12}v_2)^2v_1^2(v_1v_2)`$
$`{\displaystyle \frac{139}{16}}v_1^4(v_1v_2){\displaystyle \frac{3}{4}}(n_{12}v_1)^2(v_1v_2)^2+{\displaystyle \frac{5}{16}}(n_{12}v_1)(n_{12}v_2)(v_1v_2)^2+{\displaystyle \frac{41}{8}}v_1^2(v_1v_2)^2`$
$`+{\displaystyle \frac{1}{16}}(v_1v_2)^3{\displaystyle \frac{45}{16}}(n_{12}v_1)^2v_1^2v_2^2{\displaystyle \frac{23}{32}}(n_{12}v_1)(n_{12}v_2)v_1^2v_2^2+{\displaystyle \frac{79}{16}}v_1^4v_2^2{\displaystyle \frac{161}{32}}v_1^2(v_1v_2)v_2^2)`$
$`+{\displaystyle \frac{G^2m_1^2m_2}{r_{12}^2}}({\displaystyle \frac{49}{8}}(n_{12}v_1)^4+{\displaystyle \frac{75}{8}}(n_{12}v_1)^3(n_{12}v_2){\displaystyle \frac{187}{8}}(n_{12}v_1)^2(n_{12}v_2)^2`$
$`+{\displaystyle \frac{247}{24}}(n_{12}v_1)(n_{12}v_2)^3+{\displaystyle \frac{49}{8}}(n_{12}v_1)^2v_1^2+{\displaystyle \frac{81}{8}}(n_{12}v_1)(n_{12}v_2)v_1^2{\displaystyle \frac{21}{4}}(n_{12}v_2)^2v_1^2+{\displaystyle \frac{11}{2}}v_1^4`$
$`{\displaystyle \frac{15}{2}}(n_{12}v_1)^2(v_1v_2){\displaystyle \frac{3}{2}}(n_{12}v_1)(n_{12}v_2)(v_1v_2)+{\displaystyle \frac{21}{4}}(n_{12}v_2)^2(v_1v_2)27v_1^2(v_1v_2)`$
$`+{\displaystyle \frac{55}{2}}(v_1v_2)^2+{\displaystyle \frac{49}{4}}(n_{12}v_1)^2v_2^2{\displaystyle \frac{27}{2}}(n_{12}v_1)(n_{12}v_2)v_2^2+{\displaystyle \frac{3}{4}}(n_{12}v_2)^2v_2^2+{\displaystyle \frac{55}{4}}v_1^2v_2^2`$
$`28(v_1v_2)v_2^2+{\displaystyle \frac{135}{16}}v_2^4)+{\displaystyle \frac{3G^4m_1^4m_2}{8r_{12}^4}}+{\displaystyle \frac{G^4m_1^3m_2^2}{r_{12}^4}}({\displaystyle \frac{5809}{280}}{\displaystyle \frac{11}{3}}\lambda {\displaystyle \frac{22}{3}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right))`$
$`+{\displaystyle \frac{G^3m_1^2m_2^2}{r_{12}^3}}({\displaystyle \frac{547}{12}}(n_{12}v_1)^2{\displaystyle \frac{3115}{48}}(n_{12}v_1)(n_{12}v_2){\displaystyle \frac{123}{64}}(n_{12}v_1)^2\pi ^2`$
$`+{\displaystyle \frac{123}{64}}(n_{12}v_1)(n_{12}v_2)\pi ^2{\displaystyle \frac{575}{18}}v_1^2+{\displaystyle \frac{41}{64}}\pi ^2v_1^2+{\displaystyle \frac{4429}{144}}(v_1v_2){\displaystyle \frac{41}{64}}\pi ^2(v_1v_2))`$
$`+{\displaystyle \frac{G^3m_1^3m_2}{r_{12}^3}}({\displaystyle \frac{44627}{840}}(n_{12}v_1)^2+{\displaystyle \frac{32027}{840}}(n_{12}v_1)(n_{12}v_2)+{\displaystyle \frac{3}{2}}(n_{12}v_2)^2+{\displaystyle \frac{24187}{2520}}v_1^2`$
$`{\displaystyle \frac{27967}{2520}}(v_1v_2)+{\displaystyle \frac{5}{4}}v_2^2+22(n_{12}v_1)^2\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right)22(n_{12}v_1)(n_{12}v_2)\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right)`$
$`{\displaystyle \frac{22}{3}}v_1^2\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right)+{\displaystyle \frac{22}{3}}(v_1v_2)\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_1^{}}}\right))\}+12+𝒪(7).`$
This energy is conserved in the sense that its time-derivative computed with the 3PN equations of motion equals the radiation reaction effect at the 2.5PN order, namely
$`{\displaystyle \frac{dE}{dt}}`$ $`=`$ $`{\displaystyle \frac{4}{5}}{\displaystyle \frac{G^2m_1^2m_2}{c^5r_{12}^3}}[(v_1v_{12})(v_{12}^2+2{\displaystyle \frac{Gm_1}{r_{12}}}8{\displaystyle \frac{Gm_2}{r_{12}}})`$ (489)
$`+(n_{12}v_1)(n_{12}v_{12})(3v_{12}^26{\displaystyle \frac{Gm_1}{r_{12}}}+{\displaystyle \frac{52}{3}}{\displaystyle \frac{Gm_2}{r_{12}}})]+12+𝒪(7).`$
The rather complicated expressions (423)-(465) simplify drastically in the case where the orbit is circular \[apart from the gradual inspiral associated with the balance equation (489)\] and where we place ourselves in the center-of-mass frame. The circular orbit corresponds to the physical situation of the inspiralling compact binaries which motivate our work . Here, we give the result concerning circular orbits without proof (see also ). The relative acceleration reads
$$\frac{d𝐯_{12}}{dt}=\omega ^2𝐲_{12}+𝐅_{\mathrm{reac}}+𝒪(7),$$
(490)
where $`𝐅_{\mathrm{reac}}`$ is the standard radiation-reaction force in harmonic coordinates,
$$𝐅_{\mathrm{reac}}=\frac{32}{5}\frac{G^3m^3\nu }{c^5r_{12}^4}𝐯_{12}$$
(491)
($`\nu =\frac{m_1m_2}{m^2}`$ being the symmetric mass ratio), and where the orbital frequency $`\omega `$ of the relative circular motion is given to the 3PN order by
$`\omega ^2`$ $`=`$ $`{\displaystyle \frac{Gm}{r_{12}^3}}\{1+(3+\nu )\gamma +(6+{\displaystyle \frac{41}{4}}\nu +\nu ^2)\gamma ^2`$ (492)
$`+`$ $`(10+[{\displaystyle \frac{67759}{840}}+{\displaystyle \frac{41}{64}}\pi ^2+22\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_0^{}}}\right)+{\displaystyle \frac{44}{3}}\lambda ]\nu +{\displaystyle \frac{19}{2}}\nu ^2+\nu ^3)\gamma ^3+𝒪(8)\}.`$ (493)
The post-Newtonian parameter is defined by $`\gamma =\frac{Gm}{r_{12}c^2}`$, and we recall that $`r_{12}=|𝐲_1𝐲_2|`$ is the orbital separation in harmonic coordinates. The constant $`r_0^{}`$ appearing in the logarithm is related to the two constants $`r_1^{}`$ and $`r_2^{}`$ by
$$\mathrm{ln}r_0^{}=\frac{m_1}{m}\mathrm{ln}r_1^{}+\frac{m_2}{m}\mathrm{ln}r_2^{}.$$
(494)
The 3PN energy $`E`$ in the center of mass of the particles, which is such that $`\frac{dE}{dt}=0`$ as a consequence of the conservative equations of motion, is obtained as
$`E`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mu c^2\gamma \{1+({\displaystyle \frac{7}{4}}+{\displaystyle \frac{1}{4}}\nu )\gamma +({\displaystyle \frac{7}{8}}+{\displaystyle \frac{49}{8}}\nu +{\displaystyle \frac{1}{8}}\nu ^2)\gamma ^2`$ (495)
$`+`$ $`({\displaystyle \frac{235}{64}}+[{\displaystyle \frac{106301}{6720}}{\displaystyle \frac{123}{64}}\pi ^2+{\displaystyle \frac{22}{3}}\mathrm{ln}\left({\displaystyle \frac{r_{12}}{r_0^{}}}\right){\displaystyle \frac{22}{3}}\lambda ]\nu +{\displaystyle \frac{27}{32}}\nu ^2+{\displaystyle \frac{5}{64}}\nu ^3)\gamma ^3+𝒪(8)\}.`$ (496)
The invariant 3PN energy follows from the replacement of the post-Newtonian parameter $`\gamma `$ by its expression in terms of the frequency $`\omega `$ as deduced from computing the inverse of (492). We find
$`E`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mu c^2x\{1+({\displaystyle \frac{3}{4}}{\displaystyle \frac{1}{12}}\nu )x+({\displaystyle \frac{27}{8}}+{\displaystyle \frac{19}{8}}\nu {\displaystyle \frac{1}{24}}\nu ^2)x^2`$ (498)
$`+`$ $`({\displaystyle \frac{675}{64}}+[{\displaystyle \frac{209323}{4032}}{\displaystyle \frac{205}{96}}\pi ^2{\displaystyle \frac{110}{9}}\lambda ]\nu {\displaystyle \frac{155}{96}}\nu ^2{\displaystyle \frac{35}{5184}}\nu ^3)x^3+𝒪(8)\},`$ (499)
where the parameter $`x`$ is defined by
$$x=\left(\frac{Gm\omega }{c^3}\right)^{2/3}.$$
(500)
Note that the logarithm disappeared from the invariant expression of the energy (498), in agreement with the fact that it is pure gauge. However, the constant $`\lambda `$ stays in the final formula; the static ambiguity constant $`\omega _s`$ of Jaranowski and Schäfer is related to it by $`\omega _s=\frac{11}{3}\lambda \frac{1987}{840}`$ (see ).
###### Acknowledgements.
Most of the algebraic manipulations reported in this article have been done with the help of the softwares Mathematica and MathTensor. One of us (G.F.) would like to acknowledge Jean-Marc Conan and Vincent Michau for letting him complete this work in the context of his military service.
## A Sum of Leibniz terms
In this appendix we give the sum of all the terms of the type $`\delta _{\mathrm{Leibniz}}T`$ introduced by (68) that we have encountered during the process of simplification of the 3PN potentials. The reduction of these terms using the distributional derivative is done in Section VI.
$`\delta _{\mathrm{Leibniz}}\left({\displaystyle \frac{h^{00}+h^{ii}}{2}}\right)=\mathrm{}_{}^1\{{\displaystyle \frac{8}{c^4}}(_i\mathrm{Pf}V_i\mathrm{Pf}V+V\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V^2){\displaystyle \frac{1}{c^2}}(_t\mathrm{Pf}V)^2)`$ (A1)
$`{\displaystyle \frac{8}{c^6}}\left(_i\mathrm{Pf}V_i\mathrm{Pf}\widehat{W}+{\displaystyle \frac{1}{2}}V\mathrm{}\mathrm{Pf}\widehat{W}+{\displaystyle \frac{1}{2}}\widehat{W}\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V\widehat{W}){\displaystyle \frac{1}{c^2}}_t\mathrm{Pf}V_t\mathrm{Pf}\widehat{W}\right)`$ (A2)
$`{\displaystyle \frac{32}{c^6}}\left(V_i\mathrm{Pf}V_i\mathrm{Pf}V+{\displaystyle \frac{1}{2}}V^2\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{c^2}}V(_t\mathrm{Pf}V)^2{\displaystyle \frac{1}{6}}\mathrm{}(\mathrm{Pf}V^3)\right)`$ (A3)
$`{\displaystyle \frac{32}{c^6}}\left(_i\mathrm{Pf}V_i(\mathrm{Pf}V^2)2V_i\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A4)
$`{\displaystyle \frac{16}{c^8}}\left(V_t^2(\mathrm{Pf}V^2)2V(_t\mathrm{Pf}V)^22V^2_t^2\mathrm{Pf}V\right)`$ (A5)
$`{\displaystyle \frac{16}{c^8}}\left(_t\mathrm{Pf}V_t(\mathrm{Pf}V^2)2V(_t\mathrm{Pf}V)^2\right)`$ (A6)
$`{\displaystyle \frac{32}{c^8}}\left(V_i_t_i(\mathrm{Pf}V^2)2VV_i_t_i\mathrm{Pf}V2V_i_t\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A7)
$`{\displaystyle \frac{64}{c^8}}\left(V^2_i\mathrm{Pf}V_i\mathrm{Pf}V+{\displaystyle \frac{1}{3}}V^3\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{12}}\mathrm{}(\mathrm{Pf}V^4)\right)`$ (A8)
$`{\displaystyle \frac{32}{c^8}}\left(_i(\mathrm{Pf}V^2)_i(\mathrm{Pf}V^2)4V^2_i\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A9)
$`+{\displaystyle \frac{144}{c^8}}\left(V_i\mathrm{Pf}V_i(\mathrm{Pf}V^2)2V^2_i\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A10)
$`{\displaystyle \frac{128}{3c^8}}\left(_i\mathrm{Pf}V_i(\mathrm{Pf}V^3)3V^2_i\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A11)
$`+{\displaystyle \frac{64}{c^8}}\left(_i\mathrm{Pf}V_j_j(\mathrm{Pf}VV_i)V_i\mathrm{Pf}V_j_j\mathrm{Pf}V_iV_i_i\mathrm{Pf}V_j_j\mathrm{Pf}V\right)`$ (A12)
$`{\displaystyle \frac{16}{c^8}}\left(\widehat{W}_{ij}_{ij}(\mathrm{Pf}V^2)2V\widehat{W}_{ij}_{ij}\mathrm{Pf}V2\widehat{W}_{ij}_i\mathrm{Pf}V_j\mathrm{Pf}V\right)`$ (A13)
$`{\displaystyle \frac{4}{c^8}}\left(_i\mathrm{Pf}\widehat{W}_i\mathrm{Pf}\widehat{W}+\widehat{W}\mathrm{}\mathrm{Pf}\widehat{W}{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}\widehat{W}^2)\right)`$ (A14)
$`{\displaystyle \frac{16}{c^8}}\left(\widehat{W}_i\mathrm{Pf}V_i\mathrm{Pf}V+2V_i\mathrm{Pf}V_i\mathrm{Pf}\widehat{W}+V\widehat{W}\mathrm{}\mathrm{Pf}V+{\displaystyle \frac{1}{2}}V^2\mathrm{}\mathrm{Pf}\widehat{W}{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V^2\widehat{W})\right)`$ (A15)
$`+{\displaystyle \frac{8}{c^8}}\left(\widehat{W}\mathrm{}(\mathrm{Pf}V^2)2V\widehat{W}\mathrm{}\mathrm{Pf}V2\widehat{W}_i\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A16)
$`{\displaystyle \frac{16}{c^8}}\left(_i\mathrm{Pf}\widehat{W}_i(\mathrm{Pf}V^2)2V_i\mathrm{Pf}\widehat{W}_i\mathrm{Pf}V\right)`$ (A17)
$`{\displaystyle \frac{32}{c^8}}\left(_i\mathrm{Pf}V_i(\mathrm{Pf}V\widehat{W})V_i\mathrm{Pf}V_i\mathrm{Pf}\widehat{W}\widehat{W}_i\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A18)
$`+{\displaystyle \frac{8}{c^8}}\left(_k\mathrm{Pf}\widehat{W}_{ij}_k\mathrm{Pf}\widehat{W}_{ij}+\widehat{W}_{ij}\mathrm{}\mathrm{Pf}\widehat{W}_{ij}{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}\widehat{W}_{ij}\widehat{W}_{ij})\right)`$ (A19)
$`{\displaystyle \frac{64}{c^8}}(_i\mathrm{Pf}V_i\mathrm{Pf}(\widehat{X}+{\displaystyle \frac{1}{2}}\widehat{Z})+{\displaystyle \frac{1}{2}}V\mathrm{}\mathrm{Pf}(\widehat{X}+{\displaystyle \frac{1}{2}}\widehat{Z})+{\displaystyle \frac{1}{2}}(\widehat{X}+{\displaystyle \frac{1}{2}}\widehat{Z})\mathrm{}\mathrm{Pf}V`$ (A20)
$`{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V(\widehat{X}+{\displaystyle \frac{1}{2}}\widehat{Z})))\}+𝒪(10),`$ (A21)
$`\delta _{\mathrm{Leibniz}}h^{0i}=`$ (A22)
$`\mathrm{}_{}^1\{{\displaystyle \frac{16}{c^5}}(_j\mathrm{Pf}V_j\mathrm{Pf}V_i+{\displaystyle \frac{1}{2}}V\mathrm{}\mathrm{Pf}V_i+{\displaystyle \frac{1}{2}}V_i\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}VV_i){\displaystyle \frac{1}{c^2}}_t\mathrm{Pf}V_t\mathrm{Pf}V_i)`$ (A23)
$`+{\displaystyle \frac{24}{c^7}}\left(_t\mathrm{Pf}V_i(\mathrm{Pf}V^2)2V_t\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A24)
$`+{\displaystyle \frac{24}{c^7}}\left(_i\mathrm{Pf}V_t(\mathrm{Pf}V^2)2V_i\mathrm{Pf}V_t\mathrm{Pf}V\right)`$ (A25)
$`{\displaystyle \frac{32}{c^7}}\left(_j\mathrm{Pf}V_j\mathrm{Pf}\widehat{R}_i+{\displaystyle \frac{1}{2}}V\mathrm{}\mathrm{Pf}\widehat{R}_i+{\displaystyle \frac{1}{2}}\widehat{R}_i\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V\widehat{R}_i)\right)`$ (A26)
$`{\displaystyle \frac{8}{c^7}}\left(_j\mathrm{Pf}V_i_j\mathrm{Pf}\widehat{W}+{\displaystyle \frac{1}{2}}V_i\mathrm{}\mathrm{Pf}\widehat{W}+{\displaystyle \frac{1}{2}}\widehat{W}\mathrm{}\mathrm{Pf}V_i{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V_i\widehat{W})\right)`$ (A27)
$`+{\displaystyle \frac{32}{c^8}}\left(_j\mathrm{Pf}V_i(\mathrm{Pf}VV_j)V_j\mathrm{Pf}V_i\mathrm{Pf}V_jV_j_j\mathrm{Pf}V_i\mathrm{Pf}V\right)`$ (A28)
$`{\displaystyle \frac{32}{c^8}}\left(_j\mathrm{Pf}V_j(\mathrm{Pf}VV_i)V_j\mathrm{Pf}V_j\mathrm{Pf}V_iV_i_j\mathrm{Pf}V_j\mathrm{Pf}V\right)`$ (A29)
$`+{\displaystyle \frac{32}{c^7}}\left(_i\mathrm{Pf}V_j_j(\mathrm{Pf}V^2)2V_i\mathrm{Pf}V_j_j\mathrm{Pf}V\right)`$ (A30)
$`{\displaystyle \frac{32}{c^7}}\left(_j\mathrm{Pf}V_i_j(\mathrm{Pf}V^2)2V_j\mathrm{Pf}V_i_j\mathrm{Pf}V\right)`$ (A31)
$`+{\displaystyle \frac{16}{c^7}}\left(_k\mathrm{Pf}V_j_k\mathrm{Pf}\widehat{W}_{ij}+{\displaystyle \frac{1}{2}}V_j\mathrm{}\mathrm{Pf}\widehat{W}_{ij}+{\displaystyle \frac{1}{2}}\widehat{W}_{ij}\mathrm{}\mathrm{Pf}V_j{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V_j\widehat{W}_{ij})\right)`$ (A32)
$`{\displaystyle \frac{32}{c^7}}(V_i_j\mathrm{Pf}V_j\mathrm{Pf}V+2V_j\mathrm{Pf}V_j\mathrm{Pf}V_i+{\displaystyle \frac{1}{2}}V^2\mathrm{}\mathrm{Pf}V_i+VV_i\mathrm{}\mathrm{Pf}V{\displaystyle \frac{1}{2}}\mathrm{}(\mathrm{Pf}V^2V_i))\}`$ (A33)
$`+𝒪(9),`$ (A34)
$`\delta _{\mathrm{Leibniz}}h^{ij}=`$ (A35)
$`\mathrm{}_{}^1\{{\displaystyle \frac{16}{c^6}}(_{(i}\mathrm{Pf}V_{j)}(\mathrm{Pf}V^2)2V_{(i}\mathrm{Pf}V_{j)}\mathrm{Pf}V)`$ (A36)
$`{\displaystyle \frac{8}{c^6}}\delta _{ij}(_k\mathrm{Pf}V_k(\mathrm{Pf}V^2)2V_k\mathrm{Pf}V_k\mathrm{Pf}V)\}+𝒪(8).`$ (A37)
|
warning/0007/hep-ph0007354.html
|
ar5iv
|
text
|
# 1 Illustrating the variation of the effective coupling Λ of RS graviton towers to electrons as a function of 𝑥_𝑡=√{-𝑡}/𝑚₀.
TIFR/TH/00-38
IITK/PHY/2000/13
hep-ph/0007354
Testing the Randall-Sundrum Model at a High Energy $`𝐞^{}𝐞^{}`$ Collider
Dilip Kumar Ghosh
Department of Theoretical Physics,
Tata Institute of Fundamental Research,
Homi Bhabha Road, Mumbai 400 005, India.
E-mail: dghosh@theory.tifr.res.in
Sreerup Raychaudhuri<sup>1</sup><sup>1</sup>1On leave of absence from the Department of Physics, Indian Institute of Technology, Kanpur 208 016, India. E-mail: sreerup@iitk.ac.in
Theory Division, CERN,
CH-1211 Geneva 23, Switzerland.
E-mail: sreerup@mail.cern.ch
ABSTRACT
We study the process $`e^{}e^{}e^{}e^{}`$ at a high energy $`e^{}e^{}`$ collider including the effect of graviton exchanges in the warped gravity model of Randall and Sundrum. Discovery limits for gravitons are established and the effects of polarization are discussed.
July 2000
A great deal of recent interest centres around the physics possibilities of a high energy linear collider with $`e^\pm `$ beams. Such a machine can be run in $`e^+e^{}`$ or $`e^{}e^{}`$ collision modes modes. Th principal scattering processes at these are, respectively, Bhabha scattering $`e^+e^{}e^+e^{}`$ and M$`\overline{)}\mathrm{o}`$ller scattering $`e^{}e^{}e^{}e^{}`$.
One of the useful features of M$`\overline{)}\mathrm{o}`$ller scattering $`e^{}e^{}e^{}e^{}`$ at a high energy $`e^{}e^{}`$ collider is that it can receive only a limited number of contributions from physics options which go beyond the Standard Model (SM). Among the interesting beyond-Standard-Model (BSM) options are the exchange of multiple gravitons in models with low-scale quantum gravity. The exchange of multiple gravitons, in the $`t`$-channel as well as the $`u`$-channel, can affect the process $`e^{}e^{}e^{}e^{}`$ in two ways:
* by changing (increasing or decreasing) the total cross-section from the SM value — this being the usual effect of BSM physics;
* by changing the kinematic distributions of the final state electrons — this being the effect of exchanging particles with higher spin.
Two different scenarios of low-scale quantum gravity have attracted a great deal of recent attention. In one of these, due to Arkani-Hamed, Dimopoulos and Dvali (ADD), one envisages a spacetime with 4+$`d`$ dimensions, where the extra $`d`$ dimensions are compactified with radii $`R_c`$ as large as a millimetre. In the ADD scenario, in four dimensions there is a tower of massive Kaluza-Klein modes of the graviton, whose masses are so densely-spaced (by as little as $`10^{13}`$ GeV) as to form a quasi-continuum. Though each graviton mode couples to electrons with the feeble strength of Newtonian gravity, the collective effect of all the gravitons contributes to interactions of almost electroweak strength. Effects of multiple exchange of gravitons in M$`\overline{)}\mathrm{o}`$ller scattering, within the ADD scenario, have been studied in Ref..
The other popular scenario of low-scale quantum gravity is that due to Randall and Sundrum, who write a non-factorizable spacetime metric
$$ds^2=e^{𝒦R_c\varphi }\eta _{\mu \nu }dx^\mu dx^\nu +R_c^2d\varphi ^2$$
(1)
involving one extra dimension $`\varphi `$ compactified with a radius $`R_c`$, which is assumed to be marginally greater than the Planck length $`10^{33}`$ cm, and an extra mass scale $`𝒦`$, which is related to the Planck scale $`M_P^{(5)}`$ in the five-dimensional bulk by $`𝒦\left[M_P^{(4)}\right]^2\left[M_P^{(5)}\right]^3`$. Such a ‘warped’ geometry is motivated by compactifying the extra dimension on a $`𝐒^1/𝐙_2`$ orbifold, with two $`D`$-branes at the orbifold fixed points, viz., one at $`\varphi =0`$ (‘Planck brane’ or ‘invisible brane’), and one at $`\varphi =\pi `$ (‘TeV brane’ or ‘visible brane’). The interesting physical consequence of this geometry is that any mass scale on the Planck brane gets scaled by the ‘warp factor’ $`e^{\pi 𝒦R_c}`$ on the TeV brane. It now requires $`𝒦R_c12`$ — which is hardly unnatural — to obtain the hierarchy between the Planck scale and the electroweak scale. There still remains a minor problem: that of stabilizing the radius $`R_c`$ (which is marginally smaller than the Planck scale) against quantum fluctuations. A simple extension of the RS construction involving an extra bulk scalar field has been proposed to stabilize $`R_c`$ and this predicts light radion excitations with possible collider signatures . Alternatively, supersymmetry on the branes can also act as a stabilizing effect. Models with SM gauge bosons and fermions in the bulk have also been considered, but will not be discussed in this work.
The mass spectrum and couplings of the graviton in the RS model have been worked out, in Refs. . We do not describe the details of this calculation, but refer the reader to the original literature. It suffices here to note the following points.
1. There is a tower of massive Kaluza-Klein modes of the graviton, with masses
$$M_n=x_n𝒦e^{\pi 𝒦R_c}x_nm_o$$
(2)
where $`m_0=𝒦e^{\pi 𝒦R_c}`$ sets the scale of graviton masses and is essentially a free parameter of the theory. The $`x_n`$ are the zeros of the Bessel function $`J_1(x)`$ of order unity.
2. The massless Kaluza-Klein mode couples to matter with gravitational strength; consequently its effects can be ignored for all practical purposes.
3. Couplings of the massive Kaluza-Klein modes are gravitational, scaled by the warp factor $`e^{\pi 𝒦R_c}`$ and are consequently of electroweak strength.
Feynman rules (to the lowest order) for these modes have been worked out in Refs. ) and in the context of ADD-like scenarios. Each graviton couples to matter with strength $`\kappa =\sqrt{16\pi G_N}`$. All that we need to do to get the corresponding Feynman rules in the RS model is to multiply the coupling constant $`\kappa `$ by the warp factor $`e^{\pi 𝒦R_c}`$ wherever necessary. It is convenient to write
$$\kappa e^{\pi 𝒦R_c}=\sqrt{32\pi }\frac{c_0}{m_0}$$
(3)
where $`\kappa =\sqrt{16\pi G_N}`$, using Eqn. (2) and introducing another undetermined parameter $`c_0𝒦/M_P^{(4)}`$. Thus ($`c_0,m_0`$) may conveniently be taken as the free parameters of the theory<sup>2</sup><sup>2</sup>2 Though we differ from the exact choice of parameters in Ref. , a translation is easily made using the formulae $`c_0=\frac{1}{8\pi }\left(𝒦/\overline{M}_P\right)`$ and $`m_0=\mathrm{\Lambda }_\pi \left(𝒦/\overline{M}_P\right)`$. It follows that $`c_0`$ is roughly an order of magnitude less than $`𝒦/\overline{M}_P`$ and $`m_0`$ can be one or two orders of magnitude smaller than $`\mathrm{\Lambda }_\pi `$.. Though $`c_0`$ and $`m_0`$ are not precisely known, one can make estimates of their magnitude. The RS construction requires $`𝒦`$ to be at least an order of magnitude less than $`M_P^{(4)}`$, which means that the range of interest for $`c_0`$ is about 0.01 to 0.1 (the lower value being determined by naturalness considerations). $`m_0`$, which is of electroweak scale, may be considered in the range of a few tens of GeV to a few TeV. Eq. (2) tells us that the first massive graviton lies at $`M_1=x_1m_03.83m_0`$. Since no graviton resonances have been seen at LEP-2, running at energies around 200 GeV, it is clear that we should expect $`m_0>50`$ GeV at least.
In this letter, we examine the effects of multiple graviton exchange in M$`\overline{)}\mathrm{o}`$ller scattering in the RS scenario. We focus on the possibility of observing an excess in $`e^{}e^{}`$ events over the SM prediction, and comment on possible refinements using the the kinematic distributions of the final-state electrons. As earlier calculations have shown, in the case when $`c_0`$ is large, the resonance structure in Bhabha scattering is lost and there is not much difference, qualitatively speaking, between Bhabha and M$`\overline{)}\mathrm{o}`$ller scattering in the RS model. In other words, M$`\overline{)}\mathrm{o}`$ller scattering is as good a probe of this model as Bhabha scattering in this case. It is on this option that our interest is focussed.
The calculation of the Feynman amplitude involves, for the diagrams with graviton exchange, a sum over graviton propagators of the form
$$\underset{n}{}\frac{1}{tM_n^2}\frac{1}{m_0^2}\mathrm{\Lambda }\left(\frac{\sqrt{t}}{m_0}\right)$$
(4)
and a similar sum with $`tu`$. Using the properties of the zeros of Bessel functions, the function $`\mathrm{\Lambda }(x_t)`$ can be written, to a very good approximation, as
$$\mathrm{\Lambda }(x_t)=\frac{1}{\pi x_t}\mathrm{Im}\psi \left(1.2331+i\frac{x_t}{\pi }\right)+\frac{0.32586}{220.345+29.6898x_t^2+x_t^4}$$
(5)
where $`\psi (z)`$ is the well-known digamma function. The variation of $`\mathrm{\Lambda }(x_t)`$ with $`x_t`$ is illustrated in Figure 1. It is immediately obvious that the effective coupling of the gravitons varies according to the scattering angle, except in the case when $`\sqrt{t}m_0`$, i.e. $`x_t0`$. This is a feature quite different from that observed in the related ADD model, where it is possible to take a limit in which a similarly-defined $`\lambda (x_t)`$ is either constant or a slowly-varying function. This is also a feature which can potentially change the angular distribution of the final-state electrons.
There are six Feynman diagrams corresponding to M$`\overline{)}\mathrm{o}`$ller scattering
$$e^{}(p_1,\lambda _1)+e^{}(p_2,\lambda _2)e^{}(p_3,\lambda _3)+e^{}(p_4,\lambda _4)$$
(6)
including the Standard Model as well as graviton-exchange diagrams. Evaluation of these, using the Feynman rules for the RS model, and summing over the final-state helicities $`\lambda _3,\lambda _4`$, is straightforward and leads to a squared matrix element $`|(\lambda _1,\lambda _2)|^2`$, whose explicit form is not given here in the interests of brevity. If the initial-state electrons have a left-handed longitudinal polarization $`P`$, the differential cross-section is given by
$`{\displaystyle \frac{d\sigma }{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{64\pi s^2}}[(1P)^2|(+,+)|^2+(1+P)^2|(,)|^2`$ (7)
$`+(1P^2)\{|(+,)|^2+|(,+)|^2\}]`$
assuming that both the beams are identically polarized. The importance of the polarization factor $`P`$ is considerable, since it can be used, among other things, to enhance or decrease the SM contribution to the cross-section. In fact, polarization studies form an important part of the physics program at a linear collider.
In order to make a numerical estimate of the cross-section, we have incorporated the calculated cross-section into a Monte Carlo event generator, by means of which we calculate the cross-section for $`e^{}e^{}e^{}e^{}`$ subject to the following kinematic cuts.
* The scattering angle of the electron(s) should not lie within $`10^0`$ of the beam pipe.
* The transverse momentum of the electron(s) should not be less than 10 GeV.
These ‘acceptance’ cuts are more-or-less basic ones for any process at a high-energy collider with electron and/or positrons. Though further selection cuts will become appropriate when a more detailed analysis is done, it suffices for our analysis, which is no more than a preliminary study, to take the above cuts. We then calculate the cross-section in the SM and in the RS Model (including interference effects) for a fixed polarization $`P`$ and given input parameters $`c_0`$ and $`m_0`$ of the RS Model. Our results are given in Fig. 2.
In Fig. 2($`a`$), we present the total cross-section for the unpolarized case $`P=0`$ as a function of machine energy for three different values of the RS mass scale $`m_0=150,250`$ and 500 GeV. The dashed line represents the SM prediction and this exhibits the expected falling-off with machine energy. For large values of the graviton mass $`m_0`$, this behaviour is preserved, since the graviton contribution is very small anyway. However, when the graviton mass is smaller, the cross-sections show a marked increase with energy, which reflects the well-known behaviour of gravity. Obviously, at energies of 3–4 TeV, the gravitational contribution is huge if the mass scale $`m_0`$ is small; however, a discernible difference exists even when $`m_0=500`$ GeV. Thus, we can expect larger effects — or, conversely, stronger bounds — on the RS Model as the machine energy increases.
In Fig. 2($`b`$), we present the variation of the cross-section with the polarization $`P`$, at a 1 TeV machine, for the parameter set marked in the inset box. The solid curves correspond to the SM and the RS model predictions, for a fixed set of parameters $`(c_0,m_0)`$, while the dashed line represents the difference between the two. It is obvious that there is a modest advantage to be gained from polarizing the beams, and there is little difference between the cases when the beam is dominantly left- or right-handed. This is also expected, since graviton exchanges are non-chiral; in fact, the small difference arises from the interference between diagrams with graviton and $`Z`$-exchange.
In order to estimate the discovery reach of a linear collider, we adopt the following strategy. Discovery limits will be reached if the total experimental cross-section agrees — within the experimental precision — with the SM. Any excess or deficit must be attributed to BSM physics. Thus, for a given energy $`\sqrt{s}`$, a given polarization $`P`$ and a fixed set of parameters $`(c_0,m_0)`$, we calculate the total cross-section in the RS model. A corresponding calculation of the SM cross-section, multiplied by the luminosity, would lead to a predicted number of events. We then estimate the errors assuming that the statistical errors are Gaussian and that there are no systematic errors. While this certainly makes our estimates of the discovery limits over-optimistic, we can argue that electron detection efficiencies are generally high enough to allow us to make a reasonable estimate in this approximation. In any case, before more detailed studies of the detector design and systematic effects are undertaken, any estimate of systematic errors must be pure guesswork. We choose, therefore, to neglect such effects. Finally the search reach of the collider is given in terms of $`3\sigma `$ discovery limits.
Fig. 3 shows the search reach for the RS model at linear colliders running at 500 GeV and 1 TeV respectively, as a function of the integrated luminosity, for three different values of the coupling constant $`c_0`$ (marked along the curves). It may be seen that a linear collider could easily probe $`m_0`$ up to at least 300 GeV — which corresponds to a lightest graviton mass of around 1.3 TeV — if 500 pb<sup>-1</sup> of data are collected. A slight improvement is possible with polarized beams, as the dotted lines show. If the energy of the collider be increased to 1 TeV, the reach goes up almost by a factor of 2. It may, then, be possible to discover or exclude graviton resonances of mass 2.2 TeV or more.
While a 500 GeV or a 1 TeV collider will almost certainly be built, there has been much interest in having a collider which probes the high energy frontier. In particular, it is possible that the CLIC machine at CERN will be able to achieve a centre-of-mass energy as high as 3 TeV. Moreover, the possibility of a muon collider operating at a centre-of-mass energy of 3–4 TeV has also received serious consideration. For these machines, luminosities as high as $`10^3`$ fb<sup>-1</sup> per year have also been considered. Gravitational effects in $`e^{}e^{}e^{}e^{}`$ are, of course, identical to those in $`\mu ^{}\mu ^{}\mu ^{}\mu ^{}`$. In view of these possibilities, we have explored the discovery reach of a 3 TeV machine for the RS model. Our results are exhibited in Fig. 4. It may be seen that this can easily probe $`m_0`$ as high as 1 TeV, which corresponds to gravitons of mass nearly 4 TeV or more.
It is worth noting that if graviton masses are pushed up to 5 TeV or more, then, given that the scale $`𝒦`$ must be roughly an order of magnitude smaller than $`M_P^{(4)}`$, it follows that the warp factor $`e^{\pi 𝒦R_c}`$ must be somewhat larger than is possible now. This would either push up the Higgs boson mass to unacceptable values, or require some mechanism to have a smaller mass scale origin for the Higgs boson mass on the Planck brane. This would be a somewhat uncomfortable situation for the RS model, since the original simplicity — and therefore elegance — will be lost.
Finally we comment on the possibility of observing/constraining graviton effects using the angular distribution of the final state electrons. Since this form of BSM physics involves exchange of spin-2 particles, rather than spin-1 particles, as in the SM, one can, in principle, expect a rather different angular distribution for the electrons in the final state. In order to test this prediction, we have made a $`\chi ^2`$-analysis of the electron angular distribution in the cases when there is graviton exchange and when there is no graviton exchange. It turns out that the difference in the distributions is rather small and confined to the central region. We find that one cannot get better discovery limits by considering the angular distributions than those which can be obtained by simply considering the total cross-section. If indeed an excess or deficit over the SM prediction is found, angular distributions might then become useful in determining the type of BSM physics responsible, e.g. in distinguishing between spin-1 and spin-0 exchanges. However, this would require high statistics and fine resolutions. Accordingly, in this preliminary study, we do not pursue the question of angular distributions any further.
In conclusion, therefore, an $`e^{}e^{}`$ collider would be a useful laboratory to look for graviton exchange mechanisms, since there are very few competing BSM processes. We find that a simple study of the total cross-section for $`e^{}e^{}e^{}e^{}`$, subject to some minimal acceptance cuts, leads to a prediction of rather optimistic discovery limits. It is more useful to consider the total cross-section than the angular distribution, which is rather similar to that in the SM. Polarization of the beams can improve the search reach by a few percent, irrespective of whether the beams are left- or right-polarized. At a high energy collider, running at 3 or 4 TeV, the search limits can be taken as far as graviton masses of 5 TeV or more, which is more-or-less the frontier as far as the simplest version of the RS Model is concerned.
Acknowledgements: The authors would like to acknowledge Prasanta Das and Saswati Sarkar for useful discussions, and the Theory Division, CERN for hospitality while this work was being done. DKG would also like to thank F. Boudjema and LAPP, Annecy for hospitality.
|
warning/0007/hep-th0007049.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The pioneering work of Ooguri and Vafa revealed an intimate connection between self-dual field theories and (classical) $`N=2`$ string theories, formulated in four spacetime dimensions. In particular, non-abelian gauge fields on Kleinian flat space $`^{2,2}`$ of ultrahyperbolic signature $`(++)`$ and a self-dual field strength arise as exact (to all orders in $`\alpha ^{}`$) classical background configurations for the open $`N=2`$ string. Indeed, the only physical string degree of freedom in this case is a massless Lie-algebra-valued scalar field whose (tree-level) dynamics takes the form of Leznov’s or Yang’s equation, which both describe self-dual Yang-Mills (SDYM), albeit in different gauges. Although the absence of an infinite tower of massive excitations indicates a sort of caricature of a string, this quality makes it amenable to exact solutions, a fact quite rare in string theory. Yet $`N=2`$ strings may not only serve as a testing ground for certain issues in string theory in general but, being consistent quantum theories, they can also help us guiding the quantization of self-dual Yang-Mills theory.
To set the stage for the comparison of string theory with field theory, we review in Sections 2 and 3 the twistor description of the self-dual Yang-Mills equations, their symmetries and hierarchy, on flat Kleinian space $`^{2,2}`$. Although our treatment mainly follows Refs. , we reformulate their results in a language amenable to string theory. In particular, a real form of the integrable SDYM hierarchy corresponding to affine extensions of spacetime translations is characterized. Notice that the importance of the SDYM equations in $`^{2,2}`$ is also motivated by the conjecture that the SDYM model may be a universal integrable model. Indeed, it has been shown that most (if not all) integrable equations in $`1D3`$ dimensions can be obtained from the SDYM (or their hierarchy) equations by suitable reductions (see and references therein). Therefore open $`N=2`$ strings can also provide a consistent quantization of integrable models in $`1D3`$ dimensions.
If $`N=2`$ string theory “predicts” self-dual Yang-Mills, its wealth of symmetries should be obtainable from the stringy description. More precisely, we expect the SDYM hierarchy related to the non-local abelian symmetries of the SDYM equations to be visible in $`N=2`$ open string quantum mechanics. An analogous connection should exist between the symmetries of the closed $`N=2`$ string and the self-dual gravity hierarchy. Indeed, one of the authors (together with Jünemann and Popov) has recently identified part of these hidden closed string symmetries and has succeeded in relating them to the self-dual gravity hierarchy . Quite surprisingly, the stringy root of such symmetries is technically the somewhat obscure picture phenomenon which is present whenever covariant quantization meets worldsheet supersymmetry. Global symmetries unbroken by the string background under consideration may be classified with the help of BRST cohomology, and the latter unexpectedly displays a picture dependence (see also ).
In Sections 4 and 5, we briefly review this issue in the context of the open $`N=2`$ string. Although the field-theoretic description of SDYM is simpler than that of self-dual gravity, its string-theoretic version is more involved because open string mechanics must take into account worldsheet boundaries, cross-caps, and boundary punctures. Nevertheless, the hidden symmetries emerge as in the closed-string case. Furthermore, they are seen to be responsible for the vanishing of almost all (tree-level) open-string scattering amplitudes.
After concluding, two Appendices collect standard material about line bundles over the Riemann sphere and about twistor spaces.
## 2 Holomorphic bundles and self-dual Yang-Mills fields
In this Section, we use some facts about line bundles over the Riemann sphere and geometry of the twistor spaces which are recalled in two Appendices.
2.1 Vector bundles over $`𝒫`$ and the Penrose-Ward correspondence
By a holomorphic rank $`r`$ vector bundle we mean a collection of five objects $`(E,X,p,𝔘,f)`$, where $`E`$ and $`X`$ are complex manifolds, $`p:EX`$ is a holomorphic projection, $`𝔘=\{𝒰_i\}`$ is a covering of the manifold $`X`$ and $`f=\{f_{ij}\}`$ is a collection of holomorphic transition functions $`f_{ij}`$ on $`𝒰_i𝒰_j`$ taking values in complex $`r\times r`$ matrices. In particular, for the twistor space $`𝒫`$, one can always introduce a covering $`𝔘=\{\overline{𝒰}_+,\overline{𝒰}_{}\}`$ such that
$$\overline{𝒰}_+=\overline{H}_+^2\times ^2,\overline{𝒰}_{}=\overline{H}_{}^2\times ^2,𝒫_0=\overline{𝒰}_+\overline{𝒰}_{}S^1\times ^4,$$
(2.1)
where
$$\overline{H}_+^2:=H_+^2S^1=\{\zeta \{\mathrm{}\}:\text{Im}\zeta 0\},$$
$$\overline{H}_{}^2:=H_{}^2S^1=\{\zeta \{\mathrm{}\}:\text{Im}\zeta 0\},$$
(2.2)
and $`S^1=\overline{H}_+^2\overline{H}_{}^2=\{\zeta \{\mathrm{}\}:\text{Im}\zeta =0\}`$. For the covering $`𝔘`$ of $`𝒫`$, holomorphic bundles $`E𝒫`$ are defined by real-analytic transition functions $`f_+`$ on $`𝒫_0=\overline{𝒰}_+\overline{𝒰}_{}`$ annihilated by the vector fields
$$V_{\stackrel{\text{.}}{\alpha }}=\frac{}{x^{1\stackrel{\text{.}}{\alpha }}}\zeta \frac{}{x^{0\stackrel{\text{.}}{\alpha }}}$$
(2.3)
with real $`\zeta `$ and $`\stackrel{\text{.}}{\alpha }=0,1`$. These transition functions extend holomorphically in local complex coordinates $`\eta ^{\stackrel{\text{.}}{\alpha }}`$ and $`\zeta `$ to an open neighbourhood $`𝒰`$ of $`𝒫_0`$ in $`𝒫`$.
Let us consider holomorphic rank $`r`$ vector bundles $`E`$ over the twistor space $`𝒫`$ and suppose that bundles $`E`$ satisfy the following conditions:
(i) restriction $`E|_{\sigma _x}`$ to every real holomorphic section $`\sigma _x\mathrm{\Gamma }_{}(𝒫)`$ is trivial,
(ii) $`detE`$ is trivial,
(iii) $`E`$ has a real structure $`\tau ^{}`$.
We shall show that such bundles correspond to self-dual gauge fields on $`^{2,2}`$ .
In terms of transition functions $`f_+`$, the condition (i) means that $`f_+(\eta ^{\stackrel{\text{.}}{\alpha }},\zeta )|_{\sigma _x}`$ is the transition function for a trivial bundle over $`P_x^1`$ and therefore it can be factorized:
$$f_+(x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }},\zeta )=\psi _+^1(x,\zeta )\psi _{}(x,\zeta ),$$
(2.4)
where $`x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }}=\eta ^{\stackrel{\text{.}}{\alpha }}|_{\sigma _x},`$ $`x=\{x^{\alpha \stackrel{\text{.}}{\alpha }}|\alpha =0,1\}^{2,2}`$, and the matrices $`\psi _+(x,\zeta )`$ and $`\psi _{}(x,\zeta )`$ are holomorphic with respect to $`\zeta `$ in the upper and lower half-planes, respectively. The condition (ii) means that the structure group $`GL(r,)`$ of the bundle $`E`$ is reduced to the structure group $`SL(r,)`$. In terms of transition functions $`f_+`$ this condition has the form
$$detf_+=1.$$
(2.5)
The real structure $`\tau ^{}`$ in (iii) is induced from the real structure $`\tau `$ on $`𝒫`$ described in Appendix A.2. Namely, we put
$$\tau ^{}(f_+(\eta ^{\stackrel{\text{.}}{\alpha }},\zeta ))=f_+^{}(\overline{\eta ^{\stackrel{\text{.}}{\alpha }}},\overline{\zeta }),$$
where $``$ means the Hermitian conjugation. Then stable “points” of the map $`\tau ^{}`$ are transition functions satisfying
$$\tau ^{}(f_+)=f_+f_+^{}(\overline{\eta ^{\stackrel{\text{.}}{\alpha }}},\overline{\zeta })=f_+(\eta ^{\stackrel{\text{.}}{\alpha }},\zeta ).$$
(2.6)
It is not difficult to see that (2.6) takes place if we impose the following conditions on $`\psi _\pm `$:
$$det\psi _+=det\psi _{}=1,\psi _+^1(x,\zeta )=\psi _{}^{}(x,\overline{\zeta }).$$
(2.7)
Then $`f_+|_{\sigma _x}=\psi _{}^{}(x,\overline{\zeta })\psi _{}(x,\zeta )`$ is Hermitian on real sections $`\sigma _x\mathrm{\Gamma }_{}(𝒫)`$. Notice that if the matrix $`f_+`$ satisfies the reality conditions (2.6), then the factorization (2.4) exists for any $`x^{2,2}`$, by the results of Gohberg and Krein (see e.g. ).
Matrices $`f_+|_{\sigma _x}=f_+(x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }},\zeta )f_+(x,\zeta )`$ are annihilated by the vector fields (2.3),
$$V_{\stackrel{\text{.}}{\alpha }}f_+(x,\zeta )=0.$$
(2.8)
Substituting (2.4) into (2.8), we obtain
$$(V_{\stackrel{\text{.}}{\alpha }}\psi _+)\psi _+^1=(V_{\stackrel{\text{.}}{\alpha }}\psi _{})\psi _{}^1.$$
(2.9)
The left-hand side of (2.9) is holomorphic in finite $`\zeta \overline{H}_+^2`$, the right-hand side is holomorphic in finite $`\zeta \overline{H}_{}^2`$ and both sides have the simple pole at $`\zeta =\mathrm{}`$. Hence by an extension to Liouville’s theorem both the sides of (2.9) must be linear in $`\zeta `$,
$$(V_{\stackrel{\text{.}}{\alpha }}\psi _+)\psi _+^1=(V_{\stackrel{\text{.}}{\alpha }}\psi _{})\psi _{}^1=A_{1\stackrel{\text{.}}{\alpha }}+\zeta A_{0\stackrel{\text{.}}{\alpha }},$$
(2.10)
where $`A_{\alpha \stackrel{\text{.}}{\alpha }}`$ are some $`r\times r`$ matrices which depend only on $`x`$, $`\alpha =0,1,\stackrel{\text{.}}{\alpha }=0,1`$. From the conditions (2.7) it follows that $`A_{\alpha \stackrel{\text{.}}{\alpha }}`$ are trace-free anti-Hermitian $`r\times r`$ matrices and they can be identified with components of a Yang-Mills gauge potential $`A=A_{\alpha \stackrel{\text{.}}{\alpha }}dx^{\alpha \stackrel{\text{.}}{\alpha }}`$ on $`^{2,2}`$ taking values in the Lie algebra $`su(r)`$.
Let us rewrite (2.10) in the form
$$D_{\stackrel{\text{.}}{\alpha }}\psi _+=0,$$
(2.11)
where
$$D_{\stackrel{\text{.}}{\alpha }}:=D_{1\stackrel{\text{.}}{\alpha }}\zeta D_{0\stackrel{\text{.}}{\alpha }}=\frac{}{x^{1\stackrel{\text{.}}{\alpha }}}+A_{1\stackrel{\text{.}}{\alpha }}\zeta (\frac{}{x^{0\stackrel{\text{.}}{\alpha }}}+A_{0\stackrel{\text{.}}{\alpha }}),$$
(2.12)
and $`D_{\alpha \stackrel{\text{.}}{\alpha }}=_{\alpha \stackrel{\text{.}}{\alpha }}+A_{\alpha \stackrel{\text{.}}{\alpha }}`$ are covariant derivatives, $`_{\alpha \stackrel{\text{.}}{\alpha }}:=/x^{\alpha \stackrel{\text{.}}{\alpha }}`$. Similar equations hold for $`\psi _{}`$. The compatibility condition of the linear system of differential equations (2.11) is
$$[D_{\stackrel{\text{.}}{0}},D_{\stackrel{\text{.}}{1}}]=0,$$
(2.13)
and equating the coefficients of $`\zeta ^0,\zeta ^1`$ and $`\zeta ^2`$ in (2.13) to zero, we obtain the self-dual Yang-Mills (SDYM) equations which in the coordinates $`x^{\alpha \stackrel{\text{.}}{\alpha }}`$ have the form
$$[D_{0\stackrel{\text{.}}{0}},D_{0\stackrel{\text{.}}{1}}]=0,[D_{1\stackrel{\text{.}}{0}},D_{1\stackrel{\text{.}}{1}}]=0,[D_{0\stackrel{\text{.}}{0}},D_{1\stackrel{\text{.}}{1}}]+[D_{1\stackrel{\text{.}}{0}},D_{0\stackrel{\text{.}}{1}}]=0.$$
(2.14)
So, transition functions $`f_+`$ determining holomorphic bundles $`E`$ over $`𝒫`$ and satisfying conditions (2.4)-(2.6) encode all the information about self-dual gauge fields in $`^{2,2}`$.
From formula (2.10) it follows that the gauge potential $`A`$ does not change its form under the transformations $`\psi _+\psi _+h_+(x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }},\zeta )`$, $`\psi _{}\psi _{}h_{}(x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }},\zeta )`$, since such $`h_\pm `$ are annihilated by $`V_{\stackrel{\text{.}}{\alpha }}`$. Under these transformations $`f_+`$ transforms into the transition function $`h_+^1f_+h_{}`$ of a bundle equivalent to $`E`$. At the same time, gauge transformations $`Ag^1Ag+g^1dg`$ with $`g=g(x)SU(r)`$ correspond to transformations $`\psi _\pm g^1\psi _\pm `$ under which $`f_+`$ does not change. So, there is a one-to-one correspondence between gauge equivalence classes of $`su(r)`$-valued solutions to the SDYM equations (2.14) on the Kleinian space $`^{2,2}`$ and equivalence classes of holomorphic rank $`r`$ vector bundles $`E`$ over $`𝒫`$ satisfying conditions (i)-(iii).
2.2 Gauge fixing
Let us consider Eqs. (2.11) rewritten in the form
$$(D_{1\stackrel{\text{.}}{\alpha }}\zeta D_{0\stackrel{\text{.}}{\alpha }})\psi _+=0,$$
(2.15)
and analogous equations for $`\psi _{}`$. Recall that $`\psi _\pm `$ are smooth functions for $`x^{2,2},\zeta =\mathrm{cot}\frac{\theta }{2}\{\mathrm{}\}=S^1`$. Considering $`\zeta \mathrm{}`$ in (2.15), one obtains that
$$A_{0\stackrel{\text{.}}{\alpha }}(x)=\psi _+(x,\zeta )_{0\stackrel{\text{.}}{\alpha }}\psi _+^1(x,\zeta )|_{\zeta =\mathrm{}},$$
(2.16)
where $`\zeta =\mathrm{}`$ corresponds to $`\theta =0`$. Using a gauge transformation $`\psi _+g^1\psi _+`$ generated by $`g(x)=\psi _+(x,\zeta =\mathrm{})`$, one may transform $`A_{0\stackrel{\text{.}}{\alpha }}`$ to zero, which is equivalent to imposing conditions $`\psi _+(x,\zeta =\mathrm{})=\psi _{}(x,\zeta =\mathrm{})=1`$. Moreover, we impose the standard asymptotic conditions on $`\psi _+`$ and $`\psi _{}`$,
$$\psi _+(x,\zeta )=1+\zeta ^1\mathrm{\Psi }(x)+O(\zeta ^2),\psi _{}(x,\zeta )=1+\zeta ^1\mathrm{\Psi }(x)+O(\zeta ^2)$$
(2.17)
as $`\zeta \mathrm{}`$ in their respective domains (cf. ). At the same time, when $`\zeta 0`$ we have
$$\psi _+(x,\zeta )=\mathrm{\Phi }^1(x)+O(\zeta ),\psi _{}(x,\zeta )=\mathrm{\Phi }^1(x)+O(\zeta ),$$
(2.18)
where $`\mathrm{\Phi }(x):=\psi _+^1(x,\zeta =0)=\psi _{}^1(x,\zeta =0)`$.
By substituting (2.18) into (2.15), we have
$$A_{0\stackrel{\text{.}}{\alpha }}=0,A_{1\stackrel{\text{.}}{\alpha }}=\mathrm{\Phi }^1_{1\stackrel{\text{.}}{\alpha }}\mathrm{\Phi }.$$
(2.19)
This is the Yang gauge for which the SDYM equations (2.14) are replaced by the Yang equation ,
$$_{0\stackrel{\text{.}}{0}}(\mathrm{\Phi }^1_{1\stackrel{\text{.}}{1}}\mathrm{\Phi })_{0\stackrel{\text{.}}{1}}(\mathrm{\Phi }^1_{1\stackrel{\text{.}}{0}}\mathrm{\Phi })=0.$$
(2.20)
Analogously, by substituting (2.17) into (2.15), we find that
$$A_{0\stackrel{\text{.}}{\alpha }}=0,A_{1\stackrel{\text{.}}{\alpha }}=_{0\stackrel{\text{.}}{\alpha }}\mathrm{\Psi },$$
(2.21)
and Eqs. (2.14) are reduced to the Leznov equation ,
$$_{1\stackrel{\text{.}}{0}}_{0\stackrel{\text{.}}{1}}\mathrm{\Psi }_{0\stackrel{\text{.}}{0}}_{1\stackrel{\text{.}}{1}}\mathrm{\Psi }+[_{0\stackrel{\text{.}}{0}}\mathrm{\Psi },_{0\stackrel{\text{.}}{1}}\mathrm{\Psi }]=0.$$
(2.22)
## 3 Symmetries and hierarchy of the SDYM equations
3.1 Deformations of bundles and symmetries of the SDYM equations
In Section 2 we have introduced the covering $`𝔘=\{\overline{𝒰}_+,\overline{𝒰}_{}\}`$ of the twistor space $`𝒫`$ and holomorphic bundles $`E𝒫`$ defined by transition functions $`f_+`$. The transition functions $`f_+`$ are holomorphic on an open neighbourhood $`𝒰`$ of $`𝒫_0=\overline{𝒰}_+\overline{𝒰}_{}`$ and real-analytic on the real twistor space $`𝒯`$ (see Appendix A.2). It has been shown that transition functions $`f_+`$ of holomorphic bundles $`E𝒫`$ encode all the information about self-dual gauge potentials $`A`$ on the Kleinian space $`^{2,2}`$. We have written down the explicit formulae expressing $`A`$ through matrix-valued functions $`\psi _\pm (x,\zeta )`$ defining a trivialization of the bundle $`E`$ on real holomorphic sections of the bundle $`𝒫P^1`$.
Any holomorphic perturbation of $`f_+`$ on $`𝒰`$ preserving conditions (2.5) and (2.6) is allowed since small enough deformations of the bundle $`E`$ preserve the property (2.4) of its trivializability on $`P_x^1𝒫`$. Using the Penrose-Ward correspondence $`f_+A`$, to each infinitesimal change $`\delta f_+`$ of the transition function $`f_+`$ of the bundle $`E`$ one can correspond an infinitesimal change $`\delta A`$ of the self-dual gauge potential $`A`$. By construction, such $`\delta A`$ satisfy the linearized SDYM equations and called infinitesimal symmetries of the SDYM equations. In this Section we describe all infinitesimal symmetries of the SDYM equations in order to compare them later with hidden symmetries of open $`N=2`$ strings.
We consider a bundle $`E𝒫`$, the covering $`𝔘=\{\overline{𝒰}_+,\overline{𝒰}_{}\}`$ of $`𝒫`$, a transition function $`f_+`$ for $`E`$ satisfying the conditions (2.4)-(2.6) and the self-dual gauge potential $`A`$ corresponding to $`f_+`$. On real holomorphic sections $`\sigma _x\mathrm{\Gamma }_{}(𝒫)`$ we have $`f_+|_{\sigma _x}=\psi _+^1(x,\zeta )\psi _{}(x,\zeta )`$, $`x^{2,2}`$. Let us consider a perturbation $`\delta f_+`$ of $`f_+`$ and the factorization of the perturbed transition function,
$$f_++\delta f_+=(\psi _++\delta \psi _+)^1(\psi _{}+\delta \psi _{}),$$
(3.1)
supposing that $`f_++\delta f_+`$ satisfies the conditions (2.5) and (2.6) up to the first order in $`\delta f_+`$. Then introduce a Lie-algebra-valued function
$$\phi _+:=\psi _+(\delta f_+)f_+^1\psi _+^1=\psi _+(\delta f_+)\psi _{}^1,$$
(3.2)
defining an infinitesimal deformation of the bundle $`E𝒫`$.
For fixed $`x^{2,2}`$ we have a function $`\phi _+(x,\zeta )`$ on $`S_x^1P_x^1`$ with $`\zeta S^1P^1`$. As usual, $`\phi _+(x,\zeta )`$ can be factorized,
$$\phi _+(x,\zeta )=\phi _+(x,\zeta )\phi _{}(x,\zeta ),$$
(3.3)
where Lie-algebra-valued functions $`\phi _+`$ and $`\phi _{}`$ can be extended to functions holomorphic in $`\zeta `$ in upper and lower half-planes, respectively. To find $`\phi _\pm `$ means to solve the infinitesimal variant of the Riemann-Hilbert problem and a solution to (3.3) always exists . Suppose $`\chi _\pm (\eta ^{\stackrel{\text{.}}{\alpha }},\zeta )`$ are matrix-valued real-analytic functions on the real twistor space $`𝒯`$ extendible to holomorphic functions on $`\overline{𝒰}_\pm 𝒫_0`$. Then $`\phi _+`$ of the form
$$\phi _+=\psi _+\chi _+\psi _+^1\psi _{}\chi _{}\psi _{}^1,$$
(3.4)
defines a trivial perturbation of the transition function $`f_+`$.
Remark. Functions $`\phi _+`$ defined by formula (3.2) are elements of the space of 1-cocycles $`Z^1(𝔘,\text{ad}E)`$ of the covering $`𝔘`$ with values in the bundle $`\text{ad}E`$ of endomorphisms. Functions (3.4) form a subspace $`B^1(𝔘,\text{ad}E)`$ of 1-coboundaries (trivial 1-cocycles). Non-trivial infinitesimal deformations of the bundle $`E𝒫`$ are defined by the first cohomology group $`H^1(𝔘,\text{ad}E)=Z^1(𝔘,\text{ad}E)/B^1(𝔘,\text{ad}E)`$. For cohomological description of symmetries to the SDYM equations see .
Notice that
$$\phi _+=\psi _+(\delta f_+)\psi _{}^1=(\delta \psi _+)\psi _+^1+(\delta \psi _{})\psi _{}^1=\phi _+\phi _{},$$
(3.5)
and therefore
$$\delta \psi _+=\phi _+\psi _+,\delta \psi _{}=\phi _{}\psi _{}.$$
(3.6)
At the same time, from formulae (2.10) it follows that
$$\delta A_{1\stackrel{\text{.}}{\alpha }}\zeta \delta A_{0\stackrel{\text{.}}{\alpha }}=\delta \psi _+(_{1\stackrel{\text{.}}{\alpha }}\zeta _{0\stackrel{\text{.}}{\alpha }})\psi _+^1+\psi _+(_{1\stackrel{\text{.}}{\alpha }}\zeta _{0\stackrel{\text{.}}{\alpha }})\delta \psi _+^1=$$
$$=\delta \psi _{}(_{1\stackrel{\text{.}}{\alpha }}\zeta _{0\stackrel{\text{.}}{\alpha }})\psi _{}^1+\psi _{}(_{1\stackrel{\text{.}}{\alpha }}\zeta _{0\stackrel{\text{.}}{\alpha }})\delta \psi _{}^1.$$
(3.7)
Substituting (3.6) into (3.7), we obtain
$$\delta A_{1\stackrel{\text{.}}{\alpha }}\zeta \delta A_{0\stackrel{\text{.}}{\alpha }}=(D_{1\stackrel{\text{.}}{\alpha }}\zeta D_{0\stackrel{\text{.}}{\alpha }})\phi _+=(D_{1\stackrel{\text{.}}{\alpha }}\zeta D_{0\stackrel{\text{.}}{\alpha }})\phi _{}.$$
(3.8)
From (3.8) it follows that
$$(D_{1\stackrel{\text{.}}{\alpha }}\zeta D_{0\stackrel{\text{.}}{\alpha }})\phi _+=0,$$
(3.9)
which can also be obtained from the definition (3.2) and equations (2.10) on $`\psi _\pm `$.
It is easy to see that for $`\phi _\pm =\psi _\pm \chi _\pm \psi _\pm ^1`$ defining the 1-coboundary (3.4) we have $`\delta A_{\alpha \stackrel{\text{.}}{\alpha }}=0`$ (trivial symmetries). At the same time, for infinitesimal gauge transformations $`\delta A_{\alpha \stackrel{\text{.}}{\alpha }}=D_{\alpha \stackrel{\text{.}}{\alpha }}\phi `$ we have $`\phi _+=\phi _{}=\phi (x)`$ (i.e. $`\phi _\pm `$ do not depend on $`\zeta `$) and therefore $`\phi _+=\phi _+\phi _{}0`$ that leads to $`\delta f_+=0`$. For the case of non-trivial symmetries we have
$$\delta A_{0\stackrel{\text{.}}{\alpha }}=D_{0\stackrel{\text{.}}{\alpha }}\phi _+(x,\zeta =\mathrm{})=D_{0\stackrel{\text{.}}{\alpha }}\phi _{}(x,\zeta =\mathrm{}),$$
(3.10)
$$\delta A_{1\stackrel{\text{.}}{\alpha }}=D_{1\stackrel{\text{.}}{\alpha }}\phi _+(x,\zeta =0)=D_{1\stackrel{\text{.}}{\alpha }}\phi _{}(x,\zeta =0).$$
(3.11)
Thus, to each transformation $`f_+f_++\delta f_+`$ of the transition function $`f_+`$ in the bundle $`E𝒫`$, formulae (3.2), (3.3), (3.6), (3.10) and (3.11) correspond a symmetry transformation $`AA+\delta A`$ of the self-dual gauge potential $`A`$. By construction, such $`\delta A`$ satisfy the linearized SDYM equations.
In Subsection 2.2 we considered gauge fixing conditions for $`\psi _\pm `$ and $`\{A_{\alpha \stackrel{\text{.}}{\alpha }}\}`$. To preserve these conditions, it is enough to impose the asymptotic conditions
$$\phi _+0,\phi _{}0$$
(3.12)
as $`\zeta \mathrm{}`$ in their respective domains. Then from (3.10) we obtain $`\delta A_{0\stackrel{\text{.}}{\alpha }}=0`$ and therefore gauge fixing conditions $`A_{0\stackrel{\text{.}}{\alpha }}=0`$ are preserved. Transformations of Leznov’s prepotentials $`\mathrm{\Psi }`$ can be obtained from formulae (2.17), (3.6) and have the form
$$\delta \mathrm{\Psi }(x)=\underset{\zeta \mathrm{}}{lim}\zeta \phi _+(x,\zeta ).$$
(3.13)
Analogously, transformations of Yang’s prepotential $`\mathrm{\Phi }`$ follow from formulae (2.18), (3.6) and have the form
$$\mathrm{\Phi }^1\delta \mathrm{\Phi }=\phi _+(x,\zeta =0).$$
(3.14)
Thus, the knowledge of $`\phi _+(x,\zeta )`$ permits one to find $`\delta A_{\alpha \stackrel{\text{.}}{\alpha }},\delta \mathrm{\Psi }`$ and $`\delta \mathrm{\Phi }`$.
3.2 Hierarchy of the SDYM equations
For the twistor space $`𝒫=𝒪(1)𝒪(1)`$ with the covering $`𝔘=\{\overline{𝒰}_+,\overline{𝒰}_{}\}`$, on an open set $`𝒰\overline{𝒰}_+\overline{𝒰}_{}`$ one may define an infinite number of matrix-valued holomorphic functions $`f_+`$ satisfying conditions (2.4)-(2.6). To each such matrix $`f_+`$ there corresponds a holomorphic vector bundle $`E`$ over $`𝒫`$ and a gauge equivalence class of self-dual gauge potentials $`A`$ on $`^{2,2}`$. So, we have infinite-dimensional spaces of matrices $`f_+`$ and self-dual gauge potentials $`A`$. The moduli space of holomorphic vector bundles $`E`$ over $`𝒫`$ defined by $`f_+`$ is bijective to the moduli space $``$ of self-dual gauge fields on $`^{2,2}`$. Recall that $`=𝒩/𝒢`$, where $`𝒩`$ is the solution space and $`𝒢`$ is the group of gauge transformations. Non-trivial perturbations $`\delta f_+`$ of transition functions $`f_+`$ are vector fields on the moduli space of holomorphic bundles $`E`$, and non-trivial perturbations $`\delta A`$ of self-dual gauge potentials $`A`$ are vector fields on the moduli space $``$. The Penrose-Ward correspondence $`\delta f_+\delta A`$ described in Subsection 3.1 determines an isomorphism between the space $`H^1(𝒫,\text{ad}E)`$ of infinitesimal deformations of the bundle $`E𝒫`$ and the space $`\text{Vect}()`$ of vector fields on the moduli space $``$ of self-dual gauge fields.
Symmetries of the SDYM equations were considered in many papers (see e.g. and references therein). In particular, homomorphisms of various Kac-Moody-Virasoro type algebras into the algebra $`\text{Vect}(𝒩)`$ of vector fields on the solution space $`𝒩`$ of the SDYM equations have been described. In this paper we are mainly interested in affine extensions of spacetime translations and in a hierarchy of the SDYM equations corresponding to them. Later we show that just these symmetries correspond to abelian string symmetries.
The above-mentioned non-local abelian symmetries of the SDYM equations can be defined in the following way. Consider translations in $`^{2,2}`$ generated by a vector field
$$\stackrel{~}{T}=\underset{\stackrel{\text{.}}{\alpha }=0}{\overset{1}{}}(t^{0\stackrel{\text{.}}{\alpha }}\frac{}{x^{0\stackrel{\text{.}}{\alpha }}}+t^{1\stackrel{\text{.}}{\alpha }}\frac{}{x^{1\stackrel{\text{.}}{\alpha }}}),$$
(3.15)
where $`t^{\alpha \stackrel{\text{.}}{\alpha }}`$ are constant parameters, $`\alpha =0,1,\stackrel{\text{.}}{\alpha }=0,1.`$ The induced action of $`\stackrel{~}{T}`$ on $`f_+(x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }},\zeta )`$ is
$$\stackrel{~}{T}f_+=\underset{\stackrel{\text{.}}{\alpha }=0}{\overset{1}{}}(t^{0\stackrel{\text{.}}{\alpha }}+\zeta t^{1\stackrel{\text{.}}{\alpha }})\frac{}{\eta ^{\stackrel{\text{.}}{\alpha }}}f_+=:Tf_+,$$
(3.16)
where $`\eta ^{\stackrel{\text{.}}{\alpha }}=x^{0\stackrel{\text{.}}{\alpha }}+\zeta x^{1\stackrel{\text{.}}{\alpha }}`$ and $`\zeta `$ are the local coordinates on the real twistor space $`𝒯`$ (see Appendix A.2). So, $`T`$ is a local vector field on $`𝒯𝒫`$. Now let us consider vector fields $`T_{n\stackrel{\text{.}}{\alpha }}`$ on $`𝒯`$,
$$T_{n\stackrel{\text{.}}{\alpha }}:=\zeta ^n\frac{}{\eta ^{\stackrel{\text{.}}{\alpha }}},n=0,1,\mathrm{},2J,$$
(3.17)
where $`2J`$ is any positive integer or infinity. When $`n=0,1`$ these vector fields correspond to the translations $`/x^{0\stackrel{\text{.}}{\alpha }}`$, $`/x^{1\stackrel{\text{.}}{\alpha }}`$ in $`^{2,2}`$.
Let us define the transformations
$$f_+\delta _{n\stackrel{\text{.}}{\alpha }}f_+:=\zeta ^n\frac{}{\eta ^{\stackrel{\text{.}}{\alpha }}}f_+$$
(3.18)
of transition functions $`f_+`$ in the bundle $`E`$ restricted to $`𝒯`$. From (3.18) it is easy to see that $`[\delta _{m\stackrel{\text{.}}{\alpha }},\delta _{n\stackrel{\text{.}}{\beta }}]f_+=0`$ (commutativity). Further, by the algorithm from Subsection 3.1, one may correspond the perturbations $`\delta _{n\stackrel{\text{.}}{\alpha }}A,\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Phi }`$ and $`\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }`$ to the perturbations (3.18). Recall that $`\delta _{n\stackrel{\text{.}}{\alpha }}f_+`$, $`\delta _{n\stackrel{\text{.}}{\alpha }}A`$, $`\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Phi }`$ and $`\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }`$ are components of vector fields on the space of matrices $`f_+`$ and on the solution spaces to Eqs. (2.14), (2.20) and (2.22), respectively. To all these vector fields one may correspond dynamical systems on the solution spaces and try to solve these differential equations.
Integral trajectories of dynamical systems on the space of transition functions can be described explicitly. Namely, consider the following system of differential equations:
$$\frac{}{t^{n\stackrel{\text{.}}{\alpha }}}f_+=\zeta ^n\frac{}{\eta ^{\stackrel{\text{.}}{\alpha }}}f_+,$$
(3.19)
where $`t^{n\stackrel{\text{.}}{\alpha }}`$ are real parameters, $`\stackrel{\text{.}}{\alpha }=0,1,n=0,\mathrm{},2J`$ and $`2J`$ is any positive integer number or $`2J=\mathrm{}`$. Equations (3.19) can be easily integrated and we obtain
$$f_+(x,t,\zeta )=f_+(x^{0\stackrel{\text{.}}{\alpha }}+t^{0\stackrel{\text{.}}{\alpha }}+\zeta (x^{1\stackrel{\text{.}}{\alpha }}+t^{1\stackrel{\text{.}}{\alpha }})+\underset{n=2}{\overset{2J}{}}t^{n\stackrel{\text{.}}{\alpha }}\zeta ^n,\zeta ),$$
(3.20)
where $`t=(t^{0\stackrel{\text{.}}{\alpha }},t^{1\stackrel{\text{.}}{\alpha }},\mathrm{},t^{2J\stackrel{\text{.}}{\alpha }})`$. Any point of the space $`^{2,2}`$ can be obtained by the shift of the point $`x^{\alpha \stackrel{\text{.}}{\alpha }}=0`$ and therefore in (3.20) one may put $`x^{\alpha \stackrel{\text{.}}{\alpha }}=0`$. Then we have
$$f_+(t,\zeta )=f_+(\underset{n=0}{\overset{2J}{}}t^{n\stackrel{\text{.}}{\alpha }}\zeta ^n,\zeta ),$$
(3.21)
where $`t^{0\stackrel{\text{.}}{\alpha }}`$ and $`t^{1\stackrel{\text{.}}{\alpha }}`$ are coordinates in $`^{2,2}`$, and $`t^{n\stackrel{\text{.}}{\alpha }}`$ with $`n2`$ are extra (moduli) parameters.
Notice that for finite $`2J`$ the polynomials
$$\eta _t^{\stackrel{\text{.}}{\alpha }}(\zeta )=\underset{n=0}{\overset{2J}{}}t^{n\stackrel{\text{.}}{\alpha }}\zeta ^n$$
(3.22)
define real sections of the bundle $`𝒪(2J)𝒪(2J)P^1`$, and matrices (3.21) are transition functions of the bundle $`E𝒪(2J)𝒪(2J)`$ restricted to real sections $`S_t^1𝒪(2J)𝒪(2J)`$. In other words, local vector fields (3.17) generate the change of topology of the twistor space $`𝒫`$ from $`𝒪(1)𝒪(1)`$ to $`𝒪(2J)𝒪(2J)`$. Using the homogeneous coordinates on $`S^1`$ from Appendix A.1, one can rewrite Eqs. (3.19) for finite $`J`$ in the form
$$\frac{}{t^{n\stackrel{\text{.}}{\alpha }}}f_+=g^J\left(\genfrac{}{}{0pt}{}{2J}{n}\right)\mathrm{sin}^{2Jn}\frac{\theta }{2}\mathrm{cos}^n\frac{\theta }{2}\frac{}{\eta ^{\stackrel{\text{.}}{\alpha }}}f_+,$$
(3.23)
where parameters $`t^{n\stackrel{\text{.}}{\alpha }}`$ differ from those in (3.19) by the multipliers, but we shall not introduce new notations for them. The general solution of Eqs. (3.23) has the form
$$f_+(t,\theta )=f_+(g^J\underset{n=0}{\overset{2J}{}}\left(\genfrac{}{}{0pt}{}{2J}{n}\right)t^{n\stackrel{\text{.}}{\alpha }}\mathrm{sin}^{2Jn}\frac{\theta }{2}\mathrm{cos}^n\frac{\theta }{2},\theta )=$$
$$=f_+(g^J\underset{M=J}{\overset{J}{}}\left(\genfrac{}{}{0pt}{}{2J}{J+M}\right)t^{J+M\stackrel{\text{.}}{\alpha }}\mathrm{sin}^{JM}\frac{\theta }{2}\mathrm{cos}^{J+M}\frac{\theta }{2},\theta ),$$
(3.24)
where $`0\theta 2\pi `$.
Now one should use the Penrose-Ward transformation starting from the factorization
$$f_+(t,\zeta )=\psi _+^1(t,\zeta )\psi _{}(t,\zeta ),$$
(3.25)
where now $`\psi _\pm `$ depend on $`t=(t^{n\stackrel{\text{.}}{\alpha }})`$ and $`\zeta ,\stackrel{\text{.}}{\alpha }=0,1,n=0,\mathrm{},2J`$. As in Subsection 2.2, we use the gauge in which
$$\psi _\pm (t,\zeta )=\mathrm{\Phi }^1(t)+O(\zeta )\mathrm{for}\zeta 0,$$
(3.26)
$$\psi _\pm (t,\zeta )=1+\zeta ^1\mathrm{\Psi }(t)+O(\zeta ^2)\mathrm{for}\zeta \mathrm{},$$
(3.27)
The group-valued function $`\mathrm{\Phi }(t)`$ and the algebra-valued function $`\mathrm{\Psi }(t)`$ give (implicit) solutions of the differential equations
$$_{n\stackrel{\text{.}}{\alpha }}\mathrm{\Phi }=\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Phi },$$
(3.28)
$$_{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }=\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi },$$
(3.29)
and describe commuting flows on the space of solutions to Eqs. (2.20) and (2.22), respectively. These flows are integral curves for the dynamical systems (3.28) and (3.29), where $`_{n\stackrel{\text{.}}{\alpha }}:=/t^{n\stackrel{\text{.}}{\alpha }}`$.
Notice that the transition functions (3.21) (and (3.24)) are annihilated by the vector fields
$$V_{n\stackrel{\text{.}}{\alpha }}:=\frac{}{t^{n+1\stackrel{\text{.}}{\alpha }}}\zeta \frac{}{t^{n\stackrel{\text{.}}{\alpha }}},$$
(3.30)
i.e. we have
$$V_{n\stackrel{\text{.}}{\alpha }}f_+(t,\zeta )=0,$$
(3.31)
where $`n=0,\mathrm{},2J1,\stackrel{\text{.}}{\alpha }=0,1`$. Substituting (3.25) into Eqs. (3.31) and applying the standard arguments (see Subsection 2.1), we obtain
$$\psi _+(t,\zeta )V_{n\stackrel{\text{.}}{\alpha }}\psi _+^1(t,\zeta )=\psi _{}(t,\zeta )V_{n\stackrel{\text{.}}{\alpha }}\psi _{}^1(t,\zeta )=A_{n+1\stackrel{\text{.}}{\alpha }}(t)\zeta \stackrel{~}{A}_{n\stackrel{\text{.}}{\alpha }}(t),$$
(3.32)
where $`A_{n+1\stackrel{\text{.}}{\alpha }}(t)`$ and $`\stackrel{~}{A}_{n\stackrel{\text{.}}{\alpha }}(t)`$ are some $`su(r)`$-valued functions of $`t=(t^{n\stackrel{\text{.}}{\alpha }})`$. We remark that $`\{\stackrel{~}{A}_{0\stackrel{\text{.}}{\alpha }}(t),A_{1\stackrel{\text{.}}{\alpha }}(t)\}`$ coincide with components $`\{A_{\alpha \stackrel{\text{.}}{\alpha }}\}`$ of gauge potential on $`^{2,2}`$. The compatibility conditions of Eqs. (3.32) have the form
$$[D_{m+1\stackrel{\text{.}}{\alpha }}\zeta \stackrel{~}{D}_{m\stackrel{\text{.}}{\alpha }},D_{n+1\stackrel{\text{.}}{\beta }}\zeta \stackrel{~}{D}_{n\stackrel{\text{.}}{\beta }}]=0$$
$$[D_{m+1\stackrel{\text{.}}{\alpha }},D_{n+1\stackrel{\text{.}}{\beta }}]=0,[\stackrel{~}{D}_{m\stackrel{\text{.}}{\alpha }},\stackrel{~}{D}_{n\stackrel{\text{.}}{\beta }}]=0,[D_{m+1\stackrel{\text{.}}{\alpha }},\stackrel{~}{D}_{n\stackrel{\text{.}}{\beta }}]+[\stackrel{~}{D}_{m\stackrel{\text{.}}{\alpha }},D_{n+1\stackrel{\text{.}}{\beta }}]=0,$$
(3.33)
where $`D_{m+1\stackrel{\text{.}}{\alpha }}:=_{m+1\stackrel{\text{.}}{\alpha }}+A_{m+1\stackrel{\text{.}}{\alpha }},\stackrel{~}{D}_{m\stackrel{\text{.}}{\alpha }}:=_{m\stackrel{\text{.}}{\alpha }}+\stackrel{~}{A}_{m\stackrel{\text{.}}{\alpha }},_{m\stackrel{\text{.}}{\alpha }}:=/t^{m\stackrel{\text{.}}{\alpha }}`$. Equations (3.33) are equations of the truncated SDYM hierarchy. The SDYM hierarchy equations are obtained when $`J\mathrm{}`$ .
It is not difficult to verify that for the gauge fixing conditions (3.26), (3.27) we have
$$\stackrel{~}{A}_{m\stackrel{\text{.}}{\alpha }}=0,A_{m+1\stackrel{\text{.}}{\alpha }}(t)=\mathrm{\Phi }^1(t)_{m+1\stackrel{\text{.}}{\alpha }}\mathrm{\Phi }(t)=_{m\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }(t).$$
(3.34)
When we represent $`A_{m+1\stackrel{\text{.}}{\alpha }}`$ by $`\mathrm{\Phi }`$, Eqs. (3.33) reduce to
$$_{m\stackrel{\text{.}}{\alpha }}(\mathrm{\Phi }^1_{n+1\stackrel{\text{.}}{\beta }}\mathrm{\Phi })_{n\stackrel{\text{.}}{\beta }}(\mathrm{\Phi }^1_{m+1\stackrel{\text{.}}{\alpha }}\mathrm{\Phi })=0,$$
(3.35)
and when we represent $`A_{m+1\stackrel{\text{.}}{\alpha }}`$ by $`\mathrm{\Psi }`$, Eqs. (3.33) reduce to
$$_{m+1\stackrel{\text{.}}{\alpha }}_{n\stackrel{\text{.}}{\beta }}\mathrm{\Psi }_{n+1\stackrel{\text{.}}{\beta }}_{m\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }+[_{m\stackrel{\text{.}}{\alpha }}\mathrm{\Psi },_{n\stackrel{\text{.}}{\beta }}\mathrm{\Psi }]=0,$$
(3.36)
where $`m,n=0,\mathrm{},2J1`$. When $`m=n=0`$, Eqs. (3.36) coincide with the SDYM equations in the Leznov form (2.22). When $`n1,m=0`$, Eqs. (3.36) are equations on symmetries $`\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }=_{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }`$,
$$_{0\stackrel{\text{.}}{\beta }}\delta _{n+1\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }_{1\stackrel{\text{.}}{\beta }}\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }+[\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi },_{0\stackrel{\text{.}}{\beta }}\mathrm{\Psi }]=0,$$
(3.37)
where $`\stackrel{\text{.}}{\alpha },\stackrel{\text{.}}{\beta }=0,1`$.
## 4 Review of the open N=2 string
From the worldsheet point of view, critical open $`N=2`$ strings in flat Kleinian space $`^{2,2}`$ are a theory of $`N=2`$ supergravity $`(h,\chi ,A)`$ on a (pseudo) Riemann surface with boundaries, coupled to two chiral $`N=2`$ massless matter multiplets $`(y,\psi )`$. The latter’s components are complex scalars (the four string coordinates) and $`SO(1,1)`$ Dirac spinors (their four NSR partners). The $`N=2`$ string Lagrangian, as first written down by Brink and Schwarz , reads
$``$ $`=`$ $`\sqrt{h}\{\frac{1}{2}h^{pq}_p\overline{y}^{\overline{A}}_qy^A+\frac{i}{2}\overline{\psi }^{\overline{A}}\gamma ^q\underset{q}{\overset{}{D}}\psi ^{+A}+A_q\overline{\psi }^{\overline{A}}\gamma ^q\psi ^{+A}`$ (4.1)
$`+(_p\overline{y}^{\overline{A}}+\overline{\psi }^{\overline{A}}\chi _p^+)\overline{\chi }_q^{}\gamma ^p\gamma ^q\psi ^{+A}+\overline{\psi }^{\overline{A}}\gamma ^p\gamma ^q\chi _p^+(_qy^A+\overline{\chi }_q^{}\psi ^{+A})\}\eta _{\overline{A}A},`$
where $`h_{pq}`$ and $`A_q`$, with $`p,q=0,1`$, are the (real) worldsheet metric and $`U(1)`$ gauge connection, respectively. The worldsheet gravitino $`\chi _q^+`$ as well as the matter fields $`y^A`$ and $`\psi ^{+A}`$ are complex valued, so that the spacetime index $`A,\overline{A}=1,2`$ runs over two values only. Complex conjugation reads
$$(y^A)^{}=\overline{y}^{\overline{A}}\mathrm{but}(\psi ^{+A})^{}=\psi ^{\overline{A}}\mathrm{and}(\chi _q^+)^{}=\chi _q^{},$$
(4.2)
and $`\eta _{\overline{A}A}=\mathrm{diag}(+)`$ is the flat metric in $`^{1,1}`$. As usual, $`\{\gamma ^q\}`$ are a set of $`SO(1,1)`$ worldsheet gamma matrices, $`\overline{\psi }=\psi ^{}\gamma ^0`$, and $`D_q`$ denotes the worldsheet gravitationally covariant derivative.
Since open-string world-sheets have boundaries (and possibly cross-caps), boundary conditions are to be specified. As usual, the auxiliary supergravity fields remain free, while the string coordinates $`y`$ are subject to
$$_{\mathrm{normal}}y^A|_{\mathrm{boundary}}=0\mathrm{or}y^A|_{\mathrm{boundary}}=y_0^A=\mathrm{constant},$$
(4.3)
parallel resp. orthogonal to whatever D-branes are present. The two components of the spinors $`\psi `$ are related at each boundary segment $`\mathrm{\Gamma }_s`$ by multiplication with a phase $`e^{i\rho _s}`$. A boundary puncture (vertex) separating segments $`\mathrm{\Gamma }_s`$ and $`\mathrm{\Gamma }_t`$ thus carries a “twist” $`\rho _{st}=\rho _s\rho _t`$. The latter is a property of the corresponding asymptotic string state and interpolates between the traditional Neveu-Schwarz ($`\rho _{st}=1`$) and Ramond ($`\rho _{st}=1`$) sectors. The abelian R gauge symmetry of the $`N=2`$ worldsheet supergravity allows one to rotate all these phases to unity; in the superconformal gauge, its remnant is known as the spectral flow of the $`N=2`$ superconformal constraint algebra. Hence, we may restrict ourselves to the Neveu-Schwarz sector.
The Brink-Schwarz formulation (4.1) entails the choice of a complex structure on the Kleinian target space. A given complex structure breaks the global “Lorentz” invariance of $`^{2,2}`$,
$$\mathrm{Spin}(2,2)=SU(1,1)\times SU(1,1)^{}U(1)\times SU(1,1)^{}U(1,1).$$
(4.4)
The moduli space of complex structures is the two-sheeted hyperboloid $`H^2=H_+^2H_{}^2`$ with $`H_\pm ^2SU(1,1)/U(1)`$. It can be completed to $`CP^1`$ by sewing the two sheets together along a circle,
$$CP^1=H_+^2S^1H_{}^2.$$
(4.5)
Instead of using complex coordinates adapted to $`SU(1,1)^{}`$, one may alternatively choose a basis appropriate for $`SL(2,)^{}`$ and employ a real notation for the string coordinates,
$$y^1=x^1+ix^2,y^2=x^3+ix^4,$$
(4.6)
by expressing the real coordinates $`x^\mu `$, $`\mu ,\nu ,\mathrm{}=1,2,3,4`$, in $`SL(2,)\times SL(2,)^{}`$ spinor notation,
$$x^{\alpha \stackrel{\text{.}}{\alpha }}=\sigma _\mu ^{\alpha \stackrel{\text{.}}{\alpha }}x^\mu =\left(\begin{array}{cc}x^4+x^2& x^1x^3\\ x^1+x^3& x^4x^2\end{array}\right),\alpha \{0,1\},\stackrel{\text{.}}{\alpha }\{\stackrel{\text{.}}{0},\stackrel{\text{.}}{1}\},$$
(4.7)
with the help of chiral gamma matrices $`\sigma _\mu `$ appropriate for the spacetime metric $`\eta _{\mu \nu }=\mathrm{diag}(++)`$.
In the real formulation, the tangent space at any point of $`^{2,2}`$ can be split to $`^2^2`$ which defines a real polarization or cotangent structure $`^{2,2}=T^{}^2^2\times ^2`$. Such a polarization is characterized by a pair of null planes $`^2`$, and the latter are determined by a real null two-form modulo scale or, equivalently, by a real $`SL(2,)`$ spinor $`v`$ modulo scale. Indeed, each null vector $`(u_{\alpha \stackrel{\text{.}}{\alpha }})`$ factorizes into two real spinors, $`u_{\alpha \stackrel{\text{.}}{\alpha }}=v_\alpha w_{\stackrel{\text{.}}{\alpha }}`$. Choosing coordinates such that $`\left(\genfrac{}{}{0pt}{}{v_0}{v_1}\right)=\left(\genfrac{}{}{0pt}{}{t}{0}\right)`$, it becomes clear that a given null plane is stable under the action of
$$B_+\times SL(2,)^{},\mathrm{with}B_+:=\{\left(\genfrac{}{}{0pt}{}{ab}{0a^1}\right):a^{},b\},$$
(4.8)
where $`B_+`$ acts on $`v`$ and $`SL(2,)^{}`$ on $`w`$. The moduli space of cotangent structures thus becomes
$$\mathrm{Spin}(2,2)/[B_+\times SL(2,)^{}]SL(2,)/B_+S^1$$
(4.9)
which in fact is just the $`S^1`$ in (4.5). However, it turns out that the real spinor $`v`$ also encodes the two string couplings,
$$\left(\genfrac{}{}{0pt}{}{v_0}{v_1}\right)=\sqrt{g}\left(\genfrac{}{}{0pt}{}{\mathrm{cos}\frac{\theta }{2}}{\mathrm{sin}\frac{\theta }{2}}\right)$$
(4.10)
with $`g^+`$ being the gauge coupling and $`\theta S^1`$ the instanton angle. Since $`v`$ (including scale) is inert only under the parabolic subgroup of $`B_+`$ obtained by putting $`a=1`$, the space of string couplings is that of nonzero real $`SL(2,)`$ spinors,
$$^+\times S^1^2\{0\}\{0\}\sqrt{g}e^{i\theta /2}.$$
(4.11)
Consequently, fixing the values of the string couplings amounts to breaking the global “Lorentz” invariance of $`^{2,2}`$ in a way different from (4.4),
$$\mathrm{Spin}(2,2)=SL(2,)\times SL(2,)^{}\times SL(2,)^{},$$
(4.12)
where $`B_+(a=1)`$ from Eq. (4.8).
The $`N=2`$ supergravity multiplet defines a gravitini and a Maxwell bundle over the worldsheet Riemann surface in the presence of boundaries, cross-caps, and boundary punctures. The topology of the total space is labeled by the Euler number $`\chi `$ of this Riemann surface and the first Chern number (instanton number) $`M`$ of its Maxwell bundle. It is notationally convenient to replace the Euler number by the “spin”
$$J:=2\chi =n4+2(\mathrm{\#}\mathrm{boundaries})+2(\mathrm{\#}\mathrm{cross}\mathrm{caps})+4(\mathrm{\#}\mathrm{handles}).$$
(4.13)
The Lagrangian is to be integrated over the string worldsheet of a given topology. The first-quantized string path integral for the $`n`$-point function $`A^{(n)}`$ includes a sum over worldsheet topologies $`(J,M)`$, weighted with appropriate powers in the (dimensionless) string couplings $`(g,e^{i\theta })`$:
$`A^{(n)}(g,\theta )={\displaystyle \underset{J=n2}{\overset{\mathrm{}}{}}}g^JA_J^{(n)}(\theta )`$ $`=`$ $`{\displaystyle \underset{J=n2}{\overset{\mathrm{}}{}}}g^J{\displaystyle \underset{M=J}{\overset{+J}{}}}e^{iM\theta }A_{J,M}^{(n)c}`$
$`=`$ $`{\displaystyle \underset{J=n2}{\overset{\mathrm{}}{}}}g^J{\displaystyle \underset{M=J}{\overset{+J}{}}}\left(\genfrac{}{}{0pt}{}{2J}{J+M}\right)\mathrm{sin}^{JM}\frac{\theta }{2}\mathrm{cos}^{J+M}\frac{\theta }{2}A_{J,M}^{(n)r},`$
where the instanton sum has a finite range because bundles with $`|M|>J`$ do not contribute. The presence of Maxwell instantons breaks the explicit $`U(1)`$ factor in (4.4) but the $`SU(1,1)`$ factor (and thus the whole $`\mathrm{Spin}(2,2)`$) is fully restored if we let $`\sqrt{g}(e^{i\theta /2},e^{i\theta /2})`$ transform as an $`SU(1,1)`$ spinor. The partial amplitudes $`A_{J,M}^{(n)c}`$ (complex) and $`A_{J,M}^{(n)r}`$ (real) are integrals over the metric, gravitini, and Maxwell moduli spaces. The integrands may be obtained as correlation functions of boundary vertex operators in the $`N=2`$ superconformal field theory on the worldsheet surface of fixed shape (moduli) and topology.
The vertex operators generate from the (first-quantized) vacuum state the asymptotic string states in the scattering amplitude under consideration. They uniquely correspond to the physical states of the $`N=2`$ open string and carry their quantum numbers. The physical subspace of the $`N=2`$ string Fock space in a covariant quantization scheme turns out to be surprisingly small : Only the ground state $`|k,a`$ remains, a scalar on the massless level, i.e. for center-of-mass momentum $`k^A`$ with $`\overline{k}k:=\eta _{\overline{A}A}\overline{k}^{\overline{A}}k^A=0`$. In the presence of coincident D-branes, open strings stretch between the various branes. The open-string states encodes this information by carrying a Chan-Paton label $`a`$ which transforms in the adjoint representation of the gauge algebra. Thus, the dynamics of this string “excitation” is described by a Lie-algebra-valued massless scalar field,
$$\mathrm{\Xi }(y)=d^4ke^{i(\overline{k}y+k\overline{y})}\stackrel{~}{\mathrm{\Xi }}^a(k)T^a,$$
(4.15)
with $`T^a`$ denoting a set of Lie algebra generators. The self-interactions of this field are determined on-shell from the (amputated tree-level) string scattering amplitudes,
$$\stackrel{~}{\mathrm{\Xi }}^{a_1}(k_1)\stackrel{~}{\mathrm{\Xi }}^{a_2}(k_2)\mathrm{}\stackrel{~}{\mathrm{\Xi }}^{a_n}(k_n)_{\mathrm{tree},\theta }^{\mathrm{amp}}=:A_{n2}^{(n)}(\{k_i\};\{a_j\};\theta )=:\delta _{k_1+\mathrm{}+k_n}\stackrel{~}{A}_{n2}^{(n)}(\{k_i\};\{a_j\};\theta ).$$
(4.16)
Interestingly, it has been shown that all tree-level $`n`$-point functions vanish on-shell, except for the two- and three-point amplitudes,
$`\stackrel{~}{A}_0^{(2)}(k_1,k_2;a,b;\theta )`$ $`=`$ $`k^{ab},`$ (4.17)
$`\stackrel{~}{A}_1^{(3)}(k_1,k_2,k_3;a,b,c;\theta )`$ $`=`$ $`{\displaystyle \frac{i}{2}}f^{abc}\left(ϵ_{AB}k_1^Ak_2^Be^{i\theta }\eta _{A\overline{B}}(k_1^A\overline{k}_2^{\overline{B}}\overline{k}_1^{\overline{B}}k_2^A)ϵ_{\overline{A}\overline{B}}\overline{k}_1^{\overline{A}}\overline{k}_2^{\overline{B}}e^{i\theta }\right)`$ (4.18)
$`=`$ $`f^{abc}ϵ_{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}\left(k_1^{0\stackrel{\text{.}}{\alpha }}\mathrm{cos}\frac{\theta }{2}+k_1^{1\stackrel{\text{.}}{\alpha }}\mathrm{sin}\frac{\theta }{2}\right)\left(k_2^{0\stackrel{\text{.}}{\beta }}\mathrm{cos}\frac{\theta }{2}+k_2^{1\stackrel{\text{.}}{\beta }}\mathrm{sin}\frac{\theta }{2}\right).`$
Here, the Chan-Paton labels appear on the Killing form $`k^{ab}`$ and the (totally antisymmetric) structure constants $`f^{abc}`$, and the momenta obey $`\overline{k}_ik_j+\overline{k}_jk_i=0`$ due to $`_nk_n=0`$. Note that $`\stackrel{~}{A}_1^{(3)}`$ is totally symmetric in the external state quantum numbers $`(k_i,a_i)`$.
Since we argue that the string couplings $`(g,e^{i\theta })`$ can be changed at will by global “Lorentz” transformations, it is admissible to make a convenient choice of Lorentz frame. First, we may scale $`g1`$ (i.e. put the constant dilaton to zero). Second, the instanton angle $`\theta `$ is at our disposal. In the real notation, one sees that taking $`\theta =0`$ reduces the amplitude (4.18) to a single term ,
$$\stackrel{~}{A}_1^{(3)}(k_1,k_2,k_3;a,b,c;\theta =0)=f^{abc}ϵ_{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}k_1^{0\stackrel{\text{.}}{\alpha }}k_2^{0\stackrel{\text{.}}{\beta }},$$
(4.19)
which, renaming $`\mathrm{\Xi }\mathrm{\Psi }`$, translates to a cubic interaction<sup>1</sup><sup>1</sup>1 The $`SO(2,2)`$ transformation properties of this interaction become manifest when this term is rewritten as $`\frac{1}{6}T^{(+)\alpha \beta }ϵ^{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}\mathrm{\Psi }_{\alpha \stackrel{\text{.}}{\alpha }}\mathrm{\Psi }_{\beta \stackrel{\text{.}}{\beta }}\mathrm{\Psi }`$, with a self-dual projector $`T^{(+)}`$ having nonzero components $`T^{(+)00}=1`$ only.
$$_{\mathrm{int}}=\frac{1}{6}ϵ^{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}\mathrm{tr}\mathrm{\Psi }[_{0\stackrel{\text{.}}{\alpha }}\mathrm{\Psi },_{0\stackrel{\text{.}}{\beta }}\mathrm{\Psi }].$$
(4.20)
It is remarkable that (at least at tree-level) no quartic or higher field-theory vertices are needed to reproduce the vanishing string amplitudes, $`A_{n2}^{(n4)}=0`$, because the Feynman graphs based on (4.20) alone happen to cancel in $`2+2`$ dimensions. In this sense, the cubic Lagrangian (4.20) is tree-level exact. Its resulting equation of motion reads
$$\mathrm{}\mathrm{\Psi }+\frac{1}{2}ϵ^{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}[_{0\stackrel{\text{.}}{\alpha }}\mathrm{\Psi },_{0\stackrel{\text{.}}{\beta }}\mathrm{\Psi }]=0$$
(4.21)
which we recognize as Leznov’s equation (2.22) . It describes the dynamics of the single-helicity ($`h=+1`$) gluon in $`2+2`$ self-dual Yang-Mills theory. More precisely, the self-dual field strength $`F_{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}`$ is entirely expressed (in light-cone gauge) through Leznov’s prepotential $`\mathrm{\Psi }`$ ,
$$F_{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}=_{0\stackrel{\text{.}}{\alpha }}_{0\stackrel{\text{.}}{\beta }}\mathrm{\Psi },$$
(4.22)
which is subject to the second-order equation (4.21).
In the complex notation, the $`U(1)`$ factor in (4.4) can be restored by averaging over all cotangent structures. In this manner, $`\stackrel{~}{A}_1^{(3)}`$ simplifies to
$`{\displaystyle \frac{d\theta }{2\pi }\stackrel{~}{A}_1^{(3)}(k_1,k_2,k_3;a,b,c;\theta )}`$ $`=`$ $`\frac{1}{2}\left[\stackrel{~}{A}_1^{(3)}(\theta =0)+\stackrel{~}{A}_1^{(3)}(\theta =\pi )\right]`$ (4.23)
$`=`$ $`\frac{i}{2}f^{abc}\eta _{A\overline{B}}(k_1^A\overline{k}_2^{\overline{B}}\overline{k}_1^{\overline{B}}k_2^A)`$
$`=`$ $`\frac{1}{2}f^{abc}ϵ_{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}(k_1^{0\stackrel{\text{.}}{\alpha }}k_2^{0\stackrel{\text{.}}{\beta }}+k_1^{1\stackrel{\text{.}}{\alpha }}k_2^{1\stackrel{\text{.}}{\beta }})`$
which, renaming $`\mathrm{\Xi }\varphi `$, leads to a cubic vertex
$$_{\mathrm{int}}^{(3)}=\frac{i}{6}\eta ^{A\overline{B}}\mathrm{tr}\varphi [_A\varphi ,_{\overline{B}}\varphi ].$$
(4.24)
The Feynman rules based on this vertex yield non-vanishing $`n`$-point functions for all $`n4`$ which, however, may be cancelled recursively (at tree-level) by supplementing an infinite set of judiciously chosen higher vertices,
$$_{\mathrm{int}}=_{\mathrm{int}}^{(3)}+_{\mathrm{int}}^{(4)}+\mathrm{},$$
(4.25)
resulting in a non-polynomial but tree-level exact Lagrangian. Surprisingly, the corresponding equation of motion,
$$\mathrm{}\varphi +\frac{i}{2}\eta ^{A\overline{B}}[_A\varphi ,_{\overline{B}}\varphi ]+O(\varphi ^3)=0,$$
(4.26)
can be written in closed form after reassembling the Lie-algebra-valued field $`\varphi `$ to the group-valued field $`\mathrm{\Phi }=e^{i\varphi }`$,
$$\eta ^{A\overline{B}}_A(e^{i\varphi }_{\overline{B}}e^{i\varphi })=0.$$
(4.27)
The latter is nothing but Yang’s equation (2.20) . Like (4.21), it describes $`2+2`$ self-dual Yang-Mills but in a different parametrization. We conclude that, at tree-level, the $`N=2`$ open string is indeed identical to self-dual Yang-Mills.
## 5 Non-local symmetries of the open N=2 string
In Section 3, an infinite number of non-local symmetries of the self-dual Yang-Mills equations have been described. The latter’s intimate connection with the open $`N=2`$ string gives rise to the question where all these symmetries hide in the string description. To answer this, the BRST approach offers a systematic procedure for the construction of all conserved charges in a given quantum field theory . One should realize that this issue is a classical one from the spacetime point of view (no string loops) but involves a quantum description of the underlying (super)conformal field theory.
For closed $`N=2`$ strings, the BRST quantization has been treated exhaustively in and reviewed in the context of global symmetries in . Here, we shall only collect the facts pertinent to the open string case, especially where the treatment differs from that of the closed string. Relevant for the physics of the open $`N=2`$ string is the so-called relative chiral BRST cohomology
$$H_{\mathrm{rel}}=\frac{\mathrm{ker}Q}{\mathrm{im}Q}\mathrm{on}\mathrm{ker}b_0\mathrm{ker}b_0^{}$$
(5.1)
which is graded by<sup>2</sup><sup>2</sup>2 We use the notation of Ref. .
* ghost number $`g`$
* picture numbers $`(\pi _+,\pi _{})^2`$ (no loss of generality due to spectral flow)
* spacetime momentum $`(k^A,\overline{k}^{\overline{A}})`$ or $`k^{\alpha \stackrel{\text{.}}{\alpha }}^{2,2}`$ as well as Chan-Paton label $`a`$
so that one may restrict the analysis to Fock states of a given ghost number, built on the momentum-dressed picture vacua $`|\pi _+,\pi _{};k,a`$ of ghost number zero (by definition). It is important to distinguish the exceptional case of $`k=0`$ from the generic case $`(k0)`$ which contains the propagating modes.
It has been shown that the generic BRST cohomology is non-empty only for
* $`g=1`$
* any value for $`(\pi _+,\pi _{})`$
* any lightlike momentum, $`\overline{k}k=0`$
and is one-dimensional in each such case. Moreover, for $`k0`$ one may construct picture-raising and picture-lowering operators which commute with $`Q`$ and do not carry ghost number or momentum . Together with spectral flow, these operators may therefore be used to define an equivalence relation among all pictures. This projection of the BRST cohomology leaves us with a single, massless, physical mode taking value in the Lie algebra of the Chan-Paton group, supporting our assertion of the previous Section.
It is perhaps less well known that the exceptional (zero-momentum) BRST cohomology at ghost number one harbors all unbroken global symmetry charges of the theory . A conserved charge $`𝒜=_𝒞\mathrm{\Omega }^{(1)}`$, with the integration contour connecting the two boundaries of the free open string worldsheet, originates from a (current) one-form $`\mathrm{\Omega }^{(1)}`$ which is closed up to a BRST commutator. BRST invariance of the charge requires
$$[Q,\mathrm{\Omega }^{(1)}\}=d\mathrm{\Omega }^{(0)}$$
(5.2)
for some zero-form (function) $`\mathrm{\Omega }^{(0)}`$. Consistency then implies that
$$[Q,\mathrm{\Omega }^{(0)}\}=0\mathrm{and}\mathrm{\Omega }^{(0)}\mathrm{\Omega }^{(0)}+[Q,\mathrm{any}\}$$
(5.3)
which are precisely the defining relations for the BRST cohomology (on zero-form operators instead of Fock states). Taking as $`\mathrm{\Omega }^{(0)}`$ some cohomology class of ghost number one, we may solve the descent equation (5.2) and construct a current $`\mathrm{\Omega }^{(1)}`$ of ghost number zero which yields a conserved charge that can map physical states to physical states. In this way, the classification of global symmetries has been reduced to the computation of the $`g=1`$ exceptional relative BRST cohomology $`H_{\mathrm{rel}}^{g=1}(k=0)`$. We shall see that the picture dependence of the latter plays a crucial role.
As a simple example, consider for a moment the open bosonic string. Its ghost number one exceptional relative BRST cohomology is spanned by the operators $`\mathrm{\Omega }^{(0)}=cx^\mu `$. It is easy to see that Eq. (5.2) is solved by $`\mathrm{\Omega }^{(1)}=x^\mu dz`$. Obviously, this leads to the charge $`𝒜=p^\mu =\frac{dz}{2\pi }x^\mu `$, which is nothing but the center-of-mass momentum generating spacetime translations.
A key role in the computation of $`H_{\mathrm{rel}}^{g=1}(k=0)`$ is played by $`H_{\mathrm{rel}}^{g=0}(k=0)`$, the so-called ground ring, because it is contained in any $`H_{\mathrm{rel}}^g(k=0)`$ in the sense that
$$\omega ^gH_{\mathrm{rel}}^{g=0}(k=0)H_{\mathrm{rel}}^g(k=0)\mathrm{for}\mathrm{any}\omega ^gH_{\mathrm{rel}}^g(k=0),$$
(5.4)
where we denoted the natural product in the BRST cohomology ring by a dot. On representatives, this multiplication is nothing but the normal ordered product. In our bosonic string example, the only ghost number zero cohomology class is the unit operator, rendering the ground ring trivial. For the $`N=2`$ string, however, the ground ring was found to be infinite-dimensional, with $`\pi +1`$ generators in each picture $`\pi 1`$ . Let us briefly review this result before describing the set of conserved charges and the transformations they generate.
For convenience we change the picture labels from $`(\pi _+,\pi _{})`$ to “spin” labels $`(j,m)`$ via<sup>3</sup><sup>3</sup>3 The picture offset of the canonical ground state $`|1,1;0,a`$ is responsible for the distinction.
$`\pi _+=j+m\mathrm{and}`$ $`\pi _{}=jm`$ $`\mathrm{on}\mathrm{operators}`$
$`\pi _++1=j+m\mathrm{and}`$ $`\pi _{}+1=jm`$ $`\mathrm{on}\mathrm{states}.`$ (5.5)
One finds that the ground ring is spanned by basis elements
$$𝒪_{j,m}^{\mathrm{}}\mathrm{with}j=0,\frac{1}{2},1,\frac{3}{2},\mathrm{}m=j,j+1,\mathrm{},+j\mathrm{}=0,1,\mathrm{},2j.$$
(5.6)
which are built from the picture-raising operators $`X_\pm `$ and the spectral-flow operator $`S`$ . Under the cohomology product these operators form an infinite abelian algebra on which the picture-number operators $`\mathrm{\Pi }_\pm `$ act as derivations. This algebra and its derivations can be written more concisely in terms of polynomials in two variables $`(x,y)`$ and vector fields on the $`xy`$ plane, with the translations $`_x`$ and $`_y`$ missing.
Starting from the obvious representatives of $`H_{\mathrm{rel}}^{g=1}(k=0)`$, namely the “translations”<sup>4</sup><sup>4</sup>4 Worldsheet reparametrization and supersymmetry ghosts are denoted by $`c`$ and $`\gamma `$, respectively. Due to $`N=2`$ supersymmetry, the latter has two real components which have been collected in a matrix $`\gamma _\delta ^\beta `$ subject to $`\gamma _{0}^0=\gamma _{1}^1`$ and $`\gamma _{1}^0=\gamma _{0}^1`$.
$$P_{\alpha \stackrel{\text{.}}{\alpha }}=ϵ_{\alpha \beta }ϵ_{\stackrel{\text{.}}{\alpha }\stackrel{\text{.}}{\beta }}P^{\beta \stackrel{\text{.}}{\beta }},P^{\beta \stackrel{\text{.}}{\beta }}:=icx^{\beta \stackrel{\text{.}}{\beta }}2i\gamma _\delta ^\beta \psi ^{\delta \stackrel{\text{.}}{\beta }},$$
(5.7)
one immediately sees that
$$\mathrm{\Omega }_{j,m;\stackrel{\text{.}}{\alpha }}^{\mathrm{}(0)}:=P_{0\stackrel{\text{.}}{\alpha }}𝒪_{j,m}^{\mathrm{}}$$
(5.8)
comprise a set of $`2(2j+1)`$ independent operators in the $`(j,m)`$ sector with $`g=1`$.
In order to find the symmetry charges $`𝒜`$, we have to insert our $`g=1`$ zero-forms into the descent equation (5.2), work out the corresponding one-forms, and integrate those across the worldsheet. The result of this computation reads
$$𝒜_{j,m;\stackrel{\text{.}}{\alpha }}^{\mathrm{}}=\frac{dz}{2\pi i}\left[_z\frac{dw}{2\pi i}b(w)P_{0\stackrel{\text{.}}{\alpha }}(z)𝒪_{j,m}^{\mathrm{}}(z)\right].$$
(5.9)
Together with the derivations
$$_{j,m}^{\pm ,\mathrm{}}=𝒪_{j,m}^{\mathrm{}}(\mathrm{\Pi }_\pm +1)$$
(5.10)
these charges form an enormous non-abelian algebra.
According to Noether’s theorem the conserved charges $`𝒜`$ must generate global symmetries of the open $`N=2`$ string. The symmetry transformations of the physical state
$$|k,a\{|\pi _+,\pi _{};k,a\}/\mathrm{picture}\mathrm{changing}$$
(5.11)
(pictures are identified for $`k0`$) are found by evaluating the action of $`𝒜`$’s on $`|1,1;k,a`$,
$$𝒜_{j,m;\stackrel{\text{.}}{\alpha }}^{\mathrm{}}|k,a=𝒪_{j,m}^{\mathrm{}}p_{0\stackrel{\text{.}}{\alpha }}|k,a=h(k)^{(jm)+\mathrm{}}k_{0\stackrel{\text{.}}{\alpha }}|k,a,$$
(5.12)
with the important phase
$$h(k):=\frac{k_{0\stackrel{\text{.}}{0}}}{k_{1\stackrel{\text{.}}{0}}}=\frac{k_{0\stackrel{\text{.}}{1}}}{k_{1\stackrel{\text{.}}{1}}}.$$
(5.13)
The action of the derivations $`_{j,m}^{\pm ,\mathrm{}}`$ obtains by replacing $`k_{0\stackrel{\text{.}}{\alpha }}\pi _\pm +1`$ on the right-hand side of (5.12). The transformations (5.12) constitute an infinity of global symmetries which are unbroken in the flat Kleinian background. Their Ward identities constrain the tree-level scattering amplitudes so severely that all but the three-point function must vanish, consistent with the direct computations alluded to earlier.
To make contact with the non-local abelian symmetries of the SDYM equations (see Section 3), it suffices to consider the subalgebra of symmetry charges
$$𝒫_{n\stackrel{\text{.}}{\alpha }}:=𝒜_{j,jn;\stackrel{\text{.}}{\alpha }}^0=P_{0\stackrel{\text{.}}{\alpha }}𝒪_{j,jn}^0,$$
(5.14)
i.e. putting $`\mathrm{}=0`$. It has the important property that only non-positive powers of the phase $`h(k)`$ appear when acting by these charges on physical states,
$$\delta _{n\stackrel{\text{.}}{\alpha }}|k,a:=𝒫_{n\stackrel{\text{.}}{\alpha }}|k,a=h(k)^nk_{0\stackrel{\text{.}}{\alpha }}|k,a.$$
(5.15)
Note that for $`j=0`$ we find the translations $`\delta _{0\stackrel{\text{.}}{\alpha }}|k,a=k_{0\stackrel{\text{.}}{\alpha }}|k,a`$ as it should be. Moreover, for $`j=1/2`$ we have $`\delta _{\alpha \stackrel{\text{.}}{\alpha }}|k,a=𝒫_{\alpha \stackrel{\text{.}}{\alpha }}|k,a=k_{\alpha \stackrel{\text{.}}{\alpha }}|k,a`$. Although $`𝒫_{n\stackrel{\text{.}}{\alpha }}`$ shifts the picture $`(\pi _+,\pi _{})(\pi _++2jn,\pi _{}+n)`$ of the state representative, by virtue of the picture-changing equivalence (5.11) physical states $`|k,a`$ are eigenstates not only of the momentum $`𝒫_{\alpha \stackrel{\text{.}}{\alpha }}`$ but of all our hidden symmetry generators $`𝒫_{n\stackrel{\text{.}}{\alpha }}`$. To employ their eigenvalues $`k_{n\stackrel{\text{.}}{\alpha }}`$ as further labels of the physical state $`|k,a`$ would be customary but superfluous, since (5.15) tells us that these eigenvalues are not independent but completely given by the first two, $`k_{n\stackrel{\text{.}}{\alpha }}=h(k)^nk_{0\stackrel{\text{.}}{\alpha }}`$.
Recall that a single massless $`N=2`$ string physical state with zero instanton angle $`\theta `$ corresponds to a massless spacetime field $`\mathrm{\Psi }`$, and therefore the SDYM symmetries $`\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }`$ will correspond to the string symmetries (5.15). To compare symmetries $`\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }`$ with the string symmetries (5.15), one should turn to the momentum representation $`\mathrm{\Psi }\stackrel{~}{\mathrm{\Psi }}`$ and single out terms linear in $`\stackrel{~}{\mathrm{\Psi }}`$, since we cannot expect to see non-linear in $`\stackrel{~}{\mathrm{\Psi }}`$ transformations in the first-quantized string theory. For this consider Eqs. (3.37) in the momentum representation,
$$k_{0\stackrel{\text{.}}{\beta }}\delta _{n+1\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}k_{1\stackrel{\text{.}}{\beta }}\delta _{n\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}+O(\stackrel{~}{\mathrm{\Psi }})=0\delta _{n+1\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}h(k)^1\delta _{n\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}+O(\stackrel{~}{\mathrm{\Psi }})=0,$$
(5.16)
where $`h(k)`$ is given in (5.13). From Eqs.(5.16) we obtain
$$\delta _{n\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}=h(k)^nk_{0\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}+O(\stackrel{~}{\mathrm{\Psi }}),$$
(5.17)
since $`\delta _{0\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}=k_{0\stackrel{\text{.}}{\alpha }}\stackrel{~}{\mathrm{\Psi }}`$. The identity of (5.15) to (5.17) is evident. Thus, the transformations (5.15) of the string ground state $`|k,a`$ corresponding to the field $`\stackrel{~}{\mathrm{\Psi }}(k)`$ precisely reproduce the linear part of the symmetries (3.29) of the self-dual Yang-Mills equations. Recall that these symmetries generate flows on the moduli space of self-dual gauge fields.
## 6 Conclusion
This work lifts to a new level the identification of open $`N=2`$ strings (at tree-level) with self-dual Yang-Mills (SDYM) on $`4D`$ manifolds of signature $`(++)`$. It provides further evidence that $`N=2`$ strings inherit integrability from self-dual Yang-Mills theory. After recapitulating the SDYM equations and their symmetries in the twistor framework, as well as reviewing the open $`N=2`$ string and its rigid symmetries in the first-quantized BRST approach, we have demonstrated complete agreement on the linearized level.
Interestingly, the stringy source of those symmetries is a non-trivial ground ring of ghost number zero operators in the chiral BRST cohomology. Such a phenomenon is familiar from the non-critical $`2D`$ string, where the ground ring was exploited to investigate the global symmetries of the theory, with the result that there are more discrete states and associated symmetries in $`2D`$ string theory than had been recognized previously . The authors of Ref. have wondered if their findings “could be relevant in a realistic string theory with a macroscopic four-dimensional target space”. The outcomes of and of this paper answer their question in the affirmative by providing the first four-dimensional if not yet realistic string theory (open as well as closed) with a rich symmetry structure based on an infinite ground ring.
More concretely, the symmetry charges are constructed from zero-momentum operators of picture-raising $`X_\pm `$, picture charge $`\mathrm{\Pi }_\pm `$, spectral flow $`S`$, and momentum operators $`P_{\alpha \stackrel{\text{.}}{\alpha }}`$. The abelian subalgebra generated by $`X_\pm `$, $`S`$, and $`P_{0\stackrel{\text{.}}{\alpha }}`$ was found to coincide with the algebra of non-local abelian symmetries of the SDYM equations produced by the operators $`_{0\stackrel{\text{.}}{\alpha }}`$ and the recursion operator $`:\delta _{n\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }\delta _{n+1\stackrel{\text{.}}{\alpha }}\mathrm{\Psi }`$ defined by Eqs. (3.37).
Our results ascertain that the non-trivial picture structure of the BRST cohomology is not just an irrelevant technical detail of the BRST approach but indispensable for a deeper understanding of the theory. It is, of course, not a simple task to discover the full symmetry group of a string model. Doing so would roughly correspond to having found a useful non-perturbative definition of the theory. In this paper we have worked in the standard first-quantized formalism which is background-dependent and limits our access to unbroken linear global symmetries.
There remain a number of interesting unresolved issues which should be addressed. Prominent among them is the detection of the non-abelian symmetries of the SDYM equations in the $`N=2`$ string context. For this goal, a string field theoretic setup should be more effective. Another point concerns the quantum extension of our hidden non-local symmetries. On the string theory side, it appears that their Ward identities forbid any scattering (beyond three-point) not only at tree- but also at the loop-level . On the field theory side, such a feature would seem to select a particular quantum version of SDYM, different from the one yielding the celebrated MHV amplitudes . Finally, it would be interesting to find further examples in which the picture structure yields non-trivial information about a theory, like it happens for the relative zero-momentum cohomology of the Ramond sector of the $`N=1`$ string in flat $`9+1`$ dimensional spacetime . We hope to return to these problems soon.
### Acknowledgements
The work of T.A.I. was partially supported by the Heisenberg-Landau Program and the grant RFBR-99-01-01076. T.A.I. thanks for hospitality the Institut für Theoretische Physik der Universität Hannover, where part of this work was done. O.L. is grateful to the Erwin Schrödinger International Institute for Mathematical Physics in Vienna, where this work was completed.
## Appendix A Appendices
A.1 Line bundles over the Riemann sphere
The Riemann sphere $`P^1S^2`$ is the complex manifold obtained by patching together two coordinate patches $`\mathrm{\Omega }_+`$ and $`\mathrm{\Omega }_{}`$, with $`\mathrm{\Omega }_+`$, the neighbourhood of $`\zeta =0`$, and $`\mathrm{\Omega }_{}`$, the neighbourhood of $`\zeta =\mathrm{}`$. For example, if
$$\mathrm{\Omega }_+=\{\zeta :|\zeta |<\mathrm{}\},\mathrm{\Omega }_{}=\{\zeta \{\mathrm{}\}:|\zeta |>0\},$$
(A.1)
then we can use $`\zeta `$ as the coordinate on $`\mathrm{\Omega }_+`$ and $`\stackrel{~}{\zeta }=\zeta ^1`$ as the coordinate on $`\mathrm{\Omega }_{}`$.
Consider the holomorphic line bundle $`𝒪(n)`$ over $`P^1`$ with the transition function $`\zeta ^n`$, and the first Chern class $`c_1(𝒪(n))=n`$. The space $`𝒪(n)`$ is a two-dimensional complex manifold. It can be covered by two coordinate patches, $`𝒪(n)=U_+U_{}`$ with the coordinates $`(\gamma _+,\zeta )`$ on $`U_+`$ and $`(\gamma _{},\stackrel{~}{\zeta })`$ on $`U_{}`$. The projection
$$𝒪(n)P^1$$
(A.2)
is given by $`(\gamma _+,\zeta )\zeta ,(\gamma _{},\stackrel{~}{\zeta })\stackrel{~}{\zeta }`$ in these coordinates. On the overlap region $`U_+U_{}`$ the coordinates are related by
$$(\gamma _{},\stackrel{~}{\zeta })=(\zeta ^n\gamma _+,\zeta ^1).$$
(A.3)
The space $`\mathrm{\Gamma }(𝒪(n))`$ of global holomorphic sections of the bundle $`𝒪(n)`$ coincides with the space of polynomials of degree $`n`$ in $`\zeta `$ with complex coefficients, $`\mathrm{\Gamma }(𝒪(n))=^{n+1}`$. Points $`\sigma _t\mathrm{\Gamma }(𝒪(n))`$ are complex projective lines $`P_t^1𝒪(n)`$,
$$P_t^1=\sigma _t(\zeta )=\{\begin{array}{cc}\gamma _+=_{i=0}^nt_i\zeta ^i,& \zeta \mathrm{\Omega }_+,\\ \gamma _{}=_{i=0}^nt_i\zeta ^{in},& \zeta \mathrm{\Omega }_{},\end{array}$$
(A.4)
parametrized by points $`t=\{t_i\}^{n+1}`$.
On the complex space $`𝒪(n)`$ one can introduce a map $`\tau :𝒪(n)𝒪(n)`$ called the real structure. It is an antiholomorphic involution, defined by the formula
$$\tau (\gamma ,\zeta )=(\overline{\gamma },\overline{\zeta }),$$
(A.5)
where $`(\gamma ,\zeta )`$ are local coordinates on $`𝒪(n)`$ and the bar denotes the complex conjugate. There are fixed points of the action $`\tau `$ on $`𝒪(n)`$ and they form a two-dimensional real manifold $`𝒪_{}(n)S^1\times `$ fibred over $`S^1`$,
$$𝒪_{}(n)S^1,$$
(A.6)
and parametrized by the real coordinates $`(\gamma _+,\zeta )^2,(\gamma _{},\zeta ^1)^2`$. Here real $`\zeta `$ and $`\zeta ^1`$ parametrize the equator $`P^1=S^1\{\mathrm{}\}`$ on the sphere $`P^1`$.
Those sections $`\sigma _t\mathrm{\Gamma }(𝒪(n))`$ which are preserved under the conjugation (A.5) are called the real sections of the bundle (A.2), and form a subset $`\mathrm{\Gamma }(𝒪_{}(n))\mathrm{\Gamma }(𝒪(n))`$. Points $`\sigma _t\mathrm{\Gamma }(𝒪_{}(n))`$ are real projective lines $`P_t^1=S_t^1𝒪_{}(n)𝒪(n)`$,
$$S_t^1=\sigma _t(\zeta )=\{\begin{array}{cc}\gamma _+=_{i=0}^nt_i\zeta ^i,& \zeta ,\\ \gamma _{}=_{i=0}^nt_i\zeta ^{in},& \zeta \{\mathrm{}\}\{0\},\end{array}$$
(A.7)
parametrized by points $`t=\{t_i\}^{n+1}`$. We also consider real holomorphic sections of the bundle (A.2), which are defined by formulae (A.4) with complex $`\zeta P^1`$, real $`t=\{t_i\}^{n+1}`$ and satisfy the reality condition
$$\overline{\gamma _+(t,\overline{\zeta })}=\gamma _+(t,\zeta )\overline{(\underset{i=0}{\overset{n}{}}t_i\overline{\zeta }^i)}=\underset{i=0}{\overset{n}{}}t_i\zeta ^i$$
and analogously for $`(\gamma _{}(t,\stackrel{~}{\zeta }),\stackrel{~}{\zeta })`$. We denote the space of such sections by $`\mathrm{\Gamma }_{}(𝒪(n))`$. It is easy to see that $`\mathrm{\Gamma }_{}(𝒪(n))\mathrm{\Gamma }(𝒪_{}(n))`$.
Recall that the sphere $`P^1=\{\mathrm{}\}`$ can be decomposed into the disjoint union,
$$P^1H_+^2S^1H_{}^2,$$
(A.8)
where $`H_+^2`$ and $`H_{}^2`$ are upper and lower half-planes of the extended complex plane $`\{\mathrm{}\}`$. Moreover, (A.8) is exactly the decomposition of $`P^1`$ into three orbits $`H_+^2SL(2,)/SO(2),H_{}^2SL(2,)/SO(2)`$ and $`S^1SL(2,)/B_+`$ of the group $`SL(2,)`$ acting on $`P^1`$. Here
$$B_+:=\{\left(\begin{array}{cc}a& b\\ 0& a^1\end{array}\right):a^{},b\}$$
(A.9)
is a two-dimensional subgroup of $`SL(2,)`$ and $`S^1`$ is the equator on $`P^1`$ (the real axis $`\text{Im}\zeta =0`$ on $`\{\mathrm{}\}`$).
Notice that the linear-fractional transformation
$$\zeta \lambda =\frac{\zeta i}{\zeta +i}$$
(A.10)
carries the upper half-plane $`\text{Im}\zeta >0`$ to the unit disk $`|\lambda |<1`$, the lower half-plane $`\text{Im}\zeta <0`$ to the domain $`|\lambda |>1`$ and the real axis $`\text{Im}\zeta =0`$ to the circle $`|\lambda |=1`$. Then the conjugation operation $`\zeta \overline{\zeta }`$ is replaced by $`\lambda \overline{\lambda }^1`$ and the reality condition has a different form. If we parametrize the circle $`|\lambda |=1`$ by $`\lambda =e^{i\theta },0\theta 2\pi `$, then from (A.10) we obtain
$$\zeta =\mathrm{cot}\frac{\theta }{2}.$$
(A.11)
In some formulae it is convenient to parametrize $`P^1=\{\mathrm{}\}`$ by homogeneous coordinates
$$v_0=g^{1/2}\mathrm{sin}\frac{\theta }{2},v_1=g^{1/2}\mathrm{cos}\frac{\theta }{2}\zeta =\frac{v_1}{v_0}=\mathrm{cot}\frac{\theta }{2},$$
(A.12)
where $`g`$ is a positive constant. In these coordinates, real sections of the bundle (A.2) have the form
$$\sigma _t(\theta )=g^{n/2}\underset{i=0}{\overset{n}{}}t_i\mathrm{cos}^i\frac{\theta }{2}\mathrm{sin}^{ni}\frac{\theta }{2}=g^J\underset{M=J}{\overset{J}{}}t_{J+M}\mathrm{cos}^{J+M}\frac{\theta }{2}\mathrm{sin}^{JM}\frac{\theta }{2},$$
(A.13)
where $`J=n/2`$.
A.2 Twistor spaces
Let us consider the rank 2 holomorphic vector bundle
$$𝒫=𝒪(1)^2=𝒪(1)𝒪(1)$$
(A.14)
over the Riemann sphere $`P^1`$ with a holomorphic projection
$$\pi :𝒫P^1.$$
(A.15)
Each fibre of the bundle (A.15) is a copy of $`^2`$. The space $`𝒫`$ can be covered by two coordinate patches, $`𝒫=𝒫_+𝒫_{}`$, with the coordinates $`(\eta _+^{\stackrel{\text{.}}{0}},\eta _+^{\stackrel{\text{.}}{1}},\zeta )`$ on $`𝒫_+`$ and $`(\eta _{}^{\stackrel{\text{.}}{0}},\eta _{}^{\stackrel{\text{.}}{1}},\stackrel{~}{\zeta })`$ on $`𝒫_{}`$ related by
$$(\eta _{}^{\stackrel{\text{.}}{0}},\eta _{}^{\stackrel{\text{.}}{1}},\stackrel{~}{\zeta })=(\zeta ^1\eta _+^{\stackrel{\text{.}}{0}},\zeta ^1\eta _+^{\stackrel{\text{.}}{1}},\zeta ^1)$$
(A.16)
on the overlap $`𝒫_+𝒫_{}`$. The space $`𝒫`$ will be called the twistor space. It is an open subset of $`P^3`$, $`𝒫P^3P^1S^2\times ^4`$.
Global holomorphic sections $`\sigma _x`$ of the bundle (A.15) are complex projective lines $`P_x^1𝒫`$,
$$P_x^1=\sigma _x(\zeta )=\{\begin{array}{ccc}\eta _+^{\stackrel{\text{.}}{0}}=x^{0\stackrel{\text{.}}{0}}+\zeta x^{1\stackrel{\text{.}}{0}},& \eta _+^{\stackrel{\text{.}}{1}}=x^{0\stackrel{\text{.}}{1}}+\zeta x^{1\stackrel{\text{.}}{1}},& \zeta \mathrm{\Omega }_+,\\ \eta _{}^{\stackrel{\text{.}}{0}}=\zeta ^1x^{0\stackrel{\text{.}}{0}}+x^{1\stackrel{\text{.}}{0}},& \eta _{}^{\stackrel{\text{.}}{1}}=\zeta ^1x^{0\stackrel{\text{.}}{1}}+x^{1\stackrel{\text{.}}{1}},& \zeta \mathrm{\Omega }_{},\end{array}$$
(A.17)
parametrized by $`x=\{x^{\alpha \stackrel{\text{.}}{\alpha }}\}^4,\alpha =0,1,\stackrel{\text{.}}{\alpha }=0,1.`$
We introduce a real structure $`\tau `$ on $`𝒫`$ by the formula
$$\tau (\eta ^{\stackrel{\text{.}}{\alpha }},\zeta )=(\overline{\eta ^{\stackrel{\text{.}}{\alpha }}},\overline{\zeta }).$$
(A.18)
There are fixed points of the action $`\tau `$ on $`𝒫`$ and they form a 3-dimensional real manifold
$$𝒯=P^3P^1S^1\times ^2$$
(A.19)
with a projection
$$𝒯S^1.$$
(A.20)
The space $`𝒯𝒫`$ is called the real twistor space.
Real sections of the bundle (A.15) over $`\mathrm{\Omega }_+`$ are defined by the formulae
$$\sigma _x(\zeta )=(\eta _+^{\stackrel{\text{.}}{0}}(\zeta ),\eta _+^{\stackrel{\text{.}}{1}}(\zeta ))=(x^{0\stackrel{\text{.}}{0}}+\zeta x^{1\stackrel{\text{.}}{0}},x^{0\stackrel{\text{.}}{1}}+\zeta x^{1\stackrel{\text{.}}{1}}),\zeta ,$$
(A.21)
and over $`\mathrm{\Omega }_{}`$ by the formulae
$$\sigma _x(\stackrel{~}{\zeta })=(\eta _{}^{\stackrel{\text{.}}{0}}(\stackrel{~}{\zeta }),\eta _{}^{\stackrel{\text{.}}{1}}(\stackrel{~}{\zeta }))=(\stackrel{~}{\zeta }x^{0\stackrel{\text{.}}{0}}+x^{1\stackrel{\text{.}}{0}},\stackrel{~}{\zeta }x^{0\stackrel{\text{.}}{1}}+x^{1\stackrel{\text{.}}{1}}),\stackrel{~}{\zeta }=\zeta ^1,$$
(A.22)
with $`x=\{x^{\alpha \stackrel{\text{.}}{\alpha }}\}^4`$. Real holomorphic sections of the bundle (A.15) are defined by formulae (A.17) with real $`x=\{x^{\alpha \stackrel{\text{.}}{\alpha }}\}^4`$ and complex $`\zeta P^1`$.
On the space $`^4`$ of real sections of the bundle (A.15) isomorphic to the space of real holomorphic sections, one may introduce the metric $`\eta =diag(+1,+1,1,1)`$, and other signatures are not compatible with the real structure (A.18) on the twistor space $`𝒫`$. Namely, we have
$$ds^2=det(dx^{\alpha \stackrel{\text{.}}{\alpha }})=\eta _{\mu \nu }dx^\mu dx^\nu ,$$
(A.23)
where
$$x^1:=\frac{1}{2}(x^{0\stackrel{\text{.}}{1}}+x^{1\stackrel{\text{.}}{0}}),x^2:=\frac{1}{2}(x^{0\stackrel{\text{.}}{0}}x^{1\stackrel{\text{.}}{1}}),x^3:=\frac{1}{2}(x^{1\stackrel{\text{.}}{0}}x^{0\stackrel{\text{.}}{1}}),x^4:=\frac{1}{2}(x^{0\stackrel{\text{.}}{0}}+x^{1\stackrel{\text{.}}{1}}).$$
(A.24)
So, the space of real sections of the bundle (A.15) is the Kleinian space $`^{2,2}=(^4,\eta )`$.
Note that real sections $`\sigma _x\mathrm{\Gamma }_{}(𝒫)`$ of the bundle (A.15) define real projective lines $`S_x^1𝒯𝒫`$,
$$S_x^1=\sigma _x(\zeta )=\{\begin{array}{ccc}x^{0\stackrel{\text{.}}{0}}+\zeta x^{1\stackrel{\text{.}}{0}},& x^{0\stackrel{\text{.}}{1}}+\zeta x^{1\stackrel{\text{.}}{1}},& \zeta ,\\ \zeta ^1x^{0\stackrel{\text{.}}{0}}+x^{1\stackrel{\text{.}}{0}},& \zeta ^1x^{0\stackrel{\text{.}}{1}}+x^{1\stackrel{\text{.}}{1}},& \zeta \{\mathrm{}\}\{0\},\end{array}$$
(A.25)
parametrized by $`x=\{x^{\alpha \stackrel{\text{.}}{\alpha }}\}^{2,2}.`$ Fibres $`^2`$ of the bundle $`𝒯S^1`$ are real null 2-planes (real $`\alpha `$-planes ) in the space $`^{2,2}`$.
The polynomials $`\eta _\pm ^{\stackrel{\text{.}}{\alpha }}`$ in (A.17) are annihilated by the following differential operators:
$$V_{\stackrel{\text{.}}{\alpha }}=\frac{}{x^{1\stackrel{\text{.}}{\alpha }}}\zeta \frac{}{x^{0\stackrel{\text{.}}{\alpha }}}.$$
(A.26)
Let us introduce the subspace $`𝒫_0=S^1\times ^4`$ in $`𝒫=𝒫_+𝒫_{}S^2\times ^4`$. Then the vector fields (A.26) together with $`/\overline{\zeta }`$ form a basis of (0,1) vector fields on $`𝒫_+𝒫_0`$, and the vector fields $`\zeta ^1V_{\stackrel{\text{.}}{\alpha }},/\overline{\stackrel{~}{\zeta }}`$ form a basis of (0,1) vector fields on $`𝒫_{}𝒫_0`$. The space $`𝒫_0`$ is fibred over $`𝒯`$ by real null 2-planes (real $`\beta `$-planes ), and their basis is formed by the vector fields (A.26) with $`\zeta P^1,x^{\alpha \stackrel{\text{.}}{\alpha }},\alpha =0,1,\stackrel{\text{.}}{\alpha }=0,1`$ (real vector fields). Such vector fields annihilate the real sections (A.21), (A.22) of the bundle (A.15).
Let us introduce the basis $`d\eta _+^{\stackrel{\text{.}}{\alpha }}=dx^{0\stackrel{\text{.}}{\alpha }}+\zeta dx^{1\stackrel{\text{.}}{\alpha }}`$ of 1-forms on fibres of the bundle (A.15) and consider real $`x=\{x^{\alpha \stackrel{\text{.}}{\alpha }}\}^{2,2}`$. Then the 2-forms
$$d\eta _+^{\stackrel{\text{.}}{0}}d\eta _+^{\stackrel{\text{.}}{1}}=(dx^{0\stackrel{\text{.}}{0}}+\zeta dx^{1\stackrel{\text{.}}{0}})(dx^{0\stackrel{\text{.}}{1}}+\zeta dx^{1\stackrel{\text{.}}{1}})=$$
$$=dx^{0\stackrel{\text{.}}{0}}dx^{0\stackrel{\text{.}}{1}}+\zeta (dx^{1\stackrel{\text{.}}{0}}dx^{0\stackrel{\text{.}}{1}}+dx^{0\stackrel{\text{.}}{0}}dx^{1\stackrel{\text{.}}{1}})+\zeta ^2dx^{1\stackrel{\text{.}}{0}}dx^{1\stackrel{\text{.}}{1}}$$
(A.27)
are complex null 2-forms on $`^{2,2}`$ labelled by $`\zeta P^1`$. If $`\text{Im}\zeta =0`$, then (A.27) are real null 2-forms on $`^{2,2}`$ parametrized by $`S^1\zeta =\mathrm{cot}\frac{\theta }{2}`$.
|
warning/0007/physics0007041.html
|
ar5iv
|
text
|
# The shock-acoustic waves generated by earthquakes
## 1 Introduction
A plethora of publications have been devoted to the study of the ionospheric response to disturbances arising from impulsive forcing on the Earth’s atmosphere. It was found that in many cases a high proportion of the energy of the initial atmospheric disturbance is concentrated in the acoustic shock wave. Large earthquakes also provide a natural source of impulsive forcing.
These investigations have also important practical implications since they furnish a means of substantiating reliable signal indications of technogenic effects (among them, unauthorized), which is necessary for the construction of an effective global radiophysical system for detection and localization of these effects. Essentially, existing global systems of such a purpose use different processing techniques for infrasound and seismic signals. However, in connection with the expansion of the geography and of the types of technogenic impact on the environment, very challenging problems heretofore have been those to improve the sensitivity of detection and the reliability of measured parameters of the sources of impacts, based also on using independent measurements of the entire spectrum of signals generated during such effects.
To solve the above problems requires reliable information about fundamental parameters of the ionospheric response to the shock wave, such as the amplitude and the form, the period, the phase and group velocity of the wavetrain, as well as angular characteristics of the wave vector. Note that for naming the ionospheric response of the shock wave, the literature uses terminology incorporating a different physical interpretation, among them the term ‘shock-acoustic wave’ (SAW) (Nagorsky, 1998). For convenience, in this paper we shall use this term despite the fact that it does not reflect essentially the physical nature of the phenomenon.
There is a wide scatter of published data on the main parameters of SAWs generated during industrial explosions and earthquakes. The oscillation period of the ionospheric response of SAWs varied from 30 to 300 sec., and the propagation velocity fluctuated from 700 to 1200 m/s (Nagorsky, 1998; Fitzgerald, 1997; Calais et al., 1998; Afraimovich et al., 1984; Blanc and Jacobson, 1989).
The lack of comprehensive, reliable data on SAW parameters is due primarily to the limitations of existing experimental methods and detection facilities. The main body of data was obtained by measuring the frequency Doppler shift at vertical and oblique-incidence ionospheric soundings in the HF range (Nagorsky, 1998; Afraimovich et al., 1984; Jacobson and Carlos, 1994). In some instances the sensitivity of this method is sufficient to detect the SAW reliably; however, difficulties emerge for localizing the region where the detected signal is generated. These problems are caused by the multiple-hop character of HF signal propagation. This gives no way of deriving reliable information about the phase and group velocities of the SAW propagation, estimating angular characteristics of the wave vector, and, further still, of localizing the SAW source.
Using the method of transionospheric sounding with VHF radio signals from geostationary satellites, a number of experimental data on SAW parameters were obtained in measurements of the Faraday rotation of the plane of signal polarization which is proportional to a total electron content (TEC) along the line connecting the satellite-borne transmitter with the receiver (Li et al., 1994).
A common drawback of the above-mentioned methods when determining the SAW phase velocity is the necessity of knowing the time of events since this velocity is inferred from the SAW delay with respect to the time of events assuming that the velocity is constant along the propagation path, which is quite contrary to fact.
For determining the above-mentioned rather complete set of SAW parameters, it is necessary to have appropriate spatial and temporal resolution which cannot be ensured by existing, very sparse, networks of ionosondes, oblique-incidence radio sounding paths, and incoherent scatter radars.
The advent of the Global Positioning System (GPS) plus the subsequent creation of extensive networks of GPS stations (at least 757 sites as of November 2000), with their data being now available via the Internet, opened up a new era in remote ionospheric sensing.
Currently some authors have embarked on an intense development of methods for detecting the ionospheric response of strong earthquakes (Calais and Minster, 1995), rocket launchings (Calais and Minster, 1996), and industrial surface explosions (Fitzgerald, 1997; Calais et al., 1998). In the cited references the SAW phase velocity was determined by the ‘crossing’ method by estimating the time delay of SAW arrival at subionospheric points corresponding to different GPS satellites observed at a given time. However, the accuracy of such a method is rather low because the altitude at which the subionospheric points are specified, is determined in a crude way.
The goal of this paper is to describe a method for determining parameters of the SAW generated by earthquakes (including the phase velocity, angular characteristics of the SAW wave vector, the direction towards the source, and the source location) using GPS-arrays whose elements can be chosen out of a large set of GPS stations from the global GPS network.
Section 2 presents a description of the experimental geometry, and general information about the earthquakes under consideration. The proposed method is briefly outlined in Section 3. Results of measurements of SAW parameters from different GPS arrays during earthquakes are presented in Section 4. Section 5 is devoted to the discussion of experimental results, including analytical simulation results.
## 2 The geometry and general characterization of experiments
Detection results on two earthquakes in Turkey (August 17, and November 12, 1999), in Southern Sumatera (June 4, 2000), and off the coast of Central America (January 13, 2001) are presented below. The information about the earthquakes, was acquired via the Internet ‘http://earthquake.usgs.gov’. General information about these earthquakes is presented in Table 1 (including the time of the main shock $`t_0`$ in the universal time UT, the position of the earthquake epicenter, depth, the magnitude, as well as the level of geomagnetic disturbance from the data on $`Dst`$-variations). It was found that the deviation of $`Dst`$ for the selected days was quite moderate, which enabled the SAWs to be identified.
Fig. 1 illustrates the experimental geometry during the earthquakes in Turkey – a), off the coast of Central America – b), and in Southern Sumatera – c).
In spite of the small number of GPS stations in the earthquake area, we were able to use a sufficient number of them for the implementation of the proposed method.
Table 2 presents the geographic coordinates of the GPS stations used as GPS array elements.
## 3 Methods of determining shock-acoustic wave characteristics using GPS-arrays
The standard GPS technology provides a means for wave disturbances detecion based on phase measurements of TEC at each of spaced two-frequency GPS receivers. A method of reconstructing TEC variations was detailed and validated in a series of publications (Calais and Minster, 1995, 1996; Fitzgerald, 1997). We reproduce here only the final formula for the total electron content (I)
$$I=\frac{1}{40.308}\frac{f_1^2f_2^2}{f_1^2f_2^2}[(L_1\lambda _1L_2\lambda _2)+const+nL]$$
(1)
where $`L_1\lambda _1`$ and $`L_2\lambda _2`$ are additional paths of the radio signal caused by the phase delay in the ionosphere, (m); $`L_1`$ and $`L_2`$ represent the number of phase rotations at the frequencies $`f_1`$ and $`f_2`$; $`\lambda _1`$ and $`\lambda _2`$ stand for the corresponding wavelengths, (m); $`const`$ is the unknown initial phase ambiguity, (m); and $`nL`$ are errors in determining the phase path, (m).
Phase measurements in the GPS can be made with a high degree of accuracy corresponding to the error of TEC determination of at least $`10^{14}`$ m<sup>-2</sup> when averaged on a 30-second interval, with some uncertainty of the initial value of TEC, however. This makes possible detection of ionization irregularities and wave processes in the ionosphere over a wide range of amplitudes (up to $`10^4`$ of the diurnal TEC variation) and periods (from 24 hours to 5 min). The unit of TEC, $`TECU`$, is equal to $`10^{16}`$ m<sup>-2</sup>, it is commonly accepted in the literature, and will be used in the following.
A convenient way of determining the ionospheric response delay of the shock wave involves the frequency Doppler shift $`F`$ from TEC series obtained by formula (1). Such an approach is also useful in comparing TEC response characteristics from the GPS data with those obtained by analyzing VHF signals from geostationary satellites, as well as in detecting the shock wave in the HF range. To an approximation sufficient for the purpose of our investigation, a corresponding relationship was obtained by K. Davies (1969)
$$F=13.510^8I_t^{}/f$$
(2)
where $`I_t^{}`$ stands for the time derivative of TEC. Relevant results derived from analyzing the $`F(t)`$-variations calculated for the ‘reduced’ frequency of 136 MHz are discussed in Section 4.
The correspondence of space-time phase characteristics, obtained through transionospheric soundings, with local characteristics of disturbances in the ionosphere was considered in detail in a wide variety of publications (Afraimovich et al., 1992; Mercier and Jacobson, 1997) and is not analyzed at length in this study. The most important conclusion of the cited references is the fact that, as for the extensively exploited model of a ‘plane phase screen’ disturbances $`\mathrm{\Delta }I(x,y,t)`$ of TEC faithfully copy the horizontal part of the corresponding disturbance of local electron concentration $`\mathrm{\Delta }N(x,y,z,t)`$, independently of the angular position of the source, and can be used in experiments for measuring the TEC disturbances.
However, the TEC response amplitude experiences a strong azimuthal dependence caused by the integral character of a transionospheric sounding. As a first approximation, the transionospheric sounding method is responsive only to Traveling Ionospheric Disturbances (TIDs) with the wave vector $`K_t`$ perpendicular to the direction $`r`$ wich is along the Line-of-Sight (LOS) from the receiver to the satellite. A corresponding condition for elevation $`\theta `$ and azimuth $`\alpha `$ of an arbitrary wave vector $`K_t`$ normal to the direction $`r`$, has the form
$$\theta =\mathrm{arctan}(\mathrm{cos}(\alpha _s\alpha )/\mathrm{tan}\theta _s)$$
(3)
where $`\alpha _s`$ is the azimuthal angle measured east to north, and $`\theta _s`$ is the angle of elevation of the satellite at the receiver.
We used formula (3) to determine the elevation $`\theta `$ of $`K_t`$ from the known mean value of azimuth $`\alpha `$ by Afraimovich et al. (1998) – see Sections 3.2 and 4.
### 3.1 Detection and determination of the horizontal phase velocity and the direction of the SAW phase front along the ground by GPS-arrays
In the simplest form, space-time variations of the TEC $`\mathrm{\Delta }I(t,x,y)`$ in the ionosphere, at each given time $`t`$ can be represented in terms of the phase interference pattern that moves without a change in its shape (the solitary, plane travelling wave)
$$\mathrm{\Delta }I(t,x,y)=\delta \mathrm{sin}(\mathrm{\Omega }tK_xxK_yy+\phi _0)$$
(4)
where $`\delta `$, $`K_x`$, $`K_y`$, $`\mathrm{\Omega }`$, are the amplitude, the x- and y-projections of the wave vector $`𝑲`$, and the angular frequency of the disturbance, respectively; $`T=2\pi /\mathrm{\Omega }`$, $`\mathrm{\Lambda }=2\pi /|K|`$ is its period and wavelength; and $`\phi _0`$ is the initial phase of the disturbance. The vector $`K`$ is a horizontal projection of the full vector $`K_t`$.
At this point, it is assumed that in the case of small spatial and temporal increments (the distances between GPS-array sites are less than the typical spatial scale $`\mathrm{\Lambda }`$ of TEC variation, and the time interval between counts is less than the corresponding time scale), the influence of second derivatives can be neglected. The following choices of GPS-arrays all meet these requirements.
We now summarize briefly the sequence of data processing procedures. Out of a large number of GPS stations, three sites (A, B, C) are selected, within distances not exceeding about one-half of the expected wavelength $`\mathrm{\Lambda }`$ of the perturbation. Site B is taken to be the center of a topocentric reference frame whose axis $`x`$ is directed east, and the axis $`y`$ is directed north. The receivers in this frame of reference have the coordinates ($`x_A`$, $`y_A`$), (0,0), ($`x_C`$, $`y_C`$). Such a configuration of the GPS receivers represents the GPS-array with a minimum number of the required elements. In regions with a dense network of GPS sites, we can obtain a large variety of GPS-arrays of a different configuration enabling the acquired data to be checked for reliability; in this paper we have exploited this possibility.
The input data include series of slant TEC values $`I_A(t)`$, $`I_B(t)`$, $`I_(t)`$, as well as corresponding series of elevation values $`\theta _s(t)`$ and the azimuth $`\alpha _s(t)`$ of the LOS. For determining SAW characteristics, continuous series of measurements of $`I_A(t)`$, $`I_B(t)`$, $`I_C(t)`$ are selected with a length of at least a one-hour interval which includes the time of a earthquake.
To eliminate spatio-temporal variations of the regular ionosphere, as well as trends introduced by orbital motion of the satellite, a procedure is used to remove the trend involving a preliminary smoothing of the initial series with the selected time window. This procedure is better suited to the detection of a single pulse signal ($`N`$-wave) than the frequently used band-pass filter (Li et al., 1994; Calais and Minster, 1995, 1996; Fitzgerald, 1997; Calais et al., 1998). A limitation of the band-pass filter is the oscillatory character of the response which prevets from reconstructing the form of the $`N`$-wave.
Elevation $`\theta _s(t)`$ and azimuth $`\alpha _s(t)`$ values of the LOS are used to determine the location of the subionospheric point, as well as to calculate the elevation $`\theta `$ of the wave vector $`K_t`$ of the disturbance from the known azimuth $`\alpha `$ (see formula (3)).
The most reliable results from the determination of SAW parameters correspond to high values of elevations $`\theta _s(t)`$ of the LOS because sphericity effects become reasonably small. In addition, there is no need to convert the slant TEC $`\mathrm{\Delta }I(t)`$ to a ‘vertical’ value. In this paper, all results were obtained for elevations $`\theta _s(t)`$ larger than 30.
Since the distance between GPS-array elements (from several tens of kilometers to a few hundred kilometers) is much smaller than that to the GPS satellite (over 20000 km), the array geometry at the height of the ionosphere is identical to that on the ground.
Fig. 2a shows typical time dependencies of an slant TEC $`I(t)`$ at the GPS-array BSHM station near the area of the earthquake of August 17, 1999. (heavy curve), one day before and after the earthquake (thin lines). For the same days; panel b shows TEC variations $`\mathrm{\Delta }I(t)`$ after removal of a linear trend and smoothing by averaging over a sliding window of 5-min. Variations in frequency Doppler shift $`F(t)`$, ‘reduced’ to the sounding signal frequency of 136 MHz, for three sites of the array (KATZ BSHM GILB) on August 17, 1999, are presented in panel c.
Fig. 2 shows that fast $`N`$-shaped oscillations, with a typical period of about 390 s, are distinguished among slow TEC variations. The oscillation amplitude (up to 0.12 $`TECU`$) is far in excess of the background TEC fluctuation intensity as seen on the days before and after the earthquake. Variations in frequency Doppler shift $`F(t)`$ for spatially separated sites (KATZ BSHM GILB) are well correlated but are shifted relative to each other by an amount well below the period, which permits the SAW propagation velocity to be unambiguously determined. The 30 s sampling rate of the GPS data is not quite sufficient for determining small shifts of such signals with an adequate accuracy for different sites of the array. Therefore, we used a parabolic approximation of the $`F(t)`$-oscillations in the neighborhood of minimum $`F(t)`$, which is quite acceptable when the signal/noise ratio is high.
Taking into account the good signal/noise ratio (better than 1), and knowing the coordinates of the array sites A, B and C, we determine the horizontal projection of the phase velocity $`V_h`$ from time shifts $`t_p`$ of a maximum deviation of the frequency Doppler shift $`F(t)`$. Preliminarily measured shifts are subjected to a linear transformation with the purpose of calculating shifts for sites spaced relative to the central site northward $`N`$ and eastward $`E`$. This is followed by a calculation of the $`E`$\- and $`N`$-components of $`V_x`$ and $`V_y`$, as well as the direction $`\alpha `$ in the range of angles 0–360 and the modulus $`V_h`$ of the horizontal component of the SAW phase velocity
$$\begin{array}{cc}\hfill \alpha & =\mathrm{arctan}(V_y/V_x)\hfill \\ \hfill V_h& =|V_xV_y|(V_x^2+V_y^2)^{1/2}\hfill \end{array}$$
(5)
where $`V_y`$, $`V_x`$ are the velocities with which the phase front crosses the axes $`x`$ and $`y`$. The orientation $`\alpha `$ of the wave vector $`K`$, which is coincident with the propagation azimuth of the SAW phase front, is calculated unambiguously in the range 0–360 subject to the condition that $`arctan(V_y/V_x)`$ is calculated having regard to the sign of the numerator and denominator.
The above method for determining the SAW phase velocity neglects the correction for orbital motion of the satellite because the estimates of $`V_h`$ obtained below exceed an order of magnitude as a minimum the velocity of the subionospheric point at the height of the ionosphere for elevations $`\theta _s>30^{}`$ (Afraimovich et al., 1998).
Obviously the method presented above can be used if the distance between GPS stations is much shorter than the TEC disturbance wavelength $`\mathrm{\Lambda }`$ and the distance from the earthquake epicenter to the array. This corresponds to the detection condition in the far-field zone.
From the delay $`\mathrm{\Delta }t=t_pt_0`$ and the known path length between the earthquake focus and the subionospheric point we calculated also the SAW mean velocity $`V_a`$ in order to compare our estimates of the SAW phase velocity with the usually used method of measuring this quantity.
### 3.2 Determination of the wave vector elevation $`\theta `$ and the velocity modulus $`V_t`$ of the shock wave
Afraimovich et al. (1992) showed that for the Gaussian ionization distribution the TEC disturbance amplitude ($`M`$) is determined by the aspect angle $`\gamma `$ between the vectors $`K_t`$ and $`r`$, as well as by the ratio of the wavelength of the disturbance $`\mathrm{\Lambda }`$ to the half-thickness of the ionization maximum $`h_d`$
$$M\mathrm{exp}\left(\frac{\pi ^2h_d^2\mathrm{cos}^2\gamma }{\mathrm{\Lambda }^2\mathrm{cos}^2\theta _s}\right).$$
(6)
In the case under consideration (see below), for a phase velocity on the order of 1 km/s and for a period of about 200 s, the wavelength $`\mathrm{\Lambda }`$ is comparable with the half-thickness of the ionization maximum $`h_d`$. When the elevations $`\theta _s`$ are 30, 45, 60 , the ‘beam-width’ $`M(\gamma )`$ at 0.5 level is, respectively, 25, 22 and 15. If $`h_d`$ is twice as large as the wavelength, then the beam tapers to 14, 10 and 8, respectively.
The beam-width is sufficiently small that the aspect condition (3) restricts the number of beam trajectories to the satellite, for which it is possible to detect reliably the SAW response in the presence of noise (near the angles $`\gamma =90^{}`$ ). On the other hand, formula (3) can be used to determine the elevation $`\theta `$ of the wave vector $`K_t`$ of the shock wave at the known value of the azimuth $`\alpha `$ (Afraimovich et al., 1998). Hence the phase velocity modulus $`V_t`$ can be defined as
$$V_t=V_hcos(\theta )$$
(7)
The above values of the width $`M(\gamma )`$ determine the error of calculation of the elevations $`\theta `$ (of order 20 to the above conditions).
### 3.3 Determining the position of the SAW source without regard for refraction corrections
The ionospheric region that is responsible for the main contribution to TEC variations lies in the neighborhood of the maximum of the ionospheric $`F`$-region, which does determine the height $`h_i`$ of the penetration point. When selecting $`h_i`$, it should be taken into consideration that the decrease in electron density with height above the main maximum of the $`F_2`$-layer proceeds much slower than below this maximum. Since the density distribution with height is essentially a ‘weight function’ of the TEC response to a wave disturbance (Afraimovich et al., 1992), it is appropriate to use, as $`h_i`$, the value exceeding the true height of the layer $`h_mF2`$ maximum by about 100 km. $`h_mF2`$ varies between 250 and 350 km depending on the time of day and on some geophysical factors which, when necessary, can be taken into account if additional experimental data and current ionospheric models are available. In all calculations that follow, $`h_i=400`$ km is used.
To a first approximation, it can be assumed that the imaginary detector, which records the ionospheric SAW response in TEC variations is located at this altitude. The horizontal extent of the detection region, which can be inferred from the propagation velocity of the subionospheric point as a consequence of the orbital motion of the GPS satellite (on the order of 70 –150 m/s; see Pi at al., 1997), and from the SAW period (on the order of 200 s — see Section 4), does not exceed 20–40 km, which is far smaller than it ‘vertical size’ (on the order of the half-thickness of the ionization maximum $`h_d`$).
From the GPS data we can determine the coordinates $`X_s`$ and $`Y_s`$ of the subionospheric point in the horizontal plane $`X0Y`$ of a topocentric frame of reference centered on the point B(0,0) at the time of a maximum TEC deviation caused by the arrival of the SAW at this point. Since we know the angular coordinates $`\theta `$ and $`\alpha `$ of the wave vector $`K_t`$, it is possible to determine the location of the point at which this vector intersects the horizontal plane $`X^{}0Y^{}`$ at the height $`h_w`$ of the assumed source. Assuming a rectilinear propagation of the SAW from the source to the subionospheric point and neglecting the Earth sphericity, the coordinates $`X_w`$ and $`Y_w`$ of the source in a topocentric frame of reference can be defined as
$$X_w=X_p\left(h_ih_w\right)\frac{\mathrm{cos}\theta \mathrm{sin}\alpha }{\mathrm{sin}\theta }$$
(8)
$$Y_w=Y_p\left(h_ih_w\right)\frac{\mathrm{cos}\theta \mathrm{cos}\alpha }{\mathrm{sin}\theta }$$
(9)
The coordinates $`X_w`$ and $`Y_w`$, thus obtained, can readily be recalculated to the values of the latitude and longitude ($`\varphi _w`$ and $`\lambda _w`$) of the source.
For SAW generated during earthquakes, industrial explosions and underground tests of nuclear devices, $`h_w`$ is taken to be equal to 0 (the source lying at the ground level).
## 4 Results of measurements
Hence, using the transformations described in Section 3, we obtain the parameters set determined from TEC variations and characterizing the SAW (see Table 3).
Let us consider the results derived from analyzing the ionospheric effect of SAW during earthquake August 17, 1999 obtained at the array (KATZ, BSHM, GILB) for PRN 6 (at the left of Fig. 2, and line 2 in Table 3).
In this case the delay of the SAW response with respect to the time of the earthquake is 20 min (DAY 229). The SAW has the form of an $`N`$-wave with a period $`T`$ of about 390 s and an amplitude $`A_I=0.12TECU`$, which is an order of magnitude larger than TEC fluctuations for background days (DAY 228, DAY 230). It should be noted, however, that this time interval was characterized by a very low level of geomagnetic activity (-14 nT).
The amplitude of a maximum frequency Doppler shift $`A_F`$ at the ‘reduced’ frequency of 136 MHz was found to be 0.04 Hz. In view of the fact that the shift $`F`$ is inversely proportional to the sounding frequency squared (Davies, 1969), this corresponds to a Doppler shift at the working frequency of 13.6 MHz and the equivalent oblique-incidence sounding path of about $`A_F=4`$ Hz.
Solid curves in Fig. 1a represent the trajectories of the subionospheric points for each GPS satellite at the height $`h_i`$=400 km during the time interval 0.0–0.8 UT for August 17, 1999, and 17.0–17.4 UT for November 12, 1999. Dark diamonds along the trajectories correspond to the coordinates of the subionospheric points at time $`t_p`$ of a maximum deviation of the TEC (Fig. 2b and Fig. 2e). Crosses show the positions of the earthquake epicenters. Asterisks mark the source location at 0 km altitude inferred from the data from the GPS arrays. Numbers at the asterisks correspond to the respective day numbers. Straight dashed lines that connect the expected source and the subionospheric point represent the horizontal projection of the wave vector $`K_t`$.
The azimuth $`\alpha `$ and elevation $`\theta `$ of the wave vector $`K_t`$ whose horizontal projection is shown in Fig. 1a by a dashed line and is marked by $`K_1`$, are 154 and 26, respectively. The horizontal component and the modulus of the phase velocity were found to be $`V_h=1307`$ m/s and $`V_t=1174`$ m/s. The source coordinates at 0 km altitude were determined as $`\varphi _w=39.1^{}`$ and $`\lambda _w=25.9^{}`$. The alculated (by neglecting refraction corrections) location of the source roughly was close the earthquake epicenter.
The ‘mean’ velocity of about $`V_a=870`$ m/s, determined in a usual manner from the response delay with respect to the start, was smaller then the phase velocity $`V_t`$. Conceivably this is associated with an added delay of the response as a consequence of the refraction distortions of the SAW path along the LOS which were neglected in this study.
Similar results for the array (KABR, TELA, GILB) and PRN 30 were also obtained for the earthquake of November 12, 1999. They correspond to the projection of the vector $`K_2`$ in Fig. 1a, the time dependencies in Fig. 2 at the right, and line 7 in Table 3. The only point deserving mention here is that the SAW amplitude was somewhat smaller than that for the earthquake of August 17, 1999. With an increased level of geomagnetic activity ( -44 nT), this led to a smaller (compared with August 17, 1999) signal/noise ratio; yet this did not preclude reliable estimates of the SAW parameters.
A comparison of the data for both earthquakes showed a reasonably close agreement of SAW parameters irrespective of the level of geomagnetic disturbance, the season, and the local time.
To convince ourselves that the determination of the main parameters of the SAW form and dynamics is reliable for the earthquakes analyzed here, in the area of the earthquake we selected different combinations of three sites out of the sets of GPS stations available to us, and these data were processed with the same processing parameters. Relevant results (including the average results for the sets $`\mathrm{\Sigma }`$), presented in Table 3 and in Fig. 1a (SAW source position), show that the values of SAW parameters are similar, which indicates a good stability of the data, irrespective of the GPS-array configuration.
The aspect condition (3), corresponding to a maximum amplitude of the TEC response to the transmission of SAWs, was satisfied quite well for this geometry, simultaneously for all stations. This is confirmed by a high degree of correlation of the SAW responses at the array elements (Fig. 2c and Fig. 2f), which made it possible to obtain different sets of triangles out of the six GPS stations available to us.
The relative position of the GPS stations was highly convenient for determining the SAW parameters during the earthquakes in Turkey, and met the implementation conditions for the method described in Section 3.1. Thus, the distance between stations (100-150 km at most) did not exceed the SAW wavelength of about 200-300 km, and was far less than the distance from the epicenter to the array (1000 km).
Let us consider the results derived from analyzing the ionospheric effect of SAW during earthquake January 13, 2001, obtained at the array (TEGU, MANA, ESTI) for PRN13 (at the left of Fig. 3 and line 11 in Table 3).
In this case the delay of the SAW response with respect to the time of the earthquake is 15 min (DAY 013). The SAW has the form of oscillations with a period $`T`$ of about 270 s and an amplitude $`A_I=0.2TECU`$, which is an order of magnitude larger than TEC fluctuations for background days (DAY 012, DAY 014). It should be noted, that this time interval was characterized by a very low level of geomagnetic activity (4 nT).
Similar results for the array (SAMP, NTUS, BAKO) and PRN 03 were also obtained for the earthquake of June 4, 2000. (at the right of Fig. 3 and line 10 in Table 3).
Note that in this case the response amplitude exceeded twice, as a minimum, that for the other events under consideration. It is not improbable that this is due to the maximum magnitude of the earthquake in Southern Sumatera (see Table 1).
Unfortunately, because of the inadequately well-developed network of stations, for the earthquakes in Southern Sumatera (June 4, 2000) and off the coast of Central America (January 13, 2001) for each of these events it was impossible to select arrays meeting the applicability conditions of the method for determining the SAW wave vector parameters described in Section 3.1.
It is evident from the geometry in Fig.1b and Fig. 1c that the earthquake epicenters lay inside the GPS arrays, which does not meet the far-field zone condition. It is possible that this is also responsible for the form of the response (presence of strong oscillations) which differs from the N-form of the response for the earthquakes in Turkey. On this basis, for these earthquakes we can point out only the very fact of reliable detection of the response, and determine its amplitude $`A_I`$, typical period $`T`$, delay $`\mathrm{\Delta }t=t_pt_0`$, and velocity $`V_a`$ (see Table 3). The values of these quantities were close to the data obtained for the earthquakes in Turkey.
## 5 Discussion
The data of Table 3 are quite sufficient to estimate the position of the disturbances under discussion in the diagnostic diagram of the atmospheric waves. Specifically, the values of the characteristic periods of bipolar signals, presented in Fig. 2b and Fig. 2e, are $`T_1`$ 300 s, and $`T_2`$ 200 s. The wave vectors are at the angles $`\beta _170^{}`$ and $`\beta _250^{}`$ with vertical. At the height $`z`$ 400 km, in turn, the periods, corresponding to local frequencies of the acoustic cut-off $`\omega _a`$ and the Brunt-Väisälä $`\omega _b`$, are: $`T_a=2\pi /\omega _a`$ 950 s, $`T_b=2\pi /\omega _b`$ 1050 s.
In an isothermal atmosphere, only harmonics with periods larger than $`T_b/sin\beta `$ can be assigned to the branch of internal gravity waves. This value exceeds significantly the values of $`T_{1,2}`$. Furthermore, $`T_{1,2}`$ is considerably smaller than $`T_a`$. Hence we can contend that under conditions of the real (nonisothermal) atmosphere the disturbances under discussion almost entirely pertain to the branch of acoustic waves.
Since in (short period) acoustic waves the variation $`\mathrm{\Delta }N`$ in neutral density $`N`$ satisfies the relation $`\mathrm{\Delta }N/Nu/C`$ ($`u`$ is the gas velocity in the wave, and $`C`$ is the local velocity of sound), it is possible to make a lower estimate of the intensity $`u/C`$ of the waves at the expected altitudes $`z_{eff}350`$–400 km: $`\mathrm{\Delta }N/N\mathrm{\Delta }N_e/N_e\mathrm{\Delta }I/I0.020.04`$. In this estimation, the influence of the magnetic field is omitted, and the index $`e`$ refers to electrons. The actual intensity of the acoustic wave must be higher than the estimated value presented above.
In the case of an earthquake, the movements of the terrestrial surface are plausible sources of acoustic waves. A generally known source of the first type is the Rayleigh surface wave propagating from the epicentral zone. The ground motion in the epicentral area is the source of the second type. Let us discuss these possibilities.
### 5.1 The Rayleigh wave
The phase propagation velocity of the Rayleigh wave $`V_R3.3`$ km/s. Since $`V_RC_0`$, where $`C_00.34`$ km/s is the sound velocity at the ground, only acoustic waves can be emitted (Golitsyn and Klyatskin, 1967). At a sufficient distance from the epicenter where the curvature of the Rayleigh wave front can be neglected, and with the proviso that $`\omega \omega _a`$, acoustic waves are emitted at the angle $`\beta _Rarcsin(C_0/V_R)6^{}`$ to the vertical.
Rayleigh waves propagate generally in the form of a train consisting of several oscillations whose typical period only rarely exceeds several tens of seconds. Acoustic waves are emitted upward in the form of the same train. Because of a strong absorption of the periodic wave, the only thing that is left over in the case of the acoustic train at heights $`z`$ 350 km is the leading phase of compression. It seems likely that only in the case of strong earthquakes (Alaskian earthquake of 1964), even at large distances from the epicenter, the disturbance (the leading portion of the acoustic train) that remains from the acoustic train, can have at these altitudes a duration of about 100 s, and quite an appreciable ($`u/C>`$ 0.1) intensity (Orlov and Uralov, 1987) the nonlinear acoustics approximation.
Unfortunately, the parameters of Rayleigh waves in the neighborhood of the subionospheric points appearing in Table 3 are unknown to us. However, a most pronounced bipolar character of the main signal (Fig. 2b, Fig. 2e, Fig. 3b and Fig. 3e), its intensity, and its long duration cast some doubt upon the fact that it is the Rayleigh wave which is responsible for its origin. Such a conclusion is consistent with the observed propagation direction of the bipolar pulse, which makes a large angle with the vertical: $`\beta _{1,2}70`$$`50^{}\beta _R^\mathrm{"}15`$$`18^{}`$.
Here it is taken into consideration that the acoustic ray originating from a point on the ground at the angle of $`\beta _R6^{}`$ now makes at the heights $`z=z_{eff}`$ an angle $`\beta _R^\mathrm{"}>\beta _R`$ because of the refraction effect in a standard atmospheric model: $`C_0/sin\beta _R=C/sin\beta _R^\mathrm{"}=const=V_R`$, $`C(z=z_{eff})0.9`$–1 km/s.
The presence of strong winds at ionospheric heights can alter the value of $`\beta _R`$. However, irrespective of the atmospheric model, the phase velocity $`V_h`$ (Table 3) of the horizontal trace of the acoustic disturbance generated by the Rayleigh wave must coincide with $`V_R`$, and this is also not observed: $`V_hV_R`$.
### 5.2 The epicentral emitter
The detection of ionospheric disturbances, which are presumably generated by a vertical displacement of the terrestrial surface directly in the epicentral zone of an earthquake, using the GPS probing method, is reported in (Calais and Minster, 1995).
Results of the present study lend support to the above conjecture. However, the specific formation mechanism for the disturbance itself is still unclear. An approach to solving this problem is contained in earlier work, and involves substituting the epicentral emitter for a surface velocity point source or an explosion. In particular, the substitution of the earthquake zone for a point source turns out to be fruitful when describing long-period internal gravity waves at a very long (thousands of kilometers) distance from the epicenter (Row, 1967). The visual resemblance of ionospheric disturbances at short (hundreds of kilometers) distances from the earthquake epicenter to disturbances from surface explosions is discussed in (Calais et al., 1998).
It should be noted that ionospheric disturbances generated by industrial surface and underground nuclear explosions are also visually similar. However, the generation mechanisms for disturbances are fundamentally different in this case (Rudenko and Uralov, 1995). The radiation source in underground nuclear tests is, as in the case of earthquakes, the terrestrial surface disturbed by the explosion. The intensity and spectral composition of the generated acoustic signal reveal a strong (unlike the surface explosion) dependence on the zenith angle, and are wholly determined by the form, size and characteristics of the movement of the terrestrial surface in the epicentral zone of the underground explosion.
In this Section, we shall propose a model which, we hope, will help to understand the generation mechanism for acoustic disturbances — the subject of this paper. Because of the complexity of the problem formulated, and for lack of sufficient data on characteristics of the movement of the terrestrial surface in epicentral zones of earthquakes, the idealized model under discussion has an illustrative character. The computational scheme proposed below represents a simplified variant of the scheme used in Rudenko and Uralov, (1995) to calculate ionospheric disturbances generated by an underground confined nuclear explosion.
### 5.3 The problem of radiation of the acoustic signal
For the sake of simplicity, we consider a problem having an axial symmetry about the vertical axis $`z`$ passing through the earthquake epicenter $`r=z=0`$. The epicentral emitter is a set of plane annular velocity sources with the specified law of motion along the vertical $`U(r,t)`$. Since our interest is with the estimation of the characteristics of an acoustic disturbance at a sufficient distance from the emitter, we take advantage of the far-field approximation of a linear problem of radiation. In the approximation of linear acoustics $`\omega \omega _a`$ and with no absorption present, the gas velocity profile in the wave can be estimated by the expression:
$$u(\tau ,\beta )=\frac{A}{\pi RC_0}_0^L_{\mathrm{}}^+\mathrm{}\frac{a(t^{^{}},r)rdrdt^{^{}}}{\sqrt{y^2(\tau t^{^{}})^2}|_{|\tau t^{^{}}|y}}$$
(10)
where $`\tau =t_0^l𝑑l/C`$, $`y=\frac{rsin\beta }{C_0}`$. Here $`\beta `$ is the zenith angle of departure of the acoustic ray from the point $`r=z=0`$. $`l`$ is the group path length (the distance along the ray). $`a(t^{^{}},r)=\frac{dU(r,t)}{dt}`$ is the vertical acceleration of the terrestrial surface. $`A=A(z)=\sqrt{\frac{C_0\rho _0}{C\rho }}`$ is the acoustic factor. $`\rho _0,\rho `$ stand for the air density at the ground and at the height $`z`$, respectively, and L is a typical radius of the epicentral emitter.
In an isothermal atmosphere where the ray trajectories are straight lines, the quantity $`R=(\sqrt{r^2+z^2}=l)`$ is the radius of a divergent (in the far-field approximation) spherical wave. In this case the cross-section of the selected ray tube $`SR^2`$. In the real atmosphere the value of $`S`$ is determined from geometrical optics equations, and using the expression (10) requires a further complication of the computational scheme. Nevertheless, to make estimates we shall use only the relation (10), and, in doing so, the value of $`R`$ will be corrected.
The case $`\beta =0`$ (and also $`l=R=z`$) is a special one:
$$u(\tau ,\beta =0)=\frac{A}{RC_0}_0^La(\tau ,r)r𝑑r$$
(11)
Let us discuss the situation where all ring-type emitters ‘operate’ synchronously: $`U(r,t)=U(t)`$. In this case the epicentral zone is emitting as a round piston 2$`L`$ in diameter. Assume also that with a shock of the earthquake, the vertical displacement $`\xi `$, the velocity $`U=d\xi /dt`$ and the acceleration $`a=dU/dt`$ are time dependent, as shown in Fig. 4a. For a rectangular velocity impulse, i.e. in the limit $`\mathrm{\Delta }t\delta t0`$ (in this case $`a\delta t\pm U_0\delta (t^{^{}})`$, where $`\delta (t^{^{}})`$ is a delta-function), from (10) we can obtain
$$u(\tau ,\beta )=\frac{U_0C_0A}{\pi Rsin^2\beta }\{\sqrt{y_L^2\tau ^2}|_{|\tau |y_L}\sqrt{y_L^2(\tau \mathrm{\Delta }t)^2}|_{|\tau \mathrm{\Delta }t|y_L}\}$$
(12)
where $`y_L=\frac{Lsin\beta }{C_0}`$; $`U_0,\mathrm{\Delta }t`$ are the amplitude and duration of the rectangular velocity impulse. In this case the vertical displacement of the terrestrial surface after the earthquake shock is $`\xi _0=U_0\mathrm{\Delta }t`$. In the strict sense, the expression (12) holds true if the condition $`y_L\delta t`$ is satisfied, i.e. with zenith angles $`\beta \beta ^{}arcsin\frac{C_0\delta t}{L}`$. Here $`\delta t`$ is the operation duration of the terrestrial shock at the beginning of the movement and at the stop of the piston. When $`\beta \beta ^{}1`$ , the expression (12) and the expression (11), in view of $`\delta t0`$, yield approximately identical signals. What actually happens is that the validity of the expression (12) will break down still earlier, and we are justified in using it only for zenith angles $`\beta >\beta ^{}=arctan(\frac{L}{z})>\beta ^{}`$.
The curves in Fig. 4b give an idea of the relative amplitude and form of acoustic signals $`u(\tau ,\beta )`$ in the far-field zone of radiation of the piston $`2L=60`$ km in diameter. The duration of the rectangular velocity impulse of the piston was chosen arbitrarily, $`\mathrm{\Delta }t=10`$ s. The signals correspond to the expression (12) under the assumption of a homogeneous, $`A=1`$, $`C=C_00.34`$ km/s, atmosphere. The zenith angles are $`\beta =10^{},20^{},40^{},60^{},90^{}`$. The spherical surface $`R=const`$ serves as a reference. The abscissa axis indicates the time $`\tau `$ in seconds, and the axis of ordinates indicates the gas velocity $`u`$. In this case the amplitude of the most intense signal ($`\beta =10^{}`$) is taken to be unity.
With increasing $`\beta `$, the amplitude of the signals $`u_{max}`$ decreases, and the duration $`T`$ increases. The leading and trailing edges of the bipolar pulses in Fig. 2 and Fig. 3 have an equal duration $`\mathrm{\Delta }t`$. The positive part of the bipolar pulse corresponds to the compression phase of the acoustic wave, and the negative part refers to the rarefaction phase. The area of the compression phase (in coordinates $`u,t`$) equals the largest displacement $`\chi _+`$ of a unit volume of the atmosphere in the direction of propagation (along the ray) of the wave. A total area of a bipolar pulse is zero: $`\chi _+=\chi _{}`$. For acoustic signals, described by the expression (12), the following useful relations hold true:
$$u_{max}=U_0\frac{2AL}{\pi Rsin\beta }\sqrt{\eta \eta ^2};T=2y_L+\mathrm{\Delta }t$$
(13)
$$\chi _+=\xi _0\frac{AL}{2\pi Rsin\beta }\{\sqrt{1\eta ^2}+\frac{1}{\eta }arcsin\eta \};\eta =\frac{\mathrm{\Delta }t}{2y_L}$$
(14)
### 5.4 The problem of acoustic signal propagation
The wave vectors $`K_t`$ of the bipolar pulses make with the vertical the angles $`\beta _1=70^{}`$ and $`\beta _2=50^{}`$. With the adopted values of $`z_{eff}=350`$–400 km, the distances of the corresponding subionospheric points from the earthquake epicenters in Fig. 1a are approximately $`r_1=800`$ km and $`r_2=600`$ km. The rays, constructed in the approximation of linear geometrical acoustics (LGA) and having at the heights $`z=z_{eff}`$ ($`C(z=z_{eff})0.9`$–1 km/s) the propagation angles $`\beta _1`$ and $`\beta _2`$, correspond to the zenith angle of departure $`\beta 19`$$`21^{}`$ and $`\beta 15`$$`17^{}`$ at the level $`z=0`$.
The fact that these values of $`\beta `$ satisfy the inequality $`\beta \beta ^{}25`$$`30^{}`$, is in reasonably good agreement with the familiar picture of rays from a ground-level point source — the rays with $`\beta \beta ^{}`$ are captured by the atmospheric waveguide $`zz^{}=120`$ km, and only the rays emitted upward inside the solid angle $`\mathrm{\Omega }^{}1`$ sterad can penetrate to the heights $`z>z^{}`$. In standard models of the atmosphere, however, to the values of the angles $`\beta _{1,2}(z=z_{eff})`$ there correspond the locations of subionospheric points lying several hundreds kilometers nearer to the epicenter compared with the experimental values of $`r_1800`$ km and $`r_2600`$ km. This inconsistency can be due to two reasons.
One reason is that the value of $`z_{eff}350`$–400 km, which we are using, is too low. This is supported by the detection of velocities $`V_t1.2`$ km/s of traveling disturbances (Table 3) which exceed markedly the sound velocity $`C0.9`$–1 km/s at the heights of $`350`$–400 km. However, it seems likely that such a discrepancy may be disregarded, in view of the errors of the measurement technique used (the probability of an additional heating of the upper atmosphere prior to the earthquake cannot be ruled out, however).
The increase of the actual value of $`z_{eff}`$ can also be associated with a strong dependence (see (13), (14) and Fig. 4b) of the power of the emitted signal on the zenith angle of departure of the ray from the earthquake epicenter. Verifying this factor requires a more detailed analysis, based on particular data on vertical movements of the terrestrial surface in the epicentral zone. Such data are unavailable to us. Below is a discussion only of the second reason for the above-mentioned inconsistency as a highly probable one.
The second reason may be associated with the violation of the validity conditions of the LGA approximation at a sufficient distance from the source of the acoustic disturbances under discussion. Indeed, the utilization of this approximation is justified until the parameters of the medium and of the wave itself change substantially on the size of the first Fresnel zone $`d_F\sqrt{\lambda l}`$ that determines the physical (transverse) size of the ray. Typical wavelengths of the bipolar pulses under discussion at the heights $`z=z_{eff}`$ are large: $`\lambda 300`$–200 km. Distance $`l`$ along the expected ray is of order 900–700 km. Then $`d_F520`$–370 km, which exceeds substantially the scales of variation of atmospheric parameters. The value of $`d_F`$ is actually somewhat smaller, because the typical scale of a disturbance decreases as it approaches the source. A violation of the LGA approximation at the above-mentioned distances from the epicentral source occurs also for model signals $`\beta 20^{}`$, shown in Fig. 4b.
The increasing violation of the applicability conditions of the LGA approximation with an increase of $`l`$ implies the transport of the wave energy not strictly along the calculated rays but along the lines with a smaller curvature. With a mere estimate of the dilution factor $`R`$ in the expressions (12), (13) and (14), in mind, these lines will be assumed to be straight when $`z>z^{}`$ and to originate from an imaginary source lying at the height $`z^{}120`$ km above the earthquake epicenter. Such a situation is also clearly manifested in the LGA approximation. At the heights $`z>z^{}`$, at $`t600`$ s, for example, the surface of the wave front from the ground-level impulsive source resembles the surface of a hemisphere centered on the point $`r0`$, $`zz^{}`$.
Ultimately the energy that arrives from below inside the solid angle $`\mathrm{\Omega }^{}1`$ is scattered into a solid angle $`\mathrm{\Omega }2\pi `$. From the condition of conservation of wave energy, it is possible to find $`R_{1,2}=\sqrt{\frac{\mathrm{\Omega }}{\mathrm{\Omega }^{}}}\sqrt{(\mathrm{\Delta }z)^2+r_{1,2}^2}`$, when $`\mathrm{\Delta }z=z_{eff}z^{}`$. When using the expressions (13), (14) in the subsequent discussion, we will take the quantity $`R_1`$ rather than $`R`$, and the value of the zenith angle of departure will be taken to be $`\beta =20^{}`$. These parameters approximately correspond to the generation and propagation conditions of the signal from the first earthquake. The typical size of the epicentral zone of this earthquake was about $`2L=60`$ km (according to the USGS data: www.neic.cr.usgs.gov). This same value of $`L`$ was used in calculating the signals shown in Fig. 4b.
As is evident even from Fig. 4b (thick line), the duration of the signal, having a zenith angle of departure $`\beta =20^{}`$, is about 70 s. When the signal propagates in the approximation of linear acoustics and with no absorption, its form and duration remain unchanging, and only its amplitude changes.
In actual conditions the combined effect of the nonlinear attenuation and linear absorption factors leads to a stretching of the bipolar pulse, and to a change of its form (we do not discuss the dispersion factor). In this case the effect of the nonlinearly factor occurs in such a manner that the integral value of $`\chi _+`$, (14), calculated as an approximation of linear acoustics, remains as such in the approximation of nonlinear acoustic as well. Moreover, taking into account the finite width $`\mathrm{\Delta }T_{sh}`$ of the shock front does not change this situation until $`\mathrm{\Delta }T_{sh}<T/4`$. The linear absorption factor, in turn, somewhat reduces the true value of $`\chi _+`$ because of the mutual diffusion of the compression and rarefaction phases. Nevertheless, for a hypothetical estimation of the earthquake parameters, we shall use the assumption about the conservation of the value of $`\chi _+`$.
As is intimated by Table 3, the mean value of the $`TEC`$ disturbance amplitude after the first earthquake is $`A_I0.14`$ $`TECU`$ at the equilibrium value of $`I5`$ $`TECU`$. Assuming that on the order of magnitude $`A_I/Iu/C`$ for a maximum gas velocity in the wave, we have an estimate of $`u^{exp}30`$ m/s. For the sinusoidal form of the bipolar pulse with a duration $`T^{exp}350`$ s, we find the experimental value of gas displacement along the direction of wave propagation: $`\chi _{+}^{}{}_{}{}^{exp}=u^{exp}T^{exp}/\pi 3`$ km. Using the relation $`\chi _{+}^{}{}_{}{}^{exp}\chi _+`$ it is possible to estimate the vertical displacement $`\xi _0`$ of the terrestrial surface in the epicentral zone of the earthquake. For this purpose, it seems reasonable to introduce the assumption about a short velocity impulse which models the main earthquake shock, $`\eta 1`$, although the numerical value of $`\chi _+`$ is virtually independent of the value of $`\eta `$ (14). In view of the above considerations, we then obtain:
$$\xi _0\chi _{+}^{}{}_{}{}^{exp}\pi R_1sin\beta /(AL).$$
(15)
The uncertainty in the determination of $`\xi _0`$ is caused both by the uncertainty of the true values of the quantities involved in this relation and by the limitations of the acoustic signal generation model itself. In this case, of the greatest importance is the dependence of $`\xi _0`$ on the value of the acoustic parameter $`A`$, containing the atmospheric density $`\rho `$ at the effective height $`z=z_{eff}`$ which we have introduced artificially. The employment of the MSISE90 atmospheric model (Hedin, 1991), calculated for the location and time of the first earthquake, gives the values of $`\xi _060,40`$ and 25 cm, with the values of $`z_{eff}=350,400`$ and 450 km, respectively. In all cases it was assumed that $`R_1850\sqrt{2\pi }`$ km, $`\beta =20^{}`$, $`L=30`$ km. The possibility of a vertical displacement of the terrestrial surface in the epicentral zone by several tens of centimeters seems real. In particular, (Calais and Minster, 1995) give the value of $`\xi _040`$ cm at the epicenter of the Mw=6.7 Northridge earthquake (California, 1994).
In view of the demonstration character of the above calculation, we have intentionally excluded from consideration the effects associated with the inclination of the magnetic field lines, and with the possible presence of strong winds at ionospheric heights. The presence of a magnetic field modifies the picture of transfer of movements from the neutral gas to the electron component of the ionosphere. Since the magnetic field is not entrained by the neutral gas, the field lines can be considered fixed. In this case the acceptable approximation would be the one, in which the electron component travels only along magnetic field lines with the velocity $`ucos\psi `$, where $`\psi `$ is the angle between the magnetic field vector and the velocity vector of the neutral gas. Therefore, the quantity $`ucos\psi `$ must be involved in lieu of the quantity $`u`$ in the expression $`A_I/Iu/C`$ that was used above.
Let us estimate the value of $`\psi `$ for Turkey earthquakes. In the examples under discussion (Table 3), the horizontal projection $`K`$ of the full wave vector $`K_t`$ is virtually collinear to the horizontal component of the magnetic field. The wave vectors $`K_t`$, in turn, make with the vertical the angles from $`\beta _170^{}`$ to $`\beta _250^{}`$.
Since the magnetic dip in the middle of Turkey is about $`60^{}`$ (the angle is measured from the horizontal plane), $`\psi 20^{}`$$`40^{}`$, and the value of $`cos\psi 0.94`$$`0.77`$ differs little from 1. It should be remembered, however, that taking into account this factor can be very important in the analysis of the complete picture of TEC disturbances above the earthquake or explosion source (see, for example, Calais et al., 1998).
The presence of the zonal and meridional winds at ionospheric heights leads to a displacement and deformation of the wave front, and hence give rise to a dependence of the acoustic wave intensity on the propagation direction. The decisive role in this case is played by the wind velocity gradient. This factor can be taken into account within the framework of the ray theory. However, a corresponding model calculation would be worthwhile in the analysis of experimental data obtained for a set of subionospheric points surrounding the acoustic wave source. In the present situation, however, where the number of subionospheric points used in the analysis is too small, and the uncertainty of the parameters of the acoustic emitter itself is too large, the solution of such an unwieldy problem would be an overrun of the accuracy which is pursued by the above computational scheme.
As follows from the expressions (13), maximum values of displacements and of the velocity of the neutral atmospheric species are attained directly above the earthquake epicenter. The signal duration is minimal, and does not seem to exceed a few tens of seconds at ionospheric heights. Since the wave vector of the disturbance is directed predominantly upward, the method of oblique-incidence ionospheric sounding in this case is a technique of choice for determining the waveform.
## 6 Conclusion
In this paper we have investigated the form and dynamics of shock-acoustic waves generated during earthquakes. We have developed a method of determining the SAW parameters using GPS arrays whose elements can be chosen out of a large set of the global network GPS stations. Unlike existing radio techniques, the proposed method estimates the SAW parameters without a priori information about the location and time of the earthquake. The implementation of the method is illustrated by an analysis of ionospheric effects of the earthquakes in Turkey (August 17, and November 12, 1999), in Southern Sumatera (June 4, 2000), and off the coast of Central America (January 13, 2001).
It was found that, in spite of the difference of the earthquake characteristics, the local time, the season, and the level of geomagnetic disturbance, for four earthquakes the time period of the ionospheric response is 180-390 s, and the amplitude exceeds by a factor of two as a minimum the standard deviation of background fluctuations in total electron content in this range of periods under quiet and moderate geomagnetic conditions.
As has been pointed out in the Introduction, some investigators report markedly differing values of the SAW propagation velocity — by as much as several thousands m/s, which is beyond the values of the sound velocity at the SAW propagation heights in the atmosphere. The method proposed in this paper opens up a possibility of determining the angular characteristics of the wave vector $`K_t`$ and, accordingly, of estimating $`V_t`$. According to our data (Table 3), the elevation of the SAW wave vector varied within 20–$`44^{}`$ , and the phase velocity of the SAW varied from 1100 to 1300 m/s. We determine the phase velocity of the equal TEC line at the height of the ionospheric F-region maximum which makes the main contribution to variations of the TEC between the receiver and the GPS satellite, and corresponds to the region of maximum sensitivity of the method. Since $`V_t`$ approaches the sound velocity at these heights (Li et al., 1994) this makes it possible to identify the sound origin of the TEC disturbance. The SAW source location, calculated without taking into account the refraction corrections, approximately corresponds to the earthquake epicenter.
## 7 Acknowledgements
We are indebted to G. M. Kuznetsova and A. V. Tashchilin for their calculation of the atmospheric parameters at the time of the earthquakes, as well as to E. A. Ponomarev, V. V. Evstafiev, P. M. Nagorsky, N. N. Klimov, and A. D. Kalikhman for their interest in this study, many pieces of useful advice, and active participation in discussions. Thanks are also due to V. G. Mikhalkosky for his assistance in preparing the English version of the manuscript. Finally, the authors wish to thank the referees for valuable suggestions which greatly improved the presentation of this paper. This work was done with support from both the Russian foundation for Basic Research (grant 99-05-64753) and RFBR grant of leading scientific schools of the Russian Federation No. 00-15-98509.
|
warning/0007/nucl-ex0007020.html
|
ar5iv
|
text
|
# Multifragmentation of a very heavy nuclear system (II): bulk properties and spinodal decomposition
## 1 Introduction
The decay of highly excited and possibly compressed nuclear systems through multifragmentation (complete disassembly into several lighter nuclei in a short time scale) is, at present time, a subject of great interest in nucleus-nucleus collisions. While this process has been observed for many years, its experimental knowledge was strongly improved only recently with the advent of powerful $`4\pi `$ devices; the major experimental problem comes from the difficulty to unambiguously select and well define the relevant system or subsystem. Part I of the present work fully illustrates this fact when selecting fused systems formed in collisions between very heavy nuclei .
Our goal in studying multifragmentation of very heavy systems, which can be considered as well defined pieces of nuclear matter, was to seek out bulk properties which could be compared to models in which bulk or volume instabilities are present.
Many theories have been developed to explain multifragmentation (see for example ref. for a general review of models). In particular one can arrive at the concept of multifragmentation by considering a phase transition of excited nuclear matter. This first phase transition is of liquid-gas type due to the specific form of the nucleon-nucleon interaction, which is characterized by attraction at long and intermediate range and repulsion at short range. It is possible that, during a collision between two nuclei, a wide zone of the nuclear matter phase diagram may be explored, including the liquid-gas phase coexistence region and even more precisely the spinodal region (domain of negative incompressibility and of mechanical instability of uniform matter) where multifragmentation can occur through the growth of density fluctuations .
Among the existing models of multifragmentation some are related to statistical approaches based either on multi-body phase space calculations or on fast sequential binary decays , whereas others try to describe the dynamic evolution of systems resulting from collisions between two nuclei. In the present case, semi-classical simulations based on the nuclear Boltzmann equation, which describe the time evolution of the one-body density, are appropriate during the early phase of the collisions but become inadequate when instabilities occur. A quantum-mechanical description including N-body correlations is not yet feasible, thus dynamic scenarios taking into account the dynamics of the phase transition are simulated, with different approximations, via molecular dynamics or stochastic mean field approaches .
It is this last approach which has been used for a comparison with our experimental observables. In simulations, spinodal decomposition of hot and dilute finite nuclear systems is mimicked through a powerful tool, the Brownian One-Body (BOB) dynamics . The BOB dynamics reintroduces approximately N-body correlation effects (i.e. fluctuations) by means of a brownian force in the mean field. The magnitude of the force is adjusted to produce the same r.m.s. fluctuation amplitude as the full Boltzmann-Langevin theory for the most unstable modes in nuclear matter prepared at the corresponding density and temperature .
The paper is organized as follows : the experimental results concerning selected multifragmenting fused events from 36 A.MeV $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$ collisions are presented in section 2; the dynamical properties of fragments and their connection with an eventual expansion energy are discussed in section 3; in section 4 a bulk effect is brought to light by comparison of experimental fragment charge and multiplicity distributions with results from a lighter system. Section 5 is devoted to a presentation of the dynamical simulations and their confrontation with the experimental data. A summary is given in section 6.
## 2 Characteristics of single-source events
Single-source events, for which very heavy fused systems are formed from the majority of nucleons of a heavy projectile and target, have been isolated among very well measured 36 A.MeV $`{}_{}{}^{155}Gd`$+$`{}_{}{}^{238}U`$ collisions (see Part I). Multifragmentation of these sources comprising more than 350 nucleons leads to fragments with charges up to 60. For such heavy reaction products recombination effects in solid state detectors (silicon detectors and CsI(Tl) scintillators) are important, and thus a great effort was recently made to obtain more accurate energy calibrations (globally within 6% over the whole detection angular range) and Z identification procedures (see Part I) .
Before presenting in detail the characteristics of the isolated fused systems, let us first recall very briefly what we learned in Part I (Section 4). Single-source events are selected by imposing the condition $`Z_{tot}`$$``$120 ($``$0.77$`Z_{sys}`$) for charge completeness, and through a global shape variable, namely the polar angle $`\theta _{flow}`$, calculated from the emission properties of fragments (the detection of at least three fragments with Z$``$5 was required). From the forward-backward symmetry of reaction product Z-distributions observed in these selected events, we can infer that the single-source has nearly the centre of mass velocity. In this section we will first present and discuss observables related to fragments. Then properties of light charged particles (LCP) associated to these events, which do not enter the chosen selection, will be presented and used to derive an estimate of the characteristics (size and excitation energy) of the source.
### 2.1 Fragment emission properties
Figure 1a) shows, in the reaction centre of mass frame, the angular distribution of emitted fragments. It is isotropic, except at large angles ($`\theta _{cm}>135^o`$ ) where detection and identification thresholds come into play. On figures 1b) and c) are displayed the evolution of fragment properties (average Z and average kinetic energy) with respect to their emission angle in the centre of mass; these average values are constant within $`\pm `$15%; The slight increase around 90<sup>o</sup> probably comes from the selection of events ($`\theta _{flow}70^o`$ \- see Part I)).
Figure 2 shows the evolution of the average kinetic energy of fragments versus their atomic number. The energy first increases with Z, up to 150 MeV at Z$``$30. Then it levels off, or slightly decreases. In the same figure are also reported previously published data which did not take into account recent improvements in the energy calibration of the backward CsI(Tl) scintillators of INDRA ; differences are essentially observed for intermediate charges, Z=15-30. This correlation between kinetic energy and Z of fragments is of particular interest and will be discussed later on in the paper: it can indeed permit the extraction of qualitative information about fragment emission (sequential or simultaneous), and quantitative information related to radial expansion energy (section 3). Consequently it puts strong constraints on theoretical models (section 5).
The differential multiplicity as a function of the atomic number of the detected charged reaction products is presented in figure 3a. It extends up to Z$``$55, which corresponds to a third of the total charge of the system. The contributions to the distribution of the three heaviest fragments of each partition are also displayed in the figure, showing that only the heaviest one populates the region Z=35-60. In figure 3b the distribution for fragments is plotted in a double logarithmic scale; a power law dependence in $`Z^\tau `$ of the differential multiplicity becomes evident in the range Z=5-20, with $`\tau `$ close to 1.0. Finite size effects break this law for higher charges, and the heaviest fragment is completely excluded from this dependence. Note that such a behaviour is predicted by simulating the explosion of hot drops of classical fluid in the spinodal region, but with a somewhat higher exponent .
The distribution of the total charge emitted in fragments (Z<sub>bound</sub>) is displayed in Figure 4; on average, more than half of the total charge of the system is bound in fragments. Finally Table 1 summarizes some average measured quantities related to fragments emitted by the single source (the different multiplicity distributions associated to single-source events have been presented in Fig. 10 of the accompanying paper).
### 2.2 Coincident LCP and estimates of average size and excitation energy of the source
LCP are emitted at different stages during the collisions. First “direct” pre-equilibrium emission occurs which partially keeps a memory of the entrance channel mass asymmetry . Then particles are emitted before and during the formation of fragments from a source which may be in thermal equilibrium. Finally in the late stages statistical evaporation from hot fragments takes place. Experimentally we measure the sum of all these contributions and, without having recourse to correlation functions , only deviations from isotropy in the centre of mass can be used to give complementary information about the source.
The angular distributions of LCP and their associated average kinetic energy as a function of the centre of mass angle are displayed in figures 5 and 6. Deuterons, tritons and $`{}_{}{}^{3}He`$ suffer from isotopic identification thresholds; their contributions at low laboratory kinetic energies were estimated from the measured H and He by linearly extrapolating isotope ratios (d/H, t/H, <sup>3</sup>He/He) measured as a function of the particle laboratory kinetic energy and angle. Corrected values correspond to triangles in figure 5. The angular distributions for all LCP are very similar, being flat between 60 and 120, and presenting forward and backward peaks, the latter being lower (even if detection thresholds would be accounted for). Angular momentum effects would lead to angular distributions symmetric with respect to 90, and unrealistically large values of spin would be required to explain the measured anisotropy for protons between 0 and 90 for instance. Other arguments against a very high angular momentum of the emitter will be given in section 3. Consequently we consider that two contributions are superimposed: a forward-backward direct component (the forward-backward asymmetry arising from the asymmetry of the incident nuclei ) and an isotropic one which dominates in the angular range 60-120. A maximum estimate of the isotropic contribution coming from the single-source is obtained by doubling the multiplicities measured in this angular range. Its characteristics are summarised in Table 2. The relative abundances of the heavy hydrogen isotopes are higher than those currently observed for evaporated particles (M<sub>t</sub>/M$`{}_{p}{}^{}`$0.1); this indicates that particle emission becomes isotropic rather early in the collision process, and more particularly well before hot and well-separated fragments deexcite through evaporation.
Energy spectra of isotropic LCP emitted in the range 60-90 and 90-120 in the centre of mass are presented in Figure 7. These spectra are, as expected, identical for the two defined angular ranges. However they cannot be fitted by any statistical formula (surface or volume emission). This fact suggests again that these spectra result from emissions at different stages of the reaction, preventing us from inferring any information relative to the source temperature. However we can characterize the isotropic emissions by average quantities (multiplicity and kinetic energy) and slope parameters $`\tau `$ determined from the high-energy exponential fall-off of the spectra (Table 2). The hierarchy of the average lcp energies is a general feature of central collisions between very heavy ions: d and t have equal average energies and slope parameters, much higher than the corresponding proton quantities, and closer to that for $`\alpha `$’s; the $`{}_{}{}^{3}He`$ average energy is in turn spectacularly higher than the $`\alpha `$ energy . The same features are observed in the INDRA Xe+Sn data between 32 and 50 A.MeV, and in the FOPI Au+Au data between 150 and 250 A.MeV . A possible interpretation would be that d,t and $`{}_{}{}^{3}He`$ mostly arise from the multifragmentation stage, while p and $`\alpha `$ are in addition abundantly evaporated by colder and colder fragments, thus lowering drastically their global average energy.
The characteristics of the direct (anisotropic) emission were estimated by subtracting the isotropic contribution determined above from the total distributions. Average values for multiplicities and energies of direct particles are reported in Table 3; the indicated neutron multiplicity, was deduced by assuming that the anisotropic emission keeps the N/Z of the total system. Neutron energy was derived from proton energy by subtracting a Coulomb barrier.
The detection efficiency of INDRA for LCP being excellent (90% of $`4\pi `$), we relied on the LCP rather than on the measured fragments to estimate the single source characteristics: upper limits for the charge, mass and excitation energy of the single source were derived by subtracting the quantities removed by direct particles (without any efficiency correction) (Table 3) from the corresponding values for the total system. The average mass and charge of the source are then $`<A>=378`$, $`<Z>=150`$, which corresponds to 96% of the total system. Its average excitation energy is 6.5 MeV per nucleon to be compared with the 7.1 MeV available for the composite system in the centre of mass.
## 3 Properties of the single-source: sequential decay or simultaneous break-up? Expansion energy
Properties of multifragment systems are frequently derived from comparison with statistical models. The aim is to determine whether thermodynamical equilibrium has been reached at some stage of the collision, through the ensemble of partitions experimentally observed. These models are purely static, in the sense that they describe the system at a freeze-out instant, defined as the time where the nuclear interaction between the fragments vanishes. Fragment kinetic energy only comes from their mutual Coulomb repulsion and thermal motion. Any deviation of the measured kinetic energy from the predicted values is then attributed to some extra collective energy, such as expansion energy, or rotational energy. For instance a self-similar expansion, decoupled from thermal motion, is generally added in order to reproduce the measured kinetic energies. The values of collective energy so derived are however obviously dependent on the Coulomb energy, or in other words on the volume assumed at freeze-out . The Statistical Model for Multifragmentation (SMM ), and the Microcanonical Metropolis Monte-Carlo model (MMMC ) are widely used, for their correct statistical weighting of partitions. The SIMON event generator is chosen here for two reasons. Firstly its highly simplified algorithm for generating partitions reasonably accounts for the measured charge and multiplicity distributions. Secondly, and above all SIMON has the advantage over the models cited above to permit a rigorous treatment of space-time correlations between all fragments and emitted particles.
In SIMON, the inputs are: the mass, charge and available energy of the system; the number and spatial configuration of the primary fragments and eventually their minimum mass; the radial expansion energy at the rms radius. The hot fragments are then propagated while deexciting (through the transition state formalism in the present case). Statistical deexcitation of a hot “compound” nucleus can also be followed, by setting to 1 the number of primary fragments. An additional physical feature was recently included in this code, namely a variation of the level density with the excitation energy: the level density is indeed expected to vanish at high excitation . This is equivalent to excluding from the primary partitions levels with too short life-times. The formalism adopted here is that proposed in , where the level density at energy $`\epsilon `$ is expressed as the Fermi gas level density modified by a modulation factor.
$$\rho ^{eff}(\epsilon )=\rho ^{FG}(\epsilon )\times e^{\epsilon /T_{lim}}$$
(1)
This corresponds to having an effective intrinsic nuclear temperature:
$$T_{eff}^1=T^1+T_{lim}^1$$
(2)
Therefore the available thermal energy of the system, after removing Coulomb and collective parts, is shared between kinetic and intrinsic excitation energy of the $`M_f`$ fragments, following the equation:
$$E=3M_fT/2+\underset{1}{\overset{M_f}{}}aT_{eff}^2$$
(3)
$`a`$ being the level density parameter at zero temperature (taken here as $`a`$=A/10).
The initial system was chosen as that defined in section 2.2, after removing direct particles. A first simulation was performed to test the assumption of sequential deexcitation of the hot source as the origin of fragment emission. Indeed a modified sequential scenario based on the Expanding Emitting Source model was shown to account for some features of the system considered here . With the SIMON simulation, the calculated charge distribution does not extend to large enough values, and the fragment kinetic energies are largely underestimated, as appears in Fig 8a. In the following simulations, multifragmentation of the initial system in 6 primary fragments, with a minimum mass of 20, was assumed. Indeed, when only looking at kinematic variables, the hypotheses of light fragments and particles being produced in the explosion process itself, or evaporated by very hot and spatially close fragments, are expected to give nearly identical results. No angular momentum was considered as it was found that, for this very heavy system, an initial value of 500 $`\mathrm{}`$ (corresponding to b = 4 fm) would increase by less than 5 MeV the final fragment kinetic energy . In the two simulations shown in Fig 8, it was verified that the calculated multiplicity and charge distributions reasonably account for the experimental ones, without attempting a perfect fit. The temperature $`T_{lim}`$ and the collective energy were thus determined by adjusting the calculated fragment kinetic energies on the experimental data. $`T_{lim}`$ may be expected to lie between 5 and 12 MeV: 5 MeV is the temperature obtained from most data based on excited state population ratios, and also from a correlation study between particles and fragments for single sources formed in 50 A.MeV Xe+Sn collisions . 12 MeV is the maximum nuclear temperature predicted for nuclei when considered as liquid drops in equilibrium with their vapor . Without collective energy, the derived value of $`T_{lim}`$ is equal to 8.5 MeV; in this case the experimental variation of the fragment kinetic energy (including the largest one) with their charge is almost perfectly reproduced, independently of how the fragments are initially sitting in space. As soon as expansion energy comes into play, selected initial spatial configurations must be chosen, with the heaviest fragment close to the centre of gravity, in order to follow the experimental saturation of the kinetic energy at high Z (see Fig 2). The rising part of the curve E(Z) is however independent of the spatial configuration, and a way to avoid a configuration choice is to consider all fragments but the largest, as in Fig 8a. An equally good reproduction of experimental data is obtained with $`T_{lim}`$=10 MeV and $`\epsilon _r`$ = 0.5 A.MeV; the initial excitation energy of the fragments in this case is 4.9 A.MeV. The maximum collective energy compatible with the data is 1 A.MeV if $`T_{lim}`$ is increased to 12 MeV. Fragment spectra were then scrutinized, in order to choose between the hypotheses with 0 or 0.5 A.MeV expansion energy. In Fig 8b is shown the experimental spectrum for Z=7, and the calculated spectra with and without expansion energy. Clearly the hypothesis with expansion energy is the best, the option $`\epsilon _r`$ = 0 leading to a too broad spectrum. Note that the discrimination between the two assumptions can only be made for light fragments (Z$`<`$10), as for charges Z=12-20 the two calculated spectra are almost superimposable.
In conclusion of this section, the comparison of measured fragment kinetic energies with a simultaneous break-up scenario reveals the need for a radial collective motion at the freeze-out whose energy is $`\epsilon _r`$ = 0.5 $`\pm `$ 0.5 A.MeV. This indicates that expansion energy begins to appear for central collisions between very heavy ions around 30 A.MeV, as also found for a similar system . The origin of this expansion (compression, thermal pressure?) will be discussed in section 5, with the help of dynamical simulations.
## 4 Experimental evidence for a bulk effect: the fragment charge distribution is independent of the charge of the total system.
$`{}_{}{}^{X}129`$ \+ $`{}_{}{}^{a}n`$ reactions between 25 and 50 A.MeV were also studied with the INDRA array . From 32 A.MeV up, compact single source events could be isolated with the same flow angle selection as explained in the accompanying paper. At 32 A.MeV , the available excitation energy per nucleon for the total system is the same as for the $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$ system studied in the previous sections. An experimental effect, observed for the first time, stands out for single-source multifragmentation events from these two systems (Fig 9): the charge distributions (normalised by the average multiplicity of Z$``$3) superimpose over 3 orders of magnitude while the average fragment multiplicities are in the ratio 1.49, i.e. almost exactly the ratio of the total charges of the systems 156/104. Such a peculiar behaviour may reveal that bulk effects play a major role in the multifragmentation of these systems. Note that the same Z distribution is also observed in ref. for the 35 A.MeV Au+Au system, which has a total charge close to $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$.
As shown in section 3, the kinematical properties of single-source events are compatible with the assumption that, at some instant of the collision a freeze-out configuration is reached, i.e. the system can be described as an ensemble of (hot) charged products and neutrons in thermodynamical equilibrium, free from mutual nuclear interaction. This implies that the total system at this instant occupies a much larger volume than would a nucleus with the same mass, or in other words that it has reached a low (though inhomogeneous) density state. Statistical models like MMMC, SMM do predict that two systems with the same thermal energy (temperature) will, on average, break up into a number of fragments proportional to their mass, with the same fragment charge distributions . In these models the charge (mass) of the largest fragment essentially reflects the excitation energy; indeed the measured average charges of these fragments only differ by 7% for the two systems studied in this section, 32 A.MeV $`{}_{}{}^{X}129`$ \+ $`{}_{}{}^{a}n`$ and 36 A.MeV $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$ (Fig 12).
The dilute systems considered in the statistical models have, in the experiments, been produced through violent nuclear collisions, and have lived quite a long history before reaching this stage. How such a low density state is attained should bring information on the role of fundamental properties of nuclear matter such as incompressibility, viscosity …For similar systems of colliding heavy ions, some transport models for instance predict that in central collisions below 50 A.MeV, an irreversible evolution towards low density is obtained after a compression phase if the incompressibility parameter of nuclear matter is close to 220 MeV; conversely for larger K the system is less strongly compressed and returns to normal density . In other calculations however, multifragmentation in central Pb+Au collisions appeared to be driven by neutron/proton asymmetry rather than by the softness or stiffness of the equation of state .
Instabilities of different kinds (volume, surface …) which are dynamically generated in the course of the collision may be suspected of causing the break-up of the system . For the heavy systems considered here ($``$ 250 and 400 nucleons), bulk instabilities may particularly be expected.
## 5 Comparison with a stochastic mean-field model
The family of dynamical simulations of nuclear collisions based on the nuclear Boltzmann equation, the Landau-Vlasov (LV), Boltzmann-Uehling-Uhlenbeck (BUU) or Boltzmann-Nordheim-Vlasov (BNV) codes, were very successful in accounting for a variety of experimental findings . However, as they follow the time evolution of the one-body density, neglecting higher than binary correlations, they ignore fluctuations about the mean trajectory of the system, which becomes a severe drawback if the system happens to explore regions of instability such as the spinodal zone, for instance. Indeed, in the presence of instabilities, local density fluctuations are propagated and amplified by the mean-field leading to a multi-fragment breakup of the system. This is the so-called spinodal decomposition scenario for multifragmentation. The spinodal decomposition described in ref. , predicts a “primitive” break-up into fragments with a favoured size, connected to the wave lengths of the most unstable modes in nuclear matter. These wave lengths are roughly constant as long as the systems are large enough. Therefore break-up into equal-sized fragments is expected, with a multiplicity growing with the mass of the system. This very simple picture is in reality blurred by several effects: the beating of different modes, the coalescence of fragments while the nuclear force still acts and the finite size of the system will tend to end-up with a broad, exponentially falling distribution of fragment charges. Therefore it first appeared easier to determine whether spinodal decomposition would predict, for two different systems, the effect described in section 4 rather than to look directly for a trace of spinodal properties in a single system.
Let us now come in more detail to the predictions of stochastic mean-field simulations of nucleus-nucleus collisions, based on the Boltzmann-Langevin (BL) equation, which allows for the treatment of unstable systems. Since the application of the BL theory to 3D nuclear collisions is still too computer demanding for calculating realistic scenarios and for quantitative results, methods allowing to simulate approximately the dynamical path followed by nuclear systems crossing the spinodal region were developed. They consist of a “Boltzmann evolution” starting from an inhomogeneous initial system (Stochastic Initialisation Method: SIM) or complemented by a brownian force (Brownian One-Body Dynamics: BOB). In either case the amplitude of the initial density fluctuations or the magnitude of the stochastic force is chosen to reproduce the dynamics of the most unstable modes for infinite nuclear matter in the spinodal region.
A first comparison between experimental data and calculation was published in reference ; it was found that the multiplicity and charge distributions for both systems - and thus the identical fragment charge distributions- were correctly reproduced, but that the calculated fragment kinetic energies were too low. In this previous paper however, the poorest of the two methods for simulating the BL equation proposed was used (SIM): it consisted in initializing the correct (classical) density fluctuations when the system enters the spinodal region, but continuing afterwards a standard one-body calculation, which unfortunately leads to some damping of fluctuations even when the mean field is unstable, because of the lack of a source term. As a result, the fragment formation time was incorrectly increased, leading to smaller final kinetic energies because of the decrease with time of the expansion velocity of the unstable source. The second, and better, simulation method proposed (BOB) will be used for the comparisons presented in the following : the density fluctuations are now continuously created through the addition to the mean field of a stochastic force, and therefore maintained all along the evolution of the unstable system. A second major improvement has been implemented in the simulations: quantal fluctuations connected with collisional memory effects are now taken into account, as calculated in , with the determinant result of doubling the overall amplitude of fluctuations for the most unstable modes in our case. The dynamics of head-on 32 A.MeV $`{}_{}{}^{X}129`$ \+ $`{}_{}{}^{a}n`$ and 36A.MeV $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$ collisions was repeated with the improved simulations. The method used to select our experimental data restricts to events which are highly relaxed in form (see Part I), representing only a part of all the multifragmenting single-source events . Therefore only collisions at zero impact parameter were simulated, shape effects are neglected. It should be noticed that although the noise amplitude is computed to match the fluctuation of bulk instabilities, it may trigger any type of instabilities.
In a coherent treatment of the BL theory, fluctuations should be implemented from the beginning of the reaction, though their role only becomes crucial when instabilities appear. Indeed, they then lead to bifurcations and an ensemble of identically-prepared systems will explore a large number of different dynamical trajectories, which produce different multi-fragment partitions. Thus the stochastic mean field calculation has to be performed as many times as the number of events which one wishes to generate, which is very computer time-consuming for large statistics. Moreover the BOB dynamics is only applicable to locally-equilibrated systems and so cannot give a correct description of the very earliest, far off-equilibrium stages of the collision. Compromises therefore have to be made. As the BNV calculation of the reaction without fluctuations gives a unique evolution, it was performed only once and the results used as initial conditions for the BOB dynamics. In order to correctly simulate the growth rate of fragments in the unstable system, the amplitude of density fluctuations must be correct at the time when the spinodal region is reached. Then the BOB calculation has to begin before the onset of instability, because the brownian force cannot set up density fluctuations of the correct amplitude instantaneously, due to the test-particles’ inertia and finite relaxation times. We chose as starting point the moment of maximum compression of the system (t=40 fm/c, see Table 4), when the local equilibrium is established, and we verified that the characteristics of the system at the onset of instability in the BOB calculation were identical to those found from the ‘true’ one-body evolution of the reaction calculated with BNV.
The ingredients of the BNV/BOB simulations are as follows. The self-consistent mean field potential chosen gives a soft equation of state (K= 200 MeV) and the finite range of the nuclear interaction is taken into account using a convolution with a gaussian function with a width of 0.9 fm . The addition of a term proportional to $`\mathrm{\Delta }\rho `$ in the mean-field potential allows to well-reproduce the surface energy of ground-state nuclei . This is essential in order to correctly describe the expansion dynamics of the composite source. In the collision term a constant $`\sigma _{nn}`$ value of 41 mb, without in-medium, energy, isospin or angle dependence is used . It should be noticed that, using a stiff parameterization of the equation of state (K= 380 MeV), the dynamical evolution of the composite source follows a very different path: density oscillations remain small and the system does not enter the spinodal region. Characteristics of the Gd+U and Xe+Sn systems at two times are shown in Table 4.
A maximum density 25-30% higher than normal density is reached after 40 fm/c, corresponding to a moderate compression. Then during their expansion phase the two systems enter the spinodal region at around 80 fm/c and attain slightly later thermal equilibrium at low density, with a temperature of 4 MeV. The radial velocity at the surface ($``$0.1c) reveals the gentle expansion of the systems and the density fluctuations have time to develop leading to the formation of fragments. For Gd+U, the charge and mass of the source at 100 fm/c are close to the values determined for the experimental single-source emitting isotropically particles and fragments (end of Sect.2). An algorithm for reconstructing fragments is applied at intervals of 20 fm/c, based on a minimum density cut-off. The calculation is stopped when the fragment multiplicity becomes constant and independent of reasonable variations of the value of the cut-off density, provided it is larger than 0.01 fm<sup>-3</sup>, as shown in Fig. 10. The chosen density cut-off was 0.02 fm<sup>-3</sup> for $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$; because of the weaker Coulomb energy and radial flow for $`{}_{}{}^{X}129`$ \+ $`{}_{}{}^{a}n`$ collisions (see table 5), which lead to a different event topology at freeze-out, a higher density cut-off of 0.05 fm<sup>-3</sup> was used for this system.
The different characteristics of the fragments so defined can be calculated: mass, charge, linear and angular momentum, kinetic energy and excitation energy. A temperature is deduced from the relation T=$`\sqrt{E^{}/a}`$ using the density dependence of $`a`$ in the framework of the Fermi gas model. The calculated kinetic energy of fragments does not contain the thermal part. This is due to the fact that the amplitude of the BOB term is tuned in order to reproduce the dynamics of the most unstable modes. Hence the fluctuations that are introduced act mostly on the configuration space, leading to fragmentation, but in principle they do not automatically provide a sufficiently good description of fluctuations in momentum space. The approximate solution to this problem used here is to add to each component of the fragment velocity a random part, distributed according to a gaussian of width $`\sigma ^2=T/M`$, where $`M`$ is the mass and $`T`$ the average internal temperature of the fragment considered. This results in an increase of the average kinetic energy of fragments of $`3/2T`$. The excitation energy of fragments is then reduced by the same quantity, in order to conserve the total energy. Attention is also paid to momentum conservation, event by event.
At the end of the dynamical simulation (t$``$ 240-260 fm/c) the fragments are well separated, and still bear an average excitation energy of $``$3 A.MeV, as seen in Table 5. Note that this value is, as expected, smaller than that quoted in section 3 where the simplified assumption that “gas particles” were bound in the hot fragments was made. The mass of the fragments accounts for about 80% of the total system mass; at this stage, the experimental observations reported in section 4 are practically fulfilled: the average charge of fragments for the two systems is the same ($`<Z>13`$) and the average multiplicities are in the ratio of 1.6, which corresponds to the ratio of total masses bound in fragments. The average radial energy reported in Table 5 for the $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$ system is larger than the extra collective energy introduced in the SIMON event generator in section 3 in order to account for the average kinetic energies of the fragments. The Coulomb energy is however slightly smaller; this stresses the difficulty to experimentally determine small expansion energies, when their magnitude is similar to the amount of Coulomb energy at the freeze-out.
In a second step of the simulation, the spatial configuration of the primary fragments, with all their characteristics as given by BOB, is taken as input configuration in the SIMON code (see sect. 3): the de-excitation of the hot primary fragments is thus followed while preserving the space-time correlations of all emitted products. The last step consists in filtering the events to take into account the experimental set-up and then selecting events with large flow angles as for the data. It must be noted that the algorithm for fragment reconstruction excludes light products, and thus all light charged particles emitted before the freeze-out are lost. This has two consequences: firstly the selection of complete events is not made on the total detected charge, but on $`Z_{bound}`$ (see section 2); experimental data on Figs. 11-14 use the same selection. Comparisons with the corresponding figures from section 2 show that the two selections give identical results. Secondly, no comparison can be made between calculated and experimental light charged particle properties.
Figure 11 shows that the calculated multiplicity and charge distributions of fragments well match the experimental data. The role of secondary decay in the fragment distribution is negligible, as observed by comparing the grey (primary fragments) and the solid (final fragments) histograms. More detailed comparisons of the charge distributions of the three largest fragments display a good agreement (Fig. 12): the increasing difference between the average charges for the two systems when going from the largest to the second and the third largest fragments is accounted for by the simulation.
The shapes of the events expressed for instance through the isotropy ratio, with its multiplicity dependence, are identical in the experiment and in the simulation, as shown in Fig. 13a. Conversely the angular correlations between fragments presented in Fig. 13b are less populated at small relative angles in the calculation. This could suggest that the Coulomb effect is too strong in the calculation (longer real emission times?). The same effect would also be obtained if the freeze-out shape resembled a bubble, but in this case the event shape and the evolution of the fragment energies on their charge would be different. The experimental selection can also retain some deformed events , which are not accounted for in the simulation where spherical symmetry was kept. The disagreement between experiment and calculation on this point is not fully explained.
Finally the most crucial test is on the fragment kinetic energies, displayed in Fig. 14. Let us recall that as compared to ref. , the energy calibration of the INDRA CsI located past 45<sup>o</sup> (lab) were improved for fragments with charge larger than 15 (see Fig. 2). This correction is particularly important when the velocity of the reaction centre of mass is small, and the fragment velocities in the c.m. are large. Namely the new results are very different for the 36 A.MeV Gd+U and practically unchanged for the 32 A.MeV Xe+Sn central collisions. The improved simulations now predict the fragment energies for the Gd+U system with a good agreement; for Xe+Sn, the calculated energies fall $``$ 20% below the measured values; this remains satisfactory if one remembers that there were no adjustable parameters in the simulation.
## 6 Conclusions
The properties of the single multifragmenting source formed in central 36 A.MeV $`{}_{}{}^{G}155`$ \+ $`{}_{}{}^{a}n`$ collisions were studied in detail. The angular and energy distributions of all fragments and of most of the light charged particles are isotropic in the c.m. This indicates that the source has reached thermal equilibrium when it multifragments. Fragment kinetic energies sign the onset of expansion energy around 30 A.MeV.
The measured charge distribution is identical to that obtained for the lighter $`{}_{}{}^{X}129`$ \+ $`{}_{}{}^{a}n`$ system with the same available energy, while the average fragment multiplicities are in the ratio of the total charges of the systems. This independence of the Z distribution, experimentally observed for the first time, can be considered as strong evidence of a bulk effect in the production of these fragments. Note that this observation naturally breaks down when the total charge of the systems gets smaller than $``$60, which is the maximum measured fragment charge for the heavy systems considered here.
This experimental observation can be related either to bulk instabilities in the liquid-gas coexistence of nuclear matter (spinodal instabilities) or taken as a signature of a large exploration of phase space for such heavy systems. Indeed multiplicities, charge distributions and average fragment kinetic energies are equally well predicted by dynamical (BNV-BOB) and statistical (SMM) simulations .
Within the dynamical approach, a complete scenario of these central collisions may then be proposed: a gently compressed system expands and reaches thermal equilibrium at about the time it enters the spinodal region. There, the development of density fluctuations causes the disassembly of the system into many (hot) fragments and particles. When the fragments get free from the nuclear force, they are in a “freeze-out” configuration, and only subject to Coulomb repulsion. At that time, $``$250 fm/c in the simulations, the system has explored enough of the phase-space in order to be describable through statistical models. Within such a scenario, there is no contradiction between a “dynamical” and a “statistical” approach; the first one completely describes the time evolution of the collision, and thus helps in learning about nuclear matter and its phase diagram. The second starts from the phase diagram and has more to do with the thermodynamical description of finite nuclear systems.
Ultimate constraints on models can be expected from the study of correlations: fragment velocity and size correlations will allow to trace back to the fragment topology at freeze-out, and to look for possible fingerprints of spinodal decomposition. This work is in progress.
|
warning/0007/cond-mat0007247.html
|
ar5iv
|
text
|
# New Superhard Phases for 3D C₆₀-based Fullerites
## Abstract
We have explored new possible phases of 3D C<sub>60</sub>-based fullerites using semiempirical potentials and ab-initio density functional methods. We have found three closely related structures - two body centered orthorhombic and one body centered cubic - having 52, 56 and 60 tetracoordinated atoms per molecule. These 3D polymers result in semiconductors with bulk moduli near 300 GPa, and shear moduli around 240 GPa, which make them good candidates for new low density superhard materials.
Superhard materials are of obvious practical and theoretical interest. Among them, diamond is still the undisputed hardest substance, and will take some beatings. Its hardness comes from the stiffness of its tetravalent carbon bonds, which are also the basis of the widely used diamond-like coatings. However, natural diamond is rare and its industrial production expensive. For this reason, much experimental and theoretical work has been devoted to studying novel potential candidates for hard materials. For instance, another carbon-based crystal, 3D $`sp^2`$ carbon , was theoretically predicted to be even harder than diamond but never synthesized. Very recently, a high pressure phase of silicon clathrates was studied . Its bulk modulus resulted to be 90 GPa, very close to that of the normal silicon cubic crystal, leading to the speculation that, if constructed with carbon instead of silicon, it could be as hard as diamond. A similar idea was proposed for $`\beta `$-C<sub>3</sub>N<sub>4</sub> in relation to $`\beta `$-Si<sub>3</sub>N<sub>4</sub> but because only small quantities of the former material are available, it remains to be proven that the substance is indeed harder than diamond. A general theoretical approach to study hardness in any of its definitions is difficult, if not impossible, to perform. Traditionally, it has been linked to a large bulk modulus but it has been shown recently to correlate better with the shear modulus of the material ; diamond is the leading system in both cases, with 443 and 535 GPa, respectively.
The discovery of C<sub>60</sub> in 1985 as a third form of crystalline carbon was the kick-off for one of the most stimulating fields for research in materials science in the last decades. For the first years after its discovery and massive production, the known solid state phases of pure C<sub>60</sub> were nearly ideal molecular crystals . However, from the very beginning, there were speculations on the possibility of obtaining crystals in which the C<sub>60</sub> molecules were covalently bonded to each other, by means of high pressures and temperatures . Since then, several papers have appeared reporting a variety of experimental results in this line (see Ref. for an updated review of the field).
Nuñez-Regueiro et al. reported the crystal structures of partially covalent forms of C<sub>60</sub>. They observed a crystal containing linear chains of polymerized fullerene molecules and also two structures, one rhombohedral and one orthorhombic, of covalently bonded molecular planes (2D C<sub>60</sub>). The structures are formed by bonding through 2+2 cycloaddition reactions between neighboring molecules with a carbon hybridization change from $`sp^2`$ to $`sp^3`$; chains and planes are bound together by Van der Waals interactions. More recently, the realization of 3D C<sub>60</sub>-based fullerite has been reported . At least four crystal structures were found to be compatible with the X-ray data: rhombohedral, orthorhombic (pseudo tetragonal), tetragonal and simple cubic. Of these, the first two are formed by stacking covalent rhombohedral and tetragonal planes. Blank and coworkers have reported the obtention of 3D fullerene solids of extremely low compressibility, and very recently a superhard structure has been also reported in detail by Chernozatonskii et al. ; the crystal is body centered orthorhombic, quasi fcc, and each molecule develops two intermolecular bonds with each of its 12 neighbors, presenting corrugated surfaces. From the theoretical side, Okada et al. have reported calculations on a possible structure for pseudo-tetragonal (body centered orthorhombic) covalently bonded 3D C<sub>60</sub>, which might be obtained by applying a uniaxial pressure of about 20 GPa to pseudo-tetragonal polymerized layers of C<sub>60</sub>. This phase would be metallic with a high density of states at the Fermi level, and the C<sub>60</sub> molecules would be highly corrugated.
In this work we report new 3D C<sub>60</sub>-based fullerite phases which have a direct relation with the most probable precursors: polymerized linear chains and orthorhombic planes, while the molecules preserve their convexity. We will show that our proposed structures have low compressibilities and large shear moduli and therefore are excellent candidates for hard materials.
We have proceeded as follows: an initial search of the stable or metastable structures was done using Tersoff potentials , which are known to provide a good description of carbon fullerenes , including several polymeric phases . Once the potentially interesting structures were found, these were used as starting points for more detailed and accurate ab-initio calculations. The structure, stability and electronic properties of each phase were obtained, allowing the complete relaxation of cell parameters and atomic coordinates, and analyzed; all data we will give here (except for the calculated vibrational frequencies) are those of the final ab-initio results.
The ab-initio calculations were performed using a numerical-atomic-orbital density functional (DFT) method described in detail elsewhere . It has been already applied to large fullerene molecules and nanotubes , and many other systems . The calculations are done using the Generalized Gradient Approximation for the exchange-correlation, as parametrized by Perdew, Burke and Ernzerhof . Core electrons are replaced by nonlocal, normconserving pseudopotentials factorized in the Kleinman-Bylander form , whereas valence electrons are described using linear combinations of pseudo-atomic orbitals. In this work we have used a split-valence double-$`\zeta `$ basis set, supplemented with polarization $`d`$ orbitals (DZP) . The radial cutoffs were 4.2 and 5.0 a.u. for the $`s`$ and $`p`$ orbitals respectively. We have carried out tests using a triple-$`\zeta `$ basis set, finding only minor changes in the lattice constants (less than 0.3 %) and elastic moduli. The charge density was expanded in plane waves with a large cutoff of more than 150 Ry (the exact values vary slightly, depending on the volume). With these parameters we have obtained a bulk modulus for diamond of 430 GPa, in good agreement with known values . The relaxation of the cell parameters and atomic coordinates was performed using conjugated gradient minimizations at fixed pressure.
Starting from two experimentally observed 1D and 2D polymerized body centered orthorhombic C<sub>60</sub> structures, we obtain two new metastable 3D phases. These have 52 and 56 tetracoordinated atoms per molecule, respectively, and we will refer to them as (52-8) and (56-4). The (56-4) structure is obtained by applying uniaxial pressure along the stacking direction of a 2D polymer, while for the (52-8) one the pressure is applied in the plane perpendicular to the direction of linear chains. In those phases, the individual molecules preserve their convexity as well as the bonds between C<sub>60</sub> molecules originally present in the chains or planes (4 and 8, respectively). In both cases, six new bonds are formed with each of the new nearest neighbor molecules in the body centered position. Both structures are intimately related to one proposed by O’Keeffe, based on purely geometric arguments : it is possible to construct a body centered cubic crystal with $`T_h`$ symmetry, having all carbon atoms with coordination four, connected by bonds of equal length. We have also studied this phase, from which, once all independent degrees of freedom are relaxed, a (60-0) structure is obtained. From the electronic point of view, all the new structures are semiconductors, with bands gaps larger than 2 eV. A general view of the (56-4) structure is schematically shown in Fig. 1, where also the few remaining sp<sup>2</sup> atoms can be seen. The (52-8) phase differs from the (56-4) in that both the (001) and the (010) planes look like the (001) plane in Fig. 1. For the (60-0) structure, the (001), (010) and (001) planes, which are equivalent by symmetry, are bonded as in Fig. 1-a.
Table I shows the relevant structural data and physical parameters of the new fullerites . The specific volumes lie between those corresponding to graphite (9 $`\mathrm{\AA }^3`$/atom) and diamond (5.6 $`\mathrm{\AA }^3`$/atom) but due to the covalent intermolecular bonds are much smaller than those of Van der Waals C<sub>60</sub> and C<sub>70</sub> crystals ($``$ 11.5 $`\mathrm{\AA }^3`$/atom). The length of the bonds between tricoordinated atoms is about 1.35 $`\mathrm{\AA }`$; between tri- and tetracoordinated atoms about 1.48 $`\mathrm{\AA }`$; the remaining intramolecular bond lengths range from 1.48 to 1.70 $`\mathrm{\AA }`$. The bonds forming the hexagonal turrets which connect molecules in the body centered position are about 1.57 $`\mathrm{\AA }`$ long, and those connecting second neighbor molecules 1.62 and 1.66 $`\mathrm{\AA }`$. As expected, the bond angles are grouped around the values of graphite, diamond, and 90 degrees; the latter are associated with the turrets and intermolecular second neighbor bonds. For (60-0), where only sp<sup>3</sup> bonds are present, we obtain bonding distances of 1.475, 1.548 and 1.56 $`\mathrm{\AA }`$. Regarding the intermolecular distances between first neighbor molecules in the (110) plane (i.e., those
corresponding to the maximum number of covalent bonds between C<sub>60</sub> molecules), the typical values are around 8.4 - 8.5 $`\mathrm{\AA }`$, which agree very well with those experimentally found by Marquez et al. (8.40 and 8.80 $`\mathrm{\AA }`$).
The number of $`sp^3`$ bonds suggested the interesting possibility of low compressibility materials, and in fact our structures have bulk and shear moduli around 300 GPa and 240 GPa, respectively. They should be compared to the values calculated with the same method for diamond, 430 and 560 GPa. According to the empirical relation found by Teter between hardness and shear modulus , these values would correspond to a hardness of about 40 Vickers, which would place them among the ten or so hardest known systems, at the lower limit of superhard materials. It is also interesting that the hardness expected for these phases comes together with a considerable lightness of the materials, since the densities are about 30 % smaller than that of diamond.
Energy vs. volume curves for the new phases are shown in Fig. 2. It can be seen that the (56-4) and (52-8) structures have almost the same equilibrium energy and are more stable than the (60-0) by about 1 eV. A suggested (56-4) $``$ (60-0) transition is found at about 14 GPa, in which the (56-4) structure is compressed along one of its axes. We have also recalculated the structure of Okada et al. finding that the (56-4) crystal is lower in energy by almost 2 eV, and has a smaller volume. The structures obtained in this work (and that of Ref.20) are metastable with respect to the corresponding precursor planes and chains, and correspond to local minima in configuration space. However, as seen in Fig.2, they have a large stability range, and require quite considerable amounts of energy to be dissociated.
It is interesting to note that these phases could be recognized by vibrational spectroscopy methods. Fig. 3 shows the vibrational DOS for the isolated C<sub>60</sub> molecule, and the (52-8), (56-4) and (60-0) structures, calculated using the potentials described in Ref. ; the bands above 1500 cm<sup>-1</sup>, which are related to graphite-like vibrations in C<sub>60</sub> (the in-plane optical frequencies of graphite lie around 1590 cm<sup>-1</sup>), completely disappear in the (60-0) crystal; only four isolated optical bands remain in the (56-4) compound, and eight in the 52-8 one.
In conclusion, we have studied new phases of 3D C<sub>60</sub>-based fullerites which seem to have quite interesting features: their bulk moduli are about 70 % of diamond and their shear moduli compare with those of the hardest systems known; the phases are extremely compact, although the crystals would be light compared to most other hard materials; the original convexity of the molecules is preserved - corrugation is a costly energy consuming process; and the (52-8) and (54-6) structures are directly related to polymerized chains and layers, which seem to be most likely precursors for 3D polymerized phases, and could probably be obtained at extreme, possibly uniaxial, pressures. The present result, joined to that of the metallic phase reported by Okada et al. and the latest experimental data, suggest an exciting scenario for research on new fullerene based materials.
This work was partially supported by CONICET Grant PMT-PICT 0051. R.W. acknowledges support from Fundación Antorchas Grants No A-13622/1-103 and A-13661/1-27. P.O and R.W. acknowledge support by Motorola PSRL. Part of the research has been done using the computing resources of CESCA and CEPBA, coordinated by C<sup>4</sup>.
|
warning/0007/hep-ph0007051.html
|
ar5iv
|
text
|
# JETS IN HADRON COLLISIONS
## 1 Introduction
Jet physics is a very vibrant field and there has been a great deal of progress both experimental and theoretical in the last few years. Rather than trying to give a review of the whole field, I have chosen to go into two topics in more depth: jet definitions and jet substructure.
## 2 Jet definitions
### 2.1 Cone algorithms
The standard jet algorithms used by most hadron-collider experiments have been based on the geometric cone definition. Although the general idea of this definition is straightforward, when implementing it one is faced with myriad choices and historically each experiment implemented its own algorithm. The Snowmass Accord was an attempt to unify these and agree on one definition that theorists and experiments could use. It defines jets by finding directions in rapidity–azimuth, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$, space that maximize the amount of hadronic energy flowing through a cone of fixed radius (in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ space), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$, drawn around them. The jet momentum is defined to be massless with transverse energy, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$, rapidity and azimuth calculated from those of the particles in the cone, as
$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cone}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (1)
$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}}{\displaystyle \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cone}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2)
$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}}{\displaystyle \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cone}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (3)
#### Iterative cone algorithms
The CDF experiment, which first tried to implement the Snowmass Accord$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$, found that it was not a complete definition and had to be supplemented for two reasons. The result was an iterative cone algorithm. Subsequent experiments have followed a similar line, although the fine details vary from experiment to experiment.
The first problem is that the maximization process was not uniquely defined. A global maximization proved too costly in computer time to be practical, so they defined a local maximization which was achieved iteratively. They define a set of directions to be *seed directions*, draw a cone around each and apply Eqs. (13), which define a new direction. A cone is drawn around this direction and Eqs. (13) used again iteratively, until a stable direction is obtained. It can be shown that provided all calorimeter cells have a positive energy (which is not necessarily the case experimentally, for example in D0) a stable direction is always reached and that it gives a local maximum of the energy in the cone. The definition of the seed directions is closely tied to the details of specific detectors, but is typically every calorimeter cell above some energy threshold, eg 1 GeV.
The second problem is that the jets so defined often overlap and share energy in common, while a mapping of each calorimeter cell to only one jet was sought, so a merging/splitting algorithm was added. Again the precise details vary from experiment to experiment, but the general idea is to either merge the two overlapping jets into one, if the overlap region contains more than a given fraction of their total energy, or to split them into two along the half-way line, otherwise.
It has recently been realized$`^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}}`$ that the iterative cone algorithm is not infrared safe. The problem is that the iteration from the seed directions is not exhaustive: it is not guaranteed to produce a complete list of all local maxima. If two cones overlap in such a way that their centres can also be enclosed in one cone but there is little energy in the overlap region, then it turns out that the outcome is different depending on whether or not the overlap region contains a seed direction. This results in a logarithmic dependence on the seed cell threshold which would give a divergent cross section, if the threshold were taken to zero for the purposes of making an idealized calculation. This divergence first shows up when there can be three nearby partons, which for jets in hadron collisions is NLO in the three-jet cross section$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$ and NNLO in the two-jet or inclusive one-jet cross sections$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$.
It is also worth noting that this is mainly a problem for order-by-order perturbation theory.
As shown in Fig. 1 (Fig. 2 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$), after summing to all orders, the dependence on the cutoff is very weak. This is because physically almost every such event does in fact contain a seed direction and one gets a Sudakov form factor much less than unity. When expanded order-by-order in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$, such a form factor gives large terms at every order.
The solution$`^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}}`$ is to add an additional stage of iteration. After the first stage, but before the merging/splitting algorithm, additional seed directions are tried, defined as the mid-points of any overlapping cones.
As shown in Fig. 2 (Fig. 4b of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$), this results in much more stable cross sections, which are finite order-by-order in perturbation theory.
Although it was originally thought that, as stated above, this only became important to the inclusive cross section at NNLO, it has more recently been realized$`^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}}`$ that in DIS it appears at NLO, if the jets are analyzed in the lab frame. This is because the outgoing electron acts kinematically like a jet, against which the other jets in the event recoil, but since it is not coloured it does not contribute to the QCD corrections. We can therefore test these ideas using standard NLO calculations of two-jet production in DIS. An example is shown in Fig. 3 (Fig. 1 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$), for the dijet cross section in the lab frame at HERA.
The results in the iterative cone algorithm are clearly out of control, while those in the improved cone algorithm are considerably better. It is worth noting that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ algorithm, to be discussed shortly, is better behaved still. Similar results were found by the jets working group of the Physics at Run II workshop$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$.
#### The Improved Legacy Cone Algorithm
A recent innovation arising from the Physics at Run II workshop is a new accord on how to define cone jets, the ILCA, based on the improvement suggested in Ref. $`^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}}`$.
As shown in Fig. 4 (Figs. 4, 5 and 18 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) it is fully algorithmic, if a little cumbersome, meaning that any experiment or theorist can implement it in exactly the same way. Among the requirements it had to fulfil is that its numerical result be within 5% of the algorithms used by CDF and D0 in Run I, which is the case. Despite this small difference at the hadron level, it is finite order-by-order in perturbation theory and has smaller hadronization corrections. It is therefore a significant step forwards.
### 2.2 The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ cluster algorithm
Despite the improvements in the cone algorithm, it still has problems relative to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ cluster algorithm$`^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}}`$.
The definition of this is shown in Fig. 5 (Fig. 19 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) in the same notation: it is clearly much simpler. Its results depend on an input parameter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ (sometimes called $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$), which actually plays a similar role to the radius parameter in the cone algorithm. Among its advantages are its simplicity, the fact that it exhaustively maps every hadron in the final state to one and only one jet with no overlaps, and the fact that it is based on the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ measure, allowing the phase space for sequential soft gluon emission to be factorized and the corresponding large logarithms to be summed to all orders. It also suffers smaller hadronization and detector corrections than the cone algorithm in practice.
As shown in Fig. 6 (Fig. 2b of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$), at the level of the NLO inclusive jet cross section, the two jet definitions are essentially identical.
However, even at that order the energy spread within the jet is quite different, as shown in Fig. 7 (Fig. 1 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$).
The cluster algorithm pays more attention to the core of the jet, while the local optimization inherent in the cone algorithm does its best to suck as much soft junk into the edges of the jet as possible. This is thought to be why the cluster algorithm gives significantly cleaner reconstruction of highly boosted objects as shown in Fig. 8 (Fig. 1b of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$).
Both H1 and ZEUS now use the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ algorithm as their ‘algorithm of choice’, and CDF and D0 are planning to use it on an equal footing with the ILCA in Run II.
#### Preclustering
One small problem with the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ algorithm that has yet to be solved is the possible need for a preclustering step. This is needed by the Tevatron experiments for several reasons.
Firstly in D0 some calorimeter cells have negative energy, and it is not entirely clear whether these can be incorporated into the algorithm (although it is not clear to me that they cannot, for example by replacing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ by
$$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sign}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$
(4)
These negative energy cells would then always be clustered with their nearest neighbours in a process that would always continue until there are no negative energy clusters left, regardless of the jet resolution criterion).
Secondly, due to calorimeter segmentation and the finite transverse size of hadronic showers, it is possible for one hadron to produce two or more non-zero energy cells, or for two hadrons to shower into a single cell. Preclustering reduces the size of the detector corrections associated with these effects, particularly at small subjet resolution scales.
Finally, the clustering process takes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ time, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ is the number of initial momenta. For Tevatron events this time can be prohibitive if all calorimeter cells are used as input. It can be reduced considerably by a little preclustering$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$.
Unfortunately no theoretical implementation of preclustering has been proposed as yet. It is not even clear whether it is possible to satisfy the experimental needs with a theoretically-calculable algorithm. A possible solution is to run the inclusive $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ algorithm with a small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ parameter, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.2}$}`$, and to use the output of this algorithm as an input to the main algorithm. It seems likely that the large logarithms associated with this could be summed to all orders, but it has not been explicitly checked. Clearly it does not solve the problem of cpu time, so is not a complete solution, but if it could be shown that this has similar results to the experimental algorithm with the same precluster radius then it could be used as a common ground on which to compare theory and experiment. That is, one could correct the experimentally-preclustered calorimeter-level results to theoretically-preclustered hadron level, rather than all the way to un-preclustered hadron level.
## 3 Jet structure
### 3.1 Jet shape
The classic way to study the internal structure of jets is with the jet shape. This is inspired by the cone algorithm, although its use is not limited to cone jets. The jet shape<sup>b</sup><sup>b</sup>bNote that, perversely, the HERA experiments have defined their notation for $`\colorbox[rgb]{1,1,1}{$\mathrm{\Psi }$}`$ and $`\colorbox[rgb]{1,1,1}{$\rho $}`$ to be interchanged relative to the original definitions used by the Tevatron experiments. Here we use the HERA definition. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is defined as the fraction of the jet’s energy contained in a cone of radius $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ centred on its direction. We therefore have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, meaning that the jet’s energy is all contained within a cone of radius $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ (the relation is not exact because in neither the cluster algorithm, nor even the cone algorithm after the merging/splitting step, is the edge of a jet an exact geometric cone). Narrower jets are characterized by larger values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The jet shape is sometimes discussed in terms of the energy fractions in concentric angular annuli, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{d}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{d}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$.
It has long been known that parton shower Monte Carlo programs like HERWIG predict considerably narrower jets than are observed in the Tevatron data, for example Fig. 9 (Fig. 1 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$).
Since jet shapes in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ annihilation were known to be well described, even when separated out into quark- and gluon-jet samples$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$, possible explanations focused on the two main new ingredients in hadron collisions, initial-state radiation (ISR) and the underlying event. The ISR model is well tested by the Tevatron experiments’ measurements of colour coherence effects in two-jet events, which HERWIG describes well$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$. This leaves underlying event effects as the most likely culprit.
This hypothesis can be tested at HERA, since resolved photoproduction should have an underlying event and direct photoproduction and DIS should not. A great deal of excellent data have appeared in the last couple of years$`^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}}`$, from which we choose just one example, dijet events in DIS from H1, Fig. 10 (Fig. 5 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$).
It can be seen that HERWIG’s prediction is again too narrow, although by much less than at the Tevatron. Since there should be no underlying event correction, this clearly needs to be understood in more detail. It is worth noting that in this kinematic region the hadronization corrections are huge, as shown in Fig. 11 (Fig. 6a of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) for the LEPTO generator.
As mentioned earlier, DIS in the HERA lab frame has a special role in jet physics because the recoiling electron acts kinematically like a parton but is not coloured. This has enabled the first NLO calculation of the jet shape to be made$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$.
A comparison with ZEUS data is shown in Fig. 12 (Fig. 4a of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$). For precisely the reasons mentioned earlier this is extremely sensitive to the details of the jet algorithm. Since the iterative algorithm used by the ZEUS experiment is not infrared safe, it cannot be used in the NLO calculation, so this comparison can only be taken as indicative. To supposedly take account of this, the authors of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$ applied an additional cut in the calculation that was not applied to the data and chose a very small scale for the running coupling. Bearing this in mind, and the fact that the hadronization corrections shown in Fig. 11 (Fig. 6a of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) are about a factor of two, the claimed good agreement shown in Fig. 12 (Fig. 4a of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) must be seen as coincidental.
### 3.2 Subjet studies
The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ algorithm naturally suggests a new way of analysing the internal structure of jets that is much closer to the partonic picture of how that structure arises. After identifying a jet of a given $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$, we rerun the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ algorithm, but only on those particles that were assigned to this jet. We stop clustering when all values of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ satisfy $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, namely when all internal relative transverse momenta are greater than $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$. Analysing the jets in this way is extremely similar to analysing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ annihilation events at $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ and the same value of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}`$ and in fact it can be shown$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$ that the leading logarithms are identical. Even the ISR of gluons into the jet, which contributes at next-to-leading logarithmic accuracy, can be summed to all orders and results are shown in Fig. 13 (Fig. 1 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) for the average number of subjets.
The resummation can be seen to be extremely important for small $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}`$, while the initial-state resummation is only a relatively small correction.
The subjet multiplicity was studied in a preliminary way by D0 in $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$ and compared with the parton shower Monte Carlo programs. The results are shown in Fig. 14 (Fig. 7 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$).
I still find this figure amazing: on the left-hand side we are studying 250 GeV jets at a scale of less than 1 GeV where the hadronization corrections are huge and, at least in HERWIG, the description is perfect. It is possible that the over-production of subjets in ISAJET is related to the lack of colour coherence and angular ordering and that the deficit in PYTHIA, which starts at smaller $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}`$ is due to an over-estimate of the amount of ‘string drag’ pulling soft hadrons out of the jet, although these effects have not been studied in detail.
Subjet properties have also been studied for separate quark and gluon jet samples using an extremely neat statistical separation$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$. On the assumption that quark and gluon jet properties are each independent of $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}`$ for fixed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$, the fact that the flavour mix of jets varies strongly and that this variation is well predicted by perturbative QCD can be used to measure their individual properties without the need for an event-by-event tag.
The results for the subjet rates are shown in Fig. 15 (Fig. 3 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$), where it can be seen that as expected gluon jets contain a lot more activity than quark jets. The distributions are again well described by HERWIG.
#### All-orders resummation for subjet rates
Recently the first calculation of subjet rates in hadron collisions was performed$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$. In general the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$-subjet rate contains terms like $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{log}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}`$ at all orders of perturbation theory $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Together with the next-to-leading logarithms ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{log}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}`$) these can be summed to all orders using the same trick as was used for the subjet multiplicity in $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$, which is illustrated in Fig. 16.
The evolution of the final-state jet is process-independent and can be summed to next-to-leading log accuracy using the well-known formulæ from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ annihilation$`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$. However, the next-to-leading logs also receive a contribution from soft initial-state radiation that happens to be close enough to the jet to be combined with it. As the probability of such emission depends on the full details of the hard process through the kinematics, identities and colour-connection of all participating partons, it seems unlikely that this could be resummed analytically. However, by carefully combining the analytical result with a numerical integration of the exact matrix element to produce one additional gluon, it is possible not only to sum these logs to all orders, but also to automatically exactly reproduce the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ contribution to the one- and two-subjet rates.
Examples of the results are shown in Figs. 17 and 18 (Figs. 9, 10 and 12 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$). The first thing to note is that the general forms look very reminiscent of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ annihilation. It is possible to separate out the contribution from quark and gluon jets and test the hypothesis used by D0 that these are independent of the centre-of-mass collision energy.
As can be seen in Fig. 17 (Figs. 9 and 10 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) this is the case.
The results at fixed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cut}}$}}`$ shown in Fig. 18 (Fig. 12 of $`^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{?}}$}`$) are certainly reminiscent of D0’s but, owing to the fact mentioned earlier that they use a preclustering algorithm and the theoretical calculation does not, direct comparison is not yet possible.
It is worth noting that in order to extend this calculation to DIS or photoproduction, it is necessary only to put the appropriate matrix element into the box marked “exact $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$” in Fig. 16. All the analytically-summed contributions are then identical.
## 4 Conclusion
Precision QCD physics, and the use of jets in precision electroweak physics, requires reliable jet definitions and reliable predictions of the internal structure of jets. We have reviewed a small subset of the advances made in these areas in recent years.
## References
|
warning/0007/math0007071.html
|
ar5iv
|
text
|
# Quantum Invariants of Links and New Quantum Field Models
## 1 Introduction
In 1989 Witten derived knot invariants such as the Jones polynomial from quantum field theory based on the Chern-Simon Lagrangian . Inspired by Witten’s work in this paper we shall derive knot invariants from a gauge model of Quantum electrodynamics (QED) and its nonabelian generalization. From our approach we shall first derive the Jones polynomial and then we derive new knot and link invariants which extend the Jones polynomial and can give a complete classification of knots and links.
This paper is organized as follows. In section 2 we give a brief description of a gauge model of QED and its nonabelian extention. In this paper we shall consider a nonabelian extension with a $`SU(2)`$ gauge symmetry. With this quantum field model we introduce the partition function and the correlation of a Wilson loop which will be a knot invariant of the trivial knot (also called the unknot). This correlation of Wilson loop will later be generalized to be knot invariants of nontrivial knots and links. To investigate the properties of these partition function and correlations in section 4 we derive a chiral symmetry from the gauge transformation of this new quantum field model. From this chiral symmetry in section 5, section 6 we derive a conformal field theory which contains topics such as the affine Kac-Moody algebra and the Knizhnik-Zamolodchikov equation. This KZ equation is an equation of correlations from which in section 7 we can derive the skein relation of the Jones polynomial. A main point of our theory on the KZ equation is that we can derive two KZ equations which are dual to each other. From these two KZ equations we derive a quantum group structure for the $`W`$ matrices from which a Wilson loop is formed. Then from the correlation of these $`W`$ matrices in section 8 and section 9 we derive new knot invariants which extend the Jones polynomial and gives a complete classification of knots. In section 10 we extend the new invariants to the case of links and we compute these new link invariants for some examples of links. Then in section 11 with the new knot invariants we give a classification table of knots which is formed by using the power index $`m`$ which comes from these new knot invariants of the form $`TrR^mW(z,z)`$ where $`W(z,z)`$ denotes a Wilson loop and $`R`$ is the braiding matrix for the quantum group structure and is the monodromy of the two KZ equations.
## 2 New gauge Model of QED and Nonabelian Extensions
To begin our derivation of knot and link invariants let us first describe a quantum field model. Similar to the Wiener measure for the Brownian motion which is constructed from the integral $`_{t_0}^{t_1}\left(\frac{dx}{dt}\right)^2𝑑t`$ we construct a measure for QED from the following energy integral:
$$\frac{1}{2}_{s_0}^{s_1}[\frac{1}{2}\left(\frac{dA_1}{ds}\frac{dA_2}{ds}\right)^2+\underset{i=1}{\overset{2}{}}\left(\frac{dz}{ds}ieA_iz\right)^{}\left(\frac{dz}{ds}ieA_iz\right)]𝑑s$$
(1)
where $`s`$ denotes the proper time in relativity; $`e`$ denotes the electric charge and the complex variable $`z`$, real variables $`A_1`$, $`A_2`$ represent one electron and two photons respectively. By extending $`ds`$ to $`(ih+\beta )ds`$ we get a quantum theory of QED where $`h>0`$ denotes the Planck constant and $`\beta >0`$ is a constant related to absolute temperature.
The integral (1) has the following gauge symmetry:
$$z^{}(s)=z(s)e^{iea(s)},A_i^{}(s)=A_i(s)+\frac{da}{ds}i=1,2$$
(2)
where $`a(s)`$ is a real valued function.
We remark that a main feature of (1) is that it is not formulated with the four-dimensional space-time but is formulated with the one dimensional proper time. We refer to for the physical motivation of this new QED theory.
We can generalize the above QED model with $`U(1)`$ gauge symmetry to QCD type models with nonabelian gauge symmetry. As an illustration let us consider $`SU(2)`$ gauge symmetry. Similar to (1) we consider the following energy integral:
$$L:=\frac{1}{2}_{s_0}^{s_1}[\frac{1}{2}tr(D_1A_2D_2A_1)^{}(D_1A_2D_2A_1)+(D_1Z)^{}(D_1Z)+(D_2Z)^{}(D_2Z)]𝑑s$$
(3)
where $`Z=(z_1,z_2)^T`$ is a two dimensional complex vector; $`A_j=_{k=1}^3A_j^kt^k`$ $`(j=1,2)`$ where $`A_j^k`$ denotes a real component of a gauge field $`A^k`$; $`t^k`$ denotes a generator of $`SU(2)`$; and $`D_j=\frac{d}{ds}igA_j`$ $`(j=1,2)`$ where $`g`$ denotes the charge of interaction. From (3) we can develop a QCD type model as similar to that for the QED model. We have that (3) is invariant under the following gauge transformation:
$$\begin{array}{cc}\hfill Z^{}(s)& =U(a(s))Z(s)\hfill \\ \hfill A_j^{}(s)& =U(a(s))A_j(s)U^1(a(s))+U(a(s))\frac{dU^1(a(s))}{ds},j=1,2\hfill \end{array}$$
(4)
where $`U(a(s))=e^{a(s)}`$ and $`a(s)=_ka^k(s)t^k`$. We shall mainly consider the case that $`a`$ is a function of the form $`a(s)=\omega _1(r(s))`$ where $`\omega _1`$ and $`r`$ are analytic functions.
## 3 Knot Invariants
Since (3) is not formulated with the space-time, as analogous to the approach of Witten on knot invariants the following partition function will be shown to be a topological invariant for knots:
$$TrW_R:=DA_1DA_2DZe^LTrW_R(C)$$
(5)
where
$$W_R(C):=W(r_0,r_1):=Pe^{_CA_i𝑑x^i}$$
(6)
which may be called a Wilson loop as analogous to the usual Wilson loop where $`C`$ denotes a closed curve of the following form
$$C(s)=(x^1(r(s)),x^2(r(s))),s_0ss_1$$
(7)
where $`r`$ is an analytic function such that $`r_0:=r(s_0)=r(s_1):=r_1`$. This closed curve $`C`$ is in a two dimensional phase plane $`(x^1,x^2)`$ which is dual to $`(A_1,A_2)`$. We let this closed curve $`C`$ represents the projection of a knot in this two dimensional space. As usual the notation $`P`$ in the definition of $`W_R(C)`$ denotes a path-ordered product and $`R`$ denotes a representation of $`SU(2)`$ .
We remark that we also extend the definition of $`W_R(C)`$ to the case that $`C`$ is not a closed curve with $`r_0r_1`$.
We shall show that (5) is a topological invariant for a trivial knot. This means that in (5) the closed curve $`C`$ represents a trivial knot. We shall extend (5) to let $`C`$ represent nontrivial knot.
Our aim is to compute the above knot invariant and its generalization to knot invariants of nontrivial knots which will be defined.
## 4 Chiral Symmetry
For a given curve $`C(s)=(x^1(r(s)),x^2(r(s))),s_0ss_1`$ which may not be a closed curve we define $`W(r_0,r_1)`$ by (6) where $`r_0=r(s_0)`$ and $`r_1=r(s_1)`$. Then under an analytic gauge transformation we have the following chiral symmetry:
$$W(r_0,r_1)U(\omega (r_1))W(r_0,r_1)U^1(\omega (r_0))$$
(8)
where $`\omega `$ denotes an analytic function. This chiral symmetry is analogous to the chiral symmetry of the usual nonabelian guage theory where $`U`$ denotes an element of $`SU(2)`$ . We may extend (9) by extending $`r`$ to complex variable $`z`$ to have the following chiral symmetry:
$$W(z_0,z_1)U(\omega (z_1))W(z_0,z_1)U^1(\omega (z_0))$$
(9)
This analytic continuation corresponds to the complex transformation $`s(ih+\beta )s`$ for describing quantum physics.
From this chiral symmetry we have the following formulas for the variations $`\delta _\omega W`$ and $`\delta _\omega ^{}W`$ with respect to the chiral symmetry:
$$\delta _\omega W(z,z^{})=W(z,z^{})\omega (z)$$
(10)
and
$$\delta _\omega ^{}W(z,z^{})=\omega ^{}(z^{})W(z,z^{})$$
(11)
where $`z`$ and $`z^{}`$ are independent variables and $`\omega ^{}(z^{})=\omega (z)`$ when $`z^{}=z`$. In (10) the variation is with respect to the $`z`$ variable while in (11) the variation is with respect to the $`z^{}`$ variable. This two-side-variations is possible when $`zz^{}`$.
## 5 Affine Kac-Moody Algebra
Let us define
$$J(z):=kW^1(z,z^{})_zW(z,z^{})$$
(12)
where $`k>0`$ is a constant. As analogous to the WZW model $`J`$ is a generator of the chiral symmetry for (10).
Let us consider the following correlation
$$W_RA(z):=DA_1DA_2DZe^LW_R(C)A(z)$$
(13)
By taking a gauge transformation on this correlation and by the gauge invariance of (3) we can derive a Ward identity from which we have the following relation:
$$\delta _\omega A(z)=\frac{1}{2\pi i}_z𝑑w\omega (w)J(w)A(z)$$
(14)
where $`\delta _\omega A`$ denotes the variation of the field $`A`$ with respect to the chiral symmetry and the closed line integral $``$ is with center $`z`$ and we let the generator $`J`$ be given by (12). We remark that our approach here is analogous to the WZW model in conformal field theory .
From (9) and (12) we have that the variation $`\delta _\omega J`$ of the generator $`J`$ of the chiral symmetry is given by :
$$\delta _\omega J=[J,\omega ]k_z\omega $$
(15)
From (14) and (15) we have that $`J`$ satisfies the following relation of current algebra :
$$J^a(w)J^b(z)=\frac{k\delta _{ab}}{(wz)^2}+\underset{c}{}if_{abc}\frac{J^c(z)}{(wz)}$$
(16)
where we write
$$J(z)=\underset{a}{}J^a(z)t^a=\underset{a}{}\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}J_n^az^{n1}t^a$$
(17)
Then from (16) we can show that $`J_n^a`$ satisfy the following affine Kac-Moody algebra :
$$[J_m^a,J_n^b]=if_{abc}J_{m+n}^c+km\delta _{ab}\delta _{m+n,0}$$
(18)
where the constant $`k`$ is called the central extension or the leval of the Kac-Moody algebra.
Let us consider another generator of the chiral symmetry for (11) given by
$$J^{}(z^{})=k_z^{}W(z,z^{})W^1(z,z^{})$$
(19)
Similar to $`J`$ by the following correlation:
$$A(z^{})W_R:=DA_1DA_2DZA(z^{})W_R(C)e^L$$
(20)
we have the following formula for $`J^{}`$:
$$\delta _\omega ^{}A(z^{})=\frac{1}{2\pi i}_z^{}^{}𝑑wA(z^{})J^{}(w)(\omega ^{})(w)=\frac{1}{2\pi i}_z^{}𝑑wA(z^{})J^{}(w)\omega ^{}(w)$$
(21)
where $`^{}`$ denotes an integral with clockwise direction while $``$ denotes an integral with counterclockwise direction. We remark that this two-side variation from (13) and (20) is important for deriving the two KZ equations which are dual to each other.
Then similar to (15) we also we have
$$\delta _\omega ^{}J^{}=[\omega ^{},J^{}]k_z^{}\omega ^{}$$
(22)
Then from (21) and (22) we can derive the current algebra and the Kac-Moody algebra for $`J^{}`$ which are of the same form of (16) and (18).
## 6 Dual Knizhnik-Zamolodchikov Equation
Let us first consider (10). From (14) and (10) we have
$$J^a(z)W(w,w^{})\frac{t^aW(w,w^{})}{zw}$$
(23)
Let us define an energy-momentum tensor $`T(z)`$ by
$$T(z):=\frac{1}{k+g}\underset{a}{}:J^a(z)J^a(z):$$
(24)
where $`g`$ is the dual Coxter number. In (24) the symbol $`:\mathrm{}:`$ denotes normal ordering. This is the Sugawara construction of energy-momentum tensor where the appearing of $`g`$ is from a renormalization of quantum effect by requiring the operator product expansion of $`T`$ with itself to be of the following form :
$$T(z)T(w)=\frac{c}{2(zw)^4}+\frac{2T(w)}{(zw)^2}+\frac{T(w)}{(zw)}$$
(25)
for some constant $`c=\frac{kd}{k+g}`$ where $`d`$ denotes the dimension of $`SU(2)`$.
Then we have the following $`TW`$ operator product:
$$T(z)W(w,w^{})\frac{\mathrm{\Delta }}{(zw)^2}+\frac{1}{(zw)}L_1W(w,w^{})$$
(26)
where $`L_1W(w,w^{})=_wW(w,w^{})`$ and
$$\mathrm{\Delta }=\frac{_at^at^a}{2(k+g)}=\frac{N^21}{2N(k+N)}$$
(27)
where $`N`$ is for $`SU(N)`$.
From (24) and (26) we have the following equation :
$$L_1W(w,w^{})=\frac{1}{k+g}J_1^aJ_0^aW(w,w^{})$$
(28)
Then form (23) we have
$$J_0^aW(w,w^{})=t^aW(w,w^{})$$
(29)
By (28) and (29) we have
$$_zW(z,z^{})=\frac{1}{k+g}J_1^at^aW(z,z^{})$$
(30)
From this equation and by the $`JW`$ operator product (23) we have the following Knizhnik-Zamolodchikov equation :
$$_{z_i}W(z_1,z_1^{})\mathrm{}W(z_n,z_n^{})=\frac{1}{k+g}\underset{ji}{\overset{n}{}}\frac{_at^at^a}{z_iz_j}W(z_1,z_1^{})\mathrm{}W(z_n,z_n^{})$$
(31)
We remark that in deriving (31) we have used line integral expression of operators with counterclockwise direction .
It is interesting and important that we also have another Knizhnik-Zamolodchikov equation which will be called the dual equation of (31). The derivation of this dual equation is dual to the above derivation in that the line integral for this derivation of dual equation is with clockwise direction in contrast to counterclockwise direction in the above derivation and that the operator products and their corresponing variables are with reverse order to that in the above derivation.
From (11) and (21) we have a $`WJ^{}`$ operator product given by
$$W(w,w^{})J^a(z^{})\frac{W(w,w^{})t^a}{w^{}z^{}}$$
(32)
Similar to the above derivation of the KZ equation from (32) we can then derive the following Knizhnik-Zamolodchikov equation which is dual to (31):
$$_{z_i^{}}W(z_1,z_1^{})\mathrm{}W(z_n,z_n^{})=\frac{1}{k+g}\underset{ji}{\overset{n}{}}W(z_1,z_1^{})\mathrm{}W(z_n,z_n^{})\frac{_at^at^a}{z_j^{}z_i^{}}$$
(33)
## 7 Skein Relation for Jones Polynomial
Following the idea of Witten , if we cut a knot we get two pieces of curves crossing (or not crossing) each other once. This gives two primary fields $`W(z_1,z_2)`$ and $`W(z_3,z_4)`$ where $`W(z_1,z_2)`$ corresponds to a piece of curve with end points parametrized by $`z_1`$ and $`z_2`$ and $`W(z_3,z_4)`$ corresponds to the other piece of curve with end points parametrized by $`z_3`$ and $`z_4`$. Let us write
$$W(z_i,z_j)=W(z_i,z_i^{})W^1(z_j,z_j^{})$$
(34)
for $`i=1,3`$ and $`j=2,4`$ and for some $`z_k^{}`$ with $`z_1^{}=z_2^{}`$ and $`z_3^{}=z_4^{}`$. These two pieces of curves then correspond to the following four-point correlation function:
$$G(z_1,z_2,z_3,z_4):=W(z_1^{},z_1)W^1(z_2^{},z_2)W(z_3,z_3^{})W^1(z_4,z_4^{})$$
(35)
(In the notation $`G(z_1,z_2,z_3,z_4)`$ we have supressed the $`z^{}`$ variables for simplicity). Then we have :
$$G(z_1,z_2,z_3,z_4)=[(z_1z_3)(z_2z_4)]^{2\mathrm{\Delta }}G(x)$$
(36)
where $`\mathrm{\Delta }=\frac{N^21}{2N(N+k)}`$ and $`x=\frac{(z_1z_2)(z_3z_4)}{(z_1z_3)(z_2z_4)}`$ and from the KZ equation $`G(x)`$ satisfies the following equation:
$$\frac{dG}{dx}=[\frac{1}{x}P+\frac{1}{x1}Q]G$$
(37)
where
$$P=\frac{1}{N(N+k)}\left(\begin{array}{cc}N^21& N\\ 0& 1\end{array}\right),Q=\frac{1}{N(N+k)}\left(\begin{array}{cc}1& 0\\ N& N^21\end{array}\right)$$
(38)
This equation has two independent conformal block solutions forming a vector space of dimension 2. Let $`\psi `$ be a vector in this space and let $`B`$ denotes the braid operation. Then following Witten we have
$$a\psi bB\psi +B^2\psi =0$$
(39)
where $`a=detB`$ and $`b=TrB`$. Then following Witten from (39) we can derive the following skein relation for the Jones polynomial:
$$\frac{1}{t}V_L_{}tV_{L_+}=(t^{\frac{1}{2}}\frac{1}{t^{\frac{1}{2}}})V_{L_0}$$
(40)
where $`V_L_{}`$, $`V_{L_+}`$ and $`V_{L_0}`$ are the Jones polynomials for undercrossing, overcrossing and zero crossing respectively.
## 8 New Knot Invariants Extending Jones Polynomial
Let us consider again the correlation $`G(z_1,z_2,z_3,z_4)`$ in (35) which also have the following form:
$$G(z_1,z_2,z_3,z_4)=W(z_1,z_2)W(z_3,z_4)$$
(41)
From it in this section we shall present a method which is different from the above section to derive new knot invariants. These new knot invariants will extend the Jones polynomial and they will be defined by generalizing (5).
We have that $`G`$ satisfies the KZ equation for the variables $`z_1`$, $`z_3`$ and satisfies the dual KZ equation for the variables $`z_2`$ and $`z_4`$. By solving the KZ equation we have that $`G`$ is of the form
$$e^{t\mathrm{log}(z_1z_3)}C_1$$
(42)
where $`t:=\frac{1}{k+g}_at^at^a`$ and $`C_1`$ denotes a constant matrix which is independent of the variable $`z_1z_3`$.
Similarly by solving the dual KZ equation we have that $`G`$ is of the form
$$C_2e^{t\mathrm{log}(z_4z_2)}$$
(43)
where $`C_2`$ denotes a constant matrix which is independent of the variable $`z_4z_2`$.
From (42), (43) and we let $`C_1=Ae^{t\mathrm{log}(z_4z_2)}`$, $`C_2=e^{t\mathrm{log}(z_1z_3)}A`$ where $`A`$ is a constant matrix we have that $`G`$ is given by
$$G(z_1,z_2,z_3,z_4)=e^{t\mathrm{log}(z_1z_3)}Ae^{t\mathrm{log}(z_4z_2)}$$
(44)
Now let $`z_2=z_3`$. Then as $`z_4z_1`$ we have
$$TrG(z_1,z_2,z_2,z_1)=Tre^{i2n\pi t}A=:TrR^{2n}An=0,\pm 1,\pm 2,\mathrm{}$$
(45)
where $`R=e^{i\pi t}`$ is the monodromy of the the KZ equation . We remark that (45) is a multivalued function. From (45) we have the following relation between the partition function $`Z`$ and the matrix $`A`$:
$$A=IZ$$
(46)
where $`I`$ denotes the identity matrix. Now let $`C`$ be a closed curve in the complex plane with initial and final end points $`z_1`$. Then the following correlation function
$$TrW(z_1,z_1)=TrW(z_1,z_2)W(z_2,z_1)$$
(47)
which is the definition (5) defined along the curve $`C`$, with $`W(z_1,z_1)=W(z_1,z_2)W(z_2,z_1)`$, can be regarded as a knot invariant of the trivial knot in the three dimensional space whose porjection in the complex plane is the curve $`C`$. Indeed, from (41) and (45) we can compute (47) which is given by:
$$TrW(z_1,z_1)=ZTrR^{2n}n=0,\pm 1,\pm 2,\mathrm{}$$
(48)
From (48) we see that (47) is independent of the closed curve $`C`$ which represents the projection of a trivial knot and thus can be regarded as a knot invariant for the trivial knot.
In the following let us extend the definition (47) to knot invariants for nontrivial knots.
Since $`R`$ is the monodromy of the KZ equation, we have a branch cut such that
$$W(z_3,w)W(w,z_2)W(z_1,w)W(w,z_4)=RW(z_1,w)W(w,z_2)W(z_3,w)W(w,z_4)$$
(49)
where $`z_1`$ and $`z_3`$ denote two points on a closed curve such that along the direction of the curve the point $`z_1`$ is before the point $`z_3`$. From (49) we have
$$W(z_3,w)W(w,z_2)W(z_1,w)W(w,z_4)=RW(z_1,w)W(w,z_2)W(z_2,w)W(w,z_4)$$
(50)
Similarly for the dual KZ equation we have
$$W(z_1,w)W(w,z_4)W(z_3,w)W(w,z_2)=W(z_1,w)W(w,z_2)W(z_3,w)W(w,z_4)R^1$$
(51)
and
$$W(z_1,w)W(w,z_4)W(z_3,w)W(w,z_2)=W(z_1,w)W(w,z_2)W(z_3,w)W(w,z_4)R^1$$
(52)
where $`z_2`$ before $`z_4`$. From (50) and (52) we have
$$W(z_3,z_4)W(z_1,z_2)=RW(z_1,z_2)W(z_3,z_4)R^1$$
(53)
where $`z_1`$ and $`z_2`$ denote the end points of a curve which is before a curve with end points $`z_3`$ and $`z_4`$. From (53) we see that the algebraic structure of these $`W`$ matrices is analogous to the quasi-triangular quantum group . Now we let $`W(z_i,z_j)`$ represent a piece of curve with initial end point $`z_i`$ and final end point $`z_j`$. Then we let
$$W(z_1,z_2)W(z_3,z_4)$$
(54)
represent two pieces of uncrossing curve. Then by interchanging $`z_1`$ and $`z_3`$ we have
$$W(z_3,w)W(w,z_2)W(z_1,w)W(w,z_4)$$
(55)
represent the curve specified by $`W(z_1,z_2)`$ upcrossing the curve specified by $`W(z_3,z_4)`$ at $`z`$. Similarly by interchanging $`z_2`$ and $`z_4`$ we have
$$W(z_1,w)W(w,z_4)W(z_3,w)W(w,z_2)$$
(56)
represent the curve specified by $`W(z_1,z_2)`$ undercrossing the curve specified by $`W(z_3,z_4)`$ at $`z`$.
Now for a closed curve we may cut it into a sum of parts which are formed by two pieces of curve crossing or not crossing each other. Each of these parts is represented by (54), (55) or (56). Then we may define a correlation for a knot whose projection is this closed curve by the following form:
$$Tr\mathrm{}W(z_3,z)W(z,z_2)W(z_1,z)W(z,z_4)\mathrm{}$$
(57)
where we use (55) as an example to represent the state of the two pieces of curve specified by $`W(z_1,z_2)`$ and $`W(z_3,z_4)`$. The $`\mathrm{}`$ means multiplications of a sequence of parts represented by (54), (55) or (56) according to the state of each part. The order of the sequence in (57) follows the order of the parts given by the direction of the knot.
We shall show that (57) is a knot invariant for a given knot. In the following let us consider some examples to illustrate the way to define (57) and to show that (57) is a knot invariant. We shall also derive the three Reidemeister moves for the equivalence of knots.
Let us first consider a knot in Fig. 1. For this knot we have that (57) is given by
$$TrW(z_2,w)W(w,z_2)W(z_1,w)W(w,z_1)$$
(58)
where the product of $`W`$ is from the definition (55). In applying (55) we let $`z_1`$ as the starting and the final point. We remark that the $`W`$ matrices must be put together to follow the definition (55) and they are not separated to follow the direction of the knot.
Then we have that (58) is equal to
$$\begin{array}{cc}& TrW(w,z_2)W(z_1,w)W(w,z_1)W(z_2,w)\hfill \\ \hfill =& TrRW(z_1,w)W(w,z_2)R^1RW(z_2,w)W(w,z_1)R^1\hfill \\ \hfill =& TrW(z_1,z_2)W(z_2,z_1)\hfill \\ \hfill =& TrW(z_1,z_1)\hfill \end{array}$$
(59)
where we have used (53). We see that (59) is just the knot invariant (47) of a trivial knot. Thus the knot in Fig.1 is with the same knot invariant of a trivial knot and this agrees with the fact that this knot is topologically equivalent to a trivial knot.
Then let us derive the Reidemeister move 1. Consider the diagram in Fig.2. We have that by (55) the definition (57) for this diagram is given by:
$$\begin{array}{cc}& TrW(z_2,w)W(w,z_2)W(z_1,w)W(w,z_3)\hfill \\ \hfill =& TrW(z_2,w)RW(z_1,w)W(w,z_2)R^1W(w,z_3)\hfill \\ \hfill =& TrW(z_2,w)RW(z_1,z_2)R^1W(w,z_3)\hfill \\ \hfill =& TrR^1W(w,z_3)W(z_2,w)RW(z_1,z_2)\hfill \\ \hfill =& TrW(z_2,w)W(w,z_3)W(z_1,z_2)\hfill \\ \hfill =& TrW(z_2,z_3)W(z_1,z_2)\hfill \\ \hfill =& TrW(z_1,z_3)\hfill \end{array}$$
(60)
where$`W(z_1,z_3)`$ represent a piece of curve with initial end point $`z_1`$ and final end point $`z_3`$ which has no crossing. When Fig.2 is a part of a knot we can also derive a result similar to (60) which is for the Reidemeister move 1. This shows that the Reidemeister move 1 holds.
Then let us derive Reidemeister move 2. By (55) we have that the definition (57) for the two pieces of curve in Fig.3a is given by
$$TrW(z_5,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_6)W(z_4,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,z_5)$$
(61)
where the two products of $`W`$ separated by the $``$ are for the two crossings in Fig.3a. We have that (61) is equal to
$$\begin{array}{cc}& TrW(z_4,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,z_5)W(z_5,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_2)W(w_2,z_3)W(z_2,w_2)RW(w_1,z_2)W(w_2,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_2)RW(z_2,w_2)W(w_2,z_3)W(w_1,z_2)W(w_2,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_2)RW(z_2,z_3)W(w_1,z_2)W(w_2,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_2)W(w_1,z_2)W(z_2,z_3)RW(w_2,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_2)W(w_1,z_3)RW(w_2,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrRW(w_1,z_3)W(z_4,w_2)W(w_2,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrRW(w_1,z_3)W(z_4,w_1)R^1W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_1)W(w_1,z_3)W(z_1,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_4,w_1)RW(z_1,w_1)W(w_1,z_3)R^1W(w_1,z_6)\hfill \\ \hfill =& TrW(z_1,z_3)R^1W(w_1,z_6)W(z_4,w_1)R\hfill \\ \hfill =& TrW(z_1,z_3)W(z_4,w_1)W(w_1,z_6)\hfill \\ \hfill =& TrW(z_1,z_3)W(z_4,z_6)\hfill \end{array}$$
(62)
where we have used (53). This shows that the diagram in Fig.3a is equivalent to two uncrossing curves. When Fig.3a is a part of a knot we can also derive a result similar to (62) for the Reidemeister move 2. This shows that the Reidemeister move 2 holds.
As an illustration let us consider the knot in Fig. 3b which is related to the Reidemeister move 2. By (55) we have that the definition (57) for this knot is given by
$$\begin{array}{cc}& TrW(z_3,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,z_4)W(z_4,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_1)\hfill \\ \hfill =& TrRW(z_2,w_2)W(w_2,z_3)W(z_3,w_2)W(w_2,z_4)W(z_4,w_1)W(w_1,z_1)W(z_1,w_1)W(w_1,z_2)R^1\hfill \\ \hfill =& TrW(z_2,w_2)W(w_2,z_3)W(z_3,w_2)W(w_2,z_4)W(z_4,w_1)W(w_1,z_1)W(z_1,w_1)W(w_1,z_2)\hfill \\ \hfill =& TrW(z_2,z_2)\hfill \end{array}$$
(63)
where we let the curve be with $`z_2`$ as the initial and final end point and we have used (50) and (52). This shows that the knot in Fig.3b is with the same knot invariant of a trivial knot. This agrees with the fact that this knot is equivalent to the trivial knot.
Let us then consider a trefoil knot in Fig.4a. By (55) and similar to the above examples we have that the definition (57) for this knot is given by:
$$\begin{array}{cc}& TrW(z_4,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_5)W(z_2,w_2)W(w_2,z_6)\hfill \\ & W(z_5,w_2)W(w_2,z_3)W(z_6,w_3)W(w_3,z_4)W(z_3,w_3)W(w_3,z_1)\hfill \\ \hfill =& TrW(z_4,w_1)RW(z_1,w_1)W(w_1,z_2)R^1W(w_1,z_5)W(z_2,w_2)RW(z_5,w_2)\hfill \\ & W(w_2,z_6)R^1W(w_2,z_3)W(z_6,w_3)RW(z_3,w_3)W(w_3,z_4)R^1W(w_3,z_1)\hfill \\ \hfill =& TrW(z_4,w_1)RW(z_1,z_2)R^1W(w_1,z_5)W(z_2,w_2)RW(z_5,z_6)R^1W(w_2,z_3)\hfill \\ & W(z_6,w_3)RW(z_3,z_4)R^1W(w_3,z_1)\hfill \\ \hfill =& TrW(z_4,w_1)RW(z_1,z_2)W(z_2,w_2)W(w_1,z_5)W(z_5,z_6)R^1W(w_2,z_3)\hfill \\ & W(z_6,w_3)RW(z_3,z_4)R^1W(w_3,z_1)\hfill \\ \hfill =& TrW(z_4,w_1)RW(z_1,w_2)W(w_1,z_6)R^1W(w_2,z_3)\hfill \\ & W(z_6,w_3)RW(z_3,z_4)R^1W(w_3,z_1)\hfill \\ \hfill =& TrW(z_4,w_1)W(w_1,z_6)W(z_1,w_2)W(w_2,z_3)\hfill \\ & W(z_6,w_3)RW(z_3,z_4)R^1W(w_3,z_1)\hfill \\ \hfill =& TrW(z_4,z_6)W(z_1,z_3)W(z_6,w_3)RW(z_3,z_4)R^1W(w_3,z_1)\hfill \\ \hfill =& TrR^1W(w_3,z_1)W(z_4,z_6)W(z_1,z_3)W(z_6,w_3)RW(z_3,z_4)\hfill \\ \hfill =& TrW(z_4,z_6)W(w_3,z_1)R^1W(z_1,z_3)W(z_6,w_3)RW(z_3,z_4)\hfill \\ \hfill =& TrRW(z_3,z_6)W(w_3,z_1)R^1W(z_1,z_3)W(z_6,w_3)\hfill \\ \hfill =& TrW(w_3,z_1)W(z_3,z_6)W(z_1,z_3)W(z_6,w_3)\hfill \\ \hfill =& TrW(z_6,z_1)W(z_3,z_6)W(z_1,z_3)\hfill \end{array}$$
(64)
where we have repeatly used (53). Then we have that (64) is equal to:
$$\begin{array}{cc}& TrW(z_6,w_3)W(w_3,z_1)W(z_3,w_3)W(w_3,z_6)W(z_1,z_3)\hfill \\ \hfill =& TrRW(z_3,w_3)W(w_3,z_1)W(z_6,w_3)W(w_3,z_6)W(z_1,z_3)\hfill \\ \hfill =& TrRW(z_3,w_3)RW(z_6,w_3)W(w_3,z_1)R^1W(w_3,z_6)W(z_1,z_3)\hfill \\ \hfill =& TrW(z_3,w_3)RW(z_6,z_1)R^1W(w_3,z_6)W(z_1,z_3)R\hfill \\ \hfill =& TrW(z_3,w_3)RW(z_6,z_3)W(w_3,z_6)\hfill \\ \hfill =& TrW(w_3,z_6)W(z_3,w_3)RW(z_6,z_3)\hfill \\ \hfill =& TrRW(z_3,w_3)W(w_3,z_6)W(z_6,z_3)\hfill \\ \hfill =& TrRW(z_3,z_3)\hfill \end{array}$$
(65)
where we have used (50) and (53). We see that (65) is a knot invariant for the trefoil knot in Fig.4a.
Then let us consider the trefoil knot in Fig. 4b which is the mirror image of the trefoil knot in Fig.4a. The definition (57) for this knot is given by:
$$\begin{array}{cc}& TrW(z_1,w_1)W(w_1,z_5)W(z_4,w_1)W(w_1,z_2)W(z_5,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,z_6)\hfill \\ & W(z_3,w_3)W(w_3,z_1)W(z_6,w_3)W(w_3,z_4)\hfill \\ \hfill =& TrW(z_5,z_1)W(z_2,z_5)W(z_1,z_2)\hfill \end{array}$$
(66)
where similar to (64) we have repeatly used (53). Then we have that (66) is equal to:
$$\begin{array}{cc}& TrW(z_5,z_1)W(z_2,w_1)W(w_1,z_5)W(z_1,w_1)W(w_1,z_2)\hfill \\ \hfill =& TrW(z_5,z_1)W(z_2,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_5)R^1\hfill \\ \hfill =& TrW(z_5,z_1)W(z_2,w_1)RW(z_1,w_1)W(w_1,z_2)R^1W(w_1,z_5)R^1\hfill \\ \hfill =& TrR^1W(z_5,z_1)W(z_2,w_1)RW(z_1,z_2)R^1W(w_1,z_5)\hfill \\ \hfill =& TrW(z_2,w_1)W(z_5,z_2)R^1W(w_1,z_5)\hfill \\ \hfill =& TrW(z_5,z_2)R^1W(w_1,z_5)W(z_2,w_1)\hfill \\ \hfill =& TrW(z_5,z_2)W(z_2,w_1)W(w_1,z_5)R^1\hfill \\ \hfill =& TrW(z_5,z_5)R^1\hfill \end{array}$$
(67)
where we have used (52) and (53). We see that this is a knot invariant for the trefoil knot in Fig.4b. we notice that the knot invariants for the two trefoil knots are different. This shows that these two trefoil knots are not topologically equivalent.
Then let us derive the Reidemeister move 3. We have that the definition (57) for the diagram in Fig.5a is given by
$$\begin{array}{cc}& TrW(z_7,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_8)W(z_8,w_3)W(w_3,z_6)W(z_5,w_3)W(w_3,z_9)\hfill \\ & W(z_4,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,z_5)\hfill \end{array}$$
(68)
where we let the the curve with end points $`z_7`$, $`z_9`$ starts first, then the curve with end points $`z_1`$, $`z_3`$ starts second. Similar to the derivation of the above invariants by (50), (52), (53) we have that (68) is equivalent to the following correlation:
$$TrW(z_7,z_9)W(z_1,z_3)W(z_5,z_6)R^1W(z_4,z_5)$$
(69)
On the other hand the definition (57) for the diagram Fig.5b is given by
$$\begin{array}{cc}& TrW(z_7,w_1)W(w_1,z_5)W(z_4,w_1)W(w_1,z_8)W(z_8,w_3)W(w_3,z_3)W(z_2,w_3)W(w_3,z_9)\hfill \\ & W(z_5,w_2)W(w_2,z_2)W(z_1,w_2)W(w_2,z_6)\hfill \end{array}$$
(70)
where the ordering of the three curves is the same as that in Fig.5a. By (50), (52), (53) we have that (70) is equal to
$$TrW(z_7,z_9)W(z_4,z_5)W(z_1,z_3)R^1W(z_5,z_6)$$
(71)
Then by reversing the ordering of the curves with end points $`z_1`$, $`z_3`$ and with end points $`z_4`$, $`z_6`$ respectively we have that both (69) and (70) are equal to
$$TrW(z_7,z_9)W(z_4,z_6)W(z_1,z_3)R^1$$
(72)
This shows that Fig.5a is equivalent to Fig.5b and this gives the Reidemeister move 3.
From the above examples we see that knots can be classified by the number $`m`$ of product of $`R`$ and $`R^1`$ matrices. More calculations and examples of the above knot invariants will be given elsewhere.
## 9 Classification of Knots and Links
With the knot invariants in the above section we can now give a classification of knots. Let $`K_1`$ and $`K_2`$ be two knots. Since the two W-products $`W(z_3,w)W(w,z_2)W(z_1,w)W(w,z_4)`$ and $`W(z_1,z_2)W(z_3,z_4)`$ faithfully represent two oriented pieces of curves which are crossing or not crossing to each other we have that from the orientation of a knot $`K`$ we have that the product of sequence of W-matrices (We may call this product as a generalized Wilson loop) which are formed according to the orientation of $`K`$ in the correlation (57) of defining an invariant of the knot $`K`$ faithfully represents the knot $`K`$. From this we have that $`K_1`$ and $`K_2`$ are topological equivalent if and only if their generalized Wilson loops can be transformed to each other by using (50), (52) and (53). We note that in the above section we can derive the Reidemeister moves by using (50), (52) and (53). Thus this equivalence of $`K_1`$ and $`K_2`$ represented by their generalized Wilson loops agrees with the fact that $`K_1`$ and $`K_2`$ are topologically equivalent if and only if they can be transformed to each other by the Reidemeister moves.
Then since each knot can be changed to a trivial knot by applying braiding operation which is equivalent to (50), (52) and (53) it follows that these new knot invariants can be equivalently transformed to the form $`TrR^mW(z_1,z_1)`$ where $`m`$ is an integer and $`TrW(z_1,z_1)`$ is the knot invariant for the trivial knot. This form can also be shown by a direct computation which is similar to the computation of the invariant of the trivial knot. We notice that since this new knot invariant is of the form $`TrR^mW(z,z)`$ we have that knots can be completely classified by the power index $`m`$ of $`R`$.
Similar to the case of knots we have that the generalized Wilson loop for a link can faithfully represents this link and that the correlation (57) of the generalized Wilson loop of this link is an invariant which can completely classifies links. For the case of link as similar to the case of knot the ordering of the crossings $`W(z_3,w)W(w,z_2)W(z_1,w)W(w,z_4)`$ can be given by following the orientation of each component of a link. When a component of a link has been traced for one loop the crossings $`W(z_3,w)W(w,z_2)W(z_1,w)W(w,z_4)`$ related to this component can be ordered and the pieces $`W(w_i,z_j)`$ related to these crossings (which may come from other components of the link) have also been required being ordered. In the following section we give some examples to illustrate the formation and computations of this new link invariant.
## 10 Examples of New Link Invariants
Let us first consider the link in Fig.6a. We may let the two knots of this link be with $`z_1`$ and $`z_4`$ as the initial and final end point respectively. We let the ordering of these two knots be such that when the $`z`$ parameter goes one loop on one knot then the $`z`$ parameter for another knot also goes one loop. The correlation (57) for this link is given by:
$$TrW(z_3,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_4)W(z_4,w_2)W(w_2,z_1)W(z_2,w_2)W(w_2,z_3)$$
(73)
We let the ordering of the $`W`$-matrices in (73) be such that $`W(z_1,z_2)`$ and $`W(z_4,z_3)`$ start first. Then next $`W(z_2,z_1)`$ and $`W(z_3,z_4)`$ follows. Form this ordering we have that (73) is equal to:
$$\begin{array}{cc}& TrRW(z_1,w_1)W(w_1,z_2)W(z_3,w_1)W(w_1,z_4)W(z_4,w_2)W(w_2,z_3)W(z_2,w_2)W(w_2,z_1)R^1\hfill \\ \hfill =& TrW(z_1,z_2)W(z_3,z_4)W(z_4,z_3)W(z_2,z_1)\hfill \\ \hfill =& TrW(z_2,z_2)W(z_3,z_3)\hfill \end{array}$$
(74)
where we have used (50) and (52). Since by definition (57) we have that $`TrW(z_2,z_2)W(z_3,z_3)`$ is the knot invariant for two unlinking trivial knots, equation (74) shows that the link in Fig.6a is topologically equivalent to two unlinking trivial knots. Similarly we can show that the link in Fig.6b is topologically equivalent to two unlinking trivial knots.
Let us then consider the Hopf link in Fig.7a. The correlation (57) for this link is given by:
$$TrW(z_3,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_4)W(z_2,w_2)W(w_2,z_3)W(z_4,w_2)W(w_2,z_1)$$
(75)
The ordering of the $`W`$-matrices in (75) is such that $`W(z_1,z_2)`$ starts first and $`W(z_3,z_4)`$ follows it. Then next we let $`W(z_2,z_1)`$ starts first and $`W(z_4,z_3)`$ follows it. The ordering is such that when the $`z`$ parameter goes one loop in one knot of the link we have that the $`z`$ parameter also goes one loop on the other knot. From the ordering we have that (75) is equal to:
$$\begin{array}{cc}& TrRW(z_1,w_1)W(w_1,z_2)W(z_3,w_1)W(w_1,z_4)\hfill \\ & W(z_2,w_2)W(w_2,z_1)W(z_4,w_2)W(w_2,z_3)R^1\hfill \\ \hfill =& TrW(z_1,z_2)W(z_3,z_4)W(z_2,z_1)W(z_4,z_3)\hfill \end{array}$$
(76)
Then let us consider the following correlation:
$$TrR^2W(z_3,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_4)W(z_2,w_2)W(w_2,z_3)W(z_4,w_2)W(w_2,z_1)$$
(77)
We let the ordering of the $`W`$-matrices in (77) be such that $`W(z_1,z_2)`$ starts first and $`W(z_4,z_3)`$ follows it. Then next $`W(z_2,z_1)`$ starts first and $`W(z_3,z_4)`$ follows it. From the ordering we have that (77) is equal to:
$$\begin{array}{cc}& TrR^2RW(z_1,w_1)W(w_1,z_2)W(z_3,w_1)W(w_1,z_4)W(z_2,w_2)W(w_2,z_1)W(z_4,w_2)W(w_2,z_3)R\hfill \\ \hfill =& TrW(z_1,z_2)W(z_3,z_4)W(z_2,z_1)W(z_4,z_3)\hfill \end{array}$$
(78)
On the other hand from the ordering of (77) we have that (77) is equal to:
$$\begin{array}{cc}& TrR^2W(z_3,w_1)RW(z_1,w_1)W(w_1,z_2)R^1\hfill \\ & W(w_1,z_4)W(z_2,w_2)RW(z_4,w_2)W(w_2,z_3)R^1W(w_2,z_1)\hfill \\ \hfill =& TrR^2W(z_3,w_1)RW(z_1,z_2)R^1W(w_1,z_4)W(z_2,w_2)RW(z_4,z_3)R^1W(w_2,z_1)\hfill \\ \hfill =& TrR^2W(z_3,w_1)RW(z_1,z_2)W(z_2,w_2)W(w_1,z_4)W(z_4,z_3)R^1W(w_2,z_1)\hfill \\ \hfill =& TrR^2W(z_3,w_1)RW(z_1,w_2)W(w_1,z_3)R^1W(w_2,z_1)\hfill \\ \hfill =& TrR^2W(z_3,w_1)W(w_1,z_3)W(z_1,w_2)W(w_2,z_1)\hfill \\ \hfill =& TrR^2W(z_3,z_3)W(z_1,z_1)\hfill \end{array}$$
(79)
where we have repeatly used (53). From (76), (78) and (79) we have that the knot invariant for the Hopf link in Fig.7a is given by:
$$TrR^2W(z_3,z_3)W(z_1,z_1)$$
(80)
Then let us consider the Hopf link in Fig.7b. The correlation for this link is given by
$$TrW(z_4,w_1)W(w_1,z_2)W(z_1,w_1)W(w_1,z_3)W(z_2,w_2)W(w_2,z_4)W(z_3,w_2)W(w_2,z_1)$$
(81)
By a derivation which is dual to the above derivation for the Hopf link in Fig.7a we have that (81) is equal to
$$TrR^2W(z_4,z_4)W(z_1,z_1)$$
(82)
We see that the invariants for the above two Hopf links are different. This agrees with the fact that these two links are not topologically equivalent.
We can extend the above computations to other links. As examples let us consider the linking of two trivial knot with linking number $`n`$. Similar to the above computations we have that this link which analogous to the Hopf link in Fig.7a is with an invariant equals to $`TrR^{2n}W(z_4,z_4)W(z_1,z_1)`$. Also for this link which analogous to to the Hopf link in Fig.7b is with an invariant equals to $`TrR^{2n}W(z_4,z_4)W(z_1,z_1)`$.
More calculations and examples of these new link invariants will be given elsewhere.
## 11 A Classification Table of Knots
In this section let us give some arguments to determine the new knot invariants of prime knots and nonprime knots. We have shown that this new invariant of each knot is of the form $`TrR^mW(z_1,z_1)`$ where $`m`$ is an integer. This power index $`m`$ can be regarded as a measure of the complexity of a knot. Let us determine $`m`$ for prime and nonprime knots. We need only to determine $`m`$ for knots with positive $`m`$ since the corresponding mirror image will have negative $`m`$ if the mirror image is not equivalent to the corresponding knot. We have shown that the invariants for links of two trivial knots with linking number $`n`$ as in Fig.7a are with the product of $`R`$ of the form $`R^{2n}`$ with $`m=2n`$. This $`m`$ is an even number. Now if we insert these links into the knot table of prime knots we see that the prime knots must with $`m`$ being an odd number (We may refer to the knot table in . We may have nonprime knots which are not in this knot table having the same $`m`$ as these links. In this case our new link invariants still can distinguish them because the corresponding link invariants are of two-loop form while the corresponding knot invariants are of one-loop form).
Then we expect that for prime knots $`m`$ is an odd prime number. Computations show that for the knot $`\mathrm{𝟑}_\mathrm{𝟏}`$ we have $`m=1`$, for the knot $`\mathrm{𝟒}_\mathrm{𝟏}`$ we have $`m=3`$. Then from some arguments on the effect of $`R`$ we should have that for the knot $`\mathrm{𝟓}_\mathrm{𝟏}`$ we have $`m=5`$ and for the knot $`\mathrm{𝟓}_\mathrm{𝟐}`$ we have $`m=7`$. Then how about the knot $`\mathrm{𝟔}_\mathrm{𝟏}`$? We have that the numbers from $`1`$ to $`8`$ are ocupied by knots and links of two trivial knots. Then $`10`$ is ocupied by the link of two trivial knots with linking number $`5`$. Thus for the knot $`\mathrm{𝟔}_\mathrm{𝟏}`$ we should have $`m=9`$ or $`m=11`$. From some argument on the effect of $`R`$ we should have $`m=11`$ for $`\mathrm{𝟔}_\mathrm{𝟏}`$. Then is there a knot with $`m=9`$?
Let us first consider the granny knot (or the square knot) which is a nonprime knot composed with the knot $`\mathrm{𝟑}_\mathrm{𝟏}`$ and its mirror image. This square knot has $`6`$ crossings and $`4`$ alternating crossings and thus its complexity which is measured by the power index $`m`$ of $`R`$ is less than that of $`\mathrm{𝟓}_\mathrm{𝟏}`$ which is with $`5`$ alternating crossings. Thus this granny knot is with $`m=4`$ ($`m=3`$ has been occupied by $`\mathrm{𝟒}_\mathrm{𝟏}`$). Let us denote this granny knot by $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏}`$ where $``$ denotes the connected sum of two knots such that the resulting total number of alternating crossings is equal to the total number of alternating crossings of the two knots minus $`2`$.
Then let us consider the reef knot which is a nonprime knot composed with two identical knots $`\mathrm{𝟑}_\mathrm{𝟏}`$. This knot has $`6`$ alternating crossings which is equal to the total number of crossings as that of $`\mathrm{𝟔}_\mathrm{𝟏}`$. Since this knot is nonprime its complexity is less than that of $`\mathrm{𝟔}_\mathrm{𝟏}`$ where the complexity may be measured by the power index $`m`$ of $`R`$. Thus this reef knot is with power index $`m`$ less than $`11`$. Then if we also regard the total number of alternating crossings of a knot as a way to measure the complexity of a knot we have that the power index $`m`$ of this granny knot is greater than $`5`$ since $`\mathrm{𝟓}_\mathrm{𝟏}`$ is with $`5`$ alternating crossings and with $`m=5`$. Let us denote this reef knot by $`\mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏}`$ where $`\times `$ denotes the connected sum for two knots such that the resulting total number of alternating crossings is equal to the total number of alternating crossings of the two knots.
Now let us look for knots with $`5`$ or $`6`$ alternating crossings. Let us consider the nonprime knot $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}`$ composed with a knot $`\mathrm{𝟑}_\mathrm{𝟏}`$ and a knot $`\mathrm{𝟒}_\mathrm{𝟏}`$ with $`7`$ crossings and $`5`$ alternating crossings. In this case we have that the power index $`m`$ of $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}`$ should be greater than that of $`\mathrm{𝟓}_\mathrm{𝟏}`$ which is exactly with $`5`$ alternating crossings since $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}`$ in addtion has $`7`$ crossings. Then the power index $`m`$ of $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}`$ should be less than that of $`\mathrm{𝟓}_\mathrm{𝟐}`$ which is also with $`5`$ alternating crossings but these crossings are arranged in a more complicated way which is an effect of $`R^2`$ such that $`\mathrm{𝟓}_\mathrm{𝟐}`$ is with $`m=7`$. Thus $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}`$ is with $`m=6`$.
Then we consider the nonprime knot $`\mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})`$ with $`9`$ crossings and $`5`$ alternating crossings. The power index of this knot should be less than that of $`\mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏}`$ since it is with $`6`$ alternating crossings. Then the power index of $`\mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏}`$ should be greater than that of $`\mathrm{𝟓}_\mathrm{𝟐}`$ since it has in addition $`9`$ crossings which would be enough for a greater power index $`m`$. Then we have that $`\mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})`$ is with $`m=8`$ since it has $`9`$ crossings and $`5`$ alternating crossings and thus is with the same complexity as $`\mathrm{𝟓}_\mathrm{𝟐}`$. Then since $`\mathrm{𝟓}_\mathrm{𝟐}`$ can not have $`m=8`$ we thus have that $`\mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})`$ is with $`m=8`$.
Then we consider the nonprime knot $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟏}`$ with $`8`$ crossings and $`6`$ alternating crossings. The power index $`m`$ of this knot should be greater than that of $`\mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏}`$ which has exactly $`6`$ alternating crossings. Then the power index $`m`$ of this knot should be less than that of $`\mathrm{𝟔}_\mathrm{𝟏}`$ which also has $`6`$ alternating crossings but these crossings are arranged in a more complicated way with an effect of $`R^4`$ from that of $`5_2`$. Thus $`\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟏}`$ is with $`m=10`$ and finally we have that $`\mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏}`$ is with power index $`m=9`$.
Thus for $`m`$ from $`1`$ to $`11`$ we have fill in a suitable knot with power index $`m`$ (except the case $`m=2`$ which is filled in with the Hopf link) such that odd prime numbers are filled with prime knots. In a similar way we may determine the power index $`m`$ of other prime and nonprime knots. We list the results up to $`m=2^5`$ in a form of table.
$$\begin{array}{cccc}& & & \\ \text{Type of Knot}& \text{ Power \hspace{0.17em} Index}m& \text{Type of Knot}& \text{ Power \hspace{0.17em} Index}m\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}& 1& \mathrm{𝟔}_\mathrm{𝟑}& 17\\ & & & \\ \text{Hopf link}& 2& \mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟒}_\mathrm{𝟏}& 18\\ & & & \\ \mathrm{𝟒}_\mathrm{𝟏}& 3& \mathrm{𝟕}_\mathrm{𝟏}& 19\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏}& 4& \mathrm{𝟒}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟏}& 20\\ & & & \\ \mathrm{𝟓}_\mathrm{𝟏}& 5& \mathrm{𝟒}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏})& 21\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}& 6& \mathrm{𝟒}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟐}& 22\\ & & & \\ \mathrm{𝟓}_\mathrm{𝟐}& 7& \mathrm{𝟕}_\mathrm{𝟐}& 23\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏}& 8& \mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏})& 24\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}\times \mathrm{𝟑}_\mathrm{𝟏}& 9& \mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟏})& 25\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟏}& 10& \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟔}_\mathrm{𝟏}& 26\\ & & & \\ \mathrm{𝟔}_\mathrm{𝟏}& 11& \mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟐})& 27\\ & & & \\ \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟓}_\mathrm{𝟐}& 12& \mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟔}_\mathrm{𝟐}& 28\\ & & & \\ \mathrm{𝟔}_\mathrm{𝟐}& 13& \mathrm{𝟕}_\mathrm{𝟑}& 29\\ & & & \\ \mathrm{𝟒}_\mathrm{𝟏}\mathrm{𝟒}_\mathrm{𝟏}& 14& \mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})\mathrm{𝟒}_\mathrm{𝟏}& 30\\ & & & \\ \mathrm{𝟒}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})& 15& \mathrm{𝟕}_\mathrm{𝟒}& 31\\ & & & \\ (\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})& 16& \mathrm{𝟑}_\mathrm{𝟏}(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})(\mathrm{𝟑}_\mathrm{𝟏}\mathrm{𝟑}_\mathrm{𝟏})& 32\end{array}$$
From this table we see that comparable nonprime knots (in a sense from the table) are grouped in each of the intervals between two prime numbers. It is interesting that in each interval nonprime numbers are one-to-one filled with the comparable nonprime knots while prime numbers are filled with prime knots. This grouping property of classification reflects that the new knot invariants (and hence the power index $`m`$) give a classification of knots.
Let us find out some rules for the whole classification table. We have shown that even numbers can not be filled with prime knots. Thus even numbers (except $`2`$) can only be filled with nonprime knots. On the other hand each odd nonprime number is between two even numbers which are power indexes of nonprime knots. Thus the knot corresponding to this odd number is in the same group of these two nonprime knots and is comparable with these two nonprime knots and thus must also be a nonprime knot. From this we have that nonprime numbers are power indexes of nonprime knot. Similarly by this grouping property we have that odd prime numbers are power indexes of prime knots.
It is interesting to note that from the above knot table We see that in the $``$ product the knot $`\mathrm{𝟑}_\mathrm{𝟏}`$ plays the role of the number $`2`$ in the usual multiplication of numbers. Thus the $``$ product (or the connected sum) is a kind of multiplication corresponding to the usual multiplication of numbers. However the general rule for this multiplication is rather complicated. This reflects the fact that the numbers $`m`$ are the power indexes of $`R`$ which are with simple rule for the addition ( and not for multipication). From this generalized multiplication (which is the connected sum) of knots which corresponds to the usual multiplication of the power indexes $`m`$ we also have that prime knots are with odd prime numbers as power indexes and nonprime knots are with nonprime numbers as power indexes.
Thus we have a classification table of knots such that each prime knot corresponds to a prime power index $`m`$ and each nonprime knot corresponds to a nonprime power index $`m`$. More computations to verify this knot table shall be given alsewhere.
## 12 Conclusion
In this paper from a new quantum field model we have derived a conformal field theory and a quantum group structure for generalized Wilson loops from which we can derive the Jones polynomial and new knot and link invariants which extend the Jones polynomial. We show that these new invariants can completely classify knots and links. These new invariants are in terms of the monodromy $`R`$ of two Knizhnik-Zamolochikov equations which are dual to each other. In the case of knots these new invariants can be written in the form $`TrR^mW(z,z)`$ from which we may classify knots with the power index $`m`$ of $`R`$. A classification table of knots can then be formed where prime knots are classified with odd prime numbers $`m`$ and nonprime knots are classified with nonprime numbers $`m`$.
|
warning/0007/hep-ph0007182.html
|
ar5iv
|
text
|
# I Introduction
## I Introduction
QCD vacuum is known to possess a rich structure. A striking example is provided by the existence of (Euclidean) classical solutions – instantons , which are responsible for non-trivial topological properties of the theory . While the influence of instantons on the properties of hadrons and their interactions at low energies have been extensively studied (for a review, see ), the rôle of topological effects in high energy collisions is still an open and fascinating problem .
In electroweak theory, significant interest was excited by the possibility of baryon number non–conservation caused by instantons at the collision energy above $`10`$ TeV . In QCD, the instanton contribution to the structure functions of deep–inelastic scattering was studied in Refs. , while the cross section for gluon–gluon scattering mediated by instantons was considered in Refs. .
Recently, it has been proposed that the existence of semi–classical solutions in QCD can have important consequences also for hadron scattering at high energies, inducing a non–perturbative, and possibly dominant, contribution to the scattering amplitude. This approach was shown to lead to the “soft” pomeron with the intercept $`\alpha _P10.1`$ which has a non–trivial dependence on the strong coupling and the numbers of colors and flavors <sup>1</sup><sup>1</sup>1 Analogous non–perturbative behavior was also established in the scattering of color dipoles at low energy .. The method used in that work was based on low energy theorems of broken scale invariance , and did not assume any specific form of the semi–classical solutions. This approach was based on the low energy theorems for the trace of the QCD energy-momentum tensor taken in the chiral limit of massless quarks
$`\theta _\mu ^\mu ={\displaystyle \frac{bg^2}{32\pi ^2}}F^{a\alpha \beta }F_{\alpha \beta }^a,`$ (1)
where $`b=(11N_c2N_f)/3`$ and $`F^{a\alpha \beta }`$ is the gluonic field strength tensor. The low energy theorems derived in state that
$`i{\displaystyle d^4x(0|T\theta _\mu ^\mu (x)\theta _\nu ^\nu (0)|0)}=4(0|\theta _\mu ^\mu (0)|0).`$ (2)
Comparing Eqs. (1) and (2) one can conclude that the low energy theorem (2) can only be satisfied by a strong gluonic field <sup>2</sup><sup>2</sup>2Unfortunately the applications of low energy theorems of Eq. (2) depends very much on the way one subtracts the perturbative contribution to the correlation functions . Lattice simulations have also encountered similar problems. We thank Al Mueller for clarifying this point to us.
$`A_\mu {\displaystyle \frac{1}{g}}.`$ (3)
The strong field of Eq. (3) is usually associated with the quasi-classical solutions of QCD. This quasi-classical field has been employed in to provide gluonic interactions, which, after being iterated in the t-channel gave rise to a pomeron-like behavior of the corresponding cross sections.
Here we would like to develop the idea proposed in and explore the effects of quasi–classical fields on high energy scattering. Therefore, for the purpose of this paper we will consider a particular type of this classical field: we will assume that the field is given by the instanton solution . We will employ instanton fields, which would allow us to quantify the assumption of the quasi–classical fields being responsible for the pomeron–like behavior of the cross sections. Nevertheless we can not rule out the possibility of some other quasi–classical gluonic fields giving a significant contribution to the processes considered below.
The instanton calculations are better justified in the electroweak theory, where the coupling is small and one can hope that quantum corrections to the classical instanton solution are small. Therefore the calculation that will be presented below can also be repeated for the electroweak interactions, where it would possibly also give a cross section that rises with energy <sup>3</sup><sup>3</sup>3We thank Larry McLerran for bringing our attention to this problem..
Shuryak has argued that the explicit use of instantons could provide a useful way to interpret and extend the results of . Very recently, Shuryak and Zahed analyzed Euclidean scattering induced by the instantons, and analytically continued their results to Minkowski space using the correspondence between Euclidean angle and Minkowski rapidity . Their result is a constant cross section, which does not rise with energy.
The purpose of the present paper is to investigate the effect of the instantons on the dynamics of gluon ladders in high energy scattering. Our approach here is complementary to the one taken in Ref. ; even though it is hindered by significant numerical uncertainties, it provides a deeper insight into the mechanism of non–perturbative effects in high energy scattering. As will become clear later, our treatment of the problem is different from Ref.; while the authors of that paper consider a purely classical contribution to the scattering amplitude, we attempt at evaluating the leading quantum terms arising in perturbation theory built in the background of classical instanton fields.
The proposed mechanism of the pomeron is illustrated in Fig. 1. The pomeron is being constructed by resummation of the ladder diagrams, similarly to the well-known BFKL pomeron . The emission vertices of our pomeron ladder are given by the multiple gluon interaction vertices generated by the instantons . The vertices connect to two gluons in the t-channel and produce several gluons in the t-channel. We sum over all possible numbers of the produced t-channel gluons in each vertex . The t-channel lines in the ladder of Fig. 1 will be taken as the usual gluon lines for most of our paper. In principle one should include the virtual corrections in the t-channel, which may lead to reggeization of the gluons . We will present an estimate of these virtual corrections. However the exact calculation and the answer to the question of whether gluon reggeization happens in our case will be addressed elsewhere . For most of this paper we will work in pure gluodynamics with no quarks. However, at the end of the paper we will present an estimate of the effect of inclusion of light quarks.
The multiple gluon vertices generated by instantons in the pomeron of Fig. 1 will be calculated in Sect. III at the classical level. The vertices include a suppression factor of $`\mathrm{exp}\left(\frac{2\pi }{\alpha _s}\right)`$, which is due to the classical instanton action. If the coupling constant is sufficiently small ($`\alpha _s1`$), then this suppression factor is also small
$`\mathrm{exp}\left({\displaystyle \frac{2\pi }{\alpha _s}}\right)1`$ (4)
and can be used as a parameter justifying the perturbative expansion of Fig. 1. The scale of the running coupling in this case is set by the typical size of the instantons, so that $`\alpha _s=\alpha _s(1/\rho _0)`$. We assume that $`\alpha _s(1/\rho _0)1`$, which is only marginally satisfied in QCD. Our ladder is resumming the leading logarithms of center of mass energy $`s`$, which in this case implies resummation of all powers of the parameter
$`e^{\frac{4\pi }{\alpha _s}}\mathrm{ln}s\mathrm{\hspace{0.17em}1}.`$ (5)
Note that the dependence of our small parameter on $`\alpha _s`$ is non-analytical, which will introduce the non-analytical dependence on $`\alpha _s`$ in the calculated value of the pomeron intercept.
As energy in each rung of the ladder increases, quantum corrections to the tree level calculation of the vertices become important. The quantum corrections are known to modify the energy dependence of the $`2n`$ transition vertices . The energy dependence of the $`2n`$ total (summed over $`n`$) cross section is given by the following formula
$`\sigma _2(s)\mathrm{exp}\left[{\displaystyle \frac{1}{g^2}}F\left({\displaystyle \frac{\sqrt{s}}{E_{sph}}}\right)\right].`$ (6)
Here $`s`$ is the center of mass energy of the system, $`E_{sph}`$ is the sphaleron energy. $`F`$ is known only for the small values of its argument . The lowest order term in $`F`$ is given by the tree level calculations, and the quantum corrections provide higher order terms in the expansion of $`F`$ . It is also known that for a large number of gluon legs (large $`n`$) the value of the saddle point is shifted. Thus for large number of produced gluons the vertices may not be exactly the same instanton–induced vertices as for small $`n`$. This only substantiates our argument that the instanton field may not be an exact mechanism responsible for the multiple gluon interactions in the ladder of Fig. 1. There are many problems related to the determination of the function $`F(\sqrt{s}/E_{sph})`$ and we are not going to review all of them here. We will simply point out that the growth with energy of the cross section of Eq. (6) should eventually stop due to the unitarity constraint
$`\sigma <{\displaystyle \frac{const}{s}}\mathrm{exp}\left(const{\displaystyle \frac{2\pi }{\alpha _s}}\right).`$ (7)
Moreover, it is possible that the discussed $`2n`$ cross section will fall off at the energies of the two incoming gluons much higher than the sphaleron energy, since the instanton effects may not be important in that kinematic region. Assuming that this is true throughout the paper we will approximate the multi gluon interactions by the tree level calculation for the energies below the sphaleron energy and we will just put it to be zero for $`\sqrt{s}>E_{sph}`$. This approximation is rather crude and will be improved in the subsequent work , where several different behaviors of the $`2n`$ scattering cross section will be considered. In particular we will consider the case when the unitarity limit of Eq. (7) has been reached.
Let us discuss the (Minkowskian) space–time picture of high energy scattering amplitude that we are going to evaluate. Since at high energies the scattering occurs over large longitudinal distances, in order to construct the amplitude one may need to have several instantons involved in the process (see Fig. 1). At first glance, this seems to contradict the assumption of the dilute instanton gas that we are going to use<sup>4</sup><sup>4</sup>4A computation in a self–consistent instanton vacuum model would be a very interesting extension of this work.. The resolution of this apparent problem is in the very fact that a pomeron exchange is not an instantaneous interaction. Pomeron exchange is best represented in the wave function picture of high energy interactions . In that formalism in order to exchange a pomeron the colliding hadrons have to develop large multi–gluon fluctuations, which, in turn, interact with each other, which corresponds to the pomeron exchange in the usual t–channel language.
The space–time picture of our soft pomeron constructed in the spirit of the wave function formalism is shown in Fig. 2. Single pomeron exchange corresponds to a single gluon, which, after being emitted off the original hadron, propagates through space, inducing a chain of instanton transitions between different topological vacua and producing more gluons in each transition. The coherence length of the small-$`x`$ gluon in Fig. 2 is large, $`\tau k_+/k_{}^2`$, where $`k`$ is the gluon’s momentum. Thus a high energy gluon carrying a large “plus” component of momentum can propagate over large longitudinal distances inducing several instanton transitions. Even in the dilute instanton gas approximation the gluon should therefore be able to generate many instanton transitions while traveling over large distances in the longitudinal direction at the given value of the impact parameter.
If the process is viewed in the rest frame of one of the hadrons, then after several instanton transitions, the propagating gluon would interact with the target hadron at rest. In the center of mass frame the propagating gluon would find another gluon coming from the second hadron and would interact with it through an instanton transition, producing more gluons.
Multiple pomeron exchanges in this picture would correspond to the case when the gluons produced in the interactions of Fig. 2 in turn start propagating through the instanton gas and interacting . However, a study of that interesting scenario is beyond the scope of our paper. Here we will just argue that these multiple pomeron exchanges are suppressed compared to the single pomeron exchange contribution, for the reasons described below. Of course they will become important as the center of mass energy of the collision becomes very high.
The paper is organized as follows: after discussing the general issues related to instantons in Sect. II we will proceed to calculating the intercept of the pomeron of Fig. 1 in Sect. III. We will then calculate the slope and trajectory of the pomeron in Sect. IV and conclude by summarizing our results in Sect. V.
## II General Issues
In Minkowski space, instantons correspond to tunneling transitions between different topological vacua . The instanton calculations therefore make sense only at energies much smaller than the energy of the potential barrier separating the vacua (“sphaleron” energy $`E_{sph}`$).
The BPST instanton solution in the regular gauge is given by
$`A_\mu ^a(x)={\displaystyle \frac{2\eta _{a\mu \nu }x_\nu }{g(x^2+\rho ^2)}},`$ (8)
and the corresponding field strength is
$`(G_{\mu \nu }^a)^2={\displaystyle \frac{192\rho ^4}{g^2(x^2+\rho ^2)^4}}.`$ (9)
The energy along the instanton path can be defined as
$`E(\tau )={\displaystyle \frac{1}{8}}{\displaystyle d^3x(G_{\mu \nu }^a)^2},`$ (10)
and its maximum at $`\tau =0`$ corresponds to the sphaleron energy (see ):
$`E_{sph}=E(\tau =0)={\displaystyle \frac{3\pi ^2}{\rho }}{\displaystyle \frac{1}{g^2}}={\displaystyle \frac{3\pi }{4\alpha _s\rho }}.`$ (11)
To estimate this quantity, we take $`\rho (700\mathrm{MeV})^1`$, and $`\alpha _s(1/\rho )=g^2/4\pi 0.7`$; we get $`E_{sph}2.4÷2.6\mathrm{GeV}`$.
At first glance, this value of $`E_{sph}`$ suggests that the perturbative treatment about the instanton solution is not applicable for hadron scattering at high energies $`\sqrt{s}>>E_{sph}`$. However, this conclusion is premature. Indeed, the space–time picture of high energy scattering discussed in the Introduction implies that the scattering is described by the ladder–type diagrams, where the number of rungs is proportional to rapidity $`ylns`$, and the energy in each rung $`E_{rung}`$ is independent of the total collision energy ($`E_{rung}`$ is typically on the order of a few GeV ). The crucial condition is therefore $`E_{rung}<<E_{sph}`$, and it is likely to be satisfied (later, we will see that this is indeed the case). Therefore, the applicability of the instanton–based approach to high energy scattering is, at least a priori, plausible.
Later we will find it convenient to use the singular gauge for the instanton field (the gauge in which the singularity of the potential is shifted to the origin); the corresponding expression is
$`A_\mu ^a(x)={\displaystyle \frac{2\rho ^2\overline{\eta }_{a\mu \nu }x_\nu }{gx^2(x^2+\rho ^2)}}.`$ (12)
In momentum space, the instanton field is given by
$$A_\mu ^a(k)=\frac{i(4\pi )^2}{g}\frac{\overline{\eta }_{a\mu \nu }k_\nu }{(k^2)^2}\left[1\frac{1}{2}(k\rho )^2K_2(k\rho )\right].$$
(13)
## III Soft pomeron: the intercept
In this section we will calculate the intercept of the soft pomeron depicted in Fig. 1 to hadronic cross sections and will calculate the intercept of the soft pomeron given by that equation. We assume that the scale of the running coupling constant is being set by the typical size of the QCD instanton $`\rho _0`$, and that the coupling constant at that scale is small $`\alpha _s(1/\rho _0)\mathrm{\hspace{0.17em}1}`$. This assumption is marginally well satisfied in QCD, but is crucial for the use of QCD instantons. In the recent year there have been proposed several scenarios of how the strong coupling constant behaves at large distances suggesting that it never becomes larger than $`1`$. If the ideas stated in are correct then our small coupling assumption is likely to be justified. Otherwise we note that the small coupling assumption is of course better justified for electroweak interactions .
Assuming that $`\alpha _s\mathrm{\hspace{0.17em}1}`$ we will neglect all the usual perturbative QCD vertices as bringing extra powers of $`\alpha _s`$. We will be resumming only the instanton–induced vertices, which give powers of $`\mathrm{exp}\left(\frac{2\pi }{\alpha _s}\right)\mathrm{\hspace{0.17em}1}`$ and our model of the pomeron could be viewed as resumming powers of this parameter.
The structure of this section is the following. We will first derive the intercept of the pomeron of Fig. 1. Then we will estimate the magnitude of virtual corrections to the real emission diagrams of Fig. 1. We will then calculate the numerical value of the pomeron intercept predicted by our model.
### A Calculation of the intercept
To be able to calculate the diagrams in Fig. 1 we have to construct the basic vertex employed there — the vertex coupling several gluons to each other through an instanton. This will be done in a way similar to the calculations of the baryon number violating amplitudes involving instantons in electroweak theory done in . To calculate the multiple gluon vertex we have to construct the Green function
$$G_{\alpha \beta \mu _1\mathrm{}\mu _n}^{aba_1\mathrm{}a_n}(q_1,q_2;k_1,\mathrm{},k_n)=(2\pi )^4\delta ^4(q_1+q_2k_1\mathrm{}k_n)A_\alpha ^a(q_1)A_\beta ^b(q_2)A_{\mu _1}^{a_1}(k_1)\mathrm{}A_{\mu _n}^{a_n}(k_n),$$
(14)
where $`A_\mu ^a(k)`$ are instanton fields given by Eq. (13) and the brackets $`\mathrm{}`$ imply averaging over the instanton sizes in the single instanton sector. The Green function of Eq. (14) is depicted in Fig. 3. It combines two space-like t-channel gluons carrying momenta $`q_1`$ and $`q_2`$ with $`n`$ final state gluons with momenta $`k_1\mathrm{}k_n`$, which are taken to be on the mass shell. Thus, in Minkowski space, $`q_i^2=\underset{¯}{q}_i^2`$, where $`\underset{¯}{q}_i`$ is the transverse component of the gluon’s momentum, and $`k_i^2=0`$. We need the Green function to be in the momentum space representation because we are going to use it in the momentum space diagram calculations below. The difference between our Green function and the one usually considered in the calculations of $`2n`$ process in QCD and electroweak theory is that the two on-mass shell gluons in the initial state of the electroweak process are substituted here by two t-channel gluons which are far off mass shell.
Of course the Green function of Eq. (14) describes the multiple gluon interactions through an instanton only at not very high energies. As energy increases the quantum corrections start becoming important , as was discussed in the Introduction. At very high energies of the incoming pair of gluons the quantum corrections modify the behavior of the Green function and the corresponding $`2n`$ cross section. In our approach we assume that at high center of mass energies of the hadron–hadron (or quarkonium–quarkonium) scattering the dominant contribution comes from the pomeron diagrams in Fig. 1 with a large number of rungs in the ladder. Thus the energy per rung, or, equivalently, per vertex in the ladder remains not too large, justifying the use of the Green function of Eq. (14). We will return to quantify this issue below.
For a gluon on the mass shell the instanton field of Eq. (13) becomes
$`A_\mu ^a(k){\displaystyle \frac{i4\pi ^2\rho ^2}{g}}{\displaystyle \frac{\overline{\eta }_{a\mu \nu }k_\nu }{k^2}}\text{as}k^20.`$ (15)
Substituting the field of Eq. (15) for each $`k_i`$ line into Eq. (14) and employing the full instanton field of Eq. (13) for the $`q`$ lines we obtain
$`G_{\alpha \beta \mu _1\mathrm{}\mu _n}^{aba_1\mathrm{}a_n}(q_1,q_2;k_1,\mathrm{},k_n)=(2\pi )^4\delta ^4(q_1+q_2k_1\mathrm{}k_n){\displaystyle _0^{\mathrm{}}}𝑑\rho n(\rho )\rho ^{2n}\left({\displaystyle \frac{i4\pi ^2}{g}}\right)^n{\displaystyle \frac{(4\pi )^4}{g^2}}`$
$`\times \left({\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{\overline{\eta }_{a_i\mu _i\nu _i}k_{\nu _i}}{k_i^2}}\right){\displaystyle \frac{\overline{\eta }_{a\alpha \gamma }q_{1\gamma }}{(q_1^2)^2}}\left[1{\displaystyle \frac{1}{2}}(q_1\rho )^2K_2(q_1\rho )\right]{\displaystyle \frac{\overline{\eta }_{b\beta \delta }q_{2\delta }}{(q_2^2)^2}}\left[1{\displaystyle \frac{1}{2}}(q_2\rho )^2K_2(q_2\rho )\right],`$ (16)
where $`n(\rho )`$ is the size distribution of instantons. We want to obtain an effective vertex for multiple pomeron interactions. For that we have to truncate the external legs of the Green function of Eq. (16). To amputate the propagators of the external gluon lines one has to also remove the numerators of these propagators. For that we need to know the gauge condition of the fields of Eq. (12) and Eq. (13). Usually the field of Eq. (12) is derived using the Schwinger gauge condition $`x_\mu A_\mu =\mathrm{\hspace{0.17em}0}`$. However, we note that it also satisfies the covariant gauge condition $`_\mu A_\mu =\mathrm{\hspace{0.17em}0}`$. Therefore we will be using it as a covariant (Feynman) gauge field and will employ it in the covariant gauge calculation of diagrams. Therefore, to truncate the Green function of Eq. (16) we have to just multiply it by $`q_1^2q_2^2k_1^2\mathrm{}k_n^2`$. The resulting amputated Green function is
$`\mathrm{\Gamma }_{\alpha \beta \mu _1\mathrm{}\mu _n}^{aba_1\mathrm{}a_n}(q_1,q_2;k_1,\mathrm{},k_n)=(2\pi )^4\delta ^4(q_1+q_2k_1\mathrm{}k_n){\displaystyle _0^{\mathrm{}}}𝑑\rho n(\rho )\rho ^{2n}\left({\displaystyle \frac{i4\pi ^2}{g}}\right)^n{\displaystyle \frac{(4\pi )^4}{g^2}}`$
$`\times \left({\displaystyle \underset{i=1}{\overset{n}{}}}\overline{\eta }_{a_i\mu _i\nu _i}k_{\nu _i}\right){\displaystyle \frac{\overline{\eta }_{a\alpha \gamma }q_{1\gamma }}{q_1^2}}\left[1{\displaystyle \frac{1}{2}}(q_1\rho )^2K_2(q_1\rho )\right]{\displaystyle \frac{\overline{\eta }_{b\beta \delta }q_{2\delta }}{q_2^2}}\left[1{\displaystyle \frac{1}{2}}(q_2\rho )^2K_2(q_2\rho )\right].`$ (17)
Now we are in the position to calculate the intercept of the soft pomeron in Fig. 1. The calculation will be done in the Minkowski space. To perform it we have to first multiply the amputated Green function of Eq. (17) by the numerators of the propagators of the t-channel gluons in the light cone gauge (Weizsäcker–Williams approximation)
$`{\displaystyle \frac{q_{1\alpha }^{}}{q_{1+}}}{\displaystyle \frac{q_{2\beta }^{}}{q_2}},`$
where the gluon $`q_1`$ carries a large $`+`$ component of the light cone momentum and the gluon $`q_2`$ carries a large $``$ component. This is a standard procedure, which is described in in greater detail. Then we should multiply the gluon line factors $`k_i`$’s in Eq. (17) by the polarization vectors $`ϵ_i^{\lambda _i}(k_i)`$ and square the amputated Green function of Eq. (17) excluding the $`\delta `$ function. We should keep in mind that the color space orientation of the instanton in the complex conjugate amplitude is, in principle, different from the color space orientation of the instanton in the amplitude . Therefore we will have to average over all possible orientations, for which we will be using the approximation introduced in . We will integrate the obtained expression over the phase space of the produced gluons and sum over colors and polarizations $`\lambda _i`$
$`{\displaystyle \frac{1}{n!}}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{d^3k_i}{2(2\pi )^3\omega _i}}{\displaystyle \underset{\lambda _i}{}}`$
where $`\omega _i=k_4^i`$ and sum over $`n`$ runs from $`1`$ to infinity. We also have to multiply everything by one of each of the denominators of the propagators of the t-channel gluons
$`{\displaystyle \frac{1}{\underset{¯}{q}_1^2}}{\displaystyle \frac{1}{\underset{¯}{q}_2^2}}`$
and average over the colors of gluons $`q_1`$ and $`q_2`$, which would give a factor of $`1/(N_c^21)^2`$. We also have to include the symmetry factor of $`(n!)^2`$. We end up with the following expression for the kernel of the integral equation describing the pomeron of Fig. 1
$`K(q_1^{},q_2^{})\mathrm{ln}{\displaystyle \frac{1}{x}}={\displaystyle \frac{1}{(N_c^21)^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle d^4q\delta ^4(q\underset{i=1}{\overset{n}{}}k_i)n!}`$
$`\times \left({\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{d^3k_i}{2(2\pi )^3\omega _i}}{\displaystyle 𝑑\xi \underset{\lambda _i}{}\left[\overline{\eta }_{a_i\mu _i\nu _i}k_{\nu _i}ϵ_{\mu _i}^{\lambda _i}(k_i)\right]\left[U^{a_ib_i}(\xi )\overline{\eta }_{b_i\mu _i\nu _i}k_{\nu _i}ϵ_{\mu _i}^{\lambda _i}(k_i)\right]^{}}\right)`$
$`\times \left({\displaystyle \frac{4\pi ^2}{g}}\right)^{2n}{\displaystyle \frac{(4\pi )^8}{g^4}}{\displaystyle \frac{1}{\underset{¯}{q}_1^2\underset{¯}{q}_2^2}}\left[{\displaystyle \frac{\overline{\eta }_{a\alpha }q_{1\alpha }^{}\overline{\eta }_{b\beta +}q_{2\beta }^{}}{\underset{¯}{q}_1^2\underset{¯}{q}_2^2}}\right]\left[{\displaystyle \frac{U^{aa^{}}(\xi )\overline{\eta }_{a^{}\alpha }q_{1\alpha }^{}U^{bb^{}}(\xi )\overline{\eta }_{b^{}\beta +}q_{2\beta }^{}}{\underset{¯}{q}_1^2\underset{¯}{q}_2^2}}\right]^{}`$
$`\times \left[{\displaystyle _0^{\mathrm{}}}𝑑\rho n(\rho )\rho ^{2n}\left(1{\displaystyle \frac{1}{2}}(q_1^{}\rho )^2K_2(q_1^{}\rho )\right)\left(1{\displaystyle \frac{1}{2}}(q_2^{}\rho )^2K_2(q_2^{}\rho )\right)\right]^2,`$ (18)
with $`q=q_1+q_2`$. $`U^{ab}(\xi )`$ is the matrix of global $`SU(2)`$ rotations in the adjoint representation and it is responsible for rotating the instanton in the complex conjugate amplitude in the color space . We can perform the summation over polarizations and averaging over the color space orientations of the instantons using the approximation outlined in
$`{\displaystyle 𝑑\xi \underset{i=1}{\overset{n}{}}\underset{\lambda _i}{}\left[\overline{\eta }_{a_i\mu _i\nu _i}k_{\nu _i}ϵ_{\mu _i}^{\lambda _i}(k_i)\right]\left[U^{a_ib_i}(\xi )\overline{\eta }_{b_i\mu _i\nu _i}k_{\nu _i}ϵ_{\mu _i}^{\lambda _i}(k_i)\right]^{}}\left({\displaystyle \frac{E_{sph}^2}{3n^2}}\right)^n{\displaystyle \frac{2\pi nE_{sph}^{2n}}{3^ne^{2n}(n!)^2}}`$ (19)
for large n. Eq. (19) can be understood in the rest frame of the system of two t-channel gluons. There the dominant contribution to Eq. (19) comes from equally distributing center of mass energy among $`n`$ final state gluons, which leads to the $`(E_{sph}^2/n^2)^n`$ in Eq. (19). The estimate of Eq. (19) is somewhat crude, but was justified in . In writing Eq. (19) in the approximation outlined in we are likely to omit factors of $`n`$. However, as could be seen from the calculations below, we do not have the precision to keep all the powers of $`n`$ at this point. Extra powers of $`n`$ may arise from different ways of averaging over the instanton sizes (see Eq. (29) and Eq. (35)). The $`q^2`$ integration in Eq. (18) is sharply peaked at the upper cutoff, as will be seen from the calculations presented below, which we will put to be $`E_{sph}`$. This allowed us to just put $`E_{sph}^2`$ in Eq. (19). However, the uncertainty in our knowledge of $`E_{sph}^2`$ may also introduce significant numerical changes, which we can not control at the moment. We are going to clarify those in the subsequent work . We have also put $`U^{ab}(\xi )=\mathrm{\hspace{0.17em}1}`$ for the t-channel gluon lines. As the left hand side of Eq. (19) should of course be a Lorentz invariant expression in Minkowski space it can only depend on boost invariant quantities. Thus we can see that for $`n=1`$ the answer should be proportional to $`k^2=0`$ and should, therefore, be zero. For higher values of $`n`$ the left hand side of Eq. (19) may depend on the products of momenta of different gluons, which results in a non-zero answer.
In evaluating the expression in Eq. (18) we have to note that the perturbative size distribution of the QCD instantons $`n(\rho )`$ is divergent at large sizes $`\rho `$. However, the lattice data suggests that the distribution actually starts to fall off very steeply with increasing $`\rho `$ after some critical value of $`\rho =\rho _0`$. This behavior could be expected by the following argument based on the low energy theorem of Eq. (2), which has become known as “$`b/4`$” problem . The issue is the following: it is impossible to satisfy the low energy theorem of Eq. (2) with a purely classical field. The energy-momentum tensor of a classical field given by Eq. (1) has only one power of $`b`$ in it, since the classical field has no information about renormalization and running coupling constant in it. Thus if one tries to satisfy Eq. (2) with this energy momentum tensor one would get an extra power of $`b`$ on the left hand side which would not cancel. The commonly accepted resolution of this problem is to assume that the instantons in the instanton gas interact with each other, which leads to a different infrared cutoff on the integrals over $`\rho `$ on the left and right hand sides of Eq. (2) . This leads us to conclude that QCD regularizes itself at large distances by cutting off the growth of the instanton size distribution through some non-perturbative, but not classical mechanism. This conclusion seems to be supported by lattice data . For the purpose of this paper, in order to cure the problem we have to introduce an upper cutoff $`\rho _0`$ in the $`\rho `$ integrals in Eq. (18).
In our approximation the sum over polarizations and colors (Eq. (19)) became independent of $`k_i`$’s. Therefore we can perform the integration over the phase space of $`n`$ gluons. The integration yields
$`\delta ^4(q{\displaystyle \underset{i=1}{\overset{n}{}}}k_i){\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{d^3k_i}{2\omega _i}}={\displaystyle \frac{(\pi /2)^{n1}(q^2)^{n2}}{(n1)!(n2)!}}.`$ (20)
After some simple algebra Eq. (18) becomes
$`K(q_1^{},q_2^{})\mathrm{ln}{\displaystyle \frac{1}{x}}={\displaystyle \frac{1}{(N_c^21)^2}}{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle d^2\underset{¯}{q}\frac{dq_+}{2q_+}𝑑q^2n!\frac{1}{(2\pi )^{3n}}\frac{(\pi /2)^{n1}(q^2)^{n2}}{(n1)!(n2)!}\frac{2\pi nE_{sph}^{2n}}{3^ne^{2n}(n!)^2}\left(\frac{4\pi ^2}{g}\right)^{2n}}`$
$`\times {\displaystyle \frac{(4\pi )^8}{g^4}}{\displaystyle \frac{1}{q_1^4q_2^4}}\left[{\displaystyle _0^{\rho _0}}𝑑\rho n(\rho )\rho ^{2n}\left(1{\displaystyle \frac{1}{2}}(q_1^{}\rho )^2K_2(q_1^{}\rho )\right)\left(1{\displaystyle \frac{1}{2}}(q_2^{}\rho )^2K_2(q_2^{}\rho )\right)\right]^2.`$ (21)
We note again that the instanton size distribution is very sharply peaked around its’ maximum given by $`\rho _0`$. This statement is supported by the lattice data and by various instanton models . The size distribution for $`\rho <\rho _0`$ is very close to the perturbative results , which sharply increases near $`\rho _0`$. For $`\rho >\rho _0`$ the distribution falls off steeply , justifying our cutoff procedure in Eq. (21). Thus without loosing the accuracy of our calculations we can put $`\rho =\rho _0`$ in the factors containing the confluent Bessel functions in Eq. (21). Now we can see that the transverse momentum integration factorizes in the expression for the kernel of the equation for the pomeron–mediated forward amplitude (21). Each segment of the ladder in Fig. 1, consisting of the square of the Green function of Fig. 3, gives a factor of
$`K(q^{}){\displaystyle \frac{1}{q_{}^4}}\left(1{\displaystyle \frac{1}{2}}(q_{}\rho _0)^2K_2(q_{}\rho _0)\right)^2`$ (22)
to the rung of the ladder above it and a similar factor (with a different momentum) to the rung below it. Therefore each rung of the ladder brings in a square of the factor in Eq. (22) integrated over the transverse momentum
$`I_0\rho _0^6={\displaystyle d^2q_{}\frac{1}{q_{}^8}\left(1\frac{1}{2}(q_{}\rho _0)^2K_2(q_{}\rho _0)\right)^4}.`$ (23)
Therefore, to obtain the pomeron’s intercept out of the integral kernel in Eq. (21) we should substitute the integral $`I_0`$ instead of the transverse momentum integration in Eq. (21). $`I_0`$ is easy to evaluate numerically, which gives
$`I_0\mathrm{\hspace{0.17em}0.014}.`$ (24)
Integration over $`q_+`$ in Eq. (21) gives a factor of $`\mathrm{ln}\frac{1}{x}`$. Actually, the summation of the diagrams in Fig. 1 can be performed by the following simple integral equation
$`\varphi (q_1^{},Y=\mathrm{ln}(1/x))=K(q_1^{})+CK(q_1^{}){\displaystyle _0^Y}𝑑y{\displaystyle d^2q_2^{}K(q_2^{})\varphi (q_2^{},y)}`$ (25)
where the kernel of Eq. (21) is $`K(q_1^{},q_2^{})=CK(q_1^{})K(q_2^{})`$ and constant $`C`$ absorbs sum over $`n`$ and integral over $`q^2`$ in Eq. (21). $`\varphi (q^{},Y=\mathrm{ln}(1/x))`$ denotes the sum of all diagrams of Fig. 1–type, i. e., the pomeron–induced structure function or cross section. The initial condition is given by $`K(q_1^{})`$, which is not crucial for us and convenient for determination of the intercept. One can check that
$`\varphi (q^{},Y)=K(q^{})\mathrm{exp}(\mathrm{\Delta }_{soft}Y)`$ (26)
is the solution of Eq. (25) with
$`\mathrm{\Delta }_{soft}=C{\displaystyle d^2q^{}K^2(q^{})}.`$ (27)
A much more serious problem is posed by the integration over $`q^2`$ and summation over $`n`$ ( both absorbed in $`C`$ in Eq. (25) ) in Eq. (21), which is potentially very dangerous if one allows the upper limit of the integration to be as large as the center of mass energy of the system. This is what Eq. (21) seems to suggest if one takes it at face value. However, our Eq. (21) is valid only at relatively low energies. As energy increases quantum corrections become important. As was shown in the instanton-induced cross section for $`2n`$ process in QCD falls off with energy for energies above certain threshold energy $`E_0`$ due to quantum correction calculated in . Similar behavior results from the unitarity constraint derived in . Here we assume that this is also the case for our expression for the intercept. Namely we believe that at higher energies per rung in the ladder the quantum corrections would significantly slow down and later on completely overturn the growth of the expression in Eq. (21) with energy, as was discussed in the Introduction. These effects could be taken into account by putting an effective upper cutoff $`E_{sph}^2`$ in the $`q^2`$ integration in Eq. (21). After performing all the integrations mentioned above Eq. (21) yields the following expression for the pomeron’s intercept
$`\mathrm{\Delta }_{soft}={\displaystyle \frac{\pi I_0\rho _0^4}{(N_c^21)^2}}{\displaystyle \frac{(4\pi )^6}{\alpha ^3}}{\displaystyle \frac{E_{sph}^2\rho _0^2}{6e^2}}{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{[(n1)!]^3}}\left({\displaystyle \frac{\pi E_{sph}^4}{6\alpha e^2}}\right)^{n1}\left({\displaystyle _0^{\rho _0}}𝑑\rho n(\rho )\rho ^{2n}\right)^2.`$ (28)
Employing the same argument about the sharpness of the distribution of $`n(\rho )`$ as was used in performing the transverse momentum integral in Eq. (21) we could rewrite
$`{\displaystyle _0^{\rho _0}}𝑑\rho n(\rho )\rho ^{2n}\rho _0^{2n}N_0,`$ (29)
where we have defined
$`N_0={\displaystyle _0^{\rho _0}}𝑑\rho n(\rho ).`$ (30)
Substituting Eq. (29) into Eq. (28) we obtain
$`\mathrm{\Delta }_{soft}={\displaystyle \frac{\pi I_0(N_0\rho _0^4)^2}{(N_c^21)^2}}{\displaystyle \frac{(4\pi )^6}{\alpha ^3}}{\displaystyle \frac{E_{sph}^2\rho _0^2}{6e^2}}{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{[(n1)!]^3}}\left({\displaystyle \frac{\pi E_{sph}^4\rho _0^4}{6\alpha e^2}}\right)^{n1}.`$ (31)
Note that $`N_0`$ has the dimensions of $`M^4`$ and, therefore, Eq. (31) is, of course, dimensionless. We can rewrite Eq. (31) in a more compact form with the help of a generalized hypergeometric function
$`\mathrm{\Delta }_{soft}={\displaystyle \frac{\pi I_0(N_0\rho _0^4)^2}{(N_c^21)^2}}{\displaystyle \frac{(4\pi )^6}{\alpha ^3}}{\displaystyle \frac{E_{sph}^2\rho _0^2}{6e^2}}\left[{}_{0}{}^{}F_{2}^{}(;1,1;{\displaystyle \frac{\pi E_{sph}^4\rho _0^4}{6\alpha e^2}})1\right].`$ (32)
One could be a little more careful in evaluating Eq. (28). We can use the perturbative expression for the size distribution of instantons
$`n(\rho )=d{\displaystyle \frac{1}{\rho ^5}}\left({\displaystyle \frac{2\pi }{\alpha }}\right)^{2N_c}e^{\frac{2\pi }{\alpha }}\left({\displaystyle \frac{\rho }{\rho _0}}\right)^b,`$ (33)
where $`d`$ is given by
$`d={\displaystyle \frac{0.466e^{1.679N_c}}{(N_c1)!(N_c2)!}}\mathrm{\hspace{0.17em}1.5\hspace{0.17em}10}^3\text{for}N_c=3`$ (34)
and $`b=(11/3)N_c`$ for pure gluodynamics. We also used $`1/\rho _0`$ as a renormalization scale in Eq. (33). With the distribution of Eq. (33) we obtain
$`{\displaystyle _0^{\rho _0}}𝑑\rho n(\rho )\rho ^{2n}=d\left({\displaystyle \frac{2\pi }{\alpha }}\right)^{2N_c}e^{\frac{2\pi }{\alpha }}{\displaystyle \frac{\rho _0^{2n4}}{2n+b4}},`$ (35)
which, after plugging it into Eq. (28) and summation over $`n`$ in it yields the following expression for the soft pomeron’s intercept
$`\mathrm{\Delta }_{soft}={\displaystyle \frac{\pi I_0d^2}{(N_c^21)^2}}\left({\displaystyle \frac{2\pi }{\alpha }}\right)^{4N_c}e^{\frac{4\pi }{\alpha }}{\displaystyle \frac{(4\pi )^6}{\alpha ^3}}{\displaystyle \frac{1}{81}}{\displaystyle \frac{E_{sph}^2\rho _0^2}{6e^2}}\left[{}_{2}{}^{}F_{4}^{}({\displaystyle \frac{9}{2}},{\displaystyle \frac{9}{2}};1,1,{\displaystyle \frac{11}{2}},{\displaystyle \frac{11}{2}};{\displaystyle \frac{\pi E_{sph}^4\rho _0^4}{6\alpha e^2}})1\right],`$ (36)
where we have explicitly used $`b=11`$ for $`N_c=3`$. The perturbative instanton size distribution of Eq. (33) has been shown to fit very well the recent lattice data for $`\rho <\rho _0`$ . Therefore the estimate of the intercept given in Eq. (36) is probably more precise than the estimate from Eq. (32). Though in both cases we have to make the same assumptions to simplify the transverse momentum integration in Eq. (21). We also had to put a cutoff on $`\rho `$ integration in arriving to both Eqs. (32) and (36), which is not a very bad approximation judging from lattice data , but still is an approximation. In Sect. IIIC we will compare numerical predictions of Eq. (32) and Eq. (36) for the soft pomeron’s intercept.
It should be stressed that the approach developed here in Eq. (25) – Eq. (36) is an example of the multiperipheral approach (see Ref. and references therein). It gives all typical properties of the multiperipheral “ladder” diagrams, namely, the typical transverse momentum of produced particles (gluons) which does not depend on the total energy ($`q^{}\rho _02÷3`$ in our case, which follows from Eq. (22)) and the average energy $`\widehat{s}`$ between the produced two bunches of gluons in Fig. 1 which also does not depend on the total energy and is equal to
$`\mathrm{\Delta }_{soft}\mathrm{ln}(\widehat{s}/E_{sph}^2)\mathrm{\hspace{0.17em}1}.`$ (37)
These features of our approach cannot be spoiled by virtual corrections but, of course, they have to be taken into account when one considers the numerical value of the pomeron intercept.
### B Virtual corrections
There are two types of virtual contributions to the soft pomeron we are discussing. One is the additional gluon exchanges between instantons in t-channel, an example of which is shown in Fig. 4B. These corrections do not introduce extra logarithms of energy ($`\mathrm{ln}\frac{1}{x}`$) in the problem. The other type of virtual contributions could come from reggeization of the gluon propagator due to the instanton interactions. This effects of course may introduce additional powers of $`\mathrm{ln}\frac{1}{x}`$.
Let us first concentrate on the corrections brought in by the t-channel gluons as demonstrated in Fig. 4B. We are going to compare the contribution of the real diagram which is included in our model of the soft pomeron above, given by Fig. 4A to the contribution of the same diagram with a virtual correction added to it, which is depicted in Fig. 4B.
The difference between the graphs in Figs. 4A and 4B is just in extra gluon line, which introduces a loop integral. After some simple algebra one can see that the integral brings in a factor
$`K=\left({\displaystyle \frac{4\pi ^2}{g}}\right)^2{\displaystyle \frac{d^4k}{(2\pi )^4}\frac{\left[1\frac{1}{2}(k\rho _0)^2K_2(k\rho _0)\right]^2\left[1\frac{1}{2}(|pk|\rho _0)^2K_2(|pk|\rho _0)\right]^2}{(k^2)^2[(pk)^2]^2}},`$ (38)
where the labeling of the momenta is explained in Fig. 5. One of the t-channel virtual lines carries the 4-momentum $`k`$ and the other one carries $`pk`$. In Eq. (38) we included only the factors which make the contribution of the diagram in Fig. 4B different from the contribution of the graph in Fig. 4A. We also assumed that the instantons in the amplitude and in the complex conjugate amplitude have the same orientation in the color space. This assumption is not going to change our final answer by a large factor, and, therefore is good for the estimate we are going to perform here.
The value of the integral in Eq. (38) is easy to estimate. First we note that the expression in Eq. (38) is a decreasing function of $`p`$. Therefore one can obtain an upper bound on the value of the integral (38) by estimating it numerically for $`p=0`$. The result reads
$`K=\mathrm{\hspace{0.17em}0.0052}{\displaystyle \frac{\pi }{2\alpha }}\rho _0^4.`$ (39)
We have to compare the factor given by Eq. (39) to the factor given by the $`p`$-line in Fig. 4A, which is just $`1/(p^2)^2`$. For the estimate we can assume that approximately $`p\mathrm{\hspace{0.17em}1}/\rho _0`$ (see Eq. (22)). Therefore we conclude that the diagram in Fig. 4B is different from the diagram in Fig. 4A by the factor of $`0.0052\pi /2\alpha `$. For the values of $`\alpha `$ taken at the scale corresponding to the instanton size ($`\alpha (1/\rho _0)\mathrm{\hspace{0.17em}0.6}`$) this number is still very small, of the order of $`0.01`$. We can see that resummation of t-channel virtual corrections, which are not enhanced by logarithms of energy can not give us a large contribution and could be neglected.
One might think that the diagram of Fig. 5 taken without the s-channel gluon lines would give us a contribution to the mass of a gluon, thereby creating a non-zero gluon’s mass. However, in constructing Eq. (38) we did not impose the condition that the gluons forming the gluon loop should be in the color octet state, as one should do in calculating the correction for the gluon’s mass. That condition was not used since with the s-channel gluons present in Fig. 5 the color of the pair of t-channel gluons is not fixed. Now we have to point out that for $`p=0`$ the contribution of this diagram without the s-channel gluons is zero, which could be seen after projecting the t-channel gluons onto the color octet state.
The second type of virtual corrections are enhanced by the factors of $`\mathrm{ln}\frac{1}{x}`$ and are similar to gluon reggeization corrections for the BFKL pomeron . To include these corrections which bring in the evolution in $`\mathrm{ln}\frac{1}{x}`$ one has to calculate the intercept of the same pomeron ladder of the type shown in Fig. 1 with two t-channel gluons forming a color octet combination and with the non-zero momentum transfer in the ladder. The two gluons in the t-channel should in turn include all the virtual corrections in them. A careful inclusion of the gluon reggeization proved to be a very complicated task and will be completed elsewhere . here we are going to estimate these corrections.
Let us calculate the contribution of a single rung in the ladder of Fig. 1, with the ladder taken in the octet state with non-zero momentum transfer $`t=\underset{¯}{q}^2`$ (see also Fig. 7). A direct repetition of calculations in section IIIA leads to the following expression for the octet channel integral $`I_8(q^2)`$ which replaces integral $`I_0`$ in Eq. (32) and Eq. (36) for the intercept of the Pomeron.
$`I_8(q^2)\rho _0^6={\displaystyle d^2k_{}\frac{|\left(\underset{¯}{k}\frac{1}{2}\underset{¯}{q}\right)\times \left(\underset{¯}{k}+\frac{1}{2}\underset{¯}{q}\right)|^2}{[(\underset{¯}{k}+\frac{1}{2}\underset{¯}{q})^2]^3[(\underset{¯}{k}\frac{1}{2}\underset{¯}{q})^2]^3}\left[1\frac{1}{2}\left(\underset{¯}{k}+\frac{1}{2}\underset{¯}{q}\right)^2\rho _0^2K_2\left(|\underset{¯}{k}+\frac{1}{2}\underset{¯}{q}|\rho _0\right)\right]^2}`$
$`\times \left[1{\displaystyle \frac{1}{2}}\left(\underset{¯}{k}{\displaystyle \frac{1}{2}}\underset{¯}{q}\right)^2\rho _0^2K_2\left(|\underset{¯}{k}{\displaystyle \frac{1}{2}}\underset{¯}{q}|\rho _0\right)\right]^2,`$ (40)
We will now assume that the virtual corrections to our pomeron lead to reggeization of the t-channel gluons. Moreover, to estimate the trajectory we will assume that the expression for the gluon Regge trajectory has a form
$`\alpha _G(q^2)=\mathrm{\hspace{0.17em}\hspace{0.17em}1}+\mathrm{\Delta }_{soft}{\displaystyle \frac{I_8(q^2)}{I_0}}.`$ (41)
Eq. (41) is our estimate of the gluon’s trajectory. One can see that $`I_8(q^2)0`$ at $`q^20`$.
To simplify the expression for the gluon’s Regge trajectory we note that approximately
$`{\displaystyle \frac{1}{z^4}}\left[1{\displaystyle \frac{1}{2}}z^2K_2(z)\right]^2{\displaystyle \frac{e^{z^2/2}}{16}}.`$ (42)
Using Eq. (42) we can easily calculate function $`I_8(q^2)`$, namely, it turns out that
$`I_8(q^2)={\displaystyle \frac{\pi }{256}}\left\{\mathrm{\hspace{0.17em}2}\left(e^{\frac{\underset{¯}{q}^2\rho _0^2}{4}}e^{\frac{\underset{¯}{q}^2\rho _0^2}{2}}\right)+{\displaystyle \frac{\underset{¯}{q}^2\rho _0^2}{2}}\left[Ei\left({\displaystyle \frac{\underset{¯}{q}^2\rho _0^2}{4}}\right)\mathrm{\hspace{0.17em}\hspace{0.17em}2}Ei\left({\displaystyle \frac{\underset{¯}{q}^2\rho _0^2}{2}}\right)\right]\right\}.`$ (43)
Eq. (43) should be substituted into Eq. (41) to obtain our approximation for the gluon’s Regge trajectory.
Now we have to determine whether the reggeization corrections can significantly influence the value of the our pomeron’s intercept. It turns out that gluon reggeization can change the value of $`\mathrm{\Delta }_{soft}`$ since it leads to an additional factor in Eq. (23), which reduces this equation to the form
$`I_0\rho _0^6={\displaystyle d^2q_{}\frac{1}{q_{}^8}\left(1\frac{1}{2}(q_{}\rho _0)^2K_2(q_{}\rho _0)\right)^4e^{2[\alpha _G(q_{}^2)1]\mathrm{ln}(\widehat{s}/E_{sph}^2)}},`$ (44)
where $`\widehat{s}`$ is the energy in one rung of the ladder (see above). As we have discussed the dominant contribution in the integral of Eq. (23) comes from $`q^2\rho _0^223`$ (see Eq. (42)) where $`2[\alpha _G(q_{}^2=2/\rho _0^2)1]\mathrm{ln}(\widehat{s}/E_{sph}^2)0.4÷0.5`$, since $`\mathrm{ln}(\widehat{s}/E_{sph}^2)\mathrm{\hspace{0.17em}\hspace{0.17em}1}/\mathrm{\Delta }_{soft}`$ ( see Eq. (37) ). Therefore, the additional form factor in Eq. (44) due to gluon reggeization can suppress the value of $`I_0`$, and, consequently, the value of $`\mathrm{\Delta }_{soft}`$ by a factor of 2 $`÷`$ 3. Explicit numerical calculations confirm these expectations. A more careful analysis of the contribution of virtual corrections will be done later . In the numerical estimates which will be done below we have to keep in mind that virtual corrections can introduce a factor of 2 $`÷`$ 3 uncertainty in the value of the pomeron’s intercept.
### C Value of the intercept
The main purpose of this paper is to provide a qualitative mechanism which could account for the effect of the soft pomeron. This was done in Sect. IIIA. Nevertheless, we could try to estimate the numerical value of the soft pomeron intercept obtained in Eqs. (32) and (36) and check how it compares with the experimental value of $`0.08`$ .
At the same time one has to be very careful with the numerical estimates of the intercept in Eq. (32) and Eq. (36). It is important to realize that the intercept of Eq. (32) strongly depends on the position of the maximum of the instanton size distribution $`\rho _0`$ and on the value of the integral of this distribution over all $`\rho `$’s given by $`N_0`$. The value of the intercept in Eq. (36) depends on the parameter $`\rho _0`$ since it determines the scale of the strong coupling constant. The current knowledge of these parameters is poor . The value of $`N_0`$ seems to be undetermined up to an order of magnitude . Uncertainty in the value of $`\rho _0`$ brings in an uncertainty in the value of the strong coupling constant $`\alpha (1/\rho _0)`$. Finally a very important question concerns the value of the cutoff, which was taken to be $`E_{sph}`$ in Eqs. (32) and (36). The cutoff corresponds the maximum invariant mass of the particles produced in each vertex. In electroweak physics it correspond to the energies when the growth of the cross section of the $`2n`$ instanton-induced process slows down and, maybe, even starts decaying. In QCD it may correspond to the energy at which the unitarity limit of the $`2n`$ process is reached . The exact value of the energy at which this happens is still not determined precisely . It is one of the most crucial assumptions of this work that the turnover of the cross section does happen in QCD and that the corresponding invariant mass $`E_{sph}^2`$ is not too large.
The intercept of the pomeron calculated above (Eq. (36)) has been derived for pure gluodynamics. If we want to compare our result to the experimentally measured one we have to include quarks in the theory. They would introduce suppression of the amplitudes through the Faddeev–Popov determinant . Quark lines could be included in the scattering in two ways: they could be produced from an instanton into the final state, in which case the suppression factor they introduce is small, of the order of $`10^5`$, so this type of quark effects could be neglected. Alternatively the quarks could be absorbed in the gluon condensate . In the latter case the suppression factor in the distribution of instantons Eq. (33) is
$`{\displaystyle \underset{q=u,d,s,\mathrm{}}{}}\mathrm{\hspace{0.17em}1.3}\left(m_q\rho {\displaystyle \frac{2\pi ^2}{3}}(0|\overline{q}q|0)\rho ^3\right).`$ (45)
The suppression for the intercept of Eq. (36) is given by the square of the factor in Eq. (45). For our estimate we will just put $`\rho =\rho _0`$ in Eq. (45) and use $`(0|\overline{q}q|0)=(0.25\text{GeV})^3`$. Taking the product over three flavors we obtain a suppression factor of $`0.023`$ for the intercept.
Keeping in mind all of the above mentioned limitations we substitute the value of $`E_{sph}=\mathrm{\hspace{0.17em}2.4}\text{GeV}`$ and $`\rho _0=\mathrm{\hspace{0.17em}0.3}fm`$ from into Eq. (36) and plot the resulting value of the pomeron’s intercept as a function of the coupling constant $`\alpha _s`$. We also included suppression introduced by virtual corrections by including an extra factor $`\delta `$ on the right hand side of Eq. (36). The resulting pomeron’s intercept is related to the intercept of Eq. (36) by the following relation
$`\mathrm{\Delta }_P=\delta \left[{\displaystyle \underset{q=u,d,s,\mathrm{}}{}}\mathrm{\hspace{0.17em}1.3}\left(m_q\rho _0{\displaystyle \frac{2\pi ^2}{3}}(0|\overline{q}q|0)\rho _0^3\right)\right]^2\mathrm{\Delta }_{soft}`$ (46)
In the plot shown in Fig. 6 we have put $`\delta =0.31`$. Of course we have to admit that the plot in Fig. 6 depends on the value of the suppression factor $`\delta `$. It is also very sensitive to the cutoff $`E_{sph}`$. A work on putting some constraints on these parameters is under way . With the progress in the field of instanton models the exact numerical value of the soft pomeron intercept might be calculated with more confidence and precision.
The maximum of the intercept depicted in Fig. 6 is close to the experimental result of $`0.08`$, shown there by a horizontal line. The value of the coupling constant at the scale set by the instanton size $`\rho _0`$ is near the point on the plot of Fig. 6 where the maximum is achieved. Thus our value of the intercept should be about $`0.08`$. Also, the pomeron’s intercept seems to vanish at small distances (large momenta), in agreement with experimental data . Also our model of soft pomeron allows a non-singular continuation of the value of the intercept into the region of large distances and large coupling constant. Even if the strong coupling constant becomes very large at small momenta (large distances) the intercept of our pomeron still remains small (see Fig. 6). That way the strong interactions are able to yield us with a small pomeron’s intercept even in the kinematic region where they are very strong. Here we have to recall that strictly speaking our quasi–classical approach is applicable only when the coupling constant is small. Nevertheless, the smallness of the intercept at large coupling is an interesting result, supported by phenomenological observations .
## IV Soft pomeron: the slope and trajectory
In this Section we are going to calculate the slope and the intercept of the soft pomeron conjectured above. We will include contribution of the virtual corrections by putting a factor $`\delta `$ in the expression for pomeron’s trajectory (see Sect. IIIB). The quark contribution will also be treated similarly to Sect. IIIC. Therefore the slope and trajectory of the soft pomeron which will be derived below will result only from the real emissions part of its kernel. This will correspond to summing the diagrams of the type shown in Fig. 1.
To calculate the slope of the pomeron considered here we will follow the standard procedure. Let us consider a pomeron with non-zero momentum transfer described by Mandelstam variable $`t`$. A rung of the off-forward pomeron’s ladder is shown in Fig. 7. One t-channel gluon carries a momentum $`k+\frac{1}{2}q`$, whereas the other one carries momentum $`k\frac{1}{2}q`$. Thus the momentum transfer is $`t=\underset{¯}{q}^2`$.
It is easy to see that the only difference between the pomeron ladder with non-zero momentum transfer and the $`t=0`$ pomeron considered in Sect. III is in the transverse momentum integral. An easy calculation shows that for off-forward pomeron the integral $`I_0`$ in the expression for the forward pomeron intercept (36) should be replaced by
$`I(q^2)\rho _0^6={\displaystyle d^2k_{}\frac{(\underset{¯}{k}^2\frac{1}{4}\underset{¯}{q}^2)^2}{[(\underset{¯}{k}+\frac{1}{2}\underset{¯}{q})^2]^3[(\underset{¯}{k}\frac{1}{2}\underset{¯}{q})^2]^3}\left[1\frac{1}{2}\left(\underset{¯}{k}+\frac{1}{2}\underset{¯}{q}\right)^2\rho _0^2K_2\left(|\underset{¯}{k}+\frac{1}{2}\underset{¯}{q}|\rho _0\right)\right]^2}`$
$`\times \left[1{\displaystyle \frac{1}{2}}\left(\underset{¯}{k}{\displaystyle \frac{1}{2}}\underset{¯}{q}\right)^2\rho _0^2K_2\left(|\underset{¯}{k}{\displaystyle \frac{1}{2}}\underset{¯}{q}|\rho _0\right)\right]^2,`$ (47)
which naturally reduces to Eq. (23) when $`\underset{¯}{q}=0`$. To obtain the slope of our pomeron we need to expand Eq. (47) up to the quadratic term in $`\underset{¯}{q}^2`$. After expansion of Eq. (47) in $`\underset{¯}{q}^2`$ and integration over the angles we obtain
$`I=I_0+\pi \underset{¯}{q}^2\rho _0^2{\displaystyle _0^{\mathrm{}}}𝑑k\rho _0L(k\rho _0)`$ (48)
where
$`L(k\rho _0)={\displaystyle \frac{1}{32(k\rho _0)^9}}(1{\displaystyle \frac{1}{2}}(k\rho _0)^2K_2(k\rho _0))^2[32+(k\rho _0)^6K_1(k\rho _0)^2+4(k\rho _0)^4K_2(k\rho _0)+8(k\rho _0)^4K_2^2(k\rho _0)`$
$`2(k\rho _0)^6K_2^2(k\rho _0)+2(k\rho _0)^4K_0(k\rho _0)\left(1{\displaystyle \frac{1}{2}}(k\rho _0)^2K_2(k\rho _0)\right)20(k\rho _0)^3K_3(k\rho _0)+2(k\rho _0)^5K_2(k\rho _0)K_3(k\rho _0)`$
$`+(k\rho _0)^6K_3^2(k\rho _0)+2(k\rho _0)^3K_1(k\rho _0)\left(10+(k\rho _0)^2K_2(k\rho _0)+(k\rho _0)^3K_3(k\rho _0)\right)+2(k\rho _0)^4K_4(k\rho _0)`$
$`(k\rho _0)^6K_2(k\rho _0)K_4(k\rho _0)].`$ (49)
The integral in Eq. (48) is dominated by the small values of the argument, as could be seen from the behavior of the function $`L(k\rho _0)`$. We can expand the function $`L(k\rho _0)`$ of Eq. (49) to obtain
$`L(k\rho _0){\displaystyle \frac{1}{256k\rho _0}}`$ (50)
when the argument is small. The full integral of Eq. (47) is convergent, whereas the integral in Eq. (48) seems to be divergent in the infrared. Therefore to regularize it after we plug Eq. (50) in it we have to use $`q`$ as a lower cutoff and $`1/\rho _0`$ as an upper cutoff. The result yields
$`I=I_0{\displaystyle \frac{\pi }{256}}q^2\rho _0^2\mathrm{ln}{\displaystyle \frac{1}{q\rho _0}}.`$ (51)
Recalling that a pomeron trajectory is given in terms of its’ intercept and slope by
$`\mathrm{\Delta }(t)=\mathrm{\Delta }+\alpha ^{}t`$ (52)
and comparing (52) to Eq. (28) and Eq. (51) we can conclude that for our pomeron
$`{\displaystyle \frac{\alpha ^{}}{\mathrm{\Delta }}}={\displaystyle \frac{\pi \rho _0^2}{256I_0}}\mathrm{ln}{\displaystyle \frac{1}{q\rho _0}}.`$ (53)
Since there is still some $`q`$ dependence left in Eq. (53) we conclude that our pomeron’s trajectory is not quite linear in $`t`$ and the notion of pomeron’s slope is not very well defined for it. A similar result was obtained for the soft pomeron trajectory in using Regge theory, which seemed to confirm experimental data of . The conclusion of was drawn out of two-pion exchange model of hadronic interactions. We can obtain a numerical estimate of the pomeron’s slope of Eq. (53) by assuming that $`\mathrm{ln}\frac{1}{q\rho _0}\mathrm{\hspace{0.17em}1}`$. Substituting the values of $`I_0`$ from Eq. (24) and $`\rho _0=0.3fm=1.5GeV^1`$ into Eq. (53) we obtain
$`{\displaystyle \frac{\alpha ^{}}{\mathrm{\Delta }}}\mathrm{\hspace{0.17em}2.0}GeV^2,`$ (54)
which is remarkably close to the experimental result for the soft pomeron
$`\left({\displaystyle \frac{\alpha ^{}}{\mathrm{\Delta }}}\right)_{exp}={\displaystyle \frac{0.25}{0.08}}GeV^2\mathrm{\hspace{0.17em}3.1}GeV^2.`$ (55)
In principle the integral of Eq. (47) substituted in the Eq. (28) (or either Eq. (32) or Eq. (36)) can give us the full trajectory of our pomeron. The integral in Eq. (47) could be estimated numerically for positive values of $`\underset{¯}{q}^2`$, which corresponds to the negative values of $`t=\underset{¯}{q}^2`$. However, for positive $`t`$ the integral in Eq. (47) has a singularity and if taken literally would be simply divergent. The resolution of this problem is the following: one should first perform the integration in Eq. (47) for negative $`t`$ and then analytically continue the results into the region of positive $`t`$. This should be done analytically and the exact analytical treatment of the integral in Eq. (47) seems to be rather complicated. Instead we will use an approximation, which preserves all the main qualitative features features of the answer.
Using Eq. (42) in Eq. (47) and performing the integration over $`k`$ yields the following expression
$`I(t){\displaystyle \frac{\pi }{256}}\left\{2e^{t\rho _0^2/2}e^{t\rho _0^2/4}{\displaystyle \frac{t\rho _0^2}{2}}\left[2Ei\left({\displaystyle \frac{t\rho _0^2}{2}}\right)Ei\left({\displaystyle \frac{t\rho _0^2}{4}}\right)\right]\right\}.`$ (56)
We can check the quality of the approximation done to obtain Eq. (56) by comparing its’ value for $`t=0`$ with $`I_0`$ in Eq. (24). Eq. (56) gives
$`I(0){\displaystyle \frac{\pi }{256}}\mathrm{\hspace{0.17em}0.012},`$ (57)
which is very close to the exact value $`I_0=0.014`$ in Eq. (24). If we expand Eq. (56) for small $`t`$ up to the terms linear in $`t`$ we would exactly recover the expression for the slope of the trajectory given by Eq. (53)! Thus Eq. (56) provides us with a very good approximation of the integral in Eq. (47).
Our soft pomeron’s trajectory could be obtained by substituting $`I(t)`$ from Eq. (56) into Eq. (36) instead of $`I_0`$. This gives
$`\mathrm{\Delta }_{soft}(t)={\displaystyle \frac{\pi d^2}{(N_c^21)^2}}\left({\displaystyle \frac{2\pi }{\alpha }}\right)^{4N_c}e^{\frac{4\pi }{\alpha }}{\displaystyle \frac{(4\pi )^6}{\alpha ^3}}{\displaystyle \frac{1}{81}}{\displaystyle \frac{E_{sph}^2\rho _0^2}{6e^2}}\left[{}_{2}{}^{}F_{4}^{}({\displaystyle \frac{9}{2}},{\displaystyle \frac{9}{2}};1,2,{\displaystyle \frac{11}{2}},{\displaystyle \frac{11}{2}};{\displaystyle \frac{\pi E_{sph}^4\rho _0^4}{6\alpha e^2}})1\right]`$
$`\times {\displaystyle \frac{\pi }{256}}\left\{2e^{t\rho _0^2/2}e^{t\rho _0^2/4}{\displaystyle \frac{t\rho _0^2}{2}}\left[2Ei\left({\displaystyle \frac{t\rho _0^2}{2}}\right)Ei\left({\displaystyle \frac{t\rho _0^2}{4}}\right)\right]\right\}.`$ (58)
After taking into consideration the virtual corrections and the quark contributions the answer for the pomeron’s trajectory becomes
$`\mathrm{\Delta }_P(t)=\delta \left[{\displaystyle \underset{q=u,d,s,\mathrm{}}{}}\mathrm{\hspace{0.17em}1.3}\left(m_q\rho _0{\displaystyle \frac{2\pi ^2}{3}}(0|\overline{q}q|0)\rho _0^3\right)\right]^2\mathrm{\Delta }_{soft}(t),`$ (59)
where $`\mathrm{\Delta }_{soft}(t)`$ is given by Eq. (58)
The soft pomeron’s trajectory of Eq. (59) is depicted in Fig. 8. It is plotted for $`\rho _0=0.3fm`$, $`E_{sph}=2.4GeV`$, $`\delta =0.31`$ and $`\alpha 0.75`$, i.e., the same values as were used for the estimates of the pomeron’s intercept in Sect. IIIC.
As could be seen from Fig. 8 the soft pomeron’s trajectory never becomes negative for negative $`t`$, which means that the total cross section generated by the soft pomeron exchange is always growing with energy. For positive values of $`t`$ the trajectory first starts growing but then slows down and even turns over and falls off. It does not reproduce the linear growth behavior which would be predicted by Regge theory. However, the linear behavior of the trajectory for positive $`t`$ is usually associated with confinement in QCD. It is a well known fact that the instanton–induced effects can not account for QCD confinement . Therefore one should not expect that the pomeron constructed out of instantons should exhibit such confinement features as a linear trajectory for positive $`t`$. It is most likely that different non-perturbative physical mechanisms play an important rôle in that region.
## V Summary and discussion
In our approach we envision the soft pomeron as a chain of topological transitions induced by high momentum gluons. This is the main point of the paper. The corresponding pomeron ladder is depicted in Fig. 1. The multi–gluon vertices of the ladder are generated through a strong quasi–classical vacuum field.
To quantify the pomeron illustrated in Fig. 1 we have used the instanton fields, which have an explicit analytical form which we could employ. The resulting expressions for the pomeron intercept are given in Eqs. (32) and (36), corresponding to different ways of averaging over the instanton sizes. We have estimated the virtual corrections and showed that they are likely to change the value of the soft pomeron’s intercept by a factor of $`2÷3`$ at the most. The numerical value of the intercept has been plotted in Fig. 6 as a function of the running coupling constant. The resulting intercept at $`\alpha _s(1/\rho _0)=\mathrm{\hspace{0.17em}0.7}`$ is very close to the phenomenological value of $`0.08`$ . The intercept falls off at small distances, which agrees with phenomenological observations of . At short distances (small momenta) the dominant contribution to the structure functions comes from the hard pomeron . The work on understanding how our soft pomeron merges with hard (BFKL) pomeron in that region is under way . At large distances (small momenta) the pomeron’s intercept also decreases. That way, in the region where the coupling constant is large and the interactions are very strong the pomeron’s intercept can still be small. Thus our results could provide the resolution of the long-standing puzzle of the smallness of the soft pomeron intercept in the strong coupling regime.
Finally, we have calculated the slope of the soft pomeron in Eq. (54). The value of the slope turned out to be close to the phenomenologically suggested value of $`0.25\text{GeV}^2`$ . The full trajectory of the pomeron has also been derived (see Eq. (58) and Eq. (59)) and plotted in Fig. 8. Since the instantons can not account for confinement in QCD we do not expect a linear trajectory for positive $`t`$. For negative $`t`$, as the absolute value of $`t`$ is getting larger the trajectory ceases to be linear and exhibits some curvature.
We can also estimate the multiplicity of the produced particles in a hadron–hadron scattering event mediated by our pomeron of Fig. 1. The particles would be produced by multiperipheral mechanism. From the estimate of Eq. (37) we can conclude that the density of instantons per unit of rapidity is rather small. However, since there are several particles emitted coherently from each instanton, we may expect correlations in particle production.
Let us estimate the typical number of particles produced on an instanton in the ladder of Fig. 1. For that purpose we will consider the series of Eq. (31) and estimate the average $`n`$ in that series. That would be a typical multiplicity of gluons produced off an instanton. A simple calculation yields
$`n=\mathrm{\hspace{0.17em}1}+{\displaystyle \frac{\pi E_{sph}^4\rho _0^4}{6\alpha _se^2}}{\displaystyle \frac{{}_{0}{}^{}F_{2}^{}(;2,2;\frac{\pi E_{sph}^4\rho _0^4}{6\alpha _se^2})}{{}_{0}{}^{}F_{2}^{}(;1,1;\frac{\pi E_{sph}^4\rho _0^4}{6\alpha _se^2})}}.`$ (60)
For the values of the parameter we have used above in estimating the pomeron’s intercept, $`E_{sph}=2.4\text{GeV}`$, $`\rho _0=0.3fm`$, $`\alpha _s0.7`$ (see Fig. 6), we obtain $`n3`$. This result agrees with the estimate of .
While our numerical results are very sensitive to a number of instanton–related parameters, we believe that our approach can provide an interesting new insight in the mechanism of high–energy scattering and particle production in QCD. Further theoretical and experimental developments are clearly needed to clarify the relationship between the soft pomeron and the properties of the QCD vacuum.
## Acknowledgments
The authors are very much indebted to Ian Balitsky, Larry McLerran, Al Mueller and Edward Shuryak for numerous helpful discussions about instantons. We would like to thank James Bjorken, Greg Carter, Hirotsugu Fujii, Asher Gotsman, Lev Lipatov, Uri Maor, Rob Pisarski, Andreas Ringwald, Thomas Schäfer, Fridger Schrempp, Dam Son, Chung-I Tan, Arkady Vainshtein and Valentin Zakharov for many informative and encouraging discussions. Yu. K. would like to thank the High Energy Physics Department at Tel Aviv University for their support and hospitality during the final stages of this work. E. L. thanks BNL Nuclear Theory group for their hospitality and creative atmosphere during several stages of this work.
This manuscript has been authored under Contract No. DE-AC02-98CH10886 with the U. S. Department of Energy. The research of Yu. K. and E. L. was supported in part by the BSF grant $`\mathrm{\#}`$ 9800276 and by Israeli Science Foundation, founded by the Israeli Academy of Science and Humanities.
|
warning/0007/quant-ph0007033.html
|
ar5iv
|
text
|
# Squeezing of electromagnetic field in a cavity by electrons in Trojan states
## I Introduction
The possibility of creating stationary, nondispersive, localized, wave packets describing a highly excited electron in a hydrogen atom strongly driven by circularly polarized microwave radiation was predicted theoretically several years ago and confirmed in numerous publications (for recent reviews of the subject see ). Such electronic states are called Trojan wave packets by analogy with the cloud of Trojan asteroid in the Sun-Jupiter system.
In all previous studies the microwave field was treated as an external, classical wave. Dressing of an electron by such a wave of a suitably chosen intensity and the frequency equal to the Kepler frequency of the electron on the Rydberg orbit makes the Trojan wave packets highly stable. Their life-time is of the order of one second , which makes them an interesting object of study for theoretical and perhaps even for practical reasons.
In the present paper a similar problem of nondispersive electronic wave packets is studied for an atom interacting with the quantized electromagnetic field. Such an approach allows for a fully dynamical treatment of an autonomous atom-field system. It automatically includes a back reaction of the atom on the electromagnetic field. Thus, one can study both the dynamics and the statistical properties of the electromagnetic radiation. Our study fully confirms the existence of Trojan states of the Rydberg electron in this new regime with almost exactly the same shape of the wave packet. The back reaction of the electron on the electromagnetic field pushes the field frequency off resonance. The quantum fluctuations of the electromagnetic field exhibit strong squeezing.
## II Hydrogen atom in a cavity
Anticipating the role of highly populated, discrete modes of the microwave field in the formation of Trojan electronic states, we consider a hydrogen atom in a microwave cavity. In the presence of a cavity we can separate a finite number of relevant degrees of freedom whereas in free space we would have to deal with a continuous spectrum which precludes the existence of localized stationary states of the system.
To allow for the rotational symmetry of the atom-field system we choose a cylindrical cavity. Its dimensions will be large enough to justify the dipole approximation in the coupling of hydrogen atom with the lowest cavity modes. The atom placed in the middle of cavity interacts only with $`\mathrm{TE}_{1n}`$ modes. For definitness we choose the two (degenerate) lowest modes of this type ($`n=1`$) (labeled by $`X`$ and $`Y`$) for which the mode functions have the form
$$𝐄^X=i𝒩\omega \mathrm{sin}\frac{\pi z}{L}𝐞_z\times _{}\mathrm{J}_1(x_{11}r/R)\mathrm{sin}\phi ,$$
(2)
$$𝐁^X=\frac{𝒩\pi }{L}\mathrm{cos}\frac{\pi z}{L}_{}\mathrm{J}_1(x_{11}r/R)\mathrm{sin}\phi ,$$
(3)
$$𝐄^Y=i𝒩\omega \mathrm{sin}\frac{\pi z}{L}𝐞_z\times _{}\mathrm{J}_1(x_{11}r/R)\mathrm{cos}\phi ,$$
(4)
$$𝐁^Y=\frac{𝒩\pi }{L}\mathrm{cos}\frac{\pi z}{L}_{}\mathrm{J}_1(x_{11}r/R)\mathrm{cos}\phi ,$$
(5)
where $`R`$ and $`L`$ are the radius and the length of the cavity and $`x_{11}`$ is the first (the smallest) solution of the equation $`d\mathrm{J}_1(x)/dx=0`$. The $`z`$-axis is taken along the cylinder axis and $`_{}=(/x,/y)`$. The frequency of the modes and the effective wave vector are given as
$`\omega `$ $`=`$ $`{\displaystyle \frac{c}{R}}(x_{11}^2+(\pi R/L)^2)^{\frac{1}{2}},`$ (6)
$`k`$ $`=`$ $`\sqrt{(\omega /c)^2+(\pi /L)^2}.`$ (7)
The value of the normalization constant $`𝒩`$ has been obtained in Ref. ,
$$𝒩=\frac{x_{11}}{k^2R^2}\sqrt{\frac{\mathrm{}}{2\pi ϵ_0L\omega (11/x_{11}^2)\mathrm{J}_1^2(x_{11})}},$$
(8)
from the requirement that the energy per one photon in a mode is equal to $`\mathrm{}\omega `$.
At the position $`𝐫_A=(0,0,L/2)`$ of the center of the atom the orthogonal field vectors $`𝐄^X`$ and $`𝐄^Y`$ point respectively in $`x`$ and $`y`$ direction and are given by simple formulas:
$`𝐄^X(𝐫_A)`$ $`=`$ $`{\displaystyle \frac{i𝒩\omega x_{11}}{2R}}𝐞_x,`$ (10)
$`𝐄^Y(𝐫_A)`$ $`=`$ $`{\displaystyle \frac{i𝒩\omega x_{11}}{2R}}𝐞_y,`$ (11)
$`𝐁^X(𝐫_A)`$ $`=`$ $`0,𝐁^Y(𝐫_A)=0.`$ (12)
The relevant part of the electric and magnetic field in the cavity can be written in the form
$`𝐄`$ $`=`$ $`𝐄^Xa_X+𝐄^Xa_X^{}+𝐄^Ya_Y+𝐄^Ya_Y^{},`$ (14)
$`𝐁`$ $`=`$ $`𝐁^Xa_X+𝐁^Xa_X^{}+𝐁^Ya_Y+𝐁^Ya_Y^{},`$ (15)
where $`a_X`$ and $`a_Y`$ are the dimensionless mode expansion amplitudes.
In the laboratory frame the dynamics of the atom-field system is governed by the Hamiltonian
$`H_L={\displaystyle \frac{𝐩^2}{2m}}{\displaystyle \frac{e^2}{4\pi ϵ_0r}}`$ $``$ $`e𝐫𝐄(𝐫_A)`$ (16)
$`+`$ $`{\displaystyle \frac{1}{2}}{\displaystyle (ϵ_0𝐄^2+𝐁^2/\mu _0)d^3𝐫},`$ (17)
where $`𝐫=(x,y,z)`$ is the position of the electron relative to the center of the atom $`𝐫_A`$. The Hamiltonian $`H_L`$ describes the mutual interaction of the atomic electron with the chosen cavity modes. We can rewrite $`H_L`$ using the amplitudes $`a_X`$ and $`a_Y`$, or more conveniently, using their real combinations:
$`P_x`$ $`=`$ $`{\displaystyle \frac{i}{\sqrt{2}}}(a_Xa_X^{}),P_y={\displaystyle \frac{i}{\sqrt{2}}}(a_Ya_Y^{}),`$ (19)
$`Q_x={\displaystyle \frac{1}{\sqrt{2}}}(a_X+a_X^{}),Q_y={\displaystyle \frac{1}{\sqrt{2}}}(a_Y+a_Y^{}),`$ (20)
where the dimensionless vectors $`𝐏`$ and $`𝐐`$ represent the electric field and the magnetic induction,
$`H_L={\displaystyle \frac{𝐩^2}{2m}}{\displaystyle \frac{e^2}{4\pi ϵ_0r}}e𝐫𝐏+{\displaystyle \frac{\mathrm{}\omega }{2}}(𝐏^2+𝐐^2).`$ (21)
The field amplitude $``$ is
$$=\frac{𝒩\omega x_{11}}{R\sqrt{2}}.$$
(22)
We have found it convenient to use the natural units for our problem derived from the field frequency for the energy, length, and momentum: $`\mathrm{}\omega ,\sqrt{\mathrm{}/m\omega },\sqrt{\mathrm{}m\omega }`$. The Hamiltonian (21) in these units takes on the following form
$`H_L={\displaystyle \frac{𝐩^2}{2}}{\displaystyle \frac{\stackrel{~}{q}}{r}}\gamma 𝐫𝐏+{\displaystyle \frac{𝐏^2+𝐐^2}{2}},`$ (23)
where the dimensionless parameters $`\stackrel{~}{q}`$ and $`\gamma `$ characterizing the strength of the Coulomb field and the atom-field coupling are
$`\stackrel{~}{q}={\displaystyle \frac{e^2}{4\pi ϵ_0\mathrm{}\omega }}\sqrt{{\displaystyle \frac{m\omega }{\mathrm{}}}},\gamma ={\displaystyle \frac{e}{\mathrm{}\omega }}\sqrt{{\displaystyle \frac{\mathrm{}}{m\omega }}}.`$ (24)
## III Classical solutions
We are interested in special solutions corresponding to the Trojan states in the external electromagnetic wave introduced in Ref. . Since these states describe electronic wave packets rotating around the nucleus along circular orbits, we transform the Hamiltonian $`H_L`$ to the frame rotating around the $`z`$-axis with the angular velocity $`\mathrm{\Omega }`$. The transformed Hamiltonian is
$`H={\displaystyle \frac{𝐩^2}{2}}{\displaystyle \frac{\stackrel{~}{q}}{r}}\gamma 𝐫𝐏+{\displaystyle \frac{𝐏^2+𝐐^2}{2}}\kappa (M_z^A+M_z^F),`$ (25)
where $`\kappa =\mathrm{\Omega }/\omega `$. The $`z`$-components of the angular momenta of the electron and of the electromagnetic field are $`M_z^A=xp_yyp_x`$ and $`M_z^F=(Q_xP_yQ_yP_x)`$. In this frame, the rotational states will appear as stationary states of the Hamiltonian. We would like to stress that the Hamiltonian $`H`$ cannot be identified with the energy because of the appearance of the inertial forces in the rotating frame.
To emphasize the rotational symmetry of our problem we introduce following variables for the electromagnetic field
$`Q_+={\displaystyle \frac{Q_xP_y}{\sqrt{2}}},Q_{}={\displaystyle \frac{Q_x+P_y}{\sqrt{2}}},`$ (27)
$`P_+={\displaystyle \frac{Q_y+P_x}{\sqrt{2}}},P_{}={\displaystyle \frac{P_xQ_y}{\sqrt{2}}},`$ (28)
corresponding to the left and right circular polarization. In terms of these variables the Hamiltonian takes on the form
$`H`$ $`=`$ $`{\displaystyle \frac{𝐩^2}{2}}{\displaystyle \frac{\stackrel{~}{q}}{r}}{\displaystyle \frac{\gamma }{\sqrt{2}}}\left(x(P_++P_{})+y(Q_{}Q_+)\right)`$ (29)
$`+`$ $`{\displaystyle \frac{1+\kappa }{2}}\left(P_+^2+Q_+^2\right)+{\displaystyle \frac{1\kappa }{2}}\left(P_{}^2+Q_{}^2\right)\kappa M_z^A.`$ (30)
(31)
The kinetic part of the field Hamiltonian is made up of two terms: co-rotating and counter-rotating. Linear stability analysis shows that both parts are necessary for the existence of the nontrivial equilibrium solution. From the Hamiltonian (25) we derive the evolution equations:
$$\dot{x}=p_x+\kappa y,$$
(33)
$$\dot{y}=p_y\kappa x,$$
(34)
$$\dot{z}=p_z,$$
(35)
$$\dot{Q}_+=(1+\kappa )P_+\gamma x/\sqrt{2},$$
(36)
$$\dot{Q}_{}=(1\kappa )P_{}\gamma x/\sqrt{2},$$
(37)
$$\dot{p}_x=\frac{\stackrel{~}{q}x}{r^3}+\frac{\gamma (P_++P_{})}{\sqrt{2}}+\kappa p_y,$$
(38)
$$\dot{p}_y=\frac{\stackrel{~}{q}y}{r^3}+\frac{\gamma (Q_{}Q_+)}{\sqrt{2}}\kappa p_x,$$
(39)
$$\dot{p}_z=\frac{\stackrel{~}{q}z}{r^3},$$
(40)
$$\dot{P}_+=(1+\kappa )Q_+\gamma y/\sqrt{2},$$
(41)
$$\dot{P}_{}=(1\kappa )Q_{}+\gamma y/\sqrt{2}.$$
(42)
The time-independent solutions of these equations describe the stationary states of our system. Equating the left hand side of Eqs. (III) to zero, we obtain
$$x^{eq}=r_0\mathrm{cos}\phi ,y^{eq}=r_0\mathrm{sin}\phi ,z^{eq}=0,$$
(44)
$$p_x^{eq}=\kappa r_0\mathrm{sin}\phi ,p_y^{eq}=\kappa r_0\mathrm{cos}\phi ,p_z^{eq}=0,$$
(45)
$$Q_+^{eq}=\frac{\gamma r_0\mathrm{sin}\phi }{(\kappa +1)\sqrt{2}},Q_{}^{eq}=\frac{\gamma r_0\mathrm{sin}\phi }{(\kappa 1)\sqrt{2}},$$
(46)
$$P_+^{eq}=\frac{\gamma r_0\mathrm{cos}\phi }{(\kappa +1)\sqrt{2}},P_{}^{eq}=\frac{\gamma r_0\mathrm{cos}\phi }{(\kappa 1)\sqrt{2}}.$$
(47)
In addition, the equations (38) and (39) give the equilibrium condition:
$`{\displaystyle \frac{\stackrel{~}{q}}{r_0^3}}=\kappa ^2{\displaystyle \frac{\gamma ^2}{\kappa ^21}}.`$ (48)
This equilibrium condition can be used to express the equilibrium radius $`r_0`$ in terms of the frequency of the cavity mode $`\omega `$ and the frequency of rotation $`\mathrm{\Omega }`$
$`r_0(\mathrm{\Omega })=\stackrel{~}{q}^{1/3}\left((\mathrm{\Omega }/\omega )^2{\displaystyle \frac{\gamma ^2}{(\mathrm{\Omega }/\omega )^21}}\right)^{1/3},`$ (49)
or, alternatively, to express the frequency of rotation in terms of $`\omega `$ and $`r_0`$. The equilibrium condition (48) has two solutions for $`\mathrm{\Omega }`$, denoted by $`\mathrm{\Omega }^>(r_0)`$ and $`\mathrm{\Omega }^<(r_0)`$,
$`\mathrm{\Omega }^>(r_0)={\displaystyle \frac{\omega }{\sqrt{2}}}\sqrt{1+\stackrel{~}{q}/r_0^3+\sqrt{(1\stackrel{~}{q}/r_0^3)^2+4\gamma ^2}},`$ (51)
$`\mathrm{\Omega }^<(r_0)={\displaystyle \frac{\omega }{\sqrt{2}}}\sqrt{1+\stackrel{~}{q}/r_0^3\sqrt{(1\stackrel{~}{q}/r_0^3)^2+4\gamma ^2}}.`$ (52)
Both solutions exist for all values of $`r_0`$. The solution $`\mathrm{\Omega }^>(r_0)`$ ($`\mathrm{\Omega }^>(r_0)`$) gives the frequency that is always higher (lower) than the cavity frequency $`\omega `$ (see Fig. 1). The higher frequency (larger centrifugal force) requires the electric field to be directed towards the nucleus, whereas the lower frequency requires the field pointing outwards. The first case corresponds to the Trojan states whereas the second to the so called anti-Trojan states. In the previous study , when the electromagnetic field has been treated as a given external wave, the anti-Trojan states were found to be classically unstable. The classical stability obtained in the present study is due to the detuning from the exact resonance. Since the localization of the electronic wave packet is much worse for the anti-Trojan states (classically, the trajectories in the rotating frame are spread almost evenly around the whole circle, cf. Fig. 2), we will restrict ourselves to the Trojan states only. Hence, in what follows, we shall only consider the solution $`\mathrm{\Omega }^>(r_0)`$.
Note, that for $`\omega =\mathrm{\Omega }`$ (i.e. $`\kappa =1`$), we have only a trivial result $`𝐫^{eq}=𝐩^{eq}=𝐏^{eq}=𝐐^{eq}=0`$ which in the classical model of an atom, means that “the electron has fallen onto the nucleus” and the electric field is zero. Thus, every nontrivial solution requires the presence of a detuning ($`\omega \mathrm{\Omega }`$) between the cavity frequency and the Kepler frequency. This phenomenon is known as the frequency pushing and is a direct consequence of the mutual atom-field interaction. This detuning has been absent in all previous approaches where the atom was driven by an external wave.
The equations (III) have a continuum of time-independent solutions that can be labeled by $`r_0`$ and the angle $`\phi `$ in the $`x`$-$`y`$ plane. These solutions describe in the laboratory frame a classical electron circulating around the nucleus at the distance $`r_0`$. The orbit of the electron is confined to the $`x`$-$`y`$ plane. The electron is dressed by the classical electromagnetic field
$$𝐄=\frac{\sqrt{2}}{\kappa ^21}\frac{er_0}{\mathrm{}\omega }(\mathrm{sin}\phi 𝐞_y+\mathrm{cos}\phi 𝐞_x),$$
(53)
which has a resonance dependence on the parameter $`\kappa `$. Note that the electric field changes its sign when the frequency of rotation passes through the resonance.
Next, we expand the Hamiltonian around a time-independent solution and investigate its linear stability. The motion will be stable if all eigenfrequencies are real. The characteristic equation for this problem has the form
$`\lambda ^2(\lambda ^2q_r)(\lambda ^6(4+q_r+2)\mathrm{\Gamma })\lambda ^4`$ (54)
$`+(53/5q_r+q_r^2/2+4\gamma ^2+(4+5/2q_r\mathrm{\Gamma })\lambda ^2`$ (55)
$`(2+5/2q_r5q_r^2+q_r^3+8\gamma ^2+14q_r\gamma ^2+(2+7/2q_r+q_r^2/2+8\gamma ^2)\mathrm{\Gamma })))=0,`$ (56)
where $`q_r=\stackrel{~}{q}/r_0^3`$ and $`\mathrm{\Gamma }=\sqrt{12q+q^2+4\gamma ^2}`$. The first ($`\lambda =0`$) frequency in our problem corresponds to the rotation of the whole system and it is a reflection of the rotational symmetry. The second frequency ($`\lambda =\sqrt{\stackrel{~}{q}/r_0^3}`$) corresponds to the motion in the $`z`$-direction that (in the linear approximation) is decoupled from the motion in the $`x`$-$`y`$ plane. The remaining 3 frequencies correspond to the motion of the electron coupled to the electromagnetic field. We shall not produce the analytical expressions for these eigenfrequencies but in Fig. 3 we plot the region of stability in the $`R`$-$`r_0`$ plane.
The stability can also be studied numerically and the calculations of the classical trajectories fully confirm the stability of the equilibrium solution. In Fig. 4 we plotted the projection of a typical electron trajectory on the $`x`$-$`y`$ plane for the time interval $`(1400T,1500T)`$, where $`T=1/\omega `$. The trajectory started at the equilibrium position $`𝐫=r_0(1,0,0)`$ with the initial momenta $`𝐩=m\omega r_0(0.02,\kappa +0.07,0.02)`$. As we see, the electron follows a rather complicated, but bounded, trajectory. Obviously, if we choose $`p(t=0)`$ sufficiently large the electron will eventually leave the vicinity of the equilibrium point. In Fig. 5 we show the $`z`$-$`p_z`$ cross-section of the phase space for the same trajectory. This phase-space trajectory resembles the trajectory of a simple harmonic oscillator. Indeed, as we have seen, in the linearized evolution equations, the motion in the $`z`$-direction is purely harmonic. Thus, the interesting dynamics of the electron is found in the motion confined to the $`x`$-$`y`$ plane and in what follows we shall treat our problem as two-dimensional.
Since our system is conservative, it has a well defined energy $`H_L`$. We have calculated its value $`E(\kappa )`$ for all those solutions that in the rotating frame are determined by Eqs. (42). This energy is given by the formula
$`E(\kappa )={\displaystyle \frac{m\omega ^2r_0^2(\kappa )}{2}}\left({\displaystyle \frac{\gamma ^2}{(\kappa ^21)^2}}(5\kappa ^23)\kappa ^2\right).`$ (57)
and is plotted in Fig. 6 as a function of $`\kappa `$. The infinite growth of the energy near the resonance ($`\kappa =1)`$ expresses the phenomenon of the frequency pushing.
## IV Quantum effects
In order to study the quantum effects for the electron as well as for the electromagnetic field we will apply the procedure of the quantization around the classical solution (42). A similar quantization method has been used before, for example in nonlinear optics to describe quantum fluctuations around the classical solitons in fibers . Here, the quantization will lead to the description of the electron in terms of a quantum mechanical wave packet orbiting along the classical trajectory and, at the same time, will reveal quantum fluctuations of the electromagnetic field around its classical value.
As a starting point we choose the Hamiltonian (29) in which all variables are treated as operators and we express them as sums of their classical parts and the quantum corrections $`\widehat{𝐫}=𝐫^{eq}+𝐫`$, $`\widehat{𝐩}=𝐩^{eq}+𝐩`$, $`\widehat{𝐐}=𝐐^{eq}+𝐐`$, $`\widehat{𝐏}=𝐏^{eq}+𝐏`$. The classical parts represent equilibrium solutions (42) found in the preceding Section. In order to simplify the notation, we have not attached any labels to the operators of quantum corrections $`(𝐫,𝐩,𝐐,𝐏)`$. Next, we expand the Hamiltonian around the classical equilibrium solution neglecting all terms higher then quadratic in the quantum corrections. To proceed along these lines, we have to choose one solution, labeled by $`\phi _0`$, from the whole family of equilibrium solutions. Making this choice we break the rotational symmetry of the Hamiltonian.
This mechanism of selection of a specific classical solution resembles the spontaneous symmetry breaking. Spontaneous symmetry breaking is present in many branches of physics. It explains the appearance of deformed nuclei, the formation of magnetic domains in ferromagnetic materials, or the emergence of Higgs particles in the Glashow-Weinberg-Salam model of electroweak interactions. In all these cases the symmetry is broken by the choice of a particular ground state. In our case, however, we do not break the symmetry by choosing a ground state but by choosing an equilibrium state of the Hamiltonian that is very far from the ground state of the system.
Once we have chosen some $`\phi _0`$, we can rotate the frame of reference, so that the direction given by $`\phi _0`$ is along the $`x`$ axis. The quadratic Hamiltonian is
$`H_Q`$ $`=`$ $`{\displaystyle \frac{𝐩^2}{2}}{\displaystyle \frac{\gamma }{\sqrt{2}}}[x(P_++P_{})+y(Q_{}Q_+)]`$ (58)
$`+`$ $`{\displaystyle \frac{1+\kappa }{2}}\left(P_+^2+Q_+^2\right)+{\displaystyle \frac{1\kappa }{2}}\left(P_{}^2+Q_{}^2\right)`$ (59)
$``$ $`q\kappa ^2x^2+{\displaystyle \frac{q\kappa ^2y^2}{2}}+{\displaystyle \frac{q\kappa ^2z^2}{2}}\kappa (xp_yxp_x),`$ (60)
where the parameter $`q`$, the ratio of the Coulomb force to the centrifugal force,
$`q={\displaystyle \frac{e^2}{4\pi ϵ_0mr_0^3\mathrm{\Omega }^2}}={\displaystyle \frac{\stackrel{~}{q}}{r_0^3\kappa ^2}},`$ (61)
has been introduced to achieve the full correspondence with the notation used before in the description of Trojan states. Note, that in this Hamiltonian the quadratic term $`q\kappa ^2x^2`$ enters with the negative coefficient. If it were not for the rotational term, such a Hamiltonian would not have any stable points. In our case, however, the stability can be achieved for a particular choice of $`\gamma `$, $`q`$ and $`\kappa `$ .
We look for the fundamental solution of the Schrödinger equation with the Hamiltonian $`H_Q`$ in the form of a four-dimensional Gaussian function
$$\psi =N\mathrm{exp}\left(\frac{1}{2}𝐗A𝐗\right),$$
(62)
$`𝐗=(x,y,Q_+,Q_{})`$, and
$`A=\left(\begin{array}{cccc}a_{11}& ia_{12}& ia_{13}& ia_{14}\\ ia_{12}& a_{22}& a_{23}& a_{24}\\ ia_{13}& a_{23}& a_{33}& a_{34}\\ ia_{14}& a_{23}& a_{34}& a_{44}\end{array}\right)`$ (67)
Inserting this ansatz into the Schrödinger equation we obtain 10 algebraic, nonlinear equations for the parameters $`a_{ij}`$:
$$2\kappa ^2qa_{11}^2+2\kappa a_{12}+a_{12}^22\gamma a_{13}+a_{13}^2(1+\kappa )2\gamma a_{14}+a_{14}^2(1\kappa )=0,$$
(69)
$$a_{11}a_{12}+a_{12}a_{22}\gamma a_{23}+a_{13}a_{23}\gamma a_{24}+a_{14}a_{24}+\kappa \left(a_{11}+a_{22}+a_{13}a_{23}a_{14}a_{24}\right)=0,$$
(70)
$$a_{11}a_{13}+a_{12}a_{23}\gamma a_{33}+a_{13}a_{33}\gamma a_{34}+a_{14}a_{34}+\kappa \left(a_{23}+a_{13}a_{33}a_{14}a_{34}\right)=0,$$
(71)
$$a_{11}a_{14}+a_{12}a_{24}\gamma a_{34}+a_{13}a_{34}\gamma a_{44}+a_{14}a_{44}+\kappa \left(a_{24}+a_{13}a_{34}a_{14}a_{44}\right)=0,$$
(72)
$$\kappa ^2q2\kappa a_{12}+a_{12}^2a_{22}^2a_{23}^2(1+\kappa )a_{24}^2(1\kappa )=0,$$
(73)
$$\gamma \kappa a_{13}+a_{12}a_{13}a_{22}a_{23}a_{23}a_{33}\kappa a_{23}a_{33}a_{24}a_{34}+\kappa a_{24}a_{34}=0,$$
(74)
$$\gamma \kappa a_{14}+a_{12}a_{14}a_{22}a_{24}a_{23}a_{34}\kappa a_{23}a_{34}a_{24}a_{44}+\kappa a_{24}a_{44}=0,$$
(75)
$$1+\kappa +a_{13}^2a_{23}^2a_{33}^2(1+\kappa )a_{34}^2(1\kappa )=0,$$
(76)
$$a_{13}a_{14}a_{23}a_{24}a_{34}\left(\left(1+\kappa \right)a_{33}\left(1+\kappa \right)a_{44}\right)=0,$$
(77)
$$1\kappa +a_{14}^2a_{24}^2a_{44}^2(1\kappa )a_{34}^2(1+\kappa )=0.$$
(78)
We can easily solve these equations numerically, but first we want to find a perturbative solution. In order to do that we write the coupling constant in the form $`\gamma =\overline{\gamma }\sqrt{\kappa 1}`$. Obviously $`\overline{\gamma }=\kappa \sqrt{(1q)(\kappa +1)}`$, and we will treat $`\overline{\gamma }`$ as a small parameter. Typical values of the parameters are $`\kappa =1.0000001,q=0.95625`$ which give $`\overline{\gamma }=0.06`$. One can ask why we can not treat $`\gamma `$ (or even simpler, $`\kappa 1`$) as a perturbation parameter. However, if we do so we face a problem: the coefficients of the perturbation series are growing, since they behave as $`1/\sqrt{\kappa 1}`$. When we tend with $`\kappa 1`$ to zero we hit exactly the resonance point and the perturbation expansion becomes meaningless. On the other hand, when $`\overline{\gamma }`$ is chosen as an expansion parameter, all large contributions to the coefficients in the perturbation expansion $`a_{ij}=a_{ij}^{(0)}+\overline{\gamma }a_{ij}^{(1)}+\overline{\gamma }^2a_{ij}^{(2)}+\mathrm{}`$ cancel out.
We calculated analytically the coefficients up to the second order but we present here the analytic formulas only in the zeroth order and numerical values of the first and the second order corrections.
$`a_{11}^{(0)}`$ $`=`$ $`\kappa \sqrt{{\displaystyle \frac{(1+2q)(4q9q+88s(q))}{9q^2}}},`$ (80)
$`a_{12}^{(0)}`$ $`=`$ $`\kappa {\displaystyle \frac{2+q2s(q)}{3q}},`$ (81)
$`a_{22}^{(0)}`$ $`=`$ $`\kappa \sqrt{{\displaystyle \frac{(1q)(4q9q+88s(q))}{9q}}},`$ (82)
where $`s(q)=\sqrt{1+q2q}`$,
$`a_{13}^{(0)}`$ $`=`$ $`0,a_{23}^{(0)}=0,a_{14}^{(0)}=0,a_{24}^{(0)}=0,`$ (83)
$`a_{33}^{(0)}`$ $`=`$ $`1,a_{34}^{(0)}=0,a_{44}^{(0)}=1.`$ (84)
Thus, in the zeroth order, the electronic part of the wave packet is exactly the same as in the case of externally driven Trojan wave packet . The electromagnetic part has a form of a coherent (nonsqueezed) state.
Higher corrections are due to the mutual interaction between the field and the atom. Numerical values of the parameters $`a_{ij}`$ are calculated for the cavity parameters $`L=1`$ cm, $`R=0.32`$ cm, which give $`\omega =197`$ GHz and $`\gamma =3.24\times 10^7`$. The detuning $`\kappa `$ is chosen in such a way, that the value of $`q`$ is optimal, $`q=0.95625`$. As shown in Ref. , the wave packet is then maximally concentrated around the equilibrium point and its center is located at $`r_0=3600a_0`$ ($`a_0`$ is the atom Bohr radius). The expansion coefficients calculated up to the second order are presented in Table 1. In this order we observe the effect of the back reaction of the electron on the electromagnetic field. However, the coefficients $`a_{11},a_{12}`$, and $`a_{22}`$ characterizing the shape of the electronic wave packet are the same as in the zeroth order within the assumed accuracy.
Finally, we present in Table 2 the results of a direct numerical solution of our set of equations. As we see, almost all the coefficients have been obtained correctly already in the second order of perturbation theory. Only the $`a_{44}`$ differs from the exact numerical solution. This can be attributed to the very slowly convergent perturbation series for this particular coefficient. The coefficients $`a_{13},a_{23},a_{14}`$ and $`a_{24}`$ describing the mixing of the atomic part of the wave packet with the field part are not zero. Moreover, the electromagnetic field is strongly affected by the interaction with the atom; the coefficient $`a_{44}`$ is significantly different from its value in the zeroth order.
The four-dimensional Gaussian wave packet (62) with the coefficients $`a_{ij}`$ calculated numerically describes the fundamental state of the mutually interacting atom-field system. Owing to its Gaussian form, this state saturates the multidimensional generalized uncertainty relations for the complete atom-field system (see ). The smallness of the coefficients $`a_{13},a_{23},a_{14}`$ and $`a_{24}`$ expresses the fact that the field and the atom are only very weakly correlated in this state. As a result of this, the saturation of the uncertainty relations is almost exact, separately for both parts of the wave function. The average values of second moments the electronic variables calculated with the numerical values of the coefficients taken from the Table 2 are given in Table 3. This Table exhibits the existence of correlations between the variables in the $`x`$ and the $`y`$ directions. This requires the use of generalized uncertainty relations for a two-dimensional system in the form (cf. Ref.)
$`xxp_xp_x+xyp_xp_y`$ (85)
$``$ $`{\displaystyle \frac{1}{4}}(xp_x+p_xx^2+xp_y+p_yxp_xy+yp_x){\displaystyle \frac{\mathrm{}^2}{4}},`$ (87)
$`yyp_yp_y+xyp_xp_y`$
$``$ $`{\displaystyle \frac{1}{4}}(yp_y+p_yy^2+yp_x+p_xyp_yx+xp_y){\displaystyle \frac{\mathrm{}^2}{4}}.`$ (88)
Upon substituting the values taken from the Table 3 we find an almost exact saturation of these relations.
Now, we turn to the description of the quantum correlations for the electromagnetic field in our fundamental state of the atom-field system. The second moments for the field variables are given in the Table 4. These values of the correlations imply an almost complete decoupling between the co-rotating and counter-rotating modes so that the uncertainty is almost saturated separately for each mode. The state of the field in the counter-rotating mode is a coherent state but the state in the co-rotating mode is highly squeezed; the ratio of the correlation for the two quadratures $`Q_{}`$ and $`P_{}`$ is about $`3.5\times 10^4`$. However, the fluctuations of the field are still small as compared to the field value $`P_{}^{eq}=1.5\times 10^6`$. The plots of the Wigner function in Figs. 7 and 8 illustrates the difference in quantum fluctuations between the counter-rotating and the co-rotating modes.
## V Discussion
We have shown that the dynamical treatment of the relevant modes of the electromagnetic field enables one to reproduce exactly the properties of the Trojan states studied previously in the presence of a given, external wave. However, there appear new features totally absent in the previous studies. First, the back reaction of the electron on the electromagnetic field causes the detuning from the exact resonance. As a result, the stability region covers now also the anti-Trojan states, that were before found to be classically unstable. Second, our analysis has shown that to achieve the equilibrium state of the mutually interacting atom-field system we must take into account both polarization modes of the field: co-rotating and counter-rotating. The inclusion of only the co-rotating mode, as proposed in Ref. , is not sufficient to achieve an equilibrium state.
The quantization procedure adopted by us in this work consists in quantizing the corrections to the classical solution. These quantum corrections describe the shape of the electronic wave packet and the quantum fluctuations of the electromagnetic field around its classical value. The field fluctuations turn out to be significantly different for the two modes: for the counter-rotating mode the fluctuations are as for the vacuum state whereas for the co-rotating mode they exhibit strong squeezing.
The choice of one particular solution from the class of equivalent classical solutions spontaneously breaks the rotational symmetry of the initial Hamiltonian. The method of quantization around the classical solution used here can be also applied to a similar problem of electronic Trojan states in a polar molecule . In this case, the role of the electromagnetic field is played by the rotating molecular dipole. The application of our method would require the dynamical treatment of the relevant molecular degrees of freedom.
## Acknowledgments
We would like to thank Jan Mostowski for fruitful discussions and we acknowledge the support from the KBN Grant P03B0313.
|
warning/0007/hep-ph0007280.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The top quark, with a mass of the order of the electroweak symmetry breaking scale, is naturally considered to be related to new physics. Run 1 of the Fermilab Tevatron has small statistics on top quark events and thus leaves plenty of room for new physics to be discovered at the upgraded Tevatron in the near future. Due to higher statistics, the $`t\overline{t}`$ events at the upgraded Tevatron are expected to provide sensitive probes for new physics . The most popular model for new physics is the Minimal Supersymmetric Model (MSSM) . In this model, $`R`$-parity , defined by $`R=(1)^{2S+3B+L}`$ with spin $`S`$, baryon-number $`B`$ and lepton-number $`L`$, is often imposed on the Lagrangian to maintain the separate conservation of $`B`$ and $`L`$. As a consequence the sparticle number is conserved. Since instanton effects induce miniscule violations of baryon and lepton number , $`R`$-parity conservation is not dictated by any known fundamental principle such as gauge invariance or renormalizability. If $`R`$-parity is strictly conserved, it is conceivable that the conservation comes from some hitherto unidentified fundamental principle. Hence $`R`$-parity violation should be vigorously searched for.
The most general superpotential of the MSSM, consistent with the $`SU(3)\times SU(2)\times U(1)`$ symmetry, supersymmetry, and renomalizability also contains $`R`$-violating interactions which are given by
$`𝒲_\overline{)}R={\displaystyle \frac{1}{2}}\lambda _{ijk}L_iL_jE_k^c+\lambda _{ijk}^{}\delta ^{\alpha \beta }L_iQ_{j\alpha }D_{k\beta }^c+{\displaystyle \frac{1}{2}}\lambda _{ijk}^{\prime \prime }ϵ^{\alpha \beta \gamma }U_{i\alpha }^cD_{j\beta }^cD_{k\gamma }^c+\mu _iL_iH_2,`$ (1)
where $`L_i(Q_i)`$ and $`E_i(U_i,D_i)`$ are the left-handed lepton (quark) doublet and right-handed lepton (quark) singlet chiral superfields. The indicies $`i,j,k`$ are generation indices, $`\alpha `$, $`\beta `$ and $`\gamma `$ are the color indices, $`c`$ denotes charge conjugation, and $`ϵ^{\alpha \beta \gamma }`$ is the total antisymmetric tensor in three-dimension. $`H_{1,2}`$ are the Higgs-doublets chiral superfields. The coefficients $`\lambda `$ and $`\lambda ^{}`$ are the coupling strengths of the $`L`$-violating interactions and $`\lambda ^{\prime \prime }`$ those of the $`B`$-violating interactions. The lower bound of the proton lifetime imposes very strong conditions on the simultaneous presence of both $`L`$-violating and $`B`$-violating interactions and hence the strength of the couplings. However, the existence of either $`L`$-violating or $`B`$-violating couplings, but not both at the same time, does not induce nucleon decays and therefore the $`R`$-parity violating couplings are less constrained. This separate $`L`$ and $`B`$ violation is usually assumed in phenomenological analyses.
The study of the phenomenology of R-violating supersymmetry was started many years ago . Some constraints on the $`R`$-parity violating couplings have been obtained from various analyses, such as perturbative unitarity , $`n\overline{n}`$ oscillation , $`\nu _e`$-Majorana mass , neutrino-less double $`\beta `$ decay , charged current universality , $`e\mu \tau `$ universality , $`\nu _\mu e`$ scattering , atomic parity violation , $`\nu _\mu `$ deep-inelastic scattering , $`\mu e`$ conversion , $`K`$-decay , $`\tau `$-decay , $`D`$-decay , $`B`$-decay and $`Z`$-decay at LEP I . As reviewed in Ref. , although many such couplings have been severely constrained, the bounds on the top quark couplings are generally quite weak. This is the motivation for the phenomenological study of R-violation in processes involving the top quark.
The production mechanisms of top pairs and single top in $`R`$-violating SUSY at the upgraded Tevatron have been examined in and , respectively. In addition, the $`R`$-violating couplings can induce exotic decays for top quark at an observable level. For example, the top quark FCNC decays induced by $`R`$-violating couplings can be significantly larger than those in the MSSM with R-parity conservation . If we allow the co-existence of two $`\lambda ^{}`$ couplings, we have the new decay modes, such as $`t\mathrm{}\stackrel{~}{d}\mathrm{}^+\mathrm{}^{}u`$ . The bilinear term $`\mu _iL_iH_2`$ can also induce some new decays for the top quark, as studied in .
In this work, we focus on the explicit trilinear couplings and assume only one trilinear coupling exists at one time. Then the possible exotic top decay modes are
$`t\stackrel{~}{\overline{d}_i}\overline{d}_j,\stackrel{~}{\overline{d}}_j\overline{d}_i\overline{d}_i\overline{d}_j\stackrel{~}{\chi }_1^0`$ (2)
induced by the $`B`$-violating $`\lambda _{3ij}^{\prime \prime }`$, and
$`te_i^+\stackrel{~}{d}_j\stackrel{~}{e}_id_je_i^+d_j\stackrel{~}{\chi }_1^0`$ (3)
induced by the $`L`$-violating $`\lambda _{i3j}^{}`$. Here the subscripts $`i,j`$ are family indices and $`\stackrel{~}{\chi }_1^0`$ is the lightest neutralino which, in our analysis, is assumed to be the lightest super particle (LSP) as favored in the MSSM where the SUSY breaking is propagated to the matter sector by gravity<sup>1</sup><sup>1</sup>1If the SUSY breaking is mediated by gauge interactions, the LSP is expected to be the gravitino.. The sfermions involved in these decays can be on-shell or virtual, depending on the masses of the particles involved.
Among the exotic decays in (2) and (3), the relatively easy-to-detect modes are those induced by $`\lambda _{33j}^{\prime \prime }`$ ($`j=1,2`$)<sup>2</sup><sup>2</sup>2$`\lambda _{333}^{\prime \prime }`$ does not exist since $`\lambda _{ijk}^{\prime \prime }`$ is antisymmetric in the last two indices. and $`\lambda _{i33}^{}`$ ($`i=1,2,3`$) because their final states contain a $`b`$-quark which can be tagged. One of the $`L`$-violating channels, i.e., $`t\stackrel{~}{\tau }b(\mathrm{or}\tau \stackrel{~}{b})`$ induced by $`\lambda _{333}^{}`$ has been studied in . So in our analysis we focus on the cases of $`\lambda _{133}^{}`$ and $`\lambda _{233}^{}`$ for $`L`$-violating couplings, and $`\lambda _{331}^{\prime \prime }`$ and $`\lambda _{332}^{\prime \prime }`$ for $`B`$-violating couplings. Since the decay induced by $`\lambda _{133}^{}`$ has the similar final states to that induced by $`\lambda _{233}^{}`$, we take the presence of $`\lambda _{233}^{}`$ as an example. For the same reason, we take the presence of $`\lambda _{331}^{\prime \prime }`$ as an example in $`B`$-violating case. The Feynman diagrams for these two decays induced by $`\lambda _{331}^{\prime \prime }`$ and $`\lambda _{233}^{}`$ are shown in Figs. 1 and 2, respectively.
In our analysis we consider $`t\overline{t}`$ events where one ($`t`$ or $`\overline{t}`$) decays via R-violating coupling while the other decays by the SM interaction. The SM decay will serve as the tag of the $`\overline{t}t`$ event. Furthermore, the penalty of the suppressed $`R`$-violation coupling is paid only once. Top spin correlations are taken into account in our calculation.
Note that the LSP ($`\stackrel{~}{\chi }_1^0`$) is no longer stable when R-parity is violated. In case just one R-violating top quark coupling does not vanish, the lifetime of the LSP will be very long, depending the coupling and the masses of squarks involved in the LSP decay chain (cf. the last paper of ). Thus it is generally assumes that the LSP decays outside the detector . We will make this assumption in our analysis. This paper is orgnized as follows. In Sec. 2, we investigate the potential of observing the $`B`$-violating top quark decay at the Tevatron and LHC, and present numerical results. In Sec. 3 we present similar results for $`L`$-violating decay. Finally in Sec. 4 we present a summary and discussion.
## 2 Searching for $`B`$-violating decay
### 2.1 Signal and background
To probe the decay $`t\overline{b}\overline{d}\stackrel{~}{\chi }_1^0`$ in Fig. 1, we consider the final states given by $`t\overline{t}`$ production where one (say $`t`$) decays via the coupling $`\lambda _{331}^{\prime \prime }`$ while the other (say $`\overline{t}`$ has the SM decays to serve as the tag of the $`\overline{t}t`$ event. Due to the large QCD background at hadron colliders, we do not search for the all-jets channel despite of its higher rate. Instead, we search for the signal given by $`t\overline{t}`$ events followed by $`t\overline{b}\overline{d}\stackrel{~}{\chi }_1^0`$ and $`\overline{t}W^{}\overline{b}\mathrm{}\overline{\nu }\overline{b}`$ ($`\mathrm{}=e,\mu `$). Then the signature is a lepton, three jets containing two b-jets or two $`\overline{b}`$-jets, and missing energy ($`\mathrm{}+3j/2b+\overline{)E_T}`$). We require that two $`b`$-jets are tagged in the signal. The efficiency for double b-tagging is assumed to be $`42\%`$ .
Note that the present events have the unique signal of the two same sign b-quarks. In our analysis, to be conservative, we assume that the tagging can not distinguish a b-quark jet from $`\overline{b}`$-quark jet. Then the SM backgrounds are mainly from
* $`t\overline{t}W^{}W^+b\overline{b}`$ followed by $`W^{}\mathrm{}\overline{\nu }`$ ($`\mathrm{}=e`$,$`\mu `$) and $`W^+\tau ^+\nu `$ with the $`\tau `$ decaying into a jet plus a neutrino;
* $`t\overline{t}W^{}W^+b\overline{b}`$ followed by $`W^{}\mathrm{}\nu `$ ($`\mathrm{}=e`$,$`\mu `$) and $`W^+q\overline{q}^{}`$. This process contains an extra quark jet and can only mimic our signal if the quark misses detection by going into the beam pipe. We assume this can only happen when the light quark jet has the pseudo-rapidity greater than about 3 or the transverse momentum less than about 10 GeV.
* $`Wb\overline{b}j`$ which includes single top quark production via the quark-gluon process $`qgq^{}t\overline{b}`$ as well as non-top processes .
### 2.2 Numerical calculation and results
We calculated the signal and background cross sections with the CTEQ5L structure functions . We assume $`M_t=175`$ GeV and take $`\sqrt{s}=2`$ TeV for the upgraded Tevatron and $`\sqrt{s}=14`$ TeV for the LHC.
As shown in Fig. 1, there are two contributing graphs. Since among the down-type squarks the sbottom is most likely to be lighter than other squarks (we will elaborate on this later), we assume the first graph in Fig.1 gives the dominant contribution. (If the $`\stackrel{~}{d}`$ is as light as the sbottom, the second diagram in Fig.1 has to be taken into account. Then our results for the signal rate should be quadrupled. To be conservative, we do not consider this case.)
For the total width of the sbottom involved in our calculation, we note that since only a light sbottom is meaningful to our analysis (as will be shown in our results), its dominant (or maybe the only) decay mode is $`\stackrel{~}{b}b\stackrel{~}{\chi }_1^0`$. The charged current decay mode $`\stackrel{~}{b}t\stackrel{~}{\chi }_1^+`$ is kinematically forbidden for a light sbottom in our analysis. We do not consider the strong decay mode $`\stackrel{~}{b}b\stackrel{~}{g}`$ since the gluino $`\stackrel{~}{g}`$ is likely to be heavy .
The signal cross section is proportional to $`|\lambda _{331}^{\prime \prime }|^2`$. We will present the signal results normalized to $`|\lambda _{331}^{\prime \prime }|^2`$. The signal cross section is very sensitive to the sbottom mass. We will vary it to see how heavy it can be for the signal to be observable. Other SUSY parameters involved are the lightest neutralino mass and its coupling to sbottom, which are determined by the parameters $`M,M^{},\mu `$ and $`\mathrm{tan}\beta `$. $`M`$ is the $`SU(2)`$ gaugino mass and $`M^{}`$ is the hypercharge $`U(1)`$ gaugino mass. $`\mu `$ is the Higgs mixing term ($`\mu H_1H_2`$) in the superpotential. $`\mathrm{tan}\beta =v_2/v_1`$ is the ratio of the vacuum expectation values of the two Higgs doublets. We work in the framework of the general MSSM. But we assume the grand unification of the gaugino masses, which gives the relation $`M^{}=\frac{5}{3}M\mathrm{tan}^2\theta _W0.5M`$. Then for the three independent parameters $`M,\mu `$ and $`\mathrm{tan}\beta `$, we choose a representative set of values
$`M=100\mathrm{GeV},\mu =200\mathrm{GeV},\mathrm{tan}\beta =1.`$ (4)
They yield the lightest chargino and neutralino masses as $`m_{\stackrel{~}{\chi }_1^+}=120`$ GeV , $`m_{\stackrel{~}{\chi }_1^0}=55`$ GeV. Thus this set of values are allowed by the current experimental bounds on the chargino and neutralino masses, which are about 90 GeV and 45 GeV, respectively .
We simulate detector effects by assuming Gaussian smearing of the energy of the charged final state particles, given by:
$`\mathrm{\Delta }E/E`$ $`=`$ $`30\%/\sqrt{E}1\%,\mathrm{for}\mathrm{leptons},`$ (5)
$`=`$ $`80\%/\sqrt{E}5\%,\mathrm{for}\mathrm{hadrons},`$ (6)
where $``$ indicates that the energy dependent and independent terms are added in quadrature and $`E`$ is in GeV.
The basic selection cuts are chosen as
$`p_T^{\mathrm{}},p_T^{jet},p_T^{\mathrm{miss}}`$ $``$ $`20\mathrm{GeV},`$ (7)
$`\eta _{jet},\eta _{\mathrm{}}`$ $``$ $`2.5,`$ (8)
$`\mathrm{\Delta }R_{jj},\mathrm{\Delta }R_j\mathrm{}`$ $``$ $`0.5.`$ (9)
Here $`p_T`$ denotes transverse momentum, $`\eta `$ is the pseudo-rapidity, and $`\mathrm{\Delta }R`$ is the separation in the azimuthal angle-pseudo rapidity plane $`(\mathrm{\Delta }R=\sqrt{(\mathrm{\Delta }\varphi )^2+(\mathrm{\Delta }\eta )^2})`$ between a jet and a lepton or between two jets.
We notice that for the background process (2) and (3) the missing energy comes only from the neutrino of the W decay, while for the signal events the missing energy contains an extra neutralino. From the transverse momentum of the lepton $`\stackrel{}{P}_T^{\mathrm{}}`$ and the missing transverse momentum $`\stackrel{}{P}_T^{\mathrm{miss}}`$, we construct the transverse mass as
$$m_T(\mathrm{},p_T^{\mathrm{miss}})=\sqrt{(|\stackrel{}{P}_T^{\mathrm{}}|+|\stackrel{}{P}_T^{\mathrm{miss}}|)^2(\stackrel{}{P}_T^{\mathrm{}}+\stackrel{}{P}_T^{\mathrm{miss}})^2}.$$
(10)
As is well-known, if the two components, i.e., $`\mathrm{}`$ and $`p_T^{\mathrm{miss}}`$ in our case, are from the decay of a parent particle, the transverse mass is bound by the mass of the parent particle. So for $`Wb\overline{b}j`$ background events $`m_T(\mathrm{},p_T^{\mathrm{miss}})`$ is always less than $`M_W`$ and peaks just below $`M_W`$. However, kinematic smearings can push the bound and the peak above $`M_W`$. In order to substantially suppress the large backgrounds (2) and (3) we apply the following cut
$$m_T(\mathrm{},p_T^{\mathrm{miss}})>120\mathrm{GeV}.$$
(11)
We found that the above strong $`m_T(\mathrm{},p_T^{\mathrm{miss}})`$ cut suppresses the background process (2) and (3) by roughly three orders of magnitude for the smearing in Eqs. (5) and (6), so that they are much smaller than the other backgrounds we are considering. But since background process (1) contains three neutrinos from different parent particles, it is not supressed by the $`m_T(\mathrm{},p_T^{\mathrm{miss}})`$ cut to a negligible level. There is some model dependence involved the treatment of $`\tau `$ hadronization. To avoid having to consider each of the many hadronic decay modes separately, we assume the invariant mass of the outgoing hadrons to be distributed uniformly from $`m_\pi `$ to $`m_\tau `$. Furthermore, we assume a uniform angular distribution in the phase allowed by the invariant mass of the outgoing jet. This assumption is probably reasonable in light of the fact that the parent $`\tau `$ is heavily boosted in the lab frame.
With the above selection cuts, the signal and background cross sections are given in Table 1. We see that the signal-to-background ratio can be quite large for light sbottom mass ($`<\mathrm{\hspace{0.33em}160}`$ GeV), in which the intermediate sbottom can be materialized as a real particle. When the sbottom becomes heavier than the top quark and thus can only appear as a virtual state, the cross section is severely suppressed by the small branching ratio of the decay.
From the results for Tevatron (1.8 TeV) in Table 1 we conclude that the luminosity Run 1 (0.1 fb<sup>-1</sup>) is too low to detect such decays. However, due to the much larger statistics of Run 2 (2 fb<sup>-1</sup>) and Run 3 (30 fb<sup>-1</sup>), it is possible to observe such decays in these coming runs of the Tevatron. Using the discovery criteria $`S5\sqrt{B}`$, the discovery limits of $`\lambda _{33j}^{\prime \prime }`$ versus the sbottom mass at Run 2, Run 3 (30 fb<sup>-1</sup>) and LHC (10 fb<sup>-1</sup>) are plotted in Fig. 3. The region above each curve is the corresponding region of discovery. Since the current bounds on $`\lambda _{331}^{\prime \prime }`$ from the LEP I Z-decay are of $`𝒪(1)`$ for sfermion mass heavier than 100 GeV , we see that for a light sbottom, we have a good chance to observe such decays if $`\lambda _{331}^{\prime \prime }`$ is not far below its current upper bounds. In case of nonobservation, meaningful bounds at 95% C.L. can be set, as shown in Fig.4.
Our results for $`\lambda _{331}^{\prime \prime }`$ can be applied to the case of $`\lambda _{332}^{\prime \prime }`$. Since the current bound from $`Z`$-decay is the same on both couplings , our conclusions on $`\lambda _{331}^{\prime \prime }`$ can be applied to the case of $`\lambda _{332}^{\prime \prime }`$.
## 3 Searching for $`L`$-violating decay
For $`L`$-violating decay $`t\mu ^+b\stackrel{~}{\chi }_1^0`$, there are two contributing graphs, as shown in Fig.2. The first graph proceeds through exchanging a sbottom while the second through exchanging a slepton. As in Sec. 2, we assume sbottom can be light and thus concentrate on the first graph. In the opposite case that the slepton is light and sbottom is heavy, our following results still hold with the replacement of sbottom mass by slepton mass. If both sbottom and slepton are light and approximately degenerate (which is quite unlikely in the supergravity scenario of supersymmetry breaking, as will be elaborated on later), then our results for the signal rate should be quadrupled.
Our examination for this decay is similar to the $`B`$-violating decay in the preceding section. We search for the signal given by $`t\overline{t}`$ events where one (say $`t`$) decays via $`L`$-violating coupling, $`t\mu ^+b\stackrel{~}{\chi }_1^0`$, while the other ($`\overline{t}`$) has the SM decays, $`\overline{t}W^{}\overline{b}`$. Then there are two possible observing channels for such an event: dilepton+2-jets and single lepton+4-jets, all being associated with missing energies. The dilepton channel has the lower rate and it is difficult to find a mechanism to enhance the S/B rate so as to find the ”smoking gun” for the signal. So we search for the single lepton+4-jets channel which has a higher rate. As is shown below, we can find effective selection cuts to enhance the S/B ratio for this signal.
Among the four jets in our signal there are two b-jets (one is $`b`$, the other is $`\overline{b}`$). We require that at least one b-jet passes b-tagging. The tagging efficiency is 53% at Run 1 and expected to reach 85% at Run 2 and Run 3 . For the LHC we assume the tagging efficiency to be the same as the Tevatron Run 2.
So the signature is $`\mathrm{}+4j/b+\overline{)E_T}`$ where $`4j/b`$ represents a 4-jets event with at least one of the jets passing the $`b`$-tagging criterion. This is the same as one of the typical signatures for $`t\overline{t}`$ event in the SM, except for the different source of missing energy. To suppress the QCD background, we apply the basic selection cuts in Eqs.(7-9). Under the basic selection cuts the QCD background is reduced to about 1/12 of the SM $`t\overline{t}`$ events . However, under the basic selection cuts the number of SM $`t\overline{t}`$ events far surpasses the number of signal events. In order to extract the signal events, we turn to the transverse mass defined in Eq.(10). For the SM $`t\overline{t}`$ events and $`W`$+jets background events the missing energy comes from the neutrino of the W decay, while for the signal events the missing energy comes from the neutralino in the decay $`t\mu ^+\stackrel{~}{b}\mu ^+b\stackrel{~}{\chi }_1^0`$. Thus the transverse mass distributions of the SM background and the signal events are different, as shown in Fig.5. In order to enhance the S/B ratio, we apply the following cut, taking into account of the smearing effect,
$$m_T(\mathrm{},p_T^{\mathrm{miss}})50100\mathrm{GeV}.$$
(12)
Other details in the numerical calculation, such as the smearing of the energy of the final state particles and the choice of SUSY parameters, are the same as in Sec. 2. In Table 2 we present the signal cross section for sbottom mass of 150 GeV, with the comparison to the SM $`t\overline{t}`$ background. One sees that the transverse mass cut can enhance the S/B ratio significantly. With the increase of sbottom mass, the signal cross section drops rapidly, as shown in Table 3.
From Tables 2 and 3 one sees that Run 1 (0.1 fb<sup>-1</sup>) of the Tevatron collider is unable to detect such decays for a sbottom heavier than 150 GeV and $`\lambda _{233}^{}<1`$. The possibility of observing such a decay is enhanced at Run 2 (2 fb<sup>-1</sup>), Run 3 (30 fb<sup>-1</sup>) and the LHC. Under the discovery criteria $`S5\sqrt{B}`$, the discovery limits of $`\lambda _{233}^{}`$ versus sbottom mass are plotted in Fig.6. The nonobservation of a signal is translated to the bounds (at 95% C.L.) shown in Fig.7.
Since the current bounds on $`\lambda _{233}^{}`$ from the LEP I Z-decay are of $`𝒪(1)`$ for sfermion mass heavier than 100 GeV , the results in Figs.6 and 7 indicate that the future runs at the upgraded Tevatron and LHC could either reveal the exotic decay or set stronger constraints on the $`L`$-violating coupling $`\lambda _{233}^{}`$.
Our results for $`\lambda _{233}^{}`$ can be applied to the case of $`\lambda _{133}^{}`$. But for $`\lambda _{133}^{}`$ the current bound from the $`\nu _e`$-mass, i.e., $`\lambda _{133}^{}<0.0007`$ at the $`1\sigma `$ level , is too strong, which makes the corresponding decay $`te^+b\stackrel{~}{\chi }_1^0`$ unobservable.
## 4 Summary and discussion
We have examined the potential for the detection of top quark decays via R-violating SUSY interactions at the Fermilab Tevatron and LHC. We studied two representative decay processes: one is induced by the $`B`$-violating coupling $`\lambda _{331}^{\prime \prime }`$ and the other is induced by the $`L`$-violating coupling $`\lambda _{233}^{}`$. Both of them have a $`b`$-jet in their decay products and can proceed through the intermediate sbottom which was assumed to be light. For the $`B`$-violating decay we searched for the signal $`\mathrm{}+3j/2b+\overline{)E_T}`$ given by $`t\overline{t}`$ events, while for the $`L`$-violating decay we searched for the channel $`\mathrm{}+4j/b+\overline{)E_T}`$. We considered the possible backgrounds and performed a Monte Carlo simulation by applying suitable cuts.
The signal cross section is found to drop drastically with the increase of the intermediate sbottom mass. If the sbottom could be as light as $`160`$ GeV, then under the current bounds of the relevant R-violating couplings, these decays can be detectable at the future runs of the Tevatron and LHC. However, because of the small statistics, Run 1 of the Tevatron will not be adequate.
A few remarks are due regarding our results:
* The results are sensitive to the sbottom mass; the signal is observable only for a light sbottom. The possibility of a light sbottom is usually motivated as follows: Firstly, the neutral kaon system gives a strong constraint on the masses of the first and second generation squarks. The third generation sfermions are much less constrained so far. Secondly, in the supergravity scenario of supersymmetry breaking, mass splitting of the third-generation and the other sfermions results from the renormalization group evolution of the masses between the unification scale and the weak scale, even if the sfermions have equal masses at the unification scale. This splitting is due to the effect of the large Yukawa coupling of the top. The bottom and tau sectors are also affected. Thirdly, there are arguments that first and second generation sfermions can be as heavy as 10 TeV without conflicting the naturalness problem, while the third generation sfermions have to be rather light.
* As pointed out in Sec. (1), the two decay processes we considered resemble the favorable cases in which a $`b`$-jet is produced in the decay products. While we can apply our results directly to the cases of $`\lambda _{332}^{\prime \prime }`$ and $`\lambda _{133}^{}`$, we noticed that similar decays induced by other couplings like $`\lambda _{312}^{\prime \prime }`$ and $`\lambda _{232}^{}`$ give poor signals since there is no $`b`$-quark in their corresponding top decays.
* We noted that apart from the relevant R-violating couplings and the sbottom mass, our results are also dependent on the mass and coupling of the lightest neutralino. In our calculation we only present some illustrative results by fixing a set of SUSY parameters rather than scanning the entire allowed SUSY parameter space. In some unfavorable cases, such as when the mass of the lightest neutralino (LSP) is close to the sbottom mass so that the b-quark from the sbottom decay ($`\stackrel{~}{b}b\stackrel{~}{\chi }_1^0`$) is too soft to pass the selection cuts, these exotic decays would be unobservable even at the LHC.
* As pointed out in Sec. 2, the $`B`$-violating decay gives the unique signal of same sign b-quarks while the main SM backgrounds give the unlike sign b-quarks. To be conservative, we assumed in our analysis that the $`b`$ tagging is not of sufficient sensitivity to distinguish between a $`b`$-jet and a $`\overline{b}`$-jet. If $`b`$ charge identification can be achieved in future detectors, more stringent discovery limits than those we have presented will be possible. Additional improvements will be possible if hadronic jets from $`\tau `$ decays can be clearly identified as such, thus reducing the background from $`\tau `$ hadronization.
## Acknowledgments
We thank E. Boos and L. Dudko for the discussion of $`Wb\overline{b}j`$ background. BLY acknowledges the hospitality extended to him by Professor Zhongyuan Zhu and colleagues at the Institute of Theoretical Physics, Academia Sinica, where part of the work was performed. This work is supported in part by a grant of Chinese Academy of Science for Outstanding Young Scholars and also by a DOE grant No. DE-FG02-92ER40730.
Table 1: Signal $`\mathrm{}+3j/2b+\overline{)E_T}`$ and background cross sections in units of fb. The basic cuts are $`p_T^{\mathrm{all}}20\mathrm{GeV}`$, $`|\eta _{\mathrm{all}}|2.5`$ and $`\mathrm{\Delta }R0.5`$, and the transverse mass cut is $`m_T120`$ GeV. The signal results were calculated by assuming $`M=100`$ GeV, $`\mu =200`$ GeV and $`\mathrm{tan}\beta =1`$. The double b-jet tagging with $`42\%`$ efficiency is assumed. The charge conjugate channels have been included.
| | Sbottom mass(GeV) | 150 | 155 | 160 | 165 | 170 | 180 | 190 |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| Tevatron | Signal/$`(\lambda _{331}^{\prime \prime })^2`$ | 11 | 5.8 | 2.04 | 0.27 | 0.01 | 0.005 | 0.003 |
| (1.8 Tev) | Background | 2.07 | | | | | | |
| | Sbottom mass(GeV) | 150 | 155 | 160 | 165 | 170 | 180 | 190 |
| Tevatron | Signal/$`(\lambda _{331}^{\prime \prime })^2`$ | 16 | 8.4 | 3.0 | 0.4 | 0.02 | 0.007 | 0.004 |
| (2 Tev) | Background | 3.05 | | | | | | |
| | Sbottom mass(GeV) | 150 | 155 | 160 | 165 | 170 | 180 | 190 |
| LHC | Signal/$`(\lambda _{331}^{\prime \prime })^2`$ | 1624 | 885 | 371 | 58 | 1.7 | 0.4 | 0.3 |
| (14 Tev) | Background | 350 | | | | | | |
Table 2: Signal $`\mathrm{}+4j/2b+\overline{)E_T}`$ and the SM $`t\overline{t}`$ background cross sections for sbottom mass of 150 GeV. The basic cuts are $`p_T^{\mathrm{all}}20\mathrm{GeV}`$, $`|\eta _{\mathrm{all}}|2.5`$ and $`\mathrm{\Delta }R0.5`$, and the transverse mass cut is $`m_T(\mathrm{},p_T^{\mathrm{miss}})50100\mathrm{GeV}`$. The signal results were calculated by assuming $`M=100`$ GeV, $`\mu =200`$ GeV and $`\mathrm{tan}\beta =1`$. Tagging at least one $`b`$-jet is assumed for 53% efficiency for the Tevatron (1.8 TeV), $`85\%`$ efficiency for the upgraded Tevatron (2 TeV) and LHC. The charge conjugate channels have been included.
| | | basic cuts | basic cuts |
| --- | --- | --- | --- |
| | | | $`+`$ |
| | | | $`m_T(\mathrm{},p_T^{\mathrm{miss}})`$ cut |
| | Signal/$`(\lambda _{233}^{})^2`$ (fb) | 70 | 43 |
| Tevatron (1.8 TeV) | Background (fb) | 300 | 86 |
| | Signal/$`(\lambda _{233}^{})^2`$ (fb) | 154 | 96 |
| Tevatron (2 TeV) | Background (fb) | 662 | 193 |
| | Signal/$`(\lambda _{233}^{})^2`$ (pb) | 12.7 | 8.2 |
| LHC (14 TeV) | Background (pb) | 54 | 16 |
Table 3: Same as Table 2, but for the signal cross section versus sbottom mass under the basic plus transverse mass cut.
| Tevatron (1.8 Tev): | | | | | | | | |
| --- | --- | --- | --- | --- | --- | --- | --- | --- |
| Sbottom mass(GeV) | 150 | 155 | 160 | 165 | 170 | 180 | 190 | 200 |
| Signal/$`(\lambda _{331}^{})^2`$ (fb) | 42.8 | 23.7 | 8.0 | 0.86 | 0.04 | 0.02 | 0.01 | 0.007 |
| Tevatron (2 TeV): | | | | | | | | |
| Sbottom mass(GeV) | 150 | 155 | 160 | 165 | 170 | 180 | 190 | 200 |
| Signal/$`(\lambda _{233}^{})^2`$ (fb) | 96 | 53 | 19 | 2.2 | 0.09 | 0.04 | 0.024 | 0.016 |
| LHC (14 TeV): | | | | | | | | |
| Sbottom mass(GeV) | 150 | 155 | 160 | 165 | 170 | 180 | 190 | 200 |
| Signal/$`(\lambda _{233}^{\prime \prime })^2`$ (pb) | 8.2 | 4.8 | 1.86 | 0.26 | 0.008 | 0.003 | 0.002 | 0.001 |
|
warning/0007/cond-mat0007284.html
|
ar5iv
|
text
|
# Finite-wavevector phonon coupling to degenerate electronic states in La2-xSrxCuO₄.
## I Introduction
The existence of local intersite pairs in cuprate superconductors is inferred from the very short coherence lengths in these materials, which is in turn determined from measurements of $`H_{c2}`$. Given that the pair dimensions $`d`$ cannot exceed the coherence length, i.e. $`d\xi `$, we may infer that any possible lattice distortions associated with pairing have a finite range, defined by the extent of the pair $`d`$. However, the particular form of the electron-phonon interaction leading to these distortions and particular type of phonon mode responsible for pairing should be determined from experiments. According to inelastic neutron scattering data a large (30%) anomaly in $`\tau _1`$ phonon mode corresponding to the ($`\zeta ,0,0`$) direction has been observed in a number of doped cuprates (but not in the parent compound).
The observed anomaly takes place at an incomensurate wavevector and is suggested to be associated with pairing and/or stripe formation. The electron phonon interaction can then be written in the form:
$$g(k_0,𝐤)=g_0/((kk_0)^2+\gamma ^2)$$
(1)
where $`g_0`$ is a constant describing the strength of coupling, $`k_0`$ defines the wavevector asociated with the interaction and its range in $`k`$-space, while $`\gamma `$ defines its width. In this paper we analyse the $`ep`$ coupling of phonon modes with $`k0`$using group theory applicable to La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> and discuss the relevant phonon modes by examining the available experimental data from neutron diffraction, ESR and EXAFS. We find that the experimentally observed features can be described by the coupling of a $`\tau _1`$ \- symmetry in-plane O-mode along the $`\mathrm{\Sigma }`$ direction to holes occupying $`E_g`$ or $`E_u`$ \- symmetry planar O states.
## II E-p coupling
The Brillouin zone (BZ) corresponding to the tetragonal point group $`D_{4h}`$ applicable for La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> is shown in Figure 1. To consider local pairs and/or stripes forming along the Cu-O bond direction or along 45 to it, we need to consider the $`\mathrm{\Sigma }`$ and the $`\mathrm{\Delta }`$ points, corresponding to the ($`\zeta ,0,0)`$ and ($`\zeta ,\zeta ,0)`$ directions respectively. The special symmetry points ($`\mathrm{\Gamma },X`$ and $`M`$ etc.) give rise to commensurate stripes.
Before proceeding with analysis of the e-p coupling for the case of general $`k`$, it should be pointed out since the symmetrised cross product of the representations at the $`\mathrm{\Gamma }`$ point,
$$[E_u\times E_u]=[E_g\times E_g]=A_{1g}+B_{1g}+B_{2g}$$
(2)
the electrons can couple only with $`A_{1g}`$ modes (there are no $`B_{1g}`$ and $`B_{2g}`$ at the $`\mathrm{\Gamma }`$ point of La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub>). (There are two such modes in $`D_{4h}`$ and are associated either with apex oxygen O(2) ions or La ions.) These do not lead to any symmetry breaking. It follows that stripe formation or further symmetry breaking cannot be associated with the $`\mathrm{\Gamma }`$ point ($`k=0)`$, but can only be associated with finite $`𝐤`$. For the case of tetragonal symmetry, this corresponds to the $`\mathrm{\Sigma }`$ and the $`\mathrm{\Delta }`$ points in the BZ.
### A Coupling to non-degenerate states
Since the $`\mathrm{\Sigma }`$ and $`\mathrm{\Delta }`$ points have a four pronged star, we propose the following form of coupling to electrons in single non-degenerate electronic states:
$$H_{int}=\underset{𝐥,s}{}n_{𝐥,s}\underset{k_0=1}{\overset{4}{}}\underset{𝐤}{}g(k_0,𝐤)\mathrm{exp}(i\mathrm{𝐤𝐥})(b_𝐤^{}+b_𝐤)$$
(3)
where $`𝐥`$ is the site label and
$$g(k_0,𝐤)=g/((kk_0)^2+\gamma ^2)$$
(4)
where $`k_0`$ are the 4 wavevectors corresponding to the prongs of the star associated with stripe (pair) formation. We should stress that in this case the nondegenerate electronic states are associated with $`p_z`$-orbitals of planar oxygens which transform as $`A_{2u}`$ or $`B_{2u}`$. The Hamiltonian above on its own does not lead to symmetry breaking, but phase separation can in principle still take place as a result of competition of short-range attraction and long-range Coulomb repulsion forces. However, the symmetry of both phases would be the same.
### B Coupling to degenerate states (Jahn-Teller-like coupling)
A more interesting case arises when two-fold degenerate levels (for example the two $`E_u`$ states corresponding to the planar O $`p_x`$ and $`p_y`$ orbitals or the $`E_u`$ and $`E_g`$ states of the apical O) interacts with $`k0`$ phonons. Again, we are interested in the phonons which lead to symmetry breaking and allow the formation of stripes. The invariant form of Hamiltonian has a simple form for phonons which transforms as the $`\tau _1`$ representations of the group of wave-vector $`G_k`$. Taking into account that $`E_g`$ and $`E_u`$ representations are real and Pauli matrices $`\sigma _i`$ corresponding to the doublet of $`E_g`$ or $`E_u`$ transform as $`A_{1g}`$ ($`k_x^2+k_y^2`$) for $`\sigma _0=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)`$, $`B_{1g}`$ ($`k_x^2k_y^2`$) for $`\sigma _3=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)`$, $`B_{2g}`$ ($`k_xk_y`$) for $`\sigma _1=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)`$, and $`A_{2g}`$ ($`s_z`$) for $`\sigma _2=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right)`$representations respectively, we can construct an invariant form of the Hamiltonian:
$`H_{int}`$ $`=`$ $`{\displaystyle \underset{𝐥,s}{}}\sigma _{0,𝐥}{\displaystyle \underset{k_0=1}{\overset{4}{}}}{\displaystyle \underset{𝐤}{}}g_0(k_0,𝐤)\mathrm{exp}(i\mathrm{𝐤𝐥})(b_𝐤^{}+b_𝐤)+`$ (8)
$`{\displaystyle \underset{𝐥,s}{}}\sigma _{3,𝐥}{\displaystyle \underset{k_0=1}{\overset{4}{}}}{\displaystyle \underset{𝐤}{}}g_1(k_0,𝐤)(k_x^2k_y^2)\mathrm{exp}(i\mathrm{𝐤𝐥})(b_𝐤^{}+b_𝐤)+`$
$`{\displaystyle \underset{𝐥,s}{}}\sigma _{1,𝐥}{\displaystyle \underset{k_0=1}{\overset{4}{}}}{\displaystyle \underset{𝐤}{}}g_2(k_0,𝐤)k_xk_y\mathrm{exp}(i\mathrm{𝐤𝐥})(b_𝐤^{}+b_𝐤)`$
$`{\displaystyle \underset{𝐥,s}{}}\sigma _{2,𝐥}S_{z,𝐥}{\displaystyle \underset{k_0=1}{\overset{4}{}}}{\displaystyle \underset{𝐤}{}}g_3(k_0,𝐤)\mathrm{exp}(i\mathrm{𝐤𝐥})(b_𝐤^{}+b_𝐤)`$
where
$$g_i(k_0,𝐤)=g_i/((kk_0)^2+\gamma ^2)$$
(9)
The first term in (5) describes the symmetric coupling and is identical to the non-degenerate case. The second and third terms describe the e-p interaction corresponding to the $`\mathrm{\Sigma }`$ and $`\mathrm{\Delta }`$ directions respectively, while the last term describes the coupling to spins.
The proposed interaction (5) on its own results in splitting of the degenerate states (a Jahn-Teller-like effect), breaking tetragonal symmetry and resulting in a local orthorhombic distortion. It can therefore lead to the formation of pairs and stripes with no further interaction. Of course the stability and size of such a distortion will be determined by the balance of short-range attraction and long-range Coulomb repulsion.
## III Discussion
To determine the relevant wavevector $`k_0`$ applicable in (5) and (6) we need to look at the experimental data. It is well known that for $`k_0=0,`$ there are no significant phonon anomalies (in Raman or infrared spectroscopies for example) which can be associated either with $`T_c`$ or the pseudogap $`T_p`$. However, local probe experiments like neutron scattering, EXAFS and ESR give a different picture, and point to the existence of sizeable lattice anomalies for $`k0`$. The lattice distortions associated with these experimental observations are shown in Figure 2. Particularly the inelastic neutron data (Fig. 2a) should be singled out here because of the unambiguity of the raw data. In these experiments a large renormalisation of phonon frequency is ubiquitously observed along the $`\mathrm{\Sigma }`$ direction in many cuprates, including 214 and 123 structure superconductors, but not in their parent insulators. Moreover, its temperature dependence clearly suggests a connection with pair formation.
The phonon $`\tau _1`$ modes corresponding to the $`\mathrm{\Sigma }`$ direction which can couple to the electrons are shown in Figure 3. One of these, O(1)1 can be clearly identified as the mode observed in neutron scattering experiments and may thus be indentified as the mode responsible for pairing or stripe formation. We also note that from the experiments, the width in $`k`$-space of the anomaly is of the same order as the magnitude.
The interaction given in Eqs.(5) and (6) leads directly to phase segregation of distorted and undistorted domains, where the size of domains is determined by the interplay of short-range attraction and screened Coulomb repulsion as already mentioned. The incommensurability of the interaction leads to modulation of charge density within the stripe. Importantly since the number of charged carriers in the stripe is less than the number of sites occupied by stripe, we may expect to observe itinerant behaviour in such stripes, i.e. the charge carriers are not localised within the stripe itself. At low doping when screening is reduced, the formation of the stripes of the length 2 (pairs) is most probable, while at lower temperatures and higher doping clearly longer stripes are more probable. These pairs, contrary to the case of the Holstein interaction is more extended in space and its size is governed by $`k_0`$.
## IV Conclusion
The central idea that the pair is the ground state entity which defines the extent of the e-p interaction leads to an interaction Hamiltonian of degenerate modes with $`k0`$ phonons which is suggested to describe the pairing and stripe formation in the cuprates. The splitting of the degeneracy of planar O orbitals gives rise to a deformation either along the bond axes or along the diagonals, depending on which mode is chosen, but inelastic neutron scattering experiments suggest that the important interaction is along the Cu-O bonds, giving rise to an $`s+d`$ \- like form of the interaction as well as coupling to spins via an $`A_{2g}`$-symmetry term. Remarkably, the interaction which splits the degeneracy of the O states via a JT-like effect is in spirit, if not in detail similar as was originally suggested by Bednorz and Muller in their original paper on La<sub>2-x</sub>Ba<sub>x</sub>CuO<sub>4</sub>.
## V Figure captions
Figure 1. The Brillouin zone of the tetragonal structure corresponding to La<sub>2-x</sub>Sr<sub>x</sub>CuO$`_4.`$ According to experiment, the $`\mathrm{\Sigma }`$ directions is of relevance for pairing and stripe formation.
Figure 2. The vibrational modes causing the anomalies observed in a) neutron scattering (ref. 1), b) ESR (ref. 4) and c) EXAFS (ref. 3)
Figure 3. a) The wavevectors corresponding to the $`\tau _1`$representation at the $`\mathrm{\Sigma }`$point of the point group $`D_{4h}`$. Note that the relative magnitude of the displacements shown depends on the value of $`k_0`$. b) The $`\tau _7`$ mode corresponding to the distortion suggested in ref. 4 (Fig. 2 b) can also be viewed as a linear combination of two $`\tau _1`$ modes at the $`\mathrm{\Sigma }`$point with different $`k`$.
|
warning/0007/hep-ph0007362.html
|
ar5iv
|
text
|
# Acknowledgements
## Acknowledgements
We are grateful to J. Vermaseren for communicating the results of ref. to us prior to publication. We also thank V. Braun for pointing ref. out to us. This work has been supported by the European Community TMR research network ‘QCD and the Deep Structure of Elementary Particles’ under contract No. FMRX–CT98–0194.
Figure 1: Approximations of the $`N_f^1`$ part $`P_{qg,1}^{(2)}`$ of the three-loop splitting function $`P_{qg}^{(2)}(x)`$, as obtained from the six lowest even-integer moments together with the leading small-$`x`$ term of ref. . The full curves represent those functions selected for Eq. (11).
Figure 2: Top: Our new approximations of $`P_{\mathrm{NS}}^{(2)+}(x)`$ for $`N_f=4`$, as obtained from Eq. (7) together with Eq. (4.13) of ref. . The dotted curves represent our previous parametrizations . Bottom: The same for $`P_{\mathrm{NS}}^{(2)}(x)`$ using Eq. (6).
Figure 3: Our new approximations of the singlet splitting functions $`P_{ij}^{(2)}(x)`$ for $`N_f=4`$. $`P_{qq}^{(2)}`$ is obtained by adding $`P_{\mathrm{NS}}^{(2)+}(x)`$ of Fig. 2 (separately shown by the dash-dotted curve) and $`P_{\mathrm{PS}}^{(2)}(x)`$ of Eqs. (9) and (10). Our previous parametrizations are also shown.
Figure 4: The size and remaining uncertainties of the NNLO corrections for the scale derivatives, $`\dot{\mathrm{\Sigma }}d\mathrm{\Sigma }/d\mathrm{ln}\mu _f^2`$ and $`\dot{g}dg/d\mathrm{ln}\mu _f^2`$, of the singlet quark and gluon densities at $`\mu _f^2=\mu _r^230\text{ GeV}^2`$ ($`\alpha _s=0.2`$). The input densities are specified in Eq. (S0.Ex40).
|
warning/0007/cond-mat0007463.html
|
ar5iv
|
text
|
# Nesting Induced Precursor Effects: a Renormalization Group Approach
\[
## Abstract
We develop a controlled weak coupling renormalization group (RG) approach to itinerant electrons. Within this formalism we rederive the phase diagram for two-dimensional (2D) non-nested systems. Then we study how nesting modifies this phase diagram. We show that competition between p-p and p-h channels, leads to the manifestation of unstable precursor fixed points in the RG flow. This effect should be experimentally measurable, and may be relevant for an explanation of pseudogaps in the high temperature superconductors (HTC), as a crossover phenomenon.
\]
An important clue to understand the physics of the HTC may come from the unconventional normal state properties preceding superconducting ordering: a pseudogap is opened, which manifests through an apparent continuous reduction in the density of states at a temperature $`T^{}`$, higher than the superconducting critical temperature, $`T_c`$. The precise definition of this incipient ordering mechanism is still subject of discussion<sup>,</sup>. Two main pictures have been proposed depending on whether the correlations being built up in the pseudogap are of superconducting or (spin, or charge) density wave nature. However no experimental conclusive proof has been given so far favoring any of these pictures.
In the former, it is assumed that pre-formed pairs appear at the pseudogap energy scale $`T^{}`$, but coherence would only be established at $`T_c.`$ Some support in favor of this scenario is provided for instance by ARPES measurements. They indicate that the pseudogap shares some of the superconducting properties like the gap symmetry. Furthermore, they reveal the presence of a non-vanishing gap in the vicinity of the $`(\pi ,0)`$ point already above $`T_c`$. Nevertheless, it has been argued that NMR and heat capacity data show that superconductivity and pseudogap are competing, like antagonistic phases. Also, $`T^{}`$ merges with $`T_c`$ at a critical doping in the overdoped superconducting region, when the condensation energy, the critical currents and the superfluid density reach sharp maxima. This has been argued to be irreconcilable with a preformed pairing scenario.
The amplitude of incommensurate peaks observed in inelastic neutron scattering varies upon doping in a way highly related to the pseudogap and falls to zero in the overdoped region. ARPES has clearly shown the existence of a Fermi surface which is lying not too far from the diagonals of the first Brillouin zone, therefore providing an approximate nesting.
In the ideal case of perfect nesting, it is well known that density wave and superconducting instabilities are strongly interfering<sup>-</sup> . In this work we present a simple one-loop RG approach to electrons on 2D Fermi surfaces with flat nested regions, away from half-filling. We show that the strong competition arising from particle-particle and particle-hole corrections when nesting of the Fermi surface (FS) is included, naturally induces instabilities that can be preceded by other incipient instabilities. In the RG sense this appears as a crossover phenomenon: the RG flow passes near an unstable fixed point, capturing the physics of this fixed point at intermediate energies, before the system orders in the final superconducting state. This scenario naturally provides two energy scales and therefore bears striking similarities with the pseudogap phase of the HTC’s. The condition of perfect nesting may be relaxed provided interactions are large enough.
We parametrize a 2D Fermi surface with a set of $`4N_c`$ patches, as illustrated in figure 1. The kinetic energy dispersion is linearized around the Fermi surface, each patch being assigned a Fermi velocity, $`v_I.`$ In this work all $`v_I`$ are assumed equal as the results do not depend significantly on smooth Fermi velocity modulations. The most general spin rotational invariant interaction term which remains in an effective low energy theory only involves incoming electrons on opposite sides of the Fermi surface. This remark enables us to separate the set of patches in two complementary subsets called right ($`R`$) and left ($`L`$) patches, for convenience. We may therefore write:
$`_C`$ $`=`$ $`{\displaystyle \frac{1}{2N_cL}}{\displaystyle \underset{I,K,\delta }{}}{\displaystyle \underset{q0}{}}(C_\delta ^c(I,J)J_{q,\delta }^R\left(I\right)J_{q,\delta }^L\left(K\right)`$ (2)
$`+C_\delta ^s(I,J)𝐉_{q,\delta }^R\left(I\right)𝐉_{q,\delta }^L\left(K\right))`$
where charge $`J_\delta ^R(I)=_{\tau ,\tau ^{}}:R_{I+\delta ,\tau }^+\delta _{\tau ,\tau ^{}}R_{I,\tau ^{}}:`$ , and spin currents $`𝐉_\delta ^R(I)=_{\tau ,\tau ^{}}:R_{I+\delta ,\tau }^+\sigma _{\tau ,\tau ^{}}R_{I,\tau ^{}}:,`$ have been used, and momentum variables within each patch have been omitted to simplify the notation. Here chain indices could have been written in bold as they could represent vectors. In that case we could, for instance, consider a bilayer of 2D electron systems: the present formalism is in fact quite general allowing the treatment of many different systems. Eq. $`\left(\text{2}\right)`$ allows an extremely simple derivation of one-loop equations: only the two diagrams in fig. 1, need to be considered. Requiring the invariance of the two-particle vertex under cutoff reduction we arrive at ($`/\mathrm{ln}D`$ and $`C_iC_i/2\pi N_c`$) :
$`C_\delta ^c(I,K)`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}\delta ^c(I,K,\delta ,\alpha )(C_\alpha ^c(I,K)C_{\delta \alpha }^c(I+\alpha ,K+\alpha )`$ (6)
$`+3C_\alpha ^s(I,K)C_{\delta \alpha }^s(I+\alpha ,K+\alpha ))/(v_{I+\alpha }+v_{K+\alpha })`$
$`{\displaystyle \underset{\alpha }{}}\delta ^{ZS}(I,K,\delta ,\alpha )(C_\alpha ^c(I,K\alpha +\delta )C_{\delta \alpha }^c(I+\alpha ,K)`$
$`+3C_\alpha ^s(I,K\alpha +\delta )C_{\delta \alpha }^s(I+\alpha ,K))/(v_{I+\alpha }+v_{K\alpha +\delta })`$
for charge couplings, whereas for spin couplings we get:
$`C_\delta ^s(I,K)`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}\delta ^c(I,K,\delta ,\alpha )(C_\alpha ^s(I,K)C_{\delta \alpha }^c(I+\alpha ,K\alpha )`$ (12)
$`+C_\alpha ^c(I,K)C_{\delta \alpha }^s(I+\alpha ,K\alpha )`$
$`2C_\alpha ^s(I,K)C_{\delta \alpha }^s(I+\alpha ,K\alpha ))/(v_{I+\alpha }+v_{K+\alpha })`$
$`{\displaystyle \underset{\alpha }{}}\delta ^{ZS}(I,K,\delta ,\alpha )(C_\alpha ^c(I,K\alpha +\delta )C_{\delta \alpha }^s(I+\alpha ,K)`$
$`+C_\alpha ^s(I,K\alpha +\delta )C_{\delta \alpha }^c(I+\alpha ,K)`$
$`+2C_\alpha ^s(I,K\alpha +\delta )C_{\delta \alpha }^s(I+\alpha ,K))/(v_{I+\alpha }+v_{K\alpha +\delta })`$
Here $`\delta ^c`$ and $`\delta ^{ZS}`$ implement phase space restrictions: they are equal to 1 iff all electrons involved in the scattering are kept on the FS along the process. This happens if the electrons are on nested regions of the FS, or, in the case of the Cooper channel, if $`I=K`$. Otherwise $`\delta ^c`$ and $`\delta ^{ZS}`$ are set equal to zero.
In order to characterize the possible instabilities in the system, we introduce several response functions, of the form: $`_\delta (I,K)=`$ $`i𝑑te^{i\omega t}`$ $`T\left\{𝒪_\delta (I;t)𝒪_\delta ^+(K;0)\right\}.`$ For singlet superconductivity we denote $`_\delta (I,K)S_\delta ^S(I,K)`$ and the order parameter is: $`𝒪_{SSC,\delta }(I;t)=\frac{1}{\sqrt{2}}_\tau \tau R_{I+\delta ,\tau }\left(t\right)L_{I,\tau }\left(t\right).`$ For triplet superconductivity, $`_\delta (I,K)S_\delta ^T(I,K),`$ and $`𝒪_{TSC,\delta }(I;t)=\frac{1}{\sqrt{2}}_\tau R_{I+\delta ,\tau }\left(t\right)L_{I,\tau }\left(t\right).`$ As interactions do not connect pairs of particles with different $`\delta `$ (because of momentum conservation), the problem of finding the most divergent instabilities is reduced to that of diagonalizing matrices $`S_\delta (I,K),`$ with $`\delta `$ fixed. The eigenvectors determine the order parameter characterizing the instability.
For charge density wave we define $`_\delta (I,K)D_\delta ^c(I,K),`$ with the order parameter $`𝒪_{CDW,\delta }\left(I\right)=\frac{1}{\sqrt{2}}_\tau R_{I+\delta ,\tau }^+L_{I,\tau }`$ $`\frac{1}{\sqrt{2}}_\tau R_{M+Q,\tau }^+L_{MQ,\tau }`$(i.e., $`M=I+\frac{\delta }{2}`$ and $`Q=\frac{\delta }{2}`$). The last definition stresses that interactions connect only order parameters with equal particle-hole center of mass, $`M.`$ Thus, the problem of diagonalizing $`D_\delta ^c(I,K)`$ is once again simplified as there is a fixed parameter. Finally for spin density waves we define similarly $`𝒪_{SDW,\delta }\left(I\right)=`$ $`\frac{1}{\sqrt{2}}_\tau \tau R_{M+Q,\tau }^+L_{MQ,\tau }.`$
With these definitions it is straightforward to obtain RG equations for the response functions<sup>,</sup>, of the form $`\mathrm{ln}\overline{}_\delta (I,K)/\mathrm{ln}D=\upsilon `$ with $`\overline{}_\delta (I,K)`$= $`\pi \left(v_{I+\delta }+v_I\right)_\delta (I,K)/\mathrm{ln}\omega `$. More precisely:
$`\upsilon _{SSC}`$ $`=`$ $`{\displaystyle \frac{C_{IK}^c(K,K+\delta )3C_{IK}^s(K,K+\delta )}{\left(v_{K+\delta }+v_K\right)}}`$ (13)
$`\upsilon _{TSC}`$ $`=`$ $`{\displaystyle \frac{C_{IK}^c(K,K+\delta )+C_{IK}^s(K,K+\delta )}{\left(v_{K+\delta }+v_K\right)}}`$ (14)
$`\upsilon _{CDW}`$ $`=`$ $`{\displaystyle \frac{C_{IK}^c(K,I+\delta )+3C_{IK}^s(K,I+\delta )}{\left(v_{2IK+\delta }+v_K\right)}}`$ (15)
$`\upsilon _{SDW}`$ $`=`$ $`{\displaystyle \frac{C_{IK}^c(K,I+\delta )C_{IK}^s(K,I+\delta )}{\left(v_{2IK+\delta }+v_K\right)}}`$ (16)
In this work we are concerned with the ordering of the electronic system. As a general trend, the effective couplings in the one-loop approximation diverge for a finite value of the reduced cut-off. This is naturally interpreted as the signature of a low temperature instability towards an ordered state. However this divergency of the effective couplings also signals a breakdown of the one-loop approximation. This difficulty may be overcome by noting that in general ratios of any two coupling constants reach finite asymptotic values as the instability is approached. In this work we have thus emphasized the study of these asymptotic flows in the space of directions for the effective coupling vector. Representing the $`C_\delta (I,J)`$ by a generic $`C_i,`$ this can be done by finding the flow of non-diverging normalized couplings $`c_iC_i/𝒩,`$ with $`𝒩=\sqrt{_j\left(C_j\right)^2}`$ (the sum is over all the couplings). From the RG equations $`\left(\text{6}\right)`$ and $`\left(\text{12}\right)`$ having the form $`C_i/\mathrm{ln}D=`$ $`A_{ijk}C_jC_k`$ we obtain the evolution for the normalized couplings: $`dc_i/du=_{jkl}A_{jkl}c_k`$$`c_l\left(\delta _{ij}c_ic_j\right).`$ We have introduced a new scale parameter $`u`$ related to $`D`$ by $`du=𝒩`$ $`d\mathrm{ln}D`$. In this parametrization the magnitude $`𝒩`$ of the coupling vector diverges only as $`u\mathrm{}.`$ From the flow as a function of $`u`$, we can recover the more physical variable $`D`$ since $`d\mathrm{ln}𝒩/du`$= $`A_{ikl}c_ic_kc_l`$ which yields $`𝒩\left(u\right)`$ and therefore $`D\left(u\right)`$. A similar procedure can be applied to the computation of response functions, that we generically represent by $`\overline{}_i`$. Their RG equations have the form $`d\mathrm{ln}\overline{}_i/d\mathrm{ln}D=B_{ik}C_k`$, which becomes $`d\mathrm{ln}\overline{}_i/du=B_{ik}c_k`$. As $`u`$ goes to infinity the r.h.s. reaches a finite value. This is easily seen to imply a divergency of the form $`_i[\mathrm{ln}(D/D^{})]^{\alpha _i}`$, where $`D^{}`$ is the scale at which the instability occurs. The exponents $`\alpha _i`$ are completely determined by the fixed point values $`c_i^{}`$ of the normalized couplings.
This procedure provides a controlled framework to study the complete RG flow (not restricted to the one-loop approximation) provided the bare couplings are small enough. Certainly we expect the one-loop approach to break down if $`𝒩`$ is of the order unity. For a given choice of initial couplings with norm $`𝒩_0`$, this yields a maximal value of $`u`$ ($`u_{\mathrm{max}}`$) beyond which our approximation is no longer reliable. The main point is that $`u_{\mathrm{max}}`$ can be made arbitrarily large provided $`𝒩_0`$is small enough. In general we expect $`u_{\mathrm{max}}\mathrm{ln}(1/𝒩_0)`$. In some applications we are also interested in a finite an not infinitesimal $`𝒩_0`$ regime. Higher order corrections destroy the fixed directions found in the one-loop approximation. However, we conjecture those become fixed points in the usual sense and that the topology of the one-loop flow pattern is preserved.
As a first example of application of the present formalism we consider a non-nested Fermi surface. The spinless case is thoroughly discussed in the review paper by Shankar. Phase space restrictions set in $`\left(\text{6}\text{-}\text{12}\right)`$ through the $`\delta `$ functions, imply that only couplings of the form $`C_\delta (I,I)`$ are renormalized, due to particle-particle corrections (BCS channel). Divergence of the flow is then related to superconductivity. Indeed, inspired from (13-14), the RG flow equations can be quite simply written in terms of singlet $`C_\delta ^S(I)C_\delta ^c(I,I)3C_\delta ^s(I,I)`$ and triplet couplings$`C_\delta ^T(I)C_\delta ^c(I,I)+C_\delta ^s(I,I)`$:
$$C_\delta ^T(I)=\alpha \left(C_\delta ^T(I)\right)^2C_\delta ^S(I)=\alpha \left(C_\delta ^S(I)\right)^2$$
(17)
where $`\alpha `$ is a Fermi velocity dependent constant. The phase diagram is shown in Figure 2, for a system with initial couplings $`C_\delta ^c(I,I)=C^c`$ and $`C_\delta ^s(I,I)=C^s`$. The only region where the flow does not diverge corresponds to a Fermi Liquid phase. In the other regions, the RG couplings diverge at the energy scale that we associate with the superconducting critical temperature, $`T_c`$.
This picture changes when we introduce nesting. As already discussed flat regions dramatically enhance the number of low energy couplings to be considered. The occurrence of log singularities in both the particle-particle and the particle-hole channels induce an intricate interference mechanism, leading to the so-called parquet regime. In Figure 3, we show the new phase diagram, together with typical flow patterns (using the $`u`$ variable, and normalized couplings) for each of the different phases. These results have been obtained for a partially nested Fermi surface with $`N=6`$ patches along each flat segment and $`M=0`$ along each curved arc. Note that increasing the number of patches doesn’t change significantly the phase-diagram.
The modifications due to nesting are most obvious in the initial coupling domain corresponding to the Fermi liquid regime for the circular Fermi surface. For positive bare spin coupling, we obtain a CDW phase, as a mean field analysis predicts. The most striking result appears for repulsive couplings in the charge sector and moderate negative spin couplings. With this parametrization, the simple repulsive Hubbard model falls in this region. The flow pattern from Figure 3 clearly shows a two-step evolution as $`u`$ is increased. Before entering the final asymptotic regime, the system spends a large $`u`$-time in the proximity of an unstable fixed point (for the normalized couplings). In this intermediate phase, the most diverging response is the SDW susceptibility. The corresponding order parameter has the approximate form $`R_N^+L_1+R_{N1}^+L_2+\mathrm{}+h.c`$ so the most favorable nesting vector is perpendicular to the flat regions. Couplings involving incoming and outgoing electrons close to the endpoints of flat regions are the most enhanced. This intermediate fixed point corresponds very likely to the one found by Zheleznyak et al who worked with the original un-normalized couplings. Our approach yields the remarkable result that the system is finally attracted towards a stable low energy d-wave superconductor ($`SSC_d`$) fixed point. Again, the largest couplings involve the endpoints of the flat regions, and from the superconducting response function, the gap exhibits a strong anisotropy. It vanishes at the center of flat regions and reaches its maximal values at their extremities. The dramatic role played by these extremal regions suggests some similarity with the Van-Hove singularity scenario studied by several groups . However, we haven’t introduced any singularity in the Fermi velocity. Our model also ignores Umklapp processes, which are expected to play a crucial role in stabilizing a commensurate SDW (Néel state), and in describing the proximity to a Mott insulator .
It is important to specify the energy scales associated to both fixed points. Assuming the bare bandwidth is $`D_0`$, the correspondence between the running scale $`D`$ and the fictitious time $`u`$ takes the form: $`ln(D_0/D)=h(u)/𝒩_0`$, where $`𝒩_0`$ is the length of the initial coupling vector, and $`h(u)`$ is a function depending on the direction of this vector. In general, $`h`$ appears to be an increasing function of $`u`$ which reaches a finite limit $`h(\mathrm{})`$ as $`u`$ becomes infinite. This defines the instability scale $`D_c`$ below which the system is superconducting. Similarly, we define a scale $`D_1`$ for which the distance of the normalized coupling vector to the intermediate fixed point is minimal. Let us note by $`u_1`$ the corresponding value of $`u`$. We obtain: $`D_1/D_0=(D_c/D_0)^{h(u_1)/h(\mathrm{})}`$, which holds for a chosen direction of the initial coupling vector. $`D_c`$ and $`D_1`$ are considered here to be functions of $`𝒩_0`$. In practice, we have found that $`h(u_1)/h(\mathrm{})`$ is very close to unity. As an order of magnitude, if $`D_c100K`$ and $`D_05000K`$, $`D_1D_c1K`$. However, SDW correlations begin to build up at a much higher energy scale, as can be seen in Figure 4. We may for instance evaluate the scale $`D^{}`$ at which the SDW exponent reaches 99 percent of its maximal value.
| $`N`$ | $`3`$ | $`6`$ | $`8`$ | $`10`$ | $`12`$ | $`14`$ |
| --- | --- | --- | --- | --- | --- | --- |
| $`T_c\left(K\right)`$ | $`0`$ | $`0.0`$ | $`0.2`$ | $`15`$ | $`82`$ | $`150`$ |
| $`T^{}\left(K\right)`$ | $`0`$ | $`0.0`$ | $`9`$ | $`208`$ | $`736`$ | $`1066`$ |
| Table 1: System with N+M=14 and normalized couplings. | | | | | | |
From the table above, we conclude that there can be a large temperature range for which a SDW enhancement is observed above the superconducting instability. Both temperature scales are seen to decrease as the length of nested segments is reduced. This reduction is accompanied by an increase in the minimal distance to the unstable fixed point along the flow. This leads to a less pronounced intermediate regime, smaller exponents for the SDW response, and consequently a smaller magnitude of SDW correlations.
In summary, we have shown that precursor effects are predicted by a controlled weak coupling RG approach to fermionic systems. The flow approaches an unstable fixed point for systems with repulsive interactions, before being attracted to the final stable superconducting fixed point controlling the low temperature physics of the system. SDW response functions are enhanced and dominate near the intermediate fixed point, before the d-wave superconducting response finally diverges. We believe that Umklapp processes are likely to strengthen this picture. Close to half-filling, they are not expected to compete with SDW ordering, but with superconductivity, thereby reducing $`T_c`$ without affecting the cross-over energy scale. In spite of the weak coupling limit chosen here, we observe striking similarities with the pseudogap behavior of the HTC. This suggests that the strength of interactions in real materials may compensate the deviations of the observed Fermi surface from perfect nesting.
The authors acknowledge discussions with D. Zanchi, K. Le Hur and R. Dias. This work received traveling support from ICCTI/CNRS, Proc.423. FVA also benefited from the MCT Praxis XXI Grant No. 2/2.1/Fis/302/94.
|
warning/0007/astro-ph0007173.html
|
ar5iv
|
text
|
# THE EVOLVED RED STELLAR CONTENT OF M32
## 1 INTRODUCTION
The Local Group compact elliptical galaxy M32 has long been recognized as an important stepping-stone for interpreting the stellar content of more distant galaxies. The structural characteristics of M32 appear to be related to massive classical ellipticals (e.g. Kormendy 1985). Nevertheless, M32 is also the closest example of a class of compact elliptical galaxies that are junior partners in hierarchal systems (Prugniel, Nieto, & Simien 1987), suggesting that the high central concentration may be the result of tidal interactions (e.g. Faber 1973). On the other hand, Nieto & Prugniel (1987) find that the separation between M31 and M32 decays at a rate of $`23`$ kpc Gyr<sup>-1</sup>. As a result, the interactions between M31 and M32 were likely less intense in the past, and Nieto & Prugniel (1987) conclude that the structural characteristics and mass of M32 have not changed dramatically with time. While it is a matter of debate whether or not M32 and classical ellipticals share a common evolutionary heritage, it is clear that there is little hope of understanding the stellar contents of more distant galaxies if we can not understand the stellar content of a nearby, resolved system like M32.
M32 has been the target of numerous spectroscopic investigations, and these have detected signatures of a relatively blue main sequence turn-off (O’Connell 1980, Rose 1985), and an extended asymptotic giant branch (AGB) (Davidge 1990), both of which are indicative of an intermediate-age population. M32 also has a very red ultraviolet/visible color when compared with other elliptical galaxies (Burstein et al. 1988), and this can be explained if a large intermediate age population is present (e.g. Ohl et al. 1998, Yi et al. 1999).
During the past decade there has been increased emphasis on studies of the resolved stellar content of M32. Imaging of the outer regions of M32 at visible wavelengths failed to detect an extended AGB (Freedman 1989, Davidge & Jones 1992; Grillmair et al 1996). This is not unexpected, as the visible light from highly evolved metal-rich giants is affected by line blanketing (Bica, Barbuy, & Ortolani 1991), with the result that metal-rich AGB-tip stars can be significantly fainter at visible wavelengths than less evolved metal-poor stars. The effects of line blanketing are greatly reduced at wavelengths longward of $`1\mu `$m (e.g. Davidge & Simons 1994), and infrared imaging surveys have discovered AGB stars as bright as M$`{}_{K}{}^{}=9`$ in M32 (Freedman 1992, Elston & Silva 1992), which are much more luminous than AGB stars in Galactic globular clusters.
The outer regions of M32 contain stars spanning a range of metallicities (Freedman 1989; Davidge & Jones 1992), with the peak of the metallicity distribution function (MDF) occuring between \[Fe/H\] = – 0.2 and – 0.3 (Grillmair et al. 1996). The MDF of M32 at intermediate radii can not be fit with a simple one-zone enrichment model (Grillmair et al. 1996), suggesting a complex chemical enrichment history, possibly like that proposed for the massive elliptical galaxy NGC 5128 by Harris, Harris, & Poole (1999). Long slit spectra (Cohen 1979, Davidge, de Robertis, & Yee 1990, Davidge 1991, Hardy et al. 1994) and the spatial distribution of resolved and unresolved UV sources (Brown et al. 1998) indicate that population gradients are present, with mean metallicity being one of the parameters that likely changes with radius. While the central regions of M32 contain strong-lined stars that originate from a very metal-rich component (Rose 1994), only a modest population of UV-bright sources, which are thought to have evolved from the extreme horizontal branch (EHB) of a very metal-rich population, have been detected in M32, and Brown et al. (1998) conclude that the progenitors of these objects account for only 0.5% of all main sequence stars.
In the present study, high angular resolution near-infrared images of two fields, sampling different regions of the galaxy, are used to investigate the evolved red stellar content of M32. In addition to probing highly evolved stars, observations at infrared wavelengths also provide important information for decoupling the effects of age and metallicity in integrated light measurements (Davidge 1991). Finally, the contrast between the reddest, most evolved, stars and the unresolved, relatively blue, body of the galaxy is enhanced in the infrared with respect to the visible, making it easier to resolve evolved stars in crowded fields.
The observations, the data reduction procedures, and the photometric measurements are discussed in §2. The stellar contents of the two fields are investigated in §3, and it is demonstrated that the density of the brightest AGB stars scales with surface brightness at distances in excess of a few arcsec from the nucleus, indicating that the relative size of the intermediate age component measured with respect to the rest of the galaxy is constant over a large range of radii. However, within 1 arcsec of the nucleus the integrated near-infrared broad-band colors and the $`2.3\mu `$m CO index vary with radius. A brief summary and discussion of the results follows in §4.
## 2 OBSERVATIONS, REDUCTIONS, AND PHOTOMETRIC MEASUREMENTS
The data were recorded with the Canada-France-Hawaii Telescope (CFHT) Adaptive Optics Bonnette (AOB) and KIR imager on the nights of UT September 6 and 7, 1998. The AOB has been described by Rigaut et al. (1998). The detector in KIR is a $`1024\times 1024`$ Hg:Cd:Te array with 0.034 arcsec pixels, so that the imaged field is $`34\times 34`$ arcsec.
A field that included the nucleus of M32 (the ‘inner’ field) was observed through $`J,H,Ks,2.2\mu `$m continuum, and $`2.3\mu `$m CO filters. The galaxy nucleus, which was used as the reference source for AO compensation, was offset 10 arcsec from the field center so that the minor axis of M32 could be sampled out to a distance of 40 arcsec. A complete observing sequence for each filter consisted of 5 co-added exposures recorded at 4 dither positions, which were offset in a $`0.5\times 0.5`$ arcsec square pattern. The total exposure times were 1200 sec in each of $`J,H,`$ and $`Ks`$, 1800 sec in the $`2.2\mu `$m continuum filter, and 2400 sec in CO. The seeing conditions were mediochre when these data were recorded, and the delivered image quality is 0.4 arcsec FWHM.
Deep $`J,H,`$ and $`Ks`$ images, with 1200 sec total exposure times per filter, were also recorded of a field 2.3 arcmin from the nucleus (the ‘outer’ field) centered on the star GSC 02801 – 02082, which was used as the reference source for AO compensation. This field is slightly offset from the minor axis of the galaxy, and the data were recorded using the observing sequence described for the inner field. The seeing conditions were very good, and the delivered image quality is 0.20 arcsec FWHM in $`J`$ and 0.15 arcsec in $`H`$ and $`Ks`$.
The data were reduced using the procedure described by Davidge & Courteau (1999), which consists of the following steps: (1) subtraction of dark frames, (2) division by dome flats, (3) subtraction of a DC sky level from each image, and (4) subtraction of fringe and thermal emission patterns, using calibration frames that were constructed from blank sky fields. The images for each field were then aligned and median-combined on a filter-by-filter basis. The final $`Ks`$ images of these fields are shown in Figures 1 and 2.
Stellar brightnesses were measured with the PSF-fitting program ALLSTAR (Stetson & Harris 1988), using PSFs and target lists generated with DAOPHOT (Stetson 1987). A single PSF was constructed for each field $`+`$ filter combination. The use of a single reference star for AO-compensation produces a variable PSF (‘anisoplanaticism’), and this is a source of photometric errors if a single PSF is adopted for the entire field. However, studies of globular clusters conducted with the CFHT AOB demonstrate that the photometric uncertainties introduced by PSF variations amount to only a few percent over the KIR field (Davidge & Courteau 1999; Davidge 2000a), and an examination of the PSF-fitting residuals in the inner and outer field M32 data indicate that PSF variations do not introduce uncertainties of more than a few percent. The modest scatter in the M32 outer field CMDs (§3) is further evidence that PSF variations do not seriously affect the current photometry.
The photometric calibration was defined using standard stars from Casali & Hawarden (1992) and Elias et al. (1982), which were observed throughout the 3 night run. The unresolved body of M32 creates a non-uniform background in the inner field that complicates efforts to measure stellar brightness, and this was removed using the iterative procedure described by Davidge, Le Fevre, & Clark (1991).
## 3 RESULTS
### 3.1 The Outer Field
The $`(K,HK)`$ and $`(K,JK)`$ CMDs of the M32 outer field are shown in Figure 3. The RGB and upper portions of the AGB are clearly evident in these data: the onset of the former causes a slight broadening in the color distribution when $`K17.8`$, while the latter extends from $`K=17.5`$ to $`K=16`$, and has a scatter of $`\pm 0.05`$ mag in $`HK`$ and $`JK`$. There is also a 0.2 mag gap in the CMDs near $`K=17.6`$.
Data from Table 1 of Freedman (1992) are compared with the current observations in the right hand panel of Figure 3. Freedman observed an area four times larger than the AOB field, and so it is not surprising that she detected sources that are brighter than those in the present dataset. The CMDs constructed from the AOB data have less scatter than the Freedman observations, and extend $`11.5`$ mag fainter in $`K`$. The ridgeline of the Freedman observations, which can be traced by identifying the colors with the highest concentration of data points, is located near the left hand side of her data distribution and matches the locus of the AOB data. A gap in the CMD immediately above the RGB-tip is also evident in the Freedman data.
Freedman suggested that the objects in her dataset with $`JK1.5`$ are carbon stars. Stars with this color are not seen in the current dataset, even though a significant number would have been expected based on the Freedman observations. While stars with $`JK1.5`$ are present in the Elston & Silva (1992) $`(K,JK)`$ CMD, the relative number of these objects with respect to bluer stars is much smaller than in the Freedman (1992) dataset. Freedman (1992) notes that some stars in her study have large photometric uncertainties, which may be as great as 0.5 mag in $`JH`$ and 0.2 mag in $`HK`$. We speculate that the reddest stars in her sample are objects with large photometric errors.
The $`K`$ LF of the outer field is plotted in Figure 4. Artificial star experiments indicate that incompleteness and photometric errors become significant when $`K=20`$, and so only data brighter than this are shown in Figure 4. The dashed line shows a power-law that was fit to the LF between $`K=18`$ and $`K=19.5`$ using the method of least squares. The fitted power-law has an exponent $`0.16\pm 0.06`$, which is near the low end of what is seen in the Galactic bulge (e.g. Davidge 2000b). When $`K18`$ the LF consistently falls below the extrapolated power-law fit, indicating that the RGB-tip produces a significant discontinuity in the $`K`$ LF, even though models predict that the brightness of the RGB-tip is sensitive to age and metallicity near $`2\mu `$m (see below).
The bolometric LF of stars in the outer field, calculated using the bolometric corrections for field giants shown in Figure 1b of Frogel & Whitford (1987), is plotted in Figure 5. A distance modulus of $`\mu _0=24.3`$ has been adopted, based on an RGB-tip brightness of $`I=20.5`$ for M32 (Davidge & Jones 1992), which is 0.1 mag brighter than that in M31 (Davidge 1993), for which $`\mu _0=24.4\pm 0.1`$ (van den Bergh 2000). The data have also been corrected for a foreground reddening of E(B–V) = 0.08 (Burstein & Heiles 1984).
The most luminous stars in the M32 outer field have M$`{}_{bol}{}^{}6`$, which is comparable to what is seen in Baade’s Window (e.g. Frogel & Whitford 1987). The dashed line in Figure 5 is the LF of inner bulge giants measured by Davidge (1998), which has been scaled to match the number of stars in the M32 field with M$`{}_{bol}{}^{}3.5`$. There is excellent agreement between the bulge and M32 LFs.
The brightnesses of the AGB-tip and RGB-tip from the z = 0.001 (\[M/H\] = –1.2), z = 0.004 (\[M/H\] = –0.6), and z = 0.020 (\[M/H\] = 0.1) Bertelli et al. (1994) models are compared with the observations in Figure 6. M<sub>K</sub> was computed from the M<sub>V</sub> and $`VK`$ entries in Tables 8, 9, and 11 of Bertelli et al. (1994), while $`JK`$ colors were calculated using the field giant color relations listed in Table 3 of Bessell & Brett (1988).
The model sequences in Figure 6 consistently fall to the right of the outer field observations, although the significance of this result is low due to uncertainties in, for example, the mixing length and the adopted relation between $`VK`$ and $`JK`$. If the distance to M32 has been overestimated then the agreement along the color axis will improve slightly. While there are uncertainties in the color calibration, the relative model-to-model color variations are likely reliable, and the scatter on the upper AGB is consistent with a dispersion of $`\pm 0.3`$ dex in \[M/H\] or $`\pm 0.1`$ dex in age.
Comparisons between models and observations that are based on brightness are more reliable than those that rely on color. The inner field observations suggest that the AGB-tip occurs near $`K=15.5`$ in M32 (§3.2), which is consistent with the brightest stars detected in the Freedman (1992) and Elston & Silva (1992) datasets, and corresponds to M$`{}_{K}{}^{}=8.8\pm 0.1`$ if $`\mu _0=24.3\pm 0.1`$. It is evident from Figure 6 that models with log(t<sub>Gyr</sub>) $`9.3`$ are required to match the AGB-tip brightness. Caution should be exercised when attempting to derive absolute ages from a highly evolved state such as the AGB-tip, and the age limits inferred from Figure 6 are uncomfortably close to those defined by the visible spectrum of M32, which indicate that the galaxy does not contain a large population with log(t<sub>Gyr</sub>) $`9.0`$ (Bica et al. 1990).
### 3.2 The Circumnuclear Region
In this section the stellar content in the inner M32 field is compared with that in the outer field. Mean age and/or metallicity may change rapidly with radius near the center of M32, and so the inner field was divided into three regions, sampling distinct stellar density regimes centered on the nucleus. The radial limits and the mean surface brightnesses of these regions are listed in Table 1. The effects of crowding in Region 1 are extreme, and the stellar content in this area will be investigated using integrated colors in §3.3.
The $`K`$ LFs of Regions 2 and 3 are shown in Figure 7. The lower stellar density in Region 3 means that the brightness in $`K`$ where incompleteness becomes significant, estimated from the point at which number counts in the LFs decline, is 0.5 – 1.0 mag fainter in Region 3 than in Region 2.
Many of the sources detected in Regions 2 and 3 are not individual stars, but blends of two or more moderately faint objects. The effects of blending become less severe with increasing brightness, and the densities of stars with $`K=16`$ are sufficiently low that the sources with $`K=15.5`$ in Regions 2 and 3 are likely individual stars. This can be checked by comparing the densities of $`K=15.5`$ stars in the Freedman (1992) and Elston & Silva (1992) datasets with those in Region 2 and 3 after correcting for differences in surface brightness.
The observed densities of $`K=15.5`$ stars in the inner field are $`240\pm 50`$ per square arcmin (Region 2) and $`14\pm 8`$ per square arcmin (Region 3). Freedman (1992) detected 4 stars with $`K`$ between 15.75 and 15.25 in a 1.11 square arcmin area 2 arcmin south of the M32 nucleus, where the surface brightness is $`\mu _r=21.9`$ mag per square arcmin (Kent 1987). If the spatial distribution of $`K=15.5`$ stars follows the integrated light profile of the galaxy then the densities of these objects in the circumnuclear region should be $`200\pm 100`$ per square arcmin (Region 2) and $`50\pm 25`$ per square arcmin (Region 3). Elston & Silva (1992) detected 10 stars with $`K`$ between 15.25 and 15.75 in a 16 square arcmin field 3 arcmin east of the nucleus, where the surface brightness is $`\mu _r=23.1`$ mag per square arcmin. The densities of $`K=15.5`$ stars predicted from these data is $`110\pm 11`$ per square arcmin (Region 2) and $`27\pm 3`$ per square arcmin (Region 3). The bright star densities predicted for Regions 2 and 3 from the Freedman and Elston & Silva studies agree within the estimated uncertainties, and the unweighted means are $`150\pm 50`$ (Region 2) and $`39\pm 13`$ (Regions 3). Thus, the observed densities of bright red stars in Regions 2 and 3 differ from the densities inferred from the outer regions of M32 at less than the $`2\sigma `$ level.
The calculations described above indicate that the the brightest red stars have a spatial distribution that follows the surface brightness profile of the galaxy to within a few arcsec of the nucleus. These comparisons can be extended to fainter stars by scaling the outer field LF to match the surface brightnesses in Regions 2 and 3. The effects of crowding are significant among all but the brightest sources in Regions 2 and 3, and statistical crowding corrections appropriate for each Region were applied to the outer field LF after calculating the probability that two stars in a given brightness interval occur in the same resolution element, the diameter of which equals the FWHM of the PSF, and thus produce a blended image that would be counted in the next brightest LF bin.
The outer field LF, scaled and corrected for crowding based on the stellar densities in Regions 2 and 3, is shown as a dashed line in Figure 7. While the agreement with the Region 3 LF is excellent when $`K17`$, the comparison with the Region 2 LF is inconclusive, as incompleteness becomes significant in Region 2 when $`K>16`$. The data thus indicate that at distances in excess of 20 arcsec from the center of M32, which is the inner radius of Region 3, the evolved red stellar content scales with $`\mu _r`$. Therefore, the intermediate-age population is not more centrally concentrated than other populations at distances in excess of 20 arcsec from the nucleus. While the density of $`K=15.5`$ stars suggests that this may also be the case for radii as small as 5 arcsec, which is the inner radius of Region 2, this can not be verified with fainter stars until data with better image quality are obtained.
Although M32 contains a metallicity gradient (e.g. Hardy et al. 1994), this will not have a major impact on the comparisons discussed above, as the radial changes in mean metallicity are modest. If $`\mathrm{\Delta }`$Mg<sub>2</sub>/$`\mathrm{\Delta }`$log(r) = $`0.030\pm 0.005`$ and $`\mathrm{\Delta }<Fe>/\mathrm{\Delta }`$log(r) = $`0.57\pm 0.11`$ (Hardy et al. 1994), then the difference in indices between the outer field and Region 3 ($`\mathrm{\Delta }`$log(r) = –0.7) is $`\mathrm{\Delta }`$Mg$`{}_{2}{}^{}=0.02`$ and $`\mathrm{\Delta }<Fe>=0.36`$, while between the outer field and Region 2 ($`\mathrm{\Delta }`$log(r) = –1.0) $`\mathrm{\Delta }`$Mg$`{}_{2}{}^{}=0.03`$ and $`\mathrm{\Delta }<Fe>=0.57`$. These correspond to $`\mathrm{\Delta }`$\[Fe/H\] = 0.1 dex (Region 3) and $`\mathrm{\Delta }`$\[Fe/H\] = 0.2 dex (Region 2) with respect to the outer field according to the solar metallicity models of Worthey (1994).
### 3.3 The Nucleus of M32
The radial behaviour of $`JH`$, $`JK`$ and CO near the center of M32 is investigated in Figure 8, where radial averages of these quantities in 0.4 arcsec annuli are plotted. $`JK`$ becomes bluer when $`r1`$ arcsec, while the CO index becomes larger. Given that $`JH`$ does not change markedly with radius, it appears that the color gradients are driven primarily by changes in the spectral-energy distribution (SED) at wavelengths longward of $`2\mu `$m. The colors in Figure 8 are insensitive to uncertainties in sky brightness, which was measured in the upper left hand corner of Figure 2, because they are restricted to the high surface brightness central regions of the galaxy.
Peletier (1993) used images with an angular resolution of 1.5 arcsec FWHM to investigate near-infrared color gradients in M32. The data in Table 2 of Peletier (1993) indicate that within the central few arcsec of M32 there is a clear tendency for $`JK`$ to weaken with decreasing radius, and for CO to strengthen, in qualtitative agreement with the trends in Figure 8.
Several studies have measured the visible and UV colors near the center of M32. Michard & Nieto (1991) examined data with 0.7 – 0.9 arcsec resolution and concluded that $`BR`$ and $`UB`$ do not vary with radius near the nucleus of M32. However, the measurements listed in Table 2 of Peletier (1993), indicate that $`UR`$ and $`BR`$ may redden towards smaller radii. Lauer et al. (1998) examined images recorded with WFPC2 and concluded that $`VI`$ does not change with radius, while $`UV`$ may redden with decreasing radius. Ohl et al. (1998) find that the color of M32 in the far-ultraviolet becomes significantly redder towards smaller radii, although this trend is defined with measurements at distances in excess of 2 arcsec from the center, and so may not be related to changes in stellar content near the nucleus. The published data indicate that visible and near-infrared colors do not change dramatically with radius in M32 when $`r1`$ arcsec, and none of the studies detect a central bluing in the UV and visible portions of the spectrum. When combined with the results in Figure 8, the published data suggest that the SED of M32 undergoes complex changes close to the nucleus, becoming bluer with decreasing radius longward of 2$`\mu `$m, not changing with radius at visible wavelengths, and possibly becoming redder with decreasing radius shortward of $`0.4\mu `$m.
The $`(JH,HK)`$ and $`(CO,JK)`$ diagrams, shown in Figure 9, provide a means of probing the SED near the nucleus of M32. When $`r1`$ arcsec the near-infrared SED is similar to that of star clusters in the Magellanic Clouds. However, at smaller radii the M32 data depart from a SED appropriate for star clusters, in the sense that while $`JH`$ remains at a value appropriate for stellar systems, $`HK`$ decreases by more than 0.1 mag. When $`r1`$ arcsec $`JK`$ decreases towards smaller radii, while the CO index increases.
The large nuclear CO index of M32 is not easily explained. Mobasher & James (1996), and James & Mobasher (1999) measured the strength of $`2.3\mu `$m CO absorption in a moderately large number of elliptical galaxies. These data were recorded through 2.4 arcsec wide slits, and hence are not dominated by light from the galaxy nuclei; nevertheless, these data still provide insight into the CO index of composite metal-rich systems. After transforming the spectroscopic CO measurements into the photometric CO system using the relation defined by Doyon, Joseph, & Wright (1994), it is concluded that CO$`{}_{max}{}^{}=0.21`$, compared with CO = 0.36 near the nucleus of M32. The only other stellar system having a published CO index with a spatial resolution comparable to the M32 observations is the Galactic Center, which would have CO = 0.26 if viewed at a distance of 1 Mpc (Davidge 2000c).
Luminous red giants in the vicinity of SgrA and red supergiants in the Galaxy and the Magellanic Clouds have CO indices comparable to the nucleus of M32 (Elias et al. 1985, Kenyon 1988, Davidge 1998). However, these stars have $`JK>1`$, and hence do not provide suitable templates for the nucleus of M32. Therefore, it appears that the near-infrared SED of the M32 nucleus differs from that of benchmark systems such as nearby bright giants, star clusters, and the central few arcsec of elliptical galaxies.
## 4 SUMMARY & DISCUSSION
Near-infrared images have been used to investigate the stellar content of the Local Group compact elliptical galaxy M32. AGB stars as bright as $`K=15.5`$, which were detected in earlier infrared studies of the outer regions of the galaxy, are also found near the center of the galaxy. The number density of these objects increases towards smaller radii, indicating that they are actual members of M32, and not luminous AGB stars in the outer disk of M31. Hence, these data confirm the presence of the intermediate-age population predicted by spectroscopic investigations. The bright AGB stars in the outer regions of M32 have $`JK1.5`$, and it is suggested that the very red objects detected by Freedman (1992) have large photometric errors.
The $`K`$ LF of the brightest AGB stars scales with integrated surface brightness to within 20 arcsec of the galaxy center, suggesting that the relative frequency of bright AGB stars, as measured with respect to fainter stars, does not change over most of the galaxy. This does not rule out the presence of a subtle ($`\pm 1`$ Gyr) age gradient in M32, since the brightness of the AGB-tip changes slowly with age in all but the youngest intermediate-age populations. The UV portion of the spectrum is very age-sensitive during intermediate epochs (Yi et al. 1999), and an age gradient, in the sense that younger stars occur at smaller radii, provides one explanation for the UV color of M32 increasing with decreasing radius (O’Connell et al. 1992).
The data presented in this paper indicate that the near-infrared SED changes within a few tenths of an arcsec of the center of M32, and that the center of the galaxy has colors that differ from those of star clusters and the main bodies of elliptical galaxies. The central properties of M32 are peculiar in other ways. For example, the ratio of central x-ray flux to mass is the smallest of any known super-massive black hole, suggesting that the central object may be fuel-starved, or is not an efficient accretor (Loewenstein et al. 1998). In addition, the gravitational field from the central object influences the central arcsec of M32 (van der Marel 1999), so variations in stellar content and/or morphology might be expected to occur in this area. However, the core of M32 lacks conspicuous structure at UV (Cole et al. 1998) and visible (Lauer et al. 1998) wavelengths. M32 also lacks a centrally concentrated blue population, and this is puzzling since stellar interactions in the dense nuclear regions might be expected to produce a large number of blue stragglers (Lauer et al. 1998). The changes in near-infrared photometric properties detected in the current study occur on an angular scale that is comparable to the region influenced by the central black hole (van der Marel 1999), suggesting a possible connection. The ‘stubborn refusal of M32 to reveal any departures from utter normalacy at small radii’ noted by Lauer et al. (1998) appears to break down when data spanning a broader range of wavelengths are considered.
The relative size of the intermediate-age population with respect to the rest of the galaxy can be estimated from the integrated brightness of AGB stars brighter than the RGB-tip. The integrated brightness of the outer field based on the Kent (1987) surface brightness profile, assuming that $`Vr=0.2`$ for a typical spheroidal galaxy (Kent 1987) and that $`VK=3.1`$ for M32 (Frogel et al. 1978), is $`K=11.8`$. For comparison, the integrated brightness of stars on the upper AGB is $`K=13.5`$, so luminous AGB stars contribute 20% of the total light from the outer field in $`K`$.
Frogel, Mould, & Blanco (1990) studied the near-infrared photometric properties of AGB stars in a sample of SMC and LMC clusters, and 11 of these clusters (NGCs 411, 419, 1751, 1806, 1846, 1978, 2108, 2121, 2154, 2213, and 2231) have SWB types V, V-VI, and VI, indicating that they have ages of a few Gyr, and integrated $`K`$ brightness measurements (Persson et al. 1983). The ratio of light in $`K`$ from bright AGB stars to that of the entire cluster is a well-defined quantity among these objects, with a mean value $`0.81\pm 0.02`$. Therefore, if the intermediate age component in the M32 outer field has an age of a few Gyr then this population accounts for roughly 25% of the total light in $`K`$. The M/L ratio in $`K`$ of an intermediate age population that has a Salpeter IMF and an age of a few Gyr is one-half to one-third that of an old population (Buzzoni 1989). If the main body of M32 is relatively old, then the intermediate-age population in the outer field thus accounts for roughly 10% of the total mass, with an estimated uncertainty $`\pm 5\%`$. For comparison, Bica et al. (1990) predict that the central regions of M32 contain an intermediate-age population that accounts for 15% of the mass. The good agreement between these two mass estimates, obtained using very different techniques in different parts of the galaxy, is further evidence that the intermediate-age population in M32 is uniformly distributed throughout the galaxy, as expected based on the comparisons in §3.2.
The presence of an intermediate age population accounting for 10% of the galaxy mass indicates that M32 had a significant ISM a few Gyr in the past. What was the source of the gas from which these stars formed? M31 is a nearby reservoir of star-forming material, and there are signs that M32 has interacted recently (Cepa & Beckman 1988) with material in the disk of this galaxy (Byrd 1976, 1978; Sofue & Kato 1981). However, if M32 was able to cull large amounts of gas from M31 during intermediate epochs then it must have been a special event, as subsequent passages through the disk of M31 have not produced noticeable bursts of star formation.
Mass ejected from evolved stars is another source of an ISM. Sage, Welch, & Mitchell (1998) estimate that stellar mass loss should have produced an ISM in M32 with mass $`3\times 10^5`$ M since the last interaction with M31. This is consistent with the mass of HI within the central 3 square arcmin, which Emerson (1974) measures to be $`6\times 10^5`$ M. However, based on the non-detection of CO 1 - 0 emission from M32, Sage et al. (1998) conclude that the mass of H<sub>2</sub> is $`5.1\times 10^3`$ M within the central 160 pc of the galaxy.
The mean separation between M32 and M31 was significantly greater in the past (Nieto & Prugniel 1987), and M32 could have built up a larger ISM from stellar mass loss during earlier epochs due to decreased tidal interactions with M31. If the bright AGB stars formed from gas that originated in M32 then these objects should be among the most metal-rich in the galaxy, with metallicities within a factor of 2 of solar according to the Grillmair et al. (1996) MDF. High-quality visible and near-infrared spectra can be obtained of the brightest AGB stars using instruments on 4 or 8 metre telescopes, thereby providing a direct means of measuring the metallicities of these objects. If the AGB stars are found to be relatively metal-poor then a source of gas external to M32 would seem likely; for example, if star formation in M32 during intermediate epochs was fueled by the accretion of the high velocity clouds that have been interpreted as the remnants of the material from which the Local Group formed, then the AGB stars would have \[Fe/H\] $`1`$ (Blitz et al. 1999).
FIGURE CAPTIONS
|
warning/0007/quant-ph0007109.html
|
ar5iv
|
text
|
# Photon Localization and Vacuum Noise in Optical Measurements
## Abstract
Description of detection and emission in terms of the photon localization is discussed. It is shown that the standard representation of the plane waves of photons should be revised to take into consideration the boundary conditions caused by the presence of quantum emitters and detectors. In turn, the change of the boundary conditions causes spatially inhomogeneous structure of the electromagnetic vacuum which leads to the increase of the vacuum noise over the level predicted within the framework of the model of plane waves of photons.
Since the early days of quantum theory of light, the problem of localizing photons has attracted a great deal of interest (e.g., see for recent discussion). The point is that the photon operators of creation and destruction are defined in all space. At the same time, the intensity measurement by means of a photodetector with finite sensitive area $`\sigma `$ presupposes a kind of the photon localization, at least in vicinity of $`\sigma `$ . The transformation of photons into an electronic signal in photodetectors is not the only onion in the stew. Another example of some considerable interest is provided by the emission and absorption of radiation by atoms and molecules.
We now note that the electromagnetic field is usually quantized as though it is free and propagates in empty space. This model leads to well-known plane waves of photons, corresponding to the quantized translation invariant solutions of the homogeneous Helmholtz wave equation . The presence of atoms or surfaces which are able to interact with photons leads to the change of boundary conditions and hence to the violation of translational symmetry. For example, the presence of a single point-like atom causes the multipole structure of the field which can be described in terms of quantized spherical waves . The latter case is specified by the $`SO(3)`$ symmetry rather than translational symmetry.
In this note we show that the taking account of the boundary conditions at both emission and detection of the field from the plane waves of photons lead to more adequate picture of the photon localization. We also show that the change of the boundary conditions strongly influences the zero-point oscillations of the field strengths which causes a deterioration of quantum limit of precision of measurements.
Consider first an atom located at the origin of the reference frame spanned by the complex base vectors
$`\stackrel{}{\chi }_\pm ={\displaystyle \frac{\stackrel{}{e}_x\pm i\stackrel{}{e}_y}{\sqrt{2}}},\stackrel{}{\chi }_0=\stackrel{}{e}_z.`$ (1)
These vectors formally coincide with the states of spin $`1`$ of the photon . Since the quantum electrodynamics defines the spin states of photons as the polarization , we can choose to interpret $`\stackrel{}{\chi }_\pm `$ as the unit vectors of transversal polarization with either positive or negative helicity, while $`\stackrel{}{\chi }_0`$ is the unit vector of linear polarization in the $`z`$-direction. The third spin state is forbidden in the case of plane waves of photons due to the translational invariance, while allowed in the case of spherical waves of photons . An arbitrary vector $`\stackrel{}{A}`$ is expanded in this basis as follows
$`\stackrel{}{A}={\displaystyle \underset{\mu }{}}(1)^\mu \stackrel{}{\chi }_\mu A_\mu .`$ (2)
The positive-frequency part of the electric field strength of the monochromatic multipole field is then defined as having components
$`E_\mu (\stackrel{}{r})=ik\gamma {\displaystyle \underset{\lambda }{}}{\displaystyle \underset{j=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=j}{\overset{j}{}}}V_{\lambda jm\mu }(\stackrel{}{r})a_{\lambda jm},`$ (3)
where $`\lambda =E,M`$ denotes the type of radiation (either electric or magnetic), $`\gamma `$ is the normalization factor, $`j,m`$ are the angular momentum quantum numbers. In the classical picture, the complex field amplitudes are defined by the source . To obtain the quantum counterpart, we have to subject these amplitudes to the Weyl-Heisenberg commutation relations
$`[a_{\lambda jm},a_{\lambda ^{}j^{}m^{}}^+]=\delta _{\lambda \lambda ^{}}\delta _{jj^{}}\delta _{mm^{}}.`$ (4)
The mode functions in (1) have the form
$`V_{Ejm\mu }(\stackrel{}{r})={\displaystyle \frac{1}{\sqrt{2j+1}}}`$ (5)
$`\times [\sqrt{j}f_{j+1}1,j+1,\mu ,m\mu |jmY_{j+1,m\mu }`$ (6)
$`\sqrt{j+1}f_{j1}1,j1,\mu ,m\mu |jmY_{j1,m\mu }],`$ (7)
$`V_{Mjm\mu }(\stackrel{}{r})=f_j(kr)1,j,\mu ,m\mu |jmY_{j,m\mu },`$ (8)
where the radial function $`f_{\mathrm{}}(kr)`$ is represented either by the spherical Bessel function $`j_{\mathrm{}}(kr)`$, in the case of standing waves in a spherical cavity, or by the spherical Hankel functions of the first and the second kind, describing the outgoing and converging spherical waves respectively. Here $`\mathrm{}|jm`$ denotes the Clebsch-Gordon coefficient of vector addition of the spin and orbital parts of the total angular momentum and $`Y_{\mathrm{},m\mu }`$ is the spherical harmonics.
In view of (2), the zero-point oscillations of the electric field strength (1) have the form
$`C_E(\stackrel{}{r})0|(\stackrel{}{E}(\stackrel{}{r})+\stackrel{}{E}^+(\stackrel{}{r}))^2|0`$ (9)
$`={\displaystyle \underset{\mu }{}}0|E_\mu (\stackrel{}{r})E_\mu ^+(\stackrel{}{r})|0`$ (10)
$`=(k\gamma )^2{\displaystyle \underset{\mu }{}}{\displaystyle \underset{\lambda ,j,m}{}}|V_{\lambda jm\mu }(\stackrel{}{r})|^2.`$ (11)
To make a comparison, we remind here that the zero-point oscillations in the case of the monochromatic plane waves have the form
$`C_{plane}=2(k\gamma ^{})^2`$ (12)
everywhere.
It is seen that, unlike (5), the spherical waves of photons have the spatially inhomogeneous zero-point oscillations. It is now a straightforward matter to show that (4) strongly exceeds the standard level, given by (5), at least in some vicinity of the origin (atom), while tends to (5) as $`kr1`$. A more detailed examination shows that $`C_E(\stackrel{}{r})C_{plane}`$ at $`kr2.5`$ which gives the distance of the order of $`0.3\mathrm{\Lambda }`$ where $`\mathrm{\Lambda }`$ is the wavelength. Let us stress that this distance is of the order of typical interatomic separation in experiments with trapped Ridberg atoms .
We now stress that the above results have been obtained under the only assumption that the atom exists at the origin independent of whether we use it for emission or detection. Therefore, the strong increase of the vacuum noise in vicinity of the atom should influence both the emission and detection processes. As a simple model of complete Hertz-type optical experiment, we consider the two identical atoms separated by distance $`d`$. The first atom (source) is prepared initially in the excited state of some multipole transition, while the second atom (detector) is in the ground state. Then, the measurement consists in the emission and successive absorption of a photon.
It is clear that in order to take into account the initial localization of photon within the source, the radiation should be described in terms of the outgoing spherical wave focused on the first atom. In turn, the final localization within the detector, assumes the converging spherical wave focused on the second atom. To combine these two processes into the common picture, we have to describe the filed as the superposition of outgoing and converging waves. The coefficients of this superposed state should be defined by the boundary conditions for the real radiation field. Taking into account the recent investigation , we anticipate that this model obeys the causality principle of the electrodynamics. In view of the position dependence in (3) and (4), it is clear that, at far distances ($`d\mathrm{\Lambda }`$), the major contribution into the vacuum noise of measurement comes from the detecting atom, while, at intermediate and short distances, the noise of measurement is increased due to the influence of the source atom.
Consider now the measurement of a plane photon by a photodetector. At far distances, the photon is described by a unique wave vector $`\stackrel{}{k}`$. The Mandel’s localization in vicinity of the sensitive area $`\sigma `$ assumes that the wave converges to $`\sigma `$. This means that there is a variety of directions of the wave vectors near $`\sigma `$, although all of them have the same length. This picture, based on the taking account of the boundary conditions, can be described by a proper expansion over spherical waves. In view of the above discussion, it should lead to the increase of the vacuum noise of measurement over the level (5).
Let us briefly summarize the results. First of all, it is clear that the above results represent an extension and detailing of the Mandel’s model of the photon localization . It has been shown that the description of the photon localization in the process of detection and emission needs more adequate consideration of the boundary conditions, leading to a violation of the translation invariance inherent in the conventional model of the plane waves of photons. This violation leads to the qualitative change of the structure of the electromagnetic vacuum state. In particular, the zero-point oscillations are concentrated in vicinity of atoms, molecules, photodetectors and other local objects which are able to interact with photons. The level of the zero-point oscillations in vicinity of the emitting and measuring devices can strongly exceeds that calculated as though the field consists of the plane waves of photons. This leads to a deterioration in the estimation of the quantum limit of precision of measurement.
The above results can be important for different quantum optical measurements especially in the engineered entanglement based on the trapped atoms , in the experiments with atomic beams and single-atom lasers and in the quantum polarization measurement .
|
warning/0007/cond-mat0007067.html
|
ar5iv
|
text
|
# Duality in the Azbel-Hofstadter problem and the two-dimensional d-wave superconductivity with a magnetic field
## Abstract
A single-parameter family of lattice-fermion model is constructed. It is a deformation of the Azbel-Hofstadter problem by a parameter $`h=\mathrm{\Delta }/t`$ (quantum parameter). A topological number is attached to each energy band. A duality between the classical limit ($`h=+0`$) and the quantum limit ($`h=1`$) is revealed in the energy spectrum and the topological number.
The model has a close relation to the two-dimensional d-wave superconductivity with a magnetic field. Making use of the duality and a topological argument, we shed light on how the quasiparticles with a magnetic field behave especially in the quantum limit.
Two-dimensional Dirac fermions with a gauge field are of current interest, e.g., in the context of the vortex state in a two-dimensional d-wave superconductivity . In our study, Dirac fermions with a gauge field are realized on a two-dimensional lattice. It is a single-parameter deformation of the Azbel-Hofstadter problem. In this paper, the parameter $`h`$ is called quantum parameter. A topological number is assigned for each energy band . As the quantum parameter is varied continuously from the classical limit ($`h=+0`$) to the quantum limit ($`h=1`$), the energy spectrum is reconstructed through the change of each topological number (plateau transition). Although the two limits are not connected adiabatically, we found that there is a duality between the classical and the quantum regime.
The model has a close relation to the two-dimensional d-wave superconductivity with a magnetic field. Applying the duality and a topological argument, we provide insights into the quasiparticle spectrum. In the quantum limit, interference effects become relevant especially at zero energy. The existence of edge states is discussed as well. It reflects a non-trivial topology of each energy band.
Let us define a key Hamiltonian in our paper, which is a single-parameter family of lattice fermion model. It is a deformation of the Azbel-Hofstadter problem. The Hamiltonian is $`=_{l,m}𝐜_l^{}_{lm}𝐜_m`$ with
$`_{lm}=e^{iA_{lm}}\left(\begin{array}{cc}t_{lm}^0& \mathrm{\Delta }_{lm}^0\\ \mathrm{\Delta }_{ml}^{0}{}_{}{}^{}& t_{ml}^0\end{array}\right),`$ (1)
$`𝐜_n^{}=\left(\begin{array}{c}c_n^{}c_n^{}\end{array}\right)`$, $`𝐜_n=\left(\begin{array}{c}c_n\\ c_n\end{array}\right)`$, $`A_{lm}=A_{ml}𝐑`$ and $`{\displaystyle \underset{\text{}}{}}A_{lm}/2\pi =\varphi =p/q`$ ($`p`$ and $`q`$ are coprime integers and the summation runs over four links around a plaquette). Here $`l,m𝐙^2`$, $`t_{m\pm (1,0),m}^0=t_{m\pm (0,1),m}^0=t`$, $`\mathrm{\Delta }_{m\pm (1,0),m}^0=\mathrm{\Delta }_{m\pm (0,1),m}^0=\mathrm{\Delta }`$ $`(t,\mathrm{\Delta }𝐑)`$ and the other matrix elements are zero. $`t`$ is set to be a unit energy and a relevant parameter, quantum parameter, is defined by $`h=\mathrm{\Delta }/t`$. The relation of this model to the two-dimensional d-wave superconductivity with a magnetic field is discussed later. In the classical limit $`h=+0`$, this model decouples to two essentially equivalent Hamiltonians. It is the Azbel-Hofstadter Hamiltonian $`_{l,m}c_l^{}e^{iA_{lm}}t_{lm}^0c_m`$. The energy spectrum at $`\varphi =0`$ is given by $`E=\pm \sqrt{A(k)^2+|B(k)|^2}`$ where $`A(k)=2t(\mathrm{cos}k_x+\mathrm{cos}k_y)`$, $`B(k)=2\mathrm{\Delta }(\mathrm{cos}k_x\mathrm{cos}k_y)`$ and $`k[\pi ,\pi ]\times [\pi ,\pi ]`$. The upper and lower bands touch at four points $`(\pm \pi /2,\pm \pi /2)`$ in the Brillouin zone. The low-lying excitations around the gap-closing points are described by massless Dirac fermions.
One of the basic observables is a topological number for the $`n`$-th band, $`𝒞_n`$ . It is
$`𝒞_n={\displaystyle \frac{1}{2\pi i}}{\displaystyle }d𝐤\widehat{z}(`$ $`\times `$ $`𝐀_n),𝐀_n=u_n(𝐤)||u_n(𝐤)`$ (2)
where $`=/𝐤`$ and $`|u_n(𝐤)`$ is a Bloch vector of the $`n`$-th band. The integration $`𝑑𝐤`$ runs over the Brillouin zone (torus). The non-zero topological number results in the existence of edge states. In order to see it, put the system on a cylinder and introduce a fictitious flux through the cylinder (it is equivalent to a twist in the boundary condition). The edge states move from one boundary to the other as the fictitious flux quanta $`hc/e`$ is added adiabatically. The number of carried edge states coincides with the summation of topological numbers below the Fermi energy. Due to the topological stability, a singularity necessarily occurs with the change of the topological number (plateau transition ). The singularity is identified with an energy-gap closing on some points in the Brillouin zone.
In Figs. 1-3, the energy spectra are shown. As $`h`$ is varied continuously from the classical limit ($`h=+0`$) to the quantum limit ($`h=1`$), the energy spectrum is reconstructed through the plateau transitions. Although the two limits are not connected adiabatically, a main feature in the data is that there is a symmetry between the classical and the quantum regime. It leads to a claim that
there is a faithful correspondence between $`\varphi =(1/2+p/q)`$ in the classical limit $`h=+0`$ and $`\varphi =p/q`$ in the quantum limit $`h=1`$’.
We call this phenomena duality. As discussed above, this model reduces to a doubled Azbel-Hofstader problem in the classical limit $`h=+0`$. The duality is an analogue of a statistical transmutaion (composite fermion picture) in the fractional quantum Hall effect. In the composite fermion picture, a locally attached flux to each fermion is replaced with a global uniform flux. In our case, pairing effects (off-diagonal order) play the role of shifting a flux globally. It is reminiscent of a symmetry between the d-wave pairing and the $`\pi `$-flux, in other words, the $`SU(2)`$ symmetry. Here some comments are in order. At $`\varphi =0`$ and $`1/2`$, we proved analytically that the claim is exact. Moreover, as shown in Figs.1 and 3, topological numbers are consistent with the claim.
As an application of the duality, let us study the two-dimensional $`d`$-wave pairing model with a magnetic field especially in the quantum limit ($`h=1`$). The pairing model is $`H=_{l,m}𝐜_l^{}H_{lm}𝐜_m`$ with
$`H_{lm}=\left(\begin{array}{cc}t_{lm}& \mathrm{\Delta }_{lm}\\ \mathrm{\Delta }_{ml}^{}& t_{ml}\end{array}\right).`$ (3)
Under the unitary transformation $`c_nd_n`$, $`c_nd_n^{}`$ (for $`n`$), it becomes $`H=_{l,m}[d_l^{}t_{lm}d_m+d_l^{}t_{lm}d_m+d_l^{}\mathrm{\Delta }_{lm}d_m^{}+d_m\mathrm{\Delta }_{lm}^{}d_l].`$ It is equivalent to the Bogoliubov-de Gennes Hamiltonian for the singlet superconductivity. Here $`t_{lm}^{}=t_{ml}`$ (hermiticity) and $`\mathrm{\Delta }_{lm}=\mathrm{\Delta }_{ml}`$ (SU(2) symmetry) are imposed as well. It satisfies a relation $`(\sigma _yH_{lm}\sigma _y)^{}=H_{lm}`$, which results in a particle-hole symmetry in the energy spectrum. The two-dimensional $`d`$-wave pairing model with a magnetic field is defined by the pairing model (3) with
$`t_{lm}`$ $`=`$ $`\mathrm{exp}(iA_{lm})t_{lm}^0,`$ (4)
$`\mathrm{\Delta }_{lm}`$ $`=`$ $`\mathrm{exp}(i\phi _{lm})\mathrm{\Delta }_{lm}^0`$ (5)
where $`A_{lm}=A_{ml},\phi _{lm}=(\phi _l+\phi _m)/2𝐑`$ so that $`t_{lm}^{}=t_{ml}`$ (hermiticity) and $`\mathrm{\Delta }_{lm}=\mathrm{\Delta }_{ml}`$ (SU(2) symmetry) are satisfied. Performing a gauge transformation
$`𝐜_n\left(\begin{array}{cc}e^{i\phi _n}& 0\\ 0& 1\end{array}\right)𝐜_n,`$ (6)
we obtain
$`H_{lm}=e^{iA_{lm}}\left(\begin{array}{cc}t_{lm}^0e^{2iv_{lm}}& \mathrm{\Delta }_{lm}^0e^{iv_{lm}}\\ \mathrm{\Delta }_{ml}^{0}{}_{}{}^{}e^{iv_{lm}}& t_{ml}^0\end{array}\right)`$ (7)
where $`v_{lm}=(\phi _l\phi _m)/2A_{lm}`$. It is a lattice realization of the Hamiltonian discussed in ref. and $`v_{lm}`$ corresponds to the superfluid velocity. In the case $`v_{lm}=0`$, it reduces to the Hamiltonian (1) and the quasiparticle spectrum consists of ’Landau levels’ $`E=\pm \omega _H\sqrt{n}`$ ($`n𝐍`$) in the continuum limit .
In the following, we set a period for $`A_{lm}`$ and $`v_{lm}`$ to be $`l_x\times l_y`$. Here we emphasize that, in the context of superconductivity, $`A_{lm}`$ and $`v_{lm}`$ are determined in a self-consistent way. Moreover, the spatial variation of $`|\mathrm{\Delta }_{lm}|`$ plays a crucial role especially near a vortex core. In our study, however, $`A_{lm}`$ and $`v_{lm}`$ are treated as adjustable parameters and focus is put on the duality or topological arguments which do not depend on the detail of the potential.
In Fig.4, an example of the density of states (DoS) is shown for the $`d`$-wave pairing model with a magnetic field in the quantum limit ($`h=1`$). Although the weak-field regime ($`\varphi 0`$) is relevant to reality, we show the case $`\varphi =1/5`$ for clarity. In the weak-field regime, the number of energy bands increases but the following arguments are robust.
As the system at $`\varphi =p/q`$ approaches the quantum regime ($`h1`$), the quasiparticle spectrum becomes close to that of the Azbel-Hofstadter problem at $`\varphi =(1/2+p/q)`$. The Landau levels are mixed due to lattice effects and $`v_{lm}`$. It causes a singularity at zero energy and the broadening of each level (’Landau bands’). It is due to quantum interference effects through spatially varying potentials. It is an analogue of the vanishing DoS at zero energy in random Dirac fermions . In the case of $`\varphi =p/q`$ ($`q`$=odd), it is a natural consequence of the duality. In other words, it can be interpreted by the fact that there exists a singularity at zero energy i.e. $`\rho (E=0)=0`$ in the Azbel-Hofstadter problem at $`\varphi =(1/2+p/q)`$. It is also to be noted that, apart from the singularity and the broadening, the energy spectrum of the Azbel-Hofstadter problem at $`\varphi =(1/2+p/q)`$ $`(p=1,q1)`$ is $`\pm \omega _H\sqrt{n}`$ near the band center and $`\pm (\omega n+C)`$ near the band edge ($`n𝐍`$).
Now put the system on a cylinder (periodic in the $`y`$ direction and open in the $`x`$ direction) and let us study the basic properties of edge states in the d-wave pairing model with a magnetic field. The Schr$`\ddot{o}`$dinger equation for the Hamiltonian (3) is
$`{\displaystyle \underset{j}{}}\left(\begin{array}{cc}t_{ij}& \mathrm{\Delta }_{ij}\\ \mathrm{\Delta }_{ji}^{}& t_{ji}^{}\end{array}\right)\left(\begin{array}{c}u_j\\ v_j\end{array}\right)=E\left(\begin{array}{c}u_i\\ v_i\end{array}\right).`$ (8)
Decompose all the sites $`𝒩`$ into two sublattices $`A`$ and $`B`$ ($`AB=𝒩`$ and $`AB=\varphi `$) where $`t_{ij}`$ and $`\mathrm{\Delta }_{ij}`$ connecting the same sublattice are zero. Define $`\overline{w}_k`$ by $`\overline{w}_k=+w_k`$ ($`kA`$) and $`\overline{w}_k=w_k`$ ($`kB`$). Then it follows from the SU(2) symmetry that
$`{\displaystyle \underset{j}{}}\left(\begin{array}{cc}t_{ij}& \mathrm{\Delta }_{ij}\\ \mathrm{\Delta }_{ji}^{}& t_{ji}^{}\end{array}\right)\left(\begin{array}{c}\overline{v}_j^{}\\ \overline{u}_j^{}\end{array}\right)=E\left(\begin{array}{c}\overline{v}_i^{}\\ \overline{u}_i^{}\end{array}\right).`$ (9)
It leads to the claim that, if an edge state exists, a paired (degenerate but linearly independent) edge state can be constructed which is localized in the same side.
Next we shall discuss the existence of edge states. At first, set $`\varphi =p/q`$, $`v_{lm}=0`$ and focus on an energy gap with a non-zero topological number (see, for example, Fig.3. The duality implies that a topological number for a generic gap is non-zero even number in the quantum regime. It is consistent with the previous result ). As discussed above, a non-zero topological number leads to the existence of edge states in the gap. In other words, the existence of edge states is topologically stable. Next consider the case when $`v_{lm}`$ is finite. As the $`v_{lm}`$ is varied from zero to the finite value, the Landau bands are deformed and can be overlapped like a semimetal. We can, however, observe edge states due to the topological stability, when the plateau transition is absent in the deformation. In fact, see Fig.5. The existence of edge states can be confirmed in each energy gap. It reflects a non-trivial topology of each energy band. In the above discussion, we have employed a cylinder. It is possible to consider the same problem on a geometry with ’defects’ (e.g. annulus). In the case, the states analogous to the edge states on a cylinder may be bound to the defects.
In summary, a single-parameter family of lattice-fermion model is constructed in two dimensions. The parameter $`h=\mathrm{\Delta }/t`$ is called quantum parameter. A duality is revealed in the model between the classical limit ($`h=+0`$) and the quantum limit ($`h=1`$). Employing the duality and a topological argument, we provide insights into how the quasiparticles with a magnetic field behave especially in the quantum limit. A more detailed study including a self-consistent potential and the fluctuation is left as a future problem. It is crucial for the analysis of dynamical properties. As discussed in the context of the integer quantum Hall effect, when the potential is sufficiently strong, it can bring the system to a totally different state .
One of the authors (Y.M.) thanks H. Matsumura for valuable discussions. This work was supported in part by Grant-in-Aid from the Ministry of Education, Science and Culture of Japan. The computation has been partly done using the facilities of the Supercomputer Center, ISSP, University of Tokyo.
|
warning/0007/nucl-th0007030.html
|
ar5iv
|
text
|
# Chiral Phase Transition within Effective Models with Constituent Quarks
## I Introduction
Chiral symmetry is spontaneously broken in the QCD vacuum. Lattice QCD simulations at nonzero temperature $`T`$ and zero baryon-chemical potential $`\mu _B`$ indicate that chiral symmetry is restored above a temperature $`T150`$ MeV . Such temperatures are believed to be created in nuclear collisions at ultra-relativistic energies. Consequently, a phase where chiral symmetry is transiently restored may be formed in these collisions. The subsequent expansion cools the system and takes it to the final hadronic state, where chiral symmetry is again spontaneously broken.
It is important to determine the order of the chiral transition, as this influences the dynamical evolution of the system. For instance, it has been shown that a first order transition may lead to a deflagration wave and to a “stall” in the expansion of the system . It has also been shown that a first order transition in rapidly expanding matter may manifest itself by strong non-statistical fluctuations due to droplet formation . In the case of strong supercooling it may lead to large fluctuations due to spinodal decomposition . In a second order phase transition one may expect the appearance of critical fluctuations due to a large correlation length . Experimentally, large-acceptance detectors are now able to measure average as well as event-by-event observables, which in principle allow to distinguish between scenarios with a first order, a second order, or merely a crossover type of a phase transition.
Theoretically, the QCD phase diagram in the $`(T,\mu _B)`$ plane has recently received much attention (see ). QCD with $`N_f=2`$ flavors of massless quarks has a global $`SU(2)_L\times SU(2)_R`$ symmetry. This symmetry is spontaneously broken in the QCD vacuum, such that the order parameter $`\varphi ^{ij}\overline{q}_L^iq_R^j`$ acquires a non-vanishing expectation value, where $`q^i`$ is the quark field ($`i,j`$ are the flavor indices). At zero baryon-chemical potential, the effective theory for this order parameter is the same as the $`O(4)`$ model which has a second order phase transition. Therefore, by universality arguments , the chiral transition in $`N_f=2`$ QCD is likely to be of second order at $`\mu _B=0`$. Nonzero quark masses introduce a term in the QCD Lagrangian which explicitly breaks chiral symmetry. Then, the second order transition becomes crossover.
At nonzero baryon-chemical potential, it is more difficult to infer the order of the chiral transition from universality arguments . One commonly resorts to phenomenological models to describe the chiral transition in this case. Depending on the parameters of these models, they predict a first order, a second order, or a crossover transition. However, if there is a second order phase transition for $`\mu _B=0`$ and nonzero $`T`$ and a first order transition for small $`T`$ and nonzero $`\mu _B`$, then there exists a tricritical point in the $`(T,\mu _B)`$ plane where the line of first order phase transitions meets the line of second order phase transitions. For nonzero quark masses, this tricritical point becomes a critical point.
It has recently been proposed that this point could lead to interesting signatures in heavy-ion collisions at intermediate energies, if the evolution went through or close to this critical point. At this point, susceptibilities (e.g. the heat capacity) diverge, and the order parameter field becomes massless and consequently fluctuates strongly, which could be detected in event-by-event observables.
In this paper we investigate the thermodynamics of two popular models of chiral dynamics, the linear sigma model coupled to quarks , and the Nambu–Jona-Lasinio (NJL) model . Both models are tuned to reproduce correctly properties of the physical vacuum. Our goal is to study the chiral transition and to verify the existence of the critical point at nonzero chemical potential and temperature. We also study the behavior of isentropes in the vicinity of the phase transition line in the ($`T,\mu _B`$) plane. These results can then be used in dynamical simulations to confront the predictions of with experimental data.
The structure of the paper is as follows. In Section II we study the thermodynamics of the linear sigma model coupled to quarks. This part of the paper is an extension of our previous study in ref. . In Section III we do the same for the NJL model. Section IV presents numerical results. We conclude in Section V with a summary of our results. Our units are $`\mathrm{}=c=k_B=1`$, the metric tensor is $`g^{\mu \nu }=\mathrm{diag}(+,,,)`$.
## II Thermodynamics of the Linear Sigma Model
The Lagrangian of the linear sigma model with quark degrees of freedom reads
$$=\overline{q}\left[i\gamma ^\mu _\mu g(\sigma +i\gamma _5\stackrel{}{\tau }\stackrel{}{\pi })\right]q+\frac{1}{2}\left(_\mu \sigma ^\mu \sigma +_\mu \stackrel{}{\pi }^\mu \stackrel{}{\pi }\right)U(\sigma ,\stackrel{}{\pi }),$$
(1)
where the potential is
$$U(\sigma ,\stackrel{}{\pi })=\frac{\lambda ^2}{4}\left(\sigma ^2+\stackrel{}{\pi }^2v^2\right)^2H\sigma .$$
(2)
Here $`q`$ is the light quark field $`q=(u,d)`$. The scalar field $`\sigma `$ and the pion field $`\stackrel{}{\pi }=(\pi _1,\pi _2,\pi _3)`$ together form a chiral field $`\mathrm{\Phi }=(\sigma ,\stackrel{}{\pi })`$. This Lagrangian is invariant under chiral $`SU(2)_L\times SU(2)_R`$ transformations if the explicit symmetry breaking term $`H\sigma `$ is zero. The parameters of the Lagrangian are usually chosen such that the chiral symmetry is spontaneously broken in the vacuum and the expectation values of the meson fields are $`\sigma =f_\pi `$ and $`\stackrel{}{\pi }=0`$, where $`f_\pi =93`$ MeV is the pion decay constant. The constant $`H`$ is fixed by the PCAC relation which gives $`H=f_\pi m_\pi ^2`$, where $`m_\pi =138`$ MeV is the pion mass. Then one finds $`v^2=f_\pi ^2m_\pi ^2/\lambda ^2`$. The coupling constant $`\lambda ^2`$ is determined by the sigma mass, $`m_\sigma ^2=2\lambda ^2f_\pi ^2+m_\pi ^2`$, which we set to 600 MeV, yielding $`\lambda ^220`$. The coupling constant $`g`$ is usually fixed by the requirement that the constituent quark mass in vacuum, $`M_{\mathrm{vac}}=gf_\pi `$, is about $`1/3`$ of the nucleon mass, which gives $`g3.3`$.
Let us consider a system of quarks and antiquarks in thermodynamical equilibrium at temperature $`T`$ and quark chemical potential $`\mu \mu _B/3`$. The grand partition function reads:
$$𝒵=\mathrm{Tr}\mathrm{exp}\left[\left(\widehat{}\mu \widehat{𝒩}\right)/T\right]=𝒟\overline{q}𝒟q𝒟\sigma 𝒟\stackrel{}{\pi }\mathrm{exp}\left[_x\left(+\mu \overline{q}\gamma ^0q\right)\right].$$
(3)
Here $`_xi_0^{1/T}𝑑t_Vd^3𝐱`$, where $`V`$ is the volume of the system. We adopt the mean-field approximation, replacing $`\sigma `$ and $`\stackrel{}{\pi }`$ in the exponent by their expectation values. Then, up to an overall normalization factor:
$`𝒵`$ $`=`$ $`\mathrm{exp}\left({\displaystyle \frac{VU}{T}}\right){\displaystyle 𝒟\overline{q}𝒟q\mathrm{exp}\left\{_x\overline{q}\left[i\gamma ^\mu _\mu g(\sigma +i\gamma _5\stackrel{}{\tau }\stackrel{}{\pi })\right]q+\mu \overline{q}\gamma ^0q\right\}}`$ (4)
$`=`$ $`\mathrm{exp}\left({\displaystyle \frac{VU}{T}}\right)\mathrm{det}_p\left\{\left[p_\mu \gamma ^\mu +\mu \gamma ^0g(\sigma +i\gamma _5\stackrel{}{\tau }\stackrel{}{\pi })\right]/T\right\}.`$ (5)
All thermodynamical quantities can be obtained from the grand canonical potential
$$\mathrm{\Omega }(T,\mu )=\frac{T\mathrm{ln}𝒵}{V}=U(\sigma ,\stackrel{}{\pi })+\mathrm{\Omega }_{q\overline{q}},$$
(6)
where the quark and antiquark contribution reads:
$$\mathrm{\Omega }_{q\overline{q}}(T,\mu )=\nu _q\frac{d^3𝐩}{(2\pi )^3}\left\{E+T\mathrm{ln}\left[1+\mathrm{exp}\left(\frac{\mu E}{T}\right)\right]+T\mathrm{ln}\left[1+\mathrm{exp}\left(\frac{\mu E}{T}\right)\right]\right\}.$$
(7)
Here, $`\nu _q=2N_cN_f=12`$ is the number of internal degrees of freedom of the quarks, $`N_c=3`$, and $`E=\sqrt{p^2+M^2}`$ is the valence quark and antiquark energy. The constituent quark (antiquark) mass, $`M`$, is defined to be:
$$M^2=g^2(\sigma ^2+\stackrel{}{\pi }^2).$$
(8)
The divergent first term in eq. (7) comes from the negative energy states of the Dirac sea. As follows from the standard renormalization procedure it can be partly absorbed in the coupling constant $`\lambda ^2`$ and the constant $`v^2`$. However, a logarithmic correction from the renormalization scale remains and is neglected in the following. Similar logarithmic terms are explicitly included in calculations within the NJL model (see below). Therefore one can use the comparison of these two models to conclude about the importance of these corrections.
After integrating eq. (7) by parts the contribution of valence quarks and antiquarks can be rewritten as
$$P_{q\overline{q}}(T,\mu )=\frac{\nu _q}{6\pi ^2}_0^{\mathrm{}}𝑑p\frac{p^4}{E}\left[n_q(T,\mu )+n_{\overline{q}}(T,\mu )\right],$$
(9)
where $`n_q`$ and $`n_{\overline{q}}`$ are the quark and antiquark occupation numbers,
$$n_q(T,\mu )=\frac{1}{1+\mathrm{exp}[(E\mu )/T]},n_{\overline{q}}(T,\mu )=n_q(T,\mu ).$$
(10)
The baryon-chemical potential is determined by the net baryon density
$$n_B=\frac{1}{3}\frac{\mathrm{\Omega }}{\mu }=\frac{\nu _q}{6\pi ^2}p^2𝑑p[n_q(T,\mu )n_{\overline{q}}(T,\mu )].$$
(11)
The net quark density is obviously $`n=3n_B`$. The values for the $`\sigma `$ and $`\stackrel{}{\pi }`$ fields and thereby the quark masses in eq. (8) are obtained by minimizing $`\mathrm{\Omega }`$ with respect to $`\sigma `$ and $`\stackrel{}{\pi }`$,
$$\frac{\mathrm{\Omega }}{\sigma }=\lambda ^2(\sigma ^2+\stackrel{}{\pi }^2v^2)\sigma H+g\rho _s=0,$$
(12)
$$\frac{\mathrm{\Omega }}{\pi _i}=\lambda ^2(\sigma ^2+\stackrel{}{\pi }^2v^2)\pi _i+g\rho _{ps,i}=0.$$
(13)
The scalar and pseudoscalar densities can be expressed as:
$$\rho _s=\overline{q}q=g\sigma \nu _q\frac{d^3𝐩}{(2\pi )^3}\frac{1}{E}[n_q(T,\mu )+n_{\overline{q}}(T,\mu )],$$
(14)
$$\stackrel{}{\rho }_{ps}=\overline{q}i\gamma _5\stackrel{}{\tau }q=g\stackrel{}{\pi }\nu _q\frac{d^3𝐩}{(2\pi )^3}\frac{1}{E}[n_q(T,\mu )+n_{\overline{q}}(T,\mu )].$$
(15)
The minima of $`\mathrm{\Omega }`$ defined by eqs. (12), (13) correspond to the stable or metastable states of matter in thermodynamical equilibrium where the pressure is $`P=\mathrm{\Omega }_{min}`$. The $`\sigma `$ and pion masses are determined by the curvature of $`\mathrm{\Omega }`$ at the global minimum:
$$M_\sigma ^2=\frac{^2\mathrm{\Omega }}{\sigma ^2},M_{\pi _i}^2=\frac{^2\mathrm{\Omega }}{\pi _i^2}.$$
(16)
Explicitly they are given by the expressions
$`M_\sigma ^2`$ $`=`$ $`m_\pi ^2+\lambda ^2\left(3{\displaystyle \frac{M^2}{g^2}}f_\pi ^2\right)`$ (17)
$`+`$ $`g^2{\displaystyle \frac{\nu _q}{2\pi ^2}}{\displaystyle }dpp^2[{\displaystyle \frac{p^2}{E^3}}({\displaystyle \frac{1}{1+\mathrm{exp}[(E+\mu )/T]}}+{\displaystyle \frac{1}{1+\mathrm{exp}[(E\mu )/T]}})`$ (18)
$``$ $`{\displaystyle \frac{M^2}{TE^2}}({\displaystyle \frac{1}{2(1+\mathrm{cosh}[(E+\mu )/T])}}+{\displaystyle \frac{1}{2(1+\mathrm{cosh}[(E\mu )/T])}})],`$ (19)
$$M_\pi ^2=m_\pi ^2+\lambda ^2(\frac{M^2}{g^2}f_\pi ^2)+g^2\frac{\nu _q}{2\pi ^2}𝑑pp^2\frac{1}{E}\left[\frac{1}{1+\mathrm{exp}[(E+\mu )/T]}+\frac{1}{1+\mathrm{exp}[(E\mu )/T]}\right].$$
(20)
Here we have set the expectation value of the pion field to zero, $`\stackrel{}{\pi }=0`$, thus $`M^2=g^2\sigma ^2`$. This version of the sigma model was used earlier in ref. for thermodynamical calculations at nonzero $`T`$ and $`\mu =0`$, and at nonzero $`\mu `$ and $`T=0`$. Some useful formulae for the case of small quark mass are given in the Appendix.
## III Thermodynamics of the NJL model
The NJL model has been widely used earlier for describing hadron properties and the chiral phase transition . The simplest version of the model including only scalar and pseudoscalar 4-fermion interaction terms is given by the Lagrangian <sup>*</sup><sup>*</sup>*As demonstrated in ref. , the inclusion of the vector-axialvector terms may change significantly the parameters of the chiral phase transition, in particular, the position of the critical point. But this does not change the qualitative conclusions of the present paper.:
$$=\overline{q}\left(i\gamma ^\mu _\mu m_0\right)q+\frac{G}{2}\left[(\overline{q}q)^2+(\overline{q}i\gamma _5\stackrel{}{\tau }q)^2\right],$$
(21)
where $`m_0`$ is the small current quark mass. At vanishing $`m_0`$ this NJL Lagrangian is invariant under chiral $`SU(2)_L\times SU(2)_R`$ transformations. The coupling constant $`G`$ has dimension $`(`$energy$`)^2`$, which makes the theory non-renormalizable. Therefore, a 3-momentum cutoff $`\mathrm{\Lambda }`$ is introduced to regularize divergent integrals. It defines an upper energy limit for this effective theory. Free parameters of the model are fixed to reproduce correctly the vacuum values of the pion decay constant (93 MeV), pion mass (138 MeV), and the constituent quark mass (337 MeV). Below we use the following parameters : $`G=5.496`$ GeV<sup>-2</sup>, $`m_0=5.5`$ MeV, and $`\mathrm{\Lambda }=631`$ MeV. With these parameters the chiral transition occurs at the temperature $`T190`$ MeV (for $`\mu =0`$) which is significantly higher than in the sigma model.
The partition function for the NJL model reads:
$$𝒵=\mathrm{Tr}\mathrm{exp}\left[\left(\widehat{}\mu \widehat{𝒩}\right)\right]=𝒟\overline{q}𝒟q\mathrm{exp}\left[_x\left(+\mu \overline{q}\gamma ^0q\right)\right].$$
(22)
In the mean-field approximation the Lagrangian (21) is represented in a linearized form :
$$=\overline{q}(i\gamma ^\mu _\mu m_0)q+G\overline{q}q(\overline{q}q)\frac{G}{2}\overline{q}q^2,$$
(23)
such that the partition function becomes
$$𝒵=\mathrm{exp}\left[\frac{V}{T}\frac{G\overline{q}q^2}{2}\right]\mathrm{det}_p\left[(p_\mu \gamma ^\mu +\mu \gamma ^0M)/T\right],$$
(24)
where the constituent quark mass is determined from the gap equation
$$M=m_0G\overline{q}q.$$
(25)
The right-hand side of this equation involves the scalar density
$$\rho _s=\overline{q}q=M\nu _q_{p<\mathrm{\Lambda }}\frac{d^3𝐩}{(2\pi )^3}\frac{1}{E}[n_q(T,\mu )+n_{\overline{q}}(T,\mu )1],$$
(26)
where $`n_q`$ and $`n_{\overline{q}}`$ are the valence quark and antiquark occupation numbers defined in eq. (10). Here the last term in brackets gives the contribution from the Dirac sea (which corresponds to the vacuum part of the sigma model) and cannot be neglected. The rest comes from valence quarks and antiquarks similar to the sigma model (compare with eq. (14)).
From eq. (23), the grand canonical potential for the NJL model can be written as:
$$\mathrm{\Omega }=\frac{(Mm_0)^2}{2G}\nu _q_{p<\mathrm{\Lambda }}\frac{d^3𝐩}{(2\pi )^3}\left\{E+T\mathrm{ln}\left[1+\mathrm{exp}\left(\frac{E+\mu }{T}\right)\right]+T\mathrm{ln}\left[1+\mathrm{exp}\left(\frac{E\mu }{T}\right)\right]\right\}.$$
(27)
The minimization of $`\mathrm{\Omega }`$ with respect to $`M`$ gives the gap equation (25). The expression (27) is formally identical with eq. (7) derived for the sigma model, but now the first term in curly brackets, coming from the Dirac sea, is treated explicitly after introducing the cut-off momentum $`\mathrm{\Lambda }`$. One can calculate this vacuum contribution explicitly,
$$\mathrm{\Omega }_{vac}=\nu _q_{p<\mathrm{\Lambda }}\frac{d^3𝐩}{(2\pi )^3}\sqrt{p^2+M^2}=\frac{\nu _q\mathrm{\Lambda }^4}{8\pi ^2}\left[\sqrt{1+z^2}\left(1+\frac{z^2}{2}\right)\frac{z^4}{2}\mathrm{ln}\frac{\sqrt{1+z^2}+1}{z}\right],$$
(28)
where $`z=M/\mathrm{\Lambda }`$. Expanding this expression in powers of $`z`$ one can find contributions to $`\mathrm{\Omega }`$ of order $`M^2`$, $`M^4`$,.. . As mentioned above, in the sigma model the vacuum terms are partly absorbed in the coefficients of the effective potential $`U(\sigma ,\stackrel{}{\pi })`$. However, the logarithmic term $`M^4\mathrm{ln}\frac{\mathrm{\Lambda }}{M}`$ cannot be removed in this way. Therefore, the NJL model has additional nonlinear terms in the vacuum energy which are responsible for the differences in the thermodynamic properties of the two models.
The sigma and pion masses are not as straightforward to obtain as in the linear sigma model because in the NJL model they are not represented as dynamical fields. In this model mesons are described as collective $`q\overline{q}`$ excitations. Their masses can be obtained from the poles of the quark-antiquark scattering amplitude which can be computed, for instance, in Random Phase Approximation (RPA) . In this way, one can derive the following equations for the sigma and pion masses,
$$0=\frac{m_0}{M}+(M_\sigma ^24M^2)GI(M,M_\sigma ),$$
(29)
$$0=\frac{m_0}{M}+M_\pi ^2GI(M,M_\pi ).$$
(30)
The function $`I(x,y)`$ is the quark-antiquark propagator defined as:
$$I(x,y)=\frac{\nu _q}{2\pi ^2}𝒫_{p<\mathrm{\Lambda }}𝑑pp^2\frac{1}{E}[1n_qn_{\overline{q}}]\frac{1}{E^2\frac{1}{4}y^2},$$
(31)
where $`E=\sqrt{x^2+p^2}`$, and the occupation numbers $`n_q,n_{\overline{q}}`$ are as defined in (10). In this integral $`𝒫`$ means principal value.
## IV Numerical Results
In the case of the linear sigma model everything is determined when the gap equation (12) is solved in the $`(T,\mu )`$ plane, whereas in the NJL model the meson masses have to be solved for as well. Below we present results of our numerical calculations (see also Appendix).
### A Phase diagrams
We start this section with presenting in Fig. 1 the resulting phase diagrams in the $`(T,\mu )`$ plane calculated for the two models. The middle line corresponds to the states where the two phases co-exist in the first order phase transition. Along this line the thermodynamical potential $`\mathrm{\Omega }`$ has two minima of equal depth separated by a potential barrier which height grows towards lower temperatures. At the critical point C the barrier disappears and the transition is of second order. The other lines in Fig. 1 are spinodal lines which constrain the regions of spinodal instability where $`\left(\frac{n_B}{\mu }\right)_T<0`$. Information about the timescales of this instability can be obtained from dynamical simulations .
It is instructive to plot the thermodynamic potential as a function of the order parameter for various values of $`T`$ at $`\mu =0`$, and for various values of $`\mu `$ at $`T=0`$. The first case is shown in Fig. 2, where the left panel is for the sigma model and the right panel for the NJL model. One clearly sees the smooth crossover of the symmetry breaking pattern in both cases. Note that the effective bag constant (the energy difference between the global minimum and the local maximum of the potential in vacuum) is about $`100`$ MeV/fm<sup>3</sup> in the NJL model, whereas in the sigma model it is significantly smaller, $`60`$ MeV/fm<sup>3</sup>. To a large extent this difference is responsible for the difference in the temperatures corresponding to the crossover transition: about 140 MeV in the sigma model and about 180-190 MeV in the NJL model.
In Fig. 3 the same plot is shown for $`T=0`$ and a nonzero $`\mu `$. Here, one clearly observes the pattern characteristic for a first order phase transition: two minima corresponding to phases of restored and broken symmetry separated by a potential barrier. The barrier height is larger in the sigma model than in the NJL model, thus, indicating a weaker first order phase transition in the NJL model. It now follows that somewhere in between these two extremes, for some $`\mu _c`$ and $`T_c`$, there exists a second order phase transition (the critical point). Indeed, this point is found and shown in Fig. 1. The corresponding values are $`(T_c,\mu _c)(99,207)`$ MeV in the sigma model, and $`(T_c,\mu _c)(46,332)`$ MeV in the NJL model. The behavior of the thermodynamic potential at $`\mu =\mu _c`$ and various $`T`$ is shown in Fig. 4. One can see that the potential has only one minimum which is flattest at the critical point.
### B Effective Masses
Now let us consider the model predictions for the effective masses. The constituent quark mass is shown in Fig. 5 as function of $`T`$ and $`\mu `$. These plots, of course, show the same phase structure as discussed above. At $`\mu =0`$ in both models the quark mass falls smoothly from the respective vacuum value and approaches zero as $`T`$ goes to infinity. One could define a crossover temperature as corresponding to a steepest descent region in the variation of $`M`$. This again gives a temperature of about $`140150`$ MeV for the sigma model and about $`180190`$ MeV for the NJL model. At $`T=0`$ and nonzero $`\mu `$ the constituent quark mass shows a discontinuous behavior reflecting a first order chiral transition.
The sigma and pion masses for various $`T`$ and $`\mu `$ are shown in Figs. 6,7. In both models the sigma mass first decreases smoothly, then rebounds and grows again at high $`T`$. The pion mass does not change much at temperatures below $`T_c`$ but then increases rapidly, approaching the sigma mass and signaling the restoration of chiral symmetry. As $`T`$ goes to infinity the masses grow linearly with $`T`$. The $`\mu =\mu _c`$ case is especially interesting in the sigma model. Since the sigma field is the order parameter of the chiral phase transition, its mass must vanish at the critical point for a second order phase transition. This means that $`\mathrm{\Omega }`$ has zero curvature at this point. It is, however, not clear what the sigma mass should be at the critical point in the NJL model where the quark condensate $`\overline{q}q`$ is the order parameter. Fig. 6 indeed shows that exactly at the critical point the sigma mass is zero in the sigma model. This is not the case in the NJL model, at least within the RPA used here.
In Fig. 7 the masses are plotted as function of $`\mu `$ for $`T=0`$ and $`T=T_c`$. For $`T=0`$ one clearly sees discontinuities in the behavior of the masses characteristic for the first order phase transition.
An interesting point is that, in the linear sigma model, there is no stable phase with heavy quarks for $`T=0`$, i.e., the quark mass assumes its vacuum value all the way up to the chiral transition, and then drops to a small value in the phase where chiral symmetry is restored (see Fig. 8). This behavior is related to the appearance of a bound state at zero pressure. Within the linear sigma model this “abnormal” bound state was found by Lee and Wick a long time ago . Recently, it was shown in ref. that a similar bound state appears also in the NJL model. This behavior, however, depends on the value of the coupling constant $`g`$ or $`G`$. For our choice of $`g`$ and $`G`$, this state exists in the linear sigma model, but not in the NJL model, where there is a stable phase of heavy quarks at $`T=0`$, cf. Figs. 5 and 7. In general, if the coupling constant is sufficiently large, then the attractive force between the constituent quarks becomes large enough to counterbalance the Fermi pressure, thus giving rise to a bound state. To demonstrate this we have varied the coupling constant $`g`$ for the sigma model within reasonable limits.
The results for the quark mass are shown in Fig. 8. It is seen that, indeed, one can change the smooth crossover for $`\mu =0`$ into a first order transition by increasing the coupling constant (Fig. 8 left panel) and change the first order transition in the case of $`T=0`$ into a smooth crossover. In this way a heavy quark phase comes into existence as the coupling constant is decreased and the bound state disappears (Fig.8 right panel). An analogous investigation for the NJL model leads to similar results.
### C Adiabats
Regarding hydrodynamical simulations the entropy per baryon is an interesting quantity. One can easily calculate it using standard thermodynamic relations,
$$\frac{S}{A}=3\frac{e+p\mu n}{Tn},$$
(32)
where $`e,p,n`$ are respectively the energy density, pressure and net density of the quarks and antiquarks ($`n=3n_B`$). By studying this quantity, one can check if there is a tendency towards convergence of the adiabats towards the critical point as was claimed in ref. . If this was the case it would be easy to actually hit or go close to this point in a hydrodynamical evolution. Fig. 9 shows the contours of $`S/A`$ in the $`(T,\mu )`$ plane calculated in the sigma model (left) and in the NJL model (right). We actually observe a trend which is quite opposite to this expectation. It turns out that the adiabats turn away from the critical point when they hit the first order transition line and bend towards the critical point only when they come from the smooth crossover region. This is explained as follows. First, note that all adiabats terminate at zero temperature and $`\mu =M_{vac}`$, i.e. the $`(T,\mu )`$ combination corresponding to the vacuum. The reason is that as $`T0`$, also $`S0`$ (by the third law of thermodynamics), therefore, for fixed $`S/A`$ we have to require that $`n0`$, which is fulfilled when $`\mu =M_{vac}`$. For our choice of parameters, in the sigma model the point $`(T,\mu )=(0,M_{vac})`$ is also the endpoint of the phase transition curve at $`T=0`$, since the phase transition connects the vacuum directly with the phase of restored chiral symmetry, cf. Figs. 5 and 8. For the NJL model, the endpoint of the phase transition curve is not identical with $`(0,M_{vac})`$ but is rather close to it. Therefore, also the adiabats which hit the phase transition curve have to bend away from the critical point and approach the endpoint of the phase transition line at $`T=0`$, i.e., $`T`$ decreases and $`\mu `$ increases.
This behavior is quite opposite to the case underlying the claim in ref. , where the hadronization of a large number of quark and gluon degrees of freedom into relatively few pion degrees of freedom leads to the release of latent heat and consequently to a reheating (increase of $`T`$) through the phase transition. Remember, however, that in our case there is actually no change in the number of degrees of freedom in the two phases, only in their respective masses. Consequently, there is no “focusing” effect in the linear sigma and NJL models.
## V Conclusions
We investigated the thermodynamics of the chiral phase transition within the linear sigma model coupled to quarks and the Nambu-Jona-Lasinio model. These models have similar vacuum properties but treat the contribution of the Dirac Sea differently. In the sigma model this contribution is “renormalized out” while in the NJL model it is included explicitly up to a momentum cutoff $`\mathrm{\Lambda }`$. By comparing thermodynamic properties of these two models one can check the importance of these vacuum terms. In both models, we found for small bare quark masses a smooth crossover for nonzero temperature and zero chemical potential and a first order transition for zero temperature and nonzero chemical potential. The first order phase transition line in the ($`T,\mu `$) plane ended in the expected critical point. It has been found that the $`\sigma `$ mass is zero at the critical point in the sigma model whereas in the NJL model it always remains nonzero. The phase transition in the sigma model turned out to be of the liquid-gas type. This, however, depends on the coupling constant $`g`$ between the quarks and the chiral fields. From the comparison we conclude that the phase transition pattern is generally weaker in the NJL model than the sigma model. Certainly, it will be interesting to use both models in hydrodynamical simulations in order to confirm or disconfirm possible observable signatures of the phase transition discussed in the introduction. In particular, the sigma model which contains dynamical $`\sigma `$ and pion fields, would be suitable to study the long wavelength enhancement of the $`\sigma `$ field at the critical point. Such simulations are in progress.
## Acknowledgements
The authors thank J. Borg, L.P. Csernai, P. Ellis, A.D. Jackson, L.M. Satarov and A. Wynveen for useful discussions. O.S. thanks the Yale Relativistic Heavy Ion Group for kind hospitality and support from grant no. DE-FG02-91ER-40609. The work of Á.M. is supported by the US. Department of Energy under grant DE-FG02-87ER40328. I.N.M. thanks the Humboldt Foundation for financial support. D.H.R thanks RIKEN, BNL, and the U.S. Dept. of Energy for providing the facilities essential for the completion of this work.
## A
In the chirally symmetric phase the constituent quark (antiquark) mass $`M`$ is small, it goes to zero in the chiral limit. Therefore, it is instructive to evaluate the thermodynamic potential $`\mathrm{\Omega }(T,\mu ,M)`$ for small $`M`$. In this limit $`\mathrm{\Omega }_{q\overline{q}}`$ can be represented as a power series in $`M`$. Below we give explicit expressions for the sigma model. Taking into account that $`\mathrm{\Omega }_{q\overline{q}}`$ is an even function of $`M`$, one can write
$$\mathrm{\Omega }_{q\overline{q}}(T,\mu ;M)=\mathrm{\Omega }_0(T,\mu )+\frac{M^2}{2}\left(\frac{^2\mathrm{\Omega }_{q\overline{q}}}{M^2}\right)_{M=0}+\mathrm{}$$
(A1)
Here the first term, $`\mathrm{\Omega }_0(T,\mu )\mathrm{\Omega }_{q\overline{q}}(T,\mu ,0)`$, can be easily calculated for arbitrary $`T`$ and $`\mu `$. The well-known result is
$$\mathrm{\Omega }_0(T,\mu )=\frac{\nu _q}{2\pi ^2}\left[\frac{7\pi ^4}{180}T^4+\frac{\pi ^2}{6}T^2\mu ^2+\frac{1}{12}\mu ^4\right].$$
(A2)
The quark number and entropy densities for massless fermions are obtained by differentiating $`\mathrm{\Omega }_0(T,\mu )`$ with respect to $`\mu `$ and $`T`$, respectively,
$$n=\frac{\nu _q}{6\pi ^2}(\pi ^2T^2\mu +\mu ^3),$$
(A3)
$$s=\frac{\nu _q}{6\pi ^2}\left(\frac{7\pi ^4}{15}T^3+\pi ^2T\mu ^2\right).$$
(A4)
The second term in eq. (A1) differs only by a factor of $`M`$ from the scalar density defined in eq. (14). A straightforward calculation gives
$$\left(\frac{^2\mathrm{\Omega }}{M^2}\right)_{M=0}=\left(\frac{\rho _s}{M}\right)_{M0}=\nu _q\left(\frac{T^2}{12}+\frac{\mu ^2}{4\pi ^2}\right).$$
(A5)
This can be used to estimate the pion and sigma masses at large $`T`$ and/or $`\mu `$. Expressing $`M^2`$ in terms of mean $`\pi `$ and $`\sigma `$ fields, eq. (16), and using the definition of effective masses from eq. (8), one arrives at the following asymptotic ($`M0`$) expression for the pion and sigma masses
$$M_\pi ^2=M_\sigma ^2=g^2\nu _q\left(\frac{T^2}{12}+\frac{\mu ^2}{4\pi ^2}\right).$$
(A6)
It shows that deep in the chiral symmetric phase the pion and sigma masses are degenerate and large. At high temperatures ($`T\mu `$), $`M_\pi =M_\sigma =gT`$, where $`g3`$ in our calculations. Therefore, the contribution of pion and sigma excitations to the thermodynamical potential is negligible.
In case of the NJL model the above expressions are slightly modified due to the finite cut-off $`\mathrm{\Lambda }`$ in the momentum integration.
|
warning/0007/math0007191.html
|
ar5iv
|
text
|
# Decomposition of Marsden-Weinstein reductions for representations of quivers
## 1. Introduction
Let $`K`$ be an algebraically closed field of characteristic zero, and let $`Q`$ be a quiver with vertex set $`I`$. If $`\alpha ^I`$, the space of representations of $`Q`$ of dimension vector $`\alpha `$ is
$$\mathrm{Rep}(Q,\alpha )=\underset{aQ}{}\mathrm{Mat}(\alpha _{h(a)}\times \alpha _{t(a)},K)$$
where $`h(a)`$ and $`t(a)`$ denote the head and tail vertices of an arrow $`a`$. The group
$$\mathrm{G}(\alpha )=(\underset{iI}{}\mathrm{GL}(\alpha _i,K))/K^{}$$
acts by conjugation on $`\mathrm{Rep}(Q,\alpha )`$ and on its cotangent bundle, which may be identified with $`\mathrm{Rep}(\overline{Q},\alpha )`$, where $`\overline{Q}`$ is the double of $`Q`$, obtained from $`Q`$ by adjoining a reverse arrow $`a^{}:ji`$ for each arrow $`a:ij`$ in $`Q`$. There is a corresponding moment map
$$\mu _\alpha :\mathrm{Rep}(\overline{Q},\alpha )\mathrm{End}(\alpha )_0,\mu _\alpha (x)_i=\underset{\begin{array}{c}aQ\\ h(a)=i\end{array}}{}x_ax_a^{}\underset{\begin{array}{c}aQ\\ t(a)=i\end{array}}{}x_a^{}x_a$$
where
$$\mathrm{End}(\alpha )_0=\{\theta \underset{iI}{}\mathrm{Mat}(\alpha _i,K)\underset{iI}{}\mathrm{tr}(\theta _i)=0\}(\mathrm{Lie}\mathrm{G}(\alpha ))^{},$$
and the *Marsden-Weinstein reductions* (or *symplectic reductions*) are the affine quotient varieties
$$N(\lambda ,\alpha )=\mu _\alpha ^1(\lambda )//\mathrm{G}(\alpha ),$$
where $`\lambda `$ is an element of $`K^I`$ with $`\lambda \alpha =_{iI}\lambda _i\alpha _i`$ equal to zero, and it is identified with the element of $`\mathrm{End}(\alpha )_0`$ whose $`i`$th component is $`\lambda _i\mathrm{I}`$. (Although it is possible to equip $`N(\lambda ,\alpha )`$ with the structure of a scheme, possibly not reduced, we do not do so in this paper.)
We studied this situation in a previous paper , to which we refer for further information. We showed there that $`\mu _\alpha ^1(\lambda )`$ and $`N(\lambda ,\alpha )`$ are nonempty if and only if $`\alpha R_\lambda ^+`$, the set of sums (including 0) of elements of the set $`R_\lambda ^+`$ of positive roots $`\alpha `$ with $`\lambda \alpha =0`$ (using the root system in $`^I`$ associated to $`Q`$, see ).
The elements of $`\mu _\alpha ^1(\lambda )`$ correspond to modules for a certain algebra $`\mathrm{\Pi }^\lambda `$, the deformed preprojective algebra of , and the points of $`N(\lambda ,\alpha )`$ correspond to isomorphism classes of semisimple $`\mathrm{\Pi }^\lambda `$-modules of dimension $`\alpha `$. In we showed that the possible dimension vectors of simple $`\mathrm{\Pi }^\lambda `$-modules are the elements of the set
$$\mathrm{\Sigma }_\lambda =\{\alpha R_\lambda ^+p(\alpha )>_{t=1}^rp(\beta ^{(t)})\text{ whenever }r2\text{}\alpha =_{t=1}^r\beta ^{(t)}\text{ and }\beta ^{(t)}R_\lambda ^+\}$$
where $`p(\alpha )=1\alpha \alpha +_{aQ}\alpha _{t(a)}\alpha _{h(a)}`$. Moreover, we showed that if $`\alpha \mathrm{\Sigma }_\lambda `$ then $`\mu _\alpha ^1(\lambda )`$ and $`N(\lambda ,\alpha )`$ are irreducible varieties of dimension $`\alpha \alpha 1+2p(\alpha )`$ and $`2p(\alpha )`$ respectively. For general $`\alpha R_\lambda ^+`$ it seems that $`\mu _\alpha ^1(\lambda )`$ may be rather complicated, but we show here that $`N(\lambda ,\alpha )`$ is well-behaved. If $`X`$ is an affine variety we denote by $`S^mX`$ the symmetric product of $`m`$ copies of $`X`$. Our main result is as follows.
###### Theorem 1.1.
Any $`\alpha R_\lambda ^+`$ has a decomposition $`\alpha =\sigma ^{(1)}+\mathrm{}+\sigma ^{(r)}`$ as a sum of elements of $`\mathrm{\Sigma }_\lambda `$, with the property that any other decomposition of $`\alpha `$ as a sum of elements of $`\mathrm{\Sigma }_\lambda `$ is a refinement of this decomposition. Collecting terms and rewriting this decomposition as $`\alpha =_{t=1}^sm_t\sigma ^{(t)}`$ where $`\sigma ^{(1)},\mathrm{},\sigma ^{(s)}`$ are distinct and $`m_1,\mathrm{},m_s`$ are positive integers, we have
$$N(\lambda ,\alpha )\underset{t=1}{\overset{s}{}}S^{m_t}N(\lambda ,\sigma ^{(t)}).$$
The first part of the theorem means that if $`\alpha =_{j=1}^n\beta ^{(j)}`$ with $`\beta ^{(j)}\mathrm{\Sigma }_\lambda `$, then $`\sigma ^{(t)}=_{jP_t}\beta ^{(j)}`$ for some partition $`_{t=1}^rP_t`$ of $`\{1,\mathrm{},n\}`$.
Recall that the roots $`\beta `$ can be divided into three classes: the *real roots* which have $`p(\beta )=0`$, the *isotropic imaginary roots* which have $`p(\beta )=1`$, and the *non-isotropic imaginary roots* which have $`p(\beta )>1`$. We have some observations concerning these classes.
###### Proposition 1.2.
(1) If $`\beta `$ is a real root in $`\mathrm{\Sigma }_\lambda `$, then $`N(\lambda ,\beta )`$ is a point.
(2) If $`\beta `$ is an isotropic imaginary root in $`\mathrm{\Sigma }_\lambda `$, then it is indivisible (its components have no common divisor) and $`N(\lambda ,\beta )`$ is isomorphic to a deformation of a Kleinian singularity.
(3) If $`\beta `$ is a non-isotropic imaginary root in $`\mathrm{\Sigma }_\lambda `$ then any positive multiple of $`\beta `$ is also in $`\mathrm{\Sigma }_\lambda `$.
It follows from the proposition (or directly from the proof of the theorem) that $`m_t=1`$ whenever $`\sigma ^{(t)}`$ is a non-isotropic imaginary root. Thus the theorem actually gives
$$N(\lambda ,\alpha )\underset{\begin{array}{c}t=1\\ p(\sigma ^{(t)})=1\end{array}}{\overset{s}{}}S^{m_t}N(\lambda ,\sigma ^{(t)})\times \underset{\begin{array}{c}t=1\\ p(\sigma ^{(t)})>1\end{array}}{\overset{s}{}}N(\lambda ,\sigma ^{(t)}).$$
###### Example 1.3.
If $`Q`$ is an extended Dynkin quiver with vertex set $`\{0,1,\mathrm{},n\}`$ and $`\lambda =0`$, then $`\mathrm{\Sigma }_0=\{\delta ,ϵ_0,\mathrm{},ϵ_n\}`$ where $`\delta `$ is the minimal positive imaginary root and $`ϵ_i`$ are the coordinate vectors. Thus the decomposition of $`\alpha ^I`$ is
$$\alpha =\underset{m}{\underset{}{\delta +\mathrm{}+\delta }}+\underset{\alpha _0m\delta _0}{\underset{}{ϵ_0+\mathrm{}+ϵ_0}}+\mathrm{}+\underset{\alpha _nm\delta _n}{\underset{}{ϵ_n+\mathrm{}+ϵ_n}},$$
where $`m`$ is the largest integer with $`m\delta \alpha `$. Thus $`N(0,\alpha )S^mN(0,\delta )`$, and $`N(0,\delta )`$ is the Kleinian singularity of type $`Q`$. See for example \[1, Theorem 8.10\].
If $`\alpha R_\lambda ^+`$, we denote by $`|\alpha |_\lambda `$ the maximum value of $`_{i=1}^np(\beta ^{(i)})`$ over all decompositions $`\alpha =_{i=1}^n\beta ^{(i)}`$ with the $`\beta ^{(i)}`$ in $`R_\lambda ^+`$. In fact one may assume that all $`\beta ^{(i)}`$ are in $`\mathrm{\Sigma }_\lambda `$, for amongst all decompositions which realize the maximum, one that has as many terms as possible clearly has this property. Now by Theorem 1.1, any decomposition of $`\alpha `$ as a sum of elements of $`\mathrm{\Sigma }_\lambda `$ is a refinement of one special decomposition $`\alpha =_{t=1}^r\sigma ^{(t)}`$. The defining property of $`\mathrm{\Sigma }_\lambda `$ then shows that the maximum is only achieved by this special decomposition. In particular $`|\alpha |_\lambda =_{t=1}^rp(\sigma ^{(t)})`$.
Recall that $`N(\lambda ,\alpha )`$ classifies the semisimple $`\mathrm{\Pi }^\lambda `$-modules of dimension $`\alpha `$. If $`X`$ is a semisimple $`\mathrm{\Pi }^\lambda `$-module, one says that $`X`$ has representation type
$$(k_1,\beta ^{(1)};\mathrm{};k_n,\beta ^{(n)})$$
if it has composition factors of dimensions $`\beta ^{(i)}`$ occuring with multiplicity $`k_i`$. Now Theorem 1.1 and \[2, Theorems 1.3,1.4\] have the following immediate consequence.
###### Corollary 1.4.
If $`\alpha R_\lambda ^+`$, then $`N(\lambda ,\alpha )`$ is an irreducible variety of dimension $`2|\alpha |_\lambda `$. The general element of $`N(\lambda ,\alpha )`$ has representation type
$$(m_1,\sigma ^{(1)};\mathrm{};m_s,\sigma ^{(s)}).$$
I would like to thank A. Maffei for some useful discussions, and in particular for explaining Lemma 2.3 to me. I would like to thank E. Vasserot for pointing out the error in an earlier version of this paper which used schemes instead of varieties. This research was done in Spring 2000 while visiting first the program on ‘Noncommutative Algebra’ at MSRI, and then the Sonderforschungsbereich on ‘Discrete Structures in Mathematics’ at Bielefeld University. I would like to thank my hosts at both institutions for their hospitality.
## 2. Preliminary results
Let $`Q`$ be a quiver with vertex set $`I`$. We denote by $`(,)`$ the symmetric bilinear form on $`^I`$,
$$(\alpha ,\beta )=\underset{iI}{}2\alpha _i\beta _i\underset{a\overline{Q}}{}\alpha _{h(a)}\beta _{t(a)}$$
and by $`q(\alpha )=\frac{1}{2}(\alpha ,\alpha )`$ the corresponding quadratic form. Thus $`p(\alpha )=1q(\alpha )`$. We denote by $`ϵ_i^I`$ the coordinate vector for a vertex $`iI`$.
If $`i`$ is a loopfree vertex (so $`(ϵ_i,ϵ_i)=2`$) there is a reflection $`s_i:^I^I`$ defined by $`s_i(\alpha )=\alpha (\alpha ,ϵ_i)ϵ_i`$, and a dual reflection $`r_i:K^IK^I`$ with $`r_i(\lambda )_j=\lambda _j(ϵ_i,ϵ_j)\lambda _i`$. The reflection at vertex $`i`$ is said to be *admissible* for the pair $`(\lambda ,\alpha )`$ if $`\lambda _i0`$. In this case it is shown in that there are reflection functors relating $`\mathrm{\Pi }^\lambda `$-modules of dimension $`\alpha `$ with $`\mathrm{\Pi }^{r_i(\lambda )}`$-modules of dimension $`s_i(\alpha )`$. Let $``$ be the equivalence relation on $`K^I\times ^I`$ generated by $`(\lambda ,\alpha )(r_i(\lambda ),s_i(\lambda ))`$ whenever the reflection at $`i`$ is admissible for $`(\lambda ,\alpha )`$. We say that $`(\nu ,\beta )`$ is obtained from $`(\lambda ,\alpha )`$ by a sequence of admissible reflections if they are in the same equivalence class.
###### Lemma 2.1.
If $`(\nu ,\beta )`$ is obtained from $`(\lambda ,\alpha )`$ by a sequence of admissible reflections then $`N(\nu ,\beta )N(\lambda ,\alpha )`$.
###### Proof.
This follows from \[2, Lemma 2.2\]. ∎
If $`p`$ is an oriented cycle in $`\overline{Q}`$ then for any $`\alpha ^I`$ there is a trace function
$$\mathrm{tr}_p:\mathrm{Rep}(\overline{Q},\alpha )K,x\mathrm{tr}(x_{a_1}\mathrm{}x_a_{\mathrm{}})$$
where $`p=a_1\mathrm{}a_{\mathrm{}}`$. It is invariant under the action of $`\mathrm{G}(\alpha )`$.
###### Lemma 2.2.
If $`\lambda K^I`$ and $`\alpha ^I`$ then the ring of invariants $`K[\mu _\alpha ^1(\lambda )]^{\mathrm{GL}(\alpha )}`$ is generated by the restrictions of the trace functions $`\mathrm{tr}_p`$ where $`p`$ runs through the oriented cycles in $`\overline{Q}`$.
###### Proof.
By the ring of invariants $`K[\mathrm{Rep}(\overline{Q},\alpha )]^{\mathrm{G}(\alpha )}`$ is generated by the $`\mathrm{tr}_p`$. Now $`\mu _\alpha ^1(\lambda )`$ is a closed subvariety of $`\mathrm{Rep}(\overline{Q},\alpha )`$, so the restriction map on functions
$$K[\mathrm{Rep}(\overline{Q},\alpha )]K[\mu _\alpha ^1(\lambda )]$$
is surjective. Since $`\mathrm{G}(\alpha )`$ is reductive and the base field $`K`$ has characteristic zero, there is a Reynolds operator, and so it remains surjective on taking invariants. ∎
The following result was pointed out to the author by A. Maffei in the context of Nakajima’s quiver varieties. (The proof is our own.) If $`\lambda _i=0`$ we denote by $`S_i`$ the $`\mathrm{\Pi }^\lambda `$-module with dimension vector $`ϵ_i`$ in which all arrows are zero.
###### Lemma 2.3.
If $`i`$ is a vertex with $`\lambda _i=0`$ and $`(\alpha ,ϵ_i)>0`$, then any representation of $`\mathrm{\Pi }^\lambda `$ of dimension $`\alpha `$ has $`S_i`$ as a composition factor, and there is an isomorphism
$$N(\lambda ,\alpha ϵ_i)N(\lambda ,\alpha ).$$
###### Proof.
Since $`(\alpha ,ϵ_i)>0`$ the vertex $`i`$ must be loopfree. Now some composition factor must have dimension $`\beta `$ with $`(\beta ,ϵ_i)>0`$. Then $`\beta =ϵ_i`$ by \[2, Lemma 7.2\]. Since there is no loop at vertex $`i`$, the relevant composition factor is isomorphic to $`S_i`$. Now because $`\lambda _i=0`$, the choice of a decomposition
$$K^{\alpha _i}K^{\alpha _i1}K$$
induces an embedding
$$\mu _{\alpha ϵ_i}^1(\lambda )\mu _\alpha ^1(\lambda )$$
and hence a map $`\theta :N(\lambda ,\alpha ϵ_i)N(\lambda ,\alpha )`$ which by the observation above is a bijection. We want to prove that is is an isomorphism of varieties. For this it suffices to prove that it is a closed embedding. That is, that the map of commutative algebras
$$\theta ^{}:K[\mu _\alpha ^1(\lambda )]^{\mathrm{G}(\alpha )}K[\mu _{\alpha ϵ_i}^1(\lambda )]^{\mathrm{G}(\alpha ϵ_i)}$$
is surjective. Now it is easy to see that this map sends the trace function $`\mathrm{tr}_p`$ for dimension $`\alpha `$ to the trace function $`\mathrm{tr}_p`$ for dimension $`\alpha ϵ_i`$. Thus the assertion follows from Lemma 2.2. ∎
## 3. Symmetric products
Throughout this section $`Q`$ is an extended Dynkin quiver, $`\delta `$ is its minimal positive imaginary root, and $`\lambda K^I`$ satisfies $`\lambda \delta =0`$. We choose an extending vertex 0 for $`Q`$, which means that $`\delta _0=1`$.
We say that an element of the set $`R_\lambda ^+`$ is *indecomposable* if it is nonzero and it cannot be written as a sum of two nonzero elements of this set.
###### Lemma 3.1.
The elements of $`\mathrm{\Sigma }_\lambda `$ are $`\delta `$ and the indecomposable elements of $`R_\lambda ^+`$. All elements are $`\delta `$.
###### Proof.
Clearly any real root $`\alpha `$ in $`\mathrm{\Sigma }_\lambda `$ must be indecomposable since $`p(\alpha )=0`$. Conversely, by \[2, Lemma 5.5\] any indecomposable element is in $`\mathrm{\Sigma }_\lambda `$. If $`\alpha \mathrm{\Sigma }_\lambda \{\delta \}`$ is not $`\delta `$ then $`\alpha \delta `$ is a root with some positive component, hence a positive root. But $`\alpha =\delta +(\alpha \delta )`$, contradicting indecomposability. ∎
###### Lemma 3.2.
Any decomposition of $`m\delta `$ as a sum of elements of $`\mathrm{\Sigma }_\lambda `$ is a refinement of the decomposition
$$m\delta =\underset{m}{\underset{}{\delta +\mathrm{}+\delta }}.$$
###### Proof.
Say $`\alpha ^{(1)},\mathrm{},\alpha ^{(q)}`$ are elements of $`\mathrm{\Sigma }_\lambda `$ with $`_{t=1}^r\alpha ^{(t)}=m\delta `$. By induction it suffices to find a subset $`P`$ of $`\{1,\mathrm{},q\}`$ with $`_{tP}\alpha ^{(t)}=\delta `$. We prove this by another induction: if $`P`$ is a subset for which the sum is a root $`\beta <\delta `$, we show how to enlarge $`P`$ so that the sum is a root $`\delta `$. Now $`(\delta ,\beta )=0`$ and $`(\beta ,\beta )=2`$, so $`(\beta ,_{tP}\alpha ^{(t)})=2`$. Thus $`(\beta ,\alpha ^{(s)})1`$ for some $`sP`$. Clearly $`\alpha ^{(s)}\delta `$, so
$$q(\beta +\alpha ^{(s)})=q(\beta )+q(\alpha ^{(s)})+(\beta ,\alpha ^{(s)})1+11=1,$$
so $`\beta +\alpha ^{(s)}=_{tP\{s\}}\alpha ^{(t)}`$ is a root. Moreover $`\beta +\alpha ^{(s)}\delta `$, for otherwise $`\gamma =\beta +\alpha ^{(s)}\delta `$ is a root (since $`q(\gamma )1`$) with some positive component, hence a positive root. But then $`\alpha ^{(s)}=\gamma +(\delta \beta )`$, a sum of elements of $`R_\lambda ^+`$, which contradicts the fact that $`\alpha ^{(s)}\mathrm{\Sigma }_\lambda `$. ∎
###### Lemma 3.3.
$`K[\mu _\delta ^1(\lambda )]^{\mathrm{GL}(\delta )}`$ is generated by the trace functions for paths in $`\overline{Q}`$ which start and end at the extending vertex 0.
###### Proof.
Since $`\delta _0=1`$, the trace function $`\mathrm{tr}_p`$ for a path which starts and ends at $`0`$ involves the trace of a $`1\times 1`$ matrix, which is just the unique entry of the matrix. The assertion thus follows from \[1, Corollary 8.11\]. ∎
If $`X`$ is an affine variety, we write $`S^mX`$ for its $`m`$th *symmetric product*, the affine variety $`(X\times \mathrm{}\times X)/S_m`$. Writing $`T^mA`$ for the $`m`$th tensor power of an algebra $`A`$, we have
$$K[S^mX]=(T^mK[X])^{S_m}.$$
###### Theorem 3.4.
The direct sum map
$$\underset{j=1}{\overset{m}{}}\mu _\delta ^1(\lambda )\mu _{m\delta }^1(\lambda )$$
induces an isomorphism
$$f:S^mN(\lambda ,\delta )N(\lambda ,m\delta )$$
###### Proof.
By Lemma 3.2 we know that $`f`$ is surjective. Thus it suffices to prove that it is a closed embedding, that is, that the map on functions
$$f^{}:K[\mu _{k\delta }^1(\lambda )]^{\mathrm{GL}(k\delta )}\left(T^kK[\mu _\delta ^1(\lambda )]^{\mathrm{GL}(\delta )}\right)^{S_k}$$
is surjective.
By Lemma 3.3 the ring $`K[\mu _\delta ^1(\lambda )]^{\mathrm{GL}(\delta )}`$ is generated by the trace functions $`\mathrm{tr}_p`$ for $`p`$ a path in $`\overline{Q}`$ starting and ending at 0. Since the ring is finitely generated, a finite number of paths $`p_1,\mathrm{},p_N`$ suffices.
For $`1jm`$ let $`\pi _j`$ be the projection from the product of $`m`$ copies of $`N(\lambda ,\delta )`$ onto the $`j`$th factor. Thus the coordinate ring of this product is generated by elements $`\mathrm{tr}_{p_i}\pi _j`$.
There is a surjective map from the polynomial ring $`K[x_{ij}:1iN,1jm]`$ to $`T^m(K[\mu _\delta ^1(\lambda )]^{\mathrm{GL}(\delta )})`$ sending $`x_{ij}`$ to $`\mathrm{tr}_{p_i}\pi _j`$. This induces a surjective map
$$K[x_{ij}]^{S_m}\left(T^mK[\mu _\delta ^1(\lambda )]^{\mathrm{GL}(\delta )}\right)^{S_m}$$
Now by Lemma 3.5 below, $`K[x_{ij}]^{S_m}`$ is generated by the power sums
$$s_{r_1,\mathrm{},r_N}=\underset{j}{}x_{1j}^{r_1}\mathrm{}x_{Nj}^{r_N}.$$
Thus $`\left(T^mK[\mu _\delta ^1(\lambda )]^{\mathrm{GL}(\delta )}\right)^{S_m}`$ is generated by the elements
$$s_{r_1,\mathrm{},r_N}^{}=\underset{j}{}(\mathrm{tr}_{p_1}\pi _j)^{r_1}\mathrm{}(\mathrm{tr}_{p_N}\pi _j)^{r_N}=\underset{j}{}(\mathrm{tr}_{p_1}^{r_1}\mathrm{}\mathrm{tr}_{p_N}^{r_N})\pi _j.$$
Since $`\delta _0=1`$ we have $`\mathrm{tr}_p\mathrm{tr}_q=\mathrm{tr}_{pq}`$ for any paths $`p,q`$ which start and end at 0, so $`\mathrm{tr}_{p_1}^{r_1}\mathrm{}\mathrm{tr}_{p_N}^{r_N}=\mathrm{tr}_p`$ where $`p`$ is the path $`p_1^{r_1}\mathrm{}p_N^{r_N}`$. Thus
$$s_{r_1,\mathrm{},r_N}^{}=\underset{j}{}\mathrm{tr}_p\pi _j.$$
This shows that $`s_{r_1,\mathrm{},r_N}^{}`$ is the image under $`f^{}`$ of the trace function $`\mathrm{tr}_p`$ for $`\mu _{m\delta }^1(\lambda )`$. Thus the image of $`f^{}`$ contains a set of generators, so $`f^{}`$ is surjective, as required. ∎
###### Lemma 3.5.
If $`S_m`$ acts on the polynomial ring $`K[x_{ij}:1iN,1jm]`$ by permuting the $`x_{ij}`$ for each $`i`$, then the ring of invariants is generated by the power sums
$$s_{r_1,\mathrm{},r_N}=\underset{j}{}x_{1j}^{r_1}\mathrm{}x_{Nj}^{r_N}.$$
($`r_1,\mathrm{},r_N0`$).
###### Proof.
By \[6, Chapter II, Section 3\] the ring of invariants is generated by polarizations of the elementary symmetric polynomials, so by elements of the form
$$\varphi _{i_1,i_2,\mathrm{},i_k}=x_{i_1j_1}x_{i_2j_2}\mathrm{}x_{i_kj_k}$$
where the sum is over all distinct $`j_1,j_2,\mathrm{},j_k`$ in the range $`1`$ to $`m`$. Now the elementary symmetric polynomials can be expressed as polynomials in the power sums by Newton’s formulae, and on polarizing this expresses $`\varphi _{i_1,i_2,\mathrm{},i_k}`$ as a polynomial in the $`s_{r_1,\mathrm{},r_N}`$. For example polarizing the formula
$$\underset{j<k<\mathrm{}}{}z_jz_kz_{\mathrm{}}=\frac{1}{6}\left(\left(\underset{j}{}z_j\right)^33\left(\underset{j}{}z_j\right)\left(\underset{j}{}z_j^2\right)+2\underset{j}{}z_j^3\right)$$
with respect to the sets of variables $`x_{i_1,j}`$, $`x_{i_2,j}`$ and $`x_{i_3,j}`$ gives
$$\begin{array}{cc}\hfill \varphi _{i_1,i_2,i_3}=& \left(\underset{j}{}x_{i_1,j}\right)\left(\underset{j}{}x_{i_2,j}\right)\left(\underset{j}{}x_{i_3,j}\right)\left(\underset{j}{}x_{i_1,j}\right)\left(\underset{j}{}x_{i_2,j}x_{i_3,j}\right)\hfill \\ & \left(\underset{j}{}x_{i_2,j}\right)\left(\underset{j}{}x_{i_1,j}x_{i_3,j}\right)\left(\underset{j}{}x_{i_3,j}\right)\left(\underset{j}{}x_{i_1,j}x_{i_2,j}\right)\hfill \\ & +2\underset{j}{}x_{i_1,j}x_{i_2,j}x_{i_3,j},\hfill \end{array}$$
and all sums on the right hand side are of the form $`s_{r_1,\mathrm{},r_N}`$ for suitable $`r_1,\mathrm{},r_N`$. ∎
## 4. Adding a vertex to an extended Dynkin quiver
In this section let $`Q^{}`$ be an extended Dynkin quiver, let $`k`$ be an extending vertex for $`Q^{}`$, and let $`Q`$ be a quiver obtained from $`Q^{}`$ by adjoining one vertex $`j`$ and one arrow joining $`j`$ to $`k`$. Let $`I`$ be the vertex set of $`Q`$ and let $`\delta ^I`$ be the minimal positive imaginary root for $`Q^{}`$.
For any $`\alpha ^I`$ we define $`\alpha ^{}=\alpha \alpha _jϵ_j`$. Thus $`\alpha _j^{}=0`$ and $`\alpha _i^{}=\alpha _i`$ for $`ij`$. One can think of $`\alpha ^{}`$ as the restriction of $`\alpha `$ to $`Q^{}`$.
Throughout this section we assume that $`\lambda K^I`$ satisfies $`\lambda \delta =\lambda _j=0`$. We prove the following result which is used in the next section.
###### Proposition 4.1.
If $`\alpha \mathrm{\Sigma }_\lambda `$, $`\alpha _j=1`$ and $`m\delta \alpha ^{}R_\lambda ^+`$ for some $`m0`$ then $`\alpha =ϵ_j`$.
An example shows the necessity of the hypothesis that $`m\delta \alpha ^{}R_\lambda ^+`$.
###### Example 4.2.
Let $`Q`$ be the quiver
with vertex set $`\{1,2,3,4\}`$, so $`j=1`$, $`k=2`$, $`Q^{}`$ is of type $`\stackrel{~}{A}_2`$ and $`\delta =(0,1,1,1)`$. If $`\lambda =(0,1,2,1)`$ then $`\alpha =(1,3,2,1)\mathrm{\Sigma }_\lambda `$ since by admissible reflections at the indicated vertices the pair $`(\lambda ,\alpha )`$ transforms as
$$\begin{array}{c}\hfill ((0,1,2,1),(1,3,2,1))\stackrel{2}{}((1,1,1,2),(1,1,2,1))\stackrel{3}{}((1,2,1,1),(1,1,0,1))\\ \hfill \stackrel{4}{}((1,1,2,1),(1,1,0,0))\stackrel{2}{}((0,1,1,2),(1,0,0,0))\end{array}$$
and it is clear that $`(1,0,0,0)\mathrm{\Sigma }_{(0,1,1,2)}`$. However, it is easy to see that there is no $`m`$ with $`m\delta \alpha ^{}=(0,m3,m2,m1)`$ in $`R_\lambda ^+`$.
Before proving the proposition we need some lemmas. Observe that for a vertex $`i\{j,k\}`$ we have $`r_i(\lambda )\delta =0`$ and $`r_i(\lambda )_j=0`$. On the other hand $`r_k(\lambda )\delta =0`$, but we may have $`r_k(\lambda )_j0`$.
###### Lemma 4.3.
If $`\alpha \mathrm{\Sigma }_\lambda `$ and $`\alpha _j=1`$, then by a sequence of admissible reflections at vertices $`j`$ one can send $`(\lambda ,\alpha )`$ to $`(\nu ,ϵ_j)`$ for some $`\nu `$.
###### Proof.
We consider the pairs $`(\nu ,\beta )`$ which can be obtained from $`(\lambda ,\alpha )`$ by a sequence of such admissible reflections. Always $`\beta `$ is positive, since it is in $`\mathrm{\Sigma }_\nu `$ by \[2, Lemma 5.2\]. Thus we can choose a pair $`(\nu ,\beta )`$ with $`\beta `$ minimal. Clearly we have $`\beta _j=1`$. For a contradiction, suppose that $`\beta ^{}0`$.
Since $`\delta `$ is unchanged by these reflections, we have $`\nu \delta =\lambda \delta =0`$. Also, for each vertex $`ij`$ we have $`(\beta ,ϵ_i)0`$, for either there is a loop at $`i`$, in which case it is automatic, or $`\nu _i=0`$, in which case it follows from \[2, Lemma 7.2\], or there is an admissible reflection at $`i`$, and it follows from the minimality of $`\beta `$. We deduce that $`(\beta ^{},ϵ_i)0`$ for $`i\{j,k\}`$, and $`(\beta ^{},ϵ_k)1`$.
Suppose first that $`(\beta ^{},ϵ_k)=1`$. Then
$$0=(\beta ^{},\delta )=\underset{ij}{}(\beta ^{},ϵ_i)\delta _i=1+\underset{i\{j,k\}}{}(\beta ^{},ϵ_i)\delta _i,$$
and all terms in the second sum are $`0`$. Thus exactly one of the terms is $`1`$, and all others are zero. That is, there is a vertex $`sk`$ in $`Q^{}`$ with $`\delta _s=1`$ and $`(\beta ^{},ϵ_s)=1`$, and $`(\beta ^{},ϵ_i)=0`$ for all vertices $`i\{k,s\}`$ in $`Q^{}`$. This is impossible by \[2, Lemma 8.8\].
Thus $`(\beta ^{},ϵ_k)0`$. It follows that $`(\beta ^{},\beta ^{})0`$, so since $`Q^{}`$ is extended Dynkin we have $`\beta ^{}=m\delta `$ for some $`m>0`$. Now the decomposition $`\beta =ϵ_j+\delta +\mathrm{}+\delta `$ is easily seen to satisfy
$$p(\beta )=1q(ϵ_j+m\delta )=m(ϵ_j,\delta )=m=p(ϵ_j)+p(\delta )+\mathrm{}+p(\delta ).$$
We have seen that $`\delta R_\nu ^+`$. Also $`\nu _j=\nu ϵ_j=\nu \beta =\lambda \alpha =0`$ since $`\alpha \mathrm{\Sigma }_\lambda `$, so that $`ϵ_jR_\nu ^+`$. This contradicts the fact that $`\beta \mathrm{\Sigma }_\nu `$.
Thus $`\beta ^{}=0`$, as required. ∎
###### Lemma 4.4.
If $`\alpha \mathrm{\Sigma }_\lambda `$ and $`\alpha _j=1`$ then $`\gamma _k1(\alpha ^{},\gamma )\gamma _k`$ for any $`\gamma R_\lambda ^+`$ with $`\gamma <\delta `$.
###### Proof.
Some sequence of admissible reflections at vertices $`j`$ sends $`(\lambda ,\alpha )`$ to $`(\nu ,ϵ_j)`$. If $`\gamma R_\lambda ^+`$ and $`\gamma _j=0`$ then by \[2, Lemma 5.2\] the reflections send it to a positive root $`\beta `$, still with $`\beta _j=0`$. Thus $`(\alpha ,\gamma )=(ϵ_j,\beta )0`$, and so
$$(\alpha ^{},\gamma )=(\alpha ,\gamma )(ϵ_j,\gamma )0(\gamma _k)=\gamma _k,$$
which is one of the inequalities. The other one is obtained by replacing $`\gamma `$ with $`\delta \gamma R_\lambda ^+`$. ∎
Choose a total ordering $``$ on $`K`$ as in \[1, Section 7\]. Let $`Q^{\prime \prime }`$ be the Dynkin quiver obtained from $`Q^{}`$ by deleting the vertex $`k`$. Let $`I^{\prime \prime }`$ be the vertex set of $`Q^{\prime \prime }`$. Recall that a vector $`\mu K^{I^{\prime \prime }}`$ is said to be *dominant* if $`\mu _i0`$ for all $`iI^{\prime \prime }`$.
###### Lemma 4.5.
By a sequence of admissible reflections at vertices in $`I^{\prime \prime }`$ one can send $`(\lambda ,\alpha )`$ to a pair $`(\xi ,\beta )`$ where $`\xi `$ is a vector whose restriction to $`I^{\prime \prime }`$ is dominant.
###### Proof.
Apply \[1, Lemma 7.2\] to $`Q^{\prime \prime }`$, and then consider the sequence of reflections as reflections for $`Q`$. Of course non-admissible reflections can be omitted, for if $`\xi K^I`$ and $`\xi _i=0`$ then $`r_i(\xi )=\xi `$. ∎
###### Lemma 4.6.
If the restriction of $`\lambda `$ to $`I^{\prime \prime }`$ is dominant, and if $`\gamma R_\lambda ^+`$ has $`\gamma _j=0`$, then there is some $`r0`$ with $`\gamma _i=r\delta _i`$ for all vertices $`i`$ with $`\lambda _i0`$.
###### Proof.
Any indecomposable element of $`R_\lambda ^+`$ which vanishes at $`j`$ is $`\delta `$, so it suffices to prove that if $`\gamma ^I`$ is a vector with $`\gamma \delta `$ and $`\lambda \gamma =0`$, then either $`\gamma _i=0`$ for all $`i`$ with $`\lambda _i0`$, or $`\gamma _i=\delta _i`$ for all $`i`$ with $`\lambda _i0`$.
Since $`k`$ is an extending vertex for $`Q^{}`$ we have $`\delta _k=1`$, and so by replacing $`\gamma `$ by $`\delta \gamma `$ if necessary, we may assume that $`\gamma _k=0`$.
Now the equality $`\lambda \gamma =0`$ implies that $`_{iI^{\prime \prime }}\gamma _i\lambda _i=0`$. By the dominance condition it follows that $`\gamma _i=0`$ for any vertex $`iI^{\prime \prime }`$ with $`\lambda _i0`$. ∎
###### Proof of Proposition 4.1.
First suppose that $`\lambda =0`$. If $`\alpha ϵ_j`$ then the expression for $`\alpha `$ as a sum of coordinate vectors is a non-trivial decomposition into elements of $`R_\lambda ^+`$. Since $`p(\alpha )=0`$ by Lemma 4.3, this contradicts the fact that $`\alpha \mathrm{\Sigma }_\lambda `$.
Thus we may suppose that $`\lambda 0`$. Replacing $`(\lambda ,\alpha )`$ by the pair $`(\xi ,\beta )`$ of Lemma 4.5, we may assume that the restriction of $`\lambda `$ to $`I^{\prime \prime }`$ is dominant. Observe that the reflections involved, at vertices in $`I^{\prime \prime }`$, can change $`\alpha `$, but they do not affect the dimension vectors $`ϵ_j`$ and $`\delta `$. The standing hypotheses on $`\lambda `$ still hold, as do the hypotheses of the proposition by \[2, Lemma 5.2\].
Now the restriction of $`\lambda `$ to $`I^{\prime \prime }`$ is non-zero, for otherwise the condition that $`\lambda \delta =0`$ implies that $`\lambda _k=0`$, and then since $`\lambda _j=0`$ we have $`\lambda =0`$. Thus $`\lambda _k=_{iI^{\prime \prime }}\delta _i\lambda _i0`$.
By Lemma 4.6 there is some integer $`r`$ with $`(m\delta \alpha ^{})_i=r\delta _i`$ for all $`i`$ with $`\lambda _i0`$. Let $`\beta =\alpha ^{}(mr)\delta ^I`$. Of course $`\beta _j=0`$ and for any vertex $`i`$ with $`\lambda _i0`$ we have $`\beta _i=0`$.
Suppose that $`\beta `$ is nonzero. Consider the restriction of $`\beta `$ to a connected component of the quiver obtained from $`Q^{}`$ by deleting all vertices $`i`$ with $`\lambda _i0`$. It is actually a subquiver of $`Q^{\prime \prime }`$, so Dynkin. If $`\eta `$ is a positive root for this connected component, then $`\eta R_\lambda ^+`$, and
$$(\beta ,\eta )=(\alpha ^{},\eta )(mr)(\delta ,\eta )=(\alpha ^{},\eta ),$$
so Lemma 4.4 implies that $`1(\beta ,\eta )0`$. But this is impossible by Lemma 4.7 below.
Thus $`\beta =0`$, so $`\alpha =ϵ_j+(mr)\delta `$. Now since $`p(\alpha )=0`$ we have $`m=r`$. ∎
The above proof uses the following result about Dynkin quivers.
###### Lemma 4.7.
If $`Q^{}`$ is a Dynkin quiver with vertex set $`I^{}`$ then there is no nonzero vector $`\alpha ^I^{}`$ with $`1(\alpha ,\eta )0`$ for all positive roots $`\eta `$ for $`Q^{}`$.
###### Proof.
We cannot have $`(\alpha ,ϵ_i)=0`$ for all $`i`$, for otherwise $`(\alpha ,\alpha )=0`$, so $`\alpha =0`$ since $`Q^{}`$ is Dynkin.
Embed $`Q^{}`$ in an extended Dynkin quiver of the same type by adding an extending vertex $`s`$, and consider $`\alpha `$ as a dimension vector for this quiver. Let $`\delta `$ be the minimal positive imaginary root.
Since $`\delta ϵ_s`$ is a root for $`Q^{}`$ we have $`(\alpha ,\delta ϵ_s)1`$. Now it is equal to $`_{is}\delta _i(\alpha ,ϵ_i)`$, and all terms in the sum are $`0`$, but not all are zero. Thus exactly one term is nonzero, say for $`i=r`$, and it is equal to $`1`$. This implies that $`r`$ is an extending vertex, and $`(\alpha ,ϵ_r)=1`$. Thus the vector $`\alpha `$ and the extending vertices $`r`$ and $`s`$ contradict \[2, Lemma 8.8\]. ∎
## 5. Decomposing the quiver
In this section we suppose that $`Q`$ is a quiver whose vertex set $`I`$ is a disjoint union $`𝒥𝒦`$, and we write any $`\alpha ^I`$ as $`\alpha =\alpha _𝒥+\alpha _𝒦`$ where the summands have support in $`𝒥`$ and $`𝒦`$ respectively.
###### Lemma 5.1.
If the dimension vector of any composition factor of a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha `$ has support contained either in $`𝒥`$ or in $`𝒦`$ then
$$N(\lambda ,\alpha )N(\lambda ,\alpha _𝒥)\times N(\lambda ,\alpha _𝒦).$$
###### Proof.
We can identify
$$\mu _{\alpha _𝒥}^1(\lambda )\times \mu _{\alpha _𝒦}^1(\lambda )$$
with a $`\mathrm{G}(\alpha )`$-stable closed subvariety of $`\mu _\alpha ^1(\lambda )`$ (defined by the vanishing of all arrows with one end in $`𝒥`$ and the other end in $`𝒦`$). The inclusion thus induces a closed embedding
$$N(\lambda ,\alpha _𝒥)\times N(\lambda ,\alpha _𝒦)N(\lambda ,\alpha ),$$
and by the assumption on composition factors this is a bijection. ∎
We give some cases when this can be applied. First we need a lemma.
###### Lemma 5.2.
Suppose there is a unique arrow with one end in $`𝒥`$ and the other in $`𝒦`$, say connecting vertices $`j𝒥`$ and $`k𝒦`$. Let $`\stackrel{~}{Q}`$ be the quiver with vertex set $`𝒦\{j\}`$ containing this arrow, and all arrows with head and tail in $`𝒦`$. Let $`\mu `$ be the vector for $`\stackrel{~}{Q}`$ whose restriction to $`𝒦`$ is the same as $`\lambda `$, and with $`\mu _j=0`$.
Let $`\alpha ^I`$ and assume that $`\alpha _j=1`$ and $`\lambda \alpha _𝒥=\lambda \alpha _𝒦=0`$. Then $`\alpha R_\lambda ^+`$ if and only if $`\alpha _𝒥R_\lambda ^+`$ and $`ϵ_j+\alpha _𝒦R_\mu ^+`$.
###### Proof.
The statement does not depend on the orientation of the arrows in $`Q`$, so we may suppose that the arrow connecting $`𝒥`$ and $`𝒦`$ is $`b:kj`$.
By \[2, Theorem 4.4\] the condition that $`\alpha R_\lambda ^+`$ is that there is a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha `$. Similarly for the other two conditions.
Now if the module is given by an element $`x\mathrm{Rep}(\overline{Q},\alpha )`$, then for any vertex $`i`$ we have
$$\underset{h(a)=i}{}x_ax_a^{}\underset{t(b)=i}{}x_a^{}x_a=\lambda _i1.$$
Taking the trace and summing over all $`i𝒥`$, all but one term cancels, leaving $`\mathrm{tr}(x_bx_b^{})=0`$. Since this is a $`1\times 1`$ matrix we have $`x_bx_b^{}=0`$. It follows that the components of $`x`$ corresponding to arrows with head and tail in $`𝒥`$ define a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha _𝒥`$, and the remaining components of $`x`$ define a $`\mathrm{\Pi }^\mu (\stackrel{~}{Q})`$-module of dimension $`ϵ_j+\alpha _𝒦`$. Clearly two such modules can also be used to construct a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha `$. ∎
###### Lemma 5.3.
Suppose that $`\lambda \alpha _𝒥=0`$, there is a unique arrow $`b`$ with one end in $`𝒥`$ and the other in $`𝒦`$, say connecting vertices $`j𝒥`$ and $`k𝒦`$, and $`\alpha _j=\alpha _k=1`$. Then the dimension vector of any composition factor of a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha `$ has support contained in $`𝒥`$ or $`𝒦`$.
###### Proof.
Because of the existence of a module of dimension $`\alpha `$ we have $`\lambda \alpha =0`$, hence also $`\lambda \alpha _𝒦=0`$. For a contradiction, suppose there is a composition factor whose dimension $`\beta `$ does not have support in $`𝒥`$ or $`𝒦`$. Then $`\beta _j=\beta _k=1`$. Since the dimension vector $`\gamma `$ of any other composition factor must have support in $`𝒥`$ or $`𝒦`$, and has $`\lambda \gamma =0`$, we deduce that $`\lambda \beta _𝒥=\lambda \beta _𝒦=0`$.
By Lemma 5.2 we have $`\beta _𝒥R_\lambda ^+`$, and by symmetry also $`\beta _𝒦R_\lambda ^+`$. But clearly $`(\beta _𝒦,\beta _𝒥)=1`$, so that $`p(\beta )=p(\beta _𝒥)+p(\beta _𝒦)`$, contradicting the fact that $`\beta \mathrm{\Sigma }_\lambda `$. ∎
###### Lemma 5.4.
Suppose that $`\lambda \alpha _𝒥=0`$, there is a unique arrow with one end in $`𝒥`$ and the other in $`𝒦`$, say connecting vertices $`j𝒥`$ and $`k𝒦`$, $`\alpha _j=1`$, the restriction of $`Q`$ to $`𝒦`$ is extended Dynkin with extending vertex $`k`$ and minimal positive imaginary root $`\delta `$, and $`\alpha _𝒦=m\delta `$ with $`m2`$. Then the dimension vector of any composition factor of a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha `$ has support contained in $`𝒥`$ or $`𝒦`$.
###### Proof.
Because of the existence of a module of dimension $`\alpha `$, we have $`\lambda \alpha _𝒦=0`$. Since the field $`K`$ has characteristic zero, we deduce that $`\lambda \delta =0`$.
For a contradiction, suppose there is a composition factor whose dimension $`\beta `$ does not have support in $`𝒥`$ or $`𝒦`$. Then $`\beta _j=1`$. Since the dimension vector $`\gamma `$ of any other composition factor must have $`\gamma _j=0`$, it has support in $`𝒥`$ or $`𝒦`$, and since it has $`\lambda \gamma =0`$, we deduce that $`\lambda \beta _𝒥=\lambda \beta _𝒦=0`$. Also $`m\delta \beta _𝒦R_\lambda ^+`$.
Let $`\stackrel{~}{Q}`$ be the quiver obtained from $`Q`$ as in Lemma 5.2, and let $`\mu `$ be the corresponding vector. Since $`m\delta \beta _𝒦`$ has support in $`𝒦`$ it can be considered as an element of $`R_\mu ^+`$. By Lemma 5.2 we have $`\beta _𝒥R_\lambda ^+`$ and $`ϵ_j+\beta _𝒦R_\mu ^+`$. Now by assumption $`\beta _𝒦`$ is nonzero, so Proposition 4.1 implies that $`ϵ_j+\beta _𝒦\mathrm{\Sigma }_\mu `$. By \[2, Theorem 5.6\] this implies that there are nonzero $`\varphi ,\psi R_\mu ^+`$ with $`\varphi +\psi =ϵ_j+\beta _𝒦`$ and $`(\varphi ,\psi )_{\stackrel{~}{Q}}1`$. Without loss of generality, $`\varphi _j=0`$ and $`\psi _j=1`$. Considered as a dimension vector for $`Q`$ we clearly have $`\varphi R_\lambda ^+`$. Also, Lemma 5.2 applies to the dimension vector $`\psi +\beta _𝒥ϵ_j`$, and shows that it belongs to $`R_\lambda ^+`$. Since also
$$(\varphi ,\psi +\beta _𝒥ϵ_j)=(\varphi ,\psi )_{\stackrel{~}{Q}}1,$$
we have $`\beta =\varphi +(\psi +\beta _𝒥ϵ_j)\mathrm{\Sigma }_\lambda `$ by \[2, Theorem 5.6\]. A contradiction. ∎
## 6. Proof of the theorem
###### Proof of Theorem 1.1.
We prove this for all $`Q`$, $`\lambda `$ and $`\alpha R_\lambda ^+`$ by induction on the maximum possible number of terms in an expression for $`\alpha `$ as a sum of elements of $`R_\lambda ^+`$. If $`\alpha \mathrm{\Sigma }_\lambda `$ then the assertions are vacuous, so assume that $`\alpha \mathrm{\Sigma }_\lambda `$.
By \[2, Lemma 5.2\] and Lemma 2.1 we can always apply a sequence of admissible reflections to the pair $`(\lambda ,\alpha )`$. Let $`F_\lambda `$ be the set of \[2, Section 7\]. If $`\alpha F_\lambda `$ then by applying a sequence of admissible reflections to $`(\lambda ,\alpha )`$ we may assume that there is a loopfree vertex $`i`$ with $`\lambda _i=0`$ and $`(\alpha ,ϵ_i)>0`$. Clearly in any decomposition of $`\alpha `$ as a sum of elements of $`\mathrm{\Sigma }_\lambda `$ one of the terms, say $`\beta `$, has $`(\beta ,ϵ_i)>0`$. But by \[2, Lemma 7.2\] this implies that $`\beta =ϵ_i`$. Now $`\alpha ϵ_iR_\lambda ^+`$, and by the inductive hypothesis the assertions hold for $`\alpha ϵ_i`$. If the decomposition is
$$\alpha ϵ_i=\sigma ^{(1)}+\mathrm{}+\sigma ^{(r)}$$
then clearly
$$\alpha =ϵ_i+\sigma ^{(1)}+\mathrm{}+\sigma ^{(r)}$$
is a suitable decomposition of $`\alpha `$. Moreover, if we have
$$N(\lambda ,\alpha ϵ_i)\underset{t=1}{\overset{s}{}}S^{m_t}N(\lambda ,\sigma ^{(t)}),$$
then since $`N(\lambda ,ϵ_i)`$ is just a point, any term $`S^mN(\lambda ,ϵ_i)`$ if it occurs, can be removed, and replaced by $`S^{m+1}N(\lambda ,ϵ_i)`$ without changing the product. Thus by Lemma 2.3 we obtain the required expression for $`N(\lambda ,\alpha )`$.
Thus we are reduced to the case when $`\alpha F_\lambda \mathrm{\Sigma }_\lambda `$. By applying a sequence of admissible reflections to the pair $`(\lambda ,\alpha )`$, and then passing to the support quiver of $`\alpha `$, we may assume that one of the cases (I), (II) or (III) of \[2, Theorem 8.1\] holds. We deal with each of these in turn.
Case (I). Here $`Q`$ is extended Dynkin, $`\lambda \delta =0`$, and $`\alpha =m\delta `$ for some $`m2`$. By Lemma 3.2 and Theorem 3.4 the decomposition $`\alpha =\delta +\mathrm{}+\delta `$ has the required properties.
Case (II). Here $`Q`$ decomposes as in Lemma 5.3. In the notation of Section 5 we write $`\alpha =\alpha _𝒥+\alpha _𝒦`$. Since $`\alpha R_\lambda ^+`$ there is a $`\mathrm{\Pi }^\lambda `$-module of dimension $`\alpha `$. Since the dimension vector of any composition factor has support in $`𝒥`$ or $`𝒦`$ we deduce that $`\alpha _𝒥`$ and $`\alpha _𝒦`$ are in $`R_\lambda ^+`$. By the inductive hypothesis the conclusions of the theorem hold for $`\alpha _𝒥`$ and $`\alpha _𝒦`$. Adding together the decompositions of $`\alpha _𝒥`$ and $`\alpha _𝒦`$ we obtain a decomposition of $`\alpha `$. Obviously, since $`\alpha _𝒥`$ and $`\alpha _𝒦`$ have disjoint support, no summand occurs in both parts. The result thus follows from Lemmas 5.1 and 5.3.
Case (III). Here $`Q`$ decomposes as in Lemma 5.4. We write $`\alpha =\alpha _𝒥+m\delta `$. Again $`\alpha _𝒥`$ and $`m\delta `$ are in $`R_\lambda ^+`$ and by the inductive hypothesis the conclusions of the theorem hold for them. This gives a decomposition of $`\alpha `$ which has the required properties by Lemmas 5.1 and 5.4. ∎
###### Proof of Proposition 1.2.
(1) If $`\beta `$ is a real root in $`\mathrm{\Sigma }_\lambda `$, then $`N(\lambda ,\beta )`$ is a point by \[2, Corollary 1.4\].
(2) If $`\beta `$ is an isotropic imaginary root in $`\mathrm{\Sigma }_\lambda `$, then it is indivisible, for if $`\beta =r\gamma `$ then $`\gamma `$ is a root, it has $`\lambda \gamma =0`$ since the base field $`K`$ has characteristic zero, and the decomposition $`\beta =\gamma +\mathrm{}+\gamma `$ has $`p(\beta )<p(\gamma )+\mathrm{}+p(\gamma )`$, contrary to the definition of $`\mathrm{\Sigma }_\lambda `$.
By \[2, Theorem 5.8\], some sequence of admissible reflections sends the pair $`(\lambda ,\beta )`$ to a pair $`(\lambda ^{},\beta ^{})`$ with $`\beta ^{}`$ in the fundamental region. Since it is isotropic imaginary we have $`(\beta ^{},ϵ_i)=0`$ for any vertex $`i`$ in the support of $`\beta ^{}`$. By \[3, §1.2\] this implies that the support quiver $`Q^{}`$ of $`\beta ^{}`$ is extended Dynkin and $`\beta ^{}=\delta `$, its minimal positive imaginary root.
Finally $`N(\lambda ,\beta )N(\lambda ^{},\delta )`$ by Lemma 2.1, and this is a deformation of the Kleinian singularity of type $`Q^{}`$ by Kronheimer’s work . See for example \[1, Section 8\].
(3) Suppose that $`\beta `$ is a non-isotropic imaginary root in $`\mathrm{\Sigma }_\lambda `$ and $`m2`$. If $`F_\lambda `$ is the set of \[2, Section 7\], then \[2, Lemma 7.4\] implies that $`\beta F_\lambda `$, and hence also $`m\beta F_\lambda `$. Now in \[2, Theorem 8.1\], case (I) cannot occur since $`m\beta `$ is non-isotropic, and cases (II) and (III) cannot occur since all components of $`m\beta `$ are divisible by $`m`$. Thus $`m\beta \mathrm{\Sigma }_\lambda `$. ∎
|
warning/0007/hep-ph0007082.html
|
ar5iv
|
text
|
# 1 𝑋(2) (lines with crosses) and 𝑋(3) (lines with no symbol) as functions of 𝐿₆(𝑀_𝜌)⋅10³ and 𝑟=𝑚_𝑠/𝑚 (solid: 𝑟=20, dashed: 𝑟=25, dotted: 𝑟=30) for 𝐹₀=85 MeV.
1. Low-energy constants (LEC’s) characteristic of the effective chiral Lagrangian of QCD are order parameters of spontaneous breaking of chiral symmetry (SB$`\chi `$S) which encode informations about the chiral structure of QCD vacuum. This is why their precise determination is of theoretical importance. In QCD-like theories, the mechanism of SB$`\chi `$S is conveniently described in terms of average properties of lowest Euclidean fermionic modes . This framework singlets out the quark condensate $`\overline{q}q`$ (i.e. the average density of infrared fermionic modes) and the pion decay constant $`F_\pi `$ (conductivity) as two prominent order parameters .
Sofar, most determinations of low-energy constants have operated under two assumptions: i) Quark condensate dominates the description of symmetry breaking effects, and in particular the expansion of Goldstone boson masses. ii) Quantum fluctuations are less important and the dynamics of SB$`\chi `$S approximately fits into the large-$`N_c`$ picture of QCD. A particular consequence of the latter assumption is the expectation that most order parameters will only marginally depend on the number $`N_f`$ of massless quarks: the two-flavor condensate $`\mathrm{\Sigma }(2)=lim_{m_u,m_d0}\overline{u}u`$ should, for instance, be almost independent of the mass of the strange quark, even if the latter is set $`0`$ yielding the three-flavor condensate $`\mathrm{\Sigma }(3)`$. There should be only one condensate $`\mathrm{\Sigma }(2)\mathrm{\Sigma }(3)`$, large and not affected by the vacuum fluctuation of $`\overline{q}q`$ pairs. Simultaneously, certain LEC’s like $`L_4(\mu )`$ and $`L_6(\mu )`$ should be tiny, since they are large $`N_c`$-suppressed and they violate the Okubo-Zweig-Iizuka (OZI) rule.
During the last years , there has been a growing evidence taking its roots in the observed properties of the scalar $`0^{++}`$ channel against this simplified “mean field approximation-type” picture of SB$`\chi `$S. It has been argued that vacuum fluctuations of $`\overline{q}q`$ could affect some order parameters, that $`\mathrm{\Sigma }(3)`$ can be about one half of $`\mathrm{\Sigma }(2)`$ , and that the low-energy constant $`L_6`$ could play the role of a third important order parameter describing fluctuations of the density of small fermionic modes . The purpose of this note is to report further arguments and evidences in favor of a particular role played by the constant $`L_6(\mu )`$ and by vacuum fluctuations of $`\overline{q}q`$. The paper has two parts. First, we reconsider the standard determination of LEC’s from Goldstone boson masses and decay constants as a function of the value of $`L_6`$. Next, we develop new arguments concerning a sum rule constraint on $`L_6`$ and on the difference between the two-flavor and three-flavor quark condensates.
2. We first consider the pseudoscalar masses $`M_\pi `$, $`M_K`$, $`M_\eta `$ and the corresponding decay constants (in the sequel we put $`m_u=m_d=m`$). We want to demonstrate that the extraction of low energy parameters from these observables is amazingly sensitive to the fine tuning of the OZI- and large $`N_c`$-suppressed constant $`L_6(\mu )`$: in particular, a relatively tiny increase of $`L_6(\mu )`$ implies a substantial drop of the three-flavor condensate $`\mathrm{\Sigma }(3)=\overline{u}u_{m=m_s=0}`$. We start by Eqs. (10.7) of the classical Gasser-Leutwyler’s paper and we rewrite them as:
$`F_\pi ^2M_\pi ^2`$ $`=`$ $`2m\mathrm{\Sigma }(3)+2m(m_s+2m)Z+4m^2A+4m^2B_0^2L+F_\pi ^2\delta _\pi ,`$ (1)
$`F_K^2M_K^2`$ $`=`$ $`(m_s+m)\mathrm{\Sigma }(3)+(m_s+m)(m_s+2m)Z`$
$`+(m_s+m)^2A+m(m_s+m)B_0^2L+F_K^2\delta _K.`$
Here, $`Z`$ and $`A`$ are scale-independent constants containing the LEC’s $`L_6(\mu )`$ and $`L_8(\mu )`$,
$`Z`$ $`=`$ $`32B_0^2\left[L_6(\mu ){\displaystyle \frac{1}{512\pi ^2}}\left(\mathrm{log}{\displaystyle \frac{M_K^2}{\mu ^2}}+{\displaystyle \frac{2}{9}}\mathrm{log}{\displaystyle \frac{M_\eta ^2}{\mu ^2}}\right)\right],`$ (3)
$`A`$ $`=`$ $`16B_0^2\left[L_8(\mu ){\displaystyle \frac{1}{512\pi ^2}}\left(\mathrm{log}{\displaystyle \frac{M_K^2}{\mu ^2}}+{\displaystyle \frac{2}{3}}\mathrm{log}{\displaystyle \frac{M_\eta ^2}{\mu ^2}}\right)\right],`$ (4)
with $`B_0=\mathrm{\Sigma }(3)/F_0^2`$ and $`F_0=lim_{m,m_s0}F_\pi `$. The remaining $`O(p^4)`$ chiral logarithms are contained in $`L`$: $`32\pi ^2L=3\mathrm{log}(M_K^2/M_\pi ^2)+\mathrm{log}(M_\eta ^2/M_K^2)`$, $`L=0.0253`$. There is a similar equation (not shown here) for $`F_\eta ^2M_\eta ^2`$, which besides $`\mathrm{\Sigma }(3)`$, $`L_6(\mu )`$ and $`L_8(\mu )`$ contains the LEC $`L_7`$, as well as three equations for $`F_P^2`$ ($`P`$=$`\pi `$, $`K`$, $`\eta `$). Notice that considering $`F_P^2M_P^2`$ and $`F_P^2`$ as independent observables allows one to separate from the start the “mass-type” LEC’s $`L_6`$, $`L_7`$, $`L_8`$ from the constants $`L_4`$, $`L_5`$ which merely show up in the decay constants $`F_P^2`$.
In Eqs. (1) and (S0.Ex1), all terms linear and quadratic in quark masses are explicitely shown and all remaining contributions starting by $`O(m_{\mathrm{quark}}^3)`$ are gathered in the remainders $`\delta _P`$. We are going to imagine that the latter are given to us, allowing one to treat Eqs. (1) and (S0.Ex1) as exact algebraic identities relating the three-flavor quark condensate $`\mathrm{\Sigma }(3)`$ \[expressed in physical units: $`X(3)=2m\mathrm{\Sigma }(3)/(F_\pi M_\pi )^2`$\], the quark mass ratio $`r=m_s/m`$, and the LEC’s $`F_0`$, $`L_6(\mu )`$ and $`L_8(\mu )`$. Although no expansion will be used, we want to investigate the consequences of the assumption $`\delta _PM_P^2`$ on the LEC’s. This might be interesting in connection with the recent discussion of convergence properties of three-flavor S$`\chi `$PT beyond $`O(p^4)`$ . It should be stressed that even setting $`\delta _P=0`$, we do not follow the framework of S$`\chi `$PT at one-loop order : we do not assume that the condensate $`\mathrm{\Sigma }(3)`$ dominates in Eqs. (1) and (S0.Ex1), and consequently, we do not treat $`1X(3)`$ as a small expansion parameter and do not replace in higher order terms $`2mB_0`$ by $`M_\pi ^2`$. We do not follow G$`\chi `$PT either , since we do not treat $`B_0`$ as an expansion parameter: even with $`\delta _P=0`$, Eqs. (1) and (S0.Ex1) go beyond the tree level of G$`\chi `$PT since they include chiral logarithms.
It is convenient to rewrite Eqs. (1) and (S0.Ex1) as:
$`{\displaystyle \frac{2m}{F_\pi ^2M_\pi ^2}}[\mathrm{\Sigma }(3)+(2m+m_s)Z]`$ $`=`$ $`1\stackrel{~}{ϵ}(r){\displaystyle \frac{4m^2B_0^2}{F_\pi ^2M_\pi ^2}}{\displaystyle \frac{rL}{r1}}\delta ,`$ (5)
$`{\displaystyle \frac{4m^2A}{F_\pi ^2M_\pi ^2}}`$ $`=`$ $`\stackrel{~}{ϵ}(r)+{\displaystyle \frac{4m^2B_0^2}{F_\pi ^2M_\pi ^2}}{\displaystyle \frac{L}{r1}}+\delta ^{},`$ (6)
where $`\stackrel{~}{ϵ}(r)=2(\stackrel{~}{r}_2r)/(r^21)`$ with $`\stackrel{~}{r}_2=2(F_KM_K)^2/(F_\pi M_\pi )^21`$. $`\delta `$ and $`\delta ^{}`$ are simple linear combinations of $`\delta _\pi `$ and $`\delta _K`$:
$`\delta `$ $`=`$ $`{\displaystyle \frac{r+1}{r1}}{\displaystyle \frac{\delta _\pi }{M_\pi ^2}}\left(\stackrel{~}{ϵ}+{\displaystyle \frac{2}{r1}}\right){\displaystyle \frac{\delta _K}{M_K^2}}`$ (7)
$`\delta ^{}`$ $`=`$ $`{\displaystyle \frac{2}{r1}}{\displaystyle \frac{\delta _\pi }{M_\pi ^2}}\left(\stackrel{~}{ϵ}+{\displaystyle \frac{2}{r1}}\right){\displaystyle \frac{\delta _K}{M_K^2}}.`$ (8)
For large $`r`$, one expects $`\delta ^{}\delta \delta _\pi /M_\pi ^2`$. As in Ref. , we consider as input parameters $`F_0`$ \[or equivalently $`L_4(\mu )`$\], the large $`N_c`$-suppressed constant $`L_6(\mu )`$ and the quark mass ratio $`r=m_s/m`$. Eq. (5) then yields the following non-perturbative formula for the three-flavor GOR ratio $`X(3)`$:
$$X(3)=\frac{2}{1+[1+\kappa (1\stackrel{~}{ϵ}\delta )]^{1/2}}(1\stackrel{~}{ϵ}\delta ),$$
(9)
where $`\kappa `$ contains the constant $`L_6(\mu )`$:
$`\kappa `$ $`=`$ $`64(r+2)\left({\displaystyle \frac{F_\pi M_\pi }{F_0^2}}\right)^2\{L_6(\mu )`$
$`{\displaystyle \frac{1}{256\pi ^2}}(\mathrm{log}{\displaystyle \frac{M_K}{\mu }}+{\displaystyle \frac{2}{9}}\mathrm{log}{\displaystyle \frac{M_\eta }{\mu }})+{\displaystyle \frac{rL}{16(r1)(r+2)}}\}.`$
Eq. (9) is an exact identity which is useful to the extent that the remainder $`\delta `$ in Eq. (7) is small, i.e. if the expansion of QCD correlation functions in powers of the quark masses $`m_u`$, $`m_d`$, $`m_s`$ globally converges. For instance, the latter requirement means that $`\delta _PM_P^2`$ in Eqs. (1) and (S0.Ex1), but not necessarily that the linear (condensate) term dominates.
The parameter $`\kappa `$ describes quantum fluctuations of the condensate and, indeed, $`\kappa =0(1/N_c)`$. On the other hand, $`\kappa `$ is actually small only if $`L_6(\mu )`$ is properly chosen: $`\kappa =0`$ for $`10^3L_6=0.26`$ at the scale $`\mu =M_\rho `$, which is close to the value advocated in classical S$`\chi `$PT analysis . In this case, Eq. (9) predicts $`X(3)`$ close to 1 unless the quark mass ratio $`r`$ substantially decreases and $`\stackrel{~}{ϵ}1`$. This effect is well known in G$`\chi `$PT . Quantum fluctuations can change this picture: the number in front of the curly bracket in Eq. (S0.Ex2) is huge ($`5340`$ for $`r=26`$ and $`F_0`$ = 85 MeV) and, consequently, even a rather small positive value of $`L_6(M_\rho )`$ can imply a substantial suppression of $`X(3)`$, independently of the value of $`r=m_s/m`$. The effect is represented on Fig. 1, where $`X(3)`$ is plotted as a function of $`L_6(M_\rho )`$ for $`r=20`$, $`r=25`$, $`r=30`$ and $`F_0`$ = 85 MeV. For smaller $`F_0`$, the decrease of $`X(3)`$ is even slightly steeper.
Once $`X(3)`$ is determined one can use Eq. (6) in order to infer the constant $`L_8(\mu )`$. The latter depends on $`L_6`$ merely through $`X(3)`$: the constant $`A`$ defined in Eq. (4) contains $`B_0^2L_8`$, is given by $`r=m_s/m`$ and is very little sensitive to $`L_6`$. Using the additional equations for $`F_\eta ^2M_\eta ^2`$ and for $`F_\pi ^2`$, $`F_K^2`$, $`F_\eta ^2`$, one can obtain $`L_7`$, $`L_4`$, $`L_5`$. The results are displayed in Table 1 as a function of $`L_6`$ for two values of $`F_0`$ and $`r=25`$.
3. It has been shown elsewhere that vacuum fluctuations of $`\overline{q}q`$ imply an enhancement of the two-flavor condensate $`\mathrm{\Sigma }(2)=lim_{m0}\overline{u}u|_{m_s\mathrm{fixed}}`$ with respect to $`\mathrm{\Sigma }(3)`$, and consequently: $`X(2)>X(3)`$. The two-flavor condensate can be obtained as a limit:
$$\mathrm{\Sigma }(2)=\underset{m0}{lim}\frac{(F_\pi M_\pi )^2}{2m}=\mathrm{\Sigma }(3)+m_sZ|_{m=0}+\delta \mathrm{\Sigma }(2)$$
(11)
keeping $`m_s`$ fixed. We have $`Z|_{m=0}=Z+B_0^2\mathrm{\Delta }Z`$, with $`\mathrm{\Delta }Z=[\mathrm{log}(M_K^2/\overline{M}_K^2)+2/9\mathrm{log}(M_\eta ^2/\overline{M}_\eta ^2)]/(16\pi ^2)`$ and $`\overline{M}_P^2=lim_{m0}M_P^2`$. $`\mathrm{\Delta }Z`$ can be obtained as a corollary of the previous analysis of masses and decay constants, and it has a very little effect. It has to be compared to the logarithmic piece of $`Z`$ in Eq. (3), taken at a typical scale $`\mu M_\rho `$. $`\mathrm{\Delta }Z`$ is less than the tenth of this logarithmic term.
Eliminating $`Z`$ from Eqs. (5) and (11), one arrives at the expression for the two-flavor GOR ratio $`X(2)=2m\mathrm{\Sigma }(2)/(F_\pi M_\pi )^2`$:
$`X(2)`$ $`=`$ $`[1\stackrel{~}{ϵ}]{\displaystyle \frac{r}{r+2}}+{\displaystyle \frac{2}{r+2}}X(3)`$
$`{\displaystyle \frac{(F_\pi M_\pi )^2}{2F_0^4}}X(3)^2\left[{\displaystyle \frac{2r^2}{(r1)(r+2)}}Lr\mathrm{\Delta }Z\right]+\delta _X.`$
In the expression for $`\delta _X`$, the remainders $`\delta _P/M_P^2`$ are suppressed by $`m/m_s`$: for $`r>20`$ one expects $`|\delta _X||\delta ^{}|`$. The dependence of $`X(2)`$ on $`L_6`$ is entirely hidden in $`X(3)`$ and consequently it is rather marginal, as can be seen from Fig. 1, where $`X(2)`$ is shown together with $`X(3)`$ for $`r=20`$, 25, 30. ($`X(2)`$ however remains strongly correlated with $`r=m_s/m`$, in particular for smaller values of $`r`$.) Fig. 1 clearly exhibits the lower bound $`L_6(M_\rho )>0.2110^3`$, as a consequence of the paramagnetic inequality $`X(3)<X(2)`$. A similar inequality holds for the decay constants $`F(3)<F(2)`$, where $`F(N_f)`$ denotes $`F_\pi `$ in the limit of first $`N_f`$ massless quarks. This inequality implies a lower bound for the large $`N_c`$-suppressed constant $`L_4`$. Equivalently, it can be reexpressed as an upper bound for $`F(3)=F_0`$: $`F_0<F(2)87`$ MeV. $`F_0`$ substantially smaller than this upper bound corresponds to a larger positive value of $`L_4`$. For instance, former studies of OZI-violation effects in the slopes of scalar form factors suggest a rather strong effect of $`L_4`$ (see also Ref. ) corresponding to the values $`F_0`$ as low as 71 MeV.
4. The standard values of LEC’s (see e.g. Ref. ), including the presumed OZI-rule suppression of $`L_6(\mu )`$ and $`L_4(\mu )`$ are, of course, compatible with the previous analysis. This is illustrated in the second half of the first line of Table 1 together with corresponding values $`X(2)X(3)0.9`$. The analysis of Goldstone boson masses and decay constants alone does not allow to conclude that vacuum fluctuations of $`\overline{q}q`$ actually produce an important effect. It however reveals an extreme sensitivity of the resulting LEC’s to the input value of $`L_6(\mu )`$: a modest shift of the latter towards small positive values produces an important splitting of $`X(2)`$ and $`X(3)`$ and an important increase of $`L_8(\mu )`$ and $`L_5(\mu )`$. An insight on the actual effect of vacuum fluctuations of $`\overline{q}q`$ can be obtained from sum rules for the connected part of the large $`N_c`$-suppressed two-point function $`\overline{u}u(x)\overline{s}s(0)^c`$ first analyzed in Ref. . In the remaining part of this work, we reexamine the constraints imposed by this sum rule on $`L_6(\mu )`$ and $`X(3)`$ avoiding to use S$`\chi `$PT and, in particular, to treat $`1X(3)`$ as a small expansion parameter. Our approach will be based on the the preceding analysis of masses and decay constants. In addition, we shall use Operator Product Expansion (OPE) of QCD to better control the high-energy contribution to the sum rule.
5. Let us consider the correlator:
$$\mathrm{\Pi }(p^2)=i\frac{mm_s}{M_\pi ^2M_K^2}\underset{m0}{lim}d^4xe^{ipx}0|T\{\overline{u}u(x)\overline{s}s(0)\}|0^c,$$
(13)
which is invariant under the QCD renormalization group and violates the OZI rule in the $`0^{++}`$ sector. For $`m_s0`$, $`\mathrm{\Pi }`$ is an order parameter for $`\mathrm{SU}_L(2)\times \mathrm{SU}_R(2)`$, related to the derivative of $`\mathrm{\Sigma }(2)`$ with respect to $`m_s`$: $`mm_s\mathrm{\Sigma }(2)/m_s=M_\pi ^2M_K^2\mathrm{\Pi }(0)`$. Differentiating Eq. (11) with respect to $`m_s`$ and using Eq. (3) to compute $`Z/m_s`$, we obtain a relation between $`\mathrm{\Pi }(0)`$ and $`Z`$ :
$`X(2)X(3)=2{\displaystyle \frac{mm_s}{F_\pi ^2M_\pi ^2}}Z|_{m=0}+{\displaystyle \frac{2m}{F_\pi ^2M_\pi ^2}}\underset{m0}{lim}{\displaystyle \frac{F_\pi ^2\delta _\pi }{2m}}=2{\displaystyle \frac{M_K^2}{F_\pi ^2}}\mathrm{\Pi }(0)`$ (14)
$`+{\displaystyle \frac{r[X(3)]^2}{32\pi ^2}}{\displaystyle \frac{F_\pi ^2M_\pi ^2}{F_0^4}}\left[\overline{\lambda }_K+{\displaystyle \frac{2}{9}}\overline{\lambda }_\eta \right]+{\displaystyle \frac{2m}{F_\pi ^2M_\pi ^2}}\left[1m_s{\displaystyle \frac{}{m_s}}\right]\underset{m0}{lim}{\displaystyle \frac{F_\pi ^2\delta _\pi }{2m}},`$
where $`\overline{\lambda }_P=m_s[\mathrm{log}\overline{M}_P^2]/m_s`$. We want to bring new information by estimating $`\mathrm{\Pi }(0)`$ from a QCD sum rule. The behaviour of $`\mathrm{\Pi }`$ at large momentum is given by OPE. It will involve operators transforming as $`(\overline{u}u)(\overline{s}s)`$ under the chiral group. In the limit $`m0`$, the lowest-dimension operator is $`m_s\overline{u}u`$. At the leading order in $`\alpha _s`$, the Wilson coefficient of this operator can be computed from two-loop Feynman diagrams, one loop stemming from the $`\overline{s}s`$ source and another one from $`\overline{u}u`$ (the latter is open, due to the insertion of the quark condensate). These quark loops are connected by two gluon lines. Hence, we deal with genuine two-loop integrals, formally similar to those involved in self-energy diagrams. The asymptotic expansion technique allows to simplify the computation of the large-momentum behaviour of such integrals, by performing various Taylor expansions of the propagators . This tedious calculation ends up with the lowest-dimension term in the Operator Product Expansion for $`\mathrm{\Pi }`$:
$$\mathrm{\Pi }(p^2)=\frac{18[12\zeta (3)]}{P^2}\left(\frac{\alpha _s}{\pi }\right)^2\frac{m_s^2m\overline{u}u}{M_\pi ^2M_K^2}+\mathrm{}$$
(15)
where $`P^2=p^2`$. Taking into account the running of $`\alpha _s`$ and of the masses, we see that $`\mathrm{\Pi }`$ vanishes faster than $`1/p^2`$ and is superconvergent .
We can write a sum rule for this correlator at zero momentum:
$`\mathrm{\Pi }(0)`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle _0^{s_1}}𝑑s\mathrm{Im}\mathrm{\Pi }(s){\displaystyle \frac{1}{s}}\left(1{\displaystyle \frac{s}{s_0}}\right)`$
$`+{\displaystyle \frac{1}{\pi }}{\displaystyle _{s_1}^{s_0}}𝑑s\mathrm{Im}\mathrm{\Pi }(s){\displaystyle \frac{1}{s}}\left(1{\displaystyle \frac{s}{s_0}}\right)+{\displaystyle \frac{1}{2i\pi }}{\displaystyle _{|s|=s_0}}𝑑s\mathrm{\Pi }(s){\displaystyle \frac{1}{s}}\left(1{\displaystyle \frac{s}{s_0}}\right).`$
The integral is made of 3 pieces. Along the real axis, from 0 to $`\sqrt{s_1}1.2\mathrm{GeV}`$, we compute explicitly the spectral function $`\mathrm{Im}\mathrm{\Pi }(s)`$ by solving two-channel coupled Omnès-Muskhelishvili equations for various $`T`$-matrix models . The second integral along the real axis ($`\sqrt{s_1}\sqrt{s}\sqrt{s_0}1.6\mathrm{GeV}`$) is roughly estimated using a second sum rule. The last integration (around the complex circle) is performed supposing that duality applies in the scalar channel above $`s_0`$.
6. For $`\sqrt{s}\sqrt{s_1}1.2\mathrm{GeV}`$, $`\pi \pi `$ and $`K\overline{K}`$ are expected to dominate the spectral function. No apparent $`4\pi `$\- and $`\eta \eta `$-contributions are observed in this energy range, and hence we will neglect these channels. We will follow closely the analysis of Refs. . The spectral function can be expressed in terms of the pion and kaon scalar form-factors:
$$\stackrel{}{F}(s)=\left(\begin{array}{c}0|\overline{u}u|\pi \pi \\ 0|\overline{u}u|K\overline{K}\end{array}\right),\stackrel{}{G}(s)=\left(\begin{array}{c}0|\overline{s}s|\pi \pi \\ 0|\overline{s}s|K\overline{K}\end{array}\right).$$
(17)
$`\stackrel{}{F}`$ and $`\stackrel{}{G}`$ satisfy (separately) a set of coupled Mushkelishvili-Omnès integral equations:
$$F_i(s)=\frac{1}{\pi }\underset{i=1}{\overset{n}{}}_{4M_\pi ^2}^{\mathrm{}}𝑑s^{}\frac{1}{s^{}s}T_{ij}^{}(s^{})\sqrt{\frac{s^{}4M_j^2}{s^{}}}\theta (s^{}4M_j^2)F_j(s^{}),$$
(18)
provided that the $`T`$-matrix behaves correctly when $`s\mathrm{}`$. Obviously, new channels open when the energy increases, which invalidates the two-channel approximation. However, we want $`\stackrel{}{F}`$ and $`\stackrel{}{G}`$ for $`ss_1`$, and the shape of the spectral function at much higher energies is not important. For our purposes, we can therefore impose on our $`T`$-matrix model any convenient large-energy behaviour. If $`\mathrm{\Delta }(s)`$ denotes the sum of the diagonal phase shifts, the condition $`\mathrm{\Delta }(\mathrm{})\mathrm{\Delta }(4M_\pi ^2)=2\pi `$ insures the existence and the unicity of the solution for the linear equation (18), once the values of the form factors at zero are fixed . Any solution can be projected on a basis of two solutions $`A`$ and $`B`$ such that $`\stackrel{}{A}(0)=\left(\genfrac{}{}{0pt}{}{1}{0}\right)`$ and $`\stackrel{}{B}(0)=\left(\genfrac{}{}{0pt}{}{0}{1}\right)`$ :
$$\stackrel{}{F}(s)=F_1(0)\stackrel{}{A}(s)+F_2(0)\stackrel{}{B}(s),\stackrel{}{G}(s)=G_1(0)\stackrel{}{A}(s)+G_2(0)\stackrel{}{B}(s).$$
(19)
The values of the scalar form factors at zero, $`\stackrel{}{F}(0)`$ and $`\stackrel{}{G}(0)`$, are related to the derivatives of the pion and kaon masses with respect to $`m`$ and $`m_s`$:
$$\gamma _P=\frac{m}{M_P^2}\left(\frac{M_P^2}{m}\right)_{m=0},\lambda _P=\frac{m_s}{M_P^2}\left(\frac{M_P^2}{m_s}\right)_{m=0},P=\pi ,K.$$
(20)
In particular, $`G_1(0)=(M_\pi ^2/m_s)_{m=0}=0`$. The spectral function is the sum of two terms:
$`\mathrm{Im}\mathrm{\Pi }(s)`$ $`=`$ $`\gamma _\pi \lambda _K\left[{\displaystyle \frac{\sqrt{3}}{32\pi }}{\displaystyle \underset{i=1,2}{}}\sqrt{{\displaystyle \frac{s4M_i^2}{s}}}A_i(s)B_i^{}(s)\theta (s4M_i^2)\right]`$
$`+\gamma _K\lambda _K{\displaystyle \frac{M_K^2}{M_\pi ^2}}\left[{\displaystyle \frac{1}{16\pi }}{\displaystyle \underset{i=1,2}{}}\sqrt{{\displaystyle \frac{s4M_i^2}{s}}}B_i(s)B_i^{}(s)\theta (s4M_i^2)\right].`$
The first integral in Eq. (S0.Ex5) is therefore of the form $`\gamma _\pi \lambda _K_{AB}(s_0,s_1)+\gamma _K\lambda _K_{BB}(s_0,s_1)`$. $`_{AB}`$ and $`_{BB}`$ depend only on $`s_0`$, $`s_1`$, and on the chosen $`T`$-matrix model . On the other hand, $`\gamma _\pi `$, $`\gamma _K`$ et $`\lambda _K`$ can be expressed through the expansion of pseudoscalar masses and decay constants in powers of the quark masses. The main difference with respect to Refs. lies here in the normalization of the integrals stemming from $`\stackrel{}{F}(0)`$ and $`\stackrel{}{G}(0)`$. As already emphasized, we do not follow the standard $`O(p^4)`$ analysis of the logarithmic derivatives in Eq. (20) . Even if the three-flavor condensate does not dominate, $`\lambda _P`$ and $`\gamma _P`$ can be determined from equations for $`F_P^2M_P^2`$ and $`F_P^2`$ similar to Eqs. (1) and (S0.Ex1).
7. The contribution of the integral under $`s_1`$ is positive and dominated by the $`f_0(980)`$ peak. On the other hand, $`\mathrm{\Pi }`$ is superconvergent and the integral of the spectral function from 0 to $`\mathrm{}`$ is zero. $`\mathrm{Im}\mathrm{\Pi }(s)`$ should therefore be negative somewhere. Let us suppose that it is the case for $`s_1ss_0`$.<sup>4</sup><sup>4</sup>4If the spectral function is partially positive in this range, our assumption yields an estimate for the second integral in Eq. (14) which is lower than in reality. In this case, we will underestimate $`X(2)X(3)`$. We can then estimate the contribution of the intermediate region in Eq. (14) by:
$$\frac{1}{s_0}𝒥^{}\frac{1}{\pi }_{s_1}^{s_0}𝑑s\mathrm{Im}\mathrm{\Pi }(s)\frac{1}{s}\left(1\frac{s}{s_0}\right)\frac{1}{s_1}𝒥^{},$$
(22)
where the integral $`𝒥^{}`$ can be estimated, using a second sum rule:
$`𝒥^{}`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle _{s_1}^{s_0}}𝑑s\mathrm{Im}\mathrm{\Pi }(s)\left(1{\displaystyle \frac{s}{s_0}}\right)`$
$`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle _0^{s_1}}𝑑s\mathrm{Im}\mathrm{\Pi }(s)\left(1{\displaystyle \frac{s}{s_0}}\right)+{\displaystyle \frac{1}{2i\pi }}{\displaystyle _{|s|=s_0}}𝑑s\mathrm{\Pi }(s)\left(1{\displaystyle \frac{s}{s_0}}\right).`$
The first integral on the right hand-side is known from the previously computed spectral function (S0.Ex6), and the second one will be evaluated by replacing $`\mathrm{\Pi }`$ by its asymptotic behaviour according to OPE, cf. Eq. (15).
The third integral in Eq. (S0.Ex5) is estimated through the OPE asymptotic behaviour of $`\mathrm{\Pi }`$. The factor $`(1s/s_0)`$ ensures that this contribution is suppressed for $`s`$ in the vicinity of $`s_0`$. Notice that in Eq. (15), $`\overline{u}u`$ stands for the quark condensate with $`N_f=2`$ (we take the chiral limit $`m0`$, but $`m_s`$ remains at its physical value). Its contribution can be calculated using standard techniques, relying on the expansion of the coupling constant along the circle :
$`{\displaystyle \frac{1}{2i\pi }}{\displaystyle _{|s|=s_0}}𝑑s\mathrm{\Pi }(s){\displaystyle \frac{1}{s}}\left(1{\displaystyle \frac{s}{s_0}}\right)=9[12\zeta (3)]`$ (24)
$`\times {\displaystyle \frac{F_\pi ^2}{M_K^2}}X(2){\displaystyle \frac{m_s^2(s_0)}{s_0}}a(s_0)^2[16.5a(s_0)+48.236a^2(s_0)],`$
where $`a=\alpha _s/\pi `$. The contribution is negative, but strongly suppressed by $`\alpha _s^2`$ and $`m_s^2/s_0`$. Let us remember that, due to the probable importance of direct instanton contributions , the duality in the $`0^{++}`$ channel is not expected to extend to as low energies as in other channels.
8. Hence, we can estimate $`X(2)X(3)`$ in two different ways : the first one is Eq. (S0.Ex3), whereas the second one combines Eq. (14) and the sum rule Eq. (S0.Ex5). In both cases, $`X(2)X(3)`$ can be expressed as a function of observables and of $`[F_0,r,X(3)]`$. This overdetermination can be considered as a constraint fixing $`X(3)`$ as a function of $`r`$ and $`F_0`$. A typical result is shown on Fig. 2 for $`F_0=85`$ MeV.
There are three sources of uncertainties in this analysis. i) First, there are next-to-next-to-leading (NNL) remainders in the expressions for masses and decay constants. The control of their effect is rather clear in Eq. (S0.Ex3) ($`\delta _X\delta ^{}`$), but somewhat less transparent in Eq. (14) and in the estimates of the logarithmic derivatives $`\lambda _P`$ and $`\gamma _P`$. Within S$`\chi `$PT, authors of Ref. have found that the two-loop effects on the $`m_s`$-dependence of $`\mathrm{\Sigma }(2)`$ are likely to be small and of the same sign as the one-loop effects. A similar conclusion is reached in Ref. . NNL remainders are assumed here to be small, and they are not shown in our results.
ii) Next, the evaluation of the sum rule Eq. (S0.Ex5) involves a rough estimate of the integral between $`s_1`$ and $`s_0`$. Hence, for a given couple $`(F_0,r)`$, the present evaluation of the sum rule will not provide a single value of $`X(3)`$, but a whole range of acceptable values, depending in addition on the separators $`s_1<s_0`$. It is seen on Fig. 2 that the upper bound for $`X(3)`$ remains stable for $`\sqrt{s_0}>1.5`$ GeV, whereas the lower bound depends rather strongly on $`s_0`$. When $`s_0`$ increases, the lower bound in Eq. (22) appears too weak to be saturated. A more stringent lower bound would be welcome.
iii) The third source of uncertainty is the multi-channel $`T`$-matrix used as input to build the spectral function in Eq. (S0.Ex6) for $`s<s_1`$. Three different $`T`$-matrix models have been used , but only the results corresponding to Ref. are shown in this letter<sup>5</sup><sup>5</sup>5More complete results and discussion will be presented elsewhere .. The distinctive feature is the shape of the $`f_0(980)`$ peak, which tunes the suppression of $`X(3)`$. Among the three models of Refs. , the one corresponding to Fig. 2, Ref. , leads to the least pronounced effect. The two other models lead to a higher $`f_0(980)`$ peak, a larger value of $`\mathrm{\Pi }(0)`$ and a smaller value of $`X(3)`$ than shown in Fig. 2: typically, at $`r=25`$, the upper bound for $`X(3)`$ would be slightly below 0.5 and low values of $`r`$, such as $`r15`$ would even be excluded by the stability requirement $`X(3)>0`$.
The sum-rule results for $`X(3)`$ can be converted into corresponding results for $`L_{i=4\mathrm{}8}`$ as a function of $`r`$ and $`F_0`$ (see Fig.3). For low $`r`$, the LEC’s reach very large values: their definition includes $`1/B_0`$ factors that make them diverge when $`X(3)0`$. More interesting is the enhancement of $`L_5`$, $`L_6`$, $`L_7`$ and $`L_8`$ in the vicinity of $`r25`$: it reflects the drift of $`L_6(M_\rho )`$ to small *positive* values, which is dictated by the sum rule. Notice that the second large-$`N_c`$ suppressed LEC, $`L_4`$, is not predicted here: it is closely correlated to the input value of $`F_0`$. Strictly speaking, $`F_0`$ (and $`L_5`$) can be constrained by studying the slopes of the form factors in Eq. (19) . This procedure would yield rather low values of $`F_0`$ (between 70 and 75 MeV), supporting a larger value of the OZI-rule violating constant $`L_4`$ . On the other hand, the decrease of $`F_0`$ does not modify very much the bounds on $`X(3)`$ and the consequences of Eqs. (9) and (S0.Ex3). For this reason, this part of the discussion will be presented separately .
9. A few points to summarize and conclude. i) According to the non-perturbative formula Eq. (9), vacuum fluctuations of $`\overline{q}q`$ will suppress the three-flavor condensate $`X(3)`$, unless $`L_6(M_\rho )`$ is in a narrow band around $`0.2610^3`$, which further shrinks with increasing $`L_4`$. The effect is nearly independent of the quark mass ratio $`r=m_s/m`$. ii) Sum rules for the correlator $`\overline{u}u\overline{s}s^c`$ constrain $`L_6(M_\rho )`$ outside this narrow band, suggesting that the suppression of $`X(3)`$ due to the vacuum fluctuations of $`\overline{q}q`$ actually takes place. It is due to the importance of the $`f_0(980)`$ peak ($`I=J=0^{++}`$) and to the QCD control of higher-energy contributions. iii) The presence of at least three light (massless) flavors in the sea seems crucial for this effect: vacuum fluctuations do *not* suppress the two-flavor condensate $`X(2)`$. The latter remains large ($`0.91`$) unless $`r<20`$, see Eq. (S0.Ex3). iv) In principle, $`X(2)`$ can be measured in low-energy $`\pi \pi `$-scattering . If the data are sufficiently accurate, they could be used to determine for the first time $`L_8+2L_6`$. According to the present analysis, this combination of LEC’s could be (for $`r=25`$) 5 times larger than the values reported and used previously . On the other hand, $`\pi \pi `$-scattering can hardly tell us anything about the three-flavour condensate $`X(3)`$, the splitting $`X(2)X(3)`$ and about the vacuum fluctuations of $`\overline{q}q`$ in general.
The authors thank B. Moussallam for many helpful discussions and for providing his program solving numerically Omnes-Mushkelishvili equations, and M. Knecht, E. Oset, H. Sazdjian and P. Talavera for various discussions and comments. Work partly supported by the EU, TMR-CT98-0169, EURODA$`\mathrm{\Phi }`$NE network.
|
warning/0007/gr-qc0007074.html
|
ar5iv
|
text
|
# Gravitational Waves from a Compact Star in a Circular, Inspiral Orbit, in the Equatorial Plane of a Massive, Spinning Black Hole, as Observed by LISA
## I INTRODUCTION AND SUMMARY
Earth-based gravitational-wave detectors operate in the high-frequency band, $`1`$$`10^4`$ Hz, in which lie the waves from black holes of masses $`2`$$`10^3M_{}`$. Space-based detectors operate in the low-frequency band, $`10^4`$$`1`$ Hz populated by waves from black holes of mass $`10^3`$$`10^8M_{}`$. The high-frequency band is likely to be opened up early in the next decade by the LIGO-VIRGO network of earth-based detectors . The premier instrument for the low-frequency band is the Laser Interferometer Space Antenna (LISA) .
The European Space Agency has selected LISA as one of three “Cornerstone” missions in its “Horizon 2000+” program, NASA has appointed a mission definition team for LISA, and the ESA and NASA teams are talking to each other informally about the possibility of flying LISA as a joint ESA/NASA mission in the $`2010`$ time frame.
One of the most interesting and promising gravitational wave sources for LISA is the final epoch of inspiral of a compact, stellar-mass object into a massive black hole. In the LISA frequency band, where the central hole must have $`M10^8M_{}`$, all giant stars and main-sequence stars will be tidally disrupted before the end of their inspiral, but compact objects—white dwarfs, neutron stars, and small black holes—can survive intact. (Depending on the hole’s spin, a massive white dwarf will be disrupted before the end of inspiral if $`M<M_{\mathrm{max}}10^4`$$`10^5M_{}`$. Neutron stars and small black holes can never be tidally disrupted in the LISA frequency band.)
Sigurdsson and Rees have estimated the event rate for such compact objects to spiral into massive black holes. “Assuming most spiral galaxies have a central black hole of modest mass ($`10^6M_{}`$) and a cuspy spheroid,” and for “very conservative estimates of the black hole masses and central galactic densities,” they estimate one inspiral per year within 1Gpc distance of Earth. Most of the inspiraling objects are likely to be white dwarfs or neutron stars; the inspiral rate for stellar-mass black holes ($`\mu 6`$$`10M_{}`$) may be ten times smaller, about 3 per year out to 3Gpc, according to Sigurdsson . Sigurdsson notes, however, that that the evidence for a recent burst of star formation in the central region of our galaxy suggests that normal nucleated spirals might have such starbursts every $`10^8`$ years, which would enhance the stellar-mass black-hole density by a factor $`10`$ and would lead to stellar-mass black-hole inspirals of one per year out to one Gpc. He notes, further, that if there was just one $`50M_{}`$ black hole in the core of each galaxy now containing a $`10^6M_{}`$ central black hole, the result would be several inspirals of such $`50M_{}`$ holes per year out to a cosmological redshift $`z=1`$, all readily observable by LISA.
LISA’s observations of waves from such inspirals will have major scientific payoffs :
Ryan has shown that for circular equatorial orbits, the waves will carry, encoded in themselves, a map of the vacuum spacetime metric of the central black hole (or, equivalently, the values of the hole’s multiple moments), and he has made a first, very crude, estimate of the precision with which LISA can extract that map . Ryan’s estimate is quite promising. From the extracted map, one can determine whether the hole’s geometry is that of the Kerr metric (i.e. “test the black-hole no hair theorem”), and one can use such maps to search for other kinds of conjectured massive central bodies (e.g. soliton stars and naked singularities). It seems likely that this is true not only for circular geodesic orbits, but also for generic orbits.
The observation of many such events will provide (i) a census of the masses and spins of the massive central holes, (ii) a census of the masses of the inspiraling objects (which depend on and thus tell us about the initial stellar mass function and mass segregation in the central parsec of galactic nuclei), and (iii) a census of event rates (which depend on physical processes and on gravitational potentials in the central parsec).
In active galactic nuclei, the inspiral orbit may be significantly affected by drag in an accretion disk, producing both complications in the interpretation of the observations and opportunities for learning about the disks’ mass distribution .
In planning for the LISA mission, it is important to understand the details of the waves emitted by such inspirals. Those details are the most important factors in the choice of the mission’s noise floor and its duration, and are likely to be the principal drivers of its data analysis requirements and algorithms.
The foundations for computing the emitted waves are nearly all in place:
If the orbit is known, then the waveforms and strengths can be computed using the Teukolsky – Sasaki-Nakamura (TSN) formalism for first-order perturbations of Kerr black holes.
The orbital evolution is governed by radiation reaction (and, if there is a robust accretion disk present, by accretion-disk drag ). Most massive holes are in galaxies with normal (non-active) nuclei, and are thought to be surrounded by tenuous disks with “advection-dominated accretion flow” (ADAF). Narayan has shown that accretion drag should be totally negligible in such ADAF disks, so the orbital evolution is very cleanly governed by radiation reaction. This is the situation that we analyze in this paper; we ignore accretion-disk drag. Those few holes that are in active galactic nuclei may be surrounded by “thin” or “slim” accretion disks, for which Chakrabarti and colleagues have shown that accretion-disk drag may be significant.
The radiation reaction’s influence on the orbit can be characterized fully by the rates of change of three “constants” of the orbital motion: the orbital energy $`E`$, axial component of angular momentum $`L`$, and Carter constant $`Q`$ . From the emitted waves (computed via the TSN formalism), one can read off $`\dot{E}dE/dt`$ and $`\dot{L}`$; but the only known way to compute $`\dot{Q}`$ is directly from the radiation reaction force.
A formal expression for the radiation reaction force has been derived recently by Mino et. al. and by Quinn and Wald , and several researchers are now working hard to convert this into a practical computational tool for deducing $`\dot{Q}`$ . This will complete the necessary set of tools for computing all details of the emitted waves.
The emitted waves will be so complex and so rich in structure and in parameter dependence, that it will require extensive computations to give us the full knowledge required by the LISA mission. Those computations are proceeding in stages:
1. Initial quick surveys, based on the Newtonian or quasi-Newtonian orbits and the quadrupole-moment approximation to gravitational-wave emission. Such surveys are the foundation for the event rate estimates by Sigurdsson and Rees discussed above.
2. More detailed and accurate surveys for orbits in the massive hole’s equatorial plane, using the TSN formalism. Such surveys do not require computing $`\dot{Q}`$, since $`Q`$ vanishes for equatorial orbits. These surveys are of several types:
1. Studies of the evolution of the orbit’s eccentricity. Such studies have been carried out by Tanaka et. al. and Cutler et. al. for non-spinning holes, and by Kennefick for small eccentricities around spinning holes. These studies, coupled with estimates of the orbital eccentricities when the objects are far from the hole and are being frequently perturbed by near encounters with other objects , suggest that, despite the circularizing effect of radiation reaction, the eccentricities will still typically be large, $`e0.3`$, when the object nears the hole’s horizon.
2. Systematic computations of the details of the emitted waves and the orbital evolution for circular, equatorial orbits. This paper presents such computations and a companion paper extends them to the transition regime, near the innermost stable circular orbit (isco), during which the orbit makes a gradual transition from adiabatic inspiral to a plunge into the hole.
3. Computations of the waves’ details and orbital evolution for elliptic, equatorial orbits. First explorations have been carried out by Shibata for general ellipticity and by Kennefick for small ellipticity (though in the 1970s and 1980s there were studies for equatorial orbits that plunge from radial infinity into a hole or scatter off a hole .
3. Surveys of the orbital evolution and waves for circular orbits out of the hole’s equatorial plane. It is known that radiation reaction drives circular orbits into circular orbits, thereby causing $`\dot{Q}`$ to evolve in a manner that is fully determined by TSN-formalism calculations of $`\dot{E}`$ and $`\dot{Q}`$ . Therefore, the tools are fully in hand for these surveys, and Hughes is in the late stages of the first one. (See Shibata for an exploration of the wave emission before anyone knew how, correctly, to compute the orbital evolution, and see Shibata et. al. for studies of orbits with very small inclination angles to the equatorial plane.)
4. Surveys of orbital evolution and waves for the generic, most realistic situation: elliptic orbits outside the equatorial plane. Such surveys must await a practical computation technique for $`\dot{Q}`$.
For circular, equatorial orbits (the subject of this paper), there have been extensive previous calculations, beginning with the pioneering study by Detweiler ; for a review see Mino et. al. . However, these previous calculations have been motivated by the needs of LIGO/VIRGO observations in the high-frequency band, where (i) the ratio $`\mu /M`$ of object mass to hole mass is not very small, so finite-mass-ratio effects (omitted by the TSN formalism) are important, and (ii) almost all of the observed inspiral signal comes from radii large compared to the hole’s horizon, so post-Newtonian techniques can be used. The previous calculations have focused almost entirely on carrying the post-Newtonian calculations to very high order, on developing techniques for accelerating their convergence, and—via comparison with TSN calculations—on evaluating their convergence .
LISA’s regime and needs are quite different from this. For LISA, most of the signals are likely to come from systems with extreme mass ratios, $`\mu /M1`$, for which (a) the TSN formalism is highly accurate and (b) the object lingers for a very long time in the vicinity of the hole’s horizon before plunging into it. This means that post-Newtonian calculations are neither needed, nor appropriate.
Because of these differences between the LIGO-VIRGO regime and the LISA regime, the previous TSN-based studies do not serve LISA’s needs. The purpose of this paper is to begin filling that gap. Specifically:
In this paper we introduce a new set of functions $`𝒩`$, $`𝒯`$, $`\dot{}`$, to characterize the orbital evolution and the emitted waves. These functions are dimensionless and of order unity, and depend on the hole’s dimensionless spin parameter $`a=(`$angular momentum$`)/M^2`$ and on the orbit’s dimensionless radius $`\stackrel{~}{r}=r/M`$. We give extensive tables of these functions, as computed by one of us (LSF) using the TSN formalism. We then use those tables to compute the evolution of the waves’ frequency and signal strength in LISA for a number of instructive values of the parameters $`M=(`$hole mass), $`a=(`$hole spin parameter), $`\mu =(`$object mass), and $`m=(`$wave harmonic order$`)(`$wave frequency$`)/(`$orbital frequency). From these computations we draw a number of conclusions of importance for the LISA mission.
The paper is organized as follows. Our notation, including the dimensionless functions $`𝒩`$, $`𝒯`$, …, is introduced in Sec. II. Formulas for computing the dimensionless functions, and formulas for the orbital evolution and the waves’ properties are given in Sec. III. Tables of the dimensionless functions are given and discussed in Sec. IV. Applications to LISA are presented in Sec. V. Finally, concluding remarks are given in Sec. VI.
## II Notation
In this paper we shall adopt the following notation to describe the compact object’s inspiral and the gravitational waves it emits; throughout we use geometrized units, i.e. we set $`G(`$Newton’s gravitation constant$`)=1`$ and $`c`$(speed of light$`)=1`$.
The mass of the inspiraling object.
The black hole’s mass.
The mass ratio, assumed $`1`$.
The hole’s “rotation parameter;” here $`S`$ is the hole’s spin angular momentum.
The orbit’s Boyer-Lindquist radial coordinate; defined by $`\sqrt{r^2+a^2(1+2M/r)}=(1/2\pi )\times `$(the object’s orbital circumference)
A tilde over a quantity means that it has been made dimensionless by multiplying by the appropriate power of $`M`$ and, when the quantity is $`\mu `$, multiplying by a factor $`1/\mu `$.
The dimensionless radius of the orbit.
The object’s orbital angular velocity, as measured using Boyer-Lindquist coordinate time $`t`$ (defined below), i.e., using clocks that are far from the hole and at rest with respect to it.
The dimensionless orbital angular velocity, which is related to $`\stackrel{~}{r}`$ by $`\stackrel{~}{\mathrm{\Omega }}=1/(\stackrel{~}{r}^{3/2}+a)`$; cf. Eq. (2.16) of Ref. . When $`\stackrel{~}{\mathrm{\Omega }}`$ is small (large $`\stackrel{~}{r}`$), Kepler’s laws dictate that $`\stackrel{~}{\mathrm{\Omega }}(M/r)^{3/2}=(`$orbital velocity$`)^3`$, i.e. (orbital velocity$`)\mathrm{\Omega }^{1/3}`$.
A quantity evaluated at the object’s innermost stable circular orbit (“isco”), where the inspiral ends and the plunge begins; for example, $`\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}`$ is the value of $`\stackrel{~}{\mathrm{\Omega }}`$ at the isco.
Boyer-Lindquist coordinate time, or equivalently time as measured at “radial infinity” or on Earth.
The Boyer-Lindquist time $`\mathrm{\Delta }t`$ until the isco is reached; i.e., the total remaining duration of the inspiral.
The number of orbits remaining until the isco is reached.
The distance from the binary to Earth.
The order of a harmonic of the orbital frequency.
The frequency of gravitational waves in the $`m`$’th harmonic.
The object’s total energy including rest mass, i.e., the component $`p_t`$ of its 4-momentum. Note that, because the object is gravitationally bound to the black hole, $`E<\mu `$, its gravitational binding energy is $`\mu E>0`$.
The total rate of emission of energy into gravitational waves that go to infinity.
The total rate of emission of energy into gravitational waves that go down the horizon.
The total rate of emission of energy into gravitational waves that go both to infinity and down the hole’s horizon; and also, by energy conservation, the rate of decrease of the object’s total energy; i.e., $`\dot{E}_{\mathrm{GW}}dE_{\mathrm{GW}}/dt=dE/dt`$.
The total rate of emission of energy into the $`m`$’th harmonic of the waves that go to infinity.
The rms amplitude of the gravitational waves in harmonic $`m`$ emitted toward infinity, at a time when the wave frequency is $`f_m`$; here $`h_{m+}(t,𝐧)`$ and $`h_{m\times }(t,𝐧)`$ are the two waveforms emitted in a direction $`𝐧`$ and arriving at the Earth’s distance $`r_o`$; $`\mathrm{}`$ is an average over $`𝐧`$ and over a period of the waves; and the average over time automatically produces a factor $`1/2`$ thereby making $`h_{o,m}^2`$ be the mean value of $`\frac{1}{2}[(`$amplitude of $`h_{m+})^2+(`$amplitude of $`h_{m,\times })^2]`$.
: A characteristic amplitude for the waves in harmonic $`m`$; here and throughout this paper the dot denotes a time derivative. The significance of $`h_{c,m}`$ is discussed below.
$`h_{c,m}\mathrm{min}[1,\sqrt{3(1f_m/f_{m,\mathrm{isco}})}]`$: A modified characteristic amplitude, discussed below.
: LISA’s “sky-averaged” rms noise in a bandwidth equal to frequency $`f`$. Here $`S_h^{\mathrm{SA}}(f)`$ is the one-sided spectral density $`S_{h_+}(f)`$ for some linear polarization $`+`$, inverse averaged over source directions and polarization (“sky-averaged”), $`1/S_h^{\mathrm{SA}}1/S_{h_+}`$.
From the general relation $`dE_{\mathrm{GW}}/dtdA=(1/16\pi )(\dot{h}_{+}^{}{}_{}{}^{2}+\dot{h}_{\times }^{}{}_{}{}^{2})`$ for the energy flux in gravitational waves in terms of the time derivatives of the two waveforms (Eq. (10) of Ref. ), we infer that the rms amplitude and the energy in harmonic $`m`$ are related to each other by
$$h_{o,m}=\frac{2\sqrt{\dot{E}_\mathrm{}m}}{m\mathrm{\Omega }r_o}.$$
(1)
We shall use, as our measure of where the object is in its orbit, the dimensionless orbital angular frequency $`\stackrel{~}{\mathrm{\Omega }}`$, which is related to the gravitational-wave frequency in harmonic $`m`$ by $`f_m=(m/2\pi M)\stackrel{~}{\mathrm{\Omega }}`$. We shall write various fully relativistic, time-evolving quantities ($`\dot{E}_{\mathrm{GW}}`$, $`h_{c,m}`$, etc.) as the leading-order (“Newtonian”) term in an expansion in $`\stackrel{~}{\mathrm{\Omega }}^{1/3}(`$orbital velocity), multiplied by relativistic corrections. Our notation for the relativistic corrections will be the following:
The correction to $`\mathrm{\Omega }^2/\dot{\mathrm{\Omega }}\stackrel{~}{\mathrm{\Omega }}^2/\dot{\stackrel{~}{\mathrm{\Omega }}}`$, where the dot is a time derivative. Note that $`\mathrm{\Omega }^2/\dot{\mathrm{\Omega }}=d\mathrm{\Phi }/d\mathrm{ln}\mathrm{\Omega }`$ is the number of radians $`d\mathrm{\Phi }`$ of orbital motion required to produce (due to radiation reaction) a fractional change $`d\mathrm{\Omega }/\mathrm{\Omega }`$ in the orbital frequency.
The correction to $`N_{\mathrm{orb}}`$ (the number of orbits remaining until the end of the inspiral).
The correction to $`T`$ (the remaining time to the end of the inspiral).
The correction to $`\dot{E}_{\mathrm{GW}}`$ (the total energy loss rate).
The correction to $`\dot{E}_\mathrm{}m`$ (the energy radiated to infinity in harmonic $`m`$).
The correction to $`h_{o,m}`$ (the rms wave amplitude in harmonic $`m`$).
The correction to $`h_{c,m}`$ (the characteristic amplitude in harmonic $`m`$).
The characteristic amplitude $`h_{c,m}`$ needs some explanation. As the object spirals inward in its orbit, its $`m`$’th harmonic waves spend $`f_m^2/\dot{f}_m=d\mathrm{\Phi }_m/(2\pi d\mathrm{ln}f_m)`$ cycles in the vicinity of frequency $`f_m`$ (where $`\mathrm{\Phi }_m`$ is the harmonic’s phase). Correspondingly, in a detector that observes the waves throughout the inspiral epoch $`\mathrm{\Delta }t=f_m/\dot{f}_m`$, the signal is enhanced, in comparison to the detector noise, by approximately the square root of this quantity. The signal strength is thus approximately the same as would be produced by a broad-band burst of amplitude $`h_{c,m}h_{o,m}\sqrt{2f_m^2/\dot{f}_m}`$.
The factor $`2`$ inside the square root arises from a more precise definition of $`h_{c,m}`$ : The signal to noise ratio produced by the waves’ $`m`$’th harmonic, averaged over all possible orientations of the source and the detector, is given by
$$\left(\frac{S}{N}\right)_{\mathrm{rms}}=\sqrt{\left[\frac{h_{c,m}(f_m)}{h_\mathrm{n}(f_m)}\right]^2d\mathrm{ln}f_m},$$
(2)
where $`h_n(f_m)`$ is the detector’s rms noise at frequency $`f_m`$, in a bandwidth equal to frequency, averaged over the sky. Equation (2) serves as a definition of $`h_{c,m}(f)`$. The relation $`h_{c,m}(f_m)=h_{o,m}(f_m)\sqrt{2f_m^2/\dot{f}_m}`$ then follows from Eq. (29) of Ref. (with the factor $`2`$ changed to $`4`$ to correct an error), together with the definition of $`h_{o,m}`$ given above, Eqs. (2) and (1), and the evaluation of Fourier transforms using the stationary phase approximation.
When the inspiraling object nears the isco, the bandwidth available for building up its signal in the detector becomes less than $`\mathrm{\Delta }f=f`$. A good measure of this reduced bandwidth is $`\mathrm{\Delta }f=2(ff_{\mathrm{isco}})`$ (with half of this band below $`f`$ and half above). This is less than $`f`$ for $`2f_{\mathrm{isco}}/3<f<f_{\mathrm{isco}}`$. Correspondingly, the amplitude of the built-up signal is $`h_c\sqrt{2(f_{\mathrm{isco}}f)/(2f_{\mathrm{isco}}/3)}`$. Our modified characteristic amplitude $`h_{c,m}^{}h_{c,m}\mathrm{min}[1,\sqrt{3(1f_m/f_{m,\mathrm{isco}})}]`$ takes this signal reduction into account.
## III Formulas for Inspiral and Waves
In this section we shall give leading-order (in $`\stackrel{~}{\mathrm{\Omega }}^{1/3}`$) formulas for the various time-evolving quantities, as functions of the dimensionless orbital angular frequency $`\stackrel{~}{\mathrm{\Omega }}`$ and black-hole spin $`a`$, and thereby we shall produce exact definitions of the relativistic correction functions. To make clear the magnitudes of various quantities, we shall write some of our formulas numerically in a form relevant to LISA (for which we choose as a fiducial frequency $`f_2=0.01\mathrm{Hz}`$ and as a fiducial source, a $`\mu =10M_{}`$ black hole spiraling into a $`M=10^6M_{}`$ hole at $`r_o=1`$ Gpc distance from Earth). We shall also write our formulas in a form relevant to the LIGO-VIRGO network of high-frequency detectors (with, as our fiducial frequency, $`100\mathrm{H}\mathrm{z}`$, and our fiducial source, a $`1M_{}`$ neutron star spiraling into a $`100M_{}`$ hole at 1Gpc distance).
In the Newtonian limit, the orbital radius and orbital angular velocity are linked by the Keplerian relation
$$\frac{M}{r}\frac{1}{\stackrel{~}{r}}=(M\mathrm{\Omega })^{2/3}\stackrel{~}{\mathrm{\Omega }}^{2/3}.$$
(3)
This permits us to write the number of orbital radians spent near orbital angular frequency $`\mathrm{\Omega }`$ in the following form (cf. Eqs. (3.16) of MTW , in which $`a`$ is our orbital radius $`r`$):
$`{\displaystyle \frac{\mathrm{\Omega }^2}{\dot{\mathrm{\Omega }}}}`$ $`=`$ $`{\displaystyle \frac{d\mathrm{\Phi }}{d\mathrm{ln}\mathrm{\Omega }}}={\displaystyle \frac{5}{96}}{\displaystyle \frac{1}{\eta }}{\displaystyle \frac{1}{\stackrel{~}{\mathrm{\Omega }}^{5/3}}}𝒩`$ (4)
$`=`$ $`{\displaystyle \frac{1.17\times 10^5}{(f_2/.01\mathrm{Hz})^{5/3}}}\left({\displaystyle \frac{10M_{}}{\mu }}\right)\left({\displaystyle \frac{10^6M_{}}{M}}\right)^{\frac{2}{3}}𝒩`$ (5)
$`=`$ $`{\displaystyle \frac{117}{(f_2/100\mathrm{H}\mathrm{z})^{5/3}}}\left({\displaystyle \frac{1M_{}}{\mu }}\right)\left({\displaystyle \frac{100M_{}}{M}}\right)^{\frac{2}{3}}𝒩.`$ (6)
Here $`𝒩`$ is the general relativistic correction, which is unity in the “Newtonian” limit $`\stackrel{~}{\mathrm{\Omega }}1`$. Similarly, the total remaining time until the end of the inspiral (Eq. (36.17b) of MTW) is
$`T`$ $`=`$ $`{\displaystyle \frac{5}{256}}{\displaystyle \frac{1}{\eta }}{\displaystyle \frac{M}{\stackrel{~}{\mathrm{\Omega }}^{8/3}}}𝒯`$ (7)
$`=`$ $`{\displaystyle \frac{1.41\times 10^6\mathrm{sec}}{(f_2/.01\mathrm{Hz})^{8/3}}}\left({\displaystyle \frac{10M_{}}{\mu }}\right)\left({\displaystyle \frac{10^6M_{}}{M}}\right)^{\frac{2}{3}}𝒯`$ (8)
$`=`$ $`{\displaystyle \frac{0.141\mathrm{sec}}{(f_2/100\mathrm{H}\mathrm{z})^{8/3}}}\left({\displaystyle \frac{1M_{}}{\mu }}\right)\left({\displaystyle \frac{100M_{}}{M}}\right)^{\frac{2}{3}}𝒯,`$ (9)
and the number of orbits remaining until the end of the inspiral is
$`N_{\mathrm{orb}}`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{ln}\mathrm{\Omega }}^{\mathrm{ln}\mathrm{\Omega }_{\mathrm{isco}}}}{\displaystyle \frac{d\mathrm{\Phi }}{d\mathrm{ln}\mathrm{\Omega }}}d\mathrm{ln}\mathrm{\Omega }={\displaystyle \frac{1}{64\pi }}{\displaystyle \frac{1}{\eta }}{\displaystyle \frac{1}{\stackrel{~}{\mathrm{\Omega }}^{5/3}}}𝒩_{\mathrm{orb}}`$ (10)
$`=`$ $`{\displaystyle \frac{1.11\times 10^4}{(f_2/.01\mathrm{Hz})^{5/3}}}\left({\displaystyle \frac{10M_{}}{\mu }}\right)\left({\displaystyle \frac{10^6M_{}}{M}}\right)^{\frac{2}{3}}𝒩_{\mathrm{orb}}`$ (11)
$`=`$ $`{\displaystyle \frac{11.1}{(f_2/100\mathrm{H}\mathrm{z})^{5/3}}}\left({\displaystyle \frac{1M_{}}{\mu }}\right)\left({\displaystyle \frac{100M_{}}{M}}\right)^{\frac{2}{3}}𝒩_{\mathrm{orb}}.`$ (12)
By integrating Eq. (5) inward to the isco, one can derive the following expression for the general relativistic corrections $`𝒯`$ for $`T`$ and $`𝒩_{\mathrm{orb}}`$ for $`N_{\mathrm{orb}}`$ in terms of that $`𝒩`$ for $`\mathrm{\Omega }^2/\dot{\mathrm{\Omega }}`$:
$$𝒯=\frac{8}{3}\stackrel{~}{\mathrm{\Omega }}^{8/3}_{\stackrel{~}{\mathrm{\Omega }}}^{\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}}\frac{𝒩d\stackrel{~}{\mathrm{\Omega }}}{\stackrel{~}{\mathrm{\Omega }}^{11/3}},$$
(13)
$$𝒩_{\mathrm{orb}}=\frac{5}{3}\stackrel{~}{\mathrm{\Omega }}^{5/3}_{\stackrel{~}{\mathrm{\Omega }}}^{\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}}\frac{𝒩d\stackrel{~}{\mathrm{\Omega }}}{\stackrel{~}{\mathrm{\Omega }}^{8/3}},$$
(14)
The total energy loss rate \[Eq. (3.16) of MTW\] is
$$\dot{E}_{\mathrm{GW}}=\dot{E}=\frac{32}{5}\eta ^2\stackrel{~}{\mathrm{\Omega }}^{10/3}\dot{},$$
(15)
where $`\dot{}`$ is the general relativistic correction.
When the object is at large radii (small $`\stackrel{~}{\mathrm{\Omega }}`$), the power radiated by the system’s mass multipole moments $`I^{l,\pm m}`$ is of order $`\eta ^2\stackrel{~}{\mathrm{\Omega }}^{2+2l/3}`$, while that radiated by its current multipole moments $`S^{l,\pm m}`$ is of order $`\eta ^2\stackrel{~}{\mathrm{\Omega }}^{2+2(l+1)/3}`$ . Correspondingly, the power $`\dot{E}_\mathrm{}m`$ radiated to infinity in harmonic $`m`$ comes almost entirely from the moments of lowest allowed orders $`l`$, with the current moments of order $`l`$ being, á priori, comparable to the mass moments of order $`l+1`$.
For $`m=1`$, the lowest allowed order for either mass or current is quadrupolar, since gravitational waves are always quadrupolar or higher. For circular orbits in the equatorial plane, the $`m=\pm 1`$ components of the mass quadrupole moment vanish, so the dominant waves are current quadrupolar $`S^{2,\pm 1}`$ and mass octupolar $`I^{3,\pm 1}`$. All other multipolar contributions to $`\dot{E}_\mathrm{}1`$ are smaller than these by at least $`(\mathrm{\Omega }r)^2=\stackrel{~}{\mathrm{\Omega }}^{2/3}`$. The contributions of $`S^{2,\pm 1}`$ to the radiated power $`\dot{E}_\mathrm{}1`$ can be derived from Eqs. (4.16), (5.27), (2.18a,c), (2.7) and (2.8) of Ref. ; and those of $`I^{3,\pm 1}`$, from Eqs. (4.16), (5.27), (2.7), and (2.8) of . The sum of these dominant contributions is
$$\dot{E}_\mathrm{}1=\frac{5}{28}\eta ^2\stackrel{~}{\mathrm{\Omega }}^4_\mathrm{}1,$$
(16)
where we have tacked on the general relativistic correction factor $`\dot{}_\mathrm{}1`$ to account for contributions from all the higher-order multipoles and to make the formula be valid not just for large orbital radii $`r`$ but for all $`rr_{\mathrm{isco}}`$. In Eq. (16), the numerical factor is $`5/28=8/45+1/1260`$, where the big piece $`8/45`$ is current quadrupolar, while the tiny piece $`1/1260`$ is mass octupolar.
For harmonic $`m2`$, the lowest allowed multipoles are of order $`l=m`$, and the mass moment $`I^{m,\pm m}`$ is nonzero so it dominates. All other multipolar contributions to $`\dot{E}_m`$ are down from these by at least $`\stackrel{~}{\mathrm{\Omega }}^{2/3}`$. An expression for $`\dot{E}_m`$ can be derived from Eqs. (4.16), (5.27), (2.7), and (2.8) of Ref. . The result is
$`\dot{E}_m=`$ $`{\displaystyle \frac{2(m+1)(m+2)(2m+1)!m^{2m+1}}{(m1)[2^mm!(2m+1)!!]^2}}`$ (18)
$`\times \eta ^2\stackrel{~}{\mathrm{\Omega }}^{2+2m/3}\dot{}_\mathrm{}m,`$
where $`\dot{}_\mathrm{}m`$ is the relativistic correction and $`(2m+1)!!(2m+1)(2m1)(2m3)\mathrm{}1`$. For $`m=2`$, $`3`$, and $`4`$, this expression reduces to
$`\dot{E}_2=`$ $`{\displaystyle \frac{32}{5}}\eta ^2\stackrel{~}{\mathrm{\Omega }}^{10/3}\dot{}_\mathrm{}2,`$ (19)
$`\dot{E}_3=`$ $`{\displaystyle \frac{243}{28}}\eta ^2\stackrel{~}{\mathrm{\Omega }}^4\dot{}_\mathrm{}3,`$ (20)
$`\dot{E}_4=`$ $`{\displaystyle \frac{8192}{567}}\eta ^2\stackrel{~}{\mathrm{\Omega }}^{14/3}\dot{}_\mathrm{}4.`$ (21)
Note that the low-$`\stackrel{~}{\mathrm{\Omega }}`$ limit of $`\dot{E}_2`$ is identical to that of the total energy loss $`\dot{E}_{\mathrm{GW}}`$ \[Eq. (15)\], as it must be since the $`m=2`$ harmonic dominates at low orbital velocities.
From Eqs. (16) and (18) for $`\dot{E}_m`$ and the general relationship (1) between the waves’ amplitude and energy, we obtain the following Newtonian-order expression for the amplitude in harmonic $`m`$:
$`h_{o,1}=`$ $`\sqrt{{\displaystyle \frac{5}{7}}}{\displaystyle \frac{\eta M}{r_o}}\stackrel{~}{\mathrm{\Omega }}_{o,1},`$ (23)
$`h_{o,m}=`$ $`\sqrt{{\displaystyle \frac{8(m+1)(m+2)(2m+1)!m^{2m1}}{(m1)[2^mm!(2m+1)!!]^2}}}`$ (25)
$`\times {\displaystyle \frac{\eta M}{r_o}}\stackrel{~}{\mathrm{\Omega }}^{m/3}_{o,m}\text{for }m2,`$
where the relativistic correction is related to that for the energy by
$$_{o,m}=\sqrt{\dot{}_\mathrm{}m}.$$
(26)
For the dominant, $`m=2`$, radiation Eq. (25) becomes
$`h_{o,2}=\sqrt{{\displaystyle \frac{32}{5}}}{\displaystyle \frac{\eta M}{r_o}}\stackrel{~}{\mathrm{\Omega }}^{2/3}_{o,2}`$ (27)
$`={\displaystyle \frac{3.6\times 10^{22}}{r_o/1\mathrm{G}\mathrm{p}\mathrm{c}}}\left({\displaystyle \frac{\mu }{10M_{}}}\right)\left({\displaystyle \frac{M}{10^6M_{}}}\right)^{\frac{2}{3}}\left({\displaystyle \frac{f_2}{.01\mathrm{Hz}}}\right)^{\frac{2}{3}}_{o,2}`$ (28)
$`={\displaystyle \frac{3.6\times 10^{23}}{r_o/1\mathrm{G}\mathrm{p}\mathrm{c}}}\left({\displaystyle \frac{\mu }{M_{}}}\right)\left({\displaystyle \frac{M}{100M_{}}}\right)^{\frac{2}{3}}\left({\displaystyle \frac{f_2}{100\mathrm{H}\mathrm{z}}}\right)^{\frac{2}{3}}_{o,2}.`$ (29)
(30)
From Eqs. (III) for $`h_{o,m}`$, the definition of $`h_{c,m}`$ in terms of $`h_{o,m}`$, the relation $`f_m=(m/2\pi )\mathrm{\Omega }`$, and Eq. (5), we obtain the following expression for the characteristic amplitude in harmonic $`m`$
$`h_{c,1}=`$ $`{\displaystyle \frac{5}{\sqrt{672\pi }}}{\displaystyle \frac{\eta ^{1/2}M}{r_o}}\stackrel{~}{\mathrm{\Omega }}^{1/6}_{c,1},`$ (32)
$`h_{c,m}=`$ $`\sqrt{{\displaystyle \frac{5(m+1)(m+2)(2m+1)!m^{2m}}{12\pi (m1)[2^mm!(2m+1)!!]^2}}}`$ (34)
$`\times {\displaystyle \frac{\eta ^{1/2}M}{r_o}}\stackrel{~}{\mathrm{\Omega }}^{(2m5)/6}_{c,m}\text{for }m2,`$
where \[using Eq. (26)\] the relativistic correction is related to earlier ones by
$$_{c,m}=\sqrt{𝒩\dot{}_\mathrm{}m}.$$
(35)
For $`m=2`$, expression (34) becomes
$`h_{c,2}=\sqrt{{\displaystyle \frac{2}{3\pi }}}{\displaystyle \frac{\eta ^{1/2}M}{r_o\stackrel{~}{\mathrm{\Omega }}^{1/6}}}_{c,2}`$ (36)
$`={\displaystyle \frac{1.0\times 10^{19}}{r_o/1\mathrm{G}\mathrm{p}\mathrm{c}}}\left({\displaystyle \frac{\mu }{10M_{}}}\right)^{\frac{1}{2}}\left({\displaystyle \frac{M}{10^6M_{}}}\right)^{\frac{1}{3}}\left({\displaystyle \frac{.01\mathrm{Hz}}{f_2}}\right)^{\frac{1}{6}}_{c,2}`$ (37)
$`={\displaystyle \frac{3.2\times 10^{22}}{r_o/1\mathrm{G}\mathrm{p}\mathrm{c}}}\left({\displaystyle \frac{\mu }{M_{}}}\right)^{\frac{1}{2}}\left({\displaystyle \frac{M}{100M_{}}}\right)^{\frac{1}{3}}\left({\displaystyle \frac{100\mathrm{H}\mathrm{z}}{f_2}}\right)^{\frac{1}{6}}_{c,2}.`$ (38)
(39)
All of the relativistic correction functions can be expressed analytically in terms of $`\dot{}`$, and $`\dot{}_\mathrm{}m`$. This has almost been done already: The correction functions $`𝒯`$, $`_{o,m}`$ and $`_{c,m}`$ have been expressed in terms of $`\dot{}`$, $`\dot{}_\mathrm{}m`$, $`𝒩`$ and $`\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}`$ by Eqs. (13), (26) and (35) respectively. All that remains is to derive an expression for $`𝒩`$ in terms of $`\dot{}`$, and an expression for $`\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}`$.
The derivations are based on the Kerr-metric relations
$$E=\eta M\frac{12/\stackrel{~}{r}+a/\stackrel{~}{r}^{3/2}}{\sqrt{13/\stackrel{~}{r}+2a/\stackrel{~}{r}^{3/2}}}$$
(40)
for the object’s total energy in terms of its dimensionless orbital radius $`\stackrel{~}{r}`$ (Eq. (5.4.7b) of Ref. ), and
$$\stackrel{~}{r}=(\stackrel{~}{\mathrm{\Omega }}^1a)^{2/3}.$$
(41)
for its orbital radius in terms of its orbital angular velocity (Eq (2.16) of Ref. ). By differentiating these equations with respect to time and combining with each other and with Eqs. (5) and (15) for $`\mathrm{\Omega }^2/\dot{\mathrm{\Omega }}`$ and $`\dot{E}_{\mathrm{GW}}=\dot{E}`$, we obtain
$`𝒩=`$ $`{\displaystyle \frac{1}{\dot{}}}\left(1+{\displaystyle \frac{a}{\stackrel{~}{r}^{3/2}}}\right)^{5/3}\left(1{\displaystyle \frac{6}{\stackrel{~}{r}}}+{\displaystyle \frac{8a}{\stackrel{~}{r}^{3/2}}}{\displaystyle \frac{3a^2}{\stackrel{~}{r}^2}}\right)`$ (43)
$`\times \left(1{\displaystyle \frac{3}{\stackrel{~}{r}}}+{\displaystyle \frac{2a}{\stackrel{~}{r}^{3/2}}}\right)^{3/2}.`$
When $`\stackrel{~}{r}`$ is regarded as the function (41) of $`\stackrel{~}{\mathrm{\Omega }}`$, this becomes the desired expression for $`𝒩`$ in terms of $`\dot{}`$ and $`\stackrel{~}{\mathrm{\Omega }}`$.
The innermost stable circular orbit (isco) is at the location $`\stackrel{~}{r}_{\mathrm{isco}}`$ where the object’s total energy $`E(\stackrel{~}{r})`$ is a minimum, or equivalently where $`\dot{\mathrm{\Omega }}`$ is infinite, or equivalently where $`𝒩`$ vanishes; i.e., $`\stackrel{~}{r}_{\mathrm{isco}}`$ is that root of the quartic equation $`\stackrel{~}{r}^26\stackrel{~}{r}+8a\stackrel{~}{r}^{1/2}3a^2=0`$ which lies between 1 (when $`a=1`$) and 6 (when $`a=0`$). An analytic expression for $`r_{\mathrm{isco}}`$ has been given by Bardeen, Press, and Teukolsky:
$`\stackrel{~}{r}_{\mathrm{isco}}=`$ $`3+Z_2\mathrm{sign}(a)[(3Z_1)(3+Z_1+2Z_2)]^{1/2},`$ (46)
$`Z_11+(1a^2)^{1/3}[(1+a)^{1/3}+(1a)^{1/3}],`$
$`Z_2(3a^2+Z_{1}^{}{}_{}{}^{2})^{1/2}.`$
The dimensionless orbital angular velocity at the isco $`\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}`$ is expressed in terms of this $`\stackrel{~}{r}_{\mathrm{isco}}`$ by Eq. (41):
$$\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}=\frac{1}{\stackrel{~}{r}_{\mathrm{isco}}^{}{}_{}{}^{3/2}+a}.$$
(47)
We note, in passing, approximate analytic formulae for the relativistic corrections $`𝒯`$ and $`𝒩_{\mathrm{orb}}`$ to the time $`T`$ and number of remaining orbits $`𝒩_{\mathrm{orb}}`$ until the end of inspiral—for the special case of a nonspinning black hole, $`a=0`$. Inserting expression (43) into Eqs. (13) and (14), and noting from Table II that in each of these equations $`\dot{}`$ is a much more slowly varying function of the integration variable than the rest of the integrand, we pull $`\dot{}`$ out of the integral (i.e., we perform the first step of an integration by parts) and then perform the integration analytically. The results are
$`𝒯`$ $``$ $`{\displaystyle \frac{1}{\dot{}}}[{\displaystyle \frac{1\frac{7}{2}u\frac{147}{8}u^2\frac{2205}{16}u^3+\frac{19845}{16}u^4}{\sqrt{13u}}}`$ (49)
$`{\displaystyle \frac{4671}{8\sqrt{2}}}u^4+{\displaystyle \frac{19845}{32}}u^4\mathrm{ln}\left({\displaystyle \frac{3(1+\sqrt{2})^2u}{(1+\sqrt{13u})^2}}\right)],`$
$$𝒩_{\mathrm{orb}}\frac{1}{\dot{}}\left(\frac{14u48u^2+288u^3}{\sqrt{13u}}24\sqrt{3u^5}\right),$$
(50)
where $`u=\stackrel{~}{\mathrm{\Omega }}^{2/3}=1/\stackrel{~}{r}`$. These formulae agree with the numerical values of $`𝒯`$ and $`𝒩_{\mathrm{orb}}`$ in Tables IX and X below to within 3 per cent at $`6.02<\stackrel{~}{r}<18`$, and to within 1 per cent at $`6\stackrel{~}{r}<6.02`$ and $`\stackrel{~}{r}>18`$.
## IV Tables of Relativistic Correction Functions
We shall use two dimensionless parameters to measure the distance of an orbit from the isco: the ratio $`r/r_{\mathrm{isco}}\stackrel{~}{r}/\stackrel{~}{r}_{\mathrm{isco}}`$ of the orbit’s Boyer-Lindquist radial coordinate $`r`$ to its value at the isco, and the ratio $`\mathrm{\Omega }/\mathrm{\Omega }_{\mathrm{isco}}\stackrel{~}{\mathrm{\Omega }}/\stackrel{~}{\mathrm{\Omega }}_{\mathrm{isco}}`$ of the orbit’s angular velocity to that at the isco. The relationship between these two parameters is given by Eqs. (41), (46) and (47), and is tabulated in Table I.
We have integrated the Teukolsky-Sasaki-Nakamura equation for perturbations of a Kerr black hole to obtain the functions $`\dot{}(\stackrel{~}{\mathrm{\Omega }})`$ and $`\dot{}_\mathrm{}m(\stackrel{~}{\mathrm{\Omega }})`$, and we have then used Eqs. (41), (43), (13) and (14) to compute $`𝒩`$, $`𝒯`$ and $`𝒩_{\mathrm{orb}}`$. These functions are listed in Tables IIX and some of our numerical methods are described in the Appendix. As a byproduct of these calculations, we have inferred what fraction $`\dot{E}/\dot{E}_{\mathrm{GW}}`$ of the total rate of energy emission goes down the hole’s horizon; that fraction is shown in Table VII.
## V Applications to LISA
### A LISA Noise
Tentative error budgets for LISA are spelled out in Tables 4.1 and 4.2 of the LISA Pre-Phase-A Report . Various researchers have computed LISA noise spectra from those error budgets. It is conventional, for LISA, to characterize the noise by the sensitivity to periodic sources for one year integration time and a signal-to-noise ratio of 5, averaged over source directions and polarizations (“sky averaged”). We shall denote this quantity by $`h_{\mathrm{SN5},1\mathrm{y}\mathrm{r}}^{\mathrm{SA}}`$. It is related to the sky-averaged spectral density introduced in the paragraph before Eq. (1) by $`h_{\mathrm{SN5},1\mathrm{y}\mathrm{r}}^{\mathrm{SA}}=5\sqrt{S_h^{\mathrm{SA}}\mathrm{\Delta }f}`$, where $`\mathrm{\Delta }f=1/1\mathrm{y}\mathrm{r}`$ is the bandwidth for the one-year integration time; and correspondingly, it is related to the sky-averaged rms noise in a bandwidth equal to frequency, $`h_n(f)=\sqrt{fS_h^{\mathrm{SA}}(f)}`$ (which we use in this paper), by
$$h_n(f)=\frac{1}{5}\sqrt{\frac{f}{\mathrm{\Delta }f}}h_{\mathrm{SN5},1\mathrm{y}\mathrm{r}}^{\mathrm{SA}}(f).$$
(51)
We have deduced $`h_n(f)`$ using this equation and the values of $`h_{\mathrm{SN5},1\mathrm{y}\mathrm{r}}^{\mathrm{SA}}(f)`$ computed by various researchers; we plot it as a thick solid curve in Figs. 37 below.
It is likely that LISA’s performance will be compromised at $`f0.003`$ Hz by a stochastic background due to white-dwarf binaries. The most recent estimate of that stochastic background is by Hils and Bender ; it agrees satisfactorily with an estimate by Webbink and Han. We have used a simple piece-wise straight-line fit to the logarithm of the Hils-Bender white-dwarf-background noise curve: straight lines that join the following points in $`(\mathrm{log}_{10}f,\mathrm{log}_{10}h_{\mathrm{SN5},1\mathrm{y}\mathrm{r}}^{\mathrm{SA}})`$ where $`f`$ is measured in Hz:
$`(4,20.518),(3.62,20.737),(2.78,21.66),`$ (52)
$`(2.61,22.90),(2,23.731).`$ (53)
This white-dwarf noise, converted to our conventions via Eq. (51), is shown as a thick dashed curve in Figs. 37 below.
### B Detectable Systems
Of greatest interest, for probing the spacetime geometries of massive black holes, is the gravitational radiation emitted during the last year of inspiral of a compact object. In planning the LISA mission, it is important to know the detectability of these final-year waves, as a function of the system’s parameters: the hole’s mass $`M`$ and spin $`a`$, the object’s mass $`\mu `$, and the distance $`r_o`$ from Earth. Previous studies of this issue have assumed that the massive hole is nonspinning, $`a=0`$.
It is straightforward to compute the rms signal to noise ratio $`(S/N)_{\mathrm{rms}}`$ (averaged over detector and system orientations) from Eq. (2), using the noise amplitudes $`h_n`$ described above and the dominant $`m=2`$ characteristic amplitude $`h_{c,2}`$ of Eqs. (39) and (35), with $`𝒩`$ and $`\dot{}_\mathrm{}m`$ taken from Tables VIII and IV. In this calculation, the frequency $`f_2=\mathrm{\Omega }/\pi `$ ranges from its value at time $`T=1`$ year \[Eq. (8) and Table IX\] to its value at the isco.
In view of the complexity of the data analysis for these waves, a signal to noise ratio of about 10 may be required for their detection, and in view of the estimated event rates (Sec. I), it is necessary that LISA see out to at least $`r_o=1`$Gpc. Accordingly, we have computed the range of masses $`\mu `$ and $`M`$ and black-hole spins $`a`$ for which $`(S/N)_{\mathrm{rms}}>10`$ at a distance $`r_o=1`$ Gpc. This range of “detectable systems” is shown in Fig. 1 for LISA without the white-dwarf background (solid curves) and with the background (dashed curves).
Several features of this figure deserve comment:
* Inspiraling white dwarfs and neutron stars ($`\mu 1.4M_{}`$) are barely detectable, with $`(S/N)_{\mathrm{rms}}=10`$, at 1Gpc. It would be highly desirable to reduce LISA’s design noise floor by a factor two or three, to give greater confidence of detection.
* For $`\mu =10M_{}`$ inspiraling black holes, the detectable systems have a wide range of central black-hole masses: $`10^4M_{}M10^7M_{}`$.
* The upper limit on detectable central-hole masses $`M`$ depends strongly on the black-hole spin: for $`\mu =10M_{}`$ it ranges from $`2\times 10^6M_{}`$ to $`3\times 10^7M_{}`$ without the white-dwarf background, and $`1\times 10^6`$ to $`1.5\times 10^7`$ with the background. (The spins shown are for no rotation $`a=0`$, and for near the maximum rotation, $`a\pm 0.998`$, that can be produced by spinup via accretion from a disk .)
* The white dwarf background reduces the maximum detectable black hole mass by about a factor 2.5, independent of the spin.
* The white dwarf background and the black hole spin have little influence on the minimum detectable mass $`M`$. This is because, at low $`M`$ the object travels a large radial distance in its last year of life, so most of the signal to noise comes from radii $`rr_{\mathrm{isco}}`$ where the spin is unimportant, and (by virtue of the small $`M`$) most comes from frequencies high enough that the white-dwarf background is negligible.
For probing the immediate vicinity of the horizon, we are interested in waves with frequencies, say, $`(2/3)f_{2,\mathrm{isco}}<f_2<f_{2,\mathrm{isco}}`$. Figure 2 shows the range of systems for which $`(S/N)_{\mathrm{rms}}>10`$ in this frequency band, at a distance $`r_o=1`$Gpc, during the last year of inspiral. Note that restricting attention to this near-horizon frequency range has reduced substantially the set of detectable systems: for $`\mu =10M_{}`$, the minimum black-hole mass is increased by a factor $`20`$ to $`100`$, depending on the spin $`a`$. Nevertheless, there is still a wide range of systems accessible for study.
For distances larger than $`r_o1`$ Gpc, cosmological effects have a significant influence on the signal . At fixed $`\{a,M(1+z),\mu (1+z)\}`$ (where $`z`$ is the cosmological redshift), the characteristic amplitude and signal to noise ratio scale $`1/r_{oL}`$, where $`r_{oL}`$ is the luminosity distance to Earth. The scaling of $`(S/N)_{\mathrm{rms}}`$ with $`\mu (1+z)`$ is not so simple, because it influences the waves’ frequency evolution in complicated ways that entail the relativistic correction functions. For extremely rough estimates, one can use the leading-order (in $`\stackrel{~}{\mathrm{\Omega }}`$) expression for $`h_{c,2}`$ \[Eq. (39), $`S/Nh_{c,2}\mu ^{1/2}/r_{\mathrm{oL}}`$\] to infer $`\mu _{\mathrm{min}}(1+z)r_{oL}^2`$ for the minimum detectable object mass at fixed $`a`$ and $`M(1+z)`$, but for reliable results, one must repeat the analysis (sketched above) by which we arrived at Figs. 1 and 2.
### C Evolution of the Waves During Inspiral
To gain insight into the emitted waves and how they evolve during the inspiral, we have constructed Figs. 37. Each figure depicts the waves’ evolution for the value of object mass $`\mu `$ and hole mass $`M`$ (in solar masses) listed in bold letters in the upper right corner. The horizontal axis is the waves’ frequency $`f`$ and the vertical axis, their modified characteristic amplitude $`h_c^{}`$. As the inspiral proceeds, the waves sweep upward in frequency (left to right) along one of the thin curves. These evolutionary curves are shown for three different values of the black-hole spin, $`a=0.99`$ (retrograde orbit; short-dashed curves), $`a=0`$ (no rotation; long-dashed curves) and $`a=+0.999`$ (prograde orbit; solid curves). For each spin, three curves are shown corresponding to the three lowest harmonics $`m=1,2,3`$ of the orbital frequency. The values of $`a`$ and $`m`$ for each evolutionary curve are listed near the vertical endpoint of the curve. Also shown in each figure is the rms noise amplitude $`h_n`$ for LISA: a thick solid curve in the absence of a white-dwarf-binary background and a thick dashed curve including that background.
The range of frequency sweep is strongly dependent on the masses $`\mu `$ and $`M`$ of object and hole. Neglecting the relativistic correction factor $`𝒯`$ (which is unimportant for this purpose when the frequency sweep is substantial), Eq. (8) tells us that $`f_{\mathrm{isco}}/f_{1\mathrm{y}\mathrm{r}}(\mu /M)^{3/8}(1/M)^{3/8}`$, where $`f_{1\mathrm{y}\mathrm{r}}`$ is the frequency one year before reaching the isco. Thus, the greatest frequency sweep is for the least extreme mass ratio and the smallest hole mass, $`\mu /M=10/10^5`$ (Fig. 6) with $`f_2`$ sweeping from $`0.006`$ Hz to $`0.4`$ Hz; while the smallest sweep is for the most extreme mass ratio and largest hole mass, $`\mu /M=10/10^7`$ (Fig. 7) with $`f_2`$ sweeping only from $`0.0023`$ to $`0.0027`$ Hz.
The height of a signal curve $`h_c^{}`$ above the noise curve $`h_n`$ is about equal to the signal to noise ratio in an appropriate bandwidth $`\mathrm{\Delta }f`$: $`\mathrm{\Delta }f=f`$ well away from the endpoint of inspiral, and $`\mathrm{\Delta }f=2(ff_{\mathrm{isco}})`$ near the endpoint; cf. the discussion of the definition of $`h_c^{}`$ at the end of Sec. II. Near the endpoint of inspiral $`h_c^{}`$ plunges for three reasons: (i) because of the narrowing of our chosen bandwidth; (ii) because the rate of frequency sweep speeds up due to flattening of the effective potential for the object’s radial motion, and this produces a reduction in the number of cycles $`N_{\mathrm{cyc}}`$ in a given bandwidth and reduction in $`h_c^{}\sqrt{N_{\mathrm{cyc}}}`$; (iii) because, for large $`a`$ and prograde orbits, the orbit sinks deep into the throat of the hole’s embedding diagram, from where waves have difficulty escaping.
On each signal curve there are three solid dots. They label $`(f,h_c^{})`$ for specific times during the inspiral: $`T=1`$ year before the endpoint (leftmost dot), $`T=1`$ month before the end (center dot), and $`T=1`$ day before the end (right dot). Beside the dots for the dominant harmonic, $`m=2`$, are shown two numbers that characterize the orbit and waves at that time: the radius $`\stackrel{~}{r}=r/M`$ of the orbit in units of the black-hole mass, and the number of gravitational-wave cycles in the $`m=2`$ harmonic, from that time until the endpoint of inspiral. At the bottom end of each $`m=2`$ curve is shown the radius $`\stackrel{~}{r}`$ of the isco.
It is worthwhile to scrutinize the details of these figures, including the numbers beside the dots. Consider, for example, Fig. 4 for a $`\mu =10M_{}`$ object (black hole) spiraling into a $`M=10^6M_{}`$ hole. If the big hole is rapidly rotating and the orbit is prograde so $`a=+0.999`$, then the dominant $`m=2`$ evolutionary curve shows the object, one year before its death, at $`\stackrel{~}{r}=6.80`$ (3.4 Schwarzschild radii), with a signal to noise ratio of $`h_c/h_n100`$, and with 185,000 cycles of gravitational waves left until death. One month before death, the object is at $`\stackrel{~}{r}=3.05`$ (1.53 Schwarzschild radii), with $`h_c/h_n50`$, and with 40,000 cycles left. One day before death, it is at $`\stackrel{~}{r}=1.30`$ (compared to 1.18 for the isco), with $`h_c/h_n10`$ and with 2,320 cycles left. It is impressive how long the object lingers in the vicinity of the horizon, and how many wave cycles it emits.
For a nonspinning hole $`a=0`$, the numbers are less impressive but still remarkable: the last year is spent spiraling from $`\stackrel{~}{r}=9.46`$ (4.73 Schwarzschild radii) to the isco at $`\stackrel{~}{r}=6`$ (3 Schwarzschild radii), during which 85,000 wave cycles are emitted and $`h_c/h_n`$ drops from $`100`$ to $`10`$ at one day and then to zero.
The large number of wave cycles carry a large amount of information about the source. We shall discuss this issue in Sec. V D below.
Figures 35 illustrate the influence of the mass of the inspiraling object on the signal strength. For $`M`$ fixed at $`10^6M_{}`$, one year before merger the $`m=2`$ signal to noise ratios $`h_c/h_n`$ are $`15`$ for $`\mu =1M_{}`$, $`100`$ for $`\mu =10M_{}`$ and $`500`$ for $`\mu =100M_{}`$. This is a moderately faster growth than our crude estimate $`\mu ^{1/2}`$ in Sec. V B. Notice that $`h_c/h_n`$ drops below 10 one month before the endpoint for $`\mu =1M_{}`$, and one day before the end for $`\mu =10M_{}`$.
To maximize the exploration of the horizon’s vicinity, we want the object to spend its entire last year at radii $`\stackrel{~}{r}10`$. If the object is a $`10M_{}`$ hole, this is the case when $`M10^6M_{}`$; cf. Figs. 4, 6 and 7. For $`M<10^6M_{}`$, such exploration is debilitated by the large frequency sweep; cf. Fig. 6. We have previously met this issue in Sec. V B.
Figure 7 shows that the white-dwarf-binary background is a serious issue for hole masses $`M10^7M_{}`$, while Figs. 36 show that it is relatively unimportant for $`M10^6M_{}`$. We have previously met this in Sec. V B.
### D Information Carried by the Waves
As is well known , the waves’ highest accuracy information is carried by the time evolution of their phase. For circular, equatorial orbits, where there is no orbital precession, the phase evolution will be the same for all the harmonics as for the orbit itself, and that phase evolution is embodied in $`d\mathrm{\Phi }/d\mathrm{ln}\mathrm{\Omega }=\mathrm{\Omega }^2/\dot{\mathrm{\Omega }}`$. Equation (5) shows this quantity, at fixed frequency, to be proportional to $`𝒩/M_{\mathrm{chirp}}^2`$, where
$$M_{\mathrm{chirp}}=\mu ^{1/2}M^{1/3}$$
(54)
is the system’s chirp mass. Since a year of observations will typically entail $`N_{\mathrm{cycle}}10^5`$ cycles of waves, and by the method of matched filters one can detect a secular shift of one waveform with respect to another by a small fraction of a cycle , the “raw” precision for measuring the evolution of $`𝒩/M_{\mathrm{chirp}}`$ will be of order $`10^6`$.
If most of the last year is spent near the horizon, say at frequencies $`f/f_{\mathrm{isco}}=\mathrm{\Omega }/\mathrm{\Omega }_{\mathrm{isco}}0.1`$ (as will usually be the case), then this phase evolution will depend strongly not only on the chirp mass, but also—through the function $`𝒩(f/f_{\mathrm{isco}})`$—on the black-hole spin parameter $`a`$.
This strong $`a`$-dependence is exhibited in Fig. 8. Even for $`a<0.5`$, where the curves $`𝒩(f/f_{\mathrm{isco}})`$ for different $`a`$ look very close together, $`𝒩/a0.1`$, this $`a`$-dependence translates into $`N_{\mathrm{cycles}}/a10^4`$, which is huge. Thus, it is reasonable to expect the measured phasing to determine both $`a`$ and $`M_{\mathrm{chirp}}`$ to high precision—though a detailed parameter study is needed to be absolutely certain.
The absolute frequencies associated with the observed phase evolution (e.g., the measured frequency at the end of inspiral) are determined by a combination of $`a`$ and the hole’s mass $`M`$. This absolute frequency scale presumably will be measured much less accurately than the phasing itself, but still, probably, accurately enough to determine the mass $`M`$ to a very interesting precision. Knowing $`M_{\mathrm{chirp}}`$, $`a`$, and $`M`$, one can then compute the object’s mass $`\mu `$; and from the absolute amplitudes of the waves one can then infer the distance $`r_o`$ from the system to Earth.
Poisson has estimated the accuracies with which such phase-evolution measurements can determine $`M`$, $`\eta =\mu /M`$, and $`a`$. His estimates are based on an analytic model of the signal in which (translated into our notation) $`𝒩`$ is expanded in powers of $`\mathrm{\Omega }\mathrm{\Omega }_{\mathrm{isco}}`$ and only the leading order term is kept. Poisson assumes $`M=10^6M_{}`$, $`\mu =10M_{}`$, $`a0`$ (i.e., not close to $`\pm 1`$), and a measurement time of one year. For these parameters, our Figs. 8 and 4 suggest that, for $`a0.5`$, his expansion may be accurate to within a few tens of per cent, while for $`a0.9`$ it is seriously wrong. His estimated measurement accuracies are $`\mathrm{\Delta }a0.05/\rho `$, $`\mathrm{\Delta }M/M0.002/\rho `$, and $`\delta \eta /\eta 0.06/\rho `$, where $`\rho `$ is the amplitude signal to noise ratio.
Our Tables VIII and IX for $`𝒩`$ and $`𝒯`$ (the relativistic corrections to the orbital phase evolution rate $`d\mathrm{\Phi }/d\mathrm{ln}f`$ and the time $`T`$ to the end of inspiral) can serve as the foundation a more definitive computation of the phasing-based measurement accuracies.
Information is also carried by the relative amplitudes of the waves’ harmonics. Most promising, we think, are the amplitude ratios for the first and second harmonics and for the third and second. We plot these ratios in Fig. 9, as parametric functions of the hole’s spin $`a`$ and the orbital radius $`r/r_{\mathrm{isco}}`$. From this plot it is evident that the instantaneous amplitude ratios will give both $`a`$ and the instantaneous $`r/r_{\mathrm{isco}}`$ with moderate accuracy—though only for those systems with strong enough signals that the weakest of these harmonics, $`m=1`$, stands up strongly above the noise; cf. the short-dashed curves in Figs. 37.
In our idealized case of circular, equatorial orbits, this harmonic-ratio information is not independent of that from the orbital phasing, but it could provide a confirmation of the phasing conclusions.
In the more realistic case of noncircular, nonequatorial orbits, the waveforms will be much richer and there will be many more parameters to solve for. Our survey of the circular, equatorial case gives some rough indication of the kinds of information one can extract and by what methods.
## VI Concluding Remarks
In this paper we have tabulated the results of TSN-based computations of the waves emitted by an object spiraling into a spinning, massive black hole on a slowly shrinking, circular, equatorial orbit. Our Tables IIX can serve as a foundation for future mission-definition studies for LISA—most particularly, for studies of how changes in the mission design may affect LISA’s ability to detect such inspiral waves, for studies of the accuracies with which LISA’s data can extract the properties of the source, and for explorations of possible data analysis algorithms.
Much more important, in the long run, will be the extension of our analysis to nonequatorial and noncircular orbits. This extension is urgent, since models of active galactic nuclei predict, rather firmly, that the orbits will be nonequatorial and quite noncircular, and since the earliest possible date for LISA to fly is less than ten years in the future.
## Acknowledgments
For helpful discussions we thank Fintan Ryan and Eric Poisson. For information and advice about the LISA noise curve we thank John Armstrong, Peter Bender, Curt Cutler, Frank Estabrook, Robin (Tuck) Stebbins, and Massimo Tinto; and we thank Bender and Stebbins for providing us with a table of the noise curve from Ref. . For information and advice about white-dwarf-binary background noise, we thank Peter Bender, Sterl Phinney and Tuck Stebbins. This paper was supported in part by NASA grants NAGW-4274, NAG5-6840 and their predecessors, and in view of its future applications to LIGO, by NSF grants PHY-9800111, PHY-9996213, AST-9731698 and their predecessors.
## Numerical methods
Teukolsky found that the equations describing perturbations of the Kerr spacetime could be separated into separate radial and angular equations. For the circular, equatorial orbits studied in this paper, the challenges of solving the perturbation equations are all associated with the numerical solution of the radial equation. We have used Green function methods to solve the radial equation and determine the power radiated down the horizon and to infinity by a particle in a circular equatorial orbit. The general method of solution and formulation of the problem is well described in Ref. , and we refer the interested reader there for details. In this appendix we describe several innovations that can dramatically speed the solution of the radial equation compared to the more conventional methods applied elsewhere.
The Teukolsky radial equation is a second order, ordinary differential equation. In the form given originally by Teukolsky the equation is stiff and the solution satisfying the physical boundary conditions is difficult to obtain. Sasaki and Nakamura found, through a local change of variables, a form of the radial equation which is not stiff, and we have worked with the radial equation in that form.
In the Sasaki-Nakamura formulation, the homogeneous (source-free) radial equation takes the form
$$\left[\frac{d}{dr_{}}(r_{})\frac{d}{dr_{}}𝒰(r_{})\right]X=0.$$
(55)
Here $`r_{}`$ is the so-called tortoise coordinate, which ranges from $`\mathrm{}`$ at the (outer) horizon to $`\mathrm{}`$ at spatial infinity. (In this appendix, and in this appendix only, we express all dimensioned quantities, such as $`r`$ and $`r_{}`$, in terms of the black hole’s mass $`M`$, eschewing for convenience the superscript-tilde notation used elsewhere in this paper.) The tortoise coordinate can be expressed analytically in terms of the Boyer-Lindquist radial coordinate $`r`$ and the location of the inner and outer horizons $`r_+`$ and $`r_{}`$:
$`r_{}=r+{\displaystyle \frac{2}{r_+r_{}}}\left[r_+\mathrm{ln}\left(rr_+\right)r_{}\mathrm{ln}\left(rr_{}\right)\right]`$ (57)
where
$`(rr_+)(rr_{})`$ $`=`$ $`r^22r+a^2\text{and}`$ (58)
$`r_+`$ $``$ $`r_{}.`$ (59)
The functions $``$ and $`𝒰`$ are parameterized by the angular frequency of the perturbation $`\omega =2\pi f`$, the angular momentum of the spacetime $`a`$, and the angular separation constants $`\mathrm{}`$ and $`m`$ (with $`|m|\mathrm{}`$). (For the particular forms of $``$ and $`𝒰`$ see .) For circular, equatorial orbits $`\omega `$ is always an integer multiple of the orbital angular frequency, $`\omega =\omega _mm\mathrm{\Omega }=2\pi f_m`$ .
To obtain the Green function solution to the radial equation with source we need the two solutions to the homogeneous equation corresponding to the physical boundary conditions at infinity (no in-coming radiation) and the horizon (no up-going radiation). These solutions are determined numerically by posing the boundary conditions near infinity or the horizon and integrating the radial equation inward or outward, as appropriate. In the Sasaki-Nakamura variables, obtaining a solution to the radial equation poses no particular challenge; correspondingly, it is conventional to use a “work-horse” integrator (e.g., Runge-Kutta or Bulirsch-Stoer) to solve the equation. On the other hand, the radial equation arises from a separation of variables and is parameterized by the separation constants $`\mathrm{}`$, $`m`$ and $`\omega `$, corresponding to the resolved angular and temporal dependence of the perturbation. Consequently, it is necessary to solve the radial equation separately for every important set of angular multipoles ($`\mathrm{}`$, $`m`$) and frequency $`\omega `$. For very relativistic orbits even moderate accuracy in the total radiated power may require solving the radial equation tens of thousands of times for different angular multipoles and harmonics of the orbital frequency. Consequently, speeding the solution while preserving its accuracy is of fundamental importance. In the remainder of this appendix we address several innovations we have made in solving this equation that, depending on the details of the orbit and the desired accuracy of the solution, can result in a several order of magnitude reduction in the solution time compared to a conventional approach.
### 1 Boundary conditions at the horizon
As one approaches the horizon, the physical solution for the radial function, corresponding to down-going radiation, leads to the boundary conditions used for the numerical integration of one of the homogeneous solutions of the Sasaki-Nakamura equation:
$`\underset{r_{}\mathrm{}}{lim}X_H(r_{})`$ $`=`$ $`e^{i\omega _{}r_{}},`$ (61)
$`\underset{r_{}\mathrm{}}{lim}{\displaystyle \frac{dX_H}{dr_{}}}(r_{})`$ $`=`$ $`i\omega _{}e^{i\omega _{}r_{}},`$ (62)
where
$`\omega _{}`$ $`=`$ $`\omega {\displaystyle \frac{am}{2r_+}}`$ (63)
and $`r_+`$ is the radius of the outer horizon in Boyer-Lindquist coordinates.
As a practical matter the boundary conditions used to determine $`X_H`$ are posed at some large, negative but finite $`r_{}`$, say $`R_{}^{}`$; i.e., “close to”, but not at, the horizon. Using Eqs. (1) evaluated at finite $`R_{}^{}`$ for the boundary conditions introduces errors of fractional order $`\delta =rr_+`$ into the solution. This error can be represented as an error in the amplitude of the power radiated down the horizon and the introduction of some small component of radiation up-going from the horizon. These errors propagate to large $`r_{}`$ where they contribute to the out-going radiation and lead to errors in the calculated power radiated to infinity by the orbiting particle.
The errors introduced by using equations 1 when posing boundary conditions at finite radius can be expressed as a power series in $`\delta `$. The coefficients of that expansion can be estimated by solving the equations several times, for different $`R_{}^{}`$, and using Richardson extrapolation. To estimate the first $`N`$ terms in the error expansion requires $`N+1`$ numerical solutions of the equations, each beginning with the boundary conditions posed at a different $`R_{}^{}`$. Controlling the error requires that the radial equation be solved at least twice and often three or more times at different, large $`|R_{}^{}|`$.
To improve the convergence rate of this error estimate and allow us to pose our boundary conditions at smaller $`|R_{}^{}|`$ we have solved the Sasaki-Nakamura equation analytically about the point at $`r_{}=\mathrm{}`$, finding the first corrections in $`\delta `$ to the boundary equations given by Eqs. (1). The improved boundary conditions are given by
$`X(r_{})`$ $`=`$ $`1+\delta A^{},`$ (65)
$`{\displaystyle \frac{dX}{dr_{}}}(r_{})`$ $`=`$ $`\left[i\omega _{}\left(1+\delta A^{}\right)+{\displaystyle \frac{\delta d}{2r_+}}A^{}\right],`$ (66)
where
$`A^{}`$ $`=`$ $`\left[^{(1)}+{\displaystyle \frac{d}{4r_+^2}}\left(U_2^{(1)}+F_1^{(1)}+G_1^{(1)}\right)\right]`$ (68)
$`\times \left[{\displaystyle \frac{d^2}{4r_+^2}}+iϵ{\displaystyle \frac{d}{r_+}}\omega _{}\right]^1,`$
$`\delta `$ $``$ $`rr_+,`$ (69)
$`d`$ $``$ $`r_+r_{},`$ (70)
$`^{(1)}`$ $``$ $`i\omega _{}{\displaystyle \frac{d}{2r_+}},`$ (71)
$`U_2^{(1)}`$ $``$ $`\lambda +{\displaystyle \frac{4r_{}}{r_+}}2r_+\omega _{}\left({\displaystyle \frac{2(2+d)}{d}}{\displaystyle \frac{2r_+}{d^2}}\omega _{}\right)`$ (73)
$`\left(1{\displaystyle \frac{d(2+d)}{r_+}}\right){\displaystyle \frac{4r_+^2}{d^2}}\omega _{}^2,`$
$`F^{(1)}`$ $``$ $`2ir_+\omega _{}F(0),`$ (74)
$`G_1^{(1)}`$ $``$ $`2+d/2,`$ (75)
$`F(0)`$ $``$ $`\gamma ^{}(r_+)/\gamma (r_+),`$ (76)
$`\gamma (r)`$ $``$ $`{\displaystyle \underset{k=0}{\overset{4}{}}}\gamma _kr^k,`$ (77)
$`\gamma _0`$ $``$ $`\lambda \left(\lambda +2\right)12a\omega \left(a\omega m\right)12i\omega ,`$ (78)
$`\gamma _1`$ $``$ $`8i\left[3a^2\omega a\lambda \left(a\omega m\right)\right],`$ (79)
$`\gamma _2`$ $``$ $`12\left[2ai\left(a\omega m\right)+a^22a^2\left(a\omega m\right)^2\right],`$ (80)
$`\gamma _3`$ $``$ $`24a^2\left[1+ia\left(a\omega m\right)\right],`$ (81)
$`\gamma _4`$ $``$ $`12a^4,`$ (82)
$`\lambda `$ $``$ $`\mathrm{}(\mathrm{}+1)2am\omega +a^2\omega ^2+2.`$ (83)
The numerical solution to the radial equation using these improved boundary conditions converges upon the true solution more quickly than a solution using the boundary conditions (1). We are thus able to pose approximate horizon boundary conditions at smaller $`|R_{}^{}|`$, reducing the domain over which we must integrate the radial equation and, often the number of times we must integrate the equation for each $`(\omega ,\mathrm{},m)`$ in order to obtain a solution of controlled accuracy.
### 2 Boundary conditions at spatial infinity
As $`r_{}\mathrm{}`$, the physical solution for the radial function, corresponding to no in-going radiation, leads to the boundary conditions for the numerical integration of the other critical solution of the radial equation:
$`\underset{r_{}\mathrm{}}{lim}X_{\mathrm{}}`$ $`=`$ $`1,`$ (84)
$`\underset{r_{}\mathrm{}}{lim}{\displaystyle \frac{dX_{\mathrm{}}}{dr_{}}}`$ $`=`$ $`i\omega .`$ (85)
As with the boundary conditions at the horizon, we construct the solution $`X_{\mathrm{}}`$ beginning with boundary conditions posed at finite $`R_+^{}`$, not at infinity. Using the asymptotic form of the boundary conditions to set $`X`$ and $`X^{}`$ at finite radius leads to errors of fractional order $`1/R_+^{}`$ in the solution, which can be represented as an error in the amplitude of the out-going radiation and the introduction of some small in-going radiation component. These lead, in turn, to errors in the estimated power radiated to infinity and down the horizon. We can use Richardson extrapolation to estimate and reduce this error; however, as before, the radiated power must be determined at several different large $`R_+^{}`$ in order to estimate and reduce the error.
To permit a more accurate estimate of the radiated power from $`X`$ and $`X^{}`$ evaluated at smaller $`R_+^{}`$ we have solved the Sasaki-Nakamura equation analytically about the point at $`r_{}=\mathrm{}`$, finding the first corrections in $`1/R_+^{}`$ to the asymptotic form of the radial function $`X`$. For finite $`R_+^{}`$ we have
$`X_{\mathrm{}}(r_{})`$ $`=`$ $`\left(1+{\displaystyle \frac{a_1}{r}}\right),`$ (87)
$`{\displaystyle \frac{dX_{\mathrm{}}}{dr_{}}}(r_{})`$ $`=`$ $`i\omega \left(1+{\displaystyle \frac{a_1}{r}}\right),`$ (88)
where
$`a_1`$ $`=`$ $`{\displaystyle \frac{\gamma _1}{\gamma _0}}+{\displaystyle \frac{i}{2}}\left[\omega \left(a^2+4\right)+2am+{\displaystyle \frac{\lambda +2}{\omega }}\right].`$ (89)
We use these expressions, evaluated at finite but large $`R_+^{}`$, to set the boundary condition for the numerical solution of the homogeneous radial equation. We continue to use Richardson extrapolation to control the error of the solutions; however, each step in the extrapolation has a greater effect on the error and the extrapolation can take place at smaller $`R_+^{}`$.
### 3 A more suitable choice of variables
The solution $`X`$ to the Sasaki-Nakamura equations is a complex oscillatory function. Integrating the equations directly for $`X`$ requires a spatial resolution $`\mathrm{\Delta }r_{}`$ less than the local wavelength of $`X`$,
$$\mathrm{\Delta }r_{}\left|\frac{d\mathrm{ln}X}{dr_{}}\right|^1.$$
(90)
When solving for the radial function corresponding to a high temporal frequency $`|\omega |`$ the step-size can become quite small, with a corresponding increase in the computational time for an accurate solution.
It is advantageous in circumstances like these to reformulate the problem in action-angle variables, whose variation is both slower and smoother than the variations in $`X`$. Writing $`X`$ as
$$X\mathrm{exp}\left[i\mathrm{\Phi }(r_{})\right]$$
(92)
we define the two real functions $`\xi `$ and $`\varphi `$ as the imaginary and real parts of $`\mathrm{\Phi }`$:
$`\xi `$ $``$ $`\mathrm{}\left(\mathrm{\Phi }\right)`$ (93)
$`\varphi `$ $``$ $`\mathrm{}\left(\mathrm{\Phi }\right)`$ (94)
With this substitution the linear Sasaki-Nakamura equation for complex $`X`$ becomes a pair of coupled non-linear equations for the real $`\xi `$ and $`\varphi `$. The equation for $`\xi `$ is second order while the equation for $`\varphi `$ can be integrated immediately to obtain a first order equation. (This is expected since the solution for $`X`$ is determined only up to an overall phase.) Both $`\varphi `$ and $`\xi `$ vary slowly and smoothly compared to $`X`$. This is particularly true as one moves toward either the horizon or spatial infinity, where $`X`$ is oscillatory in $`r_{}`$ while $`\xi `$ is constant and $`\varphi `$ is linear. Correspondingly, the numerical solution of the equations for $`\varphi `$ and $`\xi `$ require much less resolution for the same numerical accuracy, dramatically speeding the integration of the radial equation.
### 4 Numerical solution of the equations for $`\xi `$ and $`\varphi `$
The local errors committed by, e.g., a fourth order Runge-Kutta integration of the radial equation are proportional to $`\mathrm{\Delta }r_{}^{}{}_{}{}^{5}`$. Reducing the step-size and increasing the number of integration steps will decrease the overall solution error algebraically, i.e., as a fixed power of $`\mathrm{\Delta }r_{}`$, while increasing the time required for a solution. A higher order computational method will increase the solution accuracy more rapidly. Exponential convergence of the solution with $`\mathrm{\Delta }r_{}`$ can be obtained if the equations are solved via collocation pseudo-spectral techniques . In a collocation pseudo-spectral method the solution for the dependent variable is approximated as a sum over a suitable set of basis functions. The differential equations, evaluated on the approximate solution at a fixed number of points, then determine the coefficients in the expansion. For problems with smooth solutions the solution accuracy increases exponentially with the number of terms in the approximation (and, correspondingly, with the number of evaluations of the differential equation, which is the analog of the spatial resolution of the integration). Our final innovation is to solve the radial equation using pseudo-spectral techniques. We have chosen a Chebyshev expansion for $`\xi `$ and $`\varphi `$ with Gauss-Lobatto collocation points. Our experience is that the best performance is obtain if the integration domain $`[R_+,R_{}]`$ is divided into two parts, at approximately the peak of the effective potential $`R_0`$: i.e., we use two expansions for $`\varphi `$ and $`\xi `$, one in the domain $`[R_+,R_0]`$ and the other in the domain $`[R_0,R_{}]`$. At $`R_0`$ we insist that the two solutions for $`\varphi `$ and $`\xi `$ agree in value, and that the solutions for $`\xi `$ agree also in their first derivative, as is appropriate for functions described by first order and second order differential equations, respectively.
|
warning/0007/cond-mat0007476.html
|
ar5iv
|
text
|
# Irrelevant Interactions without Composite Operators - A Remark on the Universality of second order Phase Transitions
## 1 Introduction
One of the great achievements of theoretical physics in the 70’s was the unification of concepts and ideas from quantum field theory and statistical mechanics through the Wilson renormalization group \[WiKo\]. In particular renormalized perturbation theory was applied successfully to the study of second order phase transitions and to the calculation of critical exponents \[BGZ, Amit, ZJ\]. One of the challenging conceptual problems was the question of universality, i.e. to realize why large classes of theories, specified essentially by the respective Hamiltonians should give rise to the same critical behaviour characterized through the critical exponents. Experimentally those depend only on dimensionality and symmetry but not on details of the dynamics. Modifications of the Hamiltonians thus should lead only to subleading corrections. We restrict our explicit presentation to one of the simplest and bestknown classes, that of the Ising model. The method is general, however. Passing to the continuous description which should be viable for correlation lengths $`\xi `$ much larger than the lattice spacing, i.e. in the vicinity of the critical point, the symmetry class of the Ising model is presented by $`\varphi ^4`$ theory, symmetric under $`\varphi \varphi `$. The standard action at the scale of the UV cutoff $`\mathrm{\Lambda }_0`$, corresponding to the inverse lattice spacing in position space, is then
$$L^0=\left(a\varphi ^2(x)+b(_\mu \varphi )^2(x)+c\varphi ^4(x)\right)d^4x.$$
(1)
If we restrict ourselves to perturbation theory the constants $`a,b,c`$ are to be viewed as power series in the renormalized coupling $`g`$ or in $`\mathrm{}`$. In the standard notation this expression is rewritten as
$$L^0=\left(\frac{Z}{2}(_\mu \varphi )^2(x)+\frac{Z}{2}(m^2+\delta m^2)\varphi ^2(x)+\frac{g_0}{4!}Z^2\varphi ^4(x)\right)d^4x,$$
(2)
where $`L^0`$ also includes the term of order 0 in perturbation theory, that is to say
$$L^0=L^0+\left(\frac{1}{2}(_\mu \varphi )^2(x)+\frac{1}{2}m^2\varphi ^2(x)\right)d^4x.$$
(3)
The field $`\varphi `$ corresponds to the renormalized field. In (2) we introduced the standard notation for the wave function renormalization $`Z`$ and the mass counterterm $`\delta m`$ as well as the bare coupling $`g_0`$. We restrict to the four dimensional theory which also serves to study lower dimensional theories through the $`\epsilon `$-expansion.<sup>2</sup><sup>2</sup>2Dimensional regularization cannot be naturally accommodated in the flow equation framework. Still the associated minimal renormalization schemes should be implementable. This has been shown for analytic regularization \[KoSm\]. Starting from the Ising model Hamiltonian on a cubic lattice one arrives at the action (1) on performing block spin transformations, expanding in local terms and passing to the continuous limit, on neglecting all irrelevant terms, i.e. those of mass dimension larger than 4. The aim of this paper is to show that this is justified indeed when analysing long distance phenomena near the critical point. This means that the dominant contributions to the correlation functions near the critical point are obtained from (1) for suitable choices of $`a,b,c`$. More precisely we will add a finite sum
$$A(\varphi )=\underset{n=1}{\overset{N}{}}\frac{Z^{2+n}}{\mathrm{\Lambda }_0^{2n}}\frac{g_{4+2n}}{(4+2n)!}\varphi ^{4+2n}(x)d^4x$$
(4)
to (1). Here the UV cutoff $`\mathrm{\Lambda }_0`$ appears naturally when expanding in local terms, by dimensional analysis. This means that the couplings $`g_{4+2n}`$ are dimensionless (in $`d=4`$).<sup>3</sup><sup>3</sup>3We choose conventions such that a factor of $`Z^{2+n}`$ appears in front of $`g_{4+2n}`$, which will somewhat simplify the notation later. Note that $`Z`$ will depend on the couplings $`g_{4+2n}`$. In particular it will also be different from $`1`$ for $`g=0`$, if some of the $`g_{4+2n}`$ do not vanish. Since the statements of renormalization theory are generally of perturbative nature, i.e. valid on formal expansion in the couplings $`g`$, they require small values of those to be reliable. When including (4) the question arises how the size of the irrelevant couplings compares to that of the original $`\varphi ^4`$ coupling $`g`$. Here of course different situations may arise and can be analysed. Later on we will regard the situation where they are chosen such that the loop expansion remains valid, which means generally that
$$g_{4+2n}g^{n+1}.$$
(5)
The expansion with respect to local terms also produces higher dimensional terms of the form $`(\varphi ^n^w\varphi ^m)(x)`$, which contain $`|w|`$ derivatives with respect to the coordinates $`x`$. Starting from a cubic lattice only terms respecting rotational symmetry, i.e. invariant under the Euclidean group, should appear in the continuum limit, i.e. when approaching the critical region. Furthermore in (4) only terms invariant under $`\varphi \varphi `$ are generated if $`\mathrm{𝖹𝖹}_2`$-symmetry is unbroken. For shortness of notation we restrict to (4), inclusion of derivative terms would only lead to minor changes. In the explicit treatment we will even limit ourselves to a single insertion $`g_6\varphi ^6`$, for simplicity of notation.
The effects of irrelevant terms have of course been studied extensively in the literature \[Weg, BGZ\] and can be found in textbooks, e.g. \[ZJ, Ch.26\]. In the field theory approach these terms were analysed by renormalization theory for composite operators, as it existed in the early seventies \[Zim\]. Treating e.g. $`\varphi ^6`$ as a composite operator insertion means that one restricts to Green functions carrying at most one insertion of this $`\varphi ^6`$ term.<sup>4</sup><sup>4</sup>4 It is possible to go beyond one insertion. But then the number of renormalization conditions one has to fix increases with the number of insertions, corresponding to the fact that a $`\varphi ^6`$ theory is nonrenormalizable. Then one has to fix renormalization conditions for the inserted Green functions up to dimension six (thus on the two, four and six point functions and on derivatives of the two and four point functions). The general and probably optimal bound on the coefficient of the term $`\varphi ^6`$ in $`L^0`$ is then of the form $`𝒫(\mathrm{log}\mathrm{\Lambda }_0)`$, i.e. a polynomial in logarithms of $`\mathrm{\Lambda }_0`$ -as has been shown e.g. in \[KeKo1\]- and not $`\mathrm{\Lambda }_0^2𝒫(\mathrm{log}\mathrm{\Lambda }_0)`$ as in (4). Otherwise stated this means that in general it is not possible to find out the renormalization conditions which would give a bound $`\mathrm{\Lambda }_0^2`$, since the associated dynamical system is unstable. From the physical point of view it seems therefore perferable to start directly from the modification of the bare action as in (4), and to perform the renormalization for this theory. We note however that it is not really possible to study the question in such a way that the only change in the bare action consists in adding the term $`\varphi ^6`$ to it. The counterterms $`\varphi ^4`$ and $`\varphi ^2`$ change at the same time if we keep the renormalization conditions fixed. This phenomenon corresponds to what is called operator mixing in the theory of composite operators. However, whereas these renormalization conditions generally are related to the physical parametrization of the theory near the critical region and thus accessible to experiment, this seems not to be the case for the Green functions carrying e.g. $`\varphi ^6`$-insertions. Our results are such that we may study an arbitrary number of irrelevant insertions for a fixed set of renormalization conditions. Thus the study of these insertions is generalized and simplified at the same time as compared to composite operator theory.
Renormalizability proofs based on the renormalization group are conceptually simple and rigorous and give a transparent view on the universality of critical behaviour, in showing that the modification of the action by irrelevant terms does not influence (the dominant part of) the critical behaviour. In this note we would like to make this explicit for the simplest case, the universality class of the Ising model. In the next section we will present the required results on the renormalizability of the theories with and without $`\varphi ^6`$-insertion, in particular in the critical region. In the last section we also use the (standard) renormalization group equations to analyse the subdominant contributions of the irrelevant insertion to the critical behaviour.
## 2 Renormalization of $`\varphi _4^4`$ theory with irrelevant terms
Renormalization theory as we are going to use it here is based on the flow equations of the renormalization group due to Wilson \[WiKo\], and particularly to Polchinski \[Pol\] as regards the application to the perturbative renormalization problem. The flow equation is obtained by successively integrating out momenta in the (regularized) theory starting from the UV cutoff $`\mathrm{\Lambda }_0`$ down to the scale $`\mathrm{\Lambda }<\mathrm{\Lambda }_0`$. The final renormalized theory is obtained on taking the limits $`\mathrm{\Lambda }_0\mathrm{}`$ and $`\mathrm{\Lambda }0`$. Its differential form can be obtained when deriving with respect to $`\mathrm{\Lambda }`$ the generating functional of the connected (free propagator) amputated Green functions (CAG) of the theory with momenta restricted to lie between $`\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }_0`$. The scales $`\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }_0`$ enter through the regularized propagator, which for the massive theory takes the form
$$C^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p)=\frac{1}{p^2+m^2}\{e^{\frac{p^2+m^2}{\mathrm{\Lambda }_0^2}}e^{\frac{p^2+m^2}{\mathrm{\Lambda }^2}}\}.$$
(6)
Its Fourier transform is
$$\widehat{C}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(x)=_pC^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p)e^{ipx}.$$
(7)
$$\text{ We use the conventions: }_p:=_{\mathrm{I}\mathrm{R}^4}\frac{d^4p}{(2\pi )^4},\varphi (x)=_p\phi (p)e^{ipx},\frac{\delta }{\delta \varphi (x)}=(2\pi )^4_p\frac{\delta }{\delta \phi (p)}e^{ipx}.$$
For finite $`\mathrm{\Lambda }_0`$ and in finite volume the theory can be given rigorous meaning starting from the functional integral
$$e^{\frac{1}{\mathrm{}}(L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\varphi )+I^{\mathrm{\Lambda },\mathrm{\Lambda }_0})}=𝑑\mu _{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\mathrm{\Phi })e^{\frac{1}{\mathrm{}}L^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}(\varphi +\mathrm{\Phi })},$$
(8)
where the factors of $`\mathrm{}`$ have been introduced to allow for a consistent loop expansion in the sequel which permits us to stay with a single expansion parameter in the presence of two coupling constants. In (8) $`d\mu _{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\mathrm{\Phi })`$ denotes the (translation invariant) Gaussian measure with covariance $`\mathrm{}\widehat{C}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(x)`$. The normalization factor $`e^{\frac{1}{\mathrm{}}I^{\mathrm{\Lambda },\mathrm{\Lambda }_0}}`$ is due to vacuum contributions. It diverges in infinite volume so that we can take the infinite volume limit only when it has been eliminated. We do not make the finite volume explicit here since it plays no role in the sequel. One may convince oneself that $`L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\varphi )`$ is equal to
$$L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\varphi )=\mathrm{ln}Z^{\mathrm{\Lambda },\mathrm{\Lambda }_0}((\widehat{C}^{\mathrm{\Lambda },\mathrm{\Lambda }_0})^1\varphi )+\mathrm{\hspace{0.17em}1}/2\varphi ,(\widehat{C}^{\mathrm{\Lambda },\mathrm{\Lambda }_0})^1\varphi .$$
(9)
Here $`Z^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(j)`$ is the (standard notation for the) generating functional of the Green functions of the (regularized) theory. By $`,`$ we denote the scalar product in $`L_2(\mathrm{I}\mathrm{R}^4,d^4x)`$ so that the second term contains the 0-loop two-point function. Thus $`L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\varphi )`$ generates the CAG, apart from the order zero contribution given by the inverted free propagator. The functional $`L^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}(\varphi )=L^0(\varphi )`$ is the bare action including counterterms, to be calculated from the renormalization conditions. On adding the 0-loop two-point function and including the $`\varphi ^6`$-insertion it takes the form (see (1, 2))
$$L^0=\left(\frac{Z}{2}(_\mu \varphi )^2(x)+\frac{Z}{2}(m^2+\delta m^2)\varphi ^2(x)+\frac{g_0}{4!}Z^2\varphi ^4(x)+\frac{Z^3}{\mathrm{\Lambda }_0^2}\frac{g_6}{6!}\varphi ^6(x)\right)d^4x.$$
(10)
Here $`Z,\delta m^2`$ and $`g_0`$ are formal power series in $`\mathrm{}`$. The Wilson flow equation (FE) is is a differential equation for the functional $`L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$, obtained from (8) on differentiating w.r.t. $`\mathrm{\Lambda }`$ :
$$_\mathrm{\Lambda }(L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}+I^{\mathrm{\Lambda },\mathrm{\Lambda }_0})=\frac{\mathrm{}}{2}\frac{\delta }{\delta \varphi },(_\mathrm{\Lambda }\widehat{C}^{\mathrm{\Lambda },\mathrm{\Lambda }_0})\frac{\delta }{\delta \varphi }L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}\frac{1}{2}\frac{\delta }{\delta \varphi }L^{\mathrm{\Lambda },\mathrm{\Lambda }_0},(_\mathrm{\Lambda }\widehat{C}^{\mathrm{\Lambda },\mathrm{\Lambda }_0})\frac{\delta }{\delta \varphi }L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}.$$
(11)
Changing to momentum space and expanding in a formal powers series w.r.t. $`\mathrm{}`$ we write (with slight abuse of notation)
$$L^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\phi )=\underset{l=0}{\overset{\mathrm{}}{}}\mathrm{}^lL_l^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\phi ).$$
(12)
From $`L_l^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\phi )`$ we then obtain the CAG of loop order $`l`$ in momentum space as <sup>6</sup><sup>6</sup>6The normalization of the $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$ is defined differently from earlier references.
$$(2\pi )^{4(n1)}\delta _{\phi (p_1)}\mathrm{}\delta _{\phi (p_n)}L_l^{\mathrm{\Lambda },\mathrm{\Lambda }_0}|_{\phi 0}=\delta ^{(4)}(p_1+\mathrm{}+p_n)_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p_1,\mathrm{},p_{n1}),$$
(13)
where we have written $`\delta _{\phi (p)}=\delta /\delta \phi (p)`$. Note again that our definition of the $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$ is such that $`_{0,2}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$ vanishes. This is important for the set-up of the inductive scheme, through which perturbative renormalizability will be established. The FE (11) rewritten in terms of the CAG (13) takes the following form
$$_\mathrm{\Lambda }^w_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p_1,\mathrm{},p_{n1})=\frac{1}{2}_k(_\mathrm{\Lambda }C^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(k))^w_{l1,n+2}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(k,k,p_1,\mathrm{},p_{n1})$$
(14)
$$\underset{\genfrac{}{}{0pt}{}{l_1+l_2=l,w_1+w_2+w_3=w}{n_1+n_2=n}}{}\frac{1}{2}\left[^{w_1}_{l_1,n_1+1}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p_1,\mathrm{},p_{n_1})(^{w_3}_\mathrm{\Lambda }C^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p^{}))^{w_2}_{l_2,n_2+1}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p_{n_1+1},\mathrm{},p_n)\right]_{ssym},$$
$$\text{where }p^{}=p_1\mathrm{}p_{n_1}=p_{n_1+1}+\mathrm{}+p_n.$$
Here we have written (14) directly in a form where also momentum derivatives of the CAG (13) are performed, and we used the shorthand notation
$$^w:=\underset{i=1}{\overset{n1}{}}\underset{\mu =0}{\overset{3}{}}(\frac{}{p_{i,\mu }})^{w_{i,\mu }}\text{ with }w=(w_{1,0},\mathrm{},w_{n1,3}),|w|=w_{i,\mu }\mathrm{I}\mathrm{N}_0.$$
(15)
The symbol $`ssym`$ (as defined in \[KMR\]) means summation over those permutations of the momenta $`p_1,\mathrm{},p_n`$, which do not leave invariant the subsets $`\{p_1,\mathrm{},p_{n_1}\}`$ and $`\{p_{n_1+1},\mathrm{},p_n\}`$. Note that the CAG are symmetric in their momentum arguments by definition. A simple inductive proof of the renormalizability of $`\varphi _4^4`$ theory has been exposed several times in the literature \[KKS, KeKo1, Kop\], and we will not repeat it in detail. The line of reasoning can be resumed as follows. The induction hypotheses to be proven are :
A) Boundedness
$$|^w_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\stackrel{}{p})|(\mathrm{\Lambda }+m)^{4n|w|}𝒫(log\frac{\mathrm{\Lambda }+m}{m})𝒫(\frac{|\stackrel{}{p}|}{\mathrm{\Lambda }+m}).$$
(16)
B) Convergence
$$|_{\mathrm{\Lambda }_0}^w_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\stackrel{}{p})|\frac{1}{\mathrm{\Lambda }_0^3}𝒫(log\frac{\mathrm{\Lambda }_0}{m})(\mathrm{\Lambda }+m)^{6n|w|}𝒫(\frac{|\stackrel{}{p}|}{\mathrm{\Lambda }+m}).$$
(17)
Here and in the following the $`𝒫`$ denote (each time they appear possibly new) polynomials with nonnegative coefficients. The coefficients depend on $`l,n,|w|,m`$, but not on $`\stackrel{}{p},\mathrm{\Lambda },\mathrm{\Lambda }_0`$. We used the shorthand $`\stackrel{}{p}=(p_1,\mathrm{},p_{n1})`$ and $`|\stackrel{}{p}|=sup\{|p_1|,\mathrm{},|p_n|\}`$. The statement (17) implies renormalizability : It proves the limits $`lim_{\mathrm{\Lambda }0,\mathrm{\Lambda }_0\mathrm{}}`$ $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\stackrel{}{p})`$ to exist to all loop orders $`l`$. But the statement (16) has to be obtained first to prove (17). To prove (16) we use an inductive scheme that proceeds upwards in $`2l+n`$, for given $`2l+n`$ upwards in $`l`$, and for given $`(l,n)`$ downwards in $`|w|`$, starting from some arbitrary $`|w_{max}|3`$. The important point to note is that the terms on the r.h.s. of the FE (14) always are prior to the ones on the l.h.s. in the inductive order. So the bound (16) may be used as an induction hypothesis on the r.h.s. Besides we also need a bound on the propagator and its momentum derivatives : It is easy to prove that
$$|^w_\mathrm{\Lambda }C^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p)|\mathrm{\Lambda }^{3|w|}𝒫(|p|/\mathrm{\Lambda })e^{\frac{p^2+m^2}{\mathrm{\Lambda }^2}}.$$
(18)
Equipped with this bound and the induction scheme, we may then integrate the FE, where terms with $`n+|w|5`$ are integrated down from $`\mathrm{\Lambda }_0`$ to $`\mathrm{\Lambda }`$, since for those terms we have the boundary conditions following from (10)<sup>7</sup><sup>7</sup>7Strictly speaking the boundary condition for the $`_{l,6}^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}`$ has to be chosen more generally first : $`_{l,6}^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}g_6\frac{1}{\mathrm{\Lambda }_0^2}𝒫(log\frac{\mathrm{\Lambda }_0}{m})`$. Once the bounds on $`_{l,2}^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}`$ have been established (which are independent of $`_{l,6}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$ at the same loop order), we specialize to $`Z^3`$ knowing that $`Z^3`$ is bounded by $`𝒫(log\frac{\mathrm{\Lambda }_0}{m})`$.
$$^w_{l,n}^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}0\text{ if }n+|w|>4\text{ and }n6,\text{ and }_{l,6}^{\mathrm{\Lambda }_0,\mathrm{\Lambda }_0}g_6\frac{1}{\mathrm{\Lambda }_0^2}Z_l^3.$$
(19)
The relevant terms (those with $`n+|w|4`$) are integrated upwards from $`0`$ to $`\mathrm{\Lambda }`$. The boundary conditions for these terms are the renormalization conditions we impose and which fix the counterterms $`Z,\delta m^2,g_0`$. We may fix for example
$$_{l,2}^{0,\mathrm{\Lambda }_0}(0)=0,_{p^2}_{l,2}^{0,\mathrm{\Lambda }_0}(0)=0,_{l,4}^{0,\mathrm{\Lambda }_0}(0)=g\delta _{l,0}.$$
(20)
To go away from the renormalization point (here chosen at zero momentum) we may use the Schlömilch or integrated Taylor formula which takes us back to the irrelevant situation.
The bound (17) holds for the $`\mathrm{𝖹𝖹}_2`$-symmetric theory only, it is sharper than the one from \[KKS,KeKo1,Kop\] (but generally assumed true in the literature). Its proof is based on the same inductive scheme as that for (16). We start from the FE (14) with $`|w|`$ momentum derivatives applied on it, integrate over $`\mathrm{\Lambda }`$ and derive w.r.t. $`\mathrm{\Lambda }_0`$. For the terms on the r.h.s., on which this derivative does not apply, we can use the bound (16). For the terms derived w.r.t. $`\mathrm{\Lambda }_0`$ we can use (17), applying our induction scheme. The best bound we can arrive at is essentially saturated by the boundary terms, we find.
We first regard the case of the irrelevant terms with $`n+|w|5`$, and here we start looking at the case $`n+|w|6`$. We have the equation (in shorthand notation)
$$_{\mathrm{\Lambda }_0}^w_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}=^w(\text{r.h.s. of the FE})|_{\mathrm{\Lambda }=\mathrm{\Lambda }_0}+_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}𝑑\mathrm{\Lambda }^{}_{\mathrm{\Lambda }_0}^w(\text{r.h.s. of the FE})(\mathrm{\Lambda }^{}).$$
(21)
Now using (16, 17) to bound the r.h.s. of (21) we verify (17) on performing the integral. In the case $`n+|w|=5`$ it suffices to regard 0 external momentum (referring again to the Schlömilch formula for deviations from 0, which takes us back to $`n+|w|=6`$). In both cases $`n=4,|w|=1`$ and $`n=2,|w|=3`$ the first term of the r.h.s. of (21) is 0 due to our boundary conditions, whereas the second vanishes due to euclidean invariance.
Terms with $`n+|w|4`$ have to be analysed at the renormalization point indicated in (20), and they are integrated from 0 to $`\mathrm{\Lambda }`$ :
$$_{\mathrm{\Lambda }_0}^w_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}=_0^\mathrm{\Lambda }𝑑\mathrm{\Lambda }^{}_{\mathrm{\Lambda }_0}^w(\text{ r.h.s. of the FE })(\mathrm{\Lambda }^{}).$$
(22)
Only the second term from (21) appears on the r.h.s. because the renormalization conditions are $`\mathrm{\Lambda }_0`$-independent. In this term we may factorize $`\mathrm{\Lambda }_0^3𝒫(log\mathrm{\Lambda }_0)`$ and verify (17) by induction. As a result of these considerations we obviously obtain the same bounds for the theory with an insertion of $`\varphi ^6`$ as for the theory with $`g_6=0`$.
To study the theory in the critical domain, and particularly the role of the irrelevant insertions in this domain, we have to analyse the correlation functions in the limit of large correlation length, i.e. in the language of field theory, (the approach to) the massless theory. Thus we regard the propagator
$$C^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p)=\frac{1}{p^2}(e^{\frac{p^2}{\mathrm{\Lambda }_0^2}}e^{\frac{p^2}{\mathrm{\Lambda }^2}}),(\mathrm{\Lambda },p)(0,0),$$
(23)
which is singular for $`sup(\mathrm{\Lambda }^2,p^2)0`$. However it has a finite limit for $`p^20`$, if $`\mathrm{\Lambda }`$ stays bounded from below so that (18) stays valid for $`\mathrm{\Lambda }m`$. Only the case $`\mathrm{\Lambda }<m`$ has to be reconsidered.
For $`\mathrm{\Lambda }0`$ in fact the CAG may become singular at certain exceptional momentum configurations, i.e. where subsums of external momenta vanish. But first, for the massless theory to exist at all, certain restrictions on the renormalization conditions have to be observed. More specifically the renormalization points have to be chosen as follows :
$$_{l,2}^{0,\mathrm{\Lambda }_0}(0)=0,(_{p_\mu }_{p_\nu }_{l,2}^{0,\mathrm{\Lambda }_0})(p=k)|_{\delta _{\mu ,\nu }}=0,_{l,4}^{0,\mathrm{\Lambda }_0}(k_1,k_2,k_3)=g\delta _{l,0}.$$
(24)
This means the mass renormalization has to be performed at 0 momentum, whereas the wave function renormalization and coupling constant renormalization have to be performed at nonexceptional external momenta, i.e. $`k^2=\mu ^20`$ and no subsum in $`k_1,k_2,k_3,k_4`$ vanishes. Since we have defined the $``$ to be symmetric functions of their arguments it is natural to make a symmetric<sup>8</sup><sup>8</sup>8But it is not necessary because the solutions of the FE come out symmetric by construction. choice, e.g. $`\stackrel{}{k_i}\stackrel{}{k_j}=\frac{\mu ^2}{4}(4\delta _{ij}1)`$ for a fixed nonvanishing momentum scale $`\mu `$. In (24) $`(_{p_\mu }_{p_\nu }_{l,2}^{0,\mathrm{\Lambda }_0})(p=k)|_{\delta _{\mu ,\nu }}`$ refers to the decomposition of the O(4) invariant tensor
$$_{p_\mu }_{p_\nu }_{l,2}^{0,\mathrm{\Lambda }_0}(p)|_{p=k}=A(\mu ^2)\delta _{\mu ,\nu }+B(\mu ^2)k_\mu k_\nu ,$$
and we have defined
$$_{p_\mu }_{p_\nu }_{l,2}^{0,\mathrm{\Lambda }_0}(p=k)|_{\delta _{\mu ,\nu }}=A(\mu ^2)$$
(25)
so that the renormalization condition implies $`A(\mu ^2)=\mathrm{\hspace{0.17em}0}`$. Note that $`B(\mu ^2)`$ is irrelevant and need not be fixed by a renormalization condition. Obviously renormalization of the massless theory introduces a new mass scale which is generally called $`\mu `$. The problem of exceptional momentum configurations can be studied in full generality and rigour with flow equations \[KeKo2\] : It is possible to define an IR index $`\gamma `$ with $`2\gamma \mathrm{I}\mathrm{N}`$, which measures the exceptionality of the momentum configuration $`P=(p_1,\mathrm{},p_n)`$. Using the shorthand notations $`=^{0,\mathrm{}}`$ and $`^\mathrm{\Lambda }=^{\mathrm{\Lambda },\mathrm{}}`$, we may phrase as follows the results from the renormalization proof \[KeKo2\] for the massless symmetric $`\varphi _4^4`$ theory :
a) The n-point CAG with for $`n>2`$ are smooth functions of the external momenta in the (open) subspace of (arbitrarily) bounded nonexceptional momentum configurations. We have
$$^w_{l,n}(p_1,\mathrm{},p_{n1})=\underset{\mathrm{\Lambda }0}{lim}^w_{l,n}^\mathrm{\Lambda }(p_1,\mathrm{},p_{n1}).$$
(26)
b) Generally one has
$$|^w_{l,n}^\mathrm{\Lambda }(p_1,\mathrm{},p_{n1})|\mu ^{4n|w|}\left(\frac{\mu }{\mathrm{\Lambda }}\right)^{2\gamma (P)+|w|}𝒫(log\frac{\mu }{\mathrm{\Lambda }}),0<\mathrm{\Lambda }\mu .$$
(27)
For the two point function at $`\mathrm{\Lambda }=0`$ one can also show that it vanishes as $`O\left(p^2𝒫(\mathrm{log}\frac{\mu ^2}{p^2})\right)`$ near $`0`$ momentum.
Since $`\mathrm{\Lambda }`$ acts as an infrared regulator the bounds (16, 17) still hold for $`\mathrm{\Lambda }>\mu `$, on replacing $`m`$ by $`\mu `$. For $`\mathrm{\Lambda }<\mu `$ these bounds also hold for nonexceptional momentum configurations. For exceptional configurations they have to be multiplied by the power of $`\mu /\mathrm{\Lambda }`$ appearing in (27). We do not enter into details of the infrared problem, since the bounds in the region $`\mathrm{\Lambda }<\mu `$ are independent of the $`\varphi ^6`$-insertion and therefore the proof from \[KeKo2\] may be taken over unaltered. As regards the term $`\mathrm{\Lambda }_0^3𝒫(log\frac{\mathrm{\Lambda }_0}{\mu })`$ appearing in (17), it does not interfer with the exceptional momentum problem and can be factored out in the inductive proof as it was done before in the massive case.
We now want to establish bounds for the difference between the theories with and without $`\varphi ^6`$-insertion. The CAG of this theory are to be called $`\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$. This means we define (in obvious notation)
$$\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(p_1,\mathrm{},p_{n1})=_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(g_6;p_1,\mathrm{},p_{n1})_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(0;p_1,\mathrm{},p_{n1}).$$
(28)
Here it is understood that the $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(g_6)`$ and the $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(0)`$ obey the same renormalization conditions, which means that all the relevant $`\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$ are imposed to vanish at the renormalization point. We may obtain the flow equations for the $`\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}`$ by taking the difference between those for $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(g_6)`$ and $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(0)`$. We only give it in shortened form without momentum arguments, the explicit form following directly from (14). We get
$$_\mathrm{\Lambda }^w\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}=\frac{1}{2}_k(_\mathrm{\Lambda }C^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(k))^w\mathrm{\Delta }_{l1,n+2}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(k,k,\mathrm{})$$
(29)
$$\underset{\genfrac{}{}{0pt}{}{l_1+l_2=l,w_1+w_2+w_3=w}{n_1+n_2=n}}{}\frac{1}{2}\left[^{w_1}\mathrm{\Delta }_{l_1,n_1+1}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(^{w_3}_\mathrm{\Lambda }C^{\mathrm{\Lambda },\mathrm{\Lambda }_0})^{w_2}\left(_{l_2,n_2+1}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(g_6)+_{l_2,n_2+1}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(0)\right)\right]_{ssym}.$$
With this system of equations we can inductively prove the following bounds for the massless theory. For nonexceptional momentum configurations one finds
$$|^w\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\stackrel{}{p})|\{\begin{array}{cc}\hfill \frac{𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })}{\mathrm{\Lambda }_0^2}\mu ^{6n|w|}𝒫(\frac{|\stackrel{}{p}|}{\mu }),& \text{ for }0\mathrm{\Lambda }\mu \hfill \\ \hfill \frac{𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })}{\mathrm{\Lambda }_0^2}\mathrm{\Lambda }^{6n|w|}𝒫(\frac{|\stackrel{}{p}|}{\mathrm{\Lambda }}),& \text{ for }\mu \mathrm{\Lambda }\mathrm{\Lambda }_0\hfill \end{array}\},$$
(30)
whereas for general momentum configurations one obtains
$$|^w\mathrm{\Delta }_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(\stackrel{}{p})|\{\begin{array}{cc}\hfill \frac{𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })}{\mathrm{\Lambda }_0^2}𝒫(\mathrm{log}\mu /\mathrm{\Lambda })\mu ^{6n|w|}(\frac{\mu }{\mathrm{\Lambda }})^{2\gamma +|w|}𝒫(\frac{|\stackrel{}{p}|}{\mu }),& \text{ for }0\mathrm{\Lambda }\mu \hfill \\ \hfill \frac{𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })}{\mathrm{\Lambda }_0^2}\mathrm{\Lambda }^{6n|w|}𝒫(\frac{|\stackrel{}{p}|}{\mathrm{\Lambda }}),& \text{ for }\mu \mathrm{\Lambda }\mathrm{\Lambda }_0\hfill \end{array}\}.$$
(31)
We do not give a proof of these bounds, since they are obtained using the same inductive scheme as before, applying also the bounds for $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(g_6)`$ and $`_{l,n}^{\mathrm{\Lambda },\mathrm{\Lambda }_0}(0)`$ obtained previously. The improvement factor $`𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })/\mathrm{\Lambda }_0^2`$ is respected in particular by the new boundary conditions: All renormalization conditions vanish, and the only nonvanishing boundary term, i.e. the term $`g_6Z^3/\mathrm{\Lambda }_0^2`$ for the six point function satisfies (30, 31). Still we would like to point out that rigorous bounds as (17, 30, 31) are hard (if not impossible) to obtain by other methods. We will use them in the next section to obtain equivalent bounds on the corrections to scaling due to irrelevant terms.
## 3 Renormalization Group Equations and Critical Behaviour
We will use the previous results to analyse the modification of critical behaviour by irrelevant terms without composite operator formalism. The advantages of this procedure have been mentioned before. In this last section we will change to the standard notation in the sense that now $`_2^{0,\mathrm{\Lambda }_0}`$ denotes the two point function including the 0 loop contribution. Our CAG $`n`$-point functions $`_n^{0,\mathrm{\Lambda }_0}`$ are defined in terms of the field variable $`\varphi `$, which is the renormalized field in standard language. Relating them to the bare functions expressed in terms of the bare field $`\varphi _B`$ which is related to $`\varphi `$ through the relation
$$\varphi _B=Z^{1/2}\varphi $$
(32)
we obtain
$$_n^b(p_i,g_0,g_6,\mathrm{\Lambda }_0)=Z^{n/2}(g,g_6,\frac{\mathrm{\Lambda }_0}{\mu })_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu ).$$
(33)
The sign in the exponent of $`Z`$ is related to the fact that the functions $`^{0,\mathrm{\Lambda }_0}`$ are the connected free propagator amputated functions. This sign changes if we use the full propagator amputated functions instead, which is of course possible, but less natural in the FE framework. Taking a derivative of (33) w.r.t. $`\mathrm{ln}\mu `$ at fixed bare parameters we obtain the (standard) renormalization group equation for the renormalized theory
$$\left[\frac{}{\mathrm{ln}\mu }+\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})\frac{}{g}+\frac{1}{2}n\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})\right]_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=0.$$
(34)
We have introduced the $`\beta `$ and $`\gamma `$ functions for the renormalized theory
$$\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})=\frac{g}{\mathrm{ln}\mu }|_{g_0,g_6,\mathrm{\Lambda }_0},\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})=\frac{\mathrm{ln}Z}{\mathrm{ln}\mu }|_{g_0,g_6,\mathrm{\Lambda }_0}.$$
(35)
Since we want to use this equation for large but nevertheless finite $`\mathrm{\Lambda }_0`$ the functions $`\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})`$ and $`\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})`$ depend also on $`\mathrm{\Lambda }_0`$. Due to (17) the $`\mathrm{\Lambda }_0`$-dependent terms are bounded by $`O((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu }))`$, since $`\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})`$ and $`\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})`$ may be expressed in terms of $`^{0,\mathrm{\Lambda }_0}`$ using (34) for fixed values of $`n`$: By dimensional analysis we transform the derivative w.r.t. $`\mu `$ into a derivative w.r.t. $`p`$ and $`\mathrm{\Lambda }_0`$ and obtain from the equations for $`n=4`$ and for $`n=2`$ :
$$\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})=$$
$$=\underset{i=1}{\overset{3}{}}p_{i,\nu }\frac{}{p_{i,\nu }}_4^{0,\mathrm{\Lambda }_0}(g,g_6)|_{r.p.}\mathrm{\hspace{0.17em}4}gp^2\frac{^2}{(p^2)^2}_2^{0,\mathrm{\Lambda }_0}(g,g_6)|_{r.p.}+O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(log\frac{\mathrm{\Lambda }_0}{\mu })\right),$$
(36)
$$\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})=2p^2\frac{^2}{(p^2)^2}_2^{0,\mathrm{\Lambda }_0}|_{r.p.}+O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })\right).$$
The functions are to be taken at the renormalization points (see (24)). The contributions $`O(\mathrm{\Lambda }_0^2𝒫\mathrm{log}\mathrm{\Lambda }_0)`$ arise when transforming the $`\mu `$-derivative into one on $`\mathrm{\Lambda }_0`$ on using the bound (17). So to be precise we rewrite (34) as
$$\left[\frac{}{\mathrm{ln}\mu }+\beta (g,g_6)\frac{}{g}+\frac{1}{2}n\gamma (g,g_6)\right]_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })\right),$$
(37)
where the whole dependence on $`\mathrm{\Lambda }_0`$ has been regrouped on the r.h.s. (with the definitions $`\beta (g,g_6)=\beta (g,g_6;0),\gamma (g,g_6)=\gamma (g,g_6;0)`$). When setting $`g_6=0`$ we obtain the corresponding equation with functions $`\beta (g,0;\frac{\mu }{\mathrm{\Lambda }_0})`$ and $`\gamma (g,0;\frac{\mu }{\mathrm{\Lambda }_0})`$ obeying the equations analogous to (36) for $`g_6=0`$. From this it follows on using (30) that
$$\mathrm{\Delta }\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0}):=\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})\beta (g,0;\frac{\mu }{\mathrm{\Lambda }_0})=O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })\right),$$
(38)
and similarly for $`\gamma `$. This bound can of course be verified in lowest orders by direct calculation of the respective $`\beta `$-functions. When the $`\varphi ^6`$-term is added, the two diagrams given in Fig.1 contribute to the relation between $`g`$ and $`g_0`$ and thus to $`\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})`$ up to two loops. <sup>9</sup><sup>9</sup>9We did not include those diagrams which are exactly cancelled by diagrams carrying an insertion of a counterterm. Since the second diagram is $`\mu `$-independent, only the first contributes to the $`\beta `$-function. The value of the diagram is
$$gg_6\frac{2}{16\pi ^4}\mathrm{ln}\frac{4}{3}+O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2\mathrm{log}(\frac{\mathrm{\Lambda }_0}{\mu })\right),$$
so that after derivation w.r.t. $`\mathrm{ln}\mu `$ its contribution is of the order given in (38).
Figure 1: Contributions $`g_6`$ to the relation between $`g`$ and $`g_0`$ up to two loops.
We refer to the textbooks \[ZJ, IZ\] for the method of solution of (34), which permits to compare $`_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )`$ to $`_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g(s),g_6,\mu )`$, the critical region corresponding to $`s\mathrm{}`$. Here the running coupling at scale $`\mu /s`$ is defined through
$$\frac{dg(s)}{dlns}=\beta (g(s));g(1)=g.$$
(39)
From (34), together with dimensional analysis, one obtains
$$_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=s^{4+n}e^{\frac{1}{2}n_g^{g(s)}\frac{\gamma (g^{},g_6;\mu /\mathrm{\Lambda }_0)}{\beta (g^{},g_6;\mu /\mathrm{\Lambda }_0)}𝑑g^{}}_n^{0,s\mathrm{\Lambda }_0}(sp_i,g(s),g_6,\mu )$$
(40)
or on replacing $`p_i`$ by $`p_i/s`$:
$$_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,g_6,\mu )=s^{4+n}e^{\frac{1}{2}n_g^{g(s)}\frac{\gamma }{\beta }𝑑g^{}}_n^{0,s\mathrm{\Lambda }_0}(p_i,g(s),g_6,\mu ).$$
(41)
For $`s>>1`$ the coupling will approach its fixed point value $`g^{}`$ for which by definition $`\beta (g^{})=0`$. In the perturbative region we have $`g^{}=0`$ in $`d=4`$ , whereas in $`d<4`$ one finds $`g^{}=O(\epsilon )`$ with $`\epsilon =4d`$. If $`g`$ is in the vicinity of the fixed point the integral $`_g^{g(s)}\frac{\gamma (g^{})}{\beta (g^{})}𝑑g^{}`$ is approximated by its value at $`g^{}`$
$$_g^{g(s)}\frac{\gamma (g^{})}{\beta (g^{})}𝑑g^{}=_0^{\mathrm{ln}s}\gamma (g(s^{}))d\mathrm{ln}s^{}\gamma (g^{})\mathrm{ln}s.$$
(42)
The neglected terms give subdominant contributions for $`s\mathrm{}`$, they are analysed in \[BGZ\]. From this we then find for the dominating behaviour
$$_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,g_6,\mu )s^{4+n(1\frac{\gamma (g^{})}{2})}_n^{0,\mathrm{}}(p_i,g^{},g_6,\mu ),$$
(43)
which shows that the fixed point value $`\gamma (g^{})`$ is to be identified with the critical exponent $`\eta `$.
The renormalization group equation for the difference functions (28) can be obtained from (34). We write it in the form
$$\left[\frac{}{\mathrm{ln}\mu }+\beta (g,0;\frac{\mu }{\mathrm{\Lambda }_0})\frac{}{g}+\frac{n}{2}\gamma (g,0;\frac{\mu }{\mathrm{\Lambda }_0})\right]\mathrm{\Delta }_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=$$
(44)
$$=\left[\mathrm{\Delta }\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})\frac{}{g}+\frac{n}{2}\mathrm{\Delta }\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})\right]_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu ).$$
For the inhomogeneous equation we make the ansatz
$$\mathrm{\Delta }_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=U_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )_n^{0,\mathrm{\Lambda }_0}(p_i,g,0,\mu ).$$
(45)
From this we obtain the following differential equation for $`U^{0,\mathrm{\Lambda }_0}`$
$$(\frac{}{\mathrm{ln}\mu }+\beta (g,0;\frac{\mu }{\mathrm{\Lambda }_0})\frac{}{g})U^{0,\mathrm{\Lambda }_0}=$$
(46)
$$=[\mathrm{\Delta }\beta (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})\frac{}{g}\mathrm{ln}_n^{0,\mathrm{\Lambda }_0}(p_i,g,0,\mu )+\frac{n}{2}\mathrm{\Delta }\gamma (g,g_6;\frac{\mu }{\mathrm{\Lambda }_0})]+O\left((\frac{\mu }{\mathrm{\Lambda }_0})^4𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })\right).$$
In the following we will negelect the last term which gives even smaller corrections, for the first two terms on the r.h.s. of this equation we write $`V_n(p_i,\mu ,g;g_6,\mathrm{\Lambda }_0)`$. Its solution is then obtained as a sum of the general soultion of the corresponding homogeneous equation -which in turn is obtained as previously for the case $`\gamma =0`$\- plus a special solution of the inhomogeneous equation, which can be written as the integral over $`V_n(p_i,\mu ,g;g_6,\mathrm{\Lambda }_0)`$. As a final result we obtain the following renormalization group relation for $`U_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )`$ :
$$U_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=U_n^{0,\mathrm{\Lambda }_0}(p_i,g(s),g_6,\mu /s)+_{lns}^0V_n(p_i,\mu e^t,g(e^t);g_6,\mathrm{\Lambda }_0)𝑑t.$$
(47)
By dimensional analysis we obtain
$$U_n^{0,\mathrm{\Lambda }_0}(p_i,g,g_6,\mu )=U_n^{0,s\mathrm{\Lambda }_0}(sp_i,g(s),g_6,\mu )+_{lns}^0V_n(sp_i,s\mu e^t,g(e^t);g_6,s\mathrm{\Lambda }_0)𝑑t,$$
(48)
since the canonical dimension of $`U_n`$ is zero. Multiplying by $`_n^{0,\mathrm{\Lambda }_0}(p_i,g,0,\mu )`$, using (41) and passing to momenta $`p_i/s`$ we thus obtain
$$\mathrm{\Delta }_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,g_6,\mu )=s^{4+n}e^{\frac{1}{2}n_g^{g(s)}\frac{\gamma (g^{},0;\mu /\mathrm{\Lambda }_0)}{\beta (g^{},0;\mu /\mathrm{\Lambda }_0)}𝑑g^{}}[\mathrm{\Delta }_n^{0,s\mathrm{\Lambda }_0}(p_i,g(s),g_6,\mu )+$$
(49)
$$+_n^{0,s\mathrm{\Lambda }_0}(p_i,g(s),0,\mu )_{lns}^0V_n(p_i,s\mu e^t,g(e^t);g_6,s\mathrm{\Lambda }_0)dt].$$
The second term can be bounded using (38) (together with (16))<sup>10</sup><sup>10</sup>10It is useful to cut the integration interval into subintervals of length $`\mathrm{ln}2`$ and sum over the bounds for the integrand in the subintervals to avoid a factor of $`\mathrm{ln}s`$ in the bound for this term., to the first term we can apply (30) to obtain the following bound on $`\mathrm{\Delta }_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,g_6,\mu )`$:
$$|\mathrm{\Delta }_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,g_6,\mu )|$$
(50)
$$s^{4+n}e^{\frac{1}{2}n_g^{g(s)}\frac{\gamma }{\beta }𝑑g^{}}\left[O\left((\frac{\mu }{s\mathrm{\Lambda }_0})^2𝒫(\mathrm{log}\frac{s\mathrm{\Lambda }_0}{\mu })\right)+|_n^{0,s\mathrm{\Lambda }_0}(p_i,g(s),0,\mu )|O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(\mathrm{log}\frac{\mathrm{\Lambda }_0}{\mu })\right)\right].$$
This bound is dominated for $`s`$ large by the second term so that we obtain
$$|\mathrm{\Delta }_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,g_6,\mu )|s^{4+n}e^{\frac{1}{2}n_g^{g(s)}\frac{\gamma }{\beta }𝑑g^{}}|_n^{0,s\mathrm{\Lambda }_0}(p_i,g(s),0,\mu )|O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2𝒫(log\frac{\mathrm{\Lambda }_0}{\mu })\right).$$
(51)
The analysis of the prefactor is the same as for $`_n^{0,\mathrm{\Lambda }_0}(\frac{p_i}{s},g,0,\mu )`$. Therefore, close to the critical region, the corrections of the long distance behaviour due to the irrelevant term are of the relative order $`O((\frac{\mu }{\mathrm{\Lambda }_0})^2)`$ up to logarithms. For this term to be negligeable we need of course $`\mu <<\mathrm{\Lambda }_0`$, that is to say, the renormalized parameters are close to the critical ones, which is a natural parametrization in the critical region. We emphasize that the corrections to scaling stem from the analysis of the terms vanishing for $`\mathrm{\Lambda }_0\mathrm{}`$, which are often neglected altogether in the literature. In the composite operator analysis one finds instead corrections $`s^2𝒫(\mathrm{log}s)`$, which would be smaller for $`s>\frac{\mathrm{\Lambda }_0}{\mu }`$. However the terms $`(\frac{\mu }{\mathrm{\Lambda }_0})^2`$ are always present, though often neglected, so that the corresponding results only hold up to $`s\frac{\mathrm{\Lambda }_0}{\mu }`$. For larger $`s`$ one has to readapt the renormalization conditions at $`\mu ^{}<<\mu `$. In terms of the bare theory the readaptation consists in adding new counterterms $`\varphi ^2`$ and $`\varphi ^4`$. This is well known from the treatment in the composite operator formalism, where such terms are introduced due to operator mixing.
In conclusion we thus realize that in our approach the corrections to scaling due to irrelevant terms are suppressed by $`O\left((\frac{\mu }{\mathrm{\Lambda }_0})^2\right)`$ to any order in the number of insertions. These irrelevant terms are introduced directly in the bare action, keeping the renormalization conditions fixed. In composite operator theory, which is completely bypassed here, the coefficient of the $`\varphi ^6`$-term in the bare action is not suppressed by $`(\frac{1}{\mathrm{\Lambda }_0})^2`$, correspondingly one does not obtain such a suppression in the corrections to scaling. Instead, on subtracting insertions of lower dimension, to be calculated from the relations for operator mixing, one obtains a suppression factor $`s^2`$, which is larger for $`s<\frac{\mathrm{\Lambda }_0}{\mu }`$, becomes of similar size for $`s\frac{\mathrm{\Lambda }_0}{\mu }`$ and unreliable beyond.
References:
* \] D.J. Amit: Field Theory, the Renormalization Group and Critical Phenomena, $`2^{nd}`$ ed. World Scientific, Singapur (1984).
* \] E. Brézin, J.C. Le Guillou and J. Zinn-Justin, in: Phase Transitions and Critical Phenomena, Vol. VI, Ch.3, C. Domb et M. S. Green, eds. Academic Press, N.Y. (1976).
* \] C. Itzykson and J.B. Zuber: Quantum Field Theory, Mc Graw Hill, N.Y. (1980).
* \] G. Keller and Ch. Kopper. Perturbative Renormalization of Composite Operators via Flow Equations I. Commun.Math.Phys.148, (1992) 445-467.
* \] G. Keller and Ch. Kopper, Perturbative Renormalization of Massless $`\varphi _4^4`$ with Flow Equations. Commun.Math.Phys.161, (1994) 515-532.
* \] G. Keller, Ch. Kopper, M. Salmhofer, Perturbative Renormalization and Effective Lagrangians in $`\varphi _4^4`$. Helv.Phys.Acta 65, (1991) 33-52.
* \] Ch. Kopper, V.F.M. Müller, Th. Reisz, Temperature Independent Renormalization of Finite Temperature Field Theory. Preprint hep-th/0003254.
* \] Ch. Kopper and V.A. Smirnov, Analytic Regularization and Minimal Subtraction of $`\varphi _4^4`$ with Flow Equations. Z.Phys.C59, (1993) 641-645.
* \] Ch. Kopper: Renormierungstheorie mit Flussgleichungen, Shaker Verlag, Aachen (1998).
* \] J. Polchinski, Renormalization and Effective Lagrangians. Nucl.Phys.B231, (1984) 269.
* \] F. Wegner, Corrections to Scaling Laws, Physical Review B5, (1972) 4529-4536, see also F. Wegner, in: Phase Transitions and Critical Phenomena, Vol. VI, Ch.2, C. Domb et M. S. Green, eds. Academic Press, N.Y. (1976).
* \] K. Wilson et J. B. Kogut, The Renormalization Group and the $`\epsilon `$-Expansion. Physics Reports, 12c, (1974) 77.
* \] W. Zimmermann, Composite Operators in the Perturbation Theory of Renormalizable Interactions, Ann.Phys. 77, (1973) 536-569.
* \] J. Zinn-Justin: Quantum Field Theory and Critical Phenomena, Clarendon Press, Oxford (1989).
|
warning/0007/hep-th0007029.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The idea that our world is a 3-brane embedded in extra spatial dimensions has been widely discussed as a solution to the weak-scale hierarchy problem . More recently, attention has been focused on the possibilities for understanding the cosmological constant problem within this setting (see for an earlier treatment). A partial solution was proposed in (ADKS–KSS), where the addition of a bulk scalar, $`\varphi `$, plays a crucial role (see also for a discussion of the physics of a scalar field coupled to gravity). The scalar is a free field in the bulk, but has nontrivial couplings to fields living on the brane. Ref. found static solutions of Einstein’s equations with the property that $`\varphi `$ becomes singular at a finite distance in the extra dimension, and the warp factor for the metric vanishes at the singularity. If one assumes that the extra dimension terminates at the singularity, or that the warp factor remains integrably small beyond it, then gravity appears to be four dimensional on large distance scales. Most importantly, the scalar field is supposed to adjust itself to any arbitrary value of the tension on the brane, which represents the four-dimensional vacuum energy–or at least that part of it which comes from nongravitational vacuum fluctuations. The fact that the metric is static means that the effective cosmological constant observed on the 3-brane is zero, regardless of the size of the brane tension. This could constitute significant progress toward the solution of the cosmological constant problem.
The self-tuning mechanism is incomplete in several ways. In its original form in , the orbifold solution requires a very particular exponential coupling of $`\varphi `$ to the matter fields on the brane, $`e^{\pm \kappa _5\varphi }`$, requiring just the right coefficient $`\kappa _5`$ in the exponent, where $`\kappa _5`$ is related to the 5-D gravity scale $`M_5`$ by $`\kappa _5^2=M_5^3`$. As understood by ADKS, and explicitly realized in refs. , different choices of the coupling function $`f(\varphi )`$ give de Sitter or anti-de Sitter branes in a $`_2`$ bulk. Furthermore, the scalar potential in the bulk was assumed to vanish. Ref. extended the analysis to non-vanishing potentials, and gave the procedure for finding solutions with arbitrary potentials in the bulk (see for an analytic solution associated to a bulk cosmological constant and see for a discussion of exponential potentials which can be associated with Neveu–Schwarz dilaton tadpoles in non-supersymmetric string theories ). Actually, as recently pointed out in Ref. , the $`_2`$ symmetric and 4D Poincaré invariant solution is unstable under bulk quantum corrections: indeed with a conformal coupling allowing flat solution with a vanishing bulk potential, the brane becomes curved as soon as a bulk potential is turned on and the jump equations relate the curvature of the brane to the value of the potential on the brane by $`R_{\text{4D}}=\kappa _5^2V(\varphi _0)`$. Supersymmetry in the bulk may prevent from such instability. Another difficulty anticipated by , and explicitly shown by , is that any procedure which regularizes (“resolves”) the singularity in the solutions causes the reintroduction of the fine-tuning which self-tuning is supposed to avoid, unless some more explicit dynamical mechanism which automatically relaxes the effects of the brane tension can be demonstrated. A further possible problem is the claim that when normal matter is added to the brane tension, the brane remains static, in contradiction with cosmology (however, see for a recent tentative to recover usual cosmology).
In this letter we demonstrate a shortcoming which is more severe than the foregoing ones; namely, starting from the very same Lagrangian which gives the static self-tuned solutions, there also exist dynamical solutions, which either begin or end with a singularity as time evolves. In section 2 we will review the static solution, and discuss the conformal symmetry which allows construction of the dynamical solutions. These constitute a family of solutions, of which the static one is a special example. In section 3 we emphasize that, even starting arbitrarily close to the static solution within this family, the brane world inevitably collapses to a singularity or else expands starting from one, with a Hubble parameter of $`H\pm 1/(4t)`$ as $`t\mathrm{}`$. The interpretation is that the static solution is a saddle point, unstable to small perturbations. The solution on the brane is shown to violate the positive energy condition, reflecting the loss of energy from the brane into the bulk via the scalar field. These remarks apply in the case when the scalar bulk potential, $`V(\varphi )`$, vanishes. In section 4 we show that for $`V(\varphi )0`$, our construction cannot be trivially applied to generate dynamical solutions. This gives further motivation for studying the stability of self-tuning solutions with a nonvanishing scalar bulk potential.
The solutions we constructed were independently found by Horowitz, Low and Zee in Ref. and were interpreted as describing a phase transition. We will argue in the final section that they actually rather signal an instability of the static solution.
## 2 Dynamical Self-Tuning Solutions
We will consider solutions arising from the action<sup>1</sup><sup>1</sup>1Our conventions correspond to a mostly positive Lorentzian signature $`(+\mathrm{}+)`$ and the definition of the curvature in terms of the metric is such that a Euclidean sphere has positive curvature. Bulk indices will be denoted by capital Latin indices and brane indices by Greek indices: $`x^\mu `$ are coordinates on the brane ($`\tau `$ or $`t`$ will be the time coordinate and $`x^i`$ the spatial ones), and $`y`$ (or $`z`$, if the metric is conformally flat) is the coordinate along the fifth dimension such that the brane is located at $`y=0`$ (or $`z=0`$).
$$S=d^{\mathrm{\hspace{0.17em}5}}x\sqrt{|g_5|}\left(\alpha R\beta _M\varphi ^M\varphi \gamma V(\varphi )\right)d^{\mathrm{\hspace{0.17em}4}}x\sqrt{|g_4|}f(\varphi _0)T,$$
(1)
where $`g_5`$ and $`g_4`$ are, respectively, the determinants the 5-D metric $`g_{MN}`$ and the 4-D metric induced on the brane, $`g_{\mu \nu }`$, and $`T`$ is the bare tension. The brane is supposed to couple to the bulk in a conformal way defined by the function $`f(\varphi )`$. The physical tension is thus given by
$$\stackrel{~}{T}=f(\varphi _0)T,$$
(2)
where $`\varphi _0`$ is the value of the scalar field on the brane. There are many conventions for the normalization of the terms in the bulk part of the action; to facilitate comparison with other papers we will leave $`\alpha ,\beta ,\gamma `$ unspecified. We will be primarily concerned with the case of vanishing bulk potential, $`V(\varphi )=0`$, but we shall also consider nonzero $`V(\varphi )`$ below. Einstein’s equations and the equation of motion for the scalar field read
$`\alpha G_{MN}=\beta _M\varphi _N\varphi \frac{1}{2}\left(\beta (\varphi )^2+\gamma V(\varphi )\right)g_{MN}\frac{1}{2}f(\varphi )Tg_{\mu \nu }\delta _M^\mu \delta _N^\nu {\displaystyle \frac{\sqrt{|g_4|}}{\sqrt{|g_5|}}}\delta (y);`$ (3)
$`2\beta {\displaystyle \frac{1}{\sqrt{|g_5|}}}_M\left(\sqrt{|g_5|}g^{MN}_N\varphi \right)\gamma {\displaystyle \frac{dV}{d\varphi }}{\displaystyle \frac{df}{d\varphi }}T{\displaystyle \frac{\sqrt{|g_4|}}{\sqrt{|g_5|}}}\delta (y)=0,`$ (4)
and, for a conformally flat metric with the form $`ds^2=\mathrm{\Omega }^2(z)(d\tau ^2+dx_i^2+dz^2)`$, the jump conditions for the derivatives of the fields at the brane are
$$\left[\mathrm{\Omega }^2\frac{d\mathrm{\Omega }}{dz}\right]_{z=0^{}}^{z=0^+}=\frac{T}{6\alpha }f(\varphi _0)\text{and}\left[\mathrm{\Omega }^1\frac{d\varphi }{dz}\right]_{z=0^{}}^{z=0^+}=\frac{T}{2\beta }\frac{df}{d\varphi }(\varphi _0).$$
(5)
For a vanishing scalar bulk potential, the self-tuning solution of with a $`_2`$–symmetric bulk orbifold is given by
$`ds^2=\mathrm{\Omega }^2(y)(d\tau ^2+dx_i^2)+dy^2=\mathrm{\Omega }^2(z)(d\tau ^2+dx_i^2+dz^2);`$ (6)
$`\varphi =\varphi _0\pm \sqrt{\frac{3\alpha }{4\beta }}\mathrm{ln}(1|y|/y_c)=\varphi _0\pm \sqrt{\frac{4\alpha }{3\beta }}\mathrm{ln}(1|z|/z_c),`$ (7)
where $`y`$ is the proper distance coordinate, $`z`$ is the conformal coordinate for the bulk and
$$\mathrm{\Omega }(y)=\left(1|y|/y_c\right)^{1/4},\mathrm{\Omega }(z)=\left(1|z|/z_c\right)^{1/3}.$$
(8)
However, this solution satisfies the jump conditions (5) only if the conformal coupling is an exponential function with the particular form
$$f(\varphi )=\mathrm{exp}(\sqrt{\frac{4\beta }{3\alpha }}\varphi ).$$
(9)
Then the integration constant, $`y_c`$ or $`z_c`$, is related to the brane tension by
$$y_c=\frac{3}{4}z_c=3\alpha \stackrel{~}{T}^1=3\alpha e^{\pm \sqrt{{\scriptscriptstyle \frac{4\beta }{3\alpha }}}\varphi _0}T^1$$
(10)
and $`\varphi _0`$ remains unconstrained—which is important for what follows. A notable peculiarity of this conformal coupling is that it satisfies the following equations everywhere in the bulk:
$$f(\varphi (y))=\frac{12\alpha }{T}\text{sgn}(y)\mathrm{\Omega }^1(y)\frac{d\mathrm{\Omega }}{dy}\text{and}\frac{df}{d\varphi }(\varphi (y))=\frac{4\beta }{T}\text{sgn}(y)\frac{d\varphi }{dy}$$
(11)
or, in conformal coordinates,
$$f(\varphi (z))=\frac{12\alpha }{T}\text{sgn}(z)\mathrm{\Omega }^2(z)\frac{d\mathrm{\Omega }}{dz}\text{and}\frac{df}{d\varphi }(\varphi (z))=\frac{4\beta }{T}\text{sgn}(z)\mathrm{\Omega }^1(z)\frac{d\varphi }{dz}.$$
(12)
Hence, the relation between the field derivatives and $`\varphi `$ required by the jump conditions at the brane are satisfied not just there, but everywhere in the bulk.
There is a simple procedure for transforming static bulk solutions into dynamical ones. As emphasized in , bulk symmetries can be used to construct new solutions involving a singularity interpreted as a brane. Starting from any regular static solution to the bulk equations of motion, written for simplicity in conformal coordinates,
$$\varphi (z)\text{and}ds^2=\mathrm{\Omega }^2(z)\eta _{MN}dx^Mdx^N,$$
(13)
we obtain a physically equivalent solution by applying a diffeomorphism
$$\stackrel{~}{\varphi }(\stackrel{~}{x}^\mu ,\stackrel{~}{z})=\varphi (z(\stackrel{~}{x}^\mu ,\stackrel{~}{z}))\text{and}ds^2=\stackrel{~}{g}_{MN}(\stackrel{~}{x}^\mu ,\stackrel{~}{z})d\stackrel{~}{x}^Md\stackrel{~}{x}^N.$$
(14)
However if we orbifold the new solution in the $`\stackrel{~}{z}`$ direction, it becomes singular and describes a brane located at $`\stackrel{~}{z}=0`$. The orbifold projection and diffeomorphisms do not commute: in general an orbifold solution constructed from (14) is not equivalent under a change of coordinates to an orbifold solution constructed from (13). The difficulty consists in finding a diffeomorphism such that the singularities introduced in the right hand side of the equations of motion by the orbifold projection can be associated with the brane components deduced from the action (1).
In order to preserve the geometry of the brane embedded in the bulk, we want to restrict ourselves to diffeomorphisms (14) that keep the metric diagonal,
$$ds^2=\stackrel{~}{n}^2(\stackrel{~}{x}^\mu ,\stackrel{~}{z})d\stackrel{~}{\tau }^2+\stackrel{~}{a}^2(\stackrel{~}{x}^\mu ,\stackrel{~}{z})d\stackrel{~}{x}_i^2+\stackrel{~}{b}^2(\stackrel{~}{x}^\mu ,\stackrel{~}{z})d\stackrel{~}{z}^2.$$
(15)
A particular subgroup is provided by the 5-D conformal transformations under which the metric remains conformally flat
$$ds^2=\stackrel{~}{\mathrm{\Omega }}^2(\stackrel{~}{x}^\mu ,\stackrel{~}{z})\eta _{MN}d\stackrel{~}{x}^Md\stackrel{~}{x}^N.$$
(16)
The infinitesimal conformal transformations are generated by the Killing vectors
$$\xi ^M=a^M+a^{[MN]}x_N+\lambda x^M+(x^Px_P\eta ^{MN}2x^Mx^N)k_N$$
(17)
where the parameters $`a^M`$, $`a^{[MN]}`$, $`\lambda `$, $`k_N`$ correspond respectively to translations, Lorentz transformations, dilations and special conformal transformations.
In the present case, a combination of a boost in the $`z`$ direction and a dilation provides a suitable diffeomorphism that will lead to a dynamical solution also satisfying the boundary conditions on the brane. If $`\varphi (z)`$ and $`\mathrm{\Omega }(z)`$ is a regular solution in the bulk, then it is simple to show that
$$\stackrel{~}{\varphi }(z,\tau )=\varphi (|z|+z_ch\tau )\text{and}\stackrel{~}{\mathrm{\Omega }}(z,\tau )=\frac{\mathrm{\Omega }(|z|+z_ch\tau )}{\sqrt{1z_c^2h^2}}$$
(18)
is a $`_2`$-symmetric solution to the bulk equations of motion. This can be checked from the explicit form of the bulk part of the action (1), which looks like
$$S_{\mathrm{bulk}}=d^{\mathrm{\hspace{0.17em}5}}x\left[4\alpha (\mathrm{\Omega }(\mathrm{\Omega })^2+2\mathrm{\Omega }^2^2\mathrm{\Omega })\beta \mathrm{\Omega }^3(\varphi )^2\gamma \mathrm{\Omega }^5V(\varphi )\right].$$
(19)
Here the contractions of $`_M`$ are performed with the Minkowski space metric. This action has the additional symmetry that, for any constant $`\zeta `$, leaves the equations of motion invariant:
$$\mathrm{\Omega }\zeta \mathrm{\Omega };\varphi \varphi ;V\zeta ^2V.$$
(20)
In the case of a vanishing bulk potential, this symmetry implies that
$$\stackrel{~}{\varphi }(z,\tau )=\varphi (|z|+z_ch\tau )\text{and}\stackrel{~}{\mathrm{\Omega }}(z,\tau )=\mathrm{\Omega }(|z|+z_ch\tau )$$
(21)
is a solution in the bulk. Applying the procedure to the regular solution corresponding to (6)–(7), we obtain
$`ds^2=\left(1|z|/z_ch\tau \right)^{2/3}(d\tau ^2+dx_i^2+dz^2);`$ (22)
$`\varphi =\varphi _0\pm \sqrt{\frac{4\alpha }{3\beta }}\mathrm{ln}(1|z|/z_ch\tau ).`$ (23)
The nontrivial step is to satisfy the jump equations at the brane, which read
$`\stackrel{~}{\mathrm{\Omega }}^2{\displaystyle \frac{d\stackrel{~}{\mathrm{\Omega }}}{dz}}|_{z=0^+}={\displaystyle \frac{T}{12\alpha }}f(\stackrel{~}{\varphi })|_{z=0^+};\stackrel{~}{\mathrm{\Omega }}^1{\displaystyle \frac{d\stackrel{~}{\varphi }}{dz}}|_{z=0^+}={\displaystyle \frac{T}{4\beta }}{\displaystyle \frac{df}{d\varphi }}(\stackrel{~}{\varphi })|_{z=0^+}.`$ (24)
These equations are more difficult to satisfy when the solutions are dynamical, because they must remain true for all conformal times $`\tau `$. However, in the present case of a vanishing bulk potential, we notice that, due to eqs. (11)–(12), the conformal coupling $`f`$ of the scalar to the brane satisfies the relations (12) for any value of $`z`$, which ensures that the dynamical solution we construct satisfies the jump equations for any $`\tau `$, as can also be explicitly verified. So from the same Lagrangian which gives the static self-tuned solutions, there also exist dynamical solutions for which the induced metric on the brane exhibits time dependence, as will be discussed in the next section. The original 4-D Poincaré invariant solution corresponds to a very particular value of the parameter $`h`$ that characterized our more general family of solutions. This evades the no-go result of which excluded the possibility of de Sitter or anti-de Sitter branes in this case.
## 3 Physical Interpretation
The dynamical solution (22)–(23) represents a singularity which is either approaching or receding from the brane, depending on the sign of the continuous parameter $`h`$. Assuming the extra dimension is simply truncated at the singularity, the strength of gravity is therefore time-dependent because of the growth or collapse of the extra dimension. The 4-D Planck mass $`M_p`$ is related to the 5-D analogue $`M_5`$ by
$$M_p^2=2M_5^3_0^{z_c(1h\tau )}\mathrm{\Omega }^3(z,\tau )𝑑z=M_5^3z_c(1h\tau ).$$
(25)
Moreover, an observer on the brane will see that his universe, although spatially flat, does not have 4-D Poincaré invariance, but it is growing or shrinking with a scale factor given by $`\mathrm{\Omega }(z=0,\tau )`$. In FRW time, $`dt=\mathrm{\Omega }d\tau `$, the scale factor and the corresponding Hubble parameter are given by
$$a(t)=(1h\tau )^{1/3}=(1\frac{4}{3}ht)^{1/4};H=\frac{\dot{a}}{a}=\frac{h}{3(1\frac{4}{3}ht)}.$$
(26)
The universe either begins or ends in a singularity, depending on whether $`h<0`$ or $`h>0`$. For the case $`h=0`$, the static solution of ADKS is recovered. Therefore one can interpret $`h`$ as the parameter determining how far away from the unstable saddle point solution one is, in the space of all solutions.
The situation is qualitatively similar to a 4-D field theory analogy, in which a cosmological “constant” $`\mathrm{\Lambda }`$ is coupled to scalar fields through the Lagrangian <sup>2</sup><sup>2</sup>2We thank Nima Arkani-Hamed for making us aware of this concept, through a related example.
$$=a^3(t)\left[\frac{1}{2}(\dot{\varphi }_1^2+\dot{\varphi }_2^2)\varphi _1(\mathrm{\Lambda }\varphi _2)\right].$$
(27)
This system also has a saddle point solution, at $`\varphi _2=\mathrm{\Lambda }`$ and $`\varphi _1=0`$, which could be construed as a self-tuning of the cosmological constant to zero. However, it is not a good solution to the cosmological constant problem because it is unstable against small perturbations. In figure 1(a) we show the time dependence of the Hubble parameter for both the 5-D self-tuning solution and the 4-D toy model, in the case of a collapsing universe, which is the generic outcome for the 4-D model. Although $`H(t)`$ looks rather similar in the two cases, for the 4-D model $`H`$ starts positive and crosses zero, while for the 5-D solution it is always negative. The differences are more clearly seen in figure 1(b), showing the scale factors $`a(t)`$ in FRW time.
Figure 1. (a) Comparison of the Hubble parameter as a function of time for the 5-D self-tuning solution and the 4-D toy model (27), for the case of a collapsing brane-world. (b) Same, but showing the respective scale factors versus time.
A further difference between the 5-D solution and any attempt to describe it in a 4-D effective theory is that the 5-D solution appears to violate energy conservation when viewed from the 3-brane. If we were in 4 dimensions, the scale factor dependence in eq. (26) would correspond to a 4-D stress energy tensor
$$T_\nu ^\mu =\frac{\rho _0}{a^8(t)}\left(\begin{array}{cccc}1& & & \\ & \frac{5}{3}& & \\ & & \frac{5}{3}& \\ & & & \frac{5}{3}\end{array}\right)$$
(28)
where $`\rho _0=M_p^2h^2/3`$. This has the equation of state $`p=(5/3)\rho `$, which implies that there exists a null vector $`\xi _\mu `$ which, when contracted with $`T_\nu ^\mu `$, gives a spacelike vector, contrary to the requirement of positivity of $`T_\nu ^\mu `$. This is due to the nonconservation of energy on the brane, which can be explicitly demonstrated in the 5-D theory, by computing the singular part of the divergence of the 5-D stress energy tensor<sup>3</sup><sup>3</sup>3 In this equation $`\rho `$ and $`p`$ are the physical energy density and pressure describing the matter living on the brane and conformally coupled to the scalar field.
$$\dot{\rho }+3H(\rho +p)=\frac{1}{4}(\rho 3p)\frac{f^{}}{f}\dot{\varphi }$$
(29)
In this context, $`\rho 3p=4\stackrel{~}{T}`$, and in the dynamical solutions, as long as $`h`$ is non-vanishing, $`\dot{\varphi }`$ is nonzero on the 3-brane, so the right hand side of (29) is nonzero. More simply put, since the physical tension $`\stackrel{~}{T}=f(\varphi )T`$ is time-dependent,<sup>4</sup><sup>4</sup>4 Curiously, if all the parameters of the Lagrangian (1) as well as the parameter $`h`$ are of the order of the 4-D Planck scale, with the age of the Universe estimated to fifteen billions years, today the physical tension would be of the order of $`(5\text{TeV})^4`$, close to the electroweak scale. energy is not conserved on the brane.
## 4 Nonvanishing Bulk Potential
An interesting question is whether our procedure can be generalized to the case when the bulk potential is nonvanishing. We do not have a definitive answer, but we can argue that, if dynamical instabilities of the static solution exist, they are much more difficult to find than when $`V(\varphi )=0`$. We are interested in orbifold solutions that can be constructed from a regular bulk solution, $`\mathrm{\Omega }(z)`$ and $`\varphi (z)`$, which satisfies the jump equations (5) at $`z=0`$. An important property of the vanishing potential solution that facilitated finding dynamic solutions was that the jump equations were actually satisfied for any $`z`$, not just at $`z=0`$. This property cannot be maintained when the scalar potential is turned on, while still having a self-tuning solution. Indeed, it is easy to verify that if the relations (5) are satisfied for any arbitrary values of $`z`$, the bulk equations of motion imply the relation
$$V(\varphi )=\frac{2}{\gamma }\left(\frac{T^2}{32\beta }\left(\frac{df}{d\varphi }\right)^2\frac{T^2}{24\alpha }f^2(\varphi )\right)$$
(30)
between the bulk scalar field and the conformal coupling. This relation is incompatible with the self-tuning mechanism unless the bulk potential vanishes; otherwise either $`V(\varphi )`$ or $`f(\varphi )`$ would have to depend on $`T`$, rather than $`\varphi _0`$. One way to overcome this difficulty might be to consider diffeomorphisms $`\tau ,z\stackrel{~}{\tau },\stackrel{~}{z}`$ such that $`\stackrel{~}{z}=0`$ implies $`z=0`$ independently of $`\stackrel{~}{\tau }`$. But it seems difficult to find such a diffeomorphism that leaves the metric diagonal. Another possibility would consist in relaxing the $`_2`$ symmetry in the bulk since a naive count of parameters shows that an integration constant remains unconstrained by the jump equations and could be promoted to a time-dependent function. However the continuity of the solution on the brane is no longer guaranteed at each time. These observations may indicate that the self-tuned solutions in the case of a nonvanishing bulk potential do not suffer from the kind of instability we have found with the ADKS-KSS solution whose pathology comes from the massless and unstabilized scalar field.
## 5 Conclusion
We have shown that the ADKS-KSS self-tuning solution is unstable against eternal expansion or singular collapse of the brane-world. Interestingly, the dynamics on the brane cannot be represented by a 4-D field theory, since energy flows off the brane into the bulk and thus appears not to be conserved on the brane. We have suggested that these problems may not occur in the presence of a potential energy in the bulk for the scalar field. If this hypothesis is correct, then further indirect evidence would consist in perturbing the brane with matter or radiation, and checking whether it behaves according to normal 4-D cosmology. We expect, in analogy to brane models without a stabilized radion, that a nonstandard Friedmann equation on the brane, such as $`H\rho `$ , will be a diagnostic of instabilities in the extra dimension, if they exist. Work along these lines is currently in progress .
In a closely related study, Horowitz, Low and Zee recently presented in a general class of plane wave solutions where the metric is parametrized as $`ds^2=e^{2A(t,y)}(dt^2+dy^2)+e^{2B(t,y)}dx_i^2`$, and $`B(t,y)=(1/3)\mathrm{ln}(f(ty)+g(t+y))`$ for arbitrary functions $`f`$ and $`g`$. The corresponding expressions for $`A(t,y)`$ and $`\varphi (t,y)`$ are generally more complicated, but it is possible to find solutions where $`A(t,y)=B(t,y)`$ and $`f`$ and $`g`$ are both linear functions, which are the same as our solutions.<sup>5</sup><sup>5</sup>5When the warp factor is normalized to one on the brane at $`t=0`$, the relation between the integration constants are $`z_c=y_0/(\xi 2)`$, $`h=\xi /y_0`$ and $`\varphi _0=d+ϵ\mathrm{ln}|y_0|`$. Interestingly, interpret these solutions as describing a phase transition claiming that it is possible to start with the static ADKS-KSS solution, at times $`t<0`$, and smoothly match it to the dynamical solutions for $`t>0`$, provided there is a sudden change $`\mathrm{\Delta }T`$ in the brane tension at time $`t=0`$. This would have demonstrated a physical mechanism for triggering the instability: an arbitrarily small change in the brane tension, such as would occur during a first order phase transition. However, it does not seem possible to have continuous time derivatives of the fields when such a gluing of the two kinds of solutions is attempted.<sup>6</sup><sup>6</sup>6This can be seen in (V.18) of , where $`y_{}^{\prime \prime }(t)`$ must have a Dirac delta function at $`t=0`$ if $`y_{}`$ goes from being a constant to being linear in $`t`$ at $`t=0`$. Such a solution would require that the tension on the brane be proportional to $`\delta (t)`$, rather than simply having a discontinuous change. On the other hand, the property of having dynamical solutions evolving to or from a Big Crunch or Bang with the same value of the brane tension translates into an instability of the static solution with respect to initial time derivatives since a small perturbation in $`\varphi _0^{}`$ drives the solution to a nonvanishing value of $`h`$, and then unavoidably leads to a singularity in the time evolution.
## Acknowledgements
We thank Nima Arkani-Hamed, Csaba Csáki, Joshua Erlich and Stéphane Lavignac for stimulating discussions. P.B. thanks the Theory Group at LBNL for its hospitality and its support. C.G. is supported in part by the Director, Office of Science, Office of High Energy and Nuclear Physics, of the US Department of Energy under Contract DE-AC03-76SF00098, and in part by the National Science Foundation under grant PHY-95-14797.
|
warning/0007/nucl-th0007061.html
|
ar5iv
|
text
|
# Effective interactions in medium heavy nuclei
## 1 Introduction
Nuclei far from the line of $`\beta `$-stability are at present in focus of the nuclear structure physics community. Considerable attention is being devoted to the experimental and theoretical study of nuclei near <sup>100</sup>Sn, from studies of the chain of Sn isotopes up to <sup>132</sup>Sn to e.g., nuclei near the proton drip line like <sup>105,106</sup>Sb. Nuclei like <sup>105</sup>Sb and <sup>109</sup>I have recently been established as ground-state proton emitters . The next to drip line nucleus for the antimony isotopes, <sup>106</sup>Sb with a proton separation energy of $`400`$ keV, was studied recently in two experiments and a level scheme for the yrast states was proposed in Ref. . Similarly, detailed spectroscopy of <sup>107</sup>Sb has also been presented recently .
In this contribution we focus on selected aspects of tin isotopes, with an emphasis on the connection to the underlying nucleon-nucleon interaction and large-scale shell-model studies with effective interactions. The derivation of a shell-model effective interaction is briefly discussed in section 2. Since the effective interactions employed in shell-model calculations are always the outcome of some truncations in the many-body expansion, the shell model may then provide a useful testing ground for the various approximations made. Furthermore, the shell-model wave function can be used to extract information on specific correlations in nuclei, such as pairing correlations. This is discussed in section 3, while section 4 contains new results from studies of Sb isotopes. Concluding remarks are presented in section 5.
## 2 Effective interactions and the shell model
Our scheme to obtain an effective two-body interaction for shell-model studies starts with a free nucleon-nucleon interaction $`V`$ which is appropriate for nuclear physics at low and intermediate energies. In this work we will thus choose to work with the charge-dependent version of the Bonn potential models, see Ref. . The next step in our many-body scheme is to handle the fact that the repulsive core of the nucleon-nucleon potential $`V`$ is unsuitable for perturbative approaches. This problem is overcome by introducing the reaction matrix $`G`$, which in a diagrammatic language represents the sum over all ladder type of diagrams. This sum is meant to renormalize the repulsive short-range part of the interaction. The physical interpretation is that the particles must interact with each other an infinite number of times in order to produce a finite interaction. We calculate $`G`$ using the double-partioning scheme discussed in e.g., Ref. . Since the $`G`$-matrix represents just the summation to all orders of particle-particle ladder diagrams, there are obviously other terms which need to be included in an effective interaction. Long-range effects represented by core-polarization terms are also needed. In order to achieve this, the $`G`$-matrix elements are renormalized by the $`\widehat{Q}`$-box method. The $`\widehat{Q}`$-box is made up of non-folded diagrams which are irreducible and valence linked. Here we include all non-folded diagrams to third order in $`G`$ . Based on the $`\widehat{Q}`$-box, we compute an effective interaction $`\stackrel{~}{H}`$ in terms of the $`\widehat{Q}`$-box, using the folded-diagram expansion method, see e.g., Ref. for further details.
The effective two-particle interaction can in turn be used in large-scale shell model calculations. The shell model problem requires the solution of a real symmetric $`n\times n`$ matrix eigenvalue equation $`\stackrel{~}{H}|\mathrm{\Psi }_k=E_k|\mathrm{\Psi }_k`$ , with $`k=1,\mathrm{},K`$. At present our basic approach to finding solutions to this equation is the Lanczos algorithm; an iterative method which gives the solution of the lowest eigenstates. The technique is described in detail in Ref. .
In our studies the shell-model space consists of the orbits $`2s_{1/2}`$, $`1d_{5/2}`$, $`1d_{3/2}`$, $`0g_{7/2}`$ and $`0h_{11/2}`$, for both protons and neutrons. For the studies of the Sb isotopes we used <sup>100</sup>Sn as closed shell core, while for the heavy tin isotopes we derived an effective interaction with <sup>132</sup>Sn as closed shell core, see e.g., Refs. for calculational details. The dimensionality $`n`$ of the eigenvalue matrix $`\stackrel{~}{H}`$ is increasing with increasing number of valence particles or holes. As an example, for <sup>116</sup>Sn the dimensionality of the hamiltonian matrix is of the order of $`n\times 10^8`$.
## 3 Selected features of tin isotopes
Of interest in this study is the fact that the chain of even tin isotopes from <sup>102</sup>Sn to <sup>130</sup>Sn exhibits a near constancy of the $`2_1^+0_1^+`$ excitation energy, a constancy which can be related to strong pairing correlations and the near degeneracy in energy of the relevant single particle orbits. As an example, we show the experimental<sup>1</sup><sup>1</sup>1We will limit our discussion to even nuclei from <sup>116</sup>Sn to <sup>130</sup>Sn, since a qualitatively similar picture is obtained from <sup>102</sup>Sn to <sup>116</sup>Sn. $`2_1^+0_1^+`$ excitation energy from <sup>116</sup>Sn to <sup>130</sup>Sn in Table 1. Our aim is to see whether partial waves which play a crucial role in superfluidity of neutron star matter , viz., $`{}_{}{}^{1}S_{0}^{}`$ and $`{}_{}{}^{3}P_{2}^{}`$, are equally important in reproducing the near constant spacing in the chain of even tin isotopes shown in Table 1.
In order to test whether the $`{}_{}{}^{1}S_{0}^{}`$ and $`{}_{}{}^{3}P_{2}^{}`$ partial waves are equally important in reproducing the near constant spacing in the chain of even tin isotopes as they are for the superfluid properties of infinite neutron star matter, we study four different approximations to the shell-model effective interaction, viz.,
1. Our best approach to the effective interaction, $`V_{\mathrm{eff}}`$, contains all one-body and two-body diagrams through third order in the $`G`$-matrix, see Ref. .
2. The effective interaction is given by the $`G`$-matrix only and inludes all partial waves up to $`l=10`$.
3. We define an effective interaction based on a $`G`$-matrix which now includes only the $`{}_{}{}^{1}S_{0}^{}`$ partial wave.
4. Finally, we use an effective interaction based on a $`G`$-matrix which does not contain the $`{}_{}{}^{1}S_{0}^{}`$ and $`{}_{}{}^{3}P_{2}^{}`$ partial waves, but all other waves up to $`l=10`$.
In all four cases the same NN interaction is used, viz., the CD-Bonn interaction described in Ref. . Table 1 lists the results.
We note from this table that the three first cases nearly produce a constant $`2_1^+0_1^+`$ excitation energy, with our most optimal effective interaction $`V_{\mathrm{eff}}`$ being closest the experimental data. The bare $`G`$-matrix interaction, with no folded diagrams as well, results in a slightly more compressed spacing. This is mainly due to the omission of the core-polarization diagrams which typically render the $`J=0`$ matrix elements more attractive. Such diagrams are included in $`V_{\mathrm{eff}}`$. Including only the $`{}_{}{}^{1}S_{0}^{}`$ partial wave in the construction of the $`G`$-matrix (case 3), yields in turn a somewhat larger spacing. This can again be understood from the fact that a $`G`$-matrix constructed with this partial wave only does not receive contributions from any entirely repulsive partial wave. It should be noted that our optimal interaction, as demonstrated in Ref. , shows a rather good reproduction of the experimental spectra for both even and odd nuclei. Although the approximations made in cases 2 and 3 produce an almost constant $`2_1^+0_1^+`$ excitation energy, they reproduce poorly the properties of odd nuclei and other excited states in the even Sn isotopes.
However, the fact that the first three approximations result in a such a good reproduction of the $`2_1^+0_1^+`$ spacing may hint to the fact that the $`{}_{}{}^{1}S_{0}^{}`$ partial wave is of paramount importance. If we now turn the attention to case 4, i.e., we omit the $`{}_{}{}^{1}S_{0}^{}`$ and $`{}_{}{}^{3}P_{2}^{}`$ partial waves in the construction of the $`G`$-matrix, the results presented in Table 1 exhibit a spectroscopic catastrophe. We do also not list eigenstates with other quantum numbers. For e.g., <sup>126</sup>Sn the ground state is no longer a $`0^+`$ state, rather it carries $`J=4^+`$ while for <sup>124</sup>Sn the ground state has $`6^+`$. The first $`0^+`$ state for this nucleus is given at an excitation energy of $`0.1`$ MeV with respect to the $`6^+`$ ground state. The general picture for other eigenstates is that of an extremely poor agreement with data. Since the agreement is so poor, even the qualitative reproduction of the $`2_1^+0_1^+`$ spacing, we defer from performing time-consuming shell-model calculations for <sup>116,118,120,122</sup>Sn.
## 4 Shell model studies of the proton drip line nuclei <sup>105,106,107</sup>Sb
We present recent results for <sup>105</sup>Sb, <sup>106</sup>Sb and <sup>107</sup>Sb in Table 2. The calculations use <sup>100</sup>Sn as closed shell core with an effective interaction for the four, five or six valence neutrons and one valence proton based on the CD-Bonn nucleon-nucleon interaction . The experimental spin assignements for <sup>105</sup>Sb and <sup>106</sup>Sb are tentative. There are also many more theoretical states than reported in the enclosed table.
The high spin level scheme of <sup>105</sup>Sb resembles the level scheme of <sup>107</sup>Sb up to $`J=19/2`$ . When compared to <sup>107</sup>Sb the <sup>105</sup>Sb level scheme shows similar trends as when going from <sup>106</sup>Sn to <sup>104</sup>Sn. That means that coupling a $`d_{5/2}`$ proton to a <sup>104</sup>Sn core is appropriate to describe the observed states. The calculation favors $`J^\pi `$=5/2<sup>+</sup> for the ground state in agreement with the suggestion from proton decay data. In this state the valence proton is mainly in the $`d_{5/2}`$ orbit and the two neutron pairs are almost evenly distributed over the $`d_{5/2}`$ and $`g_{7/2}`$ neutron orbits. The situation is very similar in the 9/2<sup>+</sup> and 13/2<sup>+</sup> states, while the $`\nu `$$`g_{5/2}^3`$$`g_{7/2}^1`$ configuration exhausts the largest parts of the wave functions of the 15/2<sup>+</sup> and 17/2<sup>+</sup> states. The neutron part of the wave function of the 19/2<sup>+</sup> state is almost identical to the 17/2<sup>+</sup> state. However, since 17/2<sup>+</sup> is the maximum spin for the $`\pi `$$`g_{5/2}^1`$$`\nu `$$`g_{5/2}^1`$$`g_{7/2}^1`$ configuration, the odd proton resides almost exclusively in the $`g_{7/2}`$ orbit in the 19/2<sup>+</sup> state. For proton degrees of freedom the $`s_{1/2}`$, $`d_{3/2}`$ and $`h_{11/2}`$ single-particle orbits give essentially negligible contributions to the wave functions and the energies of the excited states, as expected. For neutrons, although the single-particle distribution for a given state is also negligible, these orbits are important for a good describtion of the energy spectrum, as also demonstrated in large-scale shell-model calculations of tin isotopes . Similar picturer applies to <sup>106</sup>Sb and <sup>107</sup>Sb as well, see Refs. . The wave functions for the various states are to a large extent dominated by the $`g_{7/2}`$ and $`d_{5/2}`$ single-particle orbits for neutrons ($`\nu `$) and the $`d_{5/2}`$ single-particle orbit for protons ($`\pi `$). The $`\nu g_{7/2}`$ and $`\nu d_{5/2}`$ single-particle orbits represent in general more than $`90\%`$ of the total neutron single-particle occupancy, while the $`\pi d_{5/2}`$ single-particle orbits stands for $`8090\%`$ of the proton single-particle occupancy. The other single-particle orbits play an almost negligible role in the structure of the wave functions.
## 5 Conclusions
In summary, shell-model calculations with realistic effective interactions of the newly reported low-lying yrast states of the proton drip line nuclei <sup>105,106,107</sup>Sb, reproduce well the experimental data. Since the wave functions of the various states are to a large extent dominated by neutronic degrees of freedom and neutrons are well bound with a separation energy of $`8`$ MeV, this may explain why a shell-model calculation, within a restricted model space for a system close to the proton drip line, gives a satisfactory agreement with the data.
For the Sn isotopes we have shown that the $`{}_{}{}^{1}S_{0}^{}`$ and $`{}_{}{}^{3}P_{2}^{}`$ partial waves, which are crucial for our understanding of superfluidity in neutron star matter, are equally important in order to reproduce the $`2_1^+0_1^+`$ excitation energy of the even Sn isotopes. Omitting these waves, especially the $`{}_{}{}^{1}S_{0}^{}`$ wave, results in a spectrum which has essentially no correspondence with experiment.
Further analysis of nuclei such as Ag, Cd, In near $`A100`$ and Sn isotopes near $`A132`$ are in progress.
We are much indebted to Cyrus Baktash, David Dean, Hubert Grawe and Matej Lipoglavšek for many discussions on properties of nuclei near $`A100`$.
|
warning/0007/astro-ph0007463.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The study of in the formation structure in the Universe has given a renewal interest with the advent and success of extraterrestrial experiments like the Cosmic Background Explorer (COBE) and its data .Recently the investigation of the density perturbations in alternative theories of gravity such as Brans-Dicke due to the presence of scalar inflaton or dilatonic fields in inflationary scenario have been undertaken by Fabris and his group .Besides computation of density perturbations in global textures in the realm of Newtonian cosmology have been investigated by Ribeiro and Letelier .Earlier Palle and myself have independently investigate the idea of primordial density flucuations in Einstein-Cartan gravity and show that Einstein-Cartan cosmology is compatiible with COBE data.Nevertheless Palle has not continued the investigation of the primordial fluctuations providing specific examples that could be matched with future experimental data base of future experiments such as the Boomerang and other experimental devices placed on the Earth orbit.In this paper we propose to fill this gap.The paper is organized as follows:In section 2 we give a simple derivation of the evolution of density perturbations and its dynamics.In section 3 we give twoo simple examples,the first is of a two fluid Friedmann Universes embbeded on each other where one does not contain spinning particles and the other can be analyzed using Einstein-Cartan gravity and the other is a two fluid this time not closed universes as before but spatially flat and possesing dilatonic fields.In both cases the density perturbations are expressed in terms of the redshift parameter z and we show that spin-torsion density contribution is stronger when the redshifts are higher.Data from elliptical Galaxies and globular cluster are also used to provide bounds for the density perturbations and spin contributions to them.In section 4 gravitational stability of the Friedmann metric in Einstein-Cartan gravity is also discussed since this is a important topic in Galaxy formation.Besides a lower boud to the spin-torsion fluctuation is found from this analysis and COBE data.
## 2 Evolution Equation of Density Perturbation
Investigation of linear density fluctuations in the context of general relativistic cosmology have proved to be useful in the study of structure formation like galaxy formation for example.Later on we developed some of his ideas to solutions of the Einstein-Cartan cosmology with inflatons and dilatons.In this section we show that the evolution of density fluctuations can be derived from the Einstein-Cartan field equations and the associated conservation equation for the matter and spin-torsion density in analogous way that it is done in General Relativity as long as some simple assumptions are made on the galatic spinning fluid.A stationary metric here is not needed since we are not considering that the spin of the intrinsic particles affect appreciably the rotation of galaxies.Therefore the usual Friedmann metric to investigate linear perturbations in Einstein-Cartan gravity.Let us begin by considering the Friedmann metric
$$ds^2=dt^2a^2(dx^2+dy^2+dz^2)$$
(1)
where $`a(t)`$ is the cosmic scale.The Einstein-Cartan equations are given by
$$H^2=\frac{8\pi G}{3}(\rho _{eff}2\pi G\sigma ^2)$$
(2)
and
$$H^2+\dot{H}=\frac{4\pi G}{3}(\rho _{eff}8\pi G\sigma ^2)$$
(3)
where $`H=\frac{\dot{a}}{a}`$ is the Hubble parameter in terms of time.Where in short we use the following notations
$$\rho _{eff}=\dot{\varphi }^2+V$$
(4)
and
$$p_{eff}=\dot{\varphi }^2V$$
(5)
where $`\varphi `$ is the inflaton field ,$`V`$ is the inflaton potential and $`\sigma ^2=<S_{ij}S^{ij}>`$ is the averaged squared of the spin density tensor $`S_{ij}`$. By making use of the definition $`\rho =\rho _{eff}2\pi G\sigma ^2`$ and $`p=p_{eff}2\pi G\sigma ^2`$ allow us to write the conservation equation in Einstein-Cartan Gravity as
$$\frac{d}{dR}(\rho R^3)=3pR^2$$
(6)
These substitutions allow us to reduce our problem of computing density perturbations formally similar to the general relativistic one.The necessary assumption to make our task easier is to consider the pressure $`p`$ vanishes which yields
$$p_{eff}=2\pi G\sigma ^2$$
(7)
With these assumptions equation (6) reduces to
$$\frac{d}{dt}(\rho R^3)=0$$
(8)
where now we are ready to apply the traditional perturbation method on the cosmological density perturbation by making use of the following definitions
$$\delta =\frac{\rho \rho _b}{\rho _b}$$
(9)
and
$$a=R(t)+\delta R(t)$$
(10)
and
$$\sigma ^2=\sigma _b^2(1+\frac{\delta \sigma ^2}{\sigma ^2})$$
(11)
Here the index b denotes background quantities.From the equation (8) one obtains
$$\delta =3\frac{\delta R}{R}$$
(12)
following the lines and procedures of General Relativity we find the following evolution equation yields
$$\ddot{\delta }+2H_{0}^{}{}_{}{}^{2}\dot{\delta }+10\pi ^2G^2\sigma _{b}^{}{}_{}{}^{2}\delta =0$$
(13)
To solve this differential equation we need to compute the background spin-torsion density.This can be easily accomplished if one equates equations (2) and (3) and substitute (7) to obtain
$$\rho _{eff}=\frac{9}{2}\pi G\sigma _{b}^{}{}_{}{}^{2}$$
(14)
Substitution of this last expression into the conservation equation for the effective density yields
$$\dot{\sigma _{b}^{}{}_{}{}^{2}}+3H_0\sigma _{b}^{}{}_{}{}^{2}=0$$
(15)
which solution is
$$\sigma _{b}^{}{}_{}{}^{2}=e^{3H_0t}$$
(16)
to simplify matters without loosing too much physical insight one could assume that the spin-torsion background can be expanded in the form
$$\sigma _{b}^{}{}_{}{}^{2}=(1+3H_0t)^1=(3H_0t)^1$$
(17)
where we have made the hypothesis $`H_0t>>\frac{1}{3}`$.Substitution of this result into expression (13) yields the final form of the evolution equation ready now to be solved
$$\ddot{\delta }+2\alpha \dot{\delta }+\gamma \frac{\delta }{t}=0$$
(18)
where $`\alpha =2H_{0}^{}{}_{}{}^{2}`$ and $`\gamma =\frac{10\pi G}{3H_0}`$.To deduce the evolution equation (18) we assumed that the perturbed metric is the de Sitter metric where $`H_0`$ is a constant.To solve this evolution equation let us assume as usual a solution of the type $`\delta =t^m`$ where $`m`$ is a real constant to be determined from an algebraic equation.Substitution of this hint into the differential equation (18) yields the following algebraic equation
$$m^2+(1+\alpha )m+\gamma =0$$
(19)
the solution of this last equation left us with two solutions of the type $`\delta _+=t^{m_+}`$ and $`\delta _{}=t^m_{}`$ where $`m_+=\frac{5\pi G}{3H_0}`$ and $`m_{}=(1+H_{0}^{}{}_{}{}^{2})`$ are the respective elementary solutions of the algebraic equation (19). The spinning fluid of the Early Universe consists here of protons and neutrons which are fermions producing the torsion of spacetime.The fluid this time is not a dilatonic fluid and there is no potential or inflaton.The resulting evolution euqtion is simply the general relativistic equation
$$\ddot{\delta }+2\frac{\dot{R}}{R}\dot{\delta }4\pi ^2G\rho _{b}^{}{}_{eff}{}^{}\delta =0$$
(20)
where now $`\rho _{b}^{}{}_{eff}{}^{}=\rho _b2\pi G\sigma _{b}^{}{}_{}{}^{2}`$.To be able to express the evolution equation in terms of the cosmic time it is enough to remenber taht the mass density of the background is given by $`\rho _b=\frac{M}{R^3}`$ and the spin-torsion density can be obtained remenbering that the spin-torsion density tensor can be written as $`|S_{ij}|=S=\frac{nh}{R^3}`$ where n is the number of fermions and h is the Planck constant.The Friedmann equation thus becomes
$$\dot{R}^2=\frac{8\pi G}{3}(\frac{M}{R}\frac{2\pi GS^2}{R^4})$$
(21)
allow us to obtain explicitly the dependence of the cosmic scale R with time.At this point since we are mainly interested in the extremum case of the role played by torsion and spin on Galaxy formation we assume here that the first term may be dropped,(as could happens inside black holes for example) which reduces the above equation to
$$\dot{R}^2=\frac{16\pi ^2G^2S^2}{3R^6}$$
(22)
Substitution of $`R=t^p`$ where p is a real number into the equation (22) one obtains the following algebraic equation
$$4p^24p+1=0$$
(23)
Solution of this equation yields $`R(t)=t^{\frac{1}{2}}`$.Substitution of this result into formula (22) yields
$$\ddot{\delta }+2\frac{\dot{\delta }}{t}+16\pi ^2G^2\frac{n^2h^2}{3t^3}\delta =0$$
(24)
Substitution of the ansatz $`\delta =t^m`$ into equation (28) yields
$$m^2+3m4\pi GM=0$$
(25)
which yields the following solutions $`\delta _+=t^3`$ and $`\delta _{}=1`$ which possess the interesting physical feature that the second equation coincides with the general relativistic solution while the first solution seems to be a carachteristic of the Einstein-Cartan Cosmology.Our result also shows that the gravitational instability is enhanced by torsion instead of holding it.This result have already been expected by the analysis done by Hehl et al. many years ago.
## 3 Density perturbation in Spin cosmology
When D.Palle computed the evolution equation of density perturbation for a Bianchi type III model in Einstein-Cartan cosmology.It is well know in standard general relativistic cosmology that specially in the particular case of Newtonian cosmology simple models can be designed to obtain the density perturbations without solving the evolution equation.In this note we propose a two fluid very simple model in Einstein-Cartan gravity to compute the cosmological density perturbations without solving the evolution equation of density perturbations.In the absence of spin our results reduce to the ones in General Relativity .Dominance of the spin-torsion density in the early epochs of the Universe occurs in the case of density perturbations for the matter phase while at the present epoch spin-torsion stronger contribution is to the density perturbations is on radiation era.This fact seems to be physically explainable since the early epochs of the Universe are carachterize by the fact that the matter-radiation coupling is much stronger than in the present day Universe.Let us start by considering the Friedmann metric
$$ds^2=dt^2a^2(t)(\frac{dr^2}{1kr^2}+r^2(d\theta ^2+sin^2\theta d\varphi ^2))$$
(26)
where $`a(t)`$ is connected to the Hubble parameter H by the relation $`H=\frac{\dot{a}}{a}`$ and k is the spatial curvature constant which is $`k=1,0,+1`$ according the 3-space is open,flat or closed respectively.Considering a spherical region of radius $`\lambda >d_H`$ which is the Hubble radius containing spinning matter with mean density $`\rho _1`$,embedded in a $`k=0`$ Friedmann universe $`\rho _0`$ where $`\rho _1=\rho _0+\delta \rho `$, $`\delta \rho >0`$ and represents a small density perturbation.As point it out by Padmanabhan the spherical symmetry implies that the inner region is not affected by the matter outside.Thus the inner region evolves as a $`k=+1`$ Friedmann Universe .The two regions form a kind of two fluid the inner fluid being a spinning fluid obtained from the perturbation.Since the inner fluid is a spin fluid it may obey the Einstein-Cartan gravity with the following form
$$H_{1}^{}{}_{}{}^{2}+\frac{1}{a_{}^{2}{}_{1}{}^{}}=\frac{8\pi G}{3}(\rho _12\pi G\sigma ^2)$$
(27)
where $`\sigma ^2`$ represents the averaged squared value of the spin-torsion tensor.The outer fluid obey the GR equation
$$H_{0}^{}{}_{}{}^{2}=\frac{8\pi G}{3}(\rho _0)$$
(28)
where both universes are compared, the perturbed and the background universe when their expansion rates are equal are equal,or we compare their densities at the time t when $`H_1=H_0`$.Since
$$\frac{\delta \rho }{\rho _0}=\frac{\rho _1\rho _0}{\rho _0}$$
(29)
Performing the above substitution we obtain
$$\frac{\delta \rho }{\rho _0}=\frac{1}{8\pi Ga_{}^{2}{}_{1}{}^{}\rho _0}+\frac{2\pi G\sigma ^2}{\rho _0}$$
(30)
To compute this equation in terms of the cosmic expansion $`a`$ we simply remember that the spin-torsion density is proportional to $`\frac{S_{0}^{}{}_{}{}^{2}}{a^6}`$.where $`S_{0}^{}{}_{}{}^{2}=n^2h^2`$ is the spin tensor squared and n is equal to the number of nucleons in the Universe and h is the Planck constant.Substituting this value into the equation (30) and considering that $`a_1`$ is approximatly equal to $`a_0`$ allows us to see how the $`\frac{\delta \rho }{\rho _0}`$ scales with $`a`$.Since in the radiation dominated era $`\rho _0\alpha a^4`$ and $`\rho _0\alpha a^3`$ in the matter dominated phase, substitution of these values into the equation (30) yields respectively
$$\frac{\delta \rho }{\rho _0}|_R=\frac{a^2}{8\pi G}+\beta GS_{0}^{}{}_{}{}^{2}a^2$$
(31)
and
$$\frac{\delta \rho }{\rho _0}|_M=\frac{a}{8\pi G}+\beta GS_{}^{2}{}_{0}{}^{}a^3$$
(32)
others zero.From the dependence of the spin-torsion term on the coosmic factor $`a`$ we already note that the spin-torsion contributions would be redshifted with expansion.Here $`\beta `$ is a constant.To write down these expressions in terms of cosmic time to be able to check the observational results with our computations we need to solve write 36) in terms of the cosmic scale factor a in terms of time.Solution of this equation in terms of time yields the following results $`a|_R\alpha t^{\frac{1}{2}}`$ and $`a|_M\alpha t^{\frac{2}{3}}`$.Substitution of these values into the expressions (31) and (32) we obtain respectively
$$\frac{\delta \rho }{\rho _0}|_R=\frac{t}{8\pi G}+\beta GS_{0}^{}{}_{}{}^{2}t^1$$
(33)
and
$$\frac{\delta \rho }{\rho _0}|_M=\frac{t^{\frac{2}{3}}}{8\pi G}+\beta GS_{0}^{}{}_{}{}^{2}t^2$$
(34)
Notice therefore that in the case of early epochs in the Universe where $`t0`$ the spin-torsion effects appear to be stronger on matter perturbation contributing to the decoupling of matter and radiation this is explained since the torsion acts only on fermions and do not clearly interacts with radiation.Nevertheless in the present epoch the spin-torsion decays very quick which is the reason the detection of torsion is difficult froom the cosmological point of view nowadays.To have a still more simple vision of the problem we express the formula for the density perturbation of the matter in terms of the redshift z.This can be done remenber the horizon problem and using the formula
$$a_H=a(t)\frac{dt^{}}{a(t^{})}$$
(35)
and taking the value $`a\alpha t^{\frac{2}{3}}`$ obtained previously for the matter case and substituting it to expression (35) and performing the integration one obtains $`a_H=3t`$.In turn substitution of this result into the expression for the age of the Universe one obtains
$$t=\frac{1}{3}a_H=\frac{2H_0^1}{(1+z)^{\frac{3}{2}}}$$
(36)
and finally substitution of this value into equation (34)
$$\delta (z)|_M=\frac{\delta \rho }{\rho _0}|_M=\frac{(2H_0^1)^{\frac{2}{3}}}{8\pi G(1+z)}+\frac{H_0^3\beta GS_{0}^{}{}_{}{}^{2}(1+z)^3}{8}$$
(37)
From this last expresion is easy to check that spin-torsion effects are dominant for high redshifts and since the higher redshift objets are far away from us one may realize that spin-torsion effects are redshifted when the Universe expands for example in the case of inflationary cosmology.Specially strong effects can be obtained at the decoupling time where $`z=10^3`$.At this point let us consider the other example of flat spatially cosmological models embedded in each other and obeying the following field equations
$$H_{1}^{}{}_{}{}^{2}=\frac{8\pi G}{3}(\rho _1+\frac{1}{2}\dot{\varphi }^22\pi G\sigma ^2)$$
(38)
and
$$H_{0}^{}{}_{}{}^{2}=\frac{8\pi G}{3}(\rho _0)$$
(39)
After a simple algebra one obtains the expression for the density perturbation
$$\frac{\delta \rho }{\rho _0}=\frac{2\pi G\sigma ^2}{\rho _0}\frac{1}{2}\frac{\dot{\varphi }^2}{\rho _0}$$
(40)
Since from the already known solution of de Sitter inflationary cosmology with spin-torsion density we have
$$\dot{\varphi }^2=\pi G\sigma ^2$$
(41)
Substitution of this last expression into the density perturbation reduces this formula to
$$\frac{\delta \rho }{\rho _0}=\frac{\frac{3}{2}\pi G\sigma ^2}{\rho _0}$$
(42)
Now taking into account the behaviour of matter phase $`\rho _0\alpha a^3`$ and the relation between the redshift and the cosmic parameter $`a`$ as $`a=(1+z)^1`$ one obtains
$$\frac{\delta \rho }{\rho _0}=\frac{\frac{3}{2}\pi G\sigma ^2}{(1+z)^3}$$
(43)
This expression allow us to give numerical estimates to the density perturbations as far as spin-torsion effects are concerned.Since from de Sabbata and Sivaram we know that at the Planck era we have $`\frac{G\sigma ^2}{c^4}=10^{87}`$ and that at the Planck era $`t_{Pl}=10^{43}s`$ the expression
$$\frac{\delta \rho }{\rho _0}=\frac{3}{2}\pi G\sigma _{Pl}^{}{}_{}{}^{2}t_{Pl}^{}{}_{}{}^{2}=10>1$$
(44)
and the linear perturbation density approximation would break down and nonlinear perturbation theory would have to be used.On the other hand making use of the spin-torsion critical value $`\frac{G\sigma ^2}{c^4}=10^{30}`$ and the value of the redshift at the decoupling between matter and radiation $`z=10^3`$ one obtains the value of
$$\frac{\delta \rho }{\rho _0}=\frac{\frac{3}{2}\pi G\sigma ^2}{(1+z)^3}=10^{33}$$
(45)
which is an extremely low value for being detected by the COBE experimental devices.Nevertheless an estimate within the COBE capabilities may be obtained by making use of two Astronomical data, namely the spiral Galaxies and cluster energy densities of $`\rho _0=10^{25}g.cm^3`$ and $`\rho _0=10^{28}g.cm^3`$ respectively which yield for the spin-torsion contribution values of
$$\frac{\delta \rho }{\rho _0}=10^5$$
(46)
and
$$\frac{\delta \rho }{\rho _0}=10^2$$
(47)
well within the COBE data measurement capabilities.Here we also used the above critical spin-torsion density value.
## 4 Gravitational Stability of Friedmann Metric in Spin Cosmology
Now we turn to the problem of the stability of the Friedmann metric in Einstein-Cartan cosmology.Nurgaliev and Ponomariev investigated the early evolutionary stages of the Universe in the realm of Einstein-Cartan cosmology by considering an ideal fluid of nonpolarised fermions.They show that for this particular case the Friedmann solution was stable to small homogeneous and isotropic perturbations.They also conclude that the increase in entropy could lead to the evolution of the initially small stable oscillations into large ones.The Universe would begin after a definite number of oscillations.Calculations of small perturbations in their model would shown some instability which would prepare for the necessary initial conditions for the growth of pertubations in the nonlinear regime.Now we make use of their result to olace a lower limit to the spin-torsion primordial density fluctuation obtained from the Einstein-Cartan gravity and COBE satellite data .In this way structure formation like Galaxies formation would not suffer a great influence from torsion .It is also shown that at early stages of the Universe Bianchi type III models with expansion and rotation may not depend at all from torsion.Other Bianchi types like an oscillating Bianchi type IX model in Einstein-Cartan gravity have also been recently investigated.The Friedmann equation thus becomes
$$\frac{\ddot{a}}{a}=\frac{4\pi G}{3}(\rho 8\pi G\sigma ^2)$$
(48)
Making an homogeneous and isotropic small perturbation on the Friedmann yields
$$\frac{\delta \ddot{a}}{\delta a}=\frac{4\pi G}{3}(\rho 8\pi G\sigma ^2)\frac{4\pi G}{3}(\frac{\delta \rho }{\delta a}8\pi G\frac{\delta \sigma ^2}{\delta a})$$
(49)
Substitution of the well-known relation
$$\frac{\delta \rho }{\delta a}=3\frac{\rho }{a}$$
(50)
we obtain
$$\frac{\delta \ddot{a}}{\delta a}=\frac{4\pi G}{3}(\rho 8\pi G\sigma ^2)+\frac{4\pi G}{3}(\frac{\rho }{a}+8\pi G\frac{\delta \sigma ^2}{\delta a})$$
(51)
and the stability condition $`\frac{\delta \ddot{a}}{\delta a}<0`$ implies
$$(18\pi G\frac{\delta \sigma ^2}{\delta \rho })<0$$
(52)
Since the matter density $`\rho >0`$ this implies the following condition
$$\delta \sigma ^2>\frac{1}{8\pi G}\frac{\delta \rho }{\rho _0}\rho _0$$
(53)
where we take$`\rho _0=10^{31}gcm^3`$ as the matter density of the Universe and from the COBE data $`\frac{\delta \rho }{\rho _0}=10^5`$.From these data formula (53) yields the following lower limit for the spin-torsion fluctuation as
$$\delta \sigma ^2>10^{28}cgsunits$$
(54)
this result was expected since the spin-torsion density decreases with the expansion and is redshifted with inflation.This conjecture has been proposed recently by Ramos and myself .As pointed out by Nurgaliev and Ponomariev the increase in the entropy may trigger the growth in the inhomogeneities.There is no compelling reason to believe that this would not happen here.Moreover Nurgaliev and Piskareva have also investigate the structural stability of cosmological models in Einstein-Cartan gravity.A more detailed investigation of the matters discussed here including the Bianchi type IX oscilating solution in Einstein-Cartan cosmology may appear elsewhere.
## Acknowledgement
I am very much indebt to Professors P.S.Letelier,I.Shapiro and my colleague Rudnei de Oliveira for helpful discussions on the subject of this paper.Thanks are also due to an unknown referee for useful comments.Special thanks go to Dr.Andre Ribeiro for helping me with Astronomical data.Financial support from CNPq. and UERJ is gratefully acknowledged.
|
warning/0007/hep-ph0007251.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
In recent years theory necessary for precise description of a heavy quarkonium such as bottomonium or (remnant of) toponium has developed significantly. In particular, development in the computational technology of higher order corrections to the quarkonium energy spectrum and subsequent discovery of the renormalon cancellation enabled accurate determinations of the $`\overline{\mathrm{MS}}`$-mass of the bottom quark and (in the future) of the top quark . In these determinations a major part is played by the spectrum (mass) of the $`1S`$-state quarkonium.<sup>*</sup><sup>*</sup>* See e.g. for introductory reviews of the subject.
It is legitimate to consider that the present perturbative calculation of the quarkonium spectrum, when expressed in terms of the quark $`\overline{\mathrm{MS}}`$-mass, has a genuine accuracy at $`𝒪(\alpha _S^3m)`$. In fact, in formal power countings, the last known term in the relation between the $`\overline{\mathrm{MS}}`$-mass and the pole-mass of a quark is $`𝒪(\alpha _S^3m)`$, while the last known term of the binding energy (measured from twice of the quark pole-mass) is $`𝒪(\alpha _S^4m)`$. The former term includes in addition to a genuine $`𝒪(\alpha _S^3m)`$ part the leading renormalon contribution which does not become smaller than $`𝒪(\mathrm{\Lambda }_{\mathrm{QCD}})`$ . This renormalon contribution is cancelled against the renormalon contribution contained in the latter term. Therefore, after cancellation of the leading renormalons, the genuine $`𝒪(\alpha _S^3m)`$ part of the mass relation determines the accuracy of the present perturbation series relating the quark $`\overline{\mathrm{MS}}`$-mass and the quarkonium spectrum.
As stated, for the binding energy the calculation including the genuine $`𝒪(\alpha _S^4m)`$ corrections has already been completed. Also, the “large-$`\beta _0`$ approximation” is known to be a pragmatically feasible and empirically successful estimation method of the leading renormalon contributions. Taking these into account, one finds that it is sufficient There exist other methods in which the renormalon contribution contained in the pole-mass is subtracted in certain approximations (e.g. ). In our opinion our method is a most natural one, embedded in the algorithm for higher order calculations of the whole spectrum; cancellation of infrared sensitivities in the whole spectrum follows from the fact that the quarkonium is a color-singlet small-size system. to calculate further the following two corrections in order to improve the accuracy of the spectrum by one order and to achieve a genuine accuracy at $`𝒪(\alpha _S^4m)`$: (I) the $`𝒪(\alpha _S^4m)`$ relation between the $`\overline{\mathrm{MS}}`$-mass and the pole-mass, and (II) the binding energy at $`𝒪(\alpha _S^5m)`$ in the large-$`\beta _0`$ approximation. This is because the leading renormalon contribution in the full $`𝒪(\alpha _S^5m)`$ correction to the binding energy will be incorporated by the large-$`\beta _0`$ approximation and the remaining part is expected to be irrelevant at $`𝒪(\alpha _S^4m)`$.
Of these two corrections we calculate (II) analytically for the $`1S`$-state in this paper. Then we study its phenomenological applications. We can check validity of the above general argument explicitly at $`𝒪(\alpha _S^3m)`$ where we know the exact result. This will also be demonstrated.
Written in terms of the quark pole-mass $`m_{\mathrm{pole}}`$, the mass of the quarkonium $`1S`$-state is given as a series expansion in the $`\overline{\mathrm{MS}}`$ coupling constant $`\alpha _S(\mu )`$ defined in the theory with $`n_l`$ massless quarks:
$`M_{1S}=2m_{\mathrm{pole}}{\displaystyle \frac{4}{9}}\alpha _S(\mu )^2m_{\mathrm{pole}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{\alpha _S(\mu )}{\pi }}\right)^nP_n(L),`$ (1)
where $`P_n(L)`$ is an $`n`$-th-degree polynomial of $`L\mathrm{log}[3\mu /(4\alpha _S(\mu )m_{\mathrm{pole}})]`$. At each order of the perturbative expansion $`c_n=P_n(0)`$ represents a non-trivial correction, while the coefficients of $`L`$’s are determined by the renormalization-group equation. For $`n3`$, $`c_n`$’s include powers of $`\mathrm{log}\alpha _S`$ unrelated to the renormalization group, i.e. which is not accompanied by $`\mathrm{log}\mu `$ . The polynomials relevant to our analysis read
$`P_0`$ $`=`$ $`1,`$ (2)
$`P_1`$ $`=`$ $`\beta _0L+c_1,`$ (3)
$`P_2`$ $`=`$ $`{\displaystyle \frac{3}{4}}\beta _0^2L^2+\left({\displaystyle \frac{1}{2}}\beta _0^2+{\displaystyle \frac{1}{4}}\beta _1+{\displaystyle \frac{3}{2}}\beta _0c_1\right)L+c_2,`$ (4)
$`P_3`$ $`=`$ $`{\displaystyle \frac{1}{2}}\beta _0^3L^3+\left({\displaystyle \frac{7}{8}}\beta _0^3+{\displaystyle \frac{7}{16}}\beta _0\beta _1+{\displaystyle \frac{3}{2}}\beta _0^2c_1\right)L^2`$ (5)
$`+\left({\displaystyle \frac{1}{4}}\beta _0^3{\displaystyle \frac{1}{4}}\beta _0\beta _1+{\displaystyle \frac{1}{16}}\beta _2{\displaystyle \frac{3}{4}}\beta _0^2c_1+{\displaystyle \frac{3}{8}}\beta _1c_1+2\beta _0c_2\right)L+c_3.`$
$`\beta _n`$’s denote the coefficients of the QCD beta function given by
$`\beta _0=11{\displaystyle \frac{2}{3}}n_l,\beta _1=102{\displaystyle \frac{38}{3}}n_l,\beta _2={\displaystyle \frac{2857}{2}}{\displaystyle \frac{5033}{18}}n_l+{\displaystyle \frac{325}{54}}n_l^2.`$ (6)
Note that, in the renormalization-group evolution, running of the coupling included in $`\mathrm{log}\alpha _S`$ should be taken into account properly. This feature is unique to the perturbation series of a nonrelativistic boundstate spectrum. Presently $`c_n`$’s are known up to $`n=2`$ :
$`c_1`$ $`=`$ $`{\displaystyle \frac{97}{6}}{\displaystyle \frac{11}{9}}n_l,`$ (7)
$`c_2`$ $`=`$ $`{\displaystyle \frac{1793}{12}}+{\displaystyle \frac{2917\pi ^2}{216}}{\displaystyle \frac{9\pi ^4}{32}}+{\displaystyle \frac{275\zeta _3}{4}}+\left({\displaystyle \frac{1693}{72}}{\displaystyle \frac{11\pi ^2}{18}}{\displaystyle \frac{19\zeta _3}{2}}\right)n_l+\left({\displaystyle \frac{77}{108}}+{\displaystyle \frac{\pi ^2}{54}}+{\displaystyle \frac{2\zeta _3}{9}}\right)n_l^2.`$
The aim of this paper is to calculate $`c_3`$ in the large-$`\beta _0`$ approximation.
One might think that alternatively $`c_3`$ may be estimated using the asymptotic form of the series expansion of the QCD potential at large orders. We could not, however, find a justification that $`c_3`$ can be estimated with $`𝒪(\alpha _S^4m)`$ accuracy in this manner.
## 2 Outline of the calculation
In the large-$`\beta _0`$ approximation the wave functions and energy spectra (measured from $`2m_{\mathrm{pole}}`$) of quarkonium states are determined by solving the nonrelativistic Schrödinger equation
$`\left[{\displaystyle \frac{\stackrel{}{p}^2}{m}}+V_{\beta _0}(r)\right]\psi (\stackrel{}{x})=E\psi (\stackrel{}{x}).`$ (9)
Here, $`m`$ denotes the mass of the quark and it is irrelevant whether we use the pole-mass or the $`\overline{\mathrm{MS}}`$-mass for $`m`$ within our approximation; in particular it does not affect the leading renormalon cancellation. $`V_{\beta _0}`$ denotes the QCD potential in the large-$`\beta _0`$ approximation given as a perturbation series
$`V_{\beta _0}(r)=C_F{\displaystyle \frac{\alpha _S(\mu )}{r}}\times R(r),R(r)={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{\beta _0\alpha _S(\mu )}{4\pi }}\right)^nR_n(\mu r).`$ (10)
$`C_F=4/3`$ is the color factor. First few coefficients $`R_n(\mu r)`$ are listed in Table 1.
To obtain $`S`$-wave solutions, we introduce a dimensionless variable $`z=C_F\alpha _Smr`$ and set
$`\psi ={\displaystyle \frac{1}{z}}\mathrm{exp}\left[{\displaystyle 𝑑zW(z)}\right],E=(C_F\alpha _S)^2m\epsilon .`$ (11)
Then the equation becomes
$`W^{}W^2R/z=\epsilon .`$ (12)
We expand $`W`$, $`R`$ and $`\epsilon `$ in perturbation series as in (10) and
$`W={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{\beta _0\alpha _S}{4\pi }}\right)^nW_n,\epsilon ={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{\beta _0\alpha _S}{4\pi }}\right)^n\epsilon _n,`$ (13)
and substitute them to (12). We find
$`W_n^{}{\displaystyle \underset{k=0}{\overset{n}{}}}W_kW_{nk}R_n/z=\epsilon _n.`$ (14)
For the $`1S`$-state the zeroth-order solution is given by
$`W_0={\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{z}},\epsilon _0={\displaystyle \frac{1}{4}},`$ (15)
and we may solve (14) recursively for $`n1`$:
$`\epsilon _n`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _0^{\mathrm{}}}𝑑tt^2e^t\left({\displaystyle \underset{k=1}{\overset{n1}{}}}W_k(t)W_{nk}(t)+R_n(t)/t\right),`$ (16)
$`W_n(z)`$ $`=`$ $`{\displaystyle \frac{e^z}{z^2}}{\displaystyle _z^{\mathrm{}}}𝑑tt^2e^t\left({\displaystyle \underset{k=1}{\overset{n1}{}}}W_k(t)W_{nk}(t)+R_n(t)/t+\epsilon _n\right).`$ (17)
Thus, $`\epsilon _n`$’s can be obtained by evaluating multiple integrals. We may render these integrals to forms which resemble Feynman-parameter integrals that appear in calculations of multiloop Feynman diagrams. In this way we could use various techniques developed for Feynman diagram calculations.<sup>§</sup><sup>§</sup>§ We find the techniques developed in particularly useful. We obtained the first two terms of the perturbative expansion as
$`\epsilon _1`$ $`=`$ $`(\stackrel{~}{L}+1),`$ (18)
$`\epsilon _2`$ $`=`$ $`(3\stackrel{~}{L}^2+4\stackrel{~}{L}+1+{\displaystyle \frac{\pi ^2}{6}}+2\zeta _3),`$ (19)
where $`\stackrel{~}{L}=L+\frac{5}{6}=\mathrm{log}[\mu /(C_F\alpha _Sm)]+\frac{5}{6}`$. From these we confirmed the corresponding parts of $`c_1`$ and $`c_2`$ in eqs. (7,LABEL:knownres2). We also obtained a new result:
$`\epsilon _3=\left[8\stackrel{~}{L}^3+10\stackrel{~}{L}^2+\left({\displaystyle \frac{4\pi ^2}{3}}+16\zeta _3\right)\stackrel{~}{L}2+\pi ^2+16\zeta _3+{\displaystyle \frac{\pi ^4}{90}}2\pi ^2\zeta _3+24\zeta _5\right].`$ (20)
It follows that
$`c_3(\text{large-}\beta _0)=\beta _0^3({\displaystyle \frac{517}{864}}+{\displaystyle \frac{19\pi ^2}{144}}+{\displaystyle \frac{11\zeta _3}{6}}+{\displaystyle \frac{\pi ^4}{1440}}{\displaystyle \frac{\pi ^2\zeta _3}{8}}+{\displaystyle \frac{3\zeta _5}{2}}).`$ (21)
We note two points: (i) The result includes the level-5 zeta values ($`\zeta _5`$ and $`\pi ^2\zeta _3`$). (ii) The result does not include $`\gamma _E`$, that is, all $`\gamma _E`$ which appeared at intermediate stages of the calculation got cancelled. A more detailed description of our calculation will be published elsewhere.
## 3 Phenomenological applications
We examine the series expansion of the quarkonium $`1S`$-state spectrum numerically for the bottomonium and (remnant of) toponium. The input value for the coupling defined in the theory with 5 massless flavors ($`n_l=5`$) is $`\alpha _S^{(5)}(M_Z)=0.119`$. We evolve the coupling and match it to the coupling of the theory with $`n_l=4`$ following . Also we take the input values of the $`\overline{\mathrm{MS}}`$-mass, $`\overline{m}m_{\overline{\mathrm{MS}}}(m_{\overline{\mathrm{MS}}})`$, and the pole-mass as $`\overline{m}_b=4.20`$ GeV/$`m_{b,\mathrm{pole}}=4.97`$ GeV and $`\overline{m}_t=165`$ GeV/$`m_{t,\mathrm{pole}}=174.79`$ GeV. The natural size of a quarkonium system is the Bohr radius. We define a corresponding scale parameter $`\mu _B`$ such that
$`\mu _B=C_F\alpha _S(\mu _B)m_{\mathrm{pole}}`$ (22)
holds. We examine the series expansion (1) with two different choices of the scale $`\mu `$ in Table 2.
The last terms (with stars) are evaluated using the value of $`c_3`$ in the large-$`\beta _0`$ approximation (21). One sees that for the bottomonium the series expansions do not converge at all. For the toponium the series expansions converge very slowly. The numbers in square brackets represent estimates of the last terms using the asymptotic form of the series expansion of the potential $`V_{\beta _0}(r)`$ :
$`C_F{\displaystyle \frac{\alpha _S(\mu )}{r}}\times R_n(\mu r){\displaystyle \frac{2e^{5/6}C_F\alpha _S(\mu )\mu }{\pi }}\times 2^nn!\text{for}n1.`$ (23)
The series approaches its asymptotic form faster when we choose $`\mu =\mu _B`$. These features are consistent with dominance of the leading renormalon contributions.
Next we rewrite the series expansion of $`M_{1S}`$ in terms of the $`\overline{\mathrm{MS}}`$-mass instead of the pole-mass. The leading renormalon contributions cancel in this case. Presently the relation between the $`\overline{\mathrm{MS}}`$-mass and the pole-mass is known up to three loops . The only scale in this relation is the quark mass. Thus, we have two choices of scales, $`\overline{m}`$ and $`\mu _B`$, in writing the series expansion of $`M_{1S}`$; we take $`\mu =\overline{m}`$ below. We eliminate the pole-mass completely and expand in $`\alpha _S(\overline{m})`$. We should properly take into account the fact that the renormalon contributions cancel between the terms whose orders in $`\alpha _S`$ differ by one . To this end we proceed as follows. We rewrite (1) as
$`M_{1S}`$ $`=`$ $`2m_{\mathrm{pole}}\times \left[\mathrm{\hspace{0.17em}1}+{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}Q_n\alpha _S(\overline{m})^n\right]`$ (24)
$`=`$ $`2\overline{m}\times \left[\mathrm{\hspace{0.17em}1}+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}d_n\alpha _S(\overline{m})^n\right]\times \left[\mathrm{\hspace{0.17em}1}+{\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}Q_n\alpha _S(\overline{m})^n\right],`$
where $`Q_n`$’s are polynomials of $`\mathrm{log}[\alpha _S(\overline{m})]`$ and $`d_n`$’s (the $`n`$-loop coefficients of the mass relation) are just constants independent of $`\alpha _S(\overline{m})`$. We identified $`Q_n\alpha _S^n`$ as order $`\alpha _S^{n1}`$ and then reduced the last line to a single series in $`\alpha _S`$. Numerically we find The values of $`d_1d_3`$ are taken from eq. (14) of .
$`M_{1S}`$ $`=`$ $`2\times (4.20+0.36+0.13+0.040+0.0051^{\mathrm{}})\mathrm{GeV}(\mathrm{Bottomonium}),`$ (25)
$`M_{1S}`$ $`=`$ $`2\times (165.00+7.21+1.24+0.22+0.052^{\mathrm{}})\mathrm{GeV}(\mathrm{Toponium}).`$ (26)
The last terms (with sharps) are evaluated using the values of $`c_3`$ and $`d_4`$ in the large-$`\beta _0`$ approximation. Convergences of the series improve markedly in comparison to those in Table 2. (Previous analyses similar to the one presented above can be found in and references therein.)
As we argued in Section 1, parametric accuracy of the last terms in (25,26) is $`𝒪(\alpha _S^4m)`$ and we need to know further only the exact value of $`d_4`$ to make a perturbative evaluation accurate up to this order (the exact form of $`c_3`$ is not necessary). In order to verify validity of this argument, we replace $`c_2`$ by its value in the large-$`\beta _0`$ approximation. Then the $`𝒪(\alpha _S^3m)`$ terms of (25,26) change to 0.043 and 0.22, respectively. Thus, we do not lose accuracy at this order by the replacement. On the other hand, if we replace $`c_2`$ and $`d_3`$ by their values in the large-$`\beta _0`$ approximation, the same terms change to 0.056 and 0.31, respectively. Thus, we lose the accuracy at $`𝒪(\alpha _S^3m)`$. These aspects are consistent with our general argument. Also, they suggest that the last terms of (25,26) would be reasonable estimates of the orders of magnitude of the exact $`𝒪(\alpha _S^4m)`$ terms.
Finally we examine if we can use the asymptotic form of $`V_{\beta _0}(r)`$, eq. (23), to estimate $`c_3(\text{large-}\beta _0)`$. From the asymptotic values of the last terms for $`\mu =\mu _B`$ in Table 2, we may extract approximate values of $`c_3(\text{large-}\beta _0)`$ as
$`c_3(\mathrm{asympt})=\{\begin{array}{cc}2.55\times 10^3& (n_l=4)\\ 1.98\times 10^3& (n_l=5)\end{array},`$ (29)
while the corresponding values of $`c_3(\text{large-}\beta _0)`$ are, respectively, $`2.46\times 10^3`$ and $`1.91\times 10^3`$. If we substitute $`c_3(\mathrm{asympt})`$, the last terms of (25,26) change to 0.0034 and 0.051, respectively. Therefore, we see that in the case of bottomonium the use of the asymptotic form does not reproduce our result in the large-$`\beta _0`$ approximation (25) with good (relative) accuracy. Presumably it is a sign of the next-to-leading renormalon contribution, which does not become smaller than $`\mathrm{\Lambda }_{\mathrm{QCD}}(\mathrm{\Lambda }_{\mathrm{QCD}}/\mu _B)^22`$ MeV, and which is not included in $`c_3(\mathrm{asympt})`$. Namely, we conjecture that already the last term in eq. (25) stands close to the limit where an improvement of convergence of the perturbation series is possible by cancellation of the leading renormalon contributions.
## 4 Conclusions and discussions
We have calculated the $`𝒪(\alpha _S^5m)`$ correction to the quarkonium $`1S`$-state energy spectrum analytically in the large-$`\beta _0`$ approximation. As a result, in order to predict the $`1S`$-state spectrum at $`𝒪(\alpha _S^4m)`$ perturbative accuracy, only the four-loop relation between the $`\overline{\mathrm{MS}}`$-mass and the pole-mass remains to be computed. Within the present approximation, the perturbation series of the bottomonium and toponium spectra show healthy convergent behaviors up to a genuine $`𝒪(\alpha _S^4m)`$ accuracy.
In the case of bottomonium, current theoretical uncertainty due to non-perturbative effects is estimated to be of the order of 0.1 GeV . It is much larger than the size of the last term of (25). We hope that in the future the theoretical uncertainty due to non-perturbative effects will be reduced by applications of e.g. lattice calculations or combinations of operator-product-expansion and sum rules.
Part of the genuine $`𝒪(\alpha _S^5m)`$ corrections have already been calculated . Taking $`\mu =\overline{m}`$, their individual sizes are evaluated to be $`\mathrm{\Delta }M_{1S}\pm 0.05`$ GeV both for the bottomonium and toponium; if we take $`\mu =\mu _B`$, they become even an order of magnitude larger. We do not know as yet whether it may indicate a breakdown of perturbative expansion of the spectrum at $`𝒪(\alpha _S^5m)`$ or a requisiteness for some new cancellation mechanism. It is also possible that the sum of all the genuine $`𝒪(\alpha _S^5m)`$ corrections turns out to be much smaller. In any case, it would be better to separate the discussion of this problem from the determination of the genuine $`𝒪(\alpha _S^4m)`$ corrections (as we advocated in this paper), rather than to regard them as inseparable constituents of $`c_3`$.
## Acknowledgements
Y.K. was supported by the Japan Society for the Promotion of Science. Y.S. was supported in part by the Japan-German Cooperative Science Promotion Program.
|
warning/0007/hep-th0007183.html
|
ar5iv
|
text
|
# Gauge invariant derivative expansion of the effective action at finite temperature and density and the scalar field in 2+1 dimensions
## I Introduction
In their pioneering work, Deser, Jackiw and Templeton noted that gauge theories in odd-dimensional spaces naturally admit a local term of topological nature, known as Chern-Simons term (see for a recent review). One of the interesting properties of the non-Abelian Chern-Simons term is that under gauge transformations it changes proportionally to the winding number of the transformation. Thus, when the action contains a Chern-Simons term, the partition functional of the system is only well defined if the coupling constant of the Chern-Simons term is properly quantized.
It was later realized that the Chern-Simons term is induced by quantum fluctuations when gauge fields are coupled to odd-dimensional fermions. Such term comes out with the correctly quantized coupling constant and so full gauge invariance is preserved (although possibly at the price of spoiling parity invariance). Because the Chern-Simons term is a polynomial in the gauge fields and their derivatives, it can be obtained through a combination of perturbative and derivative expansions.
In the so-called imaginary time formalism for field theory at finite temperature, the space-time has a non-trivial topology, since the time is effectively compactified to a circle. This allows the existence of topologically large gauge transformations even in the Abelian case. When the problem of the induced Chern-Simons term is studied using the method just mentioned of retaining a low number of fields and of derivatives, a puzzling situation appears, namely, the coefficient of the Chern-Simons term turns out to be a smooth function of the temperature, and hence it violates the quantization condition. The situation has recently been clarified by considering a simple $`0+1`$-dimensional model which can be computed in closed form. There it is seen that full gauge invariance holds for the exact result but it is broken by perturbation theory. This is not difficult to understand, since in simple cases gauge invariance under large gauge transformations is equivalent to periodicity of the effective action as a function of the gauge field, whereas perturbation theory corresponds to a Taylor expansion of that function. Clearly, the property of being periodic is not maintained in general by a truncated Taylor expansion. In it was noted that full gauge invariance is always a property of the exact result, since it follows straightforwardly from using a $`\zeta `$-function regularization. In the exact effective action of fermions in $`2+1`$ dimensions was obtained for the case of Abelian and stationary background gauge fields.
The problem of preserving full gauge invariance at finite temperature is not tied to odd-dimensional theories nor to fermions . It appears whenever perturbation theory is involved. This is unfortunate, since, as noted in , “at finite temperature, perturbation theory is one of the few tools we have.” In this work we show that it is possible to carry out detailed calculations of the effective action fully preserving gauge invariance, without restricting oneself to particular configurations such as Abelian or stationary ones, and without choosing a particular gauge. The study of simple cases shows that the problem with gauge invariance comes through the scalar potential $`A_0(x)`$. The finite temperature effective action is non-local in time but it is local in the space variables. This suggests to consider an expansion in the number of spatial covariant derivatives only. The time component is treated non-perturbatively in order to avoid destroying gauge invariance. (See for another discussion of derivative expansions at finite temperature.)
It should be emphasized that the expansion in the number of spatial covariant derivatives is not tied to a particular method of computation, since it can be obtained from the exact result (namely, by considering an appropriate spatial dilatation of the background fields) and is fully gauge invariant. What it is shown here is that it is also amenable to explicit computation order by order, through a combination of the method of symbols and $`\zeta `$-function, for instance. This combination works very well and has been applied at zero temperature for fermions with local and non-local actions. At finite temperature it has been applied to fermions in odd-dimensions as well as in even-dimensions, however, for technical reasons, this has been done choosing a particular gauge. In the present work we remove the necessity of any choice of gauge. It turns out that previous formulas in can be reinterpreted and rewritten in a manifest gauge invariant form.
Although a naive perturbative expansion in $`A_0`$ breaks gauge invariance, the very method of calculation suggests an expansion in powers of the temporal covariant derivative in the adjoint representation which preserves gauge invariance. This yields the remarkable result that, even at finite temperature, the theory is local if expressed in the appropriate variables.
The method is explicitly applied to the case of relativistic scalar particles in $`2+1`$ dimensions. The computation is carried out through second order in the number of spatial covariant derivatives. Contact is made with the relativistic Bose gas.
## II Gauge invariant derivative expansion at finite temperature
### A The mathematical problem
The aim of this section is to present a scheme to address the computation of the one-loop effective action at finite temperature preserving gauge invariance at every step. (We note that gauge invariance in this work refers always to vector gauge transformations.) To fix ideas consider the case of scalar particles in $`d+1`$ dimensions in presence of background gauge fields. This case will be worked out later for $`d=2`$. The Euclidean action of the system is
$$S=d^{d+1}x\left((D_\mu \varphi )^{}(D_\mu \varphi )+m^2\varphi ^{}\varphi \right)$$
(1)
where $`D_\mu =_\mu +A_\mu `$ is the covariant derivative. The finite temperature condition can be implemented by using the imaginary time formalism, that is, by compactification of the Euclidean time to a circle so that the fields $`\varphi `$ and $`A_\mu `$ are periodic functions of $`x_0`$ with period $`\beta =1/T`$ ($`T`$ being the temperature). After functional integration over $`\varphi (x)`$, the Euclidean effective action is formally given by
$$W_s[m,A]=\mathrm{Tr}_b\mathrm{log}(D_\mu ^2+m^2).$$
(2)
The subindex $`b`$ recalls that the functional trace is to be taken in the Hilbert space of bosonic wave functions, i.e. with periodic boundary conditions.
Presently the mathematical problem to be addressed is the computation of quantities of the form
$$\mathrm{\Gamma }[M,A]=\mathrm{Tr}(f(M,D))$$
(3)
where $`D_\mu `$ is the covariant derivative and $`M(x)`$ collectively denote one or more matrix valued functions of $`x_\mu `$ representing other external fields in addition to the gauge fields. The trace refers to the Hilbert space $``$ of wavefunctions with space-time and internal degrees of freedom, the space-time manifold has topology $`_{d+1}=\mathrm{S}^1\times _d`$ and the wavefunctions are periodic for bosons and antiperiodic for fermions.
Although slightly pedantic, it will occasionally be convenient to regard $`M`$ not as functions but rather as multiplicative operators in $``$, i.e., operators commuting with the operators $`x_\mu `$ and otherwise with arbitrary structure in internal space. Likewise $`D_\mu `$ are differential operators of the form $`_\mu +A_\mu `$ with $`A_\mu `$ multiplicative. The quantity $`f(M,D)`$ denotes an operator constructed out of $`M`$ and $`D_\mu `$ in the algebraic sense, that is, $`f(M,D)`$ is a linear combination (or series) of products of $`M`$ and $`D_\mu `$ multiplied in any order with constant c-number coefficients. In order for $`M`$ and $`D_\mu `$ to be well-defined operators in $``$, $`M(x)`$ and $`A_\mu (x)`$ are required to be periodic functions of $`x_0`$. In addition we will assume that the fields are sufficiently convergent at infinity and the function $`f`$ is well-behaved. This means, in particular, that $`f`$ is one-valued and sufficiently convergent at infinity as a function of $`D_\mu `$ to ensure the existence of the trace (by avoiding ultraviolet divergences).
A gauge transformed configuration $`(M^U,A^U)`$ is one of the form
$$M^U(x)=U^1(x)M(x)U(x),A_\mu ^U(x)=U^1(x)_\mu U(x)+U^1(x)A_\mu (x)U(x),$$
(4)
where the gauge transformation $`U(x)`$ is a periodic function of $`x_0`$ which takes values on matrices in internal space. This corresponds to a similarity transformation of $`D_\mu `$, namely, $`D_\mu ^U=_\mu +A_\mu ^U=U^1D_\mu U`$ where $`U`$ is to be regarded as a multiplicative operator in $``$. Because $`f(M,D)`$ is constructed with $`M`$, $`D`$ and c-numbers, it follows that $`f(M,D)`$ also transforms under a similarity transformation
$$f(M^U,D^U)=U^1f(M,D)U$$
(5)
and so
$$\mathrm{\Gamma }[M^U,A^U]=\mathrm{\Gamma }[M,A]$$
(6)
using the cyclic property of the trace, which holds due to our regularity assumptions for $`f(M,D)`$.<sup>*</sup><sup>*</sup>*In practice, $`\mathrm{\Gamma }[M,A]`$ is only computed for a subset of configurations $`(M,A)`$ and only the subgroup of gauge transformations which leave invariant such a subset are relevant. For fermions, the internal space includes Dirac space as well flavor degrees of freedom, and the $`\gamma _\mu `$ matrices are included in $`M`$ (they are not c-numbers). In this case only gauge transformations in flavor space are relevant since they are the ones that preserve the form of $`\gamma _\mu `$.
### B The method of symbols
Assuming that the operator $`\widehat{f}=f(M,D)`$ admits a complete set of eigenfunctions, $`\widehat{f}|n=\lambda _n|n`$, the functional trace is simply $`\mathrm{\Gamma }[M,A]=_n\lambda _n`$. In this form gauge invariance is obvious since $`\widehat{f}`$ and $`\widehat{f}^U`$ are related by a similarity transformations and hence they have the same spectrum.
The gauge invariance of $`\mathrm{\Gamma }[M,A]`$ is also manifest computing the trace in the basis $`|x`$ of eigenfunctions of $`x_\mu `$, normalized as $`x|x^{}=\delta (xx^{})`$ (a periodic delta function in the temporal direction)
$$\mathrm{\Gamma }[M,A]=d^{d+1}x\mathrm{tr}x|f(M,D)|x$$
(7)
(where $`\mathrm{tr}`$ refers to internal space) because $`U`$ is a multiplicative operator and so $`\mathrm{tr}x|f(M,D)|x`$ is gauge invariant without integration over $`x`$. However, computationally it is more convenient to use a basis in momentum space $`|p`$
$$x|p=e^{px},p|p^{}=\beta \delta _{p_0p_0^{}}(2\pi )^d\delta (𝒑𝒑^{}),$$
(8)
(to avoid unessential factors of $`i`$ we take the convention of using purely imaginary momenta $`p_\mu `$ but $`d^d𝒑`$ below denotes the usual integral in $`^d`$ and $`\delta (𝒑𝒑^{})`$ denotes the corresponding delta function). The frequency takes the Matsubara values $`p_0=2\pi in/\beta `$ for bosons and $`p_0=2\pi i(n+\frac{1}{2})/\beta `$ for fermions. Note that we have assumed that the space manifold $`_d`$ has a topology $`^d`$. In this basis
$$\mathrm{\Gamma }[M,A]=\frac{1}{\beta }\underset{p_0}{}\frac{d^d𝒑}{(2\pi )^d}\mathrm{tr}p|f(M,D)|p.$$
(9)
At this point the symbols method can be used (see e.g. ): let $`|0`$ denote the state with $`p=0`$, then using the identities $`|p=e^{xp}|0`$ (where $`e^{xp}`$ acts as a multiplicative operator and the quantities $`p_\mu `$ are constant c-numbers) as well as $`e^{xp}D_\mu e^{xp}=D_\mu +p_\mu `$, $`e^{xp}Me^{xp}=M`$, one obtains
$$p|f(M,D)|p=0|e^{xp}f(M,D)e^{xp}|0=0|f(M,D+p)|0,$$
(10)
and so the functional trace can be cast in the form
$$\mathrm{\Gamma }[M,A]=\frac{1}{\beta }\underset{p_0}{}\frac{d^d𝒑}{(2\pi )^d}\mathrm{tr}0|f(M,D+p)|0.$$
(11)
In this expression it is clear the requirement of regularity on $`f`$: the functional trace comes after integration over momenta and sum over frequencies and this requires $`f`$ to be sufficiently convergent for large $`p_\mu `$. Let us remark that $`|0`$ is periodic rather than antiperiodic in the temporal direction. The information on whether we are dealing with bosons or fermions is now contained solely in the values taken by $`p_0`$. The state $`|0`$ satisfies
$$x|0=1,_\mu |0=0|_\mu =0,0|0=d^{d+1}x.$$
(12)
In addition, when $`\widehat{h}`$ is a multiplicative operator, $`\widehat{h}|x=h(x)|x`$,
$$0|\widehat{h}|0=d^{d+1}xh(x).$$
(13)
It follows that $`_\mu `$ appearing inside $`f(M,D+p)`$ in eq. (11) acts derivating everything to its right (or its left, by parts) and then vanishes after it reaches $`|0`$ (or $`0|`$). This is a well-defined working rule and from it stems the usefulness of the symbols method.
Unfortunately, gauge invariance is no longer manifest when using the momentum basis. In fact, $`\mathrm{tr}0|f(M,D+p)|0`$ is not gauge invariant because $`|0`$ (or more generally $`|p`$) is not covariant under local transformations. For instance, according to the rules given in eq. (12),
$$\mathrm{tr}0|D_\mu ^2|0=\mathrm{tr}0|A_\mu ^2|0=d^{d+1}x\mathrm{tr}[A_\mu ^2(x)]$$
(14)
breaks gauge invariance. However,
$$\mathrm{tr}0|[D_\mu ,D_\nu ]^2|0=d^{d+1}x\mathrm{tr}[F_{\mu \nu }^2(x)]$$
(15)
does not. Note that $`[D_\mu ,D_\nu ]`$ is a multiplicative operator whereas $`D_\mu ^2`$ is not. As a rule, when an operator $`g(M,D)`$ (a gauge covariant operator) is multiplicative, eq. (13) applies and $`\mathrm{tr}0|g(M,D)|0`$ is gauge invariant. In eq. (11) gauge invariance is only recovered after integration over momenta and sum over frequencies.An elegant method has been presented in which yields gauge invariant expressions prior to momentum integration, at the price of introducing derivatives with respect to $`p_\mu `$. The method has not yet been extended to include discrete momenta, as required at finite temperature but in can be applied to the integration over $`𝒑`$. This will be further discussed subsequently.
### C The derivative expansion at finite temperature
By computing the functional trace we essentially mean to end up with purely multiplicative operators, since this implies that the functional is expressed as the integral of a function over space-time. At zero temperature this is usually equivalent to saying that all derivative operators $`D_\mu `$ appear inside commutators. In addition it means to carry out as many implied sums and integrations (over frequencies and momenta or other parameters) as possible.
In general it is not possible to compute $`\mathrm{\Gamma }[M,A]`$ in closed form and one must resort to approximations. The standard approach is to make power expansions in one or more operators appearing in $`f(M,D)`$ while the remaining operators are treated non perturbatively. As will be clear below, a naive expansion in powers of $`D_0`$ would break gauge invariance, therefore, because it is in general difficult to work with two or more non perturbative operators unless they are commuting, and our present emphasis is in the preservation of manifest gauge invariance rather than in a particular computation, we will keep $`D_0`$ as the only operator to be treated non perturbatively, and expand in all other operators $`M`$ and $`𝑫`$.
Before proceeding, let us be more precise about the meaning of expanding in powers of $`M`$ and $`𝑫`$. A convenient way to define the expansion is by introducing constant c-number bookkeeping parameters, $`f(M,D_0,𝑫)f(\lambda _1M,D_0,\lambda _2𝑫)`$, so that counting powers of those operators is equivalent to counting powers of $`\lambda _{1,2}`$. This procedure preserves gauge invariance since it amounts to a modification of the function $`f`$. After applying the symbols method (through $`D_\mu D_\mu +p_\mu `$, cf. eq. (11)) the factor $`\lambda _2`$ will affect $`𝑫`$ and $`𝒑`$, however, this can be brought to the form $`f(\lambda _1M,D_0+p_0,\lambda _2𝑫+𝒑)`$ by a redefinition of $`𝒑`$. Therefore the expansion can be formulated as an expansion in powers of $`M`$ and $`𝑫`$ in $`f(M,D+p)`$.Of course, the actual expansion in powers of $`𝑫`$ must be done after arriving at eq. (11) (or other similar formulas in other approaches, such as Schwinger proper time method) i.e., after $`D_\mu `$ has become $`D_\mu +p_\mu `$, since otherwise powers of $`p_\mu `$ would be generated as well and that would destroy the ultraviolet convergence of the formula. Equivalently, the expansion in $`𝑫`$ can be obtained directly from the functional $`\mathrm{\Gamma }[M,A]`$ by means of a covariant spatial dilatation, namely, $`M(x_0,𝒙)M(x_0,\lambda _2𝒙)`$, $`A_0(x_0,𝒙)A_0(x_0,\lambda _2𝒙)`$ and $`𝑨(x_0,𝒙)\lambda _2𝑨(x_0,\lambda _2𝒙)`$. This guarantees that the expansion is well-defined, i.e., it depends on the functional itself and not on how it is written or computed.
The situation is completely different for an expansion in powers of $`D_0`$. A bookkeeping parameter $`D_0\lambda _3D_0`$ can be introduced in $`f(M,D)`$ and this defines a new $`\lambda _3`$-dependent gauge invariant functional. Nevertheless this functional is not useful since it presents an essential singularity at $`\lambda _3=0`$, as can be seen in the simple case of fermions in 0+1 dimensions. (The dependence on $`\lambda _{1,2}`$ is analytic or at least asymptotic under suitable regularity conditions on the fields and on the function $`f`$.) After applying the symbols method, $`f(M,\lambda _3D_0+\lambda _3p_0,𝑫)`$ is obtained. However, because $`p_0`$ is a discrete variable this is not equivalent to $`f(M,\lambda _3D_0+p_0,𝑫)`$. Therefore expanding in the explicit $`D_0`$ in $`f(M,D+p)`$ does not correspond to a modification of $`f`$ and in fact violates gauge invariance. Also, it is not possible to introduce $`\lambda _3`$ by means of a rescaling of type $`x_0\lambda _3x_0`$ of the field configuration $`(M,A)`$ since this transformation violates the periodicity condition on the wavefunctions of $``$.
After expansion there will be all kind of terms which will be products of single factors of $`M`$ and $`𝑫`$ as well as operators depending non perturbatively on $`D_0`$ (by this we merely mean that all orders of $`D_0`$ are retained). It is always possible to bring all $`𝑫`$ operators to the right producing commutators, so that we end up with two kind of terms: i) terms in which all operators $`𝑫`$ appear only in commutators (more precisely, in the form $`[𝑫,]`$) and ii) terms with unsaturated factors $`𝑫`$ at the right (i.e., $`𝑫`$ not inside a commutator). The terms of the first type are multiplicative operators regarding $`𝒙`$-space, although they are still differential (or pseudo-differential) operators with respect to $`x_0`$-space. The terms of the second type are non-multiplicative in $`𝒙`$-space. As we have argued above, these latter terms break gauge invariance and in fact they will cancel after integration over $`𝒑`$. This can be seen as follows: let us replace $`𝑫`$ by $`𝑫+𝒂`$, where $`𝒂`$ is a constant c-number. This replacement has no effect on the terms where $`𝑫`$ is in commutators, but counts the contribution from the terms with unsaturated $`𝑫`$. However, it is clear that there is no such a contribution after integration over momenta since $`𝒂`$ can be compensated by a similar shift in the integration variable $`𝒑`$. Thus at the end, all operators $`𝑫`$ appear in commutators only. (The same result is obtained directly using the method of Pletnev and Banin.) A similar argument would break down for $`D_0`$: a shift to $`D_0+a_0`$ cannot, in general, be compensated by a shift in $`p_0`$ since at finite temperature the frequency is a discrete variable.
From the previous discussion it follows that we only have to retain those terms where all operators $`𝑫`$ are in commutators. Because all operators are now multiplicative in $`𝒙`$-space, $`𝒙`$ becomes just a parameter in what follows. Let us consider a typical term:
$$\text{TT}=\frac{1}{\beta }\underset{p_0}{}\frac{d^d𝒑}{(2\pi )^d}\mathrm{tr}0|\alpha _1(D_0+p_0,𝒑)X\alpha _2(D_0+p_0,𝒑)Y\alpha _3(D_0+p_0,𝒑)|0.$$
(16)
$`X`$ and $`Y`$ are multiplicative and gauge covariant operators constructed with $`D_\mu `$ and $`M`$, and the $`\alpha _i(x,y)`$ are some functions. At this point the integration over $`p_\mu `$ is non trivial (even for the simplest forms of the functions $`\alpha _i(x,y)`$) because $`p_\mu `$ appears in different and non-commuting operators. A possible approach is to express the operators in terms of their matrix elements using as basis a complete set of eigenstates of $`D_0`$ (in the Hilbert space of time and internal degrees of freedom). These matrix elements are then ordinary functions of $`p_\mu `$. Instead of that, we will use the equivalent prescription of labelling the operators $`D_0`$ according to their position with respect to $`X`$ and $`Y`$: the symbols $`D_{01}`$, $`D_{02}`$ and $`D_{03}`$ will be used to denote the operator $`D_0`$ is positions 1 (before $`X`$), 2 (between $`X`$ and $`Y`$) and 3 (after $`Y`$) respectively. In this notation
$$\text{TT}=\frac{1}{\beta }\underset{p_0}{}\frac{d^d𝒑}{(2\pi )^d}\mathrm{tr}0|\alpha _1(D_{01}+p_0,𝒑)\alpha _2(D_{02}+p_0,𝒑)\alpha _3(D_{03}+p_0,𝒑)XY|0.$$
(17)
An immediate consequence is that the labeled operators are effectively commuting and the momentum integration and frequency summation can be carried out as for ordinary functions. The result can be written as
$$\text{TT}=\mathrm{tr}0|g(D_{01},D_{02},D_{03})XY|0$$
(18)
where the function $`g`$ is defined by
$$g(x,y,z)=\frac{1}{\beta }\underset{p_0}{}\frac{d^d𝒑}{(2\pi )^d}\alpha _1(x+p_0,𝒑)\alpha _2(y+p_0,𝒑)\alpha _3(z+p_0,𝒑).$$
(19)
(Note that there will be two versions of $`g`$, the bosonic one and the fermionic one, which are related by a shift of $`i\pi /\beta `$ in their arguments.) By construction the function $`g`$ is periodic:
$$g(x,y,z)=g(x+\frac{2\pi i}{\beta },y+\frac{2\pi i}{\beta },z+\frac{2\pi i}{\beta }).$$
(20)
This is an immediate consequence of the sum over Matsubara frequencies and ultraviolet convergence of the expressions.
This periodicity property is essential to codify the gauge invariance of the original expression. To see this, let us introduce the operation $`𝒟_\mu `$ which is defined as $`𝒟_\mu X=[D_\mu ,X]`$ for any operator $`X`$. Consistently with our previous notation, we will denote by $`𝒟_{01}`$ the action of $`𝒟_0`$ in position 1 (i.e., on $`X`$) and by $`𝒟_{02}`$ the action of $`𝒟_0`$ in position 2 (i.e., on $`Y`$). Then clearly
$$𝒟_{01}=D_{01}D_{02},𝒟_{02}=D_{02}D_{03}.$$
(21)
The interesting point is that these formulas hold for arbitrary functions of $`𝒟_{01}`$ and $`𝒟_{02}`$ as well. This follows from the well-known identity $`e^ABe^A=e^{[A,]}B`$: for any c-number $`\lambda `$
$$e^{\lambda (D_{01}D_{02})}XY=e^{\lambda D_0}Xe^{\lambda D_0}Y=\left(e^{\lambda 𝒟_0}X\right)Y=e^{\lambda 𝒟_{01}}XY,$$
(22)
and this identity immediately extends to arbitrary functions of $`D_{01}D_{02}`$, and analogously for $`𝒟_{02}`$. This allows to make everywhere the replacements
$$D_{02}=D_{01}𝒟_{01},D_{03}=D_{01}𝒟_{01}𝒟_{02}$$
(23)
and use $`D_{01}`$, $`𝒟_{01}`$ and $`𝒟_{02}`$ as the independent variables to work with. The advantage of doing this is that the action of $`𝒟_{01}`$ and $`𝒟_{02}`$ on $`X`$ and $`Y`$ produces multiplicative and gauge covariant operators. On the other hand, the presence of the operator $`D_{01}`$, which is outside commutators, combined with the gauge non-covariant operation $`0||0`$, still can introduce gauge non-invariant contributions. This is avoided thanks to the periodicity property of $`g(x,y,z)`$ as we will show now. Indeed, the periodicity property allows to write the term as
$$\text{TT}=\mathrm{tr}0|\phi (e^{\beta D_{01}},𝒟_{01},𝒟_{02})XY|0,$$
(24)
where the function $`\phi `$ is defined by
$$\phi (e^{\beta x},y,z)=g(x,xy,xyz).$$
(25)
The periodicity condition of $`g`$ ensures that the function $`\phi (\omega ,y,z)`$ is one-valued (it depends on $`\omega `$ and not just on $`\mathrm{log}(\omega )`$). In order to bring the expression into a manifestly gauge invariant form, we will use the following property:
$$e^{tD_0}=e^{t_0}Te^{_{x_0}^{x_0+t}A_0(x_0^{},𝒙)𝑑x_0^{}},t0.$$
(26)
Here $`t`$ is just a parameter and $`T`$ denotes time ordered product. (The quantities $`D_0`$, $`_0`$ and $`x_\mu `$ represent operators and the product refers to a product of operators so that $`_0`$ is not directly derivating $`x_0`$.) This equation can be easily proved by the standard procedure of showing that the two expressions satisfy the same first order differential equation in $`t`$ and coincide at $`t=0`$. The left-hand side is a manifestly gauge covariant operator. It is interesting to see how gauge covariance is realized in the right-hand side: the time-ordered product from $`x_0`$ to $`x_0+t`$ (with $`𝒙`$ fixed) transforms with $`U(x_0,𝒙)`$ at the right and $`U^1(x_0+t,𝒙)`$ at the left, and this latter factor is transformed into $`U^1(x_0,𝒙)`$ after commutation with $`e^{t_0}`$, thus the product of the two factors transforms covariantly at $`(x_0,𝒙)`$, as $`e^{tD_0}`$.
In particular, by taking $`t=\beta `$ in the previous formula, one obtains the identity
$$e^{\beta D_0}=e^{\beta _0}\mathrm{\Omega },$$
(27)
where
$$\mathrm{\Omega }(x)=T\mathrm{exp}\left(_{x_0}^{x_0+\beta }A_0(x_0^{},𝒙)𝑑x_0^{}\right).$$
(28)
(Again $`e^{\beta _0}\mathrm{\Omega }`$ is to be understood as the product of two operators.) Beyond the interval $`[0,\beta ]`$ $`A_0(x)`$ is defined as a periodic function of the time, so $`\mathrm{\Omega }(x)`$ is also periodic. Although $`\mathrm{\Omega }(x)`$ is non-local in terms of $`A_0`$, it behaves as a local field which takes values on the gauge group. In particular it transforms covariantly at $`x`$:
$$\mathrm{\Omega }^U(x)=U^1(x)\mathrm{\Omega }(x)U(x).$$
(29)
The matrices $`\mathrm{\Omega }(x)`$ at different values of $`x_0`$, but equal $`𝒙`$, are related by similarity transformations and their trace, the Polyakov loop, is independent of $`x_0`$. Another important property is
$$𝒟_0\mathrm{\Omega }=[D_0,\mathrm{\Omega }]=0.$$
(30)
On the other hand, the effect of the operator $`\mathrm{exp}(\beta _0)`$ is to produce the shift $`x_0x_0\beta `$, therefore it is equivalent to the identity operator on the space of periodic functions in which we are working (as noted the periodic wavefunction $`|0`$ appears regardless of whether we are considering bosons of fermions). So in this space
$$e^{\beta D_0}=\mathrm{\Omega }.$$
(31)
This produces the manifestly gauge invariant expression
$$\text{TT}=\mathrm{tr}0|\phi (\mathrm{\Omega }_1,𝒟_{01},𝒟_{02})XY|0.$$
(32)
(The label 1 in $`\mathrm{\Omega }`$ indicates to put this operator in position 1. The relative order between $`𝒟_{01}`$ and $`\mathrm{\Omega }_1`$ is immaterial due to eq. (30).)
It can be noted that all previous manipulations, starting from eq. (16), hold also without taking $`\mathrm{tr}0||0`$. Inside $`\mathrm{tr}0||0`$ integration by parts implies that $`𝒟_{01}`$ is equivalent to $`𝒟_{02}`$ (or equivalently, that $`D_{03}=D_{01}`$), and so we have the final formula
$$\text{TT}=\mathrm{tr}0|\phi (\mathrm{\Omega }_1,𝒟_{02})XY|0,$$
(33)
where
$$\phi (\omega ,y)=\phi (\omega ,y,y).$$
(34)
The whole point of these manipulations was to end up with a manifestly gauge covariant and multiplicative operator so that $`\mathrm{tr}[\phi (\mathrm{\Omega }_1,𝒟_{02})XY]`$ is just a gauge invariant function of $`x`$, constructed with the fields $`A_\mu (x)`$ and $`M(x)`$:
$$\text{TT}=d^{d+1}x\mathrm{tr}[\phi (\mathrm{\Omega }_1,𝒟_{02})XY],$$
(35)
The fact that $`g(x,y,z)`$ is periodic is essential to produce the gauge covariant operator $`\mathrm{\Omega }(x)`$. This periodicity property would be lost in an expansion in powers of $`D_0`$. We should remark that no restriction has been put on the field configuration, which is completely general (may be non Abelian and non stationary) and also no choice of gauge has been needed.
Perhaps it should be emphasized how exactly the lack of periodicity would break gauge invariance. To see this it is sufficient to consider $`0+1`$-dimensional fermions in presence of an Abelian configuration . The corresponding effective action is of the form $`\beta g(a)`$ with $`a=_0^\beta 𝑑x_0A_0`$. The quantity $`a`$ is invariant under topologically small gauge transformations, $`A_0^U=A_0+_0\mathrm{\Lambda }`$ ($`\mathrm{\Lambda }(x_0)`$ being a periodic function) but under a large gauge transformation, e.g. $`A_0^U=A_0+2\pi in/\beta `$ (which corresponds to $`U(x_0)=\mathrm{exp}(2\pi inx_0/\beta )`$), $`a`$ changes by an integer multiple of $`2\pi i`$, so $`g(a)`$ will not be invariant in general. When $`g`$ is periodic the effective action becomes $`\beta \phi (\mathrm{\Omega })`$ with $`\mathrm{\Omega }=\mathrm{exp}(a)`$ and it is invariant under all gauge transformations. See Section IV A for further remarks.
### D Relation with the calculation fixing the gauge
In the kind of calculation just described was carried out for fermions but fixing the gauge through the gauge condition $`_0A_0=0`$. (The idea was that the two operators treated not perturbatively, $`_0`$ and $`A_0`$, are then commuting.) No loss of generality is actually implied by this approach since such a gauge always exists. However, because it is not unique, it is necessary to find all remaining gauge transformations allowed within the $`A_0`$-stationary gauge, and then check that all of them produce the same result. This was shown to be equivalent to the periodicity condition that follows from summing over Matsubara frequencies, eq. (20). All this is unnecessary in the present approach since the gauge has not been fixed. Using gauge invariance, the results obtained within the $`A_0`$-stationary gauge can directly be taken over as follows. When $`_0A_0=0`$, the field $`\mathrm{\Omega }`$ becomes $`e^{\beta A_0}`$, so it is only necessary to replace $`e^{\beta A_0}`$ of the calculation in $`A_0`$-stationary gauge by $`\mathrm{\Omega }(x)`$ to obtain the result expressed in an arbitrary gauge. Further comments are made in IV A.
### E Expansion in space-time derivatives at finite temperature
As noted, expanding in powers of $`D_0`$ breaks periodicity and hence gauge invariance, however, in principle nothing prevents from expanding in powers of $`𝒟_0`$ in eq. (35), namely,
$$\text{TT}=\underset{n=0}{\overset{\mathrm{}}{}}d^{d+1}x\mathrm{tr}[\phi _n(\mathrm{\Omega })X𝒟_0^nY],$$
(36)
where
$$\phi (\omega ,y)=\underset{n=0}{\overset{\mathrm{}}{}}\phi _n(\omega )y^n.$$
(37)
The interest of doing that is, of course, that, at least at lower orders, the result is simpler than the full result. This can be regarded as the finite temperature generalization of the usual derivative expansion at zero temperature. Recall that $`𝑫`$ already was restricted to appear in commutators only, so this is really an expansion in powers of $`𝒟_\mu `$. As usual, higher orders are increasingly ultraviolet convergent. It can be noted that in the Abelian and stationary case $`𝒟_0`$ vanish identically (on multiplicative operators, such as $`X`$ and $`Y`$ in eq. (36)) therefore the zeroth order in the above expansion becomes exact.
Nevertheless, it should be noted that such an expansion is not as well-defined as for instance the expansion in powers of $`𝑫`$. The latter is defined from the functional itself, since it corresponds to spatial dilatations of the fields. No such transformation is known for the expansion in powers of $`𝒟_0`$. So it principle it should be expected that different ways of expressing the functional in terms of $`\mathrm{\Omega }`$ and $`𝒟_0`$ would yield different expansions, only the sum of all orders being unambiguously defined. This can be seen more clearly as follows. Recalling that inside $`0||0`$ the operators $`D_{03}`$ and $`D_{01}`$ are equivalent, the typical term considered above, eq. (18), takes the form
$$\text{TT}=\mathrm{tr}0|g(D_{01},D_{02})XY|0$$
(38)
(with $`g(x,y)=g(x,y,x)`$). Using the final form eq. (33), it is easily established that it can also be written as
$$\text{TT}=\mathrm{tr}0|g(D_{02},D_{01})YX|0,$$
(39)
because in eq. (33) all operators are multiplicative and therefore integration by parts and the cyclic symmetry can be used. Now, in the frequent case of contributions where $`X=Y`$, this implies that only the symmetric part of the function $`g(x,y)`$ is actually contributing. However, it is easy to write purely antisymmetric and periodic functions $`g(x,y)`$ such that when used in eq. (36), each order is non vanishing, although of course their full contribution vanish when summed to all orders. This particular kind of ambiguity can be fixed by imposing a symmetry restriction on $`g(x,y)`$ before carrying out the expansion in $`𝒟_0`$. This ambiguity is further discussed in Section IV B
### F Illustration of the method
To illustrate the previous manipulations in a practical case, we will consider the quantity
$$C[m,A]=\frac{1}{4}\mathrm{Tr}_b\left[\left(\frac{1}{D_\mu ^2+m^2}\frac{1}{2}\sigma _{\mu \nu }F_{\mu \nu }\right)^2\right]+O(F_{\mu \nu }^3),$$
(40)
which will appear later in the study of the scalar field in 2+1 dimensions. In this expression $`F_{\mu \nu }=[D_\mu ,D_\nu ]`$ and $`\sigma _{\mu \nu }=\frac{1}{2}[\gamma _\mu ,\gamma _\nu ]`$ (where $`\gamma _\mu `$ are Hermitian Dirac matrices in 2+1 dimensions). $`m`$ is a c-number.
$`C[m,A]`$ is ultraviolet finite and so it is a well-defined and unambiguous quantity. We will compute it through second order in an expansion in the number of spatial covariant derivatives. Clearly the expansion starts at second order and it is sufficient to retain the explicit term in eq. (40) since terms of $`O(F_{\mu \nu }^3)`$ must contain four or more spatial indices.
First, the symbols method, eq. (11), is applied. Afterwards, taking the trace in Dirac space and keeping just terms with two spatial covariant derivatives, produces
$$C_2[m,A]=\frac{1}{2}\frac{d^2𝒑}{(2\pi )^2}\frac{1}{\beta }\underset{p_0}{}0|\mathrm{tr}\left[\left(\frac{1}{(D_0+p_0)^2+𝒑^2+m^2}𝑬\right)^2\right]|0,$$
(41)
where $`𝒑^2=p_i^2`$, $`p_0=2\pi in/\beta `$, and $`𝑬=[D_0,𝑫]`$.
To proceed to the integration over momenta and sum over frequencies we use the trick of adding a label 1 or 2 to the operators $`D_0`$ to indicate their actual position in the expression, namely
$`C_2[m,A]`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^2𝒑}{(2\pi )^2}\frac{1}{\beta }\underset{p_0}{}0|\mathrm{tr}\left[\frac{1}{(D_{01}+p_0)^2+𝒑^2+m^2}\frac{1}{(D_{02}+p_0)^2+𝒑^2+m^2}𝑬^2\right]|0}`$ (42)
$`:=`$ $`0|\mathrm{tr}\left[g(D_{01},D_{02})𝑬^2\right]|0`$ (43)
The momentum integration yields
$$g(x_1,x_2)=\frac{1}{8\pi }\frac{1}{\beta }\underset{p_0}{}\frac{1}{(x_1+p_0)^2(x_2+p_0)^2}\mathrm{log}\left(\frac{m^2(x_1+p_0)^2}{m^2(x_2+p_0)^2}\right).$$
(44)
In order to sum over frequencies, it is convenient to reduce the expression to a rational form. This is achieved by derivating with respect $`m`$ and then integrating back (using that $`g(x_1,x_2)`$ vanishes as $`m\mathrm{}`$). Then the identity
$$\underset{n}{}\left(\frac{1}{x_1+i\pi n}\frac{1}{x_2+i\pi n}\right)=\mathrm{coth}(x_1)\mathrm{coth}(x_2),$$
(45)
can be applied. This produces
$$g(x_1,x_2)=\frac{1}{16\pi }\frac{1}{x_1x_2}_m^{\mathrm{}}𝑑t\left(\frac{\mathrm{coth}(\frac{\beta }{2}(t+x_1))}{2t+x_1x_2}\frac{\mathrm{coth}(\frac{\beta }{2}(t+x_2))}{2tx_1+x_2}\right)+\text{p.p.c.}$$
(46)
Where p.p.c. (which stands for pseudo-parity conjugate) refers to the same expression with the replacements $`x_1x_1`$ and $`x_2x_2`$.
Correspondingly,
$$C_2[m,A]=d^3x\mathrm{tr}\left[\phi (\mathrm{\Omega }_1,𝒟_{02})𝑬^2\right]$$
(47)
with
$$\phi (\omega ,y)=\frac{1}{16\pi }\frac{1}{y}_m^{\mathrm{}}𝑑t\left(\frac{1}{2ty}\frac{e^{\beta t}+\omega }{e^{\beta t}\omega }\frac{1}{2t+y}\frac{e^{\beta (t+y)}+\omega }{e^{\beta (t+y)}\omega }\right)+\text{p.p.c.}$$
(48)
p.p.c. corresponds to $`yy`$ and $`\omega \omega ^1`$. This is the final expression which contains all contributions to $`C[m,A]`$ with two spatial Lorentz indices and any number of zeroth indices. As expected at finite temperature, it is non-local in time but local in $`𝒙`$. Note that $`C_2[m,A]`$ is an even function of $`m`$.<sup>§</sup><sup>§</sup>§This can be shown by noting that $`\phi (\omega ,y;m)\phi (\omega ,y;m)`$ is given by the same formula (48) with replacement $`m\mathrm{}`$, and then showing that the integrand is convergent and odd as a function of $`t`$.
We can now consider a further expansion in powers of $`𝒟_0`$ since it respects gauge invariance. An explicit computation shows that $`\phi (\omega ,y)`$ is an analytic function of $`y`$. At leading (zeroth) order in $`𝒟_0`$ the result from eq. (48) is
$$C_2[m,A]=\frac{1}{16\pi }\frac{1}{2m}d^3x\mathrm{tr}\left[\left(\frac{e^{\beta m}+\mathrm{\Omega }}{e^{\beta m}\mathrm{\Omega }}+\frac{e^{\beta m}+\mathrm{\Omega }^1}{e^{\beta m}\mathrm{\Omega }^1}\right)𝑬^2\right]+O(𝒟_0).$$
(49)
Note that this result, unlike the full result in eqs. (47,48), does not contain an integral over the mass ($`_m^{\mathrm{}}𝑑t`$). This property holds to all orders in $`𝒟_0`$.
### G The effective action
In this subsection we summarize some properties of the effective action which will be needed later.
The effective action. The Euclidean effective action is defined as minus the logarithm of the partition functional. For non interacting fields it takes the form
$$W[M,A]=c\mathrm{Tr}\mathrm{log}(K(M,D))=c\mathrm{log}\mathrm{Det}(K(M,D)),$$
(50)
where $`K(M,D)`$ is a differential operator (e.g. the Klein-Gordon operator as in eq. (2) or the Dirac operator) and $`c`$ some constant.
Ultraviolet ambiguities. The previous expression needs to be regularized, and a number of methods can be used to obtain a renormalized version of it. The key observation is that all renormalized versions of the effective action must yield the same ultraviolet finite contributions and so two such versions can differ at most by a term which is a local polynomial in the external fields and their derivatives, with canonical dimension no more than $`d+1`$ (in $`d+1`$ dimensions). Note that for this discussion a mass $`m`$ plays the role of an external scalar field which happens to take a constant c-number configuration and so, in particular, the ambiguity in the renormalized action will depend polynomically on $`m`$. This implies that any sensible (that is, correctly describing the ultraviolet finite contributions) regularization plus renormalization prescription can be used to make the effective action finite; the actual effective action describing the physical system at hand will correspond to adding the appropriate local polynomial action to the previous result. Another consequence is that formal identities can be applied so long as a violation of them is allowed in the form of a local polynomial of dimension $`d+1`$ or less.
The $`\zeta `$-function method. The effective action can be defined through the $`\zeta `$-function prescription, namely
$$\mathrm{Tr}\mathrm{log}(K)=\frac{d}{ds}\mathrm{Tr}(K^s)|_{s=0}.$$
(51)
(An analytical extension in $`s`$ is understood from sufficiently negative values of $`s`$.) When $`K`$ admits a complete set of eigenvectors, $`\mathrm{Tr}(K^s)=_n\lambda _n^s`$, $`\lambda _n`$ being the eigenvalues of $`K`$. If the calculation has to be made using some expansion it is convenient to use the following formula
$$\mathrm{Tr}(K^s)=\mathrm{Tr}_\mathrm{\Gamma }\frac{dz}{2\pi i}z^s\frac{1}{zK},$$
(52)
where the path $`\mathrm{\Gamma }`$ encloses anti-clockwise the eigenvalues of $`K`$ but not $`z=0`$. This method is practical in actual calculations combined with the symbols method: after applying eq. (11) to expand the functional trace, it is straightforward to make an explicit expansion in $`M`$ and $`𝑫`$, for instance. This method has been used for fermions in and for a non-local Dirac operator in at zero temperature, and at finite temperature for odd dimensional fermions in and even dimensional fermions in .
Anomalies. If the effective action breaks a symmetry of the action there is an anomaly. In general the anomaly can be defined as the difference between the effective action of the original and the transformed configurations (of the external fields) and by construction is a local polynomial. It may happen that the symmetry can be restored by adding an appropriate local polynomial to the effective action. In this case the breaking is an unessential anomaly. When not all symmetries can be restored simultaneously the theory presents an essential anomaly. All symmetries which are implemented through a similarity transformation of $`K(M,D)`$ leave the spectrum invariant, thus, because the $`\zeta `$-function prescription defines the determinant of a differential operator by regularizing the product of its eigenvalues, it follows that for these symmetries the $`\zeta `$-function version of $`\mathrm{Det}(K)`$ is always free of anomalies. This applies in particular to vector gauge invariance. (Axial gauge transformations and scale transformations, for instance, are not implemented by similarity transformations and they can be anomalous.) Therefore the partition functional is always invariant under gauge transformations. On the other hand the effective action can change by integer multiples of $`2\pi i`$, since the logarithm is a many-valued function. (We are assuming that no zero modes are involved. They would induce changes multiples of $`i\pi `$ in the effective action.) By continuity, this can only happen for topologically large gauge transformations. For a scalar field the multivaluation cannot occur since the corresponding $`\zeta `$-function renormalized effective action is purely real. For fermions the multivaluation may take place depending on the topological numbers of the gauge transformation and the gauge field configuration and this indicates the presence of topological pieces in the effective action . Such multivaluation is indeed found in the explicit calculation for $`2+1`$-dimensional fermions of .
Locality and finite temperature. In previous subsections we have considered operators of the form $`f(M,D)`$ with $`f`$ one-valued and ultraviolet convergent. Actually, one wants to compute the effective action which contains multivaluation and ultraviolet divergences, eq. (50). In many expansions (perturbative, derivative, $`1/m`$, etc) higher orders are ultraviolet finite and thus they are also free of multivaluation. For those terms all our previous considerations hold directly. In particular, we find a remarkable result, namely, that the effective action at finite temperature can be written as (cf. eq. (36))
$$W[M,A]=\underset{n}{}d^{d+1}x\mathrm{tr}[\phi _n(\mathrm{\Omega })𝒪_n],$$
(53)
where $`\phi _n`$ are some functions and $`𝒪_n`$ are gauge covariant and local operators constructed out of $`𝒟_\mu `$ and $`M`$. In this sense the theory at finite temperature is local in the usual sense (i.e., the effective action admits an expansion in $`𝒟_\mu `$) provided that the field $`\mathrm{\Omega }(x)`$ is regarded as local.
For the ultraviolet divergent terms, some oddities appear which are necessary in order to accommodate the existence of anomalies, topological terms and multivaluation, all these issues being related. For instance, when the expansion in spatial covariant derivatives is computed for fermions in $`2+1`$ dimensions the functions $`\phi _n`$ of lower orders are many-valued (a property belonging to the exact result in $`0+1`$ dimensions and in $`2+1`$ dimension for Abelian and stationary configurations ). In addition, negative powers of $`𝒟_0`$ may appear when one goes beyond the Abelian and stationary case. This was handled in by introducing the fields $`𝓐(x)`$, defined as any solution of the equation $`𝒟_0𝑬=𝒟_0^2𝓐`$. There it was shown that the ambiguity in the definition of $`𝓐`$ cancels and all solutions yield the same effective action. In terms of the $`𝓐(x)`$ and $`\mathrm{\Omega }(x)`$ the finite temperature effective action remains local.
## III The scalar field in 2+1 dimensions
In what follows we will apply the previous ideas to the computation of the effective action of $`2+1`$-dimensional scalar particles at finite temperature and density. The Euclidean action of the system is that of eq. (1). There, the field $`\varphi (x)`$ is a Lorentz scalar and a vector in the internal symmetry space which will collectively be referred to as flavor space, with dimension $`N_f`$. The covariant derivative is defined as $`D_\mu =_\mu +A_\mu `$, where the gauge field $`A_\mu (x)`$ is an antihermitian matrix in flavor space. Correspondingly, the gauge transformations $`U(x)`$ are unitary matrices. The mass $`m`$ is a space-time constant and a real c-number in flavor space. The more general case of an arbitrary scalar field $`M(x)`$ replacing $`m`$ will not be considered here. The effective action is given by eq. (2).
Relevant symmetries of the problem are pseudo-parity and gauge transformations. Pseudo-parity corresponds to changing every Lorentz zeroth index, i.e, $`(x_0,𝒙)(x_0,𝒙)`$ and $`A_0A_0`$. Since the spectrum of the Klein-Gordon operator $`D_\mu ^2+m^2`$ is unchanged under this transformation the $`\zeta `$-function regularization prescription provides a pseudo-parity preserving effective action. Such Euclidean effective action contains only contributions with an even number of Lorentz zeroth indices, it does not contain the Levi-Civita pseudo-tensor, and thus it is purely real. Any other renormalization prescription can only produce imaginary contributions which are local polynomials. As noted, the $`\zeta `$-function regularized effective action will be strictly gauge invariant since it is real. In fact no essential anomalies are present in the case of scalar fields in 2+1 dimensions (scale anomalies are absent in odd dimensions ).
There is a third symmetry, namely, the transformation $`mm`$ which is trivial for scalar particles and again free from anomaly using $`\zeta `$-function regularization. Within other renormalization schemes there can appear terms breaking this symmetry but they will be removable by adding a local polynomial. In the case of odd-dimensional fermions neither pseudo-parity nor the transformation $`mm`$ are symmetries, however their product gives the parity transformation. Parity is a symmetry of the fermionic action but is not a similarity transformation of the Dirac operator, so it is not guaranteed to be preserved by the $`\zeta `$-function renormalization prescription. As is well-known, parity for odd dimensional fermions is in general in conflict with invariance under large gauge transformations and if the latter invariance is enforced, parity may present an anomaly, depending on the number of flavors .
### A The 0+1 dimensional model
The above-mentioned remarks can be illustrated with the 0+1 dimensional version of the system. The corresponding effective action has been computed in . Perhaps the simplest way to derive this effective action is by computing the partition function of the associated 0-dimensional Hamiltonian system in a gauge where $`A_0`$ is time independent. The energy spectrum is then obtained directly from the Klein-Gordon equation written as
$$_0\varphi =(\pm m+A_0)\varphi .$$
(54)
Thus for each flavor there are two single particle levels $`ϵ_{\pm ,a}=\pm m+A_{0,a}`$ ($`a=1,\mathrm{},N_f`$ labelling the eigenvalues of the matrix $`A_0`$ in flavor space). The standard textbook result for the partition function of a system of non-interacting bosons then applies
$$Z_s[m,A]=\underset{\sigma =\pm }{}\underset{a=1}{\overset{N_f}{}}\underset{n=0}{\overset{\mathrm{}}{}}e^{\beta (\sigma m+A_{0,a})n},$$
(55)
or equivalently
$$W_s[m,A]=\mathrm{tr}\mathrm{log}\left[\left(1e^{\beta (m+A_0)}\right)\left(1e^{\beta (m+A_0)}\right)\right].$$
(56)
The trace refers to flavor space. This result can be rewritten as
$$W_s[m,A]=\beta \mathrm{tr}(A_0)+\mathrm{\Gamma }_s[m,A].$$
(57)
The first term is the 0+1 dimensional Chern-Simons action which breaks pseudoparity and can be removed by a local polynomial counterterm. The second term is (up to a constant)
$$\mathrm{\Gamma }_s[m,A]=\mathrm{tr}\mathrm{log}\left[4\mathrm{sinh}\left(\frac{\beta }{2}(m+A_0)\right)\mathrm{sinh}\left(\frac{\beta }{2}(mA_0)\right)\right].$$
(58)
This effective action is an even function of $`m`$ and $`A_0`$, so it preserves parity and pseudo-parity. It can be written in a manifestly gauge invariant form as
$$\mathrm{\Gamma }_s[m,A]=\mathrm{tr}\mathrm{log}\left[e^{\beta m}+e^{\beta m}\mathrm{\Omega }\mathrm{\Omega }^1\right].$$
(59)
As in the case of fermions, periodicity of the effective action as a function of $`\beta A_0`$, would be lost within a perturbative expansion, i.e. an expansion in powers of $`A_0`$.
Since $`\mathrm{\Gamma }_s[m,A]`$ enjoys all symmetries of the action it coincides with the $`\zeta `$-function regularized effective action, up to a constant (since any other local polynomial must be of degree one in $`m`$ or $`A_0`$ and would break parity). The result in corresponds to $`\mathrm{\Gamma }_s[m,A]\mathrm{\Gamma }_s[m,0]`$ in Minkowski space. It is noteworthy that the partition function defined directly from the Hamiltonian breaks pseudo-parity (due to the Chern-Simons term in eq. (57)) even if no ultraviolet divergences are introduced in 0+1 dimensions in the canonical formalism.
### B Computation of the effective action in 2+1 dimensions. Relation to the fermionic case
The effective action of the 2+1 dimensional model cannot be computed in closed form for arbitrary space-time and internal symmetry configurations. Our approach will be to expand the effective action in the number of spatial covariant derivatives, or equivalently, in the number of spatial Lorentz indices. The computation will be carried out through second order, that is, we will keep terms with zero or two Lorentz indices. (There are no odd order terms in the expansion.) The zeroth Lorentz index dependency is treated exactly since this guarantees the preservation of the periodicity condition which is essential for gauge invariance.
The calculation can be done by applying the $`\zeta `$-function regularization prescription with the help of the symbols method, as described in , however, it is more economical to use the results already established for fermions in that reference. This can be done as follows. The effective action for fermions is
$$W_f[m,A]=\mathrm{Tr}_f\mathrm{log}(\gamma _\mu D_\mu +m).$$
(60)
The functional trace is taken in the space of antiperiodic wave functions and includes Dirac degrees of freedom, in addition to space-time and flavor degrees of freedom. The gamma matrices are hermitian and satisfy $`\gamma _\mu \gamma _\nu =\delta _{\mu \nu }+\sigma _{\mu \nu }`$. Actually, there are two inequivalent irreducible representations of the Dirac algebra which are distinguished by the label $`\eta =\pm 1`$ in the relation $`\gamma _\mu \gamma _\nu \gamma _\rho =i\eta ϵ_{\mu \nu \rho }`$. So if $`\gamma _\mu `$ is one of the representations, $`\gamma _\mu `$ provides another inequivalent representation of the Dirac algebra. The label $`\eta `$ is attached to the Levi-Civita pseudo-tensor and thus a change in $`\eta `$ is equivalent to a pseudo-parity transformation. Therefore the fermionic effective action can be split into two components
$$W_f[m,A]=W_f^+[m,A]+\eta W_f^{}[m,A],$$
(61)
where $`W_f^+`$ is real and even under pseudo-parity and $`W_f^{}`$ is imaginary and pseudo-parity odd. (Of course, this relation can be violated by adding a local polynomial.) Next, note that the formal identity $`\mathrm{Tr}\mathrm{log}(AB)=\mathrm{Tr}\mathrm{log}(A)+\mathrm{Tr}\mathrm{log}(B)`$ holds for the functional trace up to ultraviolet divergent contributions and so it holds modulo local polynomial terms. This implies that
$`W_f^+[m,A]`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{Tr}_f\mathrm{log}\left[(\gamma _\mu D_\mu +m)(\gamma _\mu D_\mu +m)\right]`$ (62)
$`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{Tr}_f\mathrm{log}\left[D_\mu ^2+m^2{\displaystyle \frac{1}{2}}\sigma _{\mu \nu }F_{\mu \nu }\right],(F_{\mu \nu }=[D_\mu ,D_\nu ]).`$ (63)
So $`W_s[m,A]`$ (cf. eq. (2)) is closely related to $`W_f^+[m,A]`$. The differences between both expressions are i) the Dirac degree of freedom which is absent in the scalar case, ii) the different (periodic versus antiperiodic) boundary conditions and iii) the extra term $`\frac{1}{2}\sigma _{\mu \nu }F_{\mu \nu }`$ which is not present in the Klein-Gordon operator. In addition, a local polynomial action can be further added at the end.Alternatively, one can choose to change $`mm`$ instead of $`\gamma _\mu \gamma _\mu `$ in the second factor in the logarithm in eq. (63), and then relate $`W_s[m,A]`$ to $`W_f[m,A]+W_f[m,A]`$. Up to a local polynomial, this procedure is equivalent to the one used in the text.
The first correction amounts to dividing the fermionic result by the trace of unity in Dirac space which is two in the 2+1 dimensional calculation in . The second correction can also be tackled straightforwardly. The Matsubara frequency (related to $`_0`$ in $`D_0`$) takes discrete values $`2\pi i(n+\frac{1}{2})/\beta `$ for fermions and $`2\pi in/\beta `$ for bosons, thus the functional trace computed in the fermionic Hilbert space is related to the bosonic one after the replacement $`A_0A_0\frac{i\pi }{\beta }`$. This is equivalent to $`\mathrm{\Omega }(x)\mathrm{\Omega }(x)`$. ($`\mathrm{\Omega }(x)`$ has been defined in eq. (28).)
$$W_s[m,A]=W_f^+[m,A_0\frac{i\pi }{\beta },𝑨]C[m,A].$$
(64)
The term $`C[m,A]`$ takes into account the spurious contributions coming from $`\frac{1}{2}\sigma _{\mu \nu }F_{\mu \nu }`$, which have to be removed from the fermionic result.
This formula can be illustrated in the 0+1 dimensional model, where it reads $`W_s[m,A]=2W_f^+[m,A_0\frac{i\pi }{\beta }]`$ (note that Dirac space is 1-dimensional in 0+1 dimensions so the factor 2 is not canceled in this case, and also $`C[m,A]=0`$). The simplest way to obtain the 0+1 dimensional fermionic effective action is again using the Hamiltonian formalism (fixing $`_0A_0=0`$). Since there is a single-particle level with energy $`ϵ_a=\eta m+A_{0,a}`$ for each flavor (where $`\eta =\pm 1`$ is the Dirac matrix $`\gamma _0`$ in 0+1 dimensions) it follows that
$$W_f[m,A]=\mathrm{log}\underset{a=1}{\overset{N_f}{}}\underset{n=0,1}{}e^{\beta (\eta m+A_{0,a})n}=\mathrm{tr}\mathrm{log}\left[1+e^{\beta (\eta m+A_0)}\right].$$
(65)
The result in corresponds (up to a local polynomial) to $`W_f[m,A]W_f[m,0]`$ with $`\eta =1`$. This version of the effective action does not directly satisfy eq. (61), i.e. the pseudo-parity transformation $`A_0A_0`$ is not equivalent to the transformation $`\eta \eta `$ in the previous formula. However, subtracting an appropriate $`\eta `$-dependent polynomialTo wit, $`\theta (\eta )𝑑x_0\mathrm{tr}(\eta m+A_0)`$, which is temperature independent. This is consistent with the fact that the finite temperature does not introduce new ultraviolet divergences. yields
$$W_f^{}[m,A]=\mathrm{tr}\mathrm{log}\left[1+e^{\beta (m+\eta A_0)}\right]=\mathrm{tr}\mathrm{log}\left[1+e^{\beta m}\mathrm{\Omega }^\eta \right],$$
(66)
which does satisfy eq. (61). This is the $`\zeta `$-function result . It is readily verified that $`\mathrm{\Gamma }_s[m,A]`$ in eq. (59) coincides with $`2W_f^+[m,A_0\frac{i\pi }{\beta }]`$ plus a polynomial, $`𝑑x_0\mathrm{tr}(m)`$.
In 2+1 dimensions the subtraction $`C[m,A]`$ can be computed in an expansion in powers of $`\frac{1}{2}\sigma _{\mu \nu }F_{\mu \nu }`$. The term of first order does not contribute (since the trace of $`\sigma _{\mu \nu }`$ in Dirac space vanishes) thus the leading term is that with two powers of $`F_{\mu \nu }`$, namely, the expression in eq. (40). Because $`C[m,A]`$ is ultraviolet finite it is free from anomalies, i.e., it enjoys all symmetries of the bosonic action. This term has been computed, up to two spatial derivatives, in subsection II F.
Let us quote the results for the pseudo-parity even component of the effective action of fermions in 2+1 dimensions . At zeroth order in the number of spatial covariant derivatives the result is
$$W_{f,0}[m,A]=\frac{1}{4\pi }\mathrm{tr}0|\left[\left(\frac{2}{\beta }\right)^2m\varphi _1(\frac{\beta }{2}(mD_0))\left(\frac{2}{\beta }\right)^3\varphi _2(\frac{\beta }{2}(mD_0))\right]|0+\mathrm{p}.\mathrm{p}.\mathrm{c}.$$
(67)
At second order
$`W_{f,2}^+[m,A]`$ $`=`$ $`{\displaystyle \frac{1}{8\pi }}\mathrm{tr}0|[({\displaystyle \frac{1}{2}}\left({\displaystyle \frac{2}{\beta }}\right)^2\varphi _1({\displaystyle \frac{\beta }{2}}(mD_{01})){\displaystyle \frac{2}{\beta }}\varphi _0({\displaystyle \frac{\beta }{2}}(mD_{01}))({\displaystyle \frac{1}{4}}(D_{02}D_{01})+{\displaystyle \frac{m}{2}})`$ (70)
$`{\displaystyle _m^{\mathrm{}}}dt\mathrm{tanh}({\displaystyle \frac{\beta }{2}}(tD_{01})){\displaystyle \frac{\left(\frac{1}{4}(D_{02}D_{01})^2+m^2\right)}{2t+D_{02}D_{01}}}){\displaystyle \frac{1}{(D_{02}D_{01})^3}}𝑬^2]|0`$
$`+X_{12}+\text{p.p.c.}`$
In these formulas, p.p.c. means pseudo-parity conjugate, $`D_0D_0`$, and $`X_{12}`$ means the same expression exchanging the labels 1 and 2.
The functions $`\varphi _n(z)`$ are given by
$$\varphi _n(\omega )=P_{n+1}(\omega )_\omega ^+\mathrm{}𝑑z\frac{(\omega z)^n}{n!}\left(\mathrm{tanh}(z)1\right),n=0,1,2,\mathrm{},\mathrm{Re}(\omega )>0,$$
(71)
where the integration path runs parallel to the real positive axis towards $`+\mathrm{}`$. The $`P_n(\omega )`$ are polynomials of degree $`n`$
$`P_1(\omega )`$ $`=`$ $`\omega ,`$ (72)
$`P_2(\omega )`$ $`=`$ $`{\displaystyle \frac{1}{2}}\omega ^2{\displaystyle \frac{1}{6}}\left({\displaystyle \frac{i\pi }{2}}\right)^2,`$ (73)
$`P_3(\omega )`$ $`=`$ $`{\displaystyle \frac{1}{6}}\omega ^3{\displaystyle \frac{1}{6}}\left({\displaystyle \frac{i\pi }{2}}\right)^2\omega .`$ (74)
These formulas for $`\varphi _n(\omega )`$ refer to $`\mathrm{Re}(\omega )>0`$. When $`\mathrm{Re}(\omega )<0`$ the property $`\varphi _n(\omega )=(1)^n\varphi _n(\omega )`$ can be used.
### C The effective action
We can now use eq. (64) to obtain the effective action of the $`2+1`$-dimensional scalar field. Up to two Lorentz indices, $`C[m,A]`$ equals $`C_2[m,A]`$ which is given in eq. (47). $`W_f^+[m,A]`$ is given in eqs. (67) and (70). In these latter expressions the functions $`\varphi _n(z)`$ are made explicit using eq. (71), and the variables $`\mathrm{\Omega }`$ and $`𝒟_0`$ are used instead of $`D_{01}`$ and $`D_{02}`$ (cf. subsection II C). In addition, we introduce a chemical potential by means of the replacement $`A_0A_0\mu `$ where $`\mu `$ is a real constant c-number (recall that $`A_0`$ is antihermitian). This shift is gauge invariant and it is equivalent to $`\mathrm{\Omega }e^{\beta \mu }\mathrm{\Omega }`$.
The effective action up to two spatial covariant derivatives at finite temperature and density is thus
$`W_{s,0}[m,A]`$ $`=`$ $`{\displaystyle d^3x\mathrm{tr}\left[\phi _0(e^{\beta \mu }\mathrm{\Omega })\right]},`$ (75)
$`W_{s,2}[m,A]`$ $`=`$ $`{\displaystyle d^3x\mathrm{tr}\left[\phi _2(e^{\beta \mu }\mathrm{\Omega }_1,𝒟_{02})𝑬^2\right]}.`$ (76)
The functions $`\phi _0`$ and $`\phi _2`$ are given by
$$\phi _0(\omega )=\frac{1}{4\pi }\left(\frac{1}{3}|m|^3+_{|m|}^+\mathrm{}𝑑t(t^2m^2)\frac{\omega }{e^{\beta t}\omega }\right)+\mathrm{p}.\mathrm{p}.\mathrm{c}.$$
(77)
$`\phi _2(\omega ,y)`$ $`=`$ $`{\displaystyle \frac{1}{8\pi }}{\displaystyle \frac{1}{y^3}}[{\displaystyle \frac{1}{2}}|m|y+{\displaystyle \frac{1}{2}}({\displaystyle \frac{1}{4}}y^2m^2)\mathrm{log}\left({\displaystyle \frac{2|m|+y}{2|m|y}}\right)`$ (80)
$`+{\displaystyle _{|m|}^{\mathrm{}}}dt({\displaystyle \frac{2\omega }{e^{\beta t}\omega }}({\displaystyle \frac{t^2m^2}{2ty}}{\displaystyle \frac{1}{2}}y){\displaystyle \frac{2\omega }{e^{\beta (t+y)}\omega }}({\displaystyle \frac{t^2m^2}{2t+y}}+{\displaystyle \frac{1}{2}}y))]`$
$`+\text{p.p.c.}`$
In these formulas p.p.c. corresponds to $`yy`$ and $`\omega \omega ^1`$.
The formulas (75) and (76), expanded in eqs. (77) and (80), constitute the main result of this section. They are necessarily complicated looking since an infinite number of Feynman graphs (with any number of temporal gauge fields, cf. Section IV B) are being added, and the effective action is non-local in time at finite temperature. A much simpler formula is presented below if only the lowest order is retained in a further expansion in the number of temporal covariant derivatives.
Some remarks are in order. In writing the formula we have already used the fact that the effective action is an even function of $`m`$ (since this is already true for $`W_f^+[m,A]`$). The condition $`\mathrm{Re}(\omega )>0`$ in eq. (71), implies that the previous formulas refer to the physically relevant case $`|\mu |<|m|`$ only. All polynomial contributions introduced by $`P_n(\omega )`$ in eq. (71) combine in such a way that all the polynomial dependence on $`D_{01}`$ cancels and the result is a periodic function of $`D_{01}`$. This cancellation requires the explicit and as well as the p.p.c. terms to hold and it is a non-trivial check of the formulas. (Of course, the periodicity holds also for the fermionic effective actions in the pseudo-parity even sector, $`W_f^+[m,A]`$.) The argument of the logarithm in $`\phi _2(\omega ,y)`$ is to be taken in the interval $`(\pi ,\pi )`$ and $`t`$ runs on the real positive axis. (Note that $`y`$ represents $`𝒟_0`$ and thus it is purely imaginary.)
### D Expansion in space-time derivatives at finite temperature
A simpler expression for $`W_{s,2}[m,A]`$ is obtained retaining only the leading order in an expansion in $`𝒟_0`$. This corresponds to expand in powers of $`y`$ in the function $`\phi _2(\omega ,y)`$. As noted this expansion does not break any symmetry and is a natural one in the present context. This produces
$$W_{s,2}[m,A]=d^3x\mathrm{tr}\left[\phi _{2,0}(e^{\beta \mu }\mathrm{\Omega })𝑬^2\right]+O(𝒟_0),$$
(81)
with
$`\phi _{2,0}(\omega )`$ $`=`$ $`{\displaystyle \frac{1}{96\pi }}{\displaystyle \frac{1}{m}}{\displaystyle \frac{e^{2\beta m}\omega ^22\beta me^{\beta m}\omega }{(e^{\beta m}\omega )^2}}+\text{p.p.c.}`$ (82)
$`=`$ $`{\displaystyle \frac{1}{96\pi }}{\displaystyle \frac{1}{m}}{\displaystyle \frac{d}{dm}}\left[m{\displaystyle \frac{e^{\beta m}+\omega }{e^{\beta m}\omega }}\right]+\text{p.p.c.}`$ (83)
Note that $`\phi _{2,0}(\omega )`$ is an even function of $`m`$. This formula is manifestly invariant under all symmetries of the action. Two non-trivial checks of the calculation are i) that negative powers of $`y`$ have canceled and ii) that $`\phi _{2,0}(\omega )`$ no longer contains parametric integrals (i.e., $`_{|m|}^{\mathrm{}}𝑑t`$). As noted, this expression is exact for $`W_{s,2}[m,A]`$ in the Abelian and stationary case.
### E Zero temperature limit
The zero temperature limit of eqs. (77) and (80) is straightforward: because $`|\mu |<|m|`$ and $`\mathrm{\Omega }`$ is unitary, the term $`e^{\beta t}`$ always dominates, thus the limit $`\beta \mathrm{}`$ just removes the terms with $`_{|m|}^{\mathrm{}}𝑑t`$ and the dependence in $`\mu `$ is also canceled (we are assuming $`\mu `$ fixed as $`T0`$). That is
$`W_{s,0}[m,A]`$ $`=`$ $`{\displaystyle \frac{1}{6\pi }}{\displaystyle d^3x\mathrm{tr}\left[|m|^3\right]},(T=0)`$ (84)
$`W_{s,2}[m,A]`$ $`=`$ $`{\displaystyle \frac{1}{8\pi }}{\displaystyle d^3x\mathrm{tr}\left[𝑬\left(|m|𝒟_0+\left(\frac{1}{4}𝒟_0^2m^2\right)\mathrm{log}\left(\frac{2|m|+𝒟_0}{2|m|𝒟_0}\right)\right)\frac{1}{𝒟_0^3}𝑬\right]}.`$ (85)
At zero temperature Lorentz invariance is a symmetry of the action. Such symmetry is not obvious in $`W_{s,2}[m,A]`$ since it sums all orders in $`𝒟_0`$ but not in $`𝒟_i`$, $`i=1,2`$. However, considering an expansion in inverse powers of $`m`$ (a Lorentz invariant expansion) gives
$`W_{s,2}[m,A]`$ $`=`$ $`{\displaystyle \frac{1}{48\pi }}{\displaystyle \frac{1}{|m|}}{\displaystyle d^3x\mathrm{tr}\left[𝑬^2\right]}+O(m^3).`$ (86)
The same result follows from eq. (81). This expression admits a unique Lorentz invariant completion, namely
$$W_s[m,A]=d^3x\mathrm{tr}\left[\frac{1}{6\pi }|m|^3+\frac{1}{96\pi }\frac{1}{|m|}F_{\mu \nu }^2\right]+O(m^3).$$
(87)
This formula comes from a direct application of the results in thereby being a check for our formulas. At next order in $`1/m`$ several Lorentz invariant operators of dimension 6 can appear and eq. (85) puts a constraint on their coefficients. Note that the last two formulas hold also at finite temperature since the temperature-dependent corrections are $`O(e^{\beta |m|})`$ (see also Section IV C.)
### F The partition function
The effective action at finite temperature and density is directly related to the grand-canonical potential, namely $`W_s[m,A;\beta ,\mu ]=\beta \mathrm{\Omega }(\beta ,\mu )`$. (Note that $`\mu `$ introduced by the replacement $`_0_0\mu `$ couples to the charge and not to the number of particles which is not conserved in the relativistic case ). However, strictly speaking, a system at equilibrium with temperature $`T`$ and chemical potential $`\mu `$ can only be stationary. In addition, the physical effect of a (negative) constant external scalar potential is indistinguishable from that of a (positive) chemical potential, since both add to the energy for positive charges and subtract for negative ones, thus for the partition function $`A_0`$ should not be Wick rotated, and $`A_0`$ is real instead of imaginary. All expressions depend only on the combination $`A_0\mu `$. (It would be algebraically inconsistent not to rotate $`A_0`$ to its Euclidean version in the general case, but not within the subset of stationary configurations.) Note that the fact that $`A_0`$ is real or imaginary does not affect the functional form; in any case the functional derivative of the effective action with respect to $`A_0`$ yields the charge density and the derivative with respect to $`\mu `$ yields the total charge:
$$\rho (x)=\frac{\delta W_s}{\delta A_0(x)},Q=\frac{1}{\beta }\frac{W_s}{\mu }.$$
(88)
In view of the relation between partition function and effective action, it follows that $`W_{s,0}`$ describes a two-dimensional relativistic ideal gas in presence of an almost space-time constant scalar potential $`A_0`$. An explicit calculation of the charge density using $`W_{s,0}`$ in eq. (77) and integrating by parts yields
$$\rho _0=\frac{N_f}{4\pi }_{|m|}^{\mathrm{}}𝑑tt\left(\mathrm{coth}(\frac{\beta }{2}(t\mu ))\mathrm{coth}(\frac{\beta }{2}(t+\mu ))\right),$$
(89)
$`N_f=\mathrm{tr}(1)`$ is the number of flavors. We are assuming that $`\mu `$ couples equally to all flavors and we have dropped $`A_0`$ since it can be recovered from $`\mu `$. This formula can rewritten in the standard form
$$\rho _0=\frac{d^2k}{(2\pi )^2}\left(\frac{N_f}{e^{\beta (\omega (k)\mu )}1}\frac{N_f}{e^{\beta (\omega (k)+\mu )}1}\right),$$
(90)
where $`\omega (k)=\sqrt{k^2+m^2}`$.
Likewise, the term $`W_{s,2}`$ adds a contribution to the density coupled to $`A_0`$. Let us consider the Abelian case (in addition to stationary), then $`𝒟_0`$ vanishes and eq. (81) becomes exact. There will be two contributions to the charge density, one coming from the explicit dependence on $`A_0`$ and another though the dependence in $`𝑬`$. The latter is a total derivative and does not contribute to the total charge. Both contribution can be combined to give
$`\rho _2(x)`$ $`=`$ $`{\displaystyle \frac{N_f}{96\pi }}{\displaystyle \frac{1}{m}}{\displaystyle \frac{d}{dm}}(m[{\displaystyle \frac{\beta }{2}}\mathrm{cosech}^2({\displaystyle \frac{\beta }{2}}(m\mu +A_0))𝑬^2`$ (92)
$`+2\mathrm{coth}({\displaystyle \frac{\beta }{2}}(m\mu +A_0))\mathbf{}𝑬])\text{p.p.c.}`$
(Note that, $`A_0`$, $`\mu `$ and $`𝑬`$ are pseudoparity odd.) At zero temperature (assuming $`|m|>|\mu A_0|`$, or else at finite temperature but large mass) this simplifies to
$$\rho _2(x)=\frac{N_f}{24\pi }\frac{1}{|m|}\mathbf{}𝑬,$$
(93)
which also follows from eq. (86).
It is also interesting to note the relation of our results with other formulas using polylogarithms.<sup>\**</sup><sup>\**</sup>\**The polylogarithms are defined as
$$\mathrm{Li}_n(z)=\underset{k=1}{\overset{\mathrm{}}{}}\frac{z^k}{k^n}.$$
The functions $`\varphi _n(\omega )`$ introduced in and subsection III B are directly related to polylogarithms, namely
$$\varphi _n(\omega )=P_{n+1}(\omega )(2)^n\mathrm{Li}_{n+1}(e^{2\omega }).$$
(94)
This relation is easily established from eq. (71) by noting that it verifies the following defining properties of $`\mathrm{Li}_n(z)`$:
$$\mathrm{Li}_0(z)=\frac{z}{1z},z\frac{d}{dz}\mathrm{Li}_n(z)=\mathrm{Li}_{n1}(z),\mathrm{Li}_n(0)=0.$$
(95)
In this notation, $`W_{s,0}`$ in eqs. (75) and (77) becomes
$$W_{s,0}[m,A]=\frac{1}{4\pi }d^3x\mathrm{tr}\left[\frac{|m|^3}{3}+2T^3\left(\beta |m|\mathrm{Li}_2(z)+\mathrm{Li}_3(z)\right)\right]+\mathrm{p}.\mathrm{p}.\mathrm{c}.$$
(96)
where $`z=e^{\beta (|m|\mu +A_0)}`$ is the fugacity, and the corresponding density becomes
$$\rho _0=\frac{N_fT^2}{2\pi }\left(\beta |m|\mathrm{Li}_1(z)+\mathrm{Li}_2(z)\right)\mathrm{p}.\mathrm{p}.\mathrm{c}.$$
(97)
in agreement with.
As a final comment related to the partition function, we note that the relation eq. (64) can also be written using the chemical potential instead of the scalar potential $`A_0`$. That is, if $`\mathrm{\Omega }_b(\beta ,\mu )`$ and $`\mathrm{\Omega }_f(\beta ,\mu )`$ represent the grand-canonical potentials of a system of non interacting particles in presence of external fields, treated as bosons or fermions respectively, then
$$\mathrm{\Omega }_f(\beta ,\mu )=\mathrm{\Omega }_b(\beta ,\mu +\frac{i\pi }{\beta }).$$
(98)
The minus sign comes because the functional integral with Grassman variables gives the determinant of the quadratic form instead of the inverse determinant. The shift $`\mu \mu +\frac{i\pi }{\beta }`$ accounts for the different boundary conditions. Of course a shift $`\mu \mu +\frac{2\pi i}{\beta }`$ must leave the partition function invariant, since $`\mu `$ is coupled to an integer-quantized charge. This is another manifestation of gauge invariance. For interacting particles eq. (98) can be extended using the well-known prescription of adding a minus sign for each particle loop (eq. (98) corresponds to the particular case of one-loop).<sup>††</sup><sup>††</sup>††An equivalent procedure would be to compute the grand-canonical potential for $`n`$ replicas of the particles and then set $`n`$ to $`1`$ at the end. That is, in the notation of the Hamiltonian becomes
$$H=\underset{\alpha ,\beta }{}h_{\alpha \beta }\underset{\sigma =1}{\overset{n}{}}a_{\alpha \sigma }^{}a_{\beta \sigma }+\underset{\alpha ,\beta ,\gamma ,\delta }{}v_{\alpha \beta \gamma \delta }\underset{\sigma _1,\sigma _2=1}{\overset{n}{}}a_{\alpha \sigma _1}^{}a_{\beta \sigma _2}^{}a_{\delta \sigma _2}a_{\gamma \sigma _1}.$$
(99)
## IV Further general remarks on the method
### A Remarks on gauge invariance
As noted in the introduction, topologically large gauge transformations have played a prominent role in the development of this field by putting severe constraints on the allowable forms of the effective action functional. On the other hand it seems that in our present approach such a role is played instead by the periodicity constraint which prompts the appearance of the Polyakov loop $`\mathrm{\Omega }(x)`$. In this subsection we will make some remarks to try to clarify the relation between both concepts.
Let $`G`$ be the gauge group, and let $`\mathrm{\Gamma }`$ be the relevant homotopy group controlling the existence of topologically large gauge transformations at finite temperature. In $`0+1`$-dimensions $`\mathrm{\Gamma }=\pi _1(G)`$, therefore there are large gauge transformations for the Abelian group U(1) but not for simply connected groups such as SU($`n`$). In $`d+1`$-dimensions (with $`d`$ different from zero), let us assume for this discussion that the spatial boundary conditions are such that the space becomes effectively compactified to a sphere. Hence $`\mathrm{\Gamma }`$ contains classes of mappings from the space-time manifold $`\mathrm{S}^1\times \mathrm{S}^d`$ into the gauge group $`G`$, where the factor $`\mathrm{S}^1`$ corresponds to the compactified Euclidean time and $`\mathrm{S}^d`$ to the $`d`$-dimensional space. (Note that, unless otherwise stated, we will require all functions to be continuous on the space-time manifold, and in particular, periodic as a function of time.) Theorem 4.4 of then implies that<sup>‡‡</sup><sup>‡‡</sup>‡‡When $`d=0`$ the theorem is consistent with $`\mathrm{\Gamma }=\pi _1(G)\times \pi _1(G)`$ which follows from $`\mathrm{S}^0=\{1,1\}`$. Physically, we want the spatial manifold at $`d=0`$ to be just {1} and so $`\mathrm{\Gamma }=\pi _1(G)`$.
$$\mathrm{\Gamma }/\pi _{d+1}(G)=\pi _1(G)\times \pi _d(G).$$
(100)
Let us consider the case $`d=2`$. Because the group $`\pi _2(G)`$ is always trivial for any (compact) Lie group $`G`$, this simplifies to $`\mathrm{\Gamma }/\pi _3(G)=\pi _1(G)`$. Two typical cases are
* $`G=\mathrm{U}(1)`$. In this case $`\pi _3`$ is trivial and $`\pi _1`$ is $``$, so $`\mathrm{\Gamma }=`$. There are non-trivial gauge transformations which wind $`n`$ times around the $`\mathrm{S}^1`$ factor of the space-time. They can be realized by space-independent, but time-dependent, gauge transformations.
* $`G=\mathrm{SU}(n)`$ ($`n2`$). In this case $`\pi _1`$ is trivial but $`\pi _3=`$ and so once again $`\mathrm{\Gamma }=`$. In this case the corresponding large gauge transformations must be space-time dependent.
Note that time-independent gauge transformations are controlled by the homotopy group $`\pi _2(G)`$ which is trivial, and so they are always topologically small in two spatial dimensions.
Let us now turn to the point of view used in this work. The periodicity constraint refers to the fact that a gauge invariant functional must depend on $`\mathrm{\Omega }(x)`$ and not just on $`\mathrm{log}(\mathrm{\Omega }(x))`$. The cleanest way to formalize this is by working on the gauge $`_0A_0=0`$, in which $`\mathrm{\Omega }=\mathrm{exp}(\beta A_0(𝒙))`$. Thus both $`\mathrm{\Omega }`$ and $`A_0`$ are time-independent in this gauge. Let us remark that taking $`A_0`$ to be time-independent is not merely a restriction on the set of possible gauge field configurations, it is a choice of gauge in the sense that every gauge field configuration admits a gauge transformed configuration which is $`A_0`$-stationary . In this gauge the periodicity constraint expresses that a gauge invariant functional must be a periodic functional of $`A_0`$. To be concrete consider a generic expression of the form (cf. eq. (53))
$$\mathrm{\Gamma }[M,A]=\underset{n}{}d^{d+1}x\mathrm{tr}[g_n(A_0)𝒪_n],$$
(101)
where $`𝒪_n`$ are gauge covariant local operators, then gauge invariance requires the functions $`g_n(z)`$ to be periodic with period $`2\pi i/\beta `$. The necessity of this requirement follows immediately from considering the following class of gauge transformations
$$U(x)=\mathrm{exp}(x_0\mathrm{\Lambda }(𝒙)),$$
(102)
where the time-independent function $`\mathrm{\Lambda }(𝒙)`$ takes values on the Lie algebra of $`G`$ and is restricted by the following conditions
$$[A_0(𝒙),\mathrm{\Lambda }(𝒙)]=0,\mathrm{exp}(\beta \mathrm{\Lambda }(𝒙))=1.$$
(103)
The second condition means that the eigenvalues of $`\mathrm{\Lambda }(𝒙)`$ are of the form $`\lambda _j=2\pi in_j/\beta `$, for integer $`n_j`$ (such integers are $`𝒙`$-independent by continuity) and it ensures that the corresponding $`U(x)`$ is periodic in the temporal direction. Under such a gauge transformation
$$A_0^U(𝒙)=A_0(𝒙)+\mathrm{\Lambda }(𝒙),$$
(104)
i.e. the spectrum of $`A_0`$ is shifted by multiples of $`2\pi i/\beta `$ and the functions $`g_n(z)`$ must be periodic.
Before proceeding, an important point should be noted regarding the approach used in this work. Namely, at the price of working with asymptotic expansions, we can afford to derive formulas which are “universal” in the sense that no restriction is put on the algebraic properties of the internal space, in particular, the formulas must hold for any gauge group. This means for instance that the functions $`g_n(z)`$ above are the same for all theories and configurations. The requirement of universality puts stronger constraints on the functionals that cannot be appreciated when working with concrete theories only (for instance, in particular theories some of the operators $`𝒪_n`$ can vanish identically and so the corresponding function $`g_n(z)`$ does not play a role).
The transformations introduced in eq. (102), subjected to the conditions eq. (103), have been named discrete transformations associated to $`A_0(𝒙)`$ in since in general they form a discrete set due to the condition $`\lambda _j=2\pi in_j/\beta `$. Note that this set depends on the particular time-independent field $`A_0`$, through the condition $`[A_0,\mathrm{\Lambda }]=0`$. Clearly, these transformations leave invariant the gauge condition $`_0A_0=0`$. Likewise, the gauge condition is also preserved by time-independent gauge transformations. In it is proven that these two kinds of transformations are the most general ones which preserve the $`A_0`$-stationary condition.<sup>\**</sup><sup>\**</sup>\**This is the generic case which we will assume. It holds whenever $`\mathrm{exp}(\beta A_0(𝒙))`$ is either nowhere degenerated or at least the regions of degeneracy are sufficiently small that a unique eigenbasis can be selected (up to normalization) by continuity . In this case $`A_0(𝒙)`$ is also nowhere degenerated and thus $`\mathrm{\Lambda }(𝒙)`$ is completely determined by its eigenvalues. So generically the discrete transformations form a discrete set. This means that within this gauge a functional such as $`\mathrm{\Gamma }[M,A]`$ above is gauge invariant if and only if the functions $`g_n(z)`$ are periodic. Then it can be written in a manifestly gauge invariant form (without gauge fixing) as in the right-hand side of eq. (53) with $`\phi _n(e^{\beta z})=g_n(z)`$. Note that invariance under time-independent gauge transformations do not impose further constraints on the $`g_n(z)`$.
The previous discussion suggests a comparison between large gauge transformations, on the one hand, and discrete transformations, on the other. Two questions pose themselves at this point. Are the discrete transformations large in the topological sense? Is it necessary to rely on large gauge invariance in order to arrive to an $`A_0`$-stationary gauge?
The first question can be answered as follows . For a multiply-connected group such as U(1), the non-trivial discrete transformations are topologically large since they loop once or more on the temporal circle S<sup>1</sup>. On the other hand, for a simply-connected group such as SU($`n`$) (in more than one space-time dimension) the discrete transformations associated to some gauge configuration may be large or small, depending on $`A_0(𝒙)`$. For instance if $`A_0(𝒙)`$ is everywhere diagonal, $`\mathrm{\Lambda }(𝒙)`$ is also diagonal and in fact a constant. In this case the discrete transformation describes a single loop on the gauge group for all $`x`$ and it is homotopically trivial. In general, however, discrete transformations can be topologically large. It might seem that the form of the discrete gauge transformations in eq. (102) factorizes time and space and so it is always classified by the homotopy group $`\pi _1(G)`$ being always small for a simply connected group. This is true when $`\mathrm{\Lambda }(𝒙)`$ is constant or homotopic to a constant, but in general this is not the case. The reason is that although $`\mathrm{\Lambda }(𝒙)`$ is a map from S<sup>2</sup> into the Lie algebra of $`G`$ (a vector space, and thus contractile), it cannot be contracted to a point since the spectrum of $`\mathrm{\Lambda }(𝒙)`$ is constrained to be in $`\frac{2\pi i}{\beta }`$. An explicit SU(2) example in $`2+1`$ dimensions is provided in, namely,
$$U(x)=\mathrm{exp}\left(\frac{2\pi nx_0}{\beta }i𝝉𝒙\right),n,$$
(105)
where $`𝝉`$ are the Pauli matrices and $`𝒙`$ lies on the unit sphere S<sup>2</sup> in $`^3`$. $`U(x)`$ covers SU(2) $`2n`$ times and thus it is homotopically large for non-vanishing $`n`$.
Regarding the second question, whether a given gauge configuration can be brought to a $`A_0`$-stationary gauge using only small transformations, it also depends on the group and the configuration . For the group U(1), any gauge configuration is in the same homotopy class as one which is $`A_0`$-stationary. For a simply-connected group such as SU($`n`$), it depends on the initial $`A_0(x)`$.<sup>\*†</sup><sup>\*†</sup>\*†This can be seen as follows. Let $`A_0(𝒙)`$ be some $`A_0`$-stationary configuration for which all its discrete transformations are small, and let us further assume that time-independent transformations are also small (for instance $`d=2`$). It follows that all $`A_0`$-stationary configurations related to the previous one by a gauge transformation are in the same homotopy class. Thus, if $`A_0^U(x)`$ is a gauge transformed configuration with $`U(x)`$ large, no small transformation will bring it to the $`A_0`$-stationary gauge.
An apparent paradox arises here. As emphasized in , although perturbation theory breaks large gauge invariance, it respects invariance under small transformations, or equivalent, under infinitesimal ones. (Actually, this is strictly correct for Abelian theories only. In non-Abelian theories infinitesimal gauge transformations mix different orders, however, the mixing is mild since it only involves finite sets of Feynman graphs.) On the other hand, we have just seen that expanding a functional such as $`\mathrm{\Gamma }[M,A]`$ in powers of $`A_0`$ (which is a perturbative expansion) destroys periodicity and thus gauge invariance under discrete transformations. This looks paradoxical since the breaking occurs even if the discrete transformations are topologically small, and no breaking was expected in this case.
The resolution comes from the observation that the right-hand side of eq. (101) refers solely to configurations in the gauge $`_0A_0=0`$. Within this gauge the only allowed infinitesimal transformations are the time-independent ones, for which no breaking occurs. Non trivial discrete gauge transformations, even small ones, cannot be reached continuously within the $`A_0`$-stationary gauge. In this sense they are always topologically large. It should be realized that the concept of homotopically trivial is a relative one. A transformation which is topologically large within the gauge group $`G`$ can become small if $`G`$ is regarded as a subgroup of a larger group $`G^{}`$ and deformations within $`G^{}`$ are allowed. Conversely, a small discrete gauge transformation becomes homotopically non-trivial if one insists on preserving the gauge condition $`_0A_0=0`$. The usefulness of the concept of small and large transformations remains: perturbation theory provides a functional valid in the region of small fields, this region is preserved by infinitesimal transformations and thus the perturbative functional must be small gauge invariant. Large transformations, on the other hand, necessarily move the gauge configuration away from the perturbative region and thus the response of the perturbative functional under large transformations is not trustworthy. As a consequence invariance under large transformations provides useful non-perturbative information and puts non-trivial constraints on the functional. This holds whether the transformations are large from the point of view of $`G`$ or from the point of view of the submanifold of $`A_0`$-stationary configurations.
For another argument, we can recall our previous remark that a functional such as $`\mathrm{\Gamma }[M,A]`$ in eq. (101) must hold for all theories at the same time. It is not surprising to find a breaking of gauge invariance under discrete transformations in a perturbative expansion, when such transformations happen to be large, and conclude that non-trivial non-perturbative conditions, namely, periodicity, are required to avoid the breaking. However, as we have emphasized, the property of being topologically large or small depends on the group, whereas the formula must hold in all cases, and so periodicity must follow in all cases too.
In this subsection we have considered a gauge fixing condition in order to deal with the quantity $`\mathrm{log}\mathrm{\Omega }`$ in a simple way. We must recall, however, that the expressions derived with the method studied in this work are all fully gauge invariant, provided some regularity conditions are met, since they depend on $`\mathrm{\Omega }`$. This does not mean that large or discrete gauge transformations play no role whatsoever. This is because the regularity conditions fail for the ultraviolet divergent pieces of the effective action. This translates into the fact that the functions $`\phi _n(\mathrm{\Omega })`$ (cf. eq. (53)), can be many-valued (although $`\mathrm{\Omega }=0`$ needs not be one of the branching points). A typical example is the effective action of $`0+1`$ dimensional fermions, eq. (66). Further let us consider the Abelian U(1) case, which admits large transformations and they coincide with the non-trivial discrete transformations in the $`A_0`$-stationary gauge. Choosing $`\eta =+1`$ the branching point is at $`\mathrm{\Omega }=\mathrm{exp}(\beta m)`$. Under a large transformation, the argument may go into a different Riemann sheet (depending on the sign of $`m`$), yet the functional is such that it changes, at most, by an integer multiple of $`2\pi i`$, reflecting the fact that the partition functional is one-valued. This is a general property which puts constraints on the functions $`\phi _n(\mathrm{\Omega })`$.
Perhaps this is a good place to remark that the gauge invariance implied by the use of the $`\zeta `$-function prescription, eq. (51), has two levels . One the one hand, there is the somewhat trivial gauge invariance implied by the fact that the regularization depends solely on the spectrum. Since the spectrum of the operator $`K`$ is left unchanged by gauge (or more generally similarity) transformations, gauge invariance follows. However, the definition of the $`\zeta `$-function introduces a branch cut in the manifold of operators $`K`$, each singular operators being a branching point on such a manifold. On a given Riemann sheet, the effective action functional may display a jump discontinuity along the branch cut. Because the determinant of $`K`$, being the regularized product of eigenvalues, is a smooth functional, it follows that the jump must be an integer multiple of $`2\pi i`$. This is a tighter constraint on top of the trivial gauge invariance noted above. For instance, we have noted in the introduction that perturbation theory for fermions at finite temperature yields a Chern-Simons term which is renormalized by a temperature dependent coefficient. Under large gauge transformations this would introduce an unacceptable change in the effective action by a quantity which is not a multiple of $`2\pi i`$. It would be tempting to “solve” the problem by simply replacing the Chern-Simons term by a suitable gauge invariant version of it, namely, the $`\eta `$-invariant. This prescription restores gauge invariance but introduces jumps which again are proportional to the temperature dependent coefficient and thus it can be ruled out. The exact result known in particular but non-trivial cases shows that this is not the correct mechanism and that the determinant is a continuous functional, with no jumps.
### B Feynman graphs and large gauge invariance
We have already noted in subsection III C that expressions such as those in eqs. (75,76) involve an infinite number of Feynman graphs. It seems interesting to understand which Feynman graphs are being added and gain some insight on how preservation of full gauge invariance is related to this.
To this end, we will first consider the simpler case of $`0+1`$-dimensional fermions. The corresponding exact effective action is given in eq. (66) and that formula holds for arbitrary gauge fields which need not be Abelian nor stationary. As noted, when the gauge group is not simply connected it supports large gauge transformations, which augment the value of the effective action by $`2\pi ikn`$, $`n,k`$, where $`n`$ is the winding number of the gauge transformation and $`k`$ depends on the gauge group and the sign of the mass (in higher spatial dimensions it may also depend on the homotopic class of the gauge field configuration).
Because the $`0+1`$-dimensional formula is exact, all Feynman graphs are included in this case. This suggests that all graphs are required in order to reconstruct the Polyakov loop $`\mathrm{\Omega }`$ appearing in the formula. We conclude that large gauge invariance does not act selecting a certain subset of graphs. The same conclusion is expected to hold in higher dimensional formulas since the dependence on $`\mathrm{\Omega }`$ there is qualitatively similar to that of the one-dimensional case.
To further discuss this point let us restrict ourselves to the case of an Abelian gauge group. In this case large gauge transformations correspond to discrete shifts of $`a=\mathrm{log}\mathrm{\Omega }=𝑑x_0A_0`$ and
$$W_f(a+2\pi in)=W_f(a)+2\pi ikn,$$
(106)
where the integer constant $`k`$ is known. This equation contains all the information on large gauge transformations, and it is completely equivalent to the statement
$$W_f(a)=P(a)+ka,P(a+2\pi in)=P(a).$$
(107)
Therefore, large gauge invariance is equivalent to the strict periodicity of the function $`P(a)=W_f(a)ka`$. Feynman graphs correspond to expand in powers of $`a`$, and looking for large gauge invariance in terms of Feynman graphs corresponds to detect periodicity of a function from its Taylor expansion. This seems to be a difficult task.
A related issue is studying to what extent large gauge invariance determines the effective action. In the previous $`0+1`$-dimensional Abelian case we have seen that to comply with gauge invariance, $`P(a)`$ must be periodic, i.e.,
$$P(a)=\underset{n}{}c_ne^{na},$$
(108)
for some (infinite number of) coefficients $`c_n`$. No further information can be extracted from gauge invariance, and the coefficients $`c_n`$ are not determined. In order to achieve further restrictions on the function $`P(a)`$, more information has to be provided. If, for instance, one knows that the partition function $`Z(a)=\mathrm{exp}(W_f(a))`$ (a periodic function) contains only a finite number of periodic modes, the corresponding Fourier coefficients can then be determined from a few perturbative terms. This is actually the case in $`0+1`$-dimensions , and can be traced back to the fact that the corresponding Hamiltonian contains a finite number of states, namely, the vacuum or the 1-fermion state (cf. Section III B). In higher dimensions, besides the number of fermions, there is a momentum quantum number and a corresponding kinetic energy contributing to the eigenvalues of the Hamiltonian, thus in general the partition function will contain all kinds of Fourier modes. To see how this works, it is sufficient to consider fermions in the Abelian case with $`A_i=0`$ and $`A_0`$ a space-time constant (such $`A_0`$ cannot be gauged away at finite temperature). Let $`ϵ_k^0`$ denote the single-particle levels of the Hamiltonian when $`A_0`$ is set to zero (that is, the kinetic energy only) and let $`ϵ_k`$ be the levels when $`A_0`$ is switched on. Clearly, $`ϵ_k=ϵ_k^0+A_0`$, and the label $`k`$ is related to the momentum of the fermion. The partition function is
$$Z[m,a]=\underset{k}{}(1+e^{\beta ϵ_k})=\underset{k}{}(1+\mathrm{\Omega }e^{\beta ϵ_k^0}),\mathrm{\Omega }=e^a.$$
(109)
$`Z[m,a]`$ will be periodic in $`a`$ but with many periodic modes, unless $`d=0`$. When $`d=2`$, integration over the label $`k`$ yields a particular case of the formula for $`W_{f,0}[m,A]`$ in eq. (67).
In more than one dimension we have proposed an expansion in the number of spatial covariant derivatives. The zeroth order (e.g. eq. (75)), which contains no spatial indices, corresponds to the sum of all Feynman graphs with the spatial momenta $`p_i`$ (of the external gauge fields) set to zero and no legs with spatial $`A_i`$, but any number of $`A_0`$ external legs and the full dependence in the frequency $`p_0`$. There are no odd-order terms in the expansion. The second order (e.g. eq. (76)) corresponds to all graphs with two $`A_i`$ or $`p_i`$, that is, i) graphs with $`p_i=0`$ and two spatial gauge fields, plus ii) graphs with one $`A_i`$ and $`p_i`$ kept up to first order in a Taylor expansion of the Green functions, plus iii) graphs with no legs $`A_i`$ and $`p_i`$ kept up to second order in a Taylor expansion around zero spatial momentum. Note that the method assumes that the space has topology $`^d`$. The contribution of higher orders follows a similar pattern. Eventually, all contributions, all Feynman graphs, are added up.
The previous conclusions follow from inspection of the formulas or else from making the expansion by introducing a bookkeeping parameter in order to count spatial indices, as explained in Section II C. For the simpler formulas obtained by further expanding in powers of $`𝒟_0`$, (e.g. eq. (81)), we have noted that this expansion does not seem to follow from insertion of a bookkeeping parameter and so different expansions can be obtained, all of them being equivalent when added to all orders. The problem and its interpretation in terms of Feynman graphs can be seen in a particularly simple case: let us assume that the gauge group is Abelian, that $`A_0`$ is a space-time constant and $`A_i`$ are space-independent although time-dependent. In this case the Green functions depend only on the frequencies $`\kappa _n=2\pi in/\beta `$ of the fields $`A_i`$:
$$W=f_0(\mathrm{\Omega })+\underset{n}{}f_2(\mathrm{\Omega },\kappa _n)A_{i,n}A_{i,n}+\mathrm{},A_i(x_0)=\frac{1}{\beta }\underset{n}{}e^{\kappa _nx_0}A_{i,n}.$$
(110)
In addition, $`𝒟_0`$ is equivalent to $`_0`$ and so an expansion in $`𝒟_0`$ is just a Taylor expansion in powers of $`\kappa _n`$, to be made on top of the expansion in powers of $`A_i`$. This gives the interpretation in terms of Feynman graphs. Because $`\kappa _n`$ is a discrete variable, the function $`f_2`$ (and similarly for higher orders) is only well-defined at those discrete values. The ambiguity comes when it is smoothly extended to continuous values of $`\kappa _n`$ in order to carry out the Taylor expansion. Presumably this ambiguity can be removed by using the choice suggested by Carlson’s theorem.
### C Large mass expansions
Large mass expansions of the effective action can be also considered, as done in for fermions using the heat-kernel technique. Inspection of our formulas in III C and III E, show that as the mass becomes large the temperature dependence is exponentially suppressed, being $`O(e^{\beta |m|})`$, thus the large mass expansion is an asymptotic expansion with temperature independent coefficients. This is consistent with the fact that the coefficients are local operators independent of the global topology of the space-time manifold . Therefore, in order to carry out a large mass expansion one can simplify and start with a zero temperature theory. The simplification of working at zero temperature is enormous to the point that this problem can be considered a solved one. There is a very large body of work on this subject, both for bosons and for fermions, largely summarized in and references therein. Large mass expansions for fermions with arbitrary Dirac operators and arbitrary dimension, computed along the lines of the method discussed in this paper for finite temperature, can be found in . The large mass expansion can also be used as a check of the finite temperature formulas. In III E we have already noted that eq. (87) is consistent with the result more straightforwardly obtained from the zero temperature method presented in . Likewise, for $`2+1`$-dimensional fermions, and starting from the full finite temperature calculation, one finds
$`W^+`$ $`=`$ $`{\displaystyle \frac{1}{48\pi }}{\displaystyle \frac{1}{|m|}}{\displaystyle d^3x\mathrm{tr}(F_{\mu \nu }^2)}+O\left({\displaystyle \frac{1}{m^3}}\right),`$ (111)
$`W^{}`$ $`=`$ $`\eta \sigma \mathrm{\Theta }(\sigma m)W_{\mathrm{CS}}{\displaystyle \frac{i\eta }{8\pi }}{\displaystyle \frac{\epsilon (m)}{12m^2}}{\displaystyle d^3xϵ_{\mu \nu \alpha }\mathrm{tr}(F_{\beta \mu }𝒟_\alpha F_{\beta \nu })}+O\left({\displaystyle \frac{1}{m^3}}\right).`$ (112)
for the pseudo-parity even and odd components, respectively. (In these formulas $`W_{\mathrm{CS}}`$ is the Chern-Simons term, $`\mathrm{\Theta }`$ and $`\epsilon `$ denote the step and sign functions,respectively, $`\eta =\pm 1`$ depends on the irreducible representation of the Dirac gamma matrices taken and $`\sigma =\pm 1`$ distinguish the two possible $`\zeta `$-function definitions of the effective action, depending on the branch cut in the function $`z^s`$.) These results also derive more directly from the zero temperature formulas in . The term with $`F_{\mu \nu }^2`$ in $`W^+`$ is that with $`H_4`$ in eq. (4.11) of , already noted there, whereas the term with $`F_{\beta \mu }𝒟_\alpha F_{\beta \nu }`$ is that with $`P_5`$ in eq. (4.12) of the same reference.
## V Summary and conclusions
Our findings can be summarized as follows.
1. The one-loop effective action at finite temperature and density, for bosonic or fermionic particles in presence of background fields (both gauge and non-gauge) with arbitrary internal symmetry group and arbitrary space-time dependence, can be written as a sum (an asymptotic series in general) of terms ordered by the number of spatial Lorentz indices. Each term is well-defined from the effective action functional itself and is separately gauge invariant under all gauge transformations.
2. These terms are amenable to explicit computation using a combination of $`\zeta `$-function and symbols method. We have shown that this kind of calculations can be carried out preserving full gauge invariance throughout, without assuming particular internal symmetry groups or special space-time configurations for the background fields. The same arguments show that previous calculations done fixing the gauge through the condition $`_0A_0=0`$ can be repeated lifting this condition, and this is equivalent to rewrite the final original result in a manifestly gauge invariant way.
3. A further gauge invariant expansion can be taken in the number of temporal covariant derivatives in the adjoint representation. Within this expansion, all ultraviolet finite terms and more generally, all terms not related to essential anomalies, can be written as a sum of gauge invariant local operators (i.e., constructed with $`𝒟_\mu =[D_\mu ,]`$ and $`M`$) times a function of the field $`\mathrm{\Omega }(x)`$, which also transforms locally under gauge transformations. For those terms containing anomalies, topological pieces and multivaluation, the effective action still looks local in terms of a suitable gauge covariant version of $`𝑨(x)`$, in addition to $`M`$, $`𝒟_\mu `$ and $`\mathrm{\Omega }`$.
4. The method is explicitly applied to the problem of relativistic scalar particles in $`2+1`$ dimensions. The corresponding effective action is computed up to terms with two spatial Lorentz indices. The result is checked against the known result at zero temperature and also the known partition function of a relativistic Bose gas. The corrections to the density are also computed. Finally, a simple rule is noted relating the bosonic and fermionic versions of the grand-canonical potentials of ideal or interacting systems.
## Acknowledgments
This work was supported in part by funds provided by the Spanish DGICYT grant no. PB98-1367 and Junta de Andalucía grant no. FQM-225.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.