id
stringlengths 27
33
| source
stringclasses 1
value | format
stringclasses 1
value | text
stringlengths 13
1.81M
|
---|---|---|---|
warning/0006/cond-mat0006315.html
|
ar5iv
|
text
|
# Untitled Document
SUPERSYMMETRY IN THERMO FIELD DYNAMICS
R.Parthasarathy and R.Sridhar<sup>1</sup><sup>1</sup>1e-mail address: sarathy,sridhar@imsc.ernet.in
The Institute of Mathematical Sciences
C.P.T.Campus, Taramani Post
Chennai 600 113, India.
Abstract
By considering the enlarged thermal system including the heat bath, it is shown that this system has supersymmetry which is not broken at finite temperature. The superalgebra is constructed and the Hamiltonian is expressed as the anti-commutator of two kinds of super charges. With this Hamiltonian and the thermal vaccum $`0(\beta )>`$, this supersymmetry is found to be preserved.
I. Introduction
Supersymmetry at finite temperature has been studied in detail by Girardello et.al (1981), Van Hove (1982), Das (1989), Witten (1982), Das and Kaku (1978), Teshima (1983), Umezawa, Matsumoto and Tachiki (1982), and Umezawa (1993). Nevertheless, the issue of whether supersymmetry is broken or not, at finite temperature has raised some controversy. Girardello et.al (1981) argued that supersymmetry (SUSY) is broken at positive temperature even when unbroken at $`T=0`$. In response to this, Van Hove (1982), suggested that when a change of an operator under SUSY transformation at finite temperature is considered, one should take into account the Klein operator. When this operator is incorporated, Van Hove (1982) shows that the thermal average of this change of an operator is zero for all $`T`$, thereby maintaining SUSY at finite temperature. On the other hand, Das (1989) has considered this issue within the context of the real time formalism or Thermo Field Dynamics (TFD) of Umezawa (Umezawa et.al (1982), Takahashi and Umezawa (1996)) and concluded that SUSY is broken at finite temperature, by evaluating the statistical average of the SUSY Hamiltonian at $`T=0`$ as its vacuum expectation value in the ’thermal vacuum $`0(\beta )>`$ (where $`\beta =1/kT`$, $`k`$ being the Boltzmann constant) and showing that it is non-zero at finite temperature.
It is the purpose of this work to examine first the construction of SUSY algebra in TFD and then use it to understand SUSY at finite temperature by enlarging the Fock space including that of the tilde operators. This procedure is analogous to the treatment of $`\widehat{G}`$-symmetry by Umezawa and will be explained later. We will exhibit mathematical possibilities of preserving supersymmetry at finite temperature in agreement with Van Hove.
II. Brief Outline of Thermo Field Dynamics
In thermofield dynamics (Umezawa et.al (1982), Umezawa (1993) and Takahashi and Umezawa (1996)), the ’thermal vacuum’ expectation value of an operator is equated to its statistical average. The ’thermal vacuum’ is temperature dependent in a doubled Fock space. This construction procedure leads to the introduction of ’tilde’ operators and the doubled Fock space is a direct product of the two Fock spaces for non-tilde and tilde creation and annihilation operators. This doubling nature is one of the most fundamental and universal feature of all of the thermal quantum field formalism. For an ensemble of free Bosons with frequency $`\omega `$, one has the Hamiltonian
$`H_B`$ $`=`$ $`\omega a^{}a,`$
$`[a,a^{}]`$ $`=`$ $`1,`$ (1)
where $`a`$ and $`a^{}`$ are the annihilation and creation operators for Bosons. One introduces the tilde fields by the Hamiltonian
$`\stackrel{~}{H}_B`$ $`=`$ $`\omega \stackrel{~}{a}^{}\stackrel{~}{a},`$
$`[\stackrel{~}{a},\stackrel{~}{a}^{}]`$ $`=`$ $`1,`$
$`[a,\stackrel{~}{a}]=[a,\stackrel{~}{a}^{}]`$ $`=`$ $`0,`$ (2)
where $`\stackrel{~}{a}`$ and $`\stackrel{~}{a}^{}`$ are the annihilation and creation operators for the tilde Bosonic fields. The thermal vacuum is given by
$`0(\beta )>`$ $`=`$ $`(1exp(\beta \omega ))^{\frac{1}{2}}exp(e^{\beta \omega /2}a^{}\stackrel{~}{a}^{})0,\stackrel{~}{0}>,`$ (3)
where a doubling of Fock space is exhibited. The operators $`a,a^{},\stackrel{~}{a},\stackrel{~}{a}^{}`$ pertain to zero temperature. The corresponding operators at finite temperature, namely, $`a(\beta ),a^{}(\beta ),\stackrel{~}{a}(\beta ),\stackrel{~}{a}^{}(\beta )`$ are obtained from $`a,a^{},\stackrel{~}{a},\stackrel{~}{a}^{}`$ by Bogoliubov transformation. It is important to note that while
$`a0(\beta )>`$ $``$ $`0,`$ (4)
we have
$`a(\beta )0(\beta )>`$ $`=`$ $`0,`$ (5)
so that the Fock space at finite temperature is spanned by
$`0(\beta )>,a^{}(\beta )0(\beta )>,\stackrel{~}{a}^{}(\beta )0(\beta )>,{\displaystyle \frac{1}{\sqrt{n!}}}{\displaystyle \frac{1}{\sqrt{m!}}}(a^{}(\beta ))^n(\stackrel{~}{a}^{}(\beta ))^m0(\beta )>.`$ (6)
For an ensemble of Fermions of frequency $`\omega `$ (say), one has similar relations with the commutator replaced by anti-commutator, and the Fock space at finite temperature will be spanned by
$`0(\beta )>,f^{}(\beta )0(\beta )>,\stackrel{~}{f}^{}(\beta )0(\beta )>,f^{}(\beta )\stackrel{~}{f}^{}(\beta )0(\beta )>.`$ (7)
A physical interpretation of the doubling of the degrees of freedom , namely,$`a,a^{}`$ and $`\stackrel{~}{a},\stackrel{~}{a}^{}`$for Bosons and/or $`f,f^{}`$ and $`\stackrel{~}{f},\stackrel{~}{f}^{}`$ for Fermions is the following. When the vacuum $`0(\beta )>`$ is required to be independent of time, as it should be, we choose the total Hamiltonian for free Boson fields in TFD as
$`\widehat{H}_B`$ $`=`$ $`{\displaystyle d^3k\omega _k\{a_k^{}a_k\stackrel{~}{a}_k^{}\stackrel{~}{a}_k\}}.`$ (8)
To an external stimulus, at $`T0`$, certain number of quantum particles are condensed in this system, which is in thermal equillibrium state with temperature $`T`$. The absorption of the external energy by the system results in (1) absorption by excitation of additional quanta and (2) excitation of quantum present in the vacuum, creating a hole. This is how one has doubling in TFD. In fact, Umezawa, Matsumoto and Tachiki (1982) attribute the excitation of holes to that in the thermal reservoir.
In studying the properties of dynamical observables, it is expected to use the non-tilde operators. However, in studying the symmetry properties of the system, one needs both the tilde and non-tilde operators. This is emphazised in Umezawa (1993), in studying the spontaneous breakdown of $`\widehat{G}`$ symmetry (that of the Bogoliubov transformation). In this Reference, the $`\widehat{G}`$ symmetry is defined to be spontaneously broken when $`\widehat{G}0(\beta )>0`$ while $`[\widehat{G},\widehat{H}_0]=0`$. It is to be noted that the hat-Hamiltonian is used which has the tilde operators as well (see sections 7.2.4 and 7.3). For a system of free Fermions, the total Hamiltonian (analogue of (8)) is
$`\widehat{H}_F`$ $`=`$ $`{\displaystyle d^3k\omega _k\{f_k^{}f_k\stackrel{~}{f}_k^{}\stackrel{~}{f}_k\}}`$ (9)
with the (Fermion) Fock space in (7).
III. Supersymmetric Algebra
Following Van Hove (1982), we expect supersymmetry to have non-trivial consequences not only at $`T=0`$ but also at finite temperature, since all excited states must also some how reflect the supersymmetry property. In view of this we wish to examine the possibility of maintaining supersymmetry at finite temperature as well. By considering the enlarged Fock space (6) and (7), and using (8) and (9), we will arrive at this possibility, agreeing with Van Hove (1982).
We demonstrate this by considering a system of free Bosons and free Fermions. At zero temperature, the Bosons are described by the creation and annihilation operators $`a^{}`$ and $`a`$ and the corresponding tilde operators satisfying the algebra
$`[a,a^{}]=1`$ ; $`[a,a]=0,`$
$`[\stackrel{~}{a},\stackrel{~}{a}^{}]=1,[\stackrel{~}{a},\stackrel{~}{a}]=0`$ ; $`[a,\stackrel{~}{a}]=0,[a,\stackrel{~}{a}^{}]=0,`$ (10)
and similarly the Fermions are described by $`f,f^{},\stackrel{~}{f},\stackrel{~}{f}^{}`$ satisfying the algebra
$`\{f,f^{}\}=1`$ , $`f^2=(f^{})^2=0,`$
$`\{\stackrel{~}{f},\stackrel{~}{f}^{}\}=1`$ , $`\stackrel{~}{f}^2=(\stackrel{~}{f}^{})^2=0,`$
$`\{f,\stackrel{~}{f}\}`$ $`=`$ $`\{f,\stackrel{~}{f}^{}\}=0.`$ (11)
We construct Fermionic(super) charge operators (generators of Supersymmetry) as
$`Q_+=af^{}`$ ; $`Q_{}=a^{}f,`$
$`q_+=\stackrel{~}{a}\stackrel{~}{f}^{}`$ ; $`q_{}=\stackrel{~}{a}^{}\stackrel{~}{f}.`$ (12)
These operators are nilpotent, namely, $`Q_+^2=Q_{}^2=q_+^2=q_{}^2=0`$ and convert boson to fermion and vice-versa when acting on the representative state $`n_B,\stackrel{~}{n}_B,n_F,\stackrel{~}{n}_F>`$. The elements of the superalgebra are
$`Q_\pm ,q_\pm ,(N_B+N_F),(\stackrel{~}{N}_B+\stackrel{~}{N}_F),`$ (13)
where $`N_B=a^{}a,N_F=f^{}f,\stackrel{~}{N}_B=\stackrel{~}{a}^{}\stackrel{~}{a},\stackrel{~}{N}_F=\stackrel{~}{f}^{}\stackrel{~}{f}`$. This algebra is closed
$`\{Q_+,Q_{}\}`$ $`=`$ $`N_B+N_F,`$
$`\{q_+,q_{}\}`$ $`=`$ $`\stackrel{~}{N}_B+\stackrel{~}{N}_F,`$
$`\{Q_+,q_+\}=\{Q_{},q_{}\}`$ $`=`$ $`\{Q_+,q_{}\}=\{Q_{},q_+\}=0,`$
$`[Q_\pm ,(N_B+N_F)]`$ $`=`$ $`0,`$
$`[Q_\pm ,(\stackrel{~}{N}_B+\stackrel{~}{N}_F]`$ $`=`$ $`0,`$
$`[q_\pm ,(N_B+N_F)]`$ $`=`$ $`0,`$
$`[q_\pm ,(\stackrel{~}{N}_B+\stackrel{~}{N}_F)]`$ $`=`$ $`0,`$ (14)
satisfying the structure $`\{O,O\}=E,[O,E]=E,[E,E]=E`$ for even (E) and odd (O) operators. The total Hamiltonian for supersymmetric oscillator $`\widehat{H}=a^{}a\stackrel{~}{a}^{}\stackrel{~}{a}+f^{}f\stackrel{~}{f}^{}\stackrel{~}{f}`$ is given by the anti-commutator
$`\widehat{H}`$ $`=`$ $`\{Q_+,Q_{}\}\{q_+,q_{}\},`$ (15)
and $`Q_\pm ,q_\pm `$ are Fermionic constants of motion, i.e.,
$`[Q_\pm ,\widehat{H}]`$ $`=`$ $`[q_\pm ,\widehat{H}]=0.`$ (16)
The supersymmetric vacuum state at zero temperature is
$`0>=0_B,\stackrel{~}{0}_B,0_F,\stackrel{~}{0}_F>`$ and since this vacuum is annihilated by $`a,\stackrel{~}{a},f,\stackrel{~}{f}`$, it follows that
$`<0\widehat{H}0>`$ $`=`$ $`0,`$ (17)
and
$`Q_\pm 0>`$ $`=`$ $`0,`$
$`q_\pm 0>`$ $`=`$ $`0.`$ (18)
Thus the supersymmetry constructed in (13) and (14) is not broken at zero temperature.
IV. Supersymmetry at finite temperature
At finite temperature, we will exhibit three mathematical possibilities to examine whether supersymmetry is broken or not. In view of the structure of vacua at finite temperature in (6) and (7), we choose the thermal vacuum for the supersymmetric case as
$`0(\beta )>`$ $`=`$ $`0_B(\beta ),\stackrel{~}{0}_B(\beta ),0_F(\beta ),\stackrel{~}{0}_F(\beta )>.`$ (19)
The zero-temperature operators $`a,\stackrel{~}{a},f,\stackrel{~}{f}`$ are related to the ’temperature dependent’ operators $`a(\beta ),\stackrel{~}{a}(\beta ),f(\beta ),\stackrel{~}{f}(\beta )`$ which annihilate the above ’thermal vacuum’, by the (inverse) Bogoliubov transformation (Takahashi and Umezawa (1996))
$`a`$ $`=`$ $`u(\beta )a(\beta )+v(\beta )\stackrel{~}{a}^{}(\beta ),`$
$`\stackrel{~}{a}`$ $`=`$ $`u(\beta )\stackrel{~}{a}(\beta )+v(\beta )a^{}(\beta ),`$
$`f`$ $`=`$ $`U(\beta )f(\beta )+V(\beta )\stackrel{~}{f}^{}(\beta ),`$
$`\stackrel{~}{f}`$ $`=`$ $`U(\beta )\stackrel{~}{f}(\beta )V(\beta )f^{}(\beta ),`$ (20)
where
$`u(\beta )`$ $`=`$ $`(1e^{\beta \omega })^{\frac{1}{2}},`$
$`v(\beta )`$ $`=`$ $`(e^{\beta \omega }1)^{\frac{1}{2}},`$
$`U(\beta )`$ $`=`$ $`(1+e^{\beta \omega })^{\frac{1}{2}},`$
$`V(\beta )`$ $`=`$ $`(1+e^{\beta \omega })^{\frac{1}{2}}.`$ (21)
Method.1
In this method, the thermal vacuum is given by (19) and the Hamiltonian by (15) (involving zero temperature operators). Since now we have $`a0(\beta )>0,\stackrel{~}{a}0(\beta )>0,f0(\beta )>0,\stackrel{~}{f}0(\beta )>0`$, we need to use the Bogoliubov transformation (20).
Then it follows
$`a0(\beta )>`$ $`=`$ $`v(\beta )\stackrel{~}{a}^{}(\beta )0(\beta )>,`$
so that,
$`<0(\beta )a^{}a0(\beta )>`$ $`=`$ $`v^2(\beta )<0(\beta )\stackrel{~}{a}(\beta )\stackrel{~}{a}^{}(\beta )0(\beta )>`$
$`=`$ $`v^2(\beta )<0(\beta )1+\stackrel{~}{a}^{}(\beta )\stackrel{~}{a}(\beta )0(\beta )>=v^2(\beta ),`$
and
$`\stackrel{~}{a}0(\beta )>`$ $`=`$ $`v(\beta )a^{}(\beta )0(\beta )>,`$
so that
$`<0(\beta )\stackrel{~}{a}^{}\stackrel{~}{a}0(\beta )>`$ $`=`$ $`v^2(\beta )<0(\beta )a(\beta )a^{}(\beta )0(\beta )>=v^2(\beta ).`$
Similarly, we have from
$`f0(\beta )>`$ $`=`$ $`V(\beta )\stackrel{~}{f}^{}(\beta )0(\beta )>,`$
$`\stackrel{~}{f}0(\beta )>`$ $`=`$ $`V(\beta )f^{}(\beta )0(\beta )>,`$
so that
$`<0(\beta )f^{}f0(\beta )>`$ $`=`$ $`V^2(\beta ),`$
$`<0(\beta )\stackrel{~}{f}^{}\stackrel{~}{f}0(\beta )>`$ $`=`$ $`V^2(\beta ).`$
It then follows from these that
$`<0(\beta )\widehat{H}0(\beta )>`$ $`=`$ $`<0(\beta )a^{}a\stackrel{~}{a}^{}\stackrel{~}{a}+f^{}f\stackrel{~}{f}^{}\stackrel{~}{f}0(\beta )>`$ (22)
$`=`$ $`0,`$
implying that $`\widehat{H}`$ is invariant under supersymmetry. On the other hand,
$`Q_\pm 0(\beta )>`$ $``$ $`0,`$
$`q_\pm 0(\beta )>`$ $``$ $`0.`$ (23)
Thus we realize a situation: while the total Hamiltonian is supersymmetric invariant, the thermal vacuum is not. We realize, ”spontaneous breakdown of supersymmetry”. This is in contrast to the result of Das (1989) in which both his Hamiltonian and the thermal vacuum are not invariant - a case of explicit breaking of supersymmetry.
Method.2
We first motivate the Method.2, by realizing that in the expressions $`<0(\beta )\widehat{H}0(\beta )>`$ and those such as $`Q_\pm 0(\beta )>`$, with $`\widehat{H}`$ given in (15) and $`Q_\pm ,q_\pm `$ in (12), the state vector $`0(\beta )>=0_B(\beta ),\stackrel{~}{0}_B(\beta ),0_F(\beta ),\stackrel{~}{0}_F(\beta )>`$ is in the doubled Fock space with the spectrum $`n_B(\beta ),\stackrel{~}{n}_B(\beta ),n_F(\beta ),\stackrel{~}{n}_F(\beta )>((a(\beta ))^{})^{n_B}((\stackrel{~}{a}(\beta ))^{})^{\stackrel{~}{n}_B}((f(\beta ))^{})^{n_F}((\stackrel{~}{f}(\beta ))^{})^{\stackrel{~}{n}_F}0(\beta )>`$. The states including the vacuum are temperature dependent. On the other hand, the operators $`\widehat{H}`$ and $`Q_\pm ,q_\pm `$ are in terms of temperature independent creation and annihilation operators. In view of this disparity, it is more appropriate to have operators also Bogoliubov transformed so that they act on the same Fock space for examining the symmetry properties of the total system at finite temperature and this is the reason for considering Method.2.
First. we consider the Bogoliubov transformed operators (Takahashi and Umezawa (1996)) which are temperature dependent, as
$`a(\beta )`$ $`=`$ $`u(\beta )av(\beta )\stackrel{~}{a}^{},`$
$`\stackrel{~}{a}(\beta )`$ $`=`$ $`u(\beta )\stackrel{~}{a}v(\beta )a^{},`$
$`f(\beta )`$ $`=`$ $`U(\beta )fV(\beta )\stackrel{~}{f}^{},`$
$`\stackrel{~}{f}(\beta )`$ $`=`$ $`U(\beta )\stackrel{~}{f}+V(\beta )f^{},`$ (24)
which is the inverse of (20) and the functions $`u(\beta ),v(\beta ),U(\beta ),V(\beta )`$ are given in (21). It can be verified that these temperature dependent operators satisfy the algebra
$`[a(\beta ),a^{}(\beta )]`$ $`=`$ $`1,`$
$`[\stackrel{~}{a}(\beta ),\stackrel{~}{a}^{}(\beta )]=1;[a(\beta ),\stackrel{~}{a}(\beta )]=[a(\beta ),\stackrel{~}{a}^{}(\beta )]=0,`$
$`\{f(\beta ),f^{}(\beta )\}=1`$ ; $`f^2(\beta )=(f^{}(\beta ))^2=0,`$
$`\{\stackrel{~}{f}(\beta ),\stackrel{~}{f}^{}(\beta )\}=1`$ ; $`\{f(\beta ),\stackrel{~}{f}(\beta )\}=\{f(\beta ),\stackrel{~}{f}^{}(\beta )\}=0.`$ (25)
With these, we introduce the temperature dependent super charges as
$`Q_+(\beta )=a(\beta )f^{}(\beta )`$ ; $`Q_{}(\beta )=a^{}(\beta )f(\beta ),`$
$`q_+(\beta )=\stackrel{~}{a}(\beta )\stackrel{~}{f}^{}(\beta )`$ ; $`q_{}(\beta )=\stackrel{~}{a}^{}(\beta )\stackrel{~}{f}(\beta ).`$ (26)
They are nil-potent. With the number operators, $`(N_B(\beta )+N_F(\beta ));(\stackrel{~}{N}_B(\beta )+\stackrel{~}{N}_F(\beta ))`$, where $`N_B(\beta )=a^{}(\beta )a(\beta ),\stackrel{~}{N}_B=\stackrel{~}{a}^{}(\beta )\stackrel{~}{a}(\beta ),N_F(\beta )=f^{}(\beta )f(\beta ),\stackrel{~}{N}_F(\beta )=\stackrel{~}{f}^{}(\beta )\stackrel{~}{f}(\beta )`$, they form a closed super algebra, namely,
$`\{Q_+(\beta ),Q_{}(\beta )\}`$ $`=`$ $`N_B(\beta )+N_F(\beta ),`$
$`\{q_+(\beta ),q_{}(\beta )\}`$ $`=`$ $`\stackrel{~}{N}_B(\beta )+\stackrel{~}{N}_F(\beta ),`$
$`\{Q_+(\beta ),q_+(\beta )\}`$ $`=`$ $`\{Q_+(\beta ),q_{}(\beta )\}=0,`$
$`\{Q_{}(\beta ),q_+(\beta )\}`$ $`=`$ $`\{Q_{}(\beta ),q_{}(\beta )\}=0,`$
$`[Q_\pm (\beta ),(N_B(\beta )+N_F(\beta )]`$ $`=`$ $`0,`$
$`[Q_\pm (\beta ),(\stackrel{~}{N}_B(\beta )+\stackrel{~}{N}_F(\beta ))]`$ $`=`$ $`0,`$
$`[q_\pm (\beta ),(N_B(\beta )+N_F(\beta ))]`$ $`=`$ $`0,`$
$`[q_\pm (\beta ),(\stackrel{~}{N}_B(\beta )+\stackrel{~}{N}_F(\beta ))]`$ $`=`$ $`0.`$ (27)
Furthermore, we realize the important relation,
$`\{Q_+(\beta ),Q_{}(\beta )\}\{q_+(\beta ),q_{}(\beta )\}=a^{}(\beta )a(\beta )\stackrel{~}{a}^{}(\beta )\stackrel{~}{a}(\beta )`$
$`+f^{}(\beta )f(\beta )\stackrel{~}{f}^{}(\beta )\stackrel{~}{f}(\beta )\widehat{H}(\beta ).`$ (28)
This important relation expresses the total Hamiltonian at finite temperature as anti-commutator of temperature dependent super charges and therefore $`Q_\pm (\beta ),q_\pm (\beta )`$ commute with $`\widehat{H}(\beta )`$ showing that they are the fermionic constants of motion.
It can be verified upon using (24) that $`\widehat{H}(\beta )`$ in (28) is the same as $`\widehat{H}`$ in (15) showing the invariance of the total Hamiltonian under Bogoliubov transformation. Using, $`a(\beta )0(\beta )>=0;\stackrel{~}{a}(\beta )0(\beta )>=0;f(\beta )0(\beta )>=0;\stackrel{~}{f}(\beta )0(\beta )>=0`$, it follows that
$`<0(\beta )\widehat{H}(\beta )0(\beta )>`$ $`=`$ $`0,`$
$`Q_\pm (\beta )0(\beta )>`$ $`=`$ $`0,`$
$`q_\pm (\beta )0(\beta )>`$ $`=`$ $`0.`$ (29)
Thus, both the total Hamiltonian $`\widehat{H}(\beta )`$ and the thermal vacuum $`0(\beta )>`$ are invariant under supersymmetry and so supersymmetry is not broken at finite temperature, in agreement with Van Hove (1982).
Method.3
The construction of the supersymmetric charges $`Q_\pm ,q_\pm `$ in (12) and
$`Q_\pm (\beta ),q_\pm (\beta )`$ in (26) in examining the supersymmetry breaking or not, using the total Hamiltonian $`\widehat{H}`$ and $`\widehat{H}(\beta )`$ is on mathematical grounds, in the sense that in thermo field dynamics, the observables are analysed in terms of non-tilde operators while the above mathematical procedure includes tilde operators as well. It is still possible to realize unbroken supersymmetry at $`T0`$ without using the tilde operators by restricting the super algebra to
$`Q_\pm (\beta )`$ , $`(N_B(\beta )+N_F(\beta )).`$ (30)
This algebra is closed, namely,
$`\{Q_+(\beta ),Q_{}(\beta )\}`$ $`=`$ $`N_B(\beta )+N_F(\beta ),`$
$`[Q_\pm (\beta ),(N_B(\beta )+N_F(\beta ))]`$ $`=`$ $`0.`$ (31)
The Hamiltonian of the system is identified with
$`H(\beta )`$ $`=`$ $`\{Q_+(\beta ),Q_{}(\beta )\}=a^{}(\beta )a(\beta )+f^{}(\beta )f(\beta ),`$ (32)
so that
$`[Q_\pm (\beta ),H(\beta )]`$ $`=`$ $`0,`$ (33)
giving the fermionic constants of motion with respect to $`H(\beta )`$. The ground state is the thermal vacuum $`0(\beta )>`$ as before. Then it follows
$`<0(\beta )H(\beta )0(\beta )>`$ $`=`$ $`0,`$
$`Q_\pm (\beta )0(\beta )>`$ $`=`$ $`0,`$ (34)
showing that supersymmetry is not broken at finite temperature. The situation at zero temperature in this case is obtained by taking the limit $`\beta \mathrm{}`$. We have from (20) and (21), $`u(\beta )1;v(\beta )0;a(\beta )a;f(\beta )f;H(\beta )H;0(\beta )>0>`$ as $`\beta \mathrm{}`$ and then we recover the zero temperature case and this has supersymmetry unbroken.
V. Summary
We have examined the supersymmetric structure in thermo field dynamics by considering the enlarged Fock space. Supersymmetric generators are constructed and three methods, in which the Hamiltonian is governed by the anti-commutator of super charges, are studied. Besides realizing spontaneous breakdown of supersymmetry in Method.1, there are two possibilities to realize unbroken supersymmetry at finite temperature. These have well defined zero temperature limit in which the supersymmetry is not broken. These results are in agreement with Van Hove. This analysis goes through in the Lagrangian formulation in TFD and since this is straightforward, we have not included this.
Acknowledgement
Useful discussions with R.Anishetty are acknowledged with thanks.
References
Das A, Physica 1989 A158 1.
Das A and Kaku M, 1978 Phys.Rev. D18 4540.
Girardello L, Grisanuand M T, and Salomonson N P, 1981 Nucl.Phys. B178 513.
Takahashi Y and Umezawa H, 1996 Int.J.Mod.Phys. B10 1755.
Teshima K, 1983 Phys.Lett. B123 226.
Umezawa H, Matsumoto H, and Tachiki M 1982 Thermo Field Dynamics, North-Holland, Amsterdam.
Umezawa H, 1993 Advanced Field Theory, AIP, New York.
Van Hove L, 1982 Nucl.Phys. B207 15.
Witten E, 1982 Nucl.Phys. B202 253.
|
warning/0006/cond-mat0006108.html
|
ar5iv
|
text
|
# Bose-Einstein condensation thermodynamics of a trapped gas with attractive interaction
## I Introduction
There has been recent experimental observation of Bose-Einstein (BE) condensation in dilute interacting bosonic atoms of <sup>87</sup>Rb , <sup>23</sup>Na , <sup>7</sup>Li , and <sup>1</sup>H employing magnetic traps at ultra-low temperatures. The interaction among the atoms could be either attractive or repulsive. Although, in both cases the condensate is well-described by the Gross-Pitaevskii (GP) equation , the nature of BE condensates in these two cases is of entirely different nature. In the repulsive case the number of atoms in the condensate can grow without bound, whereas this number is limited by a upper bound in the case of attractive interaction . Of the experimentally observed cases of BE condensation one has repulsion for <sup>1</sup>H, <sup>23</sup>Na, and <sup>87</sup>Rb atoms and attraction for <sup>7</sup>Li atoms. The existence of a maximum number of condensed atoms in the case of <sup>7</sup>Li has been noted experimentally and is consistent with the prediction of theoretical analysis based on the GP equation . In the attractive case the GP equation has no solution for the number of atoms larger than a critical number.
The GP equation is a nonlinear Schrödinger equation and it is tedious to find its converged numerical solution . For the repulsive case, an approximate solution scheme of this equation, such as the one based on the Thomas-Fermi approximation, has frequently been used for a qualitative description of the condensate . However, no such approximation scheme is known for the attractive case which requires an exact numerical solution of the GP equation. This makes the theoretical study of this case a more challenging task.
There have been several comprehensive studies on the temperature dependencies of the thermodynamic observables in the case of repulsive interaction using mean-field two-fluid models . One such mean-field scheme is provided by the so-called Popov approximation and has been considered by several authors . The physical ingredients of these mean-field models are quite similar and they lead to similar numerical results in the case of weakly repulsive interatomic interactions.
We shall use a two-fluid mean-field model to study the temperature dependencies of the thermodynamic observables of the condensate. For a condensate composed of 40000 trapped <sup>87</sup>Rb atoms the perturbative solution of the system of equations of this model converged rapidly and provided a satisfactory account of the condensate fraction, internal energy, and specific heat in agreement with experiment . It was also found that the lowest order solution already provided a very good approximation. Later the same model has been used in one and two space dimensions .
The above mean-field two-fluid model is used in this work for a theoretical description of the BE condensation thermodynamics in the case of attractive interaction appropriate for <sup>7</sup>Li. Depending on the strength of the attractive potential, the condensate in the attractive case may consist of a few thousand atoms confined by the trap potential. For a fixed trap, the maximum number of atoms in the BE condensate with attractive interaction is inversely proportional to $`|a|`$ , where $`a`$ is the scattering length of two atoms. As the temperature is lowered below the critical temperature $`T_0`$ of BE condensation, the condensate starts to form and finally at 0 K all the available atoms (limited by the maximum number mentioned above) form the condensate in the present model.
The condensate wave function in the present model is described by the GP equation. In the repulsive case usually some approximate solutions of the GP equation are used . Although we shall be using the iterative solution of the system of equations of the mean-field model, we shall employ a converged numerical solution of the GP equation in the present attractive case. As the GP equation is a nonlinear one, this amounts to a nontrivial modification of the calculational scheme.
The plan of the paper is as follows. In Sec. II we present the mean-field two-fluid model. In Sec. III we discuss the numerical scheme for its solution and present numerical results for <sup>7</sup>Li. Finally, in Sec. IV we present some concluding remarks.
## II Mean-field Two-fluid Model
We consider a system of $`N`$ bosons with attractive interaction at temperature $`T`$ under the influence of a trap potential. The condensate is described by the following GP equation for the wave function $`\mathrm{\Psi }(𝐫)`$ with eigenvalue $`\overline{\mu }`$ :
$$\left[\frac{\mathrm{}^2^2}{2m}+V_{\text{ext}}(r)+2gn_1(r)+g\mathrm{\Psi }^2(𝐫)\overline{\mu }\right]\mathrm{\Psi }(𝐫)=0.$$
(1)
Here $`V_{\text{ext}}(r)m\omega ^2r^2/2`$ is the spherically symmetric harmonic-oscillator trap potential, $`g4\pi \mathrm{}^2a/m`$ the strength of the interatomic interaction, $`m`$ the mass of a single atom, $`\omega `$ the angular frequency, $`a`$ the atom-atom scattering length, $`\mu \overline{\mu }\overline{\mu }_0`$ is the chemical potential, where $`\overline{\mu }_0`$ is the eigenvalue of Eq. (1) for the harmonic oscillator potential alone in the absence of interatomic interaction ($`g=0`$), and $`n_1(r)`$ represents the distribution function of the noncondensed bosons. Although in actual experiment the harmonic oscillator trap is not quite symmetric, the deviation from spherical symmetry is quite small. However, the converged numerical solution of the GP equation (1) for a nonsymmetric trap is quite complicated numerically and hence for a qualitative description we consider a spherically symmetric trap in the present study. An attractive (repulsive) interaction corresponds to negative (positive) values of $`a`$ and $`g`$.
The noncondensed particles are treated as non-interacting bosons in an effective potential $`V_{\text{eff}}(r)=V_{\text{ext}}(r)+2gn_1(r)+2g\mathrm{\Psi }^2(𝐫)`$ . Thermal averages are calculated with a standard Bose distribution of the noncondensed particles in chemical equilibrium with the condensate governed by the same chemical potential $`\mu `$. In particular the density $`n_1(r)`$ is given by
$$n_1(r)=\frac{1}{(2\pi \mathrm{})^3}\frac{d^3p}{\mathrm{exp}[\{p^2/2m+V_{\text{eff}}(r)\mu \}/k_BT]1},$$
(2)
where $`k_B`$ is the Boltzmann constant. Equations (1) $``$ (2) above are the principal equations of the present model.
The total number of particles $`N`$ of the system is given by
$`N=N_0+{\displaystyle \frac{\rho (E)dE}{\mathrm{exp}[(E\mu )/k_BT]1}},`$ (3)
where $`N_0\mathrm{\Psi }^2(𝐫)d^3r`$ is the total number of particles in the condensate. The semiclassical density of states $`\rho (E)`$ of Eq. (3) is given by
$$\rho (E)=\frac{(2m)^{3/2}}{4\pi ^2\mathrm{}^3}_{V_{\text{eff}}(r)<E}\sqrt{EV_{\text{eff}}(r)}d^3r.$$
(4)
The critical temperature $`T_0`$ is obtained as the solution of Eq. (3) with $`N_0`$ and $`\mu `$ set equal to 0.
The average single-particle energy of the noncondensed particles is given by
$$E_{\text{nc}}=\frac{E\rho (E)dE}{\mathrm{exp}[(E\mu )/k_BT]1}.$$
(5)
The kinetic energy of the condensate is assumed to be negligible and its interaction energy per particle is given by $`E_\text{c}=(g/2)\mathrm{\Psi }^4(𝐫)d^3r`$. The quantity of experimental interest is the average total energy $`E=[E_{\text{nc}}(NN_0)/2+E_\text{c}]/N,`$ which we calculate.
The coupling of the nonlinear equation (1) with other equations make the solution algorithm complicated and it is convenient to express this system of equation in dimensionless units. This has advantage in the numerical solution . We express energy in units of $`\mathrm{}\omega `$, and length in units of the harmonic oscillator length $`a_{\text{ho}}\sqrt{\mathrm{}/(m\omega )}`$. We shall consider in this paper only the spherically symmetric ground state solution of the condensate with $`\mathrm{\Psi }(𝐫)=\mathrm{\Psi }(r)`$. Then the GP equation (1) become
$`\left[{\displaystyle \frac{d^2}{dx^2}}+x^2+4\pi \eta \overline{n}_1(x)\pm {\displaystyle \frac{2\varphi ^2(x)}{x^2}}2\overline{\alpha }\right]\varphi (x)=0,`$ (6)
where $`\eta 4a/a_{\text{ho}}`$ is the new dimensionless strength, $`xr/a_{\text{ho}}`$, $`\mathrm{\Psi }(r)\varphi (x)/(x\sqrt{\pi |\eta |a_{\text{ho}}^3}),`$ and $`\overline{\alpha }\overline{\mu }/(\mathrm{}\omega )`$. The positive (negative) sign in Eq. (6) corresponds to repulsive (attractive) interaction. The dimensionless density $`\overline{n}_1(x)a_{\text{ho}}^3n_1(r)`$ is defined by
$$\overline{n}_1(x)=\frac{1}{2\pi ^2}\frac{k^2dk}{\mathrm{exp}[\{k^2/2+V_{\text{eff}}(x)\alpha \}/t]1},$$
(7)
where $`kpa_{\text{ho}}/\mathrm{}`$, $`V_{\text{eff}}(x)x^2/2+2\pi \eta \overline{n}_1(x)+2\varphi ^2(x)/x^2`$, $`tk_BT/(\mathrm{}\omega )`$, $`\alpha \mu /(\mathrm{}\omega )`$. For the spherically symmetric ground state $`\alpha =\overline{\alpha }\overline{\alpha }_0`$, where $`\overline{\alpha }_0\overline{\mu }_0/(\mathrm{}\omega )=1.5`$ is the energy eigenvalue of the harmonic oscillator potential alone (zero-point energy in units of $`\mathrm{}\omega `$) in the absence of interatomic interaction. In the present consideration of the chemical potential finite-size effects are excluded . Using these dimensionless variables the number equation (3) becomes
$`N=N_0+{\displaystyle \frac{\overline{\rho }(e)de}{\mathrm{exp}[(e\alpha )/t]1}},`$ (8)
where $`N_0=(4/|\eta |)\varphi ^2(x)𝑑x`$, $`eE/(\mathrm{}\omega )`$, and the dimensionless density $`\overline{\rho }(e)\mathrm{}\omega \rho (E)`$ is given by
$$\overline{\rho }(e)=\frac{2\sqrt{2}}{\pi }_{V_{\text{eff}}(x)<e}\sqrt{eV_{\text{eff}}(x)}x^2𝑑x.$$
(9)
The above set of equations (6) $``$ (9) are solved iteratively using the converged numerical solution of the GP equation (6). The iteration is started with $`\overline{n}_1(x)=0`$ at a definite temperature with a trial value for the chemical potential $`\alpha `$. Then Eq. (6) is solved and with its solution the functions $`V_{\text{eff}}(x)`$ and $`\overline{n}_1(x)`$ are calculated. Using these new $`V_{\text{eff}}(x)`$ and $`\overline{n}_1(x)`$ Eq. (6) is solved again and $`\overline{n}_1(x)`$ and $`\varphi (x)`$ are recalculated. This iterative scheme is continued until final convergence is achieved. In each order of iteration we calculate the condensate fraction $`N_0/N`$ and energy $`E`$, in addition to the chemical potential $`\alpha `$. This scheme is repeated until convergence is achieved. It is then verified if the number equation (8) is satisfied with this solution. If not, a new trial value for the chemical potential is employed. Once the number equation is satisfied the desired solution is obtained.
Next we present the solution procedure of the GP equation (6). The solution of this equation satisfies the following boundary conditions
$`\varphi (0)`$ $`=`$ $`0,\varphi ^{}(0)=\text{constant}`$ (10)
$`\underset{x\mathrm{}}{lim}\varphi (x)`$ $`=`$ $`N_C\mathrm{exp}\left[{\displaystyle \frac{x^2}{2}}+(\overline{\alpha }{\displaystyle \frac{1}{2}})\mathrm{ln}x\right],`$ (11)
where $`N_C`$ is a normalization constant. The derivative of the wave function can be obtained from Eq. (11) and one obtains the following log-derivative of the wave function in the asymptotic region
$`\underset{x\mathrm{}}{lim}{\displaystyle \frac{\varphi ^{}(x)}{\varphi (x)}}`$ $`=`$ $`\left[x+\left(\overline{\alpha }{\displaystyle \frac{1}{2}}\right){\displaystyle \frac{1}{x}}\right].`$ (12)
Equation (6) is integrated numerically for a given $`\overline{\alpha }`$ by the four-point Runge-Kutta rule starting at the origin ($`x=0`$) with the initial boundary condition (10) with a trial $`\varphi ^{}(0)`$ in steps of $`dx=0.001`$. Using Eq. (6) the integration is propagated to $`x=x_{\text{max}}`$, where the asymptotic condition (12) is valid. The agreement between the numerically calculated log-derivative of the wave function and the theoretical result (12) is enforced to four significant figures. The maximum value of $`x`$ up to which we need to integrate to obtain this precision is $`x_{\text{max}}=3`$. If for a trial $`\varphi ^{}(0)`$, this precision can not be obtained, a new value of $`\varphi ^{}(0)`$ is to be chosen. The procedure is repeated until the converged solution is obtained.
## III Numerical Results
In our numerical study we would be interested only in the case of attractive interaction (negative $`\eta `$). We consider the experimentally relevant situation of <sup>7</sup>Li . In this case the trap frequencies in the experiment of Ref. along the $`X`$, $`Y`$, and $`Z`$ directions are 150.6 Hz, 152.6 Hz, and 131.5 Hz which lead to a maximum of about 1400 trapped atoms. The deviation from spherical symmetry in this case is small and in order to have a qualitative understanding of the condensate we consider the trap to be spherically symmetric with $`N_{\text{max}}=1400`$. We reconfirm from the numerical solution of the GP equation that
$$\frac{|\eta |N_{\text{max}}}{4}=0.575.$$
(13)
Considering the known result for <sup>7</sup>Li, that $`N_{\text{max}}=1400`$ , we obtain from Eq. (13), $`\eta =0.00164`$ and we use this value of $`\eta `$ in our numerical calculation. In our calculation we use three values of $`N`$, e.g., $`N=1300,1000,`$ and 500 ($`N<N_{\text{max}}=1400`$). First we calculate the critical temperatures in these cases from the number equation (8) and obtain the values $`T_0=10.27\mathrm{}\omega /k_B,9.41\mathrm{}\omega /k_B,7.47\mathrm{}\omega /k_B`$ for $`N=1300,1000`$ and 500, respectively.
In the case of <sup>7</sup>Li, the estimated value of $`\eta (=`$0.00164) is quite small. This corresponds to a very weak coupling and we find that the lowest-order solution is graphically almost indistinguishable from the converged solution for the condensate fraction, chemical potential, and total energy. The estimated coupling $`\eta `$ for <sup>87</sup>Rb is 0.0248 , which is more than ten times larger in magnitude than the coupling for <sup>7</sup>Li. In the case of <sup>87</sup>Rb, already the lowest order result was very good . Hence the very rapid convergence of the present results is not quite unexpected. In the present study we only exhibit the converged result after two iterations.
Fig. 1. Converged condensate fraction $`N_0/N`$ as a function of $`T/T_0`$ for $`\eta =0.00164`$ and $`N=`$ 1300 (full line) and 500 (dashed-dotted line); for $`\eta =0.01`$ and $`N=200`$ (dashed-double-dotted line); and for trapped ideal Bose gas (dotted line).
In Fig. 1 we present the temperature ($`T/T_0`$) dependence of the condensate fraction $`N_0/N`$ for <sup>7</sup>Li with $`\eta =0.00164`$ for $`N=1300`$ and 500. The result for $`N=1000`$ is indistinguishable from that of $`N=1300`$. The result for $`\eta =0`$ corresponding to an ideal Bose gas in a harmonic trap is also shown in this figure. For comparison, the result in the case of a stronger attractive interaction for $`\eta =0.01`$ and $`N=200`$ is also shown. In the case of repulsive interaction, the result for $`N/N_0`$ at a particular temperature $`T/T_0`$ is smaller than that for an ideal Bose gas . In the present case of attractive interaction, the result for $`N/N_0`$ at a particular temperature $`T/T_0`$ is larger than that for the ideal Bose gas. We also find from this figure that, as expected, the result for the stronger attractive interaction ($`\eta =0.01`$) deviates more from the ideal gas result than in the case of <sup>7</sup>Li ($`\eta =0.00164`$). We did not consider a much stronger attractive coupling, as because of Eq. (13) this would correspond to an unacceptably small value for the number of particles in the condensate. This number is already small for the case $`\eta =0.01`$ considered in this work.
As the solution of the GP equation (6) in this case is nontrivial, we show in Fig. 2 the wave functions $`\varphi (x)/x`$ for $`N=1000`$ at temperatures $`T/T_0`$ = 0, 0.4, 0.6, 0.8, and 0.9. As temperature decreases, the wave function is more pronounced corresponding to an increase in the number of particles $`N_0`$ in the condensate given by the normalization $`N_0=(4/|\eta |)\varphi ^2(x)𝑑x.`$ For other values of $`N`$, the wave functions are similar to those in Fig. 2 and we do not show these wave functions here.
Fig. 2. GP wave function $`\varphi (x)/x`$ of Eq. (6) for $`N=1000`$ at different temperatures $`T/T_0`$ for $`\eta =0.00164`$.
Fig. 3. Chemical potential $`\mu /(k_BT_0)`$ as a function of $`T/T_0`$ for $`\eta =0.00164`$ and $`N=`$ 1300 (full line), 1000 (dashed line) and 500 (dashed-dotted line); and for $`\eta =0.01`$ and $`N=200`$ (dashed-double-dotted line).
In Fig. 3 we show the chemical potential of the system at different temperatures $`T/T_0`$ for $`N=1300,1000,`$ and 500 and $`\eta =0.00164`$. For comparison we also show the result for the stronger attractive interaction with $`N=200`$ and $`\eta =0.01`$. Above the critical temperature $`T>T_0`$, the chemical potential for a trapped ideal Bose gas is negative and it becomes zero at the critical temperature . The same is true in the present simplified model where the very weakly interacting noncondensed gas is taken to be noninteracting. The only effect of interaction is considered via the condensate. The effect of the interaction among the atoms of the noncondensed gas could be important for $`T`$ just below $`T_0`$ when there will be a large fraction of noncondensed gas and a small fraction of condensed gas. For $`T`$ close to 0 when most atoms are condensed such effect could be neglected. At present a correct description of BEC thermodynamics including the interaction among noncondensed atoms is beyond the scope of the present work. Hence, although it would be more appropriate to treat the noncondensed gas to be interacting, the effect of interaction in the treatment of the noncondensed gas is expected to be negligibly small in the present study of very weakly interacting Bose gas ($`\eta =0.0164`$) and is neglected. In the case of the ideal Bose gas, below the critical temperature the chemical potential is identically equal to zero. For the present case of the trapped Bose gas with attractive interaction, the chemical potential, after becoming zero at $`T=T_0`$ from a negative value, becomes negative again as the temperature is reduced below the critical temperature. The chemical potential for the stronger attractive interaction ($`\eta =0.01`$) deviates more from the trapped ideal gas result ($`\mu =0`$) below the critical temperature, than in the case of the weak attractive interaction of <sup>7</sup>Li ($`\eta =0.00164`$). In the case of repulsive interaction, the chemical potential becomes positive for temperatures below $`T=T_0`$ .
Fig. 4. Energy $`E/(Nk_BT_0)`$ as a function of $`T/T_0`$ for $`\eta =0.00164`$ and $`N=1300`$ (full line) and 500 (dashed-dotted line); for $`\eta =0.00164`$ and $`N=1300`$ (dotted line); and the classical Maxwell-Boltzmann distribution (straight line).
In Fig. 4 we plot the temperature dependence of energy $`E/(Nk_BT)`$ for $`\eta =0.00164`$ and $`N=1300`$ and 500. These two cases lead to almost identical energies as can be seen from Fig. 4. For comparison, in this figure we also show the result for the repulsive interaction for $`\eta =0.00164`$ and $`N=1300`$. The energy for the trapped ideal Bose gas should lie between the energies for the attractive and repulsive cases mentioned above. The classical Maxwell-Boltzmann result is also shown. For Bose-Einstein condensation to materialize the energy of the system should be lower that the classical result. The energy in the attractive case is smaller than the corresponding repulsive case below critical temperature .
## IV Conclusion
In conclusion, we studied the temperature dependence of condensate fraction, chemical potential, and total energy for a trapped <sup>7</sup>Li gas consisting of 1300, 1000, and 500 atoms with attractive interaction using a mean-field two-fluid model. The maximum number of atoms allowed in this case is 1400 . We employed an iterative solution scheme of the system of equations of this model as in Refs. . In the case of <sup>7</sup>Li the attractive interaction is very weak and the the system of equations leads to rapid convergence. The condensate was described by the converged numerical solution of the Gross-Pitaevskii equation . As the interaction is weak, the results for condensate fraction, chemical potential, and energy of <sup>7</sup>Li are very close to the corresponding results for the trapped ideal Bose gas. However, the deviation from the result of the ideal Bose gas in case of <sup>7</sup>Li is in the opposite direction compared to the corresponding deviation in case of repulsive interaction . For example, in all cases the chemical potential is zero at the critical temperature. For ideal Bose gas it continues to be zero below the critical temperature. Below the critical temperature the chemical potential turns negative in case of <sup>7</sup>Li, whereas it turns positive in case of <sup>87</sup>Rb where the interaction is repulsive. Although, in the present attractive case the number of particles in the condensate is small, the present results for the thermodynamic observables are quite reasonable physically. This demonstates that the mean-field two-fluid models used to study the thermodynamic observables for the BE condensate in the repulsive case are quite useful in the attractive case also.
We thank Dr. A. Gammal for useful discussion. The work is supported in part by the Conselho Nacional de Desenvolvimento Científico e Tecnológico and Fundação de Amparo à Pesquisa do Estado de São Paulo of Brazil.
|
warning/0006/math0006213.html
|
ar5iv
|
text
|
# Nonabelian mixed Hodge structures
## Nonabelian mixed Hodge structures
Ludmil Katzarkov <sup>1</sup><sup>1</sup>1University of California at Irvine, Irvine, CA 92697, USA. Partially supported by NSF Career Award DMS-9875383 and A.P. Sloan research fellowship., Tony Pantev <sup>2</sup><sup>2</sup>2University of Pennsylvania, 209 South 33rd Street Philadelphia, PA 19104-6395, USA. Partially supported by NSF Grant DMS-9800790 and A.P. Sloan Research Fellowship., Carlos Simpson <sup>3</sup><sup>3</sup>3CNRS, Université de Nice-Sophia Antipolis, Parc Valrose, 06108 Nice Cedex 2, France.
Abstract— We propose a definition of “nonabelian mixed Hodge structure” together with a construction associating to a smooth projective variety $`X`$ and to a nonabelian mixed Hodge structure $`V`$, the “nonabelian cohomology of $`X`$ with coefficients in $`V`$” which is a (pre-)nonabelian mixed Hodge structure denoted $`H=Hom(X_M,V)`$. We describe the basic definitions and then give some conjectures saying what is supposed to happen. At the end we compute an example: the case where $`V`$ has underlying homotopy type the complexified $`2`$-sphere, and mixed Hodge structure coming from its identification with $`𝐏^1`$. For this example we show that $`Hom(X_M,V)`$ is a namhs for any smooth projective variety $`X`$.
Introduction—p. Nonabelian mixed Hodge structures
Part I: Nonabelian weight filtrations
Conventions—p. Nonabelian mixed Hodge structures
Nonabelian filtrations—p. Nonabelian mixed Hodge structures
Perfect complexes and Dold-Puppe linearization—p. Nonabelian mixed Hodge structures
Further study of filtered $`n`$-stacks—p. Nonabelian mixed Hodge structures
Filtered and $`𝐆_m`$-equivariant perfect complexes—p. Nonabelian mixed Hodge structures
Weight-filtered $`n`$-stacks—p. Nonabelian mixed Hodge structures
Analytic and real structures—p. Nonabelian mixed Hodge structures
Part II: The main definitions and conjectures
Pre-nonabelian mixed Hodge structures—p. Nonabelian mixed Hodge structures
Mixed Hodge complexes and linearization—p. Nonabelian mixed Hodge structures
Nonabelian mixed Hodge structures—p. Nonabelian mixed Hodge structures
Homotopy group sheaves—p. Nonabelian mixed Hodge structures
The basic construction—p. Nonabelian mixed Hodge structures
The basic conjectures—p. Nonabelian mixed Hodge structures
Variations of nonabelian mixed Hodge structure—p. Nonabelian mixed Hodge structures
Part III: Computations
Some morphisms between Eilenberg-MacLane pre-namhs—p. Nonabelian mixed Hodge structures
Construction of a namhs $`𝒱`$ of homotopy type $`S_𝐂^2`$—p. Nonabelian mixed Hodge structures
The namhs on cohomology of a smooth projective variety with coefficients in $`𝒱`$ —p. Nonabelian mixed Hodge structures
Introduction
The nonabelian aspect of homotopy theory has been present in mixed Hodge theory from very early on. This started with the work of Deligne-Griffiths-Morgan-Sullivan on rational homotopy theory, and then Morgan put a mixed Hodge structure on the differential graded algebra which calculates the homotopy groups of a simply connected complex variety (or the nilpotent completion of the fundamental group of a non-simply connected one). This construction was later improved by Hain who in particular obtained a much better functoriality.
Many applications of the Morgan-Hain construction have been given, for example in Torelli-type theorems , , the use of the strictness property , in the study of algebraic cycles , , for the Shafarevich conjecture for nilpotent fundamental groups , etc. There have recently been new attempts to better understand what is going on, see , , and to simplify the relatively abstract nature of Hain’s construction, see , .
In an important development, Hain obtained a mixed Hodge structure on the relative Malcev completions of the fundamental group at any representation underlying a variation of Hodge structure . This significantly extends the “amount” of the fundamental group that is seen by mixed Hodge theory.
The variation of the mixed Hodge structures on homotopy groups, when the underlying variety varies in a family, has been studied by Navarro-Aznar and Wojtkowiac .
There are two basic gaps in the classical mixed Hodge homotopy theory. The first is that one should really treat a “higher functoriality”: the set of homotopy $`n`$-types forms an $`n+1`$-category, so one would like to have an $`n+1`$-category of “homotopy $`n`$-types with mixed Hodge structure”. The fact that we will then have an $`n+1`$-functor from the $`1`$-category of varieties, to this $`n+1`$-category, contains subtle additional “homotopy coherence” data.
The second gap is that up until fairly recently, very little has been done on the problem of the higher homotopy theory of non-simply connected varieties. A first look at this problem may be considered to come from the work of Green-Lazarsfeld on cohomology jump loci, although they don’t talk about mixed Hodge structures. Arapura has looked at the Green-Lazarsfeld results using mixed Hodge structures . Kapranov and Toen also have methods which go in the direction of looking at the higher homotopy theory of non-simply connected varieties, but these haven’t been fully developed yet.
The third author of the present paper has proposed, as a remedy to the above problems, to look at “nonabelian cohomology”. One chooses a “coefficient object” which is an $`n`$-stack $`T`$, and looks at the nonabelian cohomology $`H=\underset{¯}{Hom}(X_B,T)`$. The coefficient object encodes data which is much the same as that of a complex differential graded algebra, however on the level of the fundamental group of $`T`$ we have $`\pi _1(T)=G`$ an affine group scheme. The cohomology stack $`H`$ now lies over the moduli $`1`$-stack $`(X_B,G)=\underset{¯}{Hom}(X_B,BG)`$ of representations of $`\pi _1(X^{\mathrm{top}})`$ in $`G`$, and $`H`$ encodes higher homotopy data relative to the universal family of representations of $`\pi _1(X^{\mathrm{top}})`$. In this way we get an approach towards the problem of higher homotopy in the presence of a fundamental group. The construction $`X\underset{¯}{Hom}(X_B,T)`$ is automatically $`n+1`$-functorial in $`X`$ (in $`T`$ too), so this addresses the issue of “higher functoriality”. Furthermore, this higher functoriality shows up in concrete ways, for example in the case of a family of varieties one obtains a secondary Kodaira-Spencer map which detects a “higher variation” of Hodge structure even when all of the classical variations of (mixed) Hodge structure associated to the family are constant.
One of the major problems in this nonabelian cohomology approach has been the lack of a notion of weight filtration, and consequently the lack of a notion of mixed Hodge structure. The present paper aims to start to close this gap. Briefly, what we do is to propose a definition of nonabelian mixed Hodge structure and in particular, the notion of “weight filtration” that goes into this. We state a number of conjectures about the notion that we define, but only go a short way toward justifying these conjectures by explicitly treating a particular example in Part III.
Our search for a notion of nonabelian mixed Hodge structure is motivated by the possibility of obtaining applications similar to what has already been done with mixed Hodge structures on homotopy groups. One of the main things would be to obtain restrictions on the homotopy types of smooth projective varieties (or even general quasiprojective varieties). This might include generalized “quadratic singularity” statements, for nonabelian cohomology stacks. Other possibilities might include Torelli-type theorems, strictness results, and restrictions on the monodromy actions (or generalized monodromy actions) for holomorphic families of smooth projective varieties. One might get a nonabelian version of the theory of realizations of motives . Finally, we think that the techniques we use here might also be of interest in generalizing the geometric Langlands correspondence to higher nonabelian cohomology.
Before saying more about the ideas that we present here, we would like to take note of a few other directions that one could possibly take. It should be possible to remedy the problem of “higher functoriality” by looking directly at Morgan’s or Hain’s construction and consistently keeping track of what is going on in a suitable $`n+1`$-categorical or $`\mathrm{}`$-categorical way. The work of Hinich would probably be relevant here.
Arapura’s treatment of the Green-Lazarsfeld results on jump loci makes explicit use of mixed Hodge structures on dgas and discusses higher homotopy in the sense that he treats the jump loci for higher cohomology. He also has some results for support loci for higher rank local systems . These results are a step in the direction indicated in the previous paragraph.
It might be possible to construct a “big” differential graded algebra which encloses mixed-Hodge theoretic information about higher homotopy, relative to all the representations of $`\pi _1`$. The first and second authors have been thinking about this for some time now and hope to treat it in a future paper.
There is a relationship between simply connected differential graded algebras and simply connected geometric $`n`$-stacks, and we have to some extent exploited this in our research on the questions treated in the present paper; however, our calculations in this direction remain largely heuristic. It would be good to understand rigorously the foundations of these calculations; this would help to bridge the gap with the earlier results of Morgan and Hain.
The big differential graded algebra approach might also be related to the notion of higher Malcev completion with its conjectural mixed Hodge structure mentionned at the end of . On the downside, these objects seem to be too “big” (see the example given at the beginning of ) and in a certain sense, one could think of the nonabelian cohomology approach as being a way of extracting finite dimensional information from these big objects.
Another approach which uses the notion of differential graded algebra but in a slightly different manner, is Kapranov’s “extended moduli space of representations” , part of the idea for which apparently comes from Kontsevitch’s folkloric “derived deformation theory”. Kapranov’s extended moduli space visibly contains higher homotopy data, because the tangent complex at a point includes all higher cohomology groups. We have not been able to identify any further the exact nature of this homotopy-theoretic information. It seems quite likely that there should be some type of “mixed Hodge structure” on Kapranov’s space, and that this would include a mixed Hodge structure on whatever higher homotopy information is contained therein.
For us, an interesting application of Kapranov’s idea would be if one could define a notion of “extended $`n`$-stack” in order to be able to have an “extended nonabelian cohomology stack” which would extend $`\underset{¯}{Hom}(X_B,T)`$ in the same way that Kapranov’s object extends $`\underset{¯}{Hom}(X_B,BG)`$. The reason why this would be helpful is that the extended object could be expected to be formally smooth (this hope seems to be due to Kontsevitch, Deligne, Drinfeld). This smoothness would serve to avoid a large number of the problems which we have with the annihilator ideals in the present paper. Thus, while representing a further abstraction, such a combination of ideas could end up eventually simplifying things.
Yet another direction is that taken recently by Toen . He proposes to do an $`\mathrm{}`$-categorical version of the idea of Tannaka duality, to view a homotopy type as the Tannaka dual of its $`\mathrm{}`$-category of local systems of perfect complexes with monoidal structure (tensor product). In this point of view, Toen expects to be able to define Hodge and weight filtrations and to obtain some type of mixed Hodge structure.
To close this part, we note that one of the major advances in mixed Hodge theory, Saito’s theory of mixed Hodge modules , treats mostly the abelian cohomology theory. It would be good to have a “nonabelian” version of Saito’s theory directed toward a notion of Hodge theory for nonabelian cohomology with local coefficients varying constructibly with respect to a stratification. This seems to be a long way off and that could be considered as a long-range goal of any of the above approaches.
Now we get to the ideas that are contained in the present paper. The domain stack $`X_M`$ which we shall use is an object enclosing $`X_B`$, $`X_{DR}`$ and $`X_{Dol}`$, see p. Nonabelian mixed Hodge structures for more details. In the abelian theory the idea of combining together several different types of cohomology theories goes back a long way, see Huber for a compendium. Several people including B. Toen and M. Rosellen had asked us whether there could be any type of “nonabelian motive” taking into account the de Rham, Dolbeault and Betti cohomologies. While there doesn’t seem to be anything deep going on here, the notation $`X_M`$ is a useful bookkeeping device.
Our first preliminary calculations showed that in order to have any hope of having some kind of “mixed Hodge structure” on a nonabelian cohomology stack $`\underset{¯}{Hom}(X_M,T)`$, the coefficient stack $`T`$ had itself to have a “mixed Hodge structure”. Thus our search for the notion of nonabelian mixed Hodge structure concerned both the coefficient stack and the answer stack. The idea is to obtain a collection of conditions for the definition, such that if $`T`$ satisfies these conditions then so does $`\underset{¯}{Hom}(X_M,T)`$. Actually the set of conditions which we propose here is just a first attempt; it could in the future turn out to be wrong either by having too many conditions, or by having not enough.
The first and most obvious element is that, in order to integrate the “weight filtration” into the notion of $`n`$-stack, we will use the Rees space approach to the notion of filtration: a filtered object is an object over $`𝐀^1`$ together with an action of $`𝐆_m`$ covering the standard action of $`𝐆_m`$ on $`𝐀^1`$. If the objects in question are vector spaces, then to a filtered vector space one obtains a corresponding equivariant bundle over $`𝐀^1`$ which is essentially its Rees module. The origins of the notion of Rees module are actually closely related to the case of “filtered schemes” (cf Fulton pp 90-91, who cites Gerstenhaber and Rees , treating filtered rings; these are all related to the “deformation to the normal cone” which we discuss on p. Nonabelian mixed Hodge structures below).
The idea of using the Rees correspondence for the Hodge filtration, and particularly for the Hodge filtration on nonabelian cohomology, is discussed in , . The idea goes back to Deligne’s approach to the twistor space , Deninger’s formal version of the Rees correspondence for the Hodge filtration , and the Neisendorfer-Taylor deformation between de Rham and Dolbeault cohomology . Perhaps the first place where the notion of Rees bundle appears in connection with the weight filtration, is in Sabbah . Recently (at Irvine, June 1998) Voevodsky made the intriguing remark that the tensor category of mixed motives had two fiber functors, one for the underlying de Rham cohomology and one for the associated-graded of the weight filtration. These clearly correspond to the two points of the stack $`𝒜`$ defined in the following paragraph, so this remark can also be seen as a precedent for using the Rees space approach to look at the weight filtration.
Thus we fix the notion of weight filtration as being an equivariant object over $`𝐀^1`$, or equivalently an object over the quotient stack
$$𝒜:=𝐀^1/𝐆_m.$$
It follows relatively logically that the notion of a pair of filtrations, the weight filtration and the Hodge filtration, is to be interpreted as being an object over the product
$$𝒜_{wt}\times 𝒜_{hod}$$
where the superscripts indicate separate copies of $`𝒜`$, the first one for the weight filtration and the second one for the Hodge filtration. We denote a geometric $`n`$-stack provided with two filtrations, by $`(V,W,F)`$; this object really means a geometric $`n`$-stack
$$Tot^{W,F}(V)𝒜_{wt}\times 𝒜_{hod}.$$
An important element in Hodge theory is the interplay between the Hodge filtration and its complex conjugate. This reappears in the notion of “twistor space” if one wishes to avoid the Hodge filtration altogether. It would certainly be possible to do what we are doing here in the “twistor” context, to obtain a notion of nonabelian mixed twistor structure. However, this would add on yet another layer of abstraction which we don’t feel is justified at the current time. Thus, for the present purposes, we introduce the notion of real structure, which allows us literally to take the complex conjugate of the Hodge filtration for the purposes of measuring the “purity” of the Hodge filtration.
The associated-graded of the weight filtration is an object $`Gr^W(V)`$ obtained by restricting to $`[0]𝒜_{wt}`$. The “point” represented by the origin is actually the quotient stack $`0/𝐆_m`$ with $`𝐆_m`$ acting trivially on the point. This quotient stack is the classifying stack denoted $`B𝐆_m`$. Thus the origin of $`𝒜`$ is in fact the classifying stack $`B𝐆_m`$ and the restriction $`Gr^W(V)`$ is an object over $`B𝐆_m`$, which really means an $`n`$-stack with action of $`𝐆_m`$. Another way to understand this is to think of the total stack as being a $`𝐆_m`$-equivariant stack over $`𝐀^1`$; since the origin is a fixed point, the fiber over the origin is a $`𝐆_m`$-equivariant stack.
The homotopy groups of the associated-graded $`Gr^W(V)`$ are vector spaces with $`𝐆_m`$ action, so they decompose into pieces which one could denote by $`\pi _iGr_k^W(V)`$. These are of course the analogues of the classical graded pieces of the weight filtration. They are provided with Hodge filtrations, and real structures, so it makes sense to look at the complex conjugate of the Hodge filtration and ask, for example, that the Hodge filtration be $`h`$-opposed with its complex conjugate on $`\pi _iGr_k^W(V)`$. Here $`h`$ is to be chosen according to Deligne’s arithmetic of which combines the degree of the weight filtration $`k`$ and the homotopical degree $`i`$: we ask that $`h=ki`$ (cf p. 1).
The above ideas are enough to try for a sketch of definition of the notion of nonabelian mixed Hodge structure. We define an object which we call a pre-namhs (see p. Nonabelian mixed Hodge structures) which basically consists of a geometric $`n`$-stack with a bifiltration $`(V_{DR},F,W)`$ (a bifiltration is determined by its total stack $`Tot^{W,F}(V)`$ over $`𝒜\times 𝒜`$ cf p. Nonabelian mixed Hodge structures), plus a real stack $`V_{B,𝐑}`$ with a filtration $`W`$ and an analytic equivalence between $`(V_{DR},W)`$ and the complexified $`(V_{B,𝐂},W)`$. The associated graded $`Gr^W(V_{DR})`$ is a $`𝐆_m`$-equivariant geometric $`n`$-stack, so (after linearization) it breaks up into components which we denote $`Gr_k^W(V_{DR})`$. These components have Hodge filtrations $`F`$, and analytic real structures coming from their analytic identifications with the complexification of $`Gr_k^W(V_{B,𝐑})`$. In particular, the homotopy group sheaves $`\pi _iGr_k^W(V_{DR})`$ are actually just complex vector spaces (for $`i2`$), and they have Hodge filtrations and real structures. Thus it makes sense to ask for a purity condition relating the Hodge filtration and its complex conjugate (this condition also involves $`k`$ and $`i`$, see below). In fact, we don’t put on any such condition in the notion of pre-namhs, because it appears useful to manipulate a notion of bare object not necessarily satisfying any conditions.
This definition of pre-namhs is modelled in an obvious way on Deligne’s definitions of mixed Hodge structure and mixed Hodge complex. More generally, if $`OBJ`$ is any $`n+1`$-stack (with real structure and analytic extension) then one can define in a similar way a notion of “pre-mixed Hodge object of type $`OBJ`$” (cf page 1) as being a collection consisting of a bifiltered object $`(V_{DR},W,F)`$ of $`OBJ`$, a real filtered object $`(V_{B,𝐑},W)`$, and an analytic equivalence $`\zeta `$. For $`OBJ=VECT`$ we recover the notion of pre-mixed Hodge structure, and for $`OBJ=PERF`$ we recover the notion of pre-mixed Hodge complex. The case of a pre-namhs is obtained from $`OBJ=nGEOM`$.
The above notion of pre-namhs has an obvious problem, namely that $`\pi _1`$ is only an algebraic group and $`\pi _0`$ is a sheaf of sets. Thus, it isn’t clear what type of purity condition to put here, even though the $`\pi _0`$ and $`\pi _1`$ are key parts of what we would like to look at here. Nonetheless, we made some calculations and looked at some examples (in fact, preliminary versions of the computations we do in Part III) and came to the conclusion that there was also a more subtle problem even with respect to the $`\pi _i`$ for $`i2`$, in the above way of proceeding. In particular, if one tries to define a nonabelian mixed Hodge structure as being a pre-namhs which satisfies some purity conditions on the Hodge filtrations on the homotopy group sheaves, then there seems to be missing the “strictness” which one has come to expect in the mixed Hodge situation.
Our next idea was to look at a specific example where we know that there should be a “nonabelian mixed Hodge structure”. The best case we could find was that of $`S^2`$. On the one hand, this seems to be an interesting coefficient stack for nonabelian cohomology (cf ); the “schematic sphere” had previously been studied in the stable range in . On the other hand, it has a motivic structure being the homotopy type of $`𝐏^1`$ and in fact this “motive” plays an important role for example in Morel’s and Voevodsky’s theory . In particular, it is clear that whatever the definition of “nonabelian mixed Hodge structure” is, the homotopy type of $`𝐏^1`$ should carry one of these.
In doing the calculations, the main thing to recall is the “shift” of the weight filtration by the homotopical degree already mentionned above. This shift first appeared in Deligne’s “Hodge III” . Basically there are three numbers involved. The Hodge degree which is the integer $`h`$ such that the Hodge filtration and its complex conjugate are $`h`$-opposed (in a more general “mixed twistor” context, $`h`$ would be the slope of a semistable bundle over the twistor line $`𝐏^1`$). The weight $`w`$ is the degree in the weight filtration, equal to $`k`$ on the graded piece $`Gr_k^W(V)`$. And the homotopical degree $`i`$ indicates that we are looking at $`\pi _i`$. In a mixed Hodge complex, the homotopical degree is minus the cohomological degree, i.e. one should consider $`\pi _i=H^i`$. The mixed Hodge condition $`\mathrm{𝐌𝐇𝐂}`$ is a linear relationship between these parameters:
$$h=wi.$$
See our discussion on page 1 for a justification of the sign in this formula.
Now we get back to our calculation for $`𝐏^1`$. We have (in complexified homotopy theory)
$$\pi _2(𝐏^1)=H_2(𝐏^1)=𝐂,$$
which is of Hodge type $`(1,1)`$. This gives $`h(\pi _2)=2`$. The homotopical degree is of course $`i(\pi _2)=2`$, and the formula $`\mathrm{𝐌𝐇𝐂}`$ determines $`w(\pi _2)=0`$. Note of course that the degree $`w`$ in the weight filtration is the degree in the actual nonabelian filtration we will put on, which is different from the level in the end-result mixed Hodge structure on $`\pi _2`$ exactly because of the formula $`\mathrm{𝐌𝐇𝐂}`$.
On the other hand,
$$\pi _3(𝐏^1)=\pi _2(𝐏^1)_𝐂\pi _2(𝐏^1),$$
the isomorphism in question being given by the Hopf map or Whitehead product. This has Hodge type $`(2,2)`$ so $`h(\pi _3)=4`$. Again of course $`i(\pi _3)=3`$ and our formula gives $`w(\pi _3)=1`$. Now, in terms of the actual unshifted weight filtration, the Whitehead product map
$$\pi _2\pi _2\pi _3$$
goes from a $`1`$-dimensional space of weight $`w=0`$ to a space of weight $`w=1`$. In particular, and here is the key observation:
—the Whitehead products vanish on the associated-graded of the actual (unshifted) weight filtration.
This observation was confirmed by heuristic calculations with differential graded algebras.
It led to the main idea: that the associated-graded for the weight filtration in a nonabelian mixed Hodge structure should be an abelian object.
This idea fits in nicely with the whole “quantization” industry (cf q-alg or math.QA), where the idea of “deformation quantization” is to deform an abelian or classical situation, to a nonabelian one. The weight filtration considered as a $`𝐆_m`$-equivariant family of objects parametrized by $`𝐀^1`$, may be considered as a deformation of the associated-graded $`Gr^W(V)`$ which is the fiber over $`0𝐀^1`$. It should be possible to explore more deeply the relationship between our notion of weight filtration and the notion of quantization, but we haven’t done that.
This parallel raises the interesting question if any of the recent considerable progress in understanding deformation quantization, as exemplified by Kontsevich for example, might be useful in understanding the weight filtration in higher homotopy.
The notion of linearization plays an important role in Toen’s higher Tannaka duality .
On a technical level, another motivation for our main idea can be found in the construction of the Whitehead product on page Nonabelian mixed Hodge structures. The shift by the homotopical degree described above conflicts with the fact that the Whitehead product goes from $`\pi _i\pi _j`$ to $`\pi _{i+j1}`$. Without a supplementary condition, this product would be going between shifted mixed Hodge structures with shifts differing by $`1`$. The condition that the Whitehead products vanish on the associated-graded of the weight filtration means that they descend the weight filtration by one step; this extra step is just what is needed to counteract the extra $`1`$ in the homotopical degree occuring in the Whitehead product (cf p. 12). With this condition, when the mixed Hodge structures on the $`\pi _i`$ are shifted back to being actual mixed Hodge structures, the Whitehead product becomes a morphism of mixed Hodge structures. This situation provides a strong motivation for asking that the associated-graded object $`Gr^W(𝒱)`$ be an abelian object.
To put our main idea into practice, we augment the definition of “pre-namhs” described above, by adding a notion of “linearization”. If $`𝒱`$ is a pre-namhs, then one obtains a pre-namhs denoted $`Gr^W(𝒱)`$, which is split in the sense that the weight filtration comes directly from an action of $`𝐆_m`$. A linearization of $`𝒱`$ is the data of a split pre-mixed Hodge complex $`𝒞`$ with an equivalence between $`Gr^W(𝒱)`$ and the Dold-Puppe pre-namhs of $`𝒞`$. (A technical point is that we require $`𝒞`$ to be split by a $`𝐆_m`$-action splitting the weight filtration, and we require the equivalence to respect splittings.) Without going into all of the definitions (for which we refer to pages Nonabelian mixed Hodge structures, Nonabelian mixed Hodge structures, Nonabelian mixed Hodge structures, Nonabelian mixed Hodge structures, Nonabelian mixed Hodge structures, Nonabelian mixed Hodge structures) we just say here that giving $`Gr^W(𝒱)`$ the structure of being the Dold-Puppe of a complex is basically the same thing as giving it an “infinite loop-space structure” which means a “linear structure” in an appropriate sense. One effect of this structure is that $`\pi _iGr_k^W(𝒱)`$ becomes a complex vector space (with Hodge filtration, real structure) for all $`i0`$. The other effect is that the Whitehead products for $`Gr^W(𝒱)`$ vanish, putting into effect our main idea. We denote the object $`𝒞`$ by $`LGr^W(𝒱)`$, and if such a structure is provided then we say that $`𝒱`$ is a linearized pre-namhs.
Now the purity condition on a linearized pre-namhs is easy to state: it is just the condition that the pre-mixed Hodge complex $`LGr^W(𝒱)`$ should be an actual mixed Hodge complex.
This condition is still not enough. It is insufficient on the level of $`\pi _0`$. To understand this part, it will be useful to suppose that the objects we are talking about are schemes rather than $`n`$-stacks. The weight filtration is a $`𝐆_m`$-equivariant deformation from $`V`$ to its associated-graded $`Gr^W(V)`$; on the other hand, the linearization amounts to an equivalence between $`Gr^W(V)`$ and a complex vector space. In particular, if the deformation is flat then $`V`$ is smooth (at least locally near where the deformation is taking place). At non-smooth points, we could deform $`V`$ to, say, its normal cone, but that would then have to be embedded in a vector space. Thus, in general we expect that the weight filtration is not flat over $`𝒜`$.
In the case where the weight filtration is flat, which means that we are looking locally near some smooth points of $`V`$, it seems likely that the “mixed Hodge complex” condition described above, is in fact sufficient to define a good notion of nonabelian mixed Hodge structure. The purity condition on the level of the associated-graded, extends outward in the deformation in the same way as it does for mixed Hodge complexes.
In the non-flat case, we look more carefully at the algebraic geometry of the situation and end up defining the notion of annihilator ideal. Since the present introduction is already too long, we won’t go into any further details here but refer the reader directly to the text. We just mention that we isolate three conditions to put on the annihilator ideals denoted $`\mathrm{𝐀𝟏}`$, $`\mathrm{𝐀𝟐}`$, and $`\mathrm{𝐀𝟑}`$. The first two concern only the weight filtration and the third is a compatibility with the split mixed Hodge structure on the associated-graded part. These conditions amount to limitations on the way the non-flat weight filtration can constitute a “jump” when going from $`Gr^W(V)`$ to $`V`$. Condition $`\mathrm{𝐀𝟑}`$ complements the purity condition on the associated-graded $`Gr^W`$ and requires that this condition persist after the jump from $`Gr^W(V)`$ down to $`V`$.
A note of caution about the conditions on the annihilator ideals is in order. We have come up with these conditions only fairly recently in our research on this problem. We feel that the three conditions represent the right collection of conditions, but this is more of an opinion than an established fact. We support this opinion by showing how the whole theory works in a specific computation in Part III, and also give some evidence showing how these conditions already arise in the context of mixed Hodge complexes (cf p. 11). However, it remains possible that in the future it might turn out to be better to modify these conditions in some way (in particular, in order to be able to obtain the various conjectures that we make).
After presenting the definition of nonabelian mixed Hodge structure, we turn to the basic construction where this notion appears, namely the construction of a nonabelian mixed Hodge structure $`=\underset{¯}{Hom}(X_M,𝒱)`$ when the coefficient object $`𝒱`$ is itself a nonabelian mixed Hodge structure. We construct $``$ as a linearized pre-namhs, when $`𝒱`$ is a linearized pre-namhs (modulo a technical hypothesis that the fundamental group object of $`𝒱`$ should be represented by a flat linear group scheme). Then we state as Conjecture 1, that if $`𝒱`$ is a nonabelian mixed Hodge structure then $``$ should be one too. We don’t prove this conjecture in general; however in Part III we construct a specific pre-namhs $`𝒱`$ whose underlying homotopy type is the complexified $`2`$-sphere, and we show that for this specific $`𝒱`$, the nonabelian cohomology stack $``$ is a nonabelian mixed Hodge structure. This example is fairly instructive in showing what is going on.
In the section presenting our conjectures, we also state several other desireable things that should be possible to prove given enough time. We refer the reader there for the statements.
Given: (1) the length necessary just to explain the definition of nonabelian mixed Hodge structure; (2) that any proof of these conjectures would probably take a lot more time to do, a lot more space to write, and be even more highly technical; and (3) the illustrative nature of the example which we can already give in Part III, we felt that it would be reasonable to leave these main statements as conjectures, and to post what we have done up until now.
So, to conclude this introduction, the current paper is basically just a presentation of the idea of the definition of nonabelian mixed Hodge structure, with a computation of a specific example plus some indications of how the future developpement of this idea might go. Unfortunately, even with this relatively limited goal, the paper is too long.
Acknowledgements: We would like to thank A. Beilinson, A. Hirschowitz, M. Kontsevich, C. Teleman, and B. Toen for helpful conversations on topics closely connected to what is being done here. The second and third authors would like to thank UC Irvine for their hospitality during the preliminary period of research on this question.
Part I: Nonabelian weight filtrations
Conventions
We work over the site $`𝒵`$ of schemes of finite type over $`𝐂`$ with the etale topology. We look at $`n`$-stacks for $`0n<\mathrm{}`$.
Throughout the paper, all geometric $`n`$-stacks (which are $`n`$-stacks of $`n`$-groupoids over the site $`𝒵`$)—see — will be assumed to be very presentable (see ); the notation “geometric $`n`$-stack” is taken to mean “very presentable geometric $`n`$-stack”. We also recall that the definition of “geometric” in includes the condition of being “of finite type” in an appropriate sense.
Denote by $`nGEOM`$ the $`n+1`$-stack of very presentable geometric $`n`$-stacks. This $`n+1`$-stack includes all not necessarily invertible morphisms, so it is a stack of $`n+1`$-categories but not of $`n+1`$-groupoids. One can think of it as being a stack of simplicial categories or “Segal categories” cf . For a scheme $`Z`$, the $`n+1`$-category $`nGEOM(Z)`$ is by definition the $`n+1`$-category of very presentable geometric $`n`$-stacks $`TZ`$.
This description also works when the base is a stack. In fact, throughout the paper we shall use the equivalence between two different ways of thinking about families of $`n`$-stacks: if $`S`$ is an $`n`$-stack of groupoids then the $`n+1`$-stack of $`n`$-stacks over $`S`$, i.e.
$$\underset{¯}{Hom}(I,nSTACK)\times _{nSTACK}\{S\}$$
(the stack of arrows in $`nSTACK`$ whose target is $`S`$), is equivalent to $`\underset{¯}{Hom}(S,nSTACK)`$. This is discussed in ; admittedly that discussion is incomplete but we don’t currently wish to add anything to it. Furthermore, we state without proof that if $`f:SnSTACK`$ corresponds to an $`n`$-stack $`TS`$, and if $`S`$ is geometric, then $`T`$ is geometric if and only if the classifying morphism $`f`$ lands in $`nGEOM`$. Thus if $`S`$ is a geometric $`n`$-stack then a morphism $`SnGEOM`$ is the same thing as a geometric $`n`$-stack $`T`$ over $`S`$, i.e. a morphism of geometric $`n`$-stacks $`TS`$. We shall use this equivalence tacitly throughout the paper.
In the case $`n=0`$, a geometric $`0`$-stack is just an Artin algebraic space. If $`n=1`$ then a geometric $`1`$-stack is an Artin algebraic $`1`$-stack. The “very presentable” condition corresponds to asking that the stabilizer groups for points of the $`1`$-stack be affine algebraic groups.
Nonabelian filtrations
Fix an $`n+1`$-stack $`OBJ`$ which we think of as parametrizing a “type of object”. The typical example is when $`OBJ=VECT`$ is the $`1`$-stack of vector bundles. This example leads back to the classical notion of filtered vector space. The example which we shall use is when $`OBJ=nGEOM`$, the $`n+1`$-stack of geometric $`n`$-stacks. Another useful example is $`OBJ=nPERF`$, the $`n+1`$-stack of perfect complexes of length $`n`$ (see below).
Recall that a filtered object of $`OBJ`$ is a $`𝐆_m`$-equivariant morphism from $`𝐀^1`$ to $`OBJ`$, in other words a morphism of $`n+1`$-stacks from (the $`1`$-stack quotient thought of as an $`n+1`$-stack) $`𝐀^1/𝐆_m`$, to $`OBJ`$.
For the rest of the paper we establish the notation
$$𝒜:=𝐀^1/𝐆_m$$
(the quotient $`1`$-stack), and we denote the substacks:
$$𝐆_m/𝐆_m\text{by}[1]𝒜,$$
and
$$\{0\}/𝐆_mB𝐆_m\text{by}[0]𝒜.$$
We obtain the $`n+1`$-stack of filtered objects in $`OBJ`$, denoted
$$F.OBJ:=\underset{¯}{Hom}(𝒜,OBJ).$$
Note here and throughout, that whenever we write a formula of this kind using $`Hom`$ or internal $`\underset{¯}{Hom}`$ we make the tacit assumption that the range object has been replaced with a fibrant replacement, cf .
The underlying object of a filtered object, is the preimage of $`[1]𝒜`$, call it $`V`$. It is a point of $`OBJ`$, i.e. a morphism $`V:OBJ`$ or equivalently a global section $`VOBJ(Spec(𝐂))`$. If $`V`$ is a global section of $`OBJ`$ then a “filtration of $`V`$” is the specification of a filtered object in $`OBJ`$ together with an equivalence between the underlying object and $`V`$; we denote such a specification by a letter such as $`W`$, and the morphism $`𝒜OBJ`$ corresponding to $`W`$ is then denoted $`Tot^W(V)`$. Introduce also the notation
$$T^W(V):=Tot^W(V)|_{𝐀^1};$$
in terms of $`n`$-stacks over a base this could also be written
$$T^W(V):=Tot^W(V)\times _𝒜𝐀^1.$$
We denote by $`Gr^W(V)`$ the associated-graded object, which is defined to be the restriction to $`[0]𝒜`$. This is an object with an action of $`𝐆_m`$, because it is really a morphism $`B𝐆_mOBJ`$. It is occasionally more convenient to think of $`Gr^W(V)`$ as being the object itself (i.e. the section $`OBJ`$ obtained by restriction to the basepoint $`B𝐆_m`$), provided with an action of $`𝐆_m`$. With all of these notations we denote a filtered object in the usual way by a pair of letters $`(V,W)`$.
Linearization
Suppose now that $`\mathrm{\Phi }:OBJ^{}OBJ`$ is a functor of $`n+1`$-stacks. A $`\mathrm{\Phi }`$-linearized (or just linearized) filtered object is a filtered object $`(V,W)`$ together with an object $`LGr^W(V):B𝐆_mOBJ^{}`$ and an equivalence
$$Gr^W(V)\mathrm{\Phi }(LGr^W(V)).$$
We denote by $`F^\mathrm{\Phi }.OBJ`$ the $`n+1`$-stack of $`\mathrm{\Phi }`$-linearized filtered objects. The terminology “linearization” is due to the fact that we will use this with $`OBJ^{}`$ being a stack of “linear objects” (perfect complexes).
Filtered vector spaces
Before getting to the nonabelian examples which are basic to this paper, we start by examining the case of filtered vector spaces. The references for this are , , , , . As well as motivating the nonabelian definitions, this discussion is necessary for fixing our conventions regarding the indexing of filtrations.
Apply the above general discussion to the $`1`$-stack $`VECT`$ of vector bundles: for a scheme $`X`$ we put $`VECT(X)`$ equal to the category of vector bundles (of finite rank) on $`X`$. A filtered object of $`VECT`$ is thus a morphism $`𝒜VECT`$, i.e. a $`𝐆_m`$-equivariant vector bundle over $`𝐀^1`$. The fiber over $`[1]`$ is a vector space $`V`$, and the total vector bundle is denoted $`T^W(V)=Tot^W(V)|_{𝐀^1}`$. An element $`vV`$ gives rise to a unique $`𝐆_m`$-invariant section $`\stackrel{~}{v}=𝐆_mv`$ over $`𝐆_m𝐀^1`$. We say that $`vW_kV`$ if the section $`t^k\stackrel{~}{v}`$ extends to a section of $`T^W(V)`$ over $`𝐀^1`$; i.e. if $`𝐆_mv`$ vanishes to order at least $`k`$ as a section of $`T^W(V)`$. Here $`t`$ is the coordinate vanishing at the origin of $`𝐀^1`$. We obtain sub-vector spaces $`W_kVV`$ which form an increasing filtration of $`V`$:
$$\mathrm{}W_kVW_{k+1}V\mathrm{}.$$
The fact that $`T^W(V)`$ is a vector bundle over $`𝐀^1`$ means that this filtration is exhaustive (and since $`V`$ has finite rank, it has only a finite number of nonzero steps). Furthermore, the fiber of $`T^W(V)`$ over $`0𝐀^1`$ is naturally identified with the associated-graded of the actual filtration defined above:
$$T^W(V)|_0Gr_k^W(V)=W_kV/W_{k1}V.$$
The identification between $`W_kV/W_{k1}V`$ and a subspace of $`T^W(V)|_0`$ is provided by the map
$$vt^k\stackrel{~}{v}(0).$$
Now note that $`T^W(V)`$ is a $`𝐆_m`$-equivariant bundle, in particular the fiber over the origin has an action of $`𝐆_m`$. To fix conventions for this action (which has an impact on the sign of the indexing integers below), we say that the action is considered as an action on the total space covering the action $`\lambda t=\lambda t`$ on $`𝐀^1`$. If $`u`$ is a section of a $`𝐆_m`$-equivariant bundle then $`tu`$ is another section and we have $`\lambda (tu)=\lambda ^1t(\lambda u)`$ (the sign can be seen by looking at the trivial action on the trivial bundle with the unit section $`u`$). If $`\lambda 𝐆_m`$ then applying the above to the invariant section $`\stackrel{~}{v}`$ we get
$$\lambda (t^k\stackrel{~}{v})=\lambda ^kt^k\stackrel{~}{v}.$$
In particular, the image of $`W_kV/W_{k1}V`$ is the subspace of $`T^W(V)|_0`$ on which $`\lambda 𝐆_m`$ acts by the character $`\lambda ^k`$. Note that this will motivate our choice of sign in the notations of Theorem 7 and Corollary 8 below.
The above construction gives an equivalence between $`F.VECT(Spec(𝐂))`$ and the category of filtered vector spaces. The functor going in the other direction is the “Rees module” construction, see , , , , .
Filtered vector sheaves
The reader is urged to skip this subsection. Let $`VSCH`$ denote the $`1`$-stack of vector schemes, and $`VSH`$ the $`1`$-stack of vector sheaves. Recall that a vector scheme over a base $`Y`$ is a group scheme which locally looks like the kernel of a map of vector bundles, and a vector sheaf (see ) is a sheaf of groups on the big site (Zariski or etale) which locally looks like the kernel of a map of coherent sheaves. We have inclusions
$$VECTVESCHVESH.$$
We consequently obtain notions of “filtered vector schemes” and “filtered vector sheaves”. These notions are different but the distinction is somewhat subtle, because the underlying objects are the same: a vector sheaf or vector scheme over $`Spec(𝐂)`$ is always just a finite dimensional vector space. However, a vector scheme on $`𝐀^1`$ is not necessarily a vector bundle, and in turn a vector sheaf is not necessarily a vector scheme. Thus for a vector space $`V`$ we have three distinct notions of filtration on $`V`$, namely a vector bundle filtration; a vector scheme filtration; and a vector sheaf filtration.
In all three cases, the definition of the actual filtration of $`V`$ by subspaces, using the extension property of sections (i.e. that $`vW_kVt^k\stackrel{~}{v}`$ extends to a section of $`T^W(V)`$), works equally well. Thus, if $`(V,W)`$ is a filtered vector sheaf then it induces a filtration of $`V`$ by sub-vector spaces $`W_kV`$. This construction is not an equivalence of categories, however, and so should be used with care.
We will further discuss these notions when they are necessary; usually they are not. Unless otherwise indicated, a “filtered vector space” will always mean an object of $`F.VECT`$, i.e. a vector space provided with a filtration in the stack of vector bundles.
Filtered geometric $`n`$-stacks
The basic example which we use all the time is when $`OBJ:=nGEOM`$ is the $`n+1`$-stack of geometric $`n`$-stacks of groupoids (for finite $`n`$). In this case, a filtered object $`(V,W)`$ consists of the specification of a total geometric $`n`$-stack $`Tot^W(V)𝒜`$ together with an equivalence between the fiber over $`[1]`$ and $`V`$; the associated-graded $`n`$-stack $`Gr^W(V)`$ is the fiber over $`0`$ with its $`𝐆_m`$-action.
We will study this situation more closely further on below.
Split filtrations
Suppose $`UB𝐆_m`$ is a $`𝐆_m`$-equivariant geometric $`n`$-stack (let $`V`$ be the fiber $`V=U\times _{B𝐆_m}`$; then to be precise $`V`$ has an action of $`𝐆_m`$ and $`U`$ is the quotient). Then we can define the filtration of $`V`$ associated to this action in the following way. Note that we have a morphism $`𝒜B𝐆_m`$ whose fiber is $`𝐀^1`$. Set
$$Tot^W(V):=U\times _{B𝐆_m}𝒜.$$
The fiber of this over $`[1]`$ is just the fiber of $`U`$ over $``$, namely it is $`V`$; and the associated-graded or fiber over the origin, is $`Gr^W(V)=UB𝐆_m`$.
A filtered object obtained from this construction is called a split filtered object. However, when we speak of the $`n+1`$-category of split filtered objects, we usually mean that the splitting is included in the structure, in other words this is just the $`n+1`$-category of $`𝐆_m`$-equivariant objects, and one is conscious that it has a functor towards the category of filtered objects.
If we apply this discussion to the case of filtered vector spaces, we get the standard construction which takes a vector space with $`𝐆_m`$-action to the associated filtered vector space, turning the grading into a filtration.
Bifiltered objects
The next example shows why we wrote up the above definition of filtered object in a totally general way: we can apply it to $`OBJ:=F.nGEOM`$, the $`n+1`$-stack of filtered geometric $`n`$-stacks. We obtain an $`n+1`$-stack $`F.F.nGEOM`$ of “geometric $`n`$-stacks having two filtrations”. This phrase is in quotation marks because, as we shall see, the datum of an object in $`F.F.nGEOM`$ is different from the data of an object $`V`$ provided with two different filtrations $`W_1`$ and $`W_2`$; the “compatibility between the filtrations” amounts to extra structure.
In fact, let’s back up and analyse what it means to have an object in $`F.F.OBJ`$. We have
$`F.F.OBJ`$ $`:=`$ $`\underset{¯}{Hom}(𝒜,F.OBJ)`$
$`=`$ $`\underset{¯}{Hom}(𝒜,\underset{¯}{Hom}(𝒜,OBJ))`$
$`=`$ $`\underset{¯}{Hom}(𝒜\times 𝒜,OBJ).`$
Put another way, an object of $`F.F.OBJ`$ is a $`𝐆_m\times 𝐆_m`$-equivariant morphism $`𝐀^2OBJ`$. In passing note that there is an action of the symmetric group $`S_2`$ on $`𝒜\times 𝒜`$, and hence on $`F.F.OBJ`$; the transposition acts as “interchanging the two filtrations”. More generally there is an action of $`S_m`$ on $`F.\mathrm{}F.OBJ`$.
Now getting back to the case where $`OBJ=nGEOM`$, an object of $`F.F.nGEOM`$ is the same thing as a geometric $`n`$-stack $`T𝒜\times 𝒜`$. As before we shall adopt the notation $`(V,W,F)`$ for an object of $`F.F.nGEOM`$, where $`V`$ is the underlying object and $`W`$ and $`F`$ denote the two filtrations. The total stack which “is” the pair of filtrations, shall now be denoted by
$$Tot^{W,F}(V)𝒜\times 𝒜.$$
Its fiber over $`([1],[1])`$ is identified with $`V`$. Note that with our notations we have
$$Tot^{W,F}(V)=Tot^F(Tot^W(V)).$$
We denote the fiber of $`Tot^{W,F}(V)`$ over $`0`$ in the first factor, by $`(Gr^W(V),F)`$ and the fiber over $`0`$ in the second factor, by $`(Gr^F(V),W)`$; these are both filtered objects with $`𝐆_m`$-action, i.e. morphisms $`B𝐆_mF.nGEOM`$. The fiber over $`(0,0)`$ is denoted $`Gr^{W,F}(V)`$, and we have
$$Gr^{W,F}(V)Gr^F(Gr^W(V))Gr^W(Gr^F(V)).$$
It is an object with $`𝐆_m\times 𝐆_m`$-action, i.e. a morphism $`B𝐆_m\times B𝐆_mnGEOM`$. We shall sometimes call an object $`(V,W,F)`$ a bifiltered geometric $`n`$-stack.
We can see the difference between the above notion of bifiltered object, and the notion of the same object having two different filtrations. If $`(V,G)`$ and $`(V,H)`$ are two filtered objects having the same underlying object, then we have
$$Tot^G(V)𝒜,$$
and
$$Tot^H(V)𝒜.$$
These agree over $`[1]𝒜`$ so we can glue them together to get a geometric $`n`$-stack
$$T𝒜^{[1]}𝒜.$$
Note that
$$𝒜^{[1]}𝒜=(𝐀^2\{0\})/𝐆_m\times 𝐆_m.$$
Thus $`T`$ is a $`𝐆_m\times 𝐆_m`$-equivariant object over $`(𝐀^2\{0\})`$. On the other hand we have seen that a bifiltered object means a $`𝐆_m\times 𝐆_m`$-equivariant object over $`𝐀^2`$. Thus, the extra information needed to go from an object with two separate filtrations, to a bifiltered object, is an extension of the $`n`$-stack $`T(𝐀^2\{0\})`$ to an $`n`$-stack $`T^{}`$ over $`𝐀^2`$.
In the case of bifiltered vector spaces, which is the case $`OBJ=VECT`$, the extension in question is automatic; this corresponds to the “lemme de Zassenhaus” of (Deligne , 1.2.1), which says that the double associated-graded taken in one order is equal to the double associated-graded taken in the other order. In our language, an object in $`OBJ=VECT`$ with two different filtrations is a $`𝐆_m\times 𝐆_m`$-equivariant vector bundle over $`(𝐀^2\{0\})`$, whereas an object with two filtrations which are compatible in the sense that they satisfy the “lemme de Zassenhaus” is a $`𝐆_m\times 𝐆_m`$-equivariant vector bundle over $`𝐀^2`$. Recall that every algebraic vector bundle over $`(𝐀^2\{0\})`$ extends uniquely to a vector bundle over $`𝐀^2`$ (this is proved by extending the algebraic vector bundle to a reflexive sheaf by taking the direct image from the open set to all of $`𝐀^2`$, and then using the theorem that reflexive sheaves on smooth surfaces are bundles). The common double associated-graded is just the fiber over $`(0,0)`$ of this extension. The fact that one doesn’t have a “lemme de Zassenhaus” for three filtrations translates the fact that the extension property for vector bundles doesn’t hold for $`(𝐀^3\{0\})`$.
It would be interesting to be able to say something about the problem of existence and/or uniqueness extensions of families of other types of objects (schemes, geometric $`n`$-stacks, perfect complexes) from $`(𝐀^2\{0\})`$ to $`𝐀^2`$. While the “lemme de Zassenhaus” probably fails in all of these contexts, it would be good to understand just to what extent it fails.
Bifiltrations and linearization
We finish with a word about bifiltrations where one of the filtrations is linearized. In practice below we will work with “weight filtrations” which are filtered objects $`(V,W)`$ provided with linearization via a certain functor $`\mathrm{\Phi }`$ as envisaged above. We will then want to look at filtered objects in $`F^\mathrm{\Phi }.nGEOM`$ (the second filtration being the “Hodge filtration”).
We will still denote objects of $`F.F^\mathrm{\Phi }.nGEOM`$ as triples $`(V,W,F)`$, where $`(V,W)`$ is an linearized filtered object and $`F`$ is a filtration of $`(V,W)`$ considered as an object of $`F^\mathrm{\Phi }.nGEOM`$. Concretely (and denoting for the time being the linearization functor by $`\mathrm{\Phi }:OBJ^{}nGEOM`$) this means that the graded object $`(Gr^W(V),F)`$ is linearized along $`\mathrm{\Phi }`$, i.e. it is the image of a filtered object $`(LGr^W(V),F)`$ of $`OBJ^{}`$ (and also it has the $`𝐆_m`$-action because of being the associated-graded for $`W`$) or in other words,
$$(LGr^W(V),F):B𝐆_mF.OBJ^{}(Spec(𝐂)),$$
with given an equivalence
$$\mathrm{\Phi }(LGr^W(V),F)(Gr^W(V),F)$$
of $`𝐆_m`$-equivariant objects in $`F.nGEOM`$.
(Hemi-)perfect complexes and Dold-Puppe linearization
In this section we look at the specific linearization functor $`\mathrm{\Phi }`$ which will interest us. It is the forgetful functor from geometric $`n`$-stacks of spectra, to geometric $`n`$-stacks. Equivalently, it is the Dold-Puppe functor from perfect $`n`$-complexes, to geometric $`n`$-stacks. Let $`nPERF`$ denote the $`n+1`$-stack of perfect complexes of length $`n`$ (supported in the interval $`[n,0]`$). Recall from that this may be defined in the following way. For any scheme $`Z𝒵`$ we put $`nPerf(Z)`$ equal to the category of complexes of sheaves <sup>4</sup><sup>4</sup>4By it doesn’t matter whether we work with complexes of sheaves of $`𝒪`$-modules, or complexes of sheaves of abelian groups, since we are working on the big site. on $`Z`$ which are locally quasiisomorphic to complexes of vector bundles of length $`n`$ (supported in $`[n,0]`$, say). Then set $`nPERF(Z)`$ equal to the simplicial (or Segal) category obtained by localizing $`nPerf(Z)`$ by dividing out by quasiisomorphisms, in the sense of Dwyer-Kan . This simplicial category is in fact $`n+1`$-truncated, i.e. it may be considered as an $`n+1`$-category. It is proved in that the resulting $`n+1`$-prestack is an $`n+1`$-stack.
The functor
$$DP:nPERFnGEOM$$
is just the Dold-Puppe functor .
Another way of defining the $`n+1`$-stack $`nPERF`$ is as follows. Fix $`N>n`$. Then the $`n+N+1`$-substack
$$(n+N)GEOM^{ptd,Nconn}(n+N)GEOM$$
consisting of pointed, $`N`$-connected geometric (and by our convention, very presentable) $`n+N`$-stacks, is in fact $`n+1`$-truncated. (Note that an object of this substack, over a scheme $`Z`$, is an $`n+N`$-stack $`TZ`$ provided with a section $`t:ZT`$ and such that $`T`$ is relatively $`N`$-connected over $`Z`$.) This substack is equivalent to $`nPERF`$. The equivalence in one direction is obtained by taking a perfect complex supported in $`[n,0]`$ and shifting it to $`[nN,N]`$ then applying Dold-Puppe. In the other direction, take the singular cochain complex for example. The fact that these two constructions are inverses is the fact that rational homotopy theory in the stable range is equivalent to the theory of complexes of rational vector spaces \[43, Ch.8§1, Theorem 7\]. (We leave it to the reader to fill in the details of the proof of this equivalence).
In this second point of view, the functor
$$DP:nPERF=(n+N)GEOM^{ptd,Nconn}nGEOM$$
is the functor which to an object $`(T,t)`$ associates the $`N`$-th iterated loop stack of $`T`$ at the basepoint $`t`$.
Example: A perfect complex over $`Spec(𝐂)`$ is just a complex of $`𝐂`$-vector spaces. In this case, it is non-canonically quasiisomorphic to a complex with zero differential (the direct sum of its cohomology spaces) so its Dold-Puppe decomposes as a product of things of the form $`K(𝐂^{a_i},i)`$.
Hemiperfect complexes
Somewhat unfortunately, we will sometimes have to deal with a notion slightly more general than that of “perfect complex” which we call “hemiperfect complex”. This notion appears, for complexes of length $`1`$, in for example. The origin of the problem is that if $`K^{}`$ is a perfect complex (over a scheme $`Z`$, say) supported in the interval $`[n,\mathrm{})`$ then its truncation $`C^{}=\tau _0(K^{})`$ is no longer a perfect complex. Rather, the $`C^i=K^i`$ are vector bundles for $`i<0`$ but
$$C^0=\mathrm{ker}(K^0\stackrel{d}{}K^1)$$
is only a vector scheme.
Thus we introduce the following definition: a hemiperfect complex of length $`n`$ over a scheme $`Z`$ is a complex of sheaves of groups on the big etale site over $`Z`$, which is locally quasiisomorphic to a complex of the form $`C^{}`$ where $`C^i`$ is zero outside the interval $`[n,0]`$, where the $`C^i`$ are vector bundles for $`ni1`$, and where $`C^0`$ is a vector scheme. Let $`nHPerf(Z)`$ denote the $`1`$-category of such objects, and let $`nHPERF(Z)`$ denote the Dwyer-Kan localization dividing out by quasiisomorphisms. We state without proof that:
—the Segal category $`nHPERF(Z)`$ is $`n+1`$-truncated, so it may be considered as an $`n+1`$-category;
—if $`K`$ and $`L`$ are hemiperfect complexes, then the morphism $`n`$-groupoid between $`K`$ and $`L`$ in $`nHPERF(Z)`$ is calculated by the formula
$$nHPERF(Z)_{1/}(K,L)=DP(\tau _0(𝐑Hom(K,L)))$$
where the $`𝐑Hom`$ is taken in the derived category of complexes of sheaves of abelian groups or sheaves of $`𝒪`$-modules on the big Zariski site (we state that the two answers are the same); and
—the $`n+1`$-prestack $`ZnHPERF(Z)`$ is actually an $`n+1`$-stack on the site $`𝒵`$.
These statements are the analogues of statements proved for perfect complexes in , see also for the fact that the $`𝐑Hom`$ may be taken in complexes of sheaves of abelian groups or sheaves of $`𝒪`$-modules. We assume that the same proofs work for hemiperfect complexes.
If $`B`$ is an $`n`$-stack then a hemiperfect complex over $`Z`$ is defined as being a morphism $`BnHPERF`$.
It is sometimes useful to have a second point of view on perfect complexes, for hemiperfect complexes. This is somewhat problematic, because a simple transcription of that discussion cannot work. Indeed, an $`n+N`$-stack which is geometric and also $`N`$-connected relative to the base scheme $`Z`$ is automatically smooth and corresponds to a perfect complex. In particular, it cannot come from a hemiperfect complex which is not perfect. Thus we have to use the notion of very presentable $`n`$-stack.
A hemiperfect complex $`𝒞`$ may be viewed as a relatively pointed very presentable $`n+N`$-stack $`(\stackrel{~}{𝒞},p)`$ over $`Z`$ (i.e. $`p`$ is a section of $`\stackrel{~}{𝒞}/Z`$) such that $`\stackrel{~}{𝒞}`$ is relatively $`N`$-connected, and such that the $`N`$-th looping $`\mathrm{\Omega }^N(\stackrel{~}{𝒞},p)`$ is geometric. If furthermore $`\mathrm{\Omega }^N(\stackrel{~}{𝒞},p)`$ is smooth relative to $`Z`$, then $`\stackrel{~}{𝒞}`$ itself will be geometric and this will correspond to a perfect complex.
Our statement of the equivalence between the notion of truncated perfect complex and the notion of $`(\stackrel{~}{𝒞},p)`$ as above, is given with only a sketch of proof: starting with $`𝒞`$ we set $`\stackrel{~}{𝒞}`$ to be the Dold-Puppe of the shift of $`𝒞`$ by $`N`$ places to the left. On the other hand given $`\stackrel{~}{𝒞}`$ we construct $`𝒞`$ as the relative homology complex of $`\stackrel{~}{𝒞}`$ over $`Z`$, which is then shifted by $`N`$ places to the right. The proof that if $`𝒞`$ is hemiperfect then $`\mathrm{\Omega }^N(\stackrel{~}{𝒞},p)`$ is geometric, is the same as the proof for perfect complexes given in . The proof in the other direction, that if $`\mathrm{\Omega }^N(\stackrel{~}{𝒞},p)`$ is geometric then $`𝒞`$ is hemiperfect, is obtained by looking at $`𝒞`$ as the relative tangent complex of $`\mathrm{\Omega }^N(\stackrel{~}{𝒞},p)`$ over $`Z`$.
We still obtain a Dold-Puppe functor which we again denote by
$$DP:nHPERFnGEOM.$$
It sends $`𝒞`$ to the geometric $`n`$-stack $`\mathrm{\Omega }^N(\stackrel{~}{𝒞},p)`$, or equivalently it may be viewed directly as the Dold-Puppe of $`𝒞`$ see .
Cohomology and homotopy sheaves
If $`C^{}`$ is a (hemi)perfect complex of length $`n`$ on a scheme $`Z`$ then its $`i`$-th cohomology is a vector sheaf denoted $`H^i(C^{})`$ over $`Z`$, nonzero for $`0in`$ (see ). We have the following compatibility with the relative homotopy sheaves of the Dold-Puppe:
$$\pi _i(DP(C^{})/Z)=H^i(C^{}).$$
Thus, the $`H^i(C^{})`$ plays the role of the homotopy sheaf $`\pi _i`$ of the perfect complex.
In terms of the second point of view given above, if we think of $`C^{}`$ as corresponding to a relatively $`N`$-connected very presentable $`n+N`$-stack $`\stackrel{~}{C}`$ over $`Z`$, then $`H^i(C^{})`$ may be defined as being equal to the relative homotopy group sheaf $`\pi _{N+i}(\stackrel{~}{C}/Z)`$. In these terms, the above compatibility becomes tautological.
Dold-Puppe linearization
We now say that a Dold-Puppe-linearized filtered geometric $`n`$-stack is an object
$$(V,W)F^{DP}.nGEOM,$$
i.e. it is a filtered geometric $`n`$-stack with extra structure on the associated-graded via the functor
$$DP:nHPERFnGEOM.$$
Concretely, this means that $`(V,W)`$ is a quintuple
$$(V,W)=(V,ϵ,Tot^W(V)𝒜,\eta ,LGr^W(V))$$
where $`V`$ is a geometric $`n`$-stack; $`Tot^W(V)`$ is a geometric $`n`$-stack over $`𝒜`$; $`ϵ`$ is an equivalence
$$ϵ:VTot^W(V)\times _𝒜[1];$$
where
$$LGr^W(V):B𝐆_mnHPERF$$
is a hemiperfect complex with $`𝐆_m`$-action; and where $`\eta `$ is an equivalence between $`DP`$ of this latter, and the fiber of $`Tot^W(V)`$ over $`0`$:
$$\eta :DP(LGr^W(V))Tot^W(V)\times _𝒜B𝐆_m.$$
We denote by $`LF.nGEOM:=F^{DP}.nGEOM`$, the $`n+1`$-stack of Dold-Puppe-linearized filtered geometric $`n`$-stacks.
If the object $`LGr^W(V)`$ is a perfect complex then we say that the linearization is a perfect linearization. In the general case we may sometimes call it a hemiperfect linearization in order to avoid confusion, but if no specification is made then the linearization is a priori only hemiperfect.
The compatibility between cohomology groups of a complex and homotopy groups of the Dold-Puppe, gives in the case of a linearized filtration the formulae
$$\pi _iGr^W(V)=H^i(LGr^W(V)).$$
In particular, even $`\pi _0Gr^W(V)`$ and $`\pi _1Gr^W(V)`$ are vector sheaves (i.e. vector spaces when the base is a point).
Further study of filtered $`n`$-stacks
If $`(V,W)`$ is a filtered geometric $`n`$-stack then
$$T^W(V):=Tot^W(V)\times _𝒜𝐀^1$$
is a geometric $`n`$-stack mapping to $`𝐀^1`$. This situation leads to a certain amount of algebraic geometry which is basically the same as the geometry of a scheme mapping to $`𝐀^1`$. In studying this, it is important to recall that there is a scheme of finite type $`Z`$ with a surjective smooth map
$$ZT^W(V).$$
Coherent sheaves on geometric $`n`$-stacks
Before getting to our geometric study of $`T^W(V)`$ or $`Tot^W(V)`$, we digress for a moment to define the abelian category $`Coh(Tot^W(V))`$ of coherent sheaves. Let $`COH`$ be the $`1`$-stack (in additive categories) which to each scheme $`Y`$ assigns the additive category of coherent sheaves on $`Y`$. If $`X`$ is a geometric $`n`$-stack, put $`Coh(X):=\underset{¯}{Hom}(X,COH)`$.
###### Lemma 1
If $`X`$ is a geometric $`n`$-stack, then $`Coh(X)`$ is an abelian category containing a canonical object $`𝒪`$. If $`f:ZX`$ is a flat morphism of geometric $`n`$-stacks, then the induced morphism $`Coh(X)Coh(Z)`$ is exact. If $`f`$ is flat and surjective then this morphism is an exact inclusion of abelian categories.
Proof: The proof is by induction on $`n`$. Note that if $`X`$ is a scheme then $`Coh(X)=COH(X)`$ is the usual abelian category of coherent sheaves, and smooth morphisms between schemes induce exact pullback functors. The case of schemes takes care of the case $`n=1`$ in a certain sense in the induction, so we don’t bother starting with $`n=0`$. Suppose the statement is known for $`n1`$. Suppose $`X`$ is a geometric $`n`$-stack and choose a smooth surjection from a scheme $`ZX`$. By induction, we know that
$$Coh(Z),Coh(Z\times _XZ)\text{and}Coh(Z\times _XZ\times _XZ)$$
are abelian categories, and that the pullback functors $`p_i`$ and $`p_{ij}`$ between them are exact embeddings of abelian categories. Now an object $``$ of $`Coh(X)`$ is the same thing as a pair $`(𝒢,g)`$ where $`𝒢Coh(Z)`$ and $`g:p_1^{}(𝒢)p_2^{}(𝒢)`$ is an isomorphism such that the composition of $`p_{12}^{}(g)`$ and $`p_{23}^{}(g)`$ is equal to $`p_{13}^{}(g)`$. From this description and the fact that the functors $`p_i`$ and $`p_{ij}`$ involved are exact, it is clear that $`Coh(X)`$ admits kernels, cokernels, and that the coimages are equal to the images; i.e. $`Coh(X)`$ is abelian. Furthermore it is clear that the pullback functor $`Coh(X)Coh(Z)`$ is a faithful and exact.
The pair $`(𝒪,1)`$ corresponds to the object $`𝒪`$ of $`Coh(X)`$.
Suppose now that $`f:XX^{}`$ is a flat morphism (resp. flat surjection). Let $`Z^{}X^{}`$ be a smooth surjection from a scheme. Then $`X\times _X^{}Z^{}X`$ is a smooth surjection. Let $`ZX\times _X^{}Z^{}`$ be a smooth surjection from a scheme, so $`ZX`$ is a smooth surjection from a scheme. Note that $`ZZ^{}`$ is a flat morphism (resp. flat surjection) of schemes. In the diagram
$$\begin{array}{ccc}Coh(X^{})& & Coh(X)\\ & & \\ Coh(Z^{})& & Coh(Z)\end{array}$$
the vertical arrows are exact and faithful, and the bottom arrow is exact (resp. exact and faithful). Therefore the top arrow is exact (resp. exact and faithful). $`|||`$
This lemma implies, among other things, that the objects of $`Coh(X)`$ are noetherian.
For any $`n`$-stack $`X`$ one can define an additive category
$$Coh(X):=\underset{¯}{Hom}(X,COH).$$
It doesn’t look likely that $`Coh(X)`$ will be abelian for an arbitrary $`n`$-stack $`X`$, but we don’t have a counterexample. Note however that $`Coh(X)`$ depends only on the $`1`$-truncation $`\tau _1(X)`$ so any $`n`$-stack whose $`1`$-truncation is equivalent to the $`1`$-truncation of a geometric $`n`$-stack, will yield an abelian category $`Coh(X)`$.
Geometric study of $`Tot^W(V)`$
We can now continue with our study of $`T^W(V)`$ and $`Tot^W(V)`$. To start with, we say that $`(V,W)`$ is flat if $`T^W(V)𝐀^1`$ is a flat morphism. This means that the morphism of schemes $`Z𝐀^1`$ is flat.
We obtain from the previous lemma an abelian category $`Coh(T^W(V))`$ of coherent sheaves on the geometric $`n`$-stack $`T^W(V)`$. There is a canonical object, the structure sheaf $`𝒪Coh(T^W(V))`$. On the other hand, let $`t`$ denote the coordinate on $`𝐀^1`$. This pulls back to a section of the structure sheaf, which we can view as a morphism in the category $`Coh(T^W(V))`$, eventually raised to a power:
$$t^m:𝒪𝒪.$$
Let $`\mathrm{ker}(t^m)`$ denote the subobject of $`𝒪`$ which is the kernel of $`t^m`$. It is a coherent sheaf of ideals (i.e. a subobject of the structure sheaf) on $`T^W(V)`$. We have
$$\mathrm{ker}(t^m)\mathrm{ker}(t^{m+1}).$$
Since $`𝒪`$ is noetherian (look over any scheme in $`𝒵`$ with a smooth surjection to $`T^W(V)`$), this increasing sequence of ideals eventually becomes stationary.
We now note that these ideals are “$`𝐆_m`$-equivariant” i.e. they actually come from ideals (for which we use the same notation) $`\mathrm{ker}(t^m)`$ on $`Tot^W(V)=T^W(V)/𝐆_m`$. To see this, let $`𝒪\{1\}`$ be the structure sheaf $`𝒪`$ but with action of $`𝐆_m`$ twisted by the standard character of $`𝐆_m`$; it is an object in $`Coh(Tot^W(V))`$. The coordinate $`t`$ is a morphism
$$t:𝒪𝒪\{1\}$$
and its powers are morphisms
$$t^m:𝒪𝒪\{m\}.$$
Thus $`\mathrm{ker}(t^m)`$ has a meaning on $`Tot^W(V)`$.
These ideals are related to the canonical flat sub-filtration of a filtered object, defined in the following lemma:
###### Lemma 2
Suppose $`(V,W)`$ is a filtered geometric $`n`$-stack. Then the $`n+1`$-category of objects $`(V,W^{})(V,W)`$ (filtrations on $`V`$ mapping to $`W`$) such that $`(V,W^{})`$ is flat, has a final object $`(V,W^{\mathrm{fl}})`$, which we call the canonical flat subfiltration. Furthermore, $`Tot^{W^{\mathrm{fl}}}(V)`$ is the closed geometric substack of $`Tot^W(V)`$ defined by the coherent sheaf of ideals $`\mathrm{ker}(t^m)`$ (for $`m0`$).
Proof: Define $`(V,W^{\mathrm{fl}})`$ by the condition that $`Tot^{W^{\mathrm{fl}}}(V)`$ be the closed geometric substack of $`Tot^W(V)`$ defined by the coherent sheaf of ideals $`\mathrm{ker}(t^m)`$ (for $`m0`$). Note that for $`m0`$ this sequence of ideals becomes stationary, as can be verified by going to a scheme $`Z`$ with smooth surjection to $`Tot^W(V)`$. Note also that if $`Z^{\mathrm{fl}}`$ is the closed subscheme of $`Z`$ defined by the sheaf of ideals $`\mathrm{ker}(t^m)`$ on $`Z`$ (which is also the pullback of $`\mathrm{ker}(t^m)`$ from $`Tot^W(V)`$) then $`Z^{\mathrm{fl}}Tot^{W^{\mathrm{fl}}}(V)`$ is a smooth surjection.
The coordinate $`t`$ is not well-defined on $`𝒜`$ but on any scheme mapping to $`𝒜`$ we can locally chose a definition of $`t`$. On overlaps of neighborhoods of definition, the coordinates $`t`$ differ by invertible elements. Note that $`\mathrm{ker}(t^m)`$ on $`Z`$, for $`m0`$, is the ideal of $`t`$-torsion in $`𝒪(Z)`$. In particular, $`𝒪(Z^{\mathrm{fl}})`$ has no $`t`$-torsion so the map $`Z^{\mathrm{fl}}𝒜`$ is flat.
Suppose that $`(V,W^{})`$ is a flat filtration mapping to $`(V,W)`$, and let $`Z^{}`$ be a scheme with a smooth surjection to $`Tot^W^{}(V)`$, so that $`Z^{}𝐀^1`$ is flat. We may assume that there is a lifting to a map $`Z^{}Z`$. Then $`𝒪(Z^{})`$ has no $`t`$-torsion, so the $`t`$-torsion in $`𝒪(Z)`$ goes to zero in $`𝒪(Z^{})`$. Thus, the map factors through $`Z^{}Z^{\mathrm{fl}}`$. This implies that the map
$$Tot^W^{}(V)Tot^W(V)$$
factors through a map
$$Tot^W^{}(V)Tot^{W^{\mathrm{fl}}}(V).$$
The factorization is unique since
$$Tot^{W^{\mathrm{fl}}}(V)Tot^W(V)$$
is a closed substack defined by a sheaf of ideals. This completes the proof. $`|||`$
Next we note that $`Coh(Gr^W(V))`$ is equivalent to the category of objects in $`Coh(Tot^W(V))`$ on which $`t`$ acts as zero. The operation of restriction of a coherent sheaf from $`Tot^W(V)`$ to $`Gr^W(V)`$ corresponds to tensoring by $`𝒪_{Gr}:=𝒪/t𝒪\{1\}`$. If $``$ is a coherent sheaf on $`Tot^W(V)`$ then
$$|_{Gr^W(V)}=_𝒪𝒪_{Gr}.$$
All of these statements can be seen by using Lemma 1 to reduce to the case of a scheme $`Z`$ surjecting smoothly to $`Tot^W(V)`$.
Finally we define the annihilator ideals of $`(V,W)`$ to be the images
$$Ann^W(t^m;V):=im(\mathrm{ker}(t^m)_𝒪𝒪_{Gr}𝒪_{Gr}).$$
These are ideals on $`Gr^W(V)`$, in an increasing sequence,
$$Ann^W(t^m;V)Ann^W(t^{m+1};V)$$
and this sequence eventually becomes stationary (because $`𝒪_{Gr}`$ is noetherian).
###### Lemma 3
Suppose in the above situation, that for all $`m1`$ we have
$$Ann^W(t;V)=Ann^W(t^m;V).$$
Then the $`\mathrm{ker}(t^m)`$ are all equal, and they are already supported on $`Gr^W(V)`$, i.e. we have
$$\mathrm{ker}(t^m)=Ann^W(t;V).$$
Proof: It suffices to prove this for a morphism of schemes $`Z𝐀^1`$ (for the ideals $`\mathrm{ker}(\mathrm{})`$ and $`Ann(\mathrm{})`$ defined in the same way on $`Z`$). To see that this suffices, let $`Z`$ be a scheme with a smooth surjection
$$ZTot^W(V)\times _𝒜𝐀^1,$$
and note also that $`Z`$ has a smooth surjection to $`Tot^W(V)`$. The ideals defined relative to the map $`Z𝐀^1`$ are the pullbacks of the corresponding ideals on $`Tot^W(V)`$, and since $`ZTot^W(V)`$ is a smooth surjection, the formulae in question may be proven after pullback to $`Z`$.
Now the question for $`Z𝐀^1`$ is just a statement in usual algebraic geometry. To justify it, note that we have an exact sequence
$$0𝒯𝒪(Z)0$$
where $`𝒯`$ is the submodule of $`t`$-torsion in $`𝒪(Z)`$, and $``$ is flat over a neighborhood of $`0𝐀^1`$. Note that the torsion is of finite length, in other words $`𝒯`$ is really a $`𝐂[t]/t^M`$-module for some $`M`$. Then $`\mathrm{ker}(t^m)`$ and $`Ann(t^m,Z)`$ depend only on $`𝒯`$ (these objects may be defined for any $`𝐂[t]`$-module, and the statement is that the objects for $`𝒯`$ coincide with those for $`𝒪(Z)`$). On the other hand, the standard proof of the Jordan normal form works even in this context where the vector spaces are of infinite dimension but the order of nilpotency is bounded, so we may write
$$𝒯=M_i,M_i=𝐂[t]/t^{e_i}.$$
On a piece $`M_i`$, direct calculation shows that $`Ann(t^m;M_i)`$ is zero if $`m<e_i`$ and is $`𝐂`$ if $`me_i`$. Thus the condition that $`Ann(t^m;M_i)=Ann(t;M_i)`$ implies that $`e_i=1`$. On the pieces $`M_i`$ with $`e_i=1`$ (which are just of the form $`M_i=𝐂`$) one immediately gets that $`\mathrm{ker}(t^m)`$ is independent of $`m`$ and that these are equal to $`Ann(t,M_i)`$.
In this proof, the condition of the lemma (which we shall call $`\mathrm{𝐀𝟏}`$ below) is seen as equivalent to the statement that the $`t`$-torsion in $`𝒪(Z)`$ is actually a $`𝐂=𝐂[t]/(t)`$-module. $`|||`$
Conditions $`\mathrm{𝐀𝟏}`$ and $`\mathrm{𝐀𝟐}`$
We now state the first and second conditions on the annihilator ideals that will be used (for the weight filtration) in the definition of nonabelian mixed Hodge structure:
$`\mathrm{𝐀𝟏}(V,W)`$ This condition says that for all $`m1`$ we have
$$Ann^W(t;V)=Ann^W(t^m;V).$$
This says that the conclusion of the previous lemma applies.
If this condition holds, then $`Ann^W(t;V)`$ is an ideal on $`Tot^W(V)`$, and the quotient by this ideal is the structure sheaf for the canonical flat extension $`Tot^{W^{\mathrm{fl}}}(V)`$.
$`\mathrm{𝐀𝟐}(V,W)`$ This condition says that $`Ann^W(t,V)𝒪_{Gr^W(V)}`$ is contained in the ideal of $`0Gr^W(V)`$ (the point $`0`$ being defined because of the structure of complex, giving a structure of vector space on $`\pi _0Gr^W(V)`$).
###### Corollary 4
If $`(V,W)`$ satisfies conditions $`\mathrm{𝐀𝟏}(V,W)`$ and $`\mathrm{𝐀𝟐}(V,W)`$ then there is a quasi-finite flat cover $`Y𝐀^1`$, containing a lift of the origin, such that the morphism $`Tot^W(V)\times _𝒜YY`$ admits a section.
Proof: Let $`(V,W^{\mathrm{fl}})`$ denote the canonical flat subfiltration. Our conditions $`\mathrm{𝐀𝟏}`$ and $`\mathrm{𝐀𝟐}`$ imply that for $`m0`$ the ideals $`\mathrm{ker}(t^m)`$ (which are all equal) restricted to the fiber of the total space over the origin, are not $`(1)`$. This ideal defines $`Tot^{W^{\mathrm{fl}}}(V)`$, in particular $`Gr^{W^{\mathrm{fl}}}(V)`$ is nonempty. To finish the proof of the corollary, we may now assume that $`(V,W)`$ is a filtered object with $`Tot^W(V)`$ flat over $`𝒜`$ and such that $`Gr^W(V)`$ is nonempty. Thus
$$T^W(V)𝐀^1$$
is a flat morphism whose fiber over $`0`$ is nonempty. Choose a smooth surjection from a scheme $`ZT^W(V)`$ so that $`Z𝐀^1`$ is a flat morphism of schemes of finite type. In this situation it is automatic that there is $`Y𝐀^1`$ as in the statement, such that there exists a section $`YZ`$. $`|||`$
This corollary says that, if conditions $`\mathrm{𝐀𝟏}`$ and $`\mathrm{𝐀𝟐}`$ hold, then $`Gr^W(V)`$ is not “totally detached” from $`V`$.
Counterexample (1): A simple construction gives an example why we need some condition of type $`\mathrm{𝐀𝟐}`$ to insure that $`Gr^W(V)`$ isn’t detached from $`V`$. Let $`V`$ be a scheme, and let $`UB𝐆_m`$ be any scheme with $`𝐆_m`$ action. Then set
$$Tot^W(V):=VU$$
with its map to $`𝒜`$ (sending $`V`$ to $`[1]`$ and $`U`$ to the $`B𝐆_m`$ which is the origin). This defines a perfectly good filtration $`(V,W)`$ in which $`Gr^W(V)=U`$. However, $`U`$ doesn’t really have anything to do with $`V`$. In this construction, the ideals $`\mathrm{ker}(t^m)`$ and $`Ann^W(t^m;V)`$ are all equal to the unit ideal $`(1)`$ of $`𝒪_U`$. The canonical flat extension $`(V,W^{\mathrm{fl}})`$ is the filtration whose total space is just $`V`$ and whose associated-graded is empty.
Counterexample (2): One cannot conclude, in the above corollary, existence of a section over $`𝐀^1`$ (and certainly not existence of an equivariant section either) as shown by the following example. Let $`f:Y𝐀^1`$ be a double cover ramified at the origin, and let $`T:=Y\times 𝐆_m`$. Define a morphism $`p:T𝐀^1`$ by $`p(y,t):=tf(y)`$. This morphism is $`𝐆_m`$-equivariant (by construction) so it defines a morphism $`Y=T/𝐆_m𝒜`$, in other words we get a filtered object $`(V,W)`$ with $`V`$ equal to the fiber of $`T`$ over $`1𝐀^1`$ (which is also equal to the inverse image of $`𝐆_m𝐀^1`$ in $`Y`$) and with $`Tot^W(V):=T/𝐆_m=Y`$. We have
$$Tot^W(V)\times _𝒜𝐀^1=T\stackrel{p}{}𝐀^1.$$
The fiber of $`T`$ over the origin has multiplicity two, i.e. the morphism $`p`$ is ramified along the fiber; this implies that there doesn’t exist a section over $`𝐀^1`$ (or over any etale neighborhood of the origin).
Filtered and $`𝐆_m`$-equivariant perfect complexes
We can apply our discussion of filtered objects, to the stack $`nPERF`$ of length-$`n`$ perfect complexes, to obtain an $`n+1`$-stack $`F.nPERF`$ of filtered perfect complexes. The same holds for hemiperfect complexes, and we get an $`n+1`$-stack $`F.nHPERF`$ of filtered hemiperfect complexes.
We leave for another place the task of establishing the relationship between filtered complexes in the usual sense (cf Illusie ) for example, and objects of $`F.nPERF`$ or $`F.nHPERF`$.
Delooping and truncation
Suppose $`𝒞`$ is a complex of sheaves supported in $`[n,0]`$. Let $`\mathrm{\Omega }^k𝒞`$ denote the complex obtained by shifting $`𝒞`$ to the right by $`k`$ places, and doing the “intelligent” truncation in degrees $`0`$ which consists of replacing the term of degree zero by the kernel of its differential to the term of degree one. Thus $`\mathrm{\Omega }^k𝒞`$ is supported in the interval $`[kn,0]`$. If $`k>n`$ this produces the zero complex. We call this complex the $`k`$-th loop complex of $`𝒞`$. This terminology is due to the following compatibility with the Dold-Puppe operation:
$$DP(\mathrm{\Omega }^k𝒞)\mathrm{\Omega }^k(DP𝒞,0)$$
where the terminology on the right means taking $`k`$ times the loop space based at the zero section of $`DP(𝒞)`$. The equivalence is a canonical weak homotopy equivalence of $`nk`$-stacks.
If $`𝒞`$ is a hemiperfect complex then $`\mathrm{\Omega }^k𝒞`$ is again hemiperfect. The operation $`\mathrm{\Omega }`$ combines a shift and a truncation, so if $`𝒞`$ is perfect, the loop complex might be hemiperfect without being perfect. In fact this is the main motivation for introducing hemiperfect complexes: they arise as the truncations of perfect complexes.
###### Lemma 5
Suppose $`(𝒞,F)`$ is a filtered hemiperfect complex. Suppose that the cohomology sheaves $`H^i(Tot^F(𝒞)/𝒜)`$ are vector bundles over $`𝒜`$. Then there exists a splitting, namely an equivalence in $`F.nHPERF`$ between $`(𝒞,F)`$ and $`_i(H^i(𝒞),F)`$. In particular $`(𝒞,F)`$ is a filtered perfect complex, as are all of the loop complexes $`(\mathrm{\Omega }^k𝒞,F)`$. For filtered perfect complexes the condition (which we denote by $`\mathrm{𝐒𝐭𝐫}`$) that the $`H^i`$ be vector bundles over $`𝒜`$, is equivalent to strictness of the differentials or equivalently degeneration of the spectral sequence at $`E_1`$. Finally, strictness of the differentials implies that the cohomology vector sheaves are vector bundles.
Proof: Suppose the cohomology vector sheaves are vector bundles. The total complex $`Tot^F(𝒞)`$ is classified by a collection of extension classes in higher $`Ext`$ groups between the cohomology vector sheaves. If the vector sheaves are vector bundles then these $`Ext`$ groups coincide with the usual ones, and in fact are calculated cohomologically: $`Ext^i(U,V)=H^i(𝒜,U^{}V)`$. But $`𝒜`$ is cohomologically trivial for vector bundle coefficients (use the principal $`𝐆_m`$-bundle $`𝐀^1𝒜`$ to see this). Thus the higher $`Ext`$ groups vanish which implies that the complex splits.
If the complex splits, it is clear that the spectral sequence degenerates. This latter condition is equivalent by Deligne Proposition 1.3.2, to strictness of the differentials; and finally, strictness of the differentials means that the differentials of the total complex are strict morphisms of vector bundles, which in turn implies that the cohomology vector sheaves are vector bundles. $`|||`$
If $`(𝒞,W)`$ is a filtered hemiperfect complex then we can replace the vector scheme $`(Tot^W(𝒞))^0`$ by its flat subscheme (cf Lemma 2 p. 2) and obtain the total complex for a filtered perfect complex. We call this the perfect subfiltration denoted $`W^{\mathrm{fl}}W`$.
Annihilator ideals for filtered hemiperfect complexes
In the spirit of the previous remark, we can also look at the annihilator ideals of a hemiperfect complex. This discussion constitutes, of course, an abelian version of the previous discussion of annihilator ideals for $`n`$-stacks, and could have been done before as a motivation for the nonabelian version of the discussion. As the notation is actually simpler in the general nonabelian case, we have not chosen that order of things.
Suppose the base $`B`$ is $`𝐀^1`$ or $`𝒜`$. If $`C^{}`$ is a hemiperfect complex, then we define as before the annihilator ideals $`Ann(t^m;C^{})`$ which are ideals on $`H^0(C^{})`$. In this case they are ideals of linear subspaces (in particular they are contained in the maximal ideal defining the origin, in other words condition $`\mathrm{𝐀𝟐}`$ is always satisfied). To define the annihilator ideals, $`C^0C^{}`$ serves as a smooth surjection from a scheme, and the annihilator ideals are defined as descending from the corresponding ideals defined on the vector scheme $`C^0`$. In particular, if $`C^{}`$ were actually a perfect complex (i.e. if $`C^0`$ were a vector bundle) then the annihilator ideals would vanish.
It also follows that the annihilator ideals are compatible with the Dold-Puppe construction.
If the base is $`𝒜`$ so in fact $`C^{}`$ corresponds to a filtered complex $`(𝒞,W)`$ (which we call a “truncated filtered complex”) then we denote the annihilator ideals by $`Ann^W(t^m;𝒞)`$.
We have the following relationship between the annihilator ideals and the spectral sequence of a filtered complex. It generalizes Lemma 5 one step further, to spectral sequences degenerating at $`E_2`$.
###### Lemma 6
Suppose $`(𝒟,W)`$ is a filtered perfect complex of length $`n+k`$, and let $`𝒞=\mathrm{\Omega }^k𝒟`$ be the loop complex which is a filtered hemiperfect complex of length $`n`$. Suppose that the spectral sequence $`E_{}(𝒟)`$ degenerates at $`E_2`$. Then we have condition $`\mathrm{𝐀𝟏}`$ saying that
$$m1,Ann^W(t^m;𝒞)=Ann^W(t;𝒞).$$
Proof: Write $`𝒟=\{D^i,d(i):D^iD^{i+1}\}`$ with $`D^i`$ being filtered vector spaces, which we can think of as being vector bundles over $`𝒜`$. We obtain complexes $`\{E_r^i,d_r(i):E_r^iE_r^{i+1}\}`$ forming the spectral sequence.
Fix $`i`$; then the behavior of the differentials $`d_r(i)`$ for a given $`i`$ (i.e. their vanishing or not) depends only on the original differential between filtered vector spaces $`d(i)`$. To see this, note that if we consider the complex $`\stackrel{~}{D}^{}`$ consisting only of the terms $`D^i`$ and $`D^{i+1}`$ then this gives rise to a spectral sequence $`\stackrel{~}{E}_r^{}`$ and
$$\stackrel{~}{E}_r^iE_r^i$$
is surjective, and
$$E_r^{i+1}\stackrel{~}{E}_r^{i+1}$$
is injective, and the differential $`\stackrel{~}{d}_r(i)`$ factors through $`d_r(i)`$; thus $`d_r(i)=0`$ if and only if $`\stackrel{~}{d}_r(i)=0`$.
We claim that if $`d_r(0)=0`$ for all $`r2`$, then the condition of the lemma holds for the truncated complex $`𝒞`$. Note that the condition of the lemma, for the truncated complex $`𝒞`$, again depends only on the differential $`d(0):D^0D^1`$ of the original complex. In view of all of this, we can reduce to consideration of a complex $`𝒟`$ concentrated only in degrees $`0`$ and $`1`$; for the rest of the proof we make this hypothesis.
Now, our complex $`𝒟`$ breaks up into a direct sum of complexes where either one of the terms is zero, plus a sum of complexes of the form
$$d:LM$$
where $`L`$ and $`M`$ are one-dimensional filtered vector spaces and $`d`$ is a nonzero map between them. To see this decomposition, work with
$$d(0):D^0D^1$$
as a morphism of vector bundles over $`𝒜`$ and use the Gauss method to reduce the matrix of $`d(0)`$ to a diagonal form (taking note of the $`𝐆_m`$-equivariance, which is preserved if the Gauss operations are done correctly).
Now in view of this decomposition (and noting that the cases where $`D^0=0`$ or $`D^1=0`$ are trivial), it suffices to treat the case of a complex of the form $`d:LM`$ with $`L`$ and $`M`$ being one-dimensional. Let $`l`$ and $`m`$ denote the respective weights of $`L`$ and $`M`$ (i.e. the places where the single nonzero steps occur in the filtrations). Note that $`ml`$. In this case, it is easy to see that the condition of degeneration at $`E_2`$ is equivalent to the condition $`ml1`$. From this we deduce that the differential, considered as a map between trivial bundles over $`𝐀^1`$, is multiplication either by $`1`$ or by $`t`$ (here, pull back from $`𝒜`$ to $`𝐀^1`$). From this, we get the desired statement about the annihilator ideals. $`|||`$
A decomposition for $`𝐆_m`$-equivariant perfect complexes
In a Dold-Puppe linearized filtration, the object $`LGr^W(V)`$ is a $`𝐆_m`$-equivariant hemiperfect complex. In this subsection we point out that such a thing decomposes into terms $`LGr_k^W(V)`$ according to the character of the action of $`𝐆_m`$.
The choice of sign is motivated by the discussion of filtered vector spaces on p. Nonabelian mixed Hodge structures.
###### Proposition 7
Let $`X`$ be any $`n`$-stack. A hemiperfect complex $`C`$ over $`X\times B𝐆_m`$, breaks up canonically as a direct sum
$$C=\underset{k}{}C_k$$
where the $`C_k`$ are hemiperfect complexes on $`X\times B𝐆_m`$ such that the action of $`𝐆_m`$ on the cohomology vector sheaves of $`C_k`$ is via the character $`tt^k`$.
Sketch of proof: The proof uses two techniques which we don’t develop here (but which will, one hopes, be developped elsewhere). These techniques are: tensoring hemiperfect complexes by line bundles; and direct images of hemiperfect complexes. Let $`\omega ^k`$ be the rank one line bundle on $`B𝐆_m`$ corresponding to the character $`tt^k`$ (and use the same notation for its pullback to $`X\times B𝐆_m`$). Denoting by $`p:X\times B𝐆_mX`$ the projection, put
$$C_k:=\omega ^kp^{}(p_{}(C\omega ^k))$$
where $`p_{}`$ is the (full derived) direct image functor for hemiperfect complexes, which in this case gives a hemiperfect complex as an answer. We have a morphism of hemiperfect complexes
$$i:CC_k.$$
Noting that the higher cohomology of $`B𝐆_m`$ with vector sheaf coefficients vanishes, and that the global sections functor just takes the invariants, we obtain that $`i`$ is an isomorphism and that the $`C_k`$ have the desired property. $`|||`$
If the complex $`C`$ is perfect then the components $`C_k`$ are also perfect. This is because the fact that the higher cohomology of $`B𝐆_m`$ vanishes means that no truncation operation is necessary to give back complexes supported in $`[n,0]`$, and the direct image of a perfect complex remains perfect.
Remark: The construction of the $`C_k`$ and the isomorphism $`i`$, are $`n+1`$-functorial in $`C`$ and in $`X`$. In particular we obtain a morphism of $`n+1`$-stacks
$$\underset{¯}{Hom}(B𝐆_m,nHPERF)\underset{k}{}nHPERF,$$
$$C(\mathrm{},p_{}(C\omega ^k),\mathrm{})$$
which is fully faithful with essential image the “restricted product” (consisting of those elements which are zero except for a finite number of places).
Applying the theorem to the case $`X=𝒜`$, we obtain:
###### Corollary 8
If $`(C,F)`$ is a $`𝐆_m`$-equivariant filtered perfect or hemiperfect complex then it decomposes canonically as a direct sum of terms $`(C_k,F)`$ such that the action of $`𝐆_m`$ on the $`H^i(Tot^F(C_k))`$ is via the character $`tt^k`$.
Notation: If $`(V,W,LGr^W(V))`$ is a Dold-Puppe linearized filtration then we denote by $`LGr_k^W(V)`$ the component of degree $`k`$ in the decomposition given by the previous corollary (recall that $`LGr^W(V)`$ is a $`𝐆_m`$-equivariant perfect complex). Similarly, if $`(V,W)`$ is a filtered (hemi)perfect complex then $`Gr^W(V)`$ is again a $`𝐆_m`$-equivariant (hemi)perfect complex, and again the previous corollary gives a decomposition into components denoted $`Gr_k^W(V)`$. We call these components (in both cases) the “$`k`$-th graded pieces” of the associated-gradeds.
Exercise: The non-canonical decomposition of p. Nonabelian mixed Hodge structures also holds for $`𝐆_m`$-equivariant perfect complexes; in other words, if $`C:B𝐆_mnPERF`$ is a $`𝐆_m`$-equivariant perfect complex then it decomposes as
$$C=\underset{i}{}H^i(C)[i],$$
non-canonically but still into $`𝐆_m`$-equivariant pieces.
Weight-filtered $`n`$-stacks
Before getting to the full notion of nonabelian mixed Hodge structure, we will first discuss just that part which concerns the weight filtration. Note that this is what is new in the present paper: the Hodge filtration has been the subject of several previous papers, and its integration into the picture is essentially straigthforward. On the other hand, for the weight filtration, we have introduced an essentially new idea which is that of “linearization” cf page Nonabelian mixed Hodge structures. To explain this alone we introduce the definition of a “weight-filtered $`n`$-stack”.
Definition: A pre-weight-filtered $`n`$-stack is a quadruple
$$\{(V,W),LGr^W(V),\zeta \}$$
where $`(V,W)`$ is a filtered geometric $`n`$-stack (cf page Nonabelian mixed Hodge structures and page Nonabelian mixed Hodge structures), where $`LGr^W(V)`$ is a hemiperfect complex of length $`n`$ (cf page Nonabelian mixed Hodge structures), and
$$\zeta :DP(LGr^W(V))Gr^W(V)$$
is a Dold-Puppe linearization (cf page Nonabelian mixed Hodge structures).
In other words, the $`n+1`$-stack of pre-weight filtered $`n`$-stacks is just $`F^{DP}.nGEOM`$.
Recall that we have considered two conditions which one can put on a filtration:
$`\mathrm{𝐀𝟏}(V,W)`$: that the annihilator ideals $`Ann^W(t^m;V)`$ are all equal for $`m1`$; qnd
$`\mathrm{𝐀𝟐}(V,W)`$: that the annihilator ideal $`Ann^W(t;V)`$ is contained in the ideal of the zero section of $`Gr^W(Tot^F(V))`$.
A weight-filtered $`n`$-stack is a pre-weight-filtered $`n`$-stack which satisfies conditions $`\mathrm{𝐀𝟏}(V,W)`$ and $`\mathrm{𝐀𝟐}(V,W)`$.
An example: deformation to the tangent space
An important basic example of a pre-weight-filtered $`n`$-stack is the following, which puts a linearized filtration on any pointed geometric $`n`$-stack. This discussion is closely related to Fulton’s “deformation to the normal cone” see .
Let
$$p^{}:Q^{}𝐀^1$$
be the closed subscheme of $`𝐀^2`$ defined by the equation $`(tu)(t+u)=0`$ with projection $`p^{}`$ given by the coordinate $`t`$. Let $`Q:=Q^{}/𝐆_m`$ where the action of $`𝐆_m`$ on $`Q^{}`$ is by homotheties. This action covers the standard action on $`𝐀^1`$ so $`p^{}`$ induces a projection
$$p:Q𝒜.$$
Let $`P^{}=𝐀^1Q^{}`$ be the subspace defined by $`(tu)=0`$ (it defines a section of $`p^{}`$) and let $`P=P^{}/𝐆_m`$. Thus $`P`$ is the image of a section of $`p`$ (i.e. $`P=𝒜`$). The projection $`p:Q𝒜`$ is flat, projective and the fibers are finite schemes. The fiber over $`[1]`$ is a scheme with two points (and $`P_{[1]}`$ is one of the two points) whereas the fiber over $`[0]`$ is the spectrum of the dual numbers $`𝐂[u]/u^2`$ with $`P_{[0]}`$ being the basepoint and the $`𝐆_m`$-action being by homotheties.
If $`Y`$ is a geometric $`n`$-stack then
$$\underset{¯}{Hom}(\frac{Q}{𝒜},\frac{Y\times 𝒜}{𝒜})$$
is a geometric $`n`$-stack mapping to $`𝒜`$. To see this (which is an easy case of (Conjecture 1, , p. 27) or alternatively a relative version of (Theorem 8.1, op cit)), note that the construction of taking relative internal $`\underset{¯}{Hom}`$ from $`Q`$ to an object, is compatible with homotopy fiber products and takes smooth morphisms to smooth morphisms. In the case of $`Y`$ being a smooth scheme, it is clear that the $`\underset{¯}{Hom}`$ stack defined above is again a smooth scheme over $`𝒜`$. Now any scheme is obtained from smooth ones as a homotopy fiber product, so we get the desired statement for $`Y`$ a scheme; finally, it follows for any geometric $`n`$-stack $`Y`$ by applying the definition of geometric $`n`$-stack, which only involves notions of smoothness, of being a scheme, and of homotopy fiber products.
The operation of restricting a morphism to the section $`P`$ provides a morphism
$$\underset{¯}{Hom}(\frac{Q}{𝒜},\frac{Y\times 𝒜}{𝒜})\underset{¯}{Hom}(\frac{P}{𝒜},\frac{Y\times 𝒜}{𝒜})=Y\times 𝒜.$$
Suppose now that $`y:Y`$ is a point. Put
$$Tot^{W(y)}(Y):=\underset{¯}{Hom}(\frac{Q}{𝒜},\frac{Y\times 𝒜}{𝒜})\times _{Y\times 𝒜}𝒜,$$
where the first morphism is the restriction morphism described above and the second morphism is $`(y,1):𝒜Y\times 𝒜`$. This is a geometric $`n`$-stack mapping to $`𝒜`$. From the previous descriptions of the fibers of $`Q`$ over $`[1]`$ and $`[0]`$ respectively, we get immediately that the fiber of $`Tot^{W(y)}(Y)`$ over $`[1]`$ is $`Y`$ and that the fiber over $`[0]`$ is the tangent stack $`T_y(Y)`$ (see for the definition of the tangent stack). Thus, by definition $`Tot^{W(y)}(Y)`$ defines a filtration $`W(y)`$ of $`Y`$, whose associated-graded is the tangent stack to $`Y`$ at $`y`$.
Furthermore, recall from that the tangent stack $`T_y(Y)`$ actually has a structure of “spectrum”, i.e. it has a canonical infinite delooping structure. Since we are in characteristic zero, a spectrum is the same thing as a complex, thus $`T_y(Y)`$ has a structure of perfect complex (one can check in the construction that this structure is equivariant for the action of $`𝐆_m`$ by homotheties on $`Spec(𝐂[u]/u^2)`$). Letting $`LGr^{W(y)}(Y)`$ denote this perfect complex whose Dold-Puppe is $`Gr^{W(y)}(Y)`$, we obtain a linearization of the filtration $`W(y)`$, which makes $`(Y,W(y),LGr^{W(y)}(Y))`$ into a pre-weight-filtered geometric $`n`$-stack.
We obtain an $`n+1`$-functor
$$nGEOM^{ptd}F^{DP}.nGEOM.$$
Exercises: (1) Observe that condition $`\mathrm{𝐀𝟐}`$ is always satisfied (indeed $`Tot^{W(y)}(Y)`$ comes with a section.
(2) Show that if $`Y`$ is a scheme with quadratic singularities at $`y`$ (or if $`Y`$ is a geometric $`n`$-stack with a smooth surjection from such a scheme) then $`W(y)`$ satisfies condition $`\mathrm{𝐀𝟏}`$. The subscheme defined by the annihilator ideal is the normal cone.
While we don’t imagine that all weight filtrations arise in this way, we do think that this should provide an important construction of certain types of weight filtrations. The relationship between weight filtrations arising “in nature” and those arising from this construction, seems to be an interesting question for further thought.
Interaction with the Hodge filtration
We apply the discussion from page Nonabelian mixed Hodge structures to the case of the weight filtration. A weight-Hodge linearized bifiltration of a geometric $`n`$-stack is an object of $`F.LF.nGEOM=F.F^{DP}.nGEOM`$ where $`DP:nHPERFnGEOM`$ is the Dold-Puppe linearization functor considered above. This object is denoted $`(V,W,F)`$ where $`V`$ is a geometric $`n`$-stack; where $`(V,W)`$ is the “weight-filtration” on $`V`$ which is the one which is linearized, i.e. it is a filtration plus a ($`𝐆_m`$-equivariant) perfect complex $`LGr^W(V)`$ whose associated ($`𝐆_m`$-equivariant) geometric $`n`$-stack is $`Gr^W(V)`$; and finally where $`F`$ is a filtration of $`(V,W)`$ in the stack $`F^{DP}.nGEOM`$. As described above, concretely this means that $`Tot^F(V,W)`$ is a Dold-Puppe-linearized filtered geometric $`n`$-stack over $`𝒜`$. In turn this means that there is a $`𝐆_m`$-equivariant filtered perfect complex
$$(LGr^W(V),F):B𝐆_mF.nHPERF$$
whose underlying filtered geometric $`n`$-stack (Dold-Puppe) is $`(Gr^W(V),F)`$.
In order to clarify what is going on, when we look at both the weight filtration and the Hodge filtration we will use two different subscripts for the two copies of $`𝒜`$ which come into the picture: the copy which concerns the weight filtration will be denoted by $`𝒜_{wt}`$ and the copy concerning the Hodge filtration will be denoted by $`𝒜_{hod}`$. Thus a weight-Hodge linearized bifiltration is a morphism of geometric $`n`$-stacks
$$Tot^{W,F}(V)𝒜_{hod}\times 𝒜_{wt},$$
together with a perfect complex $`Tot^F(LGr^W(V))`$ over $`[0]_{wt}\times 𝒜_{hod}`$.
In the presence of both weight and Hodge filtrations, we would like to impose our two conditions $`\mathrm{𝐀𝟏}`$ and $`\mathrm{𝐀𝟐}`$ on the weight filtration, plus some additional conditions related to the Hodge filtration. These conditions are quite naturally imposed on the biggest stack for which they make sense, in the following way.
$`\mathrm{𝐀𝟏}(V,W,F)`$: that the annihilator ideals $`Ann^W(t^m;Tot^F(V))`$ satisfy the condition $`\mathrm{𝐀𝟏}(Tot^F(V),W)`$ (they are all equal for $`m1`$);
$`\mathrm{𝐀𝟐}(V,W,F)`$: that the annihilator ideal $`Ann^W(t;Tot^F(V))`$ satisfies condition $`\mathrm{𝐀𝟐}(Tot^F(V),W)`$, namely that it is contained in the ideal of the zero section of $`Gr^W(Tot^F(V))`$; and
$`\mathrm{𝐅𝐥}(V,W,F)`$: that the Hodge filtration $`F`$ is flat, by which we mean that the morphism corresponding to the Hodge filtration,
$$Tot^{W,F}(V)𝒜_{hod}$$
is flat.
These three conditions are meant to insure that the biflitration $`(V,W,F)`$ is reasonably well connected to $`V`$ itself. Condition $`\mathrm{𝐀𝟐}`$ prevents silly constructions such as in Counterexample (1) of p. 4. The flatness condition means that the only degeneracies involved are the non-flatness of the weight filtration. Note in particular that because of the flatness condition, $`Tot^FGr^W(V)𝒜_{hod}`$ is flat, which implies that the hemiperfect linearization is actually a perfect one, i.e. $`Tot^FGr^W(V)`$ is a perfect complex. Condition $`\mathrm{𝐀𝟏}`$ puts a limit on how complicated the degeneracy in the weight-filtration direction can be.
To bring things back down to earth a bit, we take note of the following remarks which follow immediately from the definitions. The homotopy stacks $`\pi _i(Gr^W(V))`$ are vector sheaves over $`[0]_{wt}B𝐆_m`$, i.e. they are finite dimensional complex vetor spaces with $`𝐆_m`$-action. This includes the cases $`i=0`$ and $`i=1`$ because they are the cohomology sheaves of the perfect complex $`LGr^W(V)`$:
$$\pi _i(Gr^W(V))=H^i(LGr^W(V)).$$
In the case $`i=0`$ this means that $`\pi _0(Gr^W(V))`$ is given the extra structure of a complex vector space with $`𝐆_m`$-action. Furthermore, a perfect complex over $`B𝐆_m`$ splits up (non-canonically) as the product of its cohomology sheaves, so $`Gr^W(V)`$ splits as a product of its Postnikov-Eilenberg-MacLane components (cf the exercise on p. 8):
$$Gr^W(V)\underset{i=0}{\overset{n}{}}K(\pi _i(Gr^W(V))/B𝐆_m,i).$$
Now the Hodge filtration induces a filtration on $`LGr^W(V)`$. This means that $`(LGr^W(V),F)`$ corresponds to a perfect complex
$$Tot^F(LGr^W(V))𝒜_{hod}.$$
In particular the cohomology sheaves of this are vector sheaves on $`𝒜_{hod}`$, so they correspond to filtrations of the vector spaces $`\pi _i(Gr^W(V))`$ in the sense that these latter are objects of $`F.VESH`$ (cf p. Nonabelian mixed Hodge structures). This leads us to our next condition, stating that these vector sheaves are actually vector bundles (and thus that the filtration is a filtration in the classical sense i.e. an object of $`F.VECT`$). We call this condition “strictness” (see Lemma 5):
$`\mathrm{𝐒𝐭𝐫}(V,W,F)`$: that the cohomology sheaves
$$H^i(Tot^F(LGr^W(V)))$$
over $`𝒜_{hod}`$, are actually vector bundles.
We can view $`(LGr^W(V),F)`$ as a filtered complex, and by Lemma 5 this condition amounts to asking that it be a strict filtered complex in the sense of . This condition means, in view of the discussion of $`𝐆_m`$-equivariant perfect complexes, that $`Tot^F(LGr^W(V))`$ (which is a $`𝐆_m`$-equivariant perfect complex over $`𝒜_{hod}`$) decomposes as a direct sum of its cohomology vector bundles. Recall from Lemma 5 that it is equivalent to degeneration of the spectral sequence for the Hodge filtration $`F`$ on $`LGr^W(V)`$, at the $`E_1`$-term.
In view of the equivalence $`Gr^W(V)=\mathrm{\Phi }(LGr^W(V))`$ the above strictness condition can be rewritten as saying that the $`\pi _i(Tot^F(Gr^W(V)))`$ are vector bundles over $`𝒜`$.
Analytic and real structures
In the main definitions to come in Part II, we shall need some basic definitions about analytic objects and real structures, particularly as they relate to filtered objects.
Algebraic to analytic objects
Let $`𝒵^{\mathrm{an}}`$ denote the site of paracompact complex analytic spaces with the etale topology. We have a functor $`\mathrm{\Phi }:𝒵𝒵^{\mathrm{an}}`$ defined by $`\mathrm{\Phi }(X):=X^{\mathrm{an}}`$. Define the operation $`^{\mathrm{an}}`$ from $`n`$-stacks on $`𝒵`$ to $`n`$-stacks on $`𝒵^{\mathrm{an}}`$, by the formula
$$^{\mathrm{an}}:=\mathrm{\Phi }_!(),$$
where we put (from )
$$\mathrm{\Phi }_{!,pre}()(Y):=\mathrm{colim}_{Y\mathrm{\Phi }(X)}(X),$$
and we define $`\mathrm{\Phi }_!()`$ to be the $`n`$-stack associated to the $`n`$-prestack $`\mathrm{\Phi }_{!,pre}()`$.
The $`colim`$ in the definition or $`\mathrm{\Phi }_{!,pre}`$ is of course the homotopy colimit of $`n`$-stacks (and since we’re talking about $`n`$-stacks of groupoids, this is the same as the homotopy colimit of the diagram of spaces ).
We have the standard notion of smooth morphism between complex analytic spaces, and by the construction of this gives rise to the notion of analytic geometric $`n`$-stack. Note however that in the analytic situation we don’t make any analogue of the “very presentable” condition.
The following lemma gives the fundamental compatibility between the analytic and algebraic notions of geometricity.
###### Lemma 9
The functor $`^{\mathrm{an}}`$ is compatible with homotopy fiber products. Furthermore, it takes a scheme to its associated analytic space. As a consequence, it takes geometric $`n`$-stacks to analytic geometric $`n`$-stacks.
Proof: The $`hocolim`$ involved in the definition of $`\mathrm{\Phi }_{!,pre}`$ is filtering (because it is taken over a category which admits finite products and fiber products), therefore it commutes with finite homotopy limits. The operation of taking the associated $`n`$-stack is also compatible with finite homotopy limits, see , so the operation $`^{\mathrm{an}}`$ is compatible with finite homotopy limits. The second sentence is standard; and the third sentence is an immediate corollary of the fact the definition of “geometric $`n`$-stack” (resp. “analytic geometric $`n`$-stack”) refers only to homotopy fiber products, schemes (resp. analytic spaces) and smoothness of maps between schemes (resp. smoothness of maps between analytic spaces). $`|||`$
To understand in a simple way what is going on in the above lemma, suppose $``$ is a geometric $`n`$-stack with a surjective morphism from a scheme $`Z`$. Let $`R=Z\times _{}Z`$, which is a geometric $`(n1)`$-stack. Thus in a certain sense $``$ is presented as the quotient of $`Z`$ by the “relation” $`R`$. Now $`Z^{\mathrm{an}}^{\mathrm{an}}`$ is a smooth surjection and
$$R^{\mathrm{an}}=Z^{\mathrm{an}}\times _{^{\mathrm{an}}}Z^{\mathrm{an}}.$$
Thus $`^{\mathrm{an}}`$ is presented as the quotient of $`Z^{\mathrm{an}}`$ by $`R^{\mathrm{an}}`$.
We define filtrations in the same way as in the algebraic case, using the analytic stack
$$𝒜^{\mathrm{an}}:=(𝐀^1)^{\mathrm{an}}/(𝐆_m)^{\mathrm{an}}.$$
(which is the analytic stack associated to $`𝒜`$). Thus for example a filtered analytic geometric $`n`$-stack is a morphism of analytic geometric $`n`$-stacks $`T𝒜^{\mathrm{an}}`$. The only construction we need is that if $`(V,W)`$ is a filtered algebraic geometric $`n`$-stack then by taking the analytic $`n`$-stack $`T^{\mathrm{an}}`$ associated to the total space $`T=Tot^W(V)`$ we obtain a filtered analytic geometric $`n`$-stack $`(V^{\mathrm{an}},W)`$.
An analytic perfect complex (resp. analytic hemiperfect complex) over a base $`B`$ is a morphism $`BnPERF^{\mathrm{an}}`$ (resp. $`BnHPERF^{\mathrm{an}}`$). If $`B`$ is a complex analytic space, this is the same thing as a complex of sheaves of $`𝒪^{\mathrm{an}}`$-modules on $`B`$ which is locally quasiisomorphic to a complex of vector bundles supported in $`[n,0]`$ (resp. the same but with $`C^0`$ a vector scheme <sup>5</sup><sup>5</sup>5 We define an analytic vector scheme as being something which is locally the kernel of a map of finite rank vector bundles. Now the $`1`$-stack of analytic vector schemes is the analytic stack associated to $`VESCH`$ because any analytic vector scheme is pulled back from a sort of “universal” algebraic vector scheme over the space of matrices, by an analytic map. One conjectures that this definition of analytic vector scheme is equivalent to the notion of “vector space in the category of complex analytic spaces over the base” but we don’t have a proof of that. ). Again, given an algebraic perfect complex over a base $`X`$ we obtain an associated analytic perfect complex over $`X^{\mathrm{an}}`$.
An important observation in what follows is that an analytic perfect complex over $`𝒜^{\mathrm{an}}`$ has a canonical algebraic structure (because of the $`𝐆_m`$-action on $`𝐀^1`$ whose orbit goes out to the only point at infinity). In particular, if $`(U_1,W)`$ and $`(U_2,W)`$ are algebraic filtered perfect complexes, then any morphism between the associated analytic filtered complexes, is actually algebraic. This remark will be used in defining the notion of linearization of a pre-namhs: whereas there is an analytic equivalence involved in the notion of pre-namhs, for the “linear” object this equivalence is taken to be algebraic because no greater generality would be obtained by using an analytic equivalence.
Real structures
We treat here the extension of ground fields $`[𝐂:𝐑]`$. Most of what we say should be valid for an arbitrary extension of ground field (although in the case of infinite field extensions, some extra work is necessary because we have to go out of the realm of schemes of finite type). In fact we state many things without proof, things which would be more appropriately dealt with in the general context of an arbitrary field extension.
Let $`𝒵_𝐑`$ denote the category of schemes of finite type over $`𝐑`$ with the etale topology. We have a complexification functor $`\mathrm{\Psi }:𝒵_𝐑𝒵`$ which sends a real scheme to its associated complex scheme. On the level of schemes, this functor is defined by
$$\mathrm{\Psi }(X):=X\times _{Spec(𝐑)}Spec(𝐂).$$
As in the previous section this gives a functor $`\mathrm{\Psi }_!`$ from $`n`$-stacks on $`𝒵_𝐑`$ to $`n`$-stacks on $`𝒵`$. We denote this operation by $`XX_𝐂`$; often an object on $`𝒵_𝐑`$ will be denoted by $`X_𝐑`$ in which case the complexified object is denoted $`X_𝐂`$.
The formula for $`\mathrm{\Psi }_!`$ is
$$\mathrm{\Psi }_!(X)(Y):=\mathrm{colim}_{Y\mathrm{\Psi }(U)}X(U),$$
where $`X`$ is a stack on $`𝒵_𝐑`$ and $`Y𝒵`$ is a complex scheme. The colimit is taken over real schemes $`U`$ with maps $`Y\mathrm{\Psi }(U)`$. Note in this case however (this is a difference with the analytification functor) that the category over which the colimit is taken has a final object. Indeed $`\mathrm{\Psi }`$ acting on schemes has a left adjoint $`\mathrm{\Psi }^!`$ defined by putting $`\mathrm{\Psi }^!(Y)`$ equal to the real scheme obtained by composing $`YSpec(𝐂)Spec(𝐑)`$. A morphism of complex schemes
$$Y\mathrm{\Psi }(U)=U\times _{Spec(𝐑)}Spec(𝐂)$$
is the same thing as a morphism of real schemes $`\mathrm{\Psi }^!(Y)U`$ (take the first projection of the previous). Thus the colimit may be seen as a colimit over maps $`\mathrm{\Psi }^!(Y)U`$ and we obtain the formula
$$\mathrm{\Psi }_!(X)(Y):=X(\mathrm{\Psi }^!(Y)).$$
On stacks the functor $`\mathrm{\Psi }_!`$ has a left adjoint again denoted $`\mathrm{\Psi }^!`$ which, heuristically, takes a complex stack $`XSpec(𝐂)`$ to the real stack obtained by composing $`XSpec(𝐂)Spec(𝐑)`$. (Technically speaking this definition doesn’t make sense because $`X`$ is a stack over the site $`Sch/𝐂`$ and $`Spec(𝐑)`$ doesn’t exist in that context; to be correct, $`\mathrm{\Psi }^!`$ must be defined by the adjoint property.) There is a natural transformation $`\mathrm{\Psi }^!(X)\mathrm{\Psi }^!Spec(𝐂)`$ and this gives an equivalence of $`n+1`$-categories between the $`n`$-stacks over $`𝒵`$ and the $`n`$-stacks over $`𝒵_𝐑`$ provided with morphism to $`\mathrm{\Psi }^!Spec(𝐂)`$. Using this point of view (plus a slight abuse of notation) it makes sense to write
$$\mathrm{\Psi }_!(X)=X\times _{Spec(𝐑)}Spec(𝐂).$$
If $`X`$ is an $`n`$-stack on $`𝒵`$ then we say that a real structure on $`X`$ is the specification of a real $`n`$-stack $`Y_𝐑`$ plus an equivalence $`XY_𝐂`$.
There is of course a “Galois” version of this notion, namely if $`X`$ is a complex stack then composing with the complex conjugation involution $`𝒵𝒵`$ we obtain a new complex stack denoted $`\overline{X}`$; and a real structure on $`X`$ is the same thing as a twisted $`𝐙/2`$-action on $`X`$ consisting principally of an equivalence $`X\overline{X}`$ plus some higher coherence data. We don’t wish to go into further detail on this at present. Note however that in terms of the “Galois” point of view, $`\mathrm{\Psi }^!(X)`$ is the scheme $`X\overline{X}`$ with real structure induced by the involution which exchanges the two pieces.
As usual there is a notion of smooth map between real schemes, and using this notion we obtain following the procedure of , the notion of real geometric $`n`$-stack (which is an $`n`$-stack on $`𝒵_𝐑`$). We impose our tacit hypothesis of the introduction, that included in the condition of being a real geometric $`n`$-stack is the condition that the associated complex geometric $`n`$-stack is also very presentable. As the condition is being made on the complexification, it concerns the homotopy group sheaves for all choices of basepoint in the complexified stack. Nonetheless, one can say that for real basepoints the homotopy group sheaves of the complex stack are the complexifications of those of the real stack. In the connected case when there is only one choice of basepoint up to equivalence, this determines the homotopy group sheaves of the complexified stack.
###### Lemma 10
The operation of complexification of a real $`n`$-stack is compatible with homotopy fiber products. Therefore it sends real geometric $`n`$-stacks to geometric $`n`$-stacks. Furthermore, if $`X_𝐑`$ is an $`n`$-stack on $`𝒵_𝐑`$ then $`X_𝐑`$ is a real geometric $`n`$-stack if and only if its complexification $`X_𝐂`$ is a geometric $`n`$-stack.
Proof: From the formula $`\mathrm{\Psi }_!(X)(Y)=X(\mathrm{\Psi }^!(Y))`$ we obtain immediately the compatibility with homotopy fiber products. This yields compatibility with the notion of geometricity (compatibility with the very presentable condition is tautological since we impose that condition on the complexified stack to begin with). To prove the last sentence, suppose $`X_𝐑`$ is a real stack and suppose $`YX_𝐂`$ is a smooth surjection. Then
$$Y\times _{X_𝐂}\overline{Y}X_𝐂$$
is a smooth surjection which admits a definition over $`𝐑`$ (the involution exchanges the two factors $`Y`$ and $`\overline{Y}`$). Using this we can prove that if $`X_𝐂`$ is geometric then so is $`X_𝐑`$. $`|||`$
Real structures on moduli stacks
Let $`nGEOM_𝐑`$ denote the stack of real geometric $`n`$-stacks over $`𝒵_𝐑`$, defined by setting $`nGEOM_𝐑(U)`$ equal to the $`n+1`$-category of real geometric $`n`$-stacks $`XU`$.
There is also a notion of real perfect (resp. hemiperfect) complex, which just means a perfect (resp. hemiperfect) complex of real vector bundles on the real scheme. This gives an $`n+1`$-stack $`nPERF_𝐑`$ (resp. $`nHPERF_𝐑`$) over $`𝒵_𝐑`$. A perfect complex on a real $`n`$-stack $`X_𝐑`$ is by definition a morphism $`X_𝐑nPERF_𝐑`$ (resp. $`X_𝐑nHPERF_𝐑`$).
###### Theorem 1
We have
$$(nGEOM_𝐑)_𝐂=nGEOM\text{and}(nHPERF_𝐑)_𝐂=nHPERF.$$
In other words, the complex moduli stacks $`nGEOM`$ and $`nHPERF`$ have canonical real structures.
Proof: Suppose $`Y`$ is a complex scheme. Then $`(nGEOM_𝐑)_𝐂(Y)`$ is by definition the $`n+1`$-category of real geometric $`n`$-stacks over $`\mathrm{\Psi }^!(Y)`$. These are thus stacks $`X`$ fitting into a diagram
$$X\mathrm{\Psi }^!Y\mathrm{\Psi }^!Spec(𝐂)Spec(𝐑).$$
Looking only at the first three terms of this diagram and applying the discussion above which gives an equivalence between $`n`$-stacks over $`𝒵`$ and $`n`$-stacks over $`𝒵_𝐑`$ provided with morphism to $`\mathrm{\Psi }^!Spec(𝐂)`$, we obtain that the data of $`X`$ is the same as the data of an object in $`nGEOM(Y)`$. This gives the equivalence in question for $`nGEOM`$. For $`nHPERF`$ we can deduce it directly by considering a hemiperfect complex as an $`n+N`$-stack which is $`N`$-connected, for $`Nn+1`$, whose $`\mathrm{\Omega }^N`$ is geometric. $`|||`$
Real structures and filtrations
Now we turn to filtrations. The stack $`𝒜:=𝐀^1/𝐆_m`$ is defined over $`𝐑`$, in other words we have a real $`n`$-stack
$$𝒜_𝐑:=𝐀_𝐑^1/𝐆_{m,𝐑}.$$
A real filtration of a real geometric $`n`$-stack $`V_𝐑`$, denoted $`(V_𝐑,W)`$, is a real geometric $`n`$-stack $`Tot^W(V_𝐑)`$ over $`𝒜_𝐑`$. The associated complex $`n`$-stack is thus a filtration of $`V_𝐂`$, we denote this associated filtration by $`(V_𝐂,W)`$, with its total space
$$Tot^W(V_𝐂):=Tot^W(V_𝐑)_𝐂.$$
The associated-graded $`Gr^W(V_𝐑)`$ is a real geometric $`n`$-stack with action of $`𝐆_{m,𝐑}`$. Again, this complexifies to $`Gr^W(V_𝐂)`$.
We take note of the fact that the canonical decomposition of $`𝐆_m`$-equivariant perfect complexes (see pages 7 or 8) is compatible with the real structure.
A weight-filtered $`n`$-stack with real structure $`(V,W)`$ means a collection of real stacks just as in the definition of page Nonabelian mixed Hodge structures. In this case, the $`\pi _iGr^W(V)H^i(LGr^W(V))`$ are real vector spaces with action of $`𝐆_{m,𝐑}`$. The components $`H^i(LGr_k^W(V))`$ are thus real subspaces. In the next part of the paper, this will be combined with a “Hodge filtration” which induces a complex filtration $`F`$ on the vector space $`H^i(LGr^W(V))_𝐂`$, i.e. an actual filtration by subspaces $`F^pH^i(LGr^W(V))_𝐂`$. It thus makes sense to take the complex conjugate of this filtration with respect to the real structure, giving a filtration by subspaces $`\overline{F}^qH^i(LGr^W(V))`$. This allows us to impose the condition that the Hodge filtration and its complex-conjugate be $`ki`$-opposed. This is the only place where the real structures which will be involved in the definitions to follow, are actually used.
Part II: The main definitions and conjectures
We now present the main definitions of pre-nonabelian mixed Hodge structure and nonabelian mixed Hodge structure, as well as a number of the basic conjectures which serve, we hope, to explain our motivation for making these definitions.
One small point to notice is that the notion of pre-namhs is not quite analogous to the notion of pre-weight-filtered $`n`$-stack which we discussed above. For pedagogical reasons, we included the notion of linearization in the notion of pre-weight-filtered $`n`$-stack. However, for practical reasons the notion of pre-namhs, without linearization, seems to be the most useful intermediate version of this definition.
Pre-nonabelian mixed Hodge structures
We can now give our first main definition. A pre-nonabelian mixed Hodge structure or pre-namhs is a collection
$$𝒱=\{(V_{DR},W,F);(V_{B,𝐑},W);\zeta _𝒱\}$$
where:
$`V_{DR}`$ is a geometric (hence by our convention, very presentable) $`n`$-stack;
$`(V_{DR},W,F)`$ is a geometric bifiltration of $`V_{DR}`$ with the filtrations parametrized by $`𝒜_{wt}\times 𝒜_{hod}`$ (but note that the weight filtration is not linearized);
$`V_{B,𝐑}`$ is a real geometric $`n`$-stack;
$`(V_{B,𝐑},W)`$ is a geometric filtration on $`V_{B,𝐑}`$; and
$`\zeta _𝒱`$ is an equivalence between the associated analytic filtered objects
$$\zeta _𝒱:(V_{DR},W)^{\mathrm{an}}(V_{B,𝐂},W)^{\mathrm{an}}.$$
We obtain an $`n+1`$-category of pre-namhs denoted $`nPNAMHS`$. Because of the presence of analytic morphisms in the definition, it is not immediately useful to consider an $`n+1`$-stack of these objects so we don’t. Constructing $`nPNAMHS`$ is straightforward following the definition of a pre-namhs. For example we have the following diagram of $`n+1`$-categories
$$\begin{array}{ccc}\hfill F.F.nGEOM()F.nGEOM()& & \\ & & \\ & & F.nGEOM^{[\mathrm{an}]}()\\ & & \\ \hfill F.nGEOM_𝐑()F.nGEOM()& & \end{array}$$
and $`nPNAMHS`$ is the limit or homotopy fiber product
$$nPNAMHS:=F.F.nGEOM()\times _{F.nGEOM^{\mathrm{an}}()}F.nGEOM_𝐑().$$
Here $`F.nGEOM^{[\mathrm{an}]}`$ is the analyic moduli $`n+1`$-stack of filtered analytic geometric $`n`$-stacks (which is different from the analytification of $`F.nGEOM`$), and $`F.nGEOM_𝐑`$ is the real moduli $`n+1`$-stack of filtered real geometric $`n`$-stacks. The top horizontal morphism is “forgetting the Hodge filtration” and the bottom horizontal morphism is complexification; the diagonal morphisms are analytification.
Concretely a $`1`$-morphism
$$\phi :\{(V_{DR},W,F),(V_{B,𝐑},W),\zeta _𝒱\}\{(V_{DR}^{},W,F),(V_{B,𝐑}^{},W),\zeta _𝒱^{}\}$$
in $`nPNAMHS`$ consists of a morphism of bifiltered objects,
$$\phi _{Hod}:(V_{DR},W,F)(V_{DR}^{},W,F),$$
a morphism of filtered real $`n`$-stacks
$$\phi _{B,𝐑}:(V_{B,𝐑},W)(V_{B,𝐑}^{},W),$$
and a homotopy or $`2`$-morphism (invertible up to equivalence) between
$$(\phi _{B,𝐂})^{\mathrm{an}}\zeta _𝒱\text{and}\zeta _𝒱^{}(\phi _{DR})^{\mathrm{an}}.$$
The $`n+1`$-categories entering into the fiber product defining $`nPNAMHS`$ admit finite limits, and the morphisms in the fiber product are compatible with these limits; therefore the fiber product $`nPNAMHS`$ also admits finite limits. In particular we can use homotopy fiber products here. This will not be the case with the notion of “namhs”, which explains the utility of isolating out a notion of “pre-namhs”.
Pre-mixed Hodge objects
Our definition of pre-namhs could be generalized in the following way. Suppose $`OBJ`$ is an $`n+1`$-stack (not of groupoids though) with real form $`OBJ_𝐑`$ and with a morphism $`OBJ^{\mathrm{an}}OBJ^{[\mathrm{an}]}`$ of analytic stacks. Then a pre-mixed Hodge object in $`OBJ`$ is a collection
$$\{(V_{DR},W,F),(V_{B,𝐑},W),\zeta \}$$
where $`V_{DR}:OBJ`$ is an object of type $`OBJ`$; where $`(W,F)`$ is a bifiltration of $`V_{DR}`$ i.e.
$$Tot^{W,F}(V_{DR}):𝒜_{wt}\times 𝒜_{hod}OBJ,$$
with fiber over $`[1]\times [1]`$ equal to $`V_{DR}`$; where $`V_{B,𝐑}:Spec(𝐑)OBJ_𝐑`$ is a real object, with filtration $`W`$; and where $`\zeta `$ is an equivalence in $`OBJ^{[\mathrm{an}]}`$ between $`Tot^W(V_{DR})`$ and the complexification of $`Tot^W(V_{B,𝐑})`$. We obtain an $`n+1`$-category $`PMH.OBJ`$ of pre-mixed Hodge objects in $`OBJ`$ (we don’t consider an $`n+1`$-stack of these objects).
A pre-namhs is just a pre-mixed Hodge object in $`nGEOM`$. Below we will see that a pre-mixed Hodge structure is an object of $`PMH.VECT`$; a pre-mixed Hodge complex is an object of $`PMH.nPERF`$; and we shall have occasion to meet objects of $`PMH.nHPERF`$, $`PMH.VESH`$, etc.
Split pre-namhs’s
Define a split pre-namhs to be a $`𝐆_m`$-equivariant pre-namhs such that the weight-filtrations arise from the construction of page Nonabelian mixed Hodge structures which to a $`𝐆_m`$-equivariant stack associates a $`𝐆_m`$-equivariant filtered stack. Denote by $`splPNAMHS`$ the $`n+1`$-category of these objects. The operation of “taking the associated-graded for the weight filtration” provides an $`n+1`$-functor
$$Gr^W:nPNAMHSsplPNAMHS.$$
Mixed Hodge complexes and linearization
Our next step is the notion of “linearization” for the weight filtrations. For this, we need to go back to the abelian yet higher cohomological case of “mixed Hodge complexes”. In order to establish notational conventions we first discuss the notion of mixed Hodge structures. A pre-mixed Hodge structure or pre-mhs (by which we always mean, in the current paper, pre-$`𝐑`$-mhs) is an object of $`PMH.VECT`$, in other words it is an uple
$$𝒱=\{(V,W,F);(V_𝐑,W),\zeta \}$$
where $`V`$ is a complex vector space with two filtrations $`F`$ and $`W`$, $`V_𝐑`$ is a real vector space with real filtration $`W`$, and $`\zeta `$ is an isomorphism of filtered objects between $`(V,W)`$ and the complexification of $`(V_𝐑,W)`$. In particular, we can think of $`(V,W,F)`$ as being an object of $`F.F.VECT`$ and $`(V_𝐑,W)`$ as an object of $`F.VECT_𝐑`$.
If $`𝒱`$ is a pre-mhs, we obtain the associated-graded $`Gr^W(𝒱)`$ which is a split pre-mhs, i.e. it splits up as a direct sum
$$Gr^W(𝒱)=Gr_k^W(𝒱)$$
of pieces on which the $`𝐆_m`$-action is by the character $`\lambda \lambda ^k`$. On these pieces, the weight filtrations are one-step (concentrated in degree $`k`$) and we have Hodge filtrations and real structures: we can denote these pieces by
$$\{(Gr_k^W(V),F),Gr_k^W(V)_𝐑\}.$$
In particular, we can take the complex conjugate of the Hodge filtration. To do this, use the old-fashioned approach to a filtration and take the complex conjugate of every subspace. According to our conventions (p. Nonabelian mixed Hodge structures), $`F`$ corresponds to an increasing filtration $`F_pGr_k^W(V)`$, but we can turn it into a decreasing filtration by the reindexation
$$F^pGr_k^W(V):=F_pGr_k^W(V).$$
The complex conjugate filtration $`\overline{F}`$ is defined by setting $`\overline{F}^pGr_k^W(V)`$ equal to the complex conjugate of $`F^pGr_k^W(V)`$ with respect to the real structure $`Gr_k^W(V)_𝐑`$. Recall from that one says that $`F`$ and $`\overline{F}`$ are $`\mu `$-opposed on $`Gr_k^W(V)`$ if
$$F^pGr_k^W(V)\overline{F}^qGr_k^W(V)=0,p+q>\mu ,$$
and
$$Gr_k^W(V)=\underset{p+q=\mu }{}F^pGr_k^W(V)\overline{F}^qGr_k^W(V).$$
Note that in this condition, we are using the upper-reindexed versions of $`F`$. In terms of the lower indexing, we get that $`F`$ and $`\overline{F}`$ are $`\mu `$-opposed on $`Gr_k^W(V)`$ if
$$F_pGr_k^W(V)\overline{F}_qGr_k^W(V)=0,\mu +p+q<0,$$
and
$$Gr_k^W(V)=\underset{p+q+\mu =0}{}F_pGr_k^W(V)\overline{F}_qGr_k^W(V).$$
Here is the place where the conventions are fixed! If $`s`$ is an integer, we say that a pre-mhs $`𝒱`$ is an $`s`$-shifted mhs if for every $`k`$, the filtrations $`F`$ and $`\overline{F}`$ are $`k+s`$-opposed on $`Gr_k^W(V)`$. We say that $`𝒱`$ is a mhs if it is a $`0`$-shifted mhs.
We now get to mixed Hodge complexes. Fix $`n0`$ which will be the length of the interval of support for the mixed Hodge complexes we are considering (we only consider ones of finite length). Define a pre-mixed Hodge complex or pre-mhc to be an uple
$$𝒞=\{(C_{DR},W,F),(C_{B,𝐑},W),\zeta _𝒞\}$$
where $`C_{DR}`$ is a perfect complex (i.e. a complex of finite dimensional $`𝐂`$-vector spaces); $`W`$ and $`F`$ form a bifiltration in the category of hemiperfect complexes (see page Nonabelian mixed Hodge structures), so on the whole $`(C_{DR},W,F)`$ is an object of $`F.F.nHPERF`$; similarly $`(C_{B,𝐑},W)`$ is a filtered real hemiperfect complex, and $`\zeta _𝒞`$ is an eqivalence between $`(C_{B,𝐂},W)`$ and $`(C_{DR},W)`$ as objects of $`F.nHPERF`$.
The use of hemiperfect complexes rather than perfect ones is a technical point which the first-time reader can safely ignore. In the definition of “mixed Hodge complex” (i.e. when we take off the prefix “pre-”) the hemiperfect complexes will be assumed to be perfect.
Notice here in particular that $`\zeta _𝒞`$ is an algebraic equivalence—we don’t speak of associated analytic objects, because in fact an analytic equivalence would automatically be algebraic, see p. 9.
Let $`PMHC`$ denote the $`n+1`$-category of pre-mixed Hodge complexes. This definition is almost equivalent to the object studied by Deligne in , the only difference is our use of hemiperfect rather than perfect complexes, which comes from the fact that we may want to consider truncations of Deligne-type pre-mhc’s.
We again have a notion of split pre-mhc, which means a $`𝐆_m`$-equivariant pre-mhc such that the $`𝐆_m`$-action splits the weight filtration. Let $`splPMHC`$ denote the $`n+1`$-category of these objects. As before, there is a functor $`𝒞Gr^W(𝒞)`$,
$$Gr^W:PMHCsplPMHC.$$
The canonical decomposition of $`𝐆_m`$-equivariant perfect complexes given by Theorem 7 (compatible with filtrations and real structures) gives a decomposition of split pre-mhc: if $`𝒞`$ is a split pre-mhc then it decomposes canonically as a direct sum of pre-mhc with trivial (but shifted) weight filtrations $`𝒞_k`$. If $`𝒞`$ is any pre-mhc then the associated-graded $`Gr^W(𝒞)`$, as a split-pre-mhc, decomposes as a direct sum which we write
$$Gr^W(𝒞)=\underset{k}{}Gr_k^W(𝒞).$$
Recall that the index $`k`$ means the piece on which $`𝐆_m`$ acts by the character $`\lambda \lambda ^k`$.
The Dold-Puppe functor from $`nHPERF`$ to $`nGEOM`$ induces in an obvious way functors
$$DP:PMHCPNAMHS$$
and
$$DP:splPMHCsplPNAMHS.$$
Linearization of pre-namhs’s
Using the previous notation we can explain what we mean by “linearization” of a pre-namhs. A linearized pre-namhs is a triple $`(𝒱,LGr^W(𝒱),\epsilon )`$ where $`𝒱`$ is a pre-namhs and $`LGr^W(𝒱)`$ is a split pre-mhc, and
$$\epsilon :DP(LGr^W(𝒱))\stackrel{}{}Gr^W(𝒱)$$
is an equivalence in $`splPNAMHS`$ of split pre-namhs. Concretely this means that for each weight-filtered object in $`𝒱`$, we have an linearized associated-graded object. For example, $`(LGr^W(V_{DR}),F)`$ is a $`𝐆_m`$-equivariant filtered hemiperfect complex whose Dold-Puppe is equivalent to $`(Gr^W(V_{DR}),F)`$ (via $`\epsilon `$), and similarly for the real Betti objects.
Define $`nLPNAMHS`$ to be the $`n+1`$-category of linearized pre-namhs’s; it may be defined as a fiber product of $`n+1`$-categories in an obvious way, with a diagram that we leave to the reader to draw.
The mixed Hodge complex conditions
We have a notion of mixed Hodge complex which (following ) is a pre-mhc which satisfies the following conditions:
—the hemiperfect complexes $`Tot^{W,F}(𝒞_{DR})`$ over $`𝒜_{wt}\times 𝒜_{hod}`$ and $`Tot^W(𝒞_{B,𝐑})`$ over $`𝒜_{wt,𝐑}`$ are actually perfect complexes;
—strictness of the Hodge filtration on $`Gr^W(𝒞)`$, a condition which we denote by $`\mathrm{𝐒𝐭𝐫}(Gr^W(𝒞),F)`$, which actually means that $`Tot^F(Gr^W(𝒞))`$ splits (non-canonically) as a product of its cohomology vector bundles over $`𝒜_{hod}`$ (cf Lemma 5); and
—the purity condition $`\mathrm{𝐌𝐇𝐂}(Gr^W(𝒞))`$ which says that on the $`k`$-th graded piece for the $`𝐆_m`$-action
$$H^i(Gr_k^W(𝒞))H^i(Gr^W(𝒞))$$
the Hodge filtration $`F`$ and its complex conjugate $`\overline{F}`$ (with respect to the real structure on the vector space in question), are $`ki`$-opposed. In other words, this condition says that the split pre-mhs’s $`H^iGr^W(𝒞)`$ are $`i`$-shifted mhs’s. This shifting depending on the cohomological degree, is the fundamental point behind Deligne’s definition . See also a recent discussion in Saito . We give a heuristic justification for the sign in the next subsection below.
If one has the strictness condition, then the $`H^iTot^F(Gr^W(𝒞))`$ are vector bundles over $`𝒜_{hod}`$, thus
$$(H^iGr_k^W(𝒞),F)F.VECT$$
are filtered vector spaces in the sense of the discussion of page Nonabelian mixed Hodge structures.
We obtain an $`n+1`$-category $`MHC`$ of mixed Hodge complexes.
More generally, we say that a pre-mhc $`𝒞`$ is an $`s`$-shifted mhc if it satisfies the strictness condition and if the $`H^iGr^W(𝒞)`$ are $`si`$-shifted mhs’s.
The sign in Deligne’s shift
Getting the signs right in the condition $`\mathrm{𝐌𝐇𝐂}`$ is a nontrivial task. Recall from the introduction that there are three numbers involved: the Hodge degree which is the integer $`h`$ such that the Hodge filtration and its complex conjugate are $`h`$-opposed; the weight $`w`$ which is the degree in the weight filtration, equal to $`k`$ on the graded piece $`Gr_k^W(V)`$; and the homotopical degree $`i`$ which indicates that we are looking at $`\pi _i`$ or $`H^i`$. Deligne’s shift in the condition $`\mathrm{𝐌𝐇𝐂}`$ is the linear relationship
$$h=wi.$$
We give a heuristic argument here which shows how to determine the sign in this formula.
We know of course that the sign of $`h`$ should be the same as that of $`w`$; the question is whether to put $`+`$ or $``$ in front of $`i`$. In order to resolve this, note that for a coefficient stack $`V`$ (such as an Eilenberg-MacLane stack $`V=K(U,m)`$) we have that
$$\pi _0\underset{¯}{Hom}(X_M,V)$$
contains (in a vague way) $`H^j(X_M,\pi _j(V))`$. Thus we would like to require that
$$h(\pi _0\underset{¯}{Hom}(X_M,V))=h(H^j(X_M,\pi _j(V)))$$
and
$$w(\pi _0\underset{¯}{Hom}(X_M,V))=w(H^j(X_M,\pi _j(V))).$$
On the other hand, we will be proposing that $`X_M`$ be constant in the weight-filtration direction, so
$$w(H^j(X_M,\pi _j(V)))=w(\pi _j(V)).$$
Taking the $`j`$-th cohomology of $`X_M`$ has the effect of adding $`j`$ to the Hodge degree:
$$h(H^j(X_M,\pi _j(V)))=h(\pi _j(V))+j.$$
Finally, note that $`i=0`$ for the $`\pi _0`$, so whatever the sign is, we will have
$$h(\pi _0\underset{¯}{Hom}(X_M,V))=w(\pi _0\underset{¯}{Hom}(X_M,V)).$$
Putting all of these together we obtain the formula:
$$h(\pi _j(V))+j=w(\pi _j(V))$$
or
$$h(\pi _j(V))=w(\pi _j(V))j.$$
As $`j`$ is the homotopical degree of the piece $`\pi _j`$, this justifies the sign in the formula for condition $`\mathrm{𝐌𝐇𝐂}`$.
Application to linearized pre-namhs’s
If $`(𝒱,LGr^W(𝒱),\epsilon )`$ is an linearized pre-namhs then we say that it satisfies the strictness condition $`\mathrm{𝐒𝐭𝐫}(LGr^W(𝒱),F)`$ if the split pre-mhc $`LGr^W(𝒱)`$ satisfies the strictness condition (and if the cohomology sheaves are vector bundles, in particular this condition implies that the hemiperfect complexes underlying $`LGr^W(𝒱)`$ are actually perfect complexes); and we say that it satisfies the mixed Hodge complex condition denoted $`\mathrm{𝐌𝐇𝐂}(LGr^W(𝒱))`$ if the split pre-mhc $`LGr^W(𝒱)`$ satisfies the “purity condition” described above.
If these two conditions are satisfied, then the $`H^i(LGr^W(𝒱))`$ are split mixed Hodge structures with weight shifted by $`i`$. In particular, for $`i=0`$ there is no shift and $`H^0(LGr^W(𝒱))`$ is a split mixed Hodge structure.
Truncated mixed Hodge complexes
While the Hodge filtration is strict on a mixed Hodge complex (on the $`Gr^W`$ this is an axiom, and on the whole mhc it is a result of Scholie 8.1.9), the weight filtration is not in general strict. This is seen by the fact that the spectral sequence for $`W`$ degenerates at $`E_2`$ but not necessarily at $`E_1`$ (the objects occuring in Proposition 14 p. 14) provide an example of this behavior). A consequence of this is that a shift and truncation of a mixed Hodge complex is no longer a mixed Hodge complex. We will see what type of structure it does have, and this provides motivation for the role of the annihilator ideals in the definition of namhs.
The definition of pre-mhc was made using the notion of hemiperfect complex; recall that this notion is stable under the truncated shift operation $`\mathrm{\Omega }`$; thus the notion of pre-mhc is stable under the truncated shift operation. The conditions $`\mathrm{𝐒𝐭𝐫}`$ and $`\mathrm{𝐌𝐇𝐂}`$ are also stable under the truncated shift operation. In particular, it follows that the associated-graded $`(Gr^W(𝒞),F)`$ is a filtered object in $`nPERF`$. Thus the only filtrations that occur in $`nHPERF`$ rather than in $`nPERF`$ are the weight filtrations.
A truncated mixed Hodge complex is the same type of thing as a mixed Hodge complex but without the condition that the underlying complexes be perfect (thus they are only hemiperfect cf p. Nonabelian mixed Hodge structures), and with the following additional requirements. First, $`Gr^W`$ is required to be an actual perfect complex and the Hodge filtration here is required to be strict. Then we also require: $`\mathrm{𝐀𝟏}`$ that the annihilator ideals are all equal,
$$m1,Ann^W(t^m;𝒞)=Ann^W(t;𝒞);$$
and $`\mathrm{𝐀𝟑}`$ that the linear subspace defined by this annihilator ideal be a sub-mixed Hodge structure of $`H^0(Gr^W(𝒞))`$.
###### Corollary 11
Suppose $`𝒟`$ is a $`k`$-shifted mixed Hodge complex of length $`n+k`$ and let $`𝒞=\mathrm{\Omega }^k𝒟`$. Then $`𝒞`$ is a truncated mixed Hodge complex.
Proof: Recall that the terms
$$H^i(Gr_l^W(𝒞))=H^{ki}(Gr_l^W(𝒟))$$
are pure Hodge structures with weight $`lik+k=li`$, the first $`ik`$ coming from the shift in the definition of condition $`\mathrm{𝐌𝐇𝐂}`$ and the second $`k`$ coming from the fact that we supposed $`𝒟`$ was $`k`$-shifted. Thus we get that $`𝒞`$ satisfies condition $`\mathrm{𝐌𝐇𝐂}`$. The sheaves entering into condition $`\mathrm{𝐒𝐭𝐫}`$ are the same as those for $`𝒟`$ so $`𝒞`$ satisfies condition $`\mathrm{𝐒𝐭𝐫}`$.
Recall from that the spectral sequence of the weight filtration on a mixed Hodge complex such as $`𝒟`$, degenerates at $`E_2`$. Therefore by Lemma 6 the truncated and shifted complex $`𝒞`$ satisfies condition $`\mathrm{𝐀𝟏}`$. For condition $`\mathrm{𝐀𝟑}`$ the subspace in question is the direct sum over weights $`l`$ of the kernel of the differential
$$d:H^k(Gr_l^W(𝒟))H^{1k}(Gr_{l1}^W(𝒟)).$$
This differential is a morphism of $`l`$-shifted pure Hodge structures, so its kernel is again an $`l`$-shifted pure Hodge structure. This provides condition $`\mathrm{𝐀𝟑}`$ and completes the proof. $`|||`$
Given a truncated mixed Hodge complex as defined above, we leave as an exercise to show that the operation of taking the flat subobject of the total spaces of the weight filtrations, gives canonically and functorially a mixed Hodge complex. Combining this exercise with the previous corollary, one obtains the result of Hain and Zucker showing how to truncate a mixed Hodge complex into another mixed Hodge complex. Certainly the key conditions $`\mathrm{𝐀𝟏}`$ and $`\mathrm{𝐀𝟑}`$ must appear in disguised form in the argument of .
Nonabelian mixed Hodge structures
A nonabelian mixed Hodge structure or namhs is a linearized pre-namhs (see page 1) which satisfies a number of conditions which we shall now explain. Before getting to the conditions, we say that the $`n+1`$-category of namhs denoted $`nNAMHS`$, is defined as the full sub-$`n+1`$-category of $`nLPNAMHS`$ (see page 1) consisting of those objects which satisfy the following conditions. In particular, a morphism of namhs is defined as a morphism of linearized pre-namhs.
Conditions: These conditions are divided into parts (I), (II) and (III).
(I)
We impose the conditions that we discussed above, namely:
$`\mathrm{𝐀𝟏}(V_{DR},W,F)`$ (resp. $`\mathrm{𝐀𝟏}(V_{B,𝐑},W)`$), that the annihilator ideals
$`Ann^W(t^m;Tot^F(V_{DR}))`$ (resp. $`Ann^W(t^m;V_{B,𝐑})`$) are all equal for $`m1`$;
$`\mathrm{𝐀𝟐}(V_{DR},W,F)`$ (resp. $`\mathrm{𝐀𝟐}(V_{B,𝐑},W,F)`$), that the annihilator ideal
$`Ann^W(t;Tot^F(V_{DR}))`$ (resp. $`Ann^W(t;V_{B,𝐑})`$) is contained in the ideal of the zero section of $`Gr^W(Tot^F(V_{DR}))`$ (resp. $`Gr^W(V_{B,𝐑})`$);
$`\mathrm{𝐅𝐥}(V_{DR},W,F)`$ that the Hodge filtration $`F`$ is flat.
(II)
These are the mixed Hodge complex conditions. First we ask that:
$`\mathrm{𝐒𝐭𝐫}(LGr^W(V_{DR}),F)`$ the sheaves
$$H^i(Tot^F(LGr^W(V_{DR}))/𝒜_{hod})\pi _i(Tot^FGr^W(V_{DR})/𝒜)$$
be vector bundles over $`𝒜_{hod}`$. In particular, by Lemma 5, the linearization of $`(Tot^F(V_{DR}),W)`$ is a perfect one cf p. Nonabelian mixed Hodge structures.
Now the object
$$LGr^W(𝒱)=\{(LGr^W(V_{DR}),F),LGr^W(V_{B,𝐑}),LGr^W(\zeta _𝒱)\}$$
is basically a graded complex with filtration $`F`$ and real structure. We can turn this into a complex with a real structure, a real filtration $`W`$ and a complex filtration $`F`$, by setting $`W`$ equal to the filtration associated to the grading, see p. Nonabelian mixed Hodge structures.
$`\mathrm{𝐌𝐇𝐂}(LGr^W(𝒱))`$ We ask that the resulting object be a mixed Hodge complex , see also . Note that this includes the strictness condition $`\mathrm{𝐒𝐭𝐫}`$ above, plus a “purity” condition. By the strictness condition, the cohomology objects $`H^i(LGr^W(𝒱))`$ are vector spaces provided with a Hodge filtration in the stack $`VECT`$, which just means filtered vector spaces in the usual sense. These vector spaces also have a grading by the $`𝐆_m`$-action and a real structure, and the purity condition says that on the $`k`$-th graded piece $`H^i(LGr_k^W(𝒱))`$ of $`H^i(LGr^W(𝒱))`$, the Hodge filtration $`F`$ and its complex-conjugate $`\overline{F}`$ (taken with respect to the real structure), are $`m`$-opposed with
$$m=ki.$$
This shift by the homotopical degree $`i`$ originates with Deligne . In our notation, see the subsection on p. 1 for an explanation of the sign.
In particular the above condition means that the vector space
$$𝒫:=H^0(LGr^W(𝒱))$$
is a split mixed Hodge structure. To fix notations for below, we have
$$𝒫=\{(P_{DR}:=H^0LGr^W(V_{DR}),F),P_{B,𝐑}:=H^0LGr^W(V_{B,𝐑}),\zeta _P\}$$
together with an action of $`𝐆_m`$ on this collection of stuff. Here note that $`\zeta _P`$ is actually an algebraic isomorphism of vector spaces between $`P_{DR}`$ and $`P_{B,𝐂}=P_{B,𝐑}_𝐑𝐂`$. Thus $`𝒫`$ consists of a graded real vector space with a complex filtration $`F`$. Turning the grading back into a weight filtration $`W`$ we obtain a pre-mixed Hodge structure, and one of the consequences of conditions $`\mathrm{𝐒𝐭𝐫}`$ and $`\mathrm{𝐌𝐇𝐂}`$ is that this object $`𝒫`$ is a mixed Hodge structure.
(II)
We finally get to our third condition. Recall that $`Ann^W(t,𝒱)`$ is a sheaf of ideals on $`\pi _0(Gr^W(𝒱))=𝒫`$. The isomorphism $`\zeta _P`$ transforms the de Rham version $`Ann^W(t,V_{DR})`$ to the complexified Betti version $`Ann^W(t,V_{B,𝐂})`$, so we don’t need to distinguish between these. Our sheaf of ideals has a Hodge filtration, a real structure, and it is compatible with the $`𝐆_m`$-action on $`𝒫`$. The $`𝐆_m`$-action may be transformed into a weight filtration on $`𝒫`$. Thus
$$Ann^W(t;𝒱)Sym^{}(𝒫^{})$$
is a sub-pre-mixed Hodge structure (i.e. subobject with two filtrations and a real structure preserving the weight filtration). The condition of part (III) is stated as our third condition on the annihilator ideals:
$`\mathrm{𝐀𝟑}(𝒱)`$, that $`Ann^W(t;𝒱)Sym^{}(𝒫^{})`$ be a sub-mixed Hodge structure.
This completes our definition of “nonabelian mixed Hodge structure”.
Remark: Conditions $`\mathrm{𝐒𝐭𝐫}`$ and $`\mathrm{𝐌𝐇𝐂}`$ are conditions on $`LGr^W(V)`$, in other words they are conditions only on the “abelian” part; conditions $`\mathrm{𝐀𝟏}`$, $`\mathrm{𝐀𝟐}`$, $`\mathrm{𝐀𝟑}`$ and $`\mathrm{𝐅𝐥}`$ concern the nonabelian part.
Homotopy group sheaves
A property which we would clearly like to have is that the homotopy group sheaves of a nonabelian mixed Hodge structure, should carry mixed Hodge structures in a natural way. This turns out to be less obvious than one might think, and requires a somewhat technical treatment. We will give a sketch here. The more technical parts of this section are optional and are not used in the computations in Part III. On the other hand, the first set of remarks—including Proposition 12 and up to the section on Whitehead products p. Nonabelian mixed Hodge structures—should be illuminating for understanding what is going on in Part III, so we suggest to read this section lightly the first time.
There is probably a close correspondence between $`1`$-connected nonabelian mixed Hodge structures and the type of “mixed Hodge dga’s” considered by Morgan and Hain . In this sense, everything that we say in the present section should probably be considered as already known by and ; and one should consider that our work in this section consists of transcribing those constructions into the context of our present definitions of nonabelian mixed Hodge structure using $`n`$-stacks.
Throughout this section, $`𝒱`$ will denote a nonabelian mixed Hodge structure which is simply connected, i.e. the $`n`$-stacks involved are $`1`$-connected relative to their base stacks $`𝒜`$ or $`𝒜\times 𝒜`$. In particular, this implies that $`𝒱`$ is smooth so the conditions on the annihilator ideals and the flatness assumption on the Hodge filtration are automatic and we don’t use them. It is clear that eventually one would like to treat the case of a non-simply connected $`𝒱`$ but this is left for a future paper.
Locally over their bases, all of the $`n`$-stacks involved in the definition of $`𝒱`$ have basepoints, and the basepoints are unique up to locally-defined homotopy which is locally unique up to a $`2`$-homotopy which itself is not unique; all of this is because of our hypothesis that $`𝒱`$ is simply connected. In particular there is no need to specify a basepoint section when speaking of the homotopy group sheaves $`\pi _i(𝒱)`$.
The vector bundle case
The first situation we consider (and this is the one which is encountered in our construction of an example in Part III) is the case when the homotopy group sheaves $`\pi _i(𝒱)`$ are vector bundles. To explain this condition recall that $`𝒱`$ consists essentially of the data of geometric $`n`$-stacks
$$Tot^{W,F}(V_{DR})𝒜_{wt}\times 𝒜_{hod},Tot^W(V_{B,𝐑})𝒜_{wt,𝐑}$$
together with an analytic equivalence and a linearization. We define $`\pi _i(𝒱)`$ as being the object consisting of sheaves in the big site
$$Tot^{W,F}(\pi _i(𝒱)_{DR}):=\pi _i(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})\text{over}𝒜_{wt}\times 𝒜_{hod}$$
and
$$Tot^W(\pi _i(𝒱)_{B,𝐑}):=\pi _i(Tot^W(V_{B,𝐑})/𝒜_{wt,𝐑})\text{over}𝒜_{wt,𝐑}$$
together with an (analytic but which is automatically algebraic) equivalence between $`Tot^W(\pi _i(𝒱)_{DR})`$ and $`Tot^W(\pi _i(𝒱)_{B,𝐂})`$. In general $`\pi _i(𝒱)`$ is a pre-mixed Hodge object (p. 1) in the stack $`VESH`$ of vector sheaves, see below. However, our hypothesis in this subsection is that the objects $`Tot^{W,F}(\pi _i(𝒱)_{DR})`$ and $`Tot^W(\pi _i(𝒱)_{B,𝐑})`$ are vector bundles over their respective base objects.
With this hypothesis, $`(\pi _i(𝒱)_{DR},W,F)`$ becomes an object of $`F.F.VECT`$ (i.e. just a bifiltered vector space) and $`(\pi _i(𝒱)_{B,𝐑},W)`$ becomes an object of $`F.VECT_𝐑`$ (i.e. a filtered real vector space); and the equivalence is just
$$(\pi _i(𝒱)_{DR},W)(\pi _i(𝒱)_{B,𝐂},W).$$
All in all this means that $`\pi _i(𝒱)`$ is a pre-mixed Hodge structure in the classical sense, i.e. an object of $`PMH.VECT`$.
A first version of our main result of this section is the following proposition.
###### Proposition 12
If $`𝒱`$ is a simply connected namhs such that the sheaves underlying the object $`\pi _i(𝒱)`$ as defined above are vector bundles, then this pre-mixed Hodge structure is a $`i`$-shifted mixed Hodge structure.
Proof: Note that
$$\pi _iGr^W(𝒱)=Gr^W(\pi _i(𝒱)),$$
indeed these two objects are obtained by restricting all of the sheaves involved, to $`[0]𝒜_{wt}`$. Note also that $`Gr^W(\pi _i(𝒱))`$ is the actual associated-graded of the pre-mixed Hodge structure $`\pi _i(𝒱)`$, in other words it is a vector space with real structure and Hodge filtration obtained by taking the associated-graded of the vector space $`\pi _i(𝒱)`$ with respect to its weight-filtration by subspaces. The indexation of the components determined by the $`𝐆_m`$-action is compatible with the usual one, i.e.
$$\pi _iGr_k^W(𝒱)=Gr_k^W(\pi _i(𝒱))=W_k\pi _i(𝒱)/W_{k1}\pi _i(𝒱)$$
(see p. Nonabelian mixed Hodge structures).
The hypothesis that $`𝒱`$ is a namhs means that it is linearized and that the pre-mhc $`LGr^W(𝒱)`$ a mixed Hodge complex. In particular, this means that the Hodge filtration and its complex conjugate are $`ki`$-opposed on
$$H^i(LGr_k^W(𝒱))=\pi _i(Gr_k^W(𝒱))=Gr_k^W(\pi _i(𝒱)).$$
But this is exactly the definition of $`\pi _i(𝒱)`$ being a $`i`$-shifted mixed Hodge structure. $`|||`$
In the context of these arguments we can briefly say what complicates the general case: in general the $`\pi _i(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})`$ are just vector sheaves rather than vector bundles over the base $`𝒜_{wt}\times 𝒜_{hod}`$. Thus while they yield actual filtrations of the underlying vector space $`\pi _i(V_{DR})`$ by subspaces (cf p. Nonabelian mixed Hodge structures), one doesn’t recover the total vector sheaves from these filtrations and in particular the fiber of say $`\pi _i(Tot^W(V_{DR})/𝒜_{wt})`$ over $`[0]`$ does not in general correspond to the associated-graded for the filtration by subspaces. The simple argument given for Proposition 12 no longer works.
Normalized homotopy objects
It is convenient to normalize the pre-mhs’s $`\pi _i(𝒱)`$ which are $`i`$-shifted mhs’s, to turn them into actual mixed Hodge structures. To do this, re-index the weight filtration shifting the indexation by $`i`$. Call the resulting mixed Hodge structures the normalized homotopy groups of $`𝒱`$, denoted $`\pi _i^\nu (𝒱)`$.
Whitehead products
Before getting on to the treatment of the general problem raised in the preceding paragraph, we discuss the “Whitehead products” in the easy case represented by Proposition 12.
If $`(X,x)`$ is a pointed space then we have the functorial Whitehead product which is a bilinear map
$$Wh_{i,j}:\pi _i(X,x)\times \pi _j(X,x)\pi _{i+j1}(X,x).$$
The same holds if $`X`$ is an $`n`$-groupoid by the equivalence of . Suppose now that $`X`$ is an $`n`$-stack over a base $`B`$ (which could be a scheme, a stack or an $`n`$-stack itself). Suppose $`x:BX`$ is a basepoint section. Then we obtain a bilinear map of sheaves over $`B`$
$$Wh_{i,j}:\pi _i(X/B,x)\times \pi _j(X/B,x)\pi _{i+j1}(X/B,x).$$
If $`X`$ is relatively $`1`$-connected over $`B`$ then locally over $`B`$ there is a basepoint section and the homotopy groups don’t depend on choice of basepoint section up to unique isomorphism, so we obtain relative $`\pi _i(X/B)`$ which are sheaves over $`B`$ (which if necessary means morphisms from $`B`$ to the $`1`$-stack $`SH`$ of sheaves). These exist even though there need not exist a global section $`x:BX`$. In this case the Whitehead product maps again give bilinear maps of sheaves over $`B`$,
$$Wh_{i,j}:\pi _i(X/B)\times \pi _j(X/B)\pi _{i+j1}(X/B).$$
Now we return to the case of a $`1`$-connected namhs $`𝒱`$. We obtain bilinear maps of the sheaves occuring in the definitions of the $`\pi _i(𝒱)`$, for example
$$Wh_{i,j}:\pi _i(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})\times \pi _j(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})$$
$$\pi _{i+j1}(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod}).$$
Idem for $`V_{B,𝐑}`$, and these maps are compatible with the equivalences $`\zeta `$.
In the case we are currently considering, the terms in the above bilinear map are assumed to be vector bundles over $`𝒜_{wt}\times 𝒜_{hod}`$. Thus the bilinear map becomes a morphism from the tensor product of the two bundles, to the third one. The correspondence between filtered vector spaces in the usual sense and bundles over $`𝒜`$ is compatible with tensor product. Taking the tensor product of all of the filtrations in question, we obtain a notion of tensor product of pre-mhs. The Whitehead product becomes a morphism of pre-mhs
$$Wh_{i,j}^{\mathrm{pre}}(𝒱):\pi _i(𝒱)\pi _j(𝒱)\pi _{i+j1}(𝒱).$$
The alert reader will have noticed that there is a problem with the shifts here: on the left we have tensored a $`i`$-shifted mhs with a $`j`$-shifted mhs, which gives a $`ij`$-shifted mhs. On the right is a $`1ij`$-shifted mhs.
Here is where our main idea, of requiring that the associated-graded be a linear object, comes in. The object $`Gr^W(𝒱)`$ is linear, so its Whitehead products vanish. In other words, the associated-graded of the above Whitehead product for $`𝒱`$ vanishes:
$$Gr^W(Wh_{i,j}^{\mathrm{pre}}(𝒱))=0.$$
In other words, $`Wh_{i,j}^{\mathrm{pre}}(𝒱)`$ lowers the degree in the weight filtration by one. Denote by $`\pi _{i+j1}(𝒱)(D_{wt})`$ the pre-mhs obtained by shifting the weight filtration by one, in the direction such that there is a morphism of pre-mhs inducing zero on the associated-graded,
$$\pi _{i+j1}(𝒱)(D_{wt})\pi _{i+j1}(𝒱).$$
Then $`\pi _{i+j1}(𝒱)(D_{wt})`$ is a $`ij`$-shifted mhs, and the Whitehead product factors through a morphism
$$Wh_{i,j}(𝒱):\pi _i(𝒱)\pi _j(𝒱)\pi _{i+j1}(𝒱)(D_{wt}).$$
This is now a morphism of $`ij`$-shifted mixed Hodge structures.
If we normalize the homotopy group sheaves, and include an extra shift by one to normalize $`\pi _{i+j1}(𝒱)(D_{wt})`$ (which gives the same answer as if we normalize $`\pi _{i+j1}(𝒱)`$), then we have ended up shifting by $`i+j`$ on both sides of the map $`Wh_{i,j}(𝒱)`$, so it normalizes to a morphism of mixed Hodge structures (the Whitehead product):
$$Wh_{i,j}^\nu (𝒱):\pi _i^\nu (𝒱)\pi _j^\nu (𝒱)\pi _{i+j1}^\nu (𝒱).$$
The fact that the Whitehead product goes from degree $`i`$ times degree $`j`$ to degree $`i+j1`$ means that in order to get a correctly shifted morphism of mixed Hodge structures, we need to pick up a shift by $`1`$ somewhere. This shift by $`1`$ comes from the fact that by assumption the Whitehead products vanish on $`Gr^W(𝒱)=DP(LGr^W(𝒱))`$. This situation may inversely be taken as one of the motivations for the idea of requiring $`Gr^W(𝒱)`$ to be a linear object.
The general case
We now turn to the general case where we don’t make the hypothesis that the sheaves underlying $`\pi _i(𝒱)`$ are vector bundles. Recall from that if $`XB`$ is a relatively $`1`$-connected geometric $`n`$-stack over a base $`B`$ then the $`\pi _i(X/B)`$ are vector sheaves over $`B`$ (which if necessary means morphisms $`BVESH`$ to the $`1`$-stack of vector sheaves). Recall also from pages Nonabelian mixed Hodge structures and Nonabelian mixed Hodge structures that if $`U`$ is a vector sheaf over $`𝒜`$ then we obtain a filtration $`W_{}`$ of the vector space $`U_{[1]}`$ by subspaces defined by the condition that $`uW_k`$ if and only if $`t^k\stackrel{~}{u}`$ extends to a section of $`U`$ over $`𝐀^1`$. This functor from vector sheaves over $`𝒜`$ to filtered vector spaces is unfortunately not an equivalence of categories, and the fiber $`U_{[0]}`$ is no longer identifiable with the usual associated-graded of the filtration (i.e. the direct sum of the subspaces $`W_k/W_{k1}`$). This is what makes the present situation more complicated.
We continue to assume that $`𝒱`$ is a nonabelian mixed Hodge structure which is $`1`$-connected. Thus the sheaves underlying $`\pi _i(𝒱)`$ are vector sheaves over the appropriate base stacks $`𝒜_{wt}\times 𝒜_{hod}`$ or $`𝒜_{wt,𝐑}`$. We obtain in particular, vector sheaves
$$Tot^F(\pi _i(V_{DR}))\text{over}𝒜_{hod},$$
and
$$Tot^W(\pi _i(V_{B,𝐑}))\text{over}𝒜_{hwt,𝐑}.$$
The fibers of these over $`[1]`$ are respectively the finite dimensional complex vector space $`\pi _i(V_{DR})`$ and the finite dimensional real vector space $`\pi _i(V_{B,𝐑})`$; and the former is isomorphic via $`\zeta `$ to the complexification of the latter. Let $`\pi _i(V)`$ denote these isomorphic spaces, together with its real structure. From the above vector sheaves we obtain filtrations $`F^{}`$ (complex) and $`W_{}`$ (real) of $`\pi _i(V)`$ by subspaces (again see pages Nonabelian mixed Hodge structures and Nonabelian mixed Hodge structures).
This yields a pre-mixed Hodge structure
$$\pi _i^\mu (𝒱):=\{\pi _i(V),\pi _i(V)_𝐑,F^{},W_{}\}.$$
Our goal is to obtain the following result.
###### Theorem 2
The pre-mixed Hodge structure $`\pi _i^\mu (𝒱)`$ which we have defined above is a $`i`$-shifted mhs. Denote by $`\pi _i^\nu (𝒱)`$ its version where the weight filtration is shifted by $`i`$ places so that it becomes a mixed Hodge structure. The Whitehead products on the homotopy groups of $`V_{DR}`$ give morphisms of mixed Hodge structures
$$Wh_{i,j}^\nu :\pi _i^\nu (𝒱)\pi _j^\nu (𝒱)\pi _{i+j1}^\nu (𝒱).$$
Along the way we will also obtain a somewhat more precise structure theorem about the vector sheaves underlying the original object $`\pi _i(𝒱)`$.
The case of mixed Hodge complexes
Suppose $`𝒞`$ is a mixed Hodge complex. Its Dold-Puppe $`𝒱=DP(𝒞)`$ is a namhs, and the homotopy groups of the latter are the same as the cohomology groups of $`𝒞`$. Taking the cohomology sheaves of all of the perfect complexes underlying $`𝒞`$ we obtain a “pre-mixed Hodge vector sheaf” $`H^i(𝒞)`$ and the Dold-Puppe compatibility says that this object is equivalent to the object $`\pi _i(𝒱)`$ considered above. In particular, we encounter the same problem as that which was outlined above, for the $`H^i(𝒞)`$. On the other hand, Deligne shows in that the cohomology groups of a mixed Hodge complex are (shifted) mixed Hodge structures. Our strategy for proving Theorem 2 will be to reduce to the case of a mixed Hodge complex. For this, we need a structure theorem which says that in a certain sense one can “split off” the behavior at the rightmost part of the mixed Hodge complex (which corresponds to the lowest homotopy degrees). From there, we will be able to use the Hurewicz fact that homotopy and homology coincide in low degrees to apply the case of mixed Hodge complexes to the general case.
The following two propositions and the resulting theorem represent our “splitting-off” results. As they are results about mixed Hodge complexes, we leave their proofs as exercises for the reader, to be done using what is known e.g. in Deligne , Hain and Zucker , Saito etc.
###### Proposition 13
Suppose $`𝒞`$ is a mixed Hodge complex which is supported in $`(\mathrm{},k]`$. Let $`C`$ denote the underlying complex (i.e. one of the two isomorphic complexes $`C_{DR}`$ or $`C_{B,𝐂}`$). If $`H^k(C)0`$ then $`H^k(C)`$ provided with its Hodge and weight filtrations and real structure, is a $`k`$-shifted mixed Hodge structure. Furthermore, if we let $`H_\mu ^k(𝒞)`$ denote this shifted mixed Hodge structure considered as a mixed Hodge complex sitting in degree $`k`$, then there is a morphism of mixed Hodge complexes inducing the identity on $`H^k(C)`$:
$$𝒞H_\mu ^0(𝒞).$$
$`|||`$
###### Proposition 14
Suppose $`𝒞`$ is a mixed Hodge complex which is supported in $`(\mathrm{},k]`$. Let $`C`$ denote the underlying complex (i.e. one of the two isomorphic complexes $`C_{DR}`$ or $`C_{B,𝐂}`$). Suppose that $`H^k(C)=0`$. Then there is a split $`k`$-shifted mixed Hodge structure $`K=H^k(Gr^W(𝒞))`$, and a mixed Hodge complex $`𝒦`$ supported in $`[k1,k]`$ such that $`H^k(Gr^W(𝒦))=K`$, such that $`H^{1k}(Gr^W(𝒦))=K^{}`$ is the split $`1k`$-shifted mixed Hodge structure obtained by shifting the weight filtration of $`K`$ by one, and such that the complex underlying $`𝒦`$ is exact. The differential of $`𝒦`$ is an isomorphism of pre-mixed Hodge complexes but which shifts the weight filtration by one. There is a morphism of mixed Hodge complexes
$$f:𝒞𝒦$$
which induces an isomorphism $`H^k(Gr^W(𝒞))H^k(Gr^W(𝒦))`$. Furthermore, $`Cone(f)`$ is quasi-isomorphic to a mixed Hodge complex supported in $`[\mathrm{},1k]`$.
$`|||`$
We show how to use these two propositions to understand the structure of $`H^i(𝒞)`$. Suppose $`𝒞`$ is supported in $`(\mathrm{},k]`$. If $`H^k(C)0`$ then we apply Proposition 13 to obtain a mixed Hodge complex $``$ supported in degree $`k`$ and a morphism
$$g:𝒞.$$
Let $`𝒞_1:=Cone(g)`$, and let $`C_1`$ denote its underlying complex. (If $`H^k(C)=0`$ then set $`𝒞_1=𝒞`$ and pass directly to the next step.) Think of the cone as a kernel of a map, so it is supported in $`(\mathrm{},k]`$. By the property obtained in Proposition 13, $`H^k(C_1)=0`$. Therefore we can apply Proposition 14 to obtain a mixed Hodge complex $`𝒦`$ supported in $`[1k,k]`$ and a morphism
$$f:𝒞_1𝒦,$$
such that $`𝒞_2:=Cone(f)`$ is quasiisomorphic to a mixed Hodge complex supported in $`(\mathrm{},1k]`$. Thus in two steps we have split off two types of mixed Hodge complexes: the first one equal to a mixed Hodge structure supported in a single degree; and the second one supported in two adjacent degrees and whose underlying complex is exact. After doing this, we move the rightmost bound for the interval of support, back by one.
We obtain the following structure theorem for the objects $`H^i(𝒞)`$.
###### Theorem 3
Suppose $`𝒞`$ is a mixed Hodge complex supported in $`(\mathrm{},0]`$. Let $`H^i(𝒞)`$ denote cohomology of $`𝒞`$ considered as a pre-mixed Hodge object in the category of vector sheaves. Let $`H_\mu ^i(𝒞)`$ denote the bifiltered vector space with real structure (i.e. pre-mhs) obtained by applying the construction of filtrations of p. Nonabelian mixed Hodge structures as above. Then there are objects $`𝒦`$ and $`𝒦^{}`$ (pre-mixed Hodge objects in the category of vector sheaves) and morphisms
$$a:H^i(𝒞)𝒦0,$$
$$0𝒦^{}\stackrel{b}{}H^i(𝒞)$$
such that $`\mathrm{ker}(a)/\mathrm{im}(b)H_\mu ^i(𝒞)`$. Furthermore, the object $`Tot^{W,F}(𝒦)`$ is a vector scheme over $`𝒜_{wt}\times 𝒜_{hod}`$, satisfying property $`\mathrm{𝐀𝟏}`$, whose fiber over $`[1]\times 𝒜_{hod}`$ is $`0`$ and whose fiber over $`[0]\times 𝒜_{hod}`$ corresponds to a split $`i`$-shifted mixed Hodge structure. Dually, the object $`Tot^{W,F}(𝒦^{})`$ is a coherent sheaf over $`𝒜_{wt}\times 𝒜_{hod}`$ whose dual vector scheme satisfies property $`\mathrm{𝐀𝟏}`$ (this means that the coherent sheaf is annihilated by the coordinate $`t`$ on $`𝒜_{wt}`$). Again its fiber over $`[1]\times 𝒜_{hod}`$ is $`0`$ and its fiber over $`[0]\times 𝒜_{hod}`$ corresponds to a split $`i`$-shifted mixed Hodge structure. Finally, the differential of the mixed Hodge complex is an isomorphism between the split shifted mhs corresponding to $`𝒦`$ for degree $`1i`$, and the split shifted mhs corresponding to $`𝒦^{}`$ for degree $`i`$, except that this isomorphism shifts the weight filtration by one.
$`|||`$
Homology and cohomology of a geometric $`n`$-stack
Suppose $`X`$ is a $`1`$-connected geometric $`n`$-stack relative to a base $`B`$. Then we obtain its cohomology object $`𝐇^{}(X/B,𝒪)`$ which is a perfect complex over $`B`$, supported in $`[0,\mathrm{})`$. This may be seen as the higher direct image complex $`𝐑p_{}(𝒪)`$ where $`p:XB`$ is the projection.
Denote the dual perfect complex by
$$𝐇_{}(X/B):=\underset{¯}{Hom}(𝐇^{}(X/B,𝒪),𝒪).$$
It is a perfect complex over $`B`$, supported in $`(\mathrm{},0]`$. Its Dold-Puppe is an $`\mathrm{}`$-stack over $`B`$ such that any truncation is “almost geometric” in the sense of p. 118. We have a morphism
$$s:XDP(𝐇_{}(X/B)).$$
This morphism is the “stabilization map”, and one could write
$$DP(𝐇_{}(X/B))=\mathrm{\Omega }^{\mathrm{}}S^{\mathrm{}}(X/B).$$
(While we don’t go into that here, one could make a precise statment of a lemma represented by the previous equation.)
The version of Hurewicz which applies to the present situation is the following lemma:
###### Lemma 15
Suppose $`X`$ is $`k1`$-connected relative to $`B`$, for $`k2`$. Then the above stabilization map $`s`$ induces an isomorphism between the homotopy vector sheaf $`\pi _k(X/B)`$ and the cohomology vector sheaf $`H^k(𝐇_{}(X/B))`$. In particular, the homotopy fiber of $`s`$ is $`k`$-connected.
$`|||`$
Homology of a namhs
We now turn back to consideration of a $`1`$-connected nonabelian mixed Hodge structure $`𝒱`$. We apply the discussion of the previous subsection to the geometric $`n`$-stacks underlying $`𝒱`$, for example $`Tot^{W,F}(V_{DR})𝒜_{wt}\times 𝒜_{hod}`$.
We obtain homology objects such as
$$𝐇_{}(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})$$
which is a perfect complex over $`𝒜_{wt}\times 𝒜_{hod}`$ supported in $`(\mathrm{},2]`$. Doing this for $`Tot^W(V_{B,𝐑})`$ and the equivalence $`\zeta `$ too, we obtain all in all a pre-mixed Hodge object in the $`\mathrm{}`$-stack of perfect complexes supported in $`(\mathrm{},2]`$. Denote this object by $`𝐇_{}(𝒱)`$. It is a pre-mixed Hodge complex supported in $`(\mathrm{},2]`$.
###### Lemma 16
Suppose $`𝒱`$ is a $`1`$-connected namhs. Then the pre-mixed Hodge complex $`𝐇_{}(𝒱)`$ constructed above is a mixed Hodge complex.
Proof: The associated-graded of $`𝐇_{}(𝒱)`$ is the homology object of $`Gr^W(𝒱)`$. But since this latter is by hypothesis linearized, and since the Hodge filtration satisfies the strictness hypothesis $`\mathrm{𝐒𝐭𝐫}`$, we obtain a noncanonical decomposition of $`Gr^W(𝒱)`$ into a product of Eilenberg-MacLane objects for split shifted mixed Hodge structures. The homology object (which is a coalgebra) is the tensor product of the homology objects for the Eilenberg-MacLane factors. Each factor looks like $`K(𝒰,i)`$ where $`𝒰`$ is a split $`i`$-shifted mixed Hodge structure. We have the formulae
$$𝐇_{}(K(𝒰,i))=Sym^{}(𝒰[i])\text{for}i\text{even},$$
$$𝐇_{}(K(𝒰,i))=\stackrel{}{}(𝒰[i])\text{for}i\text{odd},$$
where $`𝒰[i]`$ is the object $`𝒰`$ placed in degree $`i`$. For these formulae, see the “Breen calculations” in Part III below, page 7. From these, it is immediate that $`𝐇_{}(Gr^W(𝒱))`$ is a split mixed Hodge complex. The mixed Hodge complex conditions concern only $`Gr^W`$ so this implies that $`𝐇_{}(𝒱)`$ is a mixed Hodge complex. $`|||`$
The $`i`$-th cohomology of the mixed Hodge complex of cohomology $`𝐇^{}(𝒱)`$ is a pre-mixed Hodge object in $`VESH`$, and as usual we can extract from it a pre-mhs by taking the induced filtrations on the underlying vector spaces. This yields in fact an $`i`$-shifted mixed Hodge structure which we can then normalize by shifting the weight filtration to give a mixed Hodge structure denoted $`H_\nu ^i(𝒱)`$. Cup product is a morphism of mixed Hodge structures, and we expect a statement about compatibility between the higher-order Massey products and this mixed Hodge structure, but we don’t have a precise statement at the present.
Homotopy fibers of certain morphisms of namhs
In general, the condition of being a namhs is not stable under homotopy fiber products. To see this one just has to look at the case of a mixed Hodge complex equal to a single $`i`$-shifted mhs placed in degree $`i`$. Taking the loop-space of the associated namhs (which is a homotopy fiber product of two times the punctual namhs, over the namhs in question) yields the same $`i`$-shifted mhs but now placed in degree $`1i`$. It is no longer a mixed Hodge complex.
Nonetheless, we need to have a result about preservation of the namhs condition when we take fibers of certain morphisms. A quick look at the long exact sequence for homotopy groups (on the level of $`Gr^W`$, say) leads to the appropriate condition: the connecting morphisms in the long exact sequence are morphisms between shifted mhs’s with different shifts—there will be a problem if these connecting morphisms are nonzero, so we make a hypothesis that guarantees that the connecting morphisms are zero. We obtain the following statement.
###### Lemma 17
Suppose $`f:𝒱𝒰`$ is a morphism of $`1`$-connected nonabelian mixed Hodge structures. Suppose that $`𝒰`$ is pointed by a point $`u`$. Suppose that for all $`i`$ the morphism induced by $`Gr^W(f)`$ is a surjection of homotopy group objects:
$$Gr^W(f):\pi _iGr^W(𝒱)\pi _iGr^W(𝒰)0.$$
Then the homotopy fiber $`:=𝒱\times _𝒰u`$ of $`f`$ over $`u`$ is a nonabelian mixed Hodge structure.
Proof: The $`n`$-stacks underlying $``$ are geometric and relatively $`1`$-connected; therefore they are smooth over the respective base schemes so the annihilator ideals vanish and we have automatically conditions $`\mathrm{𝐀𝟏}`$, $`\mathrm{𝐀𝟐}`$, $`\mathrm{𝐀𝟑}`$. Also locally any basepoint section provides a smooth surjection from a scheme so the flatness condition $`\mathrm{𝐅𝐥}`$ holds too. Thus we just have to verify that $`LGr^W()`$ is a mixed Hodge complex, which contains conditions $`\mathrm{𝐒𝐭𝐫}`$ and $`\mathrm{𝐌𝐇𝐂}`$. The long exact sequence for the fiber of a morphism yields a long exact sequence (of split pre-mixed Hodge objects in the stack of vector sheaves)
$$\mathrm{}H^i(LGr^W())H^i(LGr^W(𝒱))H^i(LGr^W(𝒰))\mathrm{}.$$
The morphisms going from the places occupied by $`𝒱`$ to those occupied by $`𝒰`$, are surjections by hypothesis. Thus the connecting morphisms (the leftmost and rightmost arrows in the above diagram) are zero and
$$H^i(LGr^W())=\mathrm{ker}\left(H^i(LGr^W(𝒱))H^i(LGr^W(𝒰))\right).$$
The total objects of the Hodge filtrations
$$Tot^F(H^i(LGr^W(𝒱))\text{and}Tot^F(H^i(LGr^W(𝒰)))$$
are vector bundles over $`𝒜_{hod}`$, therefore the kernel is a vector bundle too. This gives condition $`\mathrm{𝐒𝐭𝐫}`$ for $``$. Furthermore the split pre-mhs corresponding to $`H^i(LGr^W())`$ is the kernel of a morphism of $`i`$-shifted mhs’s, so it is again a $`i`$-shifted mhs. This gives condition $`\mathrm{𝐌𝐇𝐂}`$. We obtain that $``$ is a namhs. $`|||`$
Conclusion: the proof of Theorem 2
We now put together all of the above statements to obtain a proof of Theorem 2. Along the way we will see that the structure result of Theorem 3 also holds for the homotopy group objects $`\pi _i(𝒱)`$.
Suppose that $`𝒱`$ is $`k1`$-connected for $`k2`$. The proof will be by descending induction on $`k`$, so we assume that Theorem 2 and indeed the structure result analogous to Theorem 3 are known for any $`k`$-connected namhs $``$. (To be technically correct in the induction, we add to the inductive hypothesis that at the bottom degree $`k+1`$ there is no quotient object in the structure result.)
Let $`𝒞:=𝐇_{}(𝒱)`$. It is a mixed Hodge complex supported in $`(\mathrm{},k)`$. Combining the constructions of Propositions 13 and 14, we obtain a mixed Hodge complex $`𝒩`$ supported in $`[1k,k]`$ and a morphism $`𝒞𝒩`$ inducing surjections on the cohomology objects of $`Gr^W(𝒞)`$ and $`Gr^W(𝒩)`$, and indeed inducing an isomorphism of cohomology objects in degree $`k`$.
The complex underlying $`𝒩`$ is exact in degree $`1k`$ and its cohomology in degree $`k`$ carries a mixed Hodge structure $`H_\mu ^k(𝒞)`$. The cohomology objects $`H^i(𝒩)`$ are as follows: in degree $`1k`$, an object such as was denoted $`𝒦`$ in Theorem 3; and in degree $`k`$ an extension with subobject as was denoted $`𝒦^{}`$ in Theorem 3 and quotient object $`H_\mu ^k(𝒞)`$.
The Dold-Puppe $`DP(𝒩)`$ has a structure of namhs (it is linearized by setting $`LGr^W(DP(𝒩)):=Gr^W(𝒩)`$). Note that, whereas $`DP(𝒞)`$ is an $`\mathrm{}`$-stack, $`DP(𝒩)`$ is an $`n`$-stack (or at most an $`n+1`$-stack if we are dealing with the limit case $`k=n`$).
Now the fact that $`k2`$ means that the nonlinear terms in the Breen calculations for the homology of $`Gr^W(𝒱)`$ start intervening only in degree $`k+2`$ or more. In particular the morphism
$$Gr^W(𝒱)DP(Gr^W(𝒞))$$
induces a surjection on $`\pi _k`$ and $`\pi _{k+1}`$. Therefore the morphism $`f:𝒱DP(𝒩)`$ induces surjections on the homotopy group objects of the associated-graded, in particular $`f`$ satisfies the hypotheses of Lemma 17. That lemma says that the fiber $``$ of $`f`$ over the zero-section of $`DP(𝒩)`$ is a nonabelian mixed Hodge structure. The long exact sequence of homotopy objects gives a long exact sequence of pre-mixed Hodge objects in the category of vector sheaves,
$$\mathrm{}\pi _i()\pi _i(𝒱)H^i(𝒩)\pi _{i1}()\mathrm{}.$$
Our inductive hypothesis says that for $`ik+2`$ the objects $`\pi _i()`$ satisfy the structure result of Theorem 3 (and in particular that the pieces corresponding to bifiltered vector spaces with real structure, are $`i`$-shifted mhs’s); and it says that for $`i=k+1`$ the same holds but with the quotient being zero in the structure result.
Note that for $`ik+2`$, $`H^i(𝒩)=0`$ so $`\pi _i(𝒱)=\pi _i()`$ and we obtain the desired structure result for $`\pi _i(𝒱)`$. For $`i=k+1`$ we have an exact sequence
$$0\pi _{k+1}()\pi _{k+1}(𝒱)H^{1k}(𝒩)0.$$
The fact that the connecting map is zero comes from the fact that we constructed $`𝒩`$ so that $`\pi _{k+1}(𝒱)H^{1k}(𝒩)`$ is surjective. From this exact sequence we obtain the desired structure result for $`\pi _{k+1}(𝒱)`$: the inductive hypothesis gives the structure result but without quotient object, for $`\pi _{k+1}()`$; and the piece $`H^{1k}(𝒩)`$ provides exactly a quotient object as forseen in Theorem 3. Finally we have by construction an isomorphism
$$\pi _k(𝒱)H^k(𝒩)$$
(which justifies why $``$ was $`k`$-connected). This gives the structure result for $`\pi _k(𝒱)`$ with no quotient object, and in particular with $`\pi _k^\mu (𝒱)`$ being a $`k`$-shifted mixed Hodge structure.
By induction we obtain the structure result of Theorem 3 for the homotopy group objects $`\pi _i(𝒱)`$. This gives the first part of the statement of Theorem 2, that the $`\pi _i^\mu (𝒱)`$ are $`i`$-shifted mixed Hodge structures. It also gives the statement that the associated-graded objects $`Gr^W\pi _i^\mu (𝒱)`$ are subquotients of the $`\pi _i(Gr^W(𝒱))`$. Now by the existence of a linearization, the Whitehead products between the $`\pi _i(Gr^W(𝒱))`$ vanish. This implies that the Whitehead products between the $`Gr^W\pi _i^\mu (𝒱)`$ vanish. The same argument as above (starting page Nonabelian mixed Hodge structures) gives that on the normalized mixed Hodge structures $`\pi _i^\nu (𝒱)`$, the Whitehead products are morphisms of mixed Hodge structures. This completes the proof of Theorem 2. $`|||`$
We conclude this section by remarking that the above Lemma 17 and its counterexample show that the the $`n+1`$-category $`nNAMHS`$ is not closed under homotopy limits, but only under very special kinds of them. We wonder about the following:
Question: Is the $`n+1`$-category $`nNAMHS`$ closed under a reasonable class of homotopy colimits?
The basic construction
The basic construction which we would like to describe is the “nonabelian cohomology” of a smooth projective variety $`X`$ with coefficients in a nonabelian mixed Hodge structure. In this section we define the cohomology with coefficients in a pre-namhs $`𝒱`$ (eventually linearized), denoted
$$=\underset{¯}{Hom}(X_M,𝒱).$$
This will be an object of the same type as a pre-namhs, i.e. an object in $`PMH.nSTACK`$. We show that it is in fact a pre-namhs (the meaning of this condition being that the $`n`$-stacks involved are geometric or equivalently that it is an object of $`PMH.nGEOM`$) under certian reasonable hypotheses on $`𝒱`$. If $`𝒱`$ was linearized then $``$ will be linearized.
In what follows $`X`$ will be a smooth projective variety over $`𝐂`$ and $`𝒱`$ will be a pre-namhs (with the same notation as in the definition page Nonabelian mixed Hodge structures). As a matter of notation, we will denote by $``$ the object $`\underset{¯}{Hom}(X_M,𝒱)`$ which we will construct, and as above we will have
$$=\{(H_{DR},W,F),(H_{B,𝐑},W),\zeta _{}\}.$$
We first describe what is meant by $`X_M`$. This is an object which is somewhat like a pre-namhs, however the $`n`$-stacks involved are not geometric, and the morphism $`\zeta `$ is not an equivalence.
Recall from that we have defined a formal category $`X_{Hod}`$ over $`𝒜_{hod}`$ which gives a $`1`$-stack
$$X_{Hod}𝒜_{hod}.$$
Technically speaking in we denoted by $`X_{Hod}`$ the $`𝐆_m`$-equivariant formal category over $`𝐀_{hod}^1`$; for our purposes here it is more convenient to take the quotient and denote by $`X_{Hod}`$ the object over $`𝒜_{hod}`$.
The fiber of $`X_{Hod}`$ over $`[1]`$ is $`X_{DR}`$ and the fiber over $`[0]=B𝐆_m`$ is $`X_{Dol}`$ with its action of $`𝐆_m`$. On the other hand, we define $`X_{B,𝐑}`$ to be the constant $`n`$-stack over $`𝒵_𝐑`$ whose value is the Poincaré $`n`$-groupoid corresponding to the $`n`$-truncation of the homotopy type of $`X^{\mathrm{top}}`$. Note that the complexified stack associated to a constant real stack is again a constant stack on $`𝒵`$ with the same value, in particular $`X_{B,𝐂}`$ is the constant stack on $`𝒵`$ whose value is the Poincaré $`n`$-groupoid of $`X^{\mathrm{top}}`$.
In order to define the morphism of analytic $`n`$-stacks
$$\zeta _X:X_{DR}^{\mathrm{an}}X_{B,𝐂}^{\mathrm{an}},$$
we need the following lemma.
###### Lemma 18
Define the $`n`$-prestack $`C^{\mathrm{pre}}(,X^{\mathrm{top}})`$ on $`𝒵^{\mathrm{an}}`$ to be the prestack which to each $`Y`$ associates the Poincaré $`n`$-groupoid of the space of continuous maps from $`Y^{\mathrm{top}}`$ to $`X^{\mathrm{top}}`$. Then the associated $`n`$-stack which we denote by $`C(,X^{\mathrm{top}})`$, is naturally equivalent to $`X_{B,𝐂}^{\mathrm{an}}`$.
Proof: Let $`X_{B,pre,𝐂}`$ be the constant $`n`$-prestack with values equal to the Poincaré $`n`$-groupoid of $`X^{\mathrm{top}}`$, and let $`X_{B,pre,𝐂}^{\mathrm{an}}`$ be the associated analytic prestack which is again a constant prestack. Then $`X_{B,𝐂}`$ (resp. $`X_{B,𝐂}^{\mathrm{an}}`$) is the $`n`$-stack associated to $`X_{B,pre,𝐂}`$ (resp. $`X_{B,pre,𝐂}^{\mathrm{an}}`$). We have a morphism of prestacks obtained by considering a point of $`X^{\mathrm{top}}`$ as a constant map from any $`Y^{\mathrm{top}}`$ to $`X^{\mathrm{top}}`$,
$$X_{B,pre,𝐂}^{\mathrm{an}}C^{\mathrm{pre}}(,X^{\mathrm{top}}).$$
This morphism is an equivalence, over any $`Y`$ such that $`Y^{\mathrm{top}}`$ is contractible. Since the contractible open sets form a base for the topology of any $`Y𝒵^{\mathrm{an}}`$, this property implies that the above morphism of prestacks induces an equivalence on associated $`n`$-stacks
$$X_{B,𝐂}^{\mathrm{an}}\stackrel{}{}C(,X^{\mathrm{top}}).$$
$`|||`$
Recall that
$$X_{DR}^{\mathrm{an}}(Y)=X(Y^{\mathrm{red}})$$
where $`Y^{\mathrm{red}}`$ is the reduced complex analytic space. A morphism from $`Y^{\mathrm{red}}`$ to $`X`$ gives a continuous map from $`Y^{\mathrm{top}}`$ to $`X^{\mathrm{top}}`$, so we get a morphism
$$X_{DR}^{\mathrm{an}}(Y)C(,X^{\mathrm{top}})(Y)X_{B,𝐂}^{\mathrm{an}}(Y).$$
Our $`\zeta _X`$ is defined to be the composed morphism $`X_{DR}^{\mathrm{an}}X_{B,𝐂}^{\mathrm{an}}`$.
The object which we denote informally by $`X_M`$ is the collection
$$X_M=\{X_{Hod}𝒜_{hod},X_{B,𝐑},\zeta _X\}.$$
Note that $`X_{Hod}`$ is an object of $`F.nSTACK`$. We can consider it as an object of $`F.F.nSTACK`$ by making it constant in the other “weight” direction; and similarly the $`n`$-stack $`X_{B,𝐑}`$ may be thought of as an object of $`F.nSTACK`$ constant over $`𝒜_{wt}`$. In this sense the object $`X_M`$ is somewhat similar to a pre-namhs. The important differences are:
–the $`n`$-stacks involved are not geometric; and
–the morphism $`\zeta _X`$ is a morphism but not an equivalence.
In spite of these differences, we still define an object $`=\underset{¯}{Hom}(X_M,𝒱)`$ as we shall now start to do.
The nonlinearized case
We will first construct $``$ in the case where the pre-namhs $`𝒱`$ is not linearized. Put:
$$H_{DR}:=\underset{¯}{Hom}(X_{DR},V_{DR}).$$
It is a geometric $`n`$-stack by , .
As indicated above we obtain objects which are constant in the weight direction
$$X_{Hod}\times 𝒜_{wt}𝒜_{hod}\times 𝒜_{wt},$$
and
$$X_{B,𝐑}\times 𝒜_{wt,𝐑}𝒜_{wt,𝐑}.$$
Restricting the first to $`[1]𝒜_{hod}`$ we obtain
$$X_{DR}\times 𝒜_{wt}𝒜_{wt}.$$
Now we define the weight and Hodge filtrations on this $`n`$-stack by defining the total space as
$$Tot^{W,F}(H_{DR}):=\underset{¯}{Hom}^{\mathrm{se},0}(\frac{X_{Hod}\times 𝒜_{wt}}{𝒜_{hod}\times 𝒜_{wt}},\frac{Tot^{W,F}(V_{DR})}{𝒜_{hod}\times 𝒜_{wt}}).$$
Here $`\underset{¯}{Hom}(\frac{A}{B},\frac{C}{B})`$ is the relative internal $`Hom`$ of $`n`$-stacks over $`B`$. It can also be viewed as the internal $`Hom`$ in the stack $`(n+1)STACK_{1/}(B,nSTACK)`$ (there is a compatibility between these two things, which we don’t get into).
The member on the left of the internal $`\underset{¯}{Hom}`$ is a formal category of smooth type over the base $`B=𝒜_{hod}\times 𝒜_{wt}`$.
The superscript means that we take the substack of semistable morphisms with vanishing Chern classes, see the discussion in §10.7 and on page Nonabelian mixed Hodge structures below. This semistability condition is automatic for Betti or de Rham cohomology, so it is only necessary for the Dolbeault cohomology, i.e. at the associated-graded for the Hodge filtration, over $`[0]_{hod}𝒜_{hod}`$. A morphism $`X_{Dol}V`$ is said to be semistable with vanishing Chern classes if the morphism $`X_{Dol}\pi _0(V)`$ is constant, say with values at a point $`t\pi _0(V)`$, and if the morphism $`X_{Dol}\tau _1(V)`$ which becomes essentially
$$X_{Dol}K(\pi _1(V,t),1),$$
is a semistable principal Higgs $`\pi _1(V,t)`$-torsor on $`X`$ with vanishing rational Chern classes. One thing that we need to know here is that the condition of semistability with vanishing Chern classes is an open condition in the $`1`$-stack of morphisms toward $`\tau _1V`$. This is Proposition 23 below.
If $`𝒱`$ is simply connected, i.e. the fundamental group object is trivial
$$\pi _1(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})=\{1\},$$
then the semistability condition is automatic. In this case we can apply Theorem 10.3.3 of to get that $`Tot^{W,F}(H_{DR})`$ is a geometric $`n`$-stack over $`𝒜_{wt}\times 𝒜_{hod}`$. For now we will restrict to the case where $`𝒱`$ is simply connected in our statement (Theorem 4) about $``$.
Later on in Proposition 25 we show that if the fundamental group object of $`𝒱`$ is a flat linear group scheme then $`Tot^{W,F}(H_{DR})`$ is geometric; this leads to the statement of Theorem 6 which extends Theorem 4 to cover this case.
The total space of the weight filtration on $`H_{DR}`$ has a formula similar to the previous one, namely:
$$Tot^W(H_{DR}):=\underset{¯}{Hom}(\frac{X_{DR}\times 𝒜_{wt}}{𝒜_{wt}},\frac{Tot^W(V_{DR})}{𝒜_{wt}}).$$
Next, we define similarly
$$H_{B,𝐑}:=\underset{¯}{Hom}(X_{B,𝐑},V_{B,𝐑}).$$
It is a real geometric $`n`$-stack cf Proposition 2.1 and the comment on page 12, which refers to Corollary 5.6; also in the present real case we use Lemma 10 above saying it suffices to check that $`H_{B,𝐂}`$ is geometric.
As before the weight filtration on $`H_{B,𝐑}`$ is defined by the formula
$$Tot^W(H_{B,𝐑}):=\underset{¯}{Hom}(\frac{X_{B,𝐑}\times 𝒜_{wt}}{𝒜_{wt}},\frac{Tot^W(V_{B,𝐑})}{𝒜_{wt}}).$$
This is geometric by Proposition 2.1, Corollary 5.6, and Lemma 10 above.
The analytic stacks associated to the algebraic $`n`$-stacks $`Tot^W(H_{DR})`$ and $`Tot^W(H_{B,𝐂})`$ are given again by the same formulae, namely:
$$Tot^W(H_{DR})^{\mathrm{an}}:=\underset{¯}{Hom}(\frac{X_{DR}^{\mathrm{an}}\times 𝒜_{wt}^{\mathrm{an}}}{𝒜_{wt}^{\mathrm{an}}},\frac{Tot^W(V_{DR})^{\mathrm{an}}}{𝒜_{wt}^{\mathrm{an}}})$$
and
$$Tot^W(H_{B,𝐂})^{\mathrm{an}}:=\underset{¯}{Hom}(\frac{X_{B,𝐂}^{\mathrm{an}}\times 𝒜_{wt}^{\mathrm{an}}}{𝒜_{wt}^{\mathrm{an}}},\frac{Tot^W(V_{B,𝐂})^{\mathrm{an}}}{𝒜_{wt}^{\mathrm{an}}}).$$
Using these last, our map $`\zeta _X`$ gives rise to a map
$$\zeta _{}:Tot^W(H_{B,𝐂})^{\mathrm{an}}Tot^W(H_{DR})^{\mathrm{an}},$$
and the “GAGA theorem” of says that this map is an equivalence. (Note that here we come across, for the first time, a case where the techniques of are needed for nonabelian cohomology with coefficients in stacks where $`\pi _0`$ is nontrivial, in this case the $`\pi _0`$ is the base of the weight filtration, $`𝒜_{wt}`$.)
We have now constructed an object $``$ in $`PMH.nSTACK`$, and if $`𝒱`$ is simply connected then $``$ is in $`PMH.nGEOM`$. In Theorem 6 we will get $``$ in $`PMH.nGEOM`$ if
$$\pi _i(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})$$
is a flat linear group scheme over $`𝒜_{wt}\times 𝒜_{hod}`$.
The case of coefficients in a pre-mhc
In order to treat the case of coefficients in a Dold-Puppe linearized object $`𝒱`$, we have to first treat the construction of an object $`\underset{¯}{Hom}(X_M,𝒞)`$ when $`𝒞`$ is a split pre-mhc (this will then be applied to $`𝒞=LGr^W(𝒱)`$). In fact it is obviously interesting to do this even when $`𝒞`$ is not split. Thus in this section we will define a pre-mhc $`\underset{¯}{Hom}(X_M,𝒞)`$ whenever $`𝒞`$ is a pre-mhc.
We start with the following general construction. Suppose $`B`$ is a base $`n`$-stack (of $`n`$-groupoids). Suppose $`B`$ is an $`n`$-stack over $`B`$ and suppose that $`𝒞`$ is a complex of sheaves of abelian groups (which are assumed to be rational vector spaces) on $`B`$ supported in $`[n,0]`$ (i.e. $`𝒞`$ is a morphism from $`B`$ to the $`n`$-stack $`nCPX`$ of complexes of sheaves of abelian groups). Then we can define the relative morphism complex
$$=\underset{¯}{Hom}(\frac{}{B},\frac{𝒞}{B})$$
which is a complex of sheaves of abelian groups (again, rational vector spaces) over $`B`$. One way of doing this is to set $``$ equal to the higher direct image complex of the restriction $`𝒞|_{}`$ via the morphism $`p:B`$, which is then truncated in positive degrees to give a complex supported in degrees $`[n,0]`$:
$$:=\tau _0𝐑p_{}(p^{}𝒞).$$
Another way of doing it is to let $`𝒞`$ correspond to a relatively $`N`$-truncated $`n+N`$-stack $`\stackrel{~}{𝒞}`$ over $`B`$, to put
$$\stackrel{~}{}^{}:=\underset{¯}{Hom}(\frac{}{B},\frac{\stackrel{~}{𝒞}}{B}),$$
and to let
$$\stackrel{~}{}:=\mathrm{fib}(\stackrel{~}{}^{}\tau _n(\stackrel{~}{}^{})).$$
The fiber is taken over the zero-morphism as basepoint section, where the zero-morphism is the constant morphism with values the basepoint section of $`\stackrel{~}{𝒞}`$. Now $`\stackrel{~}{}`$ is a relatively $`N`$-connected $`n+N`$-stack over $`B`$ so it corresponds to a complex of sheaves of abelian groups.
The above construction is compatible with the usual $`\underset{¯}{Hom}`$ of $`n`$-stacks via Dold-Puppe:
$$\underset{¯}{Hom}(\frac{}{B},\frac{DP(𝒞)}{B})=DP\underset{¯}{Hom}(\frac{}{B},\frac{𝒞}{B}).$$
The case which will interest us is when $``$ is a formal category over $`B`$. See for example for the definitions concerning this notion. Note that the definitions of formal category and the various properties of such, were defined in for the case where the base $`B`$ was a scheme. These are immediately extended to the case where $`B`$ is any $`n`$-stack by the usual trick of looking at a formal category over $`B`$ (resp. one having certain properties) as a morphism from $`B`$ into the $`2`$-stack of formal categories (resp. the substack of those having the properties in question).
The following lemma is basically the same as Theorem 10.2.5 of , except that we have added the extension to the case of hemiperfect complexes.
###### Lemma 19
Suppose $``$ is a formal category over $`B`$ which is projective and of smooth type relative to $`B`$. Suppose $`𝒞`$ is a hemiperfect complex over $`B`$. Then the morphism complex $`=\underset{¯}{Hom}(\frac{}{B},\frac{𝒞}{B})`$ is a hemiperfect complex over $`B`$.
Proof: To check that $``$ is hemiperfect it suffices (by definition) to check that this is the case over any scheme $`ZB`$. Thus we may assume that $`B`$ is a scheme, and in fact we may when necessary replace $`B`$ by an open subset (part of an open cover). With this reduction we may assume that there is a perfect complex $`𝒦`$ supported in $`[n,1]`$ such that $`𝒞=\tau _0𝒦`$ (put $`𝒦^i=𝒞^i`$ for $`i<0`$ and let $`d:𝒦^0𝒦^1`$ be a morphism which has the vector scheme $`𝒞^0`$ as its kernel). Recall that $`p`$ denotes the morphism $`B`$. We have
$$:=\tau _0𝐑p_{}(p^{}𝒞)=\tau _0𝐑p_{}(p^{}𝒦).$$
Thus it suffices to prove that $`\tau _0𝐑p_{}(p^{}𝒦)`$ is a perfect complex supported in $`[n,\mathrm{})`$. On the other hand, $`𝒦`$ is, up to quasiisomorphism, made up by taking successive cones of shifts of vector bundles. Perfection is preserved by cones, and higher direct image preserves triangles so up to quasiisomorphism it takes cones to cones. Shifts also commute with higher direct image and preserve perfection. Finally, by replacing $`B`$ by an open cover we may assume that the vector bundles are trivial bundles, so we are reduced to claiming that
$$𝐑p_{}(𝒪)$$
is a perfect complex on $`B`$. Since $``$ is of smooth type relative to $`B`$ (i.e. the morphism $`B`$ is of smooth type in the terminology of 8.6), its underlying scheme $`X`$ is flat over $`B`$, and the higher direct image is calculated as the higher direct image of the “complex of differentials” for $``$ (cf 10.2.5):
$$𝐑p_{}(𝒪)=𝐑p_{X/B,}(\mathrm{\Omega }_{X/}^{}).$$
The condition that $``$ be of smooth type (cf ) means that $`\mathrm{\Omega }_{X/}^{}`$ is a differential complex of locally free sheaves on $`X`$. Since $`X`$ is by hypothesis projective over $`B`$, it is well-known that the higher direct image is a perfect complex. This completes the proof. $`|||`$
We have an analogue of the previous lemma designed to be applied to the “Betti” case.
###### Lemma 20
Suppose $``$ is an $`n`$-stack over $`B`$ which is locally constant with fiber the homotopy type of a finite CW complex. Suppose $`𝒞`$ is a hemiperfect complex over $`B`$. Then we can define as previously a morphism complex $`=\underset{¯}{Hom}(\frac{}{B},\frac{𝒞}{B})`$ which is hemiperfect over $`B`$.
Proof: Left as an exercise. $`|||`$
Suppose $`𝒞`$ is a pre-mhc, and suppose $`X`$ is a smooth projective variety. By using exactly the same formulae as in the previous construction starting on page Nonabelian mixed Hodge structures we can define a pre-mhc $`:=\underset{¯}{Hom}(X_M,𝒞)`$. In doing this translation, replace $`𝒱`$ by $`𝒞`$ (thus, $`V_{DR}`$ by $`C_{DR}`$ and so forth), and use the morphism complexes of Lemmas 19 and 20 in place of the usual internal $`\underset{¯}{Hom}`$ stacks used in the construction starting on page Nonabelian mixed Hodge structures. We don’t write out the formulae here because they are exactly the same as in the previous construction. Again the map $`\zeta _X`$ gives rise to the required equivalence $`\zeta _{}`$ between $`H_{DR}`$ and $`H_{B,𝐂}`$. Note that in this case, analytic equivalences between hemiperfect complexes are algebraic so $`\zeta _{}`$ is actually an algebraic equivalence.
If $`𝒞`$ is a split pre-mhc then $`\underset{¯}{Hom}(X_M,𝒞)`$ is also a split pre-mhc. Indeed, the action of $`𝐆_m`$ splitting the weight filtration on $`𝒞`$ induces an action on $`\underset{¯}{Hom}(X_M,𝒞)`$ which again splits the weight filtration; this is because the object $`X_M`$ is constant in the weight-filtration direction (i.e. in the direction of $`𝒜_{wt}`$) so it doesn’t contribute to changing the degree in the weight filtration. Thus, $`\underset{¯}{Hom}(X_M,𝒞)`$ is split as a pre-mhc.
The case of coefficients in a linearized pre-namhs
We now turn to the case which combines the two previous ones. Suppose $`𝒱`$ is a linearized pre-namhs. Recall that this means that $`𝒱`$ is a pre-namhs and that we have a split pre-mhc $`LGr^W(𝒱)`$ whose Dold-Puppe is equivalent to $`Gr^W(𝒱)`$ as split pre-namhs’s. We now define a pre-namhs
$$=\underset{¯}{Hom}(X_M,𝒱)$$
as follows. The underlying pre-namhs is the one which was constructed starting on page Nonabelian mixed Hodge structures. We set
$$LGr^W():=\underset{¯}{Hom}(X_M,LGr^W(𝒱)),$$
which is a split pre-mhc. The above constructions are compatible with Dold-Puppe (i.e. the morphism complex defined in Lemmas 19 and 20 are compatible with the usual relative internal $`\underset{¯}{Hom}`$ stacks via Dold-Puppe), so we have a canonical equivalence
$$DP(LGr^W())Gr^W().$$
This completes our construction of a linearized pre-namhs $``$.
We sum up the results of this and the preceding paragraphs:
###### Theorem 4
Suppose $`X`$ is a smooth projective variety and $`𝒱`$ is a pre-namhs (resp. pre-mhc, resp. linearized pre-namhs). Suppose that $`𝒱`$ is simply connected (resp. no additional hypothesis in the pre-mhc case, resp. the same hypothesis in the linearized pre-namhs case). Then the object
$$\underset{¯}{Hom}(X_M,𝒱)==\{(H_{DR},W,F),(H_{B,𝐑},W),\zeta _{}\}$$
defined above is a pre-namhs (resp. pre-mhc, resp. linearized pre-namhs).
$`|||`$
The mixed Hodge complex conditions
The following proposition comes from Deligne’s construction in“Hodge III” .
###### Theorem 5
Suppose $`𝒞`$ is a truncated mixed Hodge complex (see page Nonabelian mixed Hodge structures) and suppose that $`X`$ is a smooth projective variety. Then the pre-mixed Hodge complex $`\underset{¯}{Hom}(X_M,𝒞)`$ that we have constructed above is a truncated mixed Hodge complex, i.e. it satisfies conditions $`\mathrm{𝐒𝐭𝐫}`$ (p. 5), $`\mathrm{𝐌𝐇𝐂}`$ (p. 1), $`\mathrm{𝐀𝟏}`$ (p. 3) and $`\mathrm{𝐀𝟑}`$ (p. Nonabelian mixed Hodge structures).
Proof: There is an actual mixed Hodge complex $`𝒟`$ (for example supported in $`[n,1]`$) such that $`𝒞`$ is the truncation of $`𝒟`$ into degrees $`0`$. Then $`\underset{¯}{Hom}(X_M,𝒞)`$ is the truncation of $`\underset{¯}{Hom}(X_M,𝒟)`$. If we show that the latter is a truncated mixed Hodge complex, then by Corollary 11 its truncation will be a truncated mixed Hodge complex. Thus we are reduced to the case where $`𝒞`$ is a mixed Hodge complex. Furthermore one shows $`()`$, in fact, that the full cohomology complex of $`X_M`$ with coefficients in $`𝒞`$ is a mixed Hodge complex supported in $`[n,\mathrm{})`$; the complex we denote $`\underset{¯}{Hom}(X_M,𝒞)`$ is the truncation of this, so it is a truncated mixed Hodge complex.
To show $`()`$ that the cohomology is a mixed Hodge complex, it suffices to treat the associated-graded; but $`Gr^W(𝒞)`$ decomposes as a direct sum of shifted split mixed Hodge structures. Thus it suffices to treat the case of cohomology with coefficients in a pure Hodge structure. It is well-known that it gives back a pure Hodge structure; the only problem is to verify that the shifts of weights work out appropriately. For this, we refer to the discussion on page 1. $`|||`$
###### Corollary 21
If $`𝒞`$ is a split mixed Hodge complex (cf p. 1), and $`X`$ is a smooth projective variety, then the pre-mhc $`\underset{¯}{Hom}(X_M,𝒞)`$ is a split mixed Hodge complex.
Proof: As remarked above, the splitting carries over to $`\underset{¯}{Hom}(X_M,𝒞)`$. We just have to show that the truncated mixed Hodge complex is actually a mixed Hodge complex. The only direction in which the hemiperfect complexes involved in $`\underset{¯}{Hom}(X_M,𝒞)`$ might not be perfect, is in the $`𝒜_{wt}`$ direction (in the $`𝒜_{hod}`$ direction they are perfect complexes because of the strictness conditon $`\mathrm{𝐒𝐭𝐫}`$). However, the splitting of the weight filtration means that in the $`𝒜_{wt}`$ direction the complexes are constant so they are perfect complexes. Thus $`\underset{¯}{Hom}(X_M,𝒞)`$ is a split mixed Hodge complex. $`|||`$
###### Corollary 22
Suppose in the situation of Theorem 4 that $`𝒱`$ is a linearized pre-namhs and satisfies the strictness condition $`\mathrm{𝐒𝐭𝐫}`$ (Lemma 5 p. 5) and the mixed Hodge complex condition $`\mathrm{𝐌𝐇𝐂}`$ (p. 1). Then the linearized pre-namhs $`=\underset{¯}{Hom}(X_M,𝒱)`$ also satisfies the strictness condition $`\mathrm{𝐒𝐭𝐫}`$ and the condition $`\mathrm{𝐌𝐇𝐂}`$.
Proof: These conditions are measured on the $`LGr^W(𝒱)`$ and $`LGr^W()`$, so Theorem 5 and the previous corollary apply. $`|||`$
A couple of further remarks
The construction which to $`X`$ and $`𝒱`$ associates $``$ is functorial in $`𝒱`$ and contravariantly functorial in $`X`$ (i.e. it gives an $`n+1`$-functor from the product of the category of smooth projective varieties with the $`n+1`$-category of connected pre-namhs’s (resp. connected linearized pre-namhs’s), to the $`n+1`$-category of pre-namhs’s (resp. linearized pre-namhs’s)).
It is relatively clear from the above construction (namely, from the fact that $`X_M`$ was taken to be constant in the $`𝒜_{wt}`$-direction) that $`𝒱`$ depends only on the data of the very presentable shape of $`X_{Hod}/𝒜_{hod}`$ together with the very presentable shape of $`X_B`$ and the equivalence between the Betti shape and the de Rham shape after composing with the analytification functor. In the future one hopes for a construction dealing with singular and open varieties, in which case there will surely be weight data contained in the appropriate generalization of $`X_M`$.
Semistable morphisms
In this section, we look more closely at $`\underset{¯}{Hom}(X_M,𝒱)`$ when $`𝒱`$ is a pre-namhs which is not necessarily simply connected. This allows us to give a statement like Theorem 4 for the case of $`𝒱`$ not simply connected. The results of this section are not used in Part III.
For the purposes of this discussion and to reduce the volume of notation, we denote $`B:=𝒜_{wt}\times 𝒜_{hod}`$ and
$$:=Tot^{W,F}(X_{DR})=X_{Hod}\times 𝒜_{wt}B$$
which is a formal category, projective and of smooth type over $`B`$. The important properties of this situation are:
(1) There is a closed substack $`B_1B`$ such that
$$|_{B_1}=X_{Dol}\times B_1,|_{BB_1}=X_{DR}\times (BB_1);$$
(2) For cohomology with vector sheaf coefficients, $`/Z`$ is of finite cohomological dimension; and
(3) There are only finitely many equivalence classes of $`Spec(𝐂)`$-valued points in $`B`$.
We shall make use of these properties to get around what seem otherwise to be some delicate questions.
To continue with our notation, we fix a pre-namhs $`𝒱`$ and we let $`V:=Tot^{W,F}(V_{DR})`$ denote the component which is a geometric $`n`$-stack over $`B`$. We assume that this is connected relative to $`B`$. Suppose $`Z`$ is a scheme mapping to $`B`$. Let $`_Z:=\times _BZ`$ and $`V_Z:=V\times _BZ`$; by going to an etale neighborhood in $`Z`$ if necessary, we may assume that there is a section $`v:ZV_Z`$. Now put $`G_Z:=\pi _1(V_Z/Z,v)`$. This is a very presentable group sheaf over $`Z`$.
Hypothesis: That the fundamental group object of $`𝒱`$ is a flat linear group scheme; this means that for every scheme $`Z`$ mapping to $`B`$ and basepoint section $`v:ZV_Z`$ as above, the group sheaf $`G_Z=\pi _1(V_Z/Z,v)`$ is represented by a flat linear group scheme over $`Z`$.
(Note that there might not exist a basepoint of $`𝒱`$ over $`B`$ so the fundamental group object doesn’t exist per se; the condition on $`G_Z`$ for every scheme $`Z`$ is thus the definition of the above hypothesis.)
We recall the definition of the sub-$`n`$-stack “of semistable morphisms with vanishing Chern classes”,
$$\underset{¯}{Hom}^{\mathrm{se},0}(/B,V/B)\underset{¯}{Hom}(/B,V/B).$$
If $`Z\stackrel{f}{}B`$ is a scheme mapping to $`B`$, we shall say when a point $`\phi :_ZV_Z`$ of the right hand $`n`$-stack, is a member of the substack. For the purposes of stating this condition we may replace $`Z`$ by an etale open neigborhood and assume that $`V_Z`$ admits a basepoint section $`v`$ (the condition will be independent of the choice of basepoint section so the condition glues to give a condition over any $`Z`$).
Given a map $`\phi `$ as above, for any closed point $`zZ`$ we obtain a morphism $`\phi _z:_zV_z`$. The first truncation in the Postnikov tower for $`V_z`$ is
$$V_z\tau _1(V_z)=K(G_z,1).$$
Thus $`\phi _z`$ projects to give a morphism $`\rho _z:_zK(G_z,1)`$, i.e. a principal $`G_z`$-bundle over $`_z`$.
Recall from (1) above that there are two cases. If $`f(z)(BB_1)`$ then $`_z=X_{DR}`$ and $`\rho _z`$ is a principal $`G_z`$-bundle with flat connection over $`X`$. In this case the condition of being semistable with vanishing Chern classes is automatic, and we admit $`\phi _z`$. If $`f(z)B_1`$ then $`_z=X_{Dol}`$ and $`\rho _z`$ is a principal $`G_z`$-Higgs bundle over $`X`$. In this case we say that $`\phi _z`$ is “semistable with vanishing Chern classes” if the rational Chern classes of the principal bundle $`\rho _z`$ vanish for all invariant polynomials on $`G_z`$, and if for every linear representation $`E`$ of $`G_z`$, the associated Higgs bundle $`E(\rho _z)`$ is semistable.
Now we say that a point $`\phi :_ZV_Z`$ is “semistable with vanishing Chern classes” if for every point $`zZ`$ the restriction $`\phi _z`$ is semistable with vanishing Chern classes in the above sense. We define
$$\underset{¯}{Hom}^{\mathrm{se},0}(/B,V/B)(Z)\underset{¯}{Hom}(/B,V/B)(Z)$$
to be the full sub-$`n`$-groupoid of such points. This defines the substack.
###### Proposition 23
Under the above notations for $`/B`$ and $`V/B`$ and with the hypothesis that the $`G_Z`$ are flat linear group schemes over $`Z`$, the sub-$`n`$-stack $`\underset{¯}{Hom}^{\mathrm{se},0}(/B,V/B)`$ is an open substack of $`\underset{¯}{Hom}(/B,V/B)`$.
Proof: Suppose $`ZB`$ is a morphism from a scheme, and suppose that $`\phi :_ZV_Z`$ is a point. We have to show that the subset of points $`zZ`$ such that $`\phi _z`$ is semistable with vanishing Chern classes, is an open subset of $`Z`$. For this we can localize on $`Z`$ so we assume that there is a basepoint section $`v`$ of $`V_Z`$ and let $`G_Z`$ be the fundamental group as above. Let $`\rho `$ be the principal $`G_Z`$-bundle over $`_Z`$ obtained by the first Postnikov projection of $`\phi `$. We have to prove that the set of points $`zZ`$ such that $`\rho _z`$ is semistable with vanishing Chern classes, is open. Let $`Z_1`$ be the inverse image of $`B_1`$; it is a closed subscheme of $`Z`$ and outside of $`Z_1`$ all of the points are semistable with vanishing Chern classes. The complement of the set in question is thus contained in $`Z_1`$; since $`Z_1`$ is closed in $`Z`$, it suffices to prove that this complement is closed in $`Z_1`$, in other words it suffices to prove that the subset of points $`zZ_1`$ where $`\rho _z`$ is semistable with vanishing Chern classes, is open in $`Z_1`$. We are reduced to this question on $`Z_1`$ so we may assume $`Z=Z_1`$. With this assumption $`_Z=X_{Dol}\times Z`$ and $`\rho `$ is a principal Higgs bundle with structure group $`G_Z`$.
When $`G_Z`$ is the constant sheaf of groups $`GL(n)`$ over $`Z`$, the openness of the set of points which are semistable with vanishing Chern classes, is well-known. The difficulty here is to extend this to the case of structure group sheaf $`G_Z`$ which is flat and linear but not necessarily constant.
Let $`G_ZGL(E/Z)`$ be the faithful representation given by the condition that $`G_Z`$ is linear. Let $`E(\rho )`$ be the associated Higgs bundle on $`X_{Dol}\times Z/Z`$. If $`zZ`$ is a point then we obtain the Higgs bundle $`E(\rho )_z`$ which is associated to the principal Higgs $`G_z`$-bundle $`\rho _z`$ by the faithful representation $`G_zGL(E_z)`$. Recall that $`\rho _z`$ is said to be semistable if the Higgs bundles associated to every representation of $`G_z`$ are semistable; however, recall from p. 86 that if $`\rho _z`$ has vanishing Chern classes, then it suffices to check this for one faithful representation of $`G_z`$. Therefore the set of points $`z`$ such that $`\rho _z`$ is semistable with vanishing Chern classes, is the same as the set of points where $`\rho _z`$ has vanishing Chern classes and where $`E(\rho )_z`$ is semistable. The set of points where $`E(\rho )_z`$ is semistable, is open cf . Thus we may work inside this set, so we may assume from now on that $`E(\rho )_z`$ is semistable for all $`zZ`$.
Let $`PX\times Z`$ be the geometric principal $`G_Z`$-bundle underlying the principal Higgs bundle $`\rho `$. We have to prove that the set of points $`zZ`$ where $`P_z=P|_{X\times \{z\}}`$ has vanishing Chern classes, is open.
It is possible to define the Chern classes in de Rham cohomology in an algebraic way, so the set of points where $`P_z`$ has vanishing Chern classes is a constructible set. To prove openness it suffices therefore to prove that it is open in the usual topology on $`Z`$ (recall that we work over $`Spec(𝐂)`$). So chose a point $`z`$ such that $`P_z`$ has vanishing Chern classes. In view of the semistability hypothesis, we obtain that the principal Higgs bundle $`\rho _z`$ corresponds to a flat principal $`G_z`$-bundle via the correspondence between Higgs bundles and local systems . This correspondence preserves the $`C^{\mathrm{}}`$ type of the bundle, so we get that the $`C^{\mathrm{}}`$ principal $`G_z`$-bundle on $`X`$ underlying $`\rho _z`$, i.e. $`P_z`$, is flat. By Deligne and Sullivan , there is a finite etale covering $`X^{}X`$ such that $`P_z|_X^{}`$ is a trivial principal bundle, i.e. has a $`C^{\mathrm{}}`$ section $`s_z`$.
Here is where we use the flatness hypothesis on $`G_Z`$. In characteristic zero flatness implies smoothness for group schemes, so $`G_Z`$ is smooth over $`Z`$. Since $`P`$ is modelled locally on $`G_Z`$, we obtain that the morphism $`PX\times Z`$ is smooth. Let $`Z^{}`$ be a neighborhood of $`z`$ in the usual topology in $`Z`$, such that there exists a retraction of $`Z^{}`$ to $`\{z\}`$ where the trajectories are smooth curves in $`Z^{}`$. Assume that the retraction is obtained by following a vector field along these curves; and assume that over $`Z^{}`$ we have lifted the vector field to a vector field on $`P`$ (projecting to zero in the direction $`X`$ of the base $`X\times Z`$). Such a neighborhood exists.
Let $`P^{}`$ be the pullback of $`P`$ to $`X^{}\times Z^{}`$. Following our lifts of vector fields on $`P^{}`$ gives a trivialization of the underlying differential manifold, $`P^{}(P_z\times _XX^{})\times Z^{}`$. Via this trivialization, we can extend our section $`s_z`$ to a section $`s`$ of the bundle $`P^{}`$. In particular, for every point $`yZ^{}`$ the $`C^{\mathrm{}}`$ principal $`G_y`$-bundle $`P_y^{}`$ on $`X^{}`$ is trivial. In particular it has vanishing Chern classes in rational cohomology. The morphism $`H^i(X,𝐐)H^i(X^{},𝐐)`$ is injective, so we obtain that for every point $`yZ^{}`$ the principal $`G_y`$-bundle $`P_y`$ has vanishing Chern classes on $`X`$. Thus $`Z^{}`$ provides the neighborhood of $`z`$ necessary to show that the set of points where $`P_y`$ has vanishing Chern classes, is open. $`|||`$
Counterexample: The hypothesis of flatness (i.e. smoothness) of $`G_Z`$ is essential in the above argument; here is a counterexample showing that we cannot remove it. Fix a stable vector bundle $`E_0`$ with vanishing Chern classes for structure group $`GL(n)`$ over a curve $`X`$. The bundle $`E`$ has a Borel reduction, i.e. a flag of sub-line-bundles, such that the degrees of the line bundles are nonzero. Let $`B`$ denote the Borel subgroup which is the structure group for this reduction. Let $`Z=𝐀^1`$ and let $`G_Z`$ be a group scheme over $`Z`$, contained in $`GL(n)\times Z`$, with generic fiber $`B`$ (over every point except the origin) and special fiber $`GL(n)`$ over the origin. Note that $`G_Z`$ isn’t flat. Let $`E`$ be the bundle $`E_0\times Z`$ over $`X\times Z`$, but considered as a principal $`G_Z`$-bundle using the $`B`$-reduction for $`E_0`$ outside the origin. The fiber over the origin is the principal $`GL(n)`$-bundle $`E_0`$ which we started with; it is semistable with vanishing Chern classes. However for any other point $`z0`$ the principal $`G_z=B`$-bundle $`E_z`$ doesn’t have vanishing Chern classes (it isn’t semistable either). Thus the set of points of $`Z`$ where the bundle is semistable with vanishing Chern classes, is reduced to $`\{0\}`$ which isn’t open.
Next we turn to the problem of local geometricity. This part of the discussion doesn’t use semistability or our special situation for $`/B`$; however we do use the hypothesis that the fundamental group sheaf is a flat linear group scheme.
###### Proposition 24
Suppose $`/B`$ is a formal category which is projective and of smooth type over a base $`n`$-stack $`B`$, with finite cohomological dimension for vector sheaf coefficients. Suppose that $`V/B`$ is a relatively connected geometric $`n`$-stack such that for any scheme $`ZB`$ with basepoint section $`v:ZV_Z`$, the fundamental group sheaf $`G_Z=\pi _1(V_Z/Z,v)`$ is represented by a flat linear group scheme. Then $`\underset{¯}{Hom}(/B,V/B)`$ is a locally geometric $`n`$-stack.
Proof: Suppose $`ZB`$ is a morphism from a scheme. By going to an etale neighborhood we may assume that $`V_Z`$ has a basepoint section $`v`$, so we obtain the group sheaf $`G_Z`$ (which by hypothesis is a flat linear group scheme over $`Z`$) and the first stage in the Postnikov tower for $`V_Z`$ is $`K(G_Z/Z,1)`$. Let
$$M_Z:=\underset{¯}{Hom}(_Z/Z,K(G_Z/Z,1)/Z),$$
which is the moduli $`1`$-stack for principal $`G_Z`$-bundles over $`_Z`$. By , Proposition 10.4.4, $`M_Z`$ is locally geometric (i.e. it is an Artin algebraic $`1`$-stack which is locally of finite type). We will show that the morphism
$$p:\underset{¯}{Hom}(/B,V/B)\times _BZM_Z$$
is geometric, yielding the desired conclusion. For this, we may suppose that $`\rho _Z:ZM_Z`$ is a point, and it suffices to show that the fiber of $`p`$ over $`\rho _Z`$ is a geometric $`n`$-stack over $`Z`$. The point $`\rho _Z`$ may also be considered as a morphism $`\rho _Z:_ZK(G_Z/Z,1)`$ or equivalently a principal $`G_Z`$-bundle over $`_Z`$. Let $`U_Z`$ be the fiber in the fibration sequence
$$U_ZV_ZK(G_Z/Z,1).$$
It is a simply connected geometric $`n`$-stack with action of the group $`G_Z`$. Twisting this $`n`$-stack by the principal $`G_Z`$-bundle $`\rho _Z`$ we obtain a relatively $`1`$-connected geometric $`n`$-stack
$$U_Z(\rho _Z)_Z.$$
The fiber of $`p`$ over $`\rho `$ is the relative section stack
$$\mathrm{\Gamma }(_Z/Z,U_Z(\rho _Z)).$$
We have to prove that this is a geometric $`n`$-stack over $`Z`$. This is a twisted version of the statement 10.3.3 of , and we follow the idea of the proof that was sketched there.
The considerations we discuss here are fundamentally well-known to algebraic topologists.
In order to write the proof in an efficient way, we work with $`\mathrm{}`$-stacks (which we assume without further stating this to be stacks of $`\mathrm{}`$-groupoids or equivalently simplicial presheaves). Suppose $`f:RZ`$ is a simply connected $`\mathrm{}`$-stack over a base scheme $`Z`$. We say that $`R`$ is cohomologically geometric if $`𝐑f_{}(𝒪)`$ is a perfect complex (supported in $`[0,\mathrm{})`$) over $`Z`$. We say that $`R`$ is residually geometric if for any $`N`$ the relative truncation $`\tau _N(R/Z)`$ is an $`n`$-almost geometric $`n`$-stack in the terminology of §7.3. Recall the result 7.3.5 of which, in terms of our present definitions, implies that if $`R`$ is cohomologically geometric then it is residually geometric. (The converse is probably also true but we don’t have a proof.) Recall also 7.3.1 saying that if $`R`$ is $`n`$-truncated relative to $`Z`$ and if $`R`$ is residually geometric then $`R`$ is in fact geometric.
We note the following canonical construction (this construction is the “Goodwillie tower of the identity functor” see ). Define
$$\mathrm{\Sigma }_0:=\mathrm{\Omega }^{\mathrm{}}S^{\mathrm{}}(R/Z)$$
to be the Dold-Puppe of the dual $`𝐑f_{}(𝒪)^{}`$. There is a natural map
$$R\mathrm{\Omega }^{\mathrm{}}S^{\mathrm{}}(R/Z)=\mathrm{\Sigma }_0.$$
If $`R`$ is cohomologically geometric then $`\mathrm{\Sigma }_0`$ is the Dold-Puppe of a perfect complex so it again is cohomologically geometric. Let $`R_1`$ be the fiber of the above map. It fits into a fibration sequence
$$\mathrm{\Omega }\mathrm{\Sigma }_0TR,$$
so an easy spectral sequence argument shows that $`T`$ is also cohomologically geometric. Iterating this construction we obtain a sequence of cohomologically geometric $`\mathrm{}`$-stacks $`R_i`$ with fibration sequences
$$R_{i+1}R_i\mathrm{\Sigma }_i$$
where the $`\mathrm{\Sigma }_i`$ are the Dold-Puppe of perfect complexes. Furthermore the $`R_i`$ eventually become more and more connected, i.e. there is a sequence $`k_i\mathrm{}`$ such that $`R_i`$ are $`k_i`$-connected.
The above constructions relative to a base scheme are canonical (with canonical homotopy coherence at all orders) so they relativize: if $`B`$ is any $`\mathrm{}`$-stack and if $`R/B`$ is an $`\mathrm{}`$-stack which is cohomologically geometric (i.e. for any scheme $`ZB`$ the pullback $`R_Z=R\times _BZ`$ is cohomologically geometric over $`Z`$) then we obtain a sequence of $`\mathrm{}`$-stacks $`R_i/B`$ fitting into fibration sequences as above, with $`\mathrm{\Sigma }_i`$ again being the relative Dold-Puppe of perfect complexes relative to $`B`$, the $`R_i`$ are cohomologically geometric, and they are more and more connected relative to $`B`$.
We now apply this construction to the study of relative section stacks. Suppose $`_ZZ`$ is a morphism and suppose $`R`$ is a cohomologically geometric $`\mathrm{}`$-stack over $`_Z`$. We obtain a morphism
$$\mathrm{\Gamma }(_Z/Z,R)\mathrm{\Gamma }(_Z/Z,\mathrm{\Sigma }_0).$$
We will treat below the case of sections of the Dold-Puppe of a perfect complex; assume for now that these are known to be residually geometric. Thus $`\mathrm{\Gamma }(_Z/Z,\mathrm{\Sigma }_0)`$ is residually geometric. The fiber of the above morphism over any section of $`\mathrm{\Gamma }(_Z/Z,\mathrm{\Sigma }_0)`$ is of the form $`\mathrm{\Gamma }(_Z/Z,R_1^{})`$ for an $`\mathrm{}`$-stack $`R_1^{}`$ over $`_Z`$ which, over any scheme mapping to $`_Z`$, locally looks like the fiber $`R_1`$ defined above. Thus the fiber $`R_1^{}`$ is cohomologically geometric relative to $`_Z`$ and we can again iterate the construction. We assumed that $`_Z`$ has finite cohomological dimension relative to $`Z`$ for vector sheaf coefficients. From this assumption, for any fixed $`N`$ the later terms in the iteration eventually become irrelevant to $`\tau _N\mathrm{\Gamma }(_Z/Z,R)`$, and up to these later terms we have expressed $`\mathrm{\Gamma }(_Z/Z,R)`$ as the result of a process of fibrations where the base is always a residually geometric $`\mathrm{}`$-stack (this remains to be treated below), and the fiber is again decomposed further in the same way. If the base and fibers are residually geometric then the total space is residually geometric (this is stated with only an indication of the proof in Lemma 7.3.3 but we leave it as an exercise here too). Finally we obtain that $`\tau _N\mathrm{\Gamma }(_Z/Z,R)`$ is a truncation of a residually geometric $`\mathrm{}`$-stack, thus it is an $`N`$-almost geometric $`N`$-stack.
If the original $`R`$ was $`n`$-truncated relative to $`_Z`$ i.e. if it was an $`n`$-stack relative to $`_Z`$ then choosing $`Nn+1`$ and applying Lemma 7.3.1, we obtain finally that
$$\mathrm{\Gamma }(_Z/Z,R)=\tau _N\mathrm{\Gamma }(_Z/Z,R)$$
is a geometric $`n`$-stack relative to $`Z`$. This applies to our twisted $`U_Z(\rho _Z)`$ to yield that
$$\mathrm{\Gamma }(_Z/Z,U_Z(\rho _Z))$$
is a geometric $`n`$-stack over $`Z`$.
To complete the proof we have to treat the case of coefficients in a perfect complex. If $`C`$ is a perfect complex over $`_Z`$ supported in $`(\mathrm{},0]`$ and if $`\mathrm{\Sigma }=DP(C/_Z)`$ then
$$\mathrm{\Gamma }(_Z/Z,\mathrm{\Sigma })=DP\tau ^0\mathrm{\Gamma }(_Z/Z,C).$$
Thus, if we can show that $`\mathrm{\Gamma }(_Z/Z,C)`$ is a perfect complex (supported in $`(\mathrm{},\mathrm{})`$) then we obtain that $`\mathrm{\Gamma }(_Z/Z,\mathrm{\Sigma })`$ is residually geometric as desired. In fact it suffices to show that $`\mathrm{\Gamma }(_Z/Z,C)`$ is residually perfect cf Definition 7.3.4.
Now $`\mathrm{\Gamma }(_Z/Z,C)`$ is calculated by the relative de Rham complex for $`_Z/Z`$, twisted by $`C`$. There is a spectral sequence whose starting terms are the cohomology of $`X/Z`$ with coefficients in $`C`$ tensored with the vector bundles appearing in the de Rham complex for $`_Z`$ (here $`X`$ denotes the scheme underlying the formal category $`_Z`$). Thus, in order to obtain the desired result it suffices to consider a flat morphism of schemes $`g:XZ`$ and to prove that if $`D`$ is a perfect complex on $`X`$ supported in $`(\mathrm{},m]`$ then the direct image $`\mathrm{\Gamma }(X/Z,D)=𝐑g_{}(D)`$ is a residually perfect complex on $`Z`$. The residually perfect condition means that for any $`N`$ the truncation $`\tau ^{[N,N]}𝐑g_{}(D)`$ into the interval $`[N,N]`$ should be quasiisomorphic to the truncation of a perfect complex. To prove this suppose $`k`$ is the cohomological dimension of $`X/Z`$ and choose an explicit resolution $`L`$ of $`D`$ by vector bundles in the interval $`[Nk,m]`$. In other words choose a complex of vector bundles $`L`$ (supported in $`[Nk1,m]`$ say) with a morphism $`LD`$ which is exact in the interval $`[Nk,m]`$. This is possible since $`X/Z`$ is flat and projective. Now
$$\tau ^{[N,N]}𝐑g_{}(L)\tau ^{[N,N]}𝐑g_{}(D)$$
is a quasiisomorphism. Since $`L`$ is a complex of vector bundles over $`X`$, the well-known fact that the higher direct image of a vector bundle under a flat morphism is a perfect complex, implies that $`\tau ^{[N,N]}𝐑g_{}(D)`$ is the truncation of a perfect complex. This being true for any $`N`$ we obtain that $`𝐑g_{}(D)`$ is a residually perfect complex. This completes the last part of the proof. $`|||`$
Finally we go back to our original situation and use the openness of semistability as well as condition (3) to get a geometric $`n`$-stack of semistable morphisms.
###### Proposition 25
Suppose $`/B`$ is a formal category which is projective and of smooth type over a base $`1`$-stack $`B`$. Suppose that we are in the situation of (1)-(3) above (with $`B`$ geometric). Suppose $`V/B`$ is a geometric $`n`$-stack which is relatively connected, and whose relative fundamental group object is a flat linear group scheme (as above this condition is made relative to every scheme $`ZB`$ such that a basepoint section $`v`$ of $`V_Z`$ exists). Then $`\underset{¯}{Hom}^{\mathrm{se},0}(/B,V/B)`$ is a geometric $`n`$-stack over $`B`$.
Proof: By Proposition 24, the $`n`$-stack $`\underset{¯}{Hom}(/B,V/B)`$ is locally geometric. Furthermore, the proof of that proposition actually gives the stronger statement that if $`\tau _1(V/B)`$ denotes the first stage in the Postnikov tower, then
$$\underset{¯}{Hom}(/B,\tau _1(V/B)/B)$$
is locally geometric, and the morphism
$$\underset{¯}{Hom}(/B,V/B)\underset{¯}{Hom}(/B,\tau _1(V/B)/B)$$
is geometric. Proposition 23 says that
$$\underset{¯}{Hom}^{\mathrm{se},0}(/B,V/B)\underset{¯}{Hom}(/B,V/B)$$
is an open substack; thus it too is locally geometric. Furthermore it is pulled back from the open substack
$$M:=\underset{¯}{Hom}^{\mathrm{se},0}(/B,\tau _1(V/B)/B)\underset{¯}{Hom}(/B,\tau _1(V/B)/B).$$
Thus we have a geometric morphism
$$\underset{¯}{Hom}^{\mathrm{se},0}(/B,V/B)\underset{¯}{Hom}(/B,\tau _1(V/B)/B)=M$$
between locally geometric $`n`$-stacks, and it suffices to prove that the target $`M`$ of this morphism is geometric, i.e. of finite type. The reader may be happy to note that $`M`$ is only a $`1`$-stack—thus it is an Artin algebraic stack locally of finite type over $`B`$ and we would like to say that it is of finite type over $`B`$.
We know that over every point $`bB`$, the fiber $`M_b`$ (which is just a moduli stack of semistable principal bundles with vanishing Chern classes on $`X_{Dol}`$ or $`X_{DR}`$) is of finite type. Normally, this type of pointwise information for each fiber isn’t sufficient to conclude that the total stack is of finite type over the base. However, we have the hypothesis (3) that there are only finitely many equivalence classes of points in the base $`B`$, and this allows us to obtain the desired conclusion.
Let $`Z=_{iI}Z_i`$ be a disjoint union of schemes $`Z_i`$ of finite type, with a smooth surjection
$$ZM.$$
Let $`B(Spec(𝐂))`$ denote the finite set of points of $`B`$, and for each $`bB(Spec(𝐂))`$ let $`M_b`$ denote the preimage in $`M`$. We know (see ) that each $`M_b`$ is of finite type, since it is a moduli stack of semistable principal Higgs $`G_b`$-bundles of vanishing Chern classes on $`X`$ (if $`bB_1`$) or of principal $`G_b`$-bundles with flat connection (if $`bBB_1`$). In particular, there is a finite subset of indices $`I^{}I`$ such that for any point $`b`$, $`M_b`$ is covered by the images of the $`X_i`$ for $`iI^{}`$. Let $`X^{}:=_{iI^{}}X_i`$ (it is a scheme of finite type). The map $`X^{}M`$ is smooth, and for any $`Spec(𝐂)`$-valued point $`mM(Spec(𝐂))`$ we have that the preimage of $`m`$ in $`X^{}`$ is nonempty. Thus, for any scheme $`Y`$ mapping to $`M`$, the map $`X^{}\times _MYY`$ has is smooth with the same pointwise surjectivity property; but $`X^{}\times _MY`$ is a geometric $`n`$-stack so this means that the map $`X^{}\times _MYY`$ is a smooth surjection. Therefore $`X^{}M`$ is a smooth surjection, and $`M`$ is of finite type. This completes the proof of the proposition. $`|||`$
As a corollary we obtain the following statement.
###### Theorem 6
Suppose $`𝒱`$ is a pre-namhs (resp. linearized pre-namhs) which is connected, and such that the fundamental group objects such as
$$\pi _1(Tot^{W,F}(V_{DR})/𝒜_{wt}\times 𝒜_{hod})$$
are flat linear group schemes. Suppose $`X`$ is a smooth projective variety. Then the object $`=\underset{¯}{Hom}(X_M,𝒱)`$ is a pre-namhs (resp. linearized pre-namhs).
Proof: In the construction of $``$ given at the start of this chapter, the only open question was whether $`Tot^{W,F}(H_{DR})`$ was a geometric $`n`$-stack over $`𝒜_{wt}\times 𝒜_{hod}`$. Recall that
$$Tot^{W,F}(H_{DR}):=\underset{¯}{Hom}^{\mathrm{se},0}(\frac{X_{Hod}\times 𝒜_{wt}}{𝒜_{wt}\times 𝒜_{hod}},\frac{Tot^{W,F}(V_{DR})}{𝒜_{wt}\times 𝒜_{hod}}).$$
The hypothesis of the theorem says exactly that the hypothesis of Proposition 25 applies, so that proposition gives geometricity of $`Tot^{W,F}(H_{DR})`$. The rest of the construction of $``$, including the treatment of the linearization, is as indicated at the start of the chapter. $`|||`$
The basic conjectures
We can now formulate our basic conjectures concerning nonabelian mixed Hodge structures. These three conjectures 1, 2 and 3 are all statements we expect that one could prove within a medium-range amount of time. We decided to post the current version of the paper as is, without proofs of these conjectures, because it seems clear (e.g. from the difficulty which we already encountered just treating the mixed Hodge structures on the homotopy groups $`\pi _i(𝒱)`$) that their proofs will require a significant amount of technical material and it didn’t seem necessary to burden the reader—or ourselves—with that at present.
###### Conjecture 1
Suppose $`X`$ is a connected smooth projective variety. Suppose $`𝒱`$ is a nonabelian mixed Hodge structure. Then the linearized pre-namhs $`=\underset{¯}{Hom}(X_M,𝒱)`$ is a nonabelian mixed Hodge structure.
Some of the main properties ($`\mathrm{𝐒𝐭𝐫}`$ and $`\mathrm{𝐌𝐇𝐂}`$ above) are given by Proposition 22, at least under a reasonable hypothesis on $`𝒱`$. The main things which remain to be shown are: (I) flatness of the Hodge filtration (condition $`\mathrm{𝐅𝐥}`$) and the conditions $`\mathrm{𝐀𝟏}`$ and $`\mathrm{𝐀𝟐}`$ on the annihilator ideals; and (III) the third Hodge-theoretic condition $`\mathrm{𝐀𝟑}`$ on the annihilator ideals.
A subsidiary conjecture is the following one.
###### Conjecture 2
Suppose $`𝒱`$ and $`𝒴`$ are nonabelian mixed Hodge structures, and suppose that $`𝒴`$ is $`1`$-connected (i.e. its $`\pi _0`$ and $`\pi _1`$ are trivial). Then there is a nonabelian mixed Hodge structure $`\underset{¯}{Hom}(𝒴,𝒱)`$ (with natural choices for the Hodge and weight filtrations), again functorial in $`𝒱`$ and (contravariantly) in $`𝒴`$.
Denote by $``$ the trivial nonabelian mixed Hodge structure where all objects are a single point (objects relative to a base such as $`𝒜_{wt}\times 𝒜_{hod}`$ are equal to that base). If $`𝒱`$ is a namhs then a morphism $`𝒱`$ is the nonabelian analogue of a Hodge class of type $`(0,0)`$ in $`𝒱`$. Thus we shall call such a point a Hodge class in $`𝒱`$.
The above internal $`\underset{¯}{Hom}`$ should be compatible with the notion of morphism, in that a morphism from $`𝒴`$ to $`𝒱`$ should be the same thing as a Hodge class in $`\underset{¯}{Hom}(𝒴,𝒱)`$. Getting back to the first conjecture, in cases where it holds we can define a morphism from $`X_M`$ to $`𝒱`$ as being a morphism $`\underset{¯}{Hom}(X_M,𝒱)`$.
We now state another conjecture about representability in the simply connected case. This conjecture (if it turns out to be true) states how to define the “mixed Hodge structure on the homotopy type of $`X`$”.
###### Conjecture 3
Suppose $`X`$ is a simply connected smooth projective variety. Then there is a universal morphism to a simply connected nonabelian mixed Hodge structure
$$X_M𝒴=𝒮(X)$$
with the property that for any nonabelian mixed Hodge structure $`𝒱`$ the resulting morphism
$$\underset{¯}{Hom}(𝒴,𝒱)\underset{¯}{Hom}(X_M,𝒱)$$
is an equivalence. Furthermore this representing object specializes to the $`n`$-stack representing the de Rham (resp. Betti) shape of $`X`$ (cf ), i.e.
$`Y_{DR}`$ is the $`n`$-stack representing the very presentable shape of $`X_{DR}`$;
$`Tot^F(Y_{DR})`$ is the Hodge filtration on the de Rham representing object, as constructed in ; and
$`Y_{B,𝐂}`$ is the $`n`$-stack representing the very presentable shape of $`X_B`$.
Finally, the mixed Hodge structures on the homotopy vector spaces $`\pi _i^{\mathrm{nu}}(𝒴)=\pi _i(X_B)𝐂`$ coincide with those defined by Morgan and Hain.
It is a consequence of the functoriality in the first two conjectures and the universality in this last conjecture that $`𝒮(X)`$ would be functorial in $`X`$. In other words, if $`XZ`$ is a morphism of smooth projective varieties then we would get a morphism of nonabelian mixed Hodge structures
$$𝒮(X)𝒮(Z)$$
or equivalently a Hodge class in $`\underset{¯}{Hom}(X_M,𝒮(Z))`$. It is natural to ask which morphisms of nonabelian mixed Hodge structures come from morphisms of varieties. This is a subtle question which could be viewed as a nonabelian analogue of the Hodge conjecture:
—do all Hodge classes in $`\underset{¯}{Hom}(X_M,𝒮(Z))`$ come from morphisms of varieties $`XZ`$?
In the non-simply connected case, there cannot exist a universal map such as given by Conjecture 3. Nonetheless, one might try to get back from the nonabelian cohomology to a mixed Hodge structure on the “higher Malcev completion” such as was proposed at the end of .
Finally we state a conjecture which is intended to show how the condition $`\mathrm{𝐀𝟏}`$ fits in with the result of Deligne-Griffiths-Morgan-Sullivan that smooth projective varieties are formal.
###### Conjecture 4
Suppose $`X`$ is a simply connected finite CW complex, and let $`X_B`$ denote the constant $`\mathrm{}`$-stack with values the Poincaré $`\mathrm{}`$-groupoid of $`X`$. Then the following two properties are equivalent:
(1) for all $`n`$ and for any $`1`$-connected namhs $`𝒱`$, the pre-weight filtered $`n`$-stack
$$\underset{¯}{Hom}(X_B,(V_{B,𝐂},W))$$
satisfies condition $`\mathrm{𝐀𝟏}`$ (cf page 3);
(2) the dga associated to $`X`$ by the theory of Sullivan and Morgan, is formal.
Some vague calculations with dga’s seem to support this conjecture.
Recall in a similar vein the result of the exercise on p. 8, saying that if $`Y`$ is a quadratic cone at a point $`y`$ then $`(Y,W(y))`$ satisfies $`\mathrm{𝐀𝟏}`$. One could also conjecture that that result is “if and only if”, and extend the above conjecture to the non-simply connected case. We hope to treat this type of question in a future paper.
If these statements turn out to be substantially true, then one could say that the properties of formality; of degeneration of spectral sequences at $`E_2`$; and the property $`\mathrm{𝐀𝟏}`$ are all sort of the same thing.
Variations of nonabelian mixed Hodge structure
Suppose $`S`$ is a smooth base scheme. We will sketch a definition of variation of nonabelian mixed Hodge structure (vnamhs) over $`S`$. The basic idea is to combine the notion of Griffiths transversality for the Hodge filtration that was defined using the object $`S_{Hod}`$ in , with the weight filtration as we are defining it in the present paper.
Most aspects of what we say in this section will be highly conjectural. Nonetheless we feel that the reader might legitimately be interested to know about a likely direction of future development.
Recall that $`S_{Hod}𝒜_{hod}`$ is a formal category whose fiber over $`[1]`$ is $`S_{DR}`$ and whose fiber over $`[0]`$ is $`S_{Dol}`$ (recall that starting from the $`𝐆_m`$-equivariant formal category over $`𝐀_{hod}^1`$ defined in , we took the quotient object by $`𝐆_m`$ to get a formal category over $`𝒜_{hod}`$ cf p. Nonabelian mixed Hodge structures). Let $`OBJ`$ be one of the $`n+1`$-stacks we consider in the present paper, and look at objects in $`OBJ`$. An object $`\underset{¯}{𝒰}S_{Hod}`$, which may also be seen as a functor
$$\underset{¯}{𝒰}:S_{Hod}OBJ,$$
restricts over $`S_{DR}`$ to an object $`\underset{¯}{U}_{DR}`$ (or $`\underset{¯}{U}_{DR}:S_{DR}OBJ`$), which is the same thing as an object over $`S`$ together with a flat connection. On the other hand, for every point $`sS`$ we obtain a morphism
$$𝒜_{hod}=\{s\}_{Hod}S_{Hod},$$
so the restriction of $`𝒰`$ to $`\{s\}_{Hod}`$ gives a filtration $`F(s)`$ of the fiber $`U_{DR}(s)`$. The fact that $`𝒰`$ is a family over $`S_{Hod}`$ encapsulates Griffiths transversality for this family of filtrations together with the connection. This is explained in somewhat more detail in .
Apply the above discussion to the case where $`OBJ=F^{DP}.nGEOM`$ is the $`n+1`$-stack of pre-weight-filtered geometric $`n`$-stacks. An object
$$S_{Hod}F^{DP}.nGEOM$$
gives a family of linearized weight filtrations with a flat connection over $`S`$, and with a varying family of Hodge filtrations (including Hodge filtrations on the perfect complexes of the linearizations) such that the Hodge filtrations satisfy Griffiths transversality for the weight filtration. Denote the above morphism by
$$(\underset{¯}{U}_{Hod},W,LGr^W):S_{Hod}F^{DP}.nGEOM.$$
The restriction over $`S_{DR}`$ is denoted $`(\underset{¯}{U}_{DR},W,LGr^W)`$.
Note that the “Hodge filtration” which we would have called $`F`$, is integrated into the object $`U_{Hod}`$. Thus it is reasonable to think of the object $`\underset{¯}{U}_{Hod}`$ as being $`(\underset{¯}{U}_{DR},F)`$. On a technical level, however, $`F`$ is not a filtration of the object $`\underset{¯}{U}_{DR}`$ as this would imply that $`F`$ were flat with respect to the connection. The object $`\underset{¯}{U}_{Hod}`$ represents a family of filtrations $`F(s)`$ on the fibers $`\underset{¯}{U}_{DR}(s)`$, satisfying Griffiths transversality with respect to the connection. We will however sometimes use the abuse of notation and call this $`(\underset{¯}{U}_{DR},F)`$. This takes care of the part $`(\underset{¯}{U}_{DR},W,F)`$ together with its linearization $`L`$, in the notion of vnamhs.
For the notion of pre-vnamhs we will want to drop the existence of a linearization, in this case we just look at a morphism
$$(\underset{¯}{U}_{Hod},W):S_{Hod}F.nGEOM.$$
If $`sS`$ is a point, then the restriction of $`(\underset{¯}{U}_{Hod},W)`$ to $`\{s\}_{Hod}𝒜_{hod}`$ becomes an object
$$𝒜_{hod}F.nGEOM,$$
i.e. an object of $`F.F.nGEOM`$. This is just the part which was denoted by $`(\underset{¯}{U}_{DR}(s),F(s),W(s))`$ in the notion of pre-namhs. In the linearized case we get the same plus a linearization $`(LGr^{W(s)}(\underset{¯}{U}_{DR}(s)),F(s))`$.
We next look at the “Betti” part of the definition. Recall again that $`S_{B,𝐑}`$ is the constant stack with values the Poincaré groupoid of $`S^{\mathrm{top}}`$. Since this will be used as the base for a family of $`n`$-stacks, we need to go to the $`n+1`$-truncation of $`S^{\mathrm{top}}`$, in other words we let $`S_{B,𝐑}`$ denote the constant $`n+1`$-stack on the real site, whose values are $`\mathrm{\Pi }_{n+1}(S^{\mathrm{top}})`$. Now the piece $`(\underset{¯}{U}_{B,𝐑},W,L)`$ in the notion of vnamhs, corresponds to a functor of $`n+1`$-stacks
$$(\underset{¯}{U}_{B,𝐑},W,LGr^W):S_{B,𝐑}F^{DP}.nGEOM.$$
This should be seen as a local system over $`S^{\mathrm{top}}`$ of linearized filtered objects. Again, for the notion of pre-vnamhs we will drop the existence of the linearization and just look at
$$(\underset{¯}{U}_{B,𝐑},W):S_{B,𝐑}F.nGEOM.$$
If $`sS`$ is a point then $`\{s\}_{B,𝐑}`$ is just $``$ (i.e. $`Spec(𝐑)`$) and the restriction of our functor to $`\{s\}_{B,𝐑}`$ is just a real filtered geometric $`n`$-stack $`(\underset{¯}{U}_{B,𝐑}(s),W(s))`$.
Finally, recall that we have a morphism
$$\zeta _S:S_{DR}^{\mathrm{an}}S_{B,𝐂}^{\mathrm{an}}.$$
Thus we can ask to have an equivalence (between pre-weight filtered $`n`$-stacks over $`S_{DR}^{\mathrm{an}}`$) of the form
$$\zeta _{\underset{¯}{𝒰}}:(\underset{¯}{U}_{DR},W,LGr^W)^{\mathrm{an}}\zeta _S^{}(\underset{¯}{U}_{B,𝐑},W,LGr^W)^{\mathrm{an}}.$$
As above, for the notion of pre-vnamhs we will drop the linearization.
We have now indicated all of the elements going into the notion of a pre-vnamhs $`\underset{¯}{𝒰}`$ over $`S`$: this consists of morphisms
$$(\underset{¯}{U}_{Hod},W):S_{Hod}F.nGEOM$$
and
$$(\underset{¯}{U}_{B,𝐑},W):S_{B,𝐑}F.nGEOM,$$
together with an analytic equivalence of filtered $`n`$-stacks over $`S_{DR}^{\mathrm{an}}`$,
$$\zeta _{\underset{¯}{𝒰}}:(\underset{¯}{U}_{DR},W)^{\mathrm{an}}\zeta _S^{}(\underset{¯}{U}_{B,𝐑},W)^{\mathrm{an}}.$$
A linearized pre-vnamhs $`\underset{¯}{𝒰}`$ is a pre-vnamhs with a linearization in the sense that the object $`Gr^W(\underset{¯}{𝒰})`$ is provided with an equivalence to the Dold-Puppe of a “split pre-vmhc” (i.e. an object similar to the above but with $`nGEOM`$ replaced by $`nHPERF`$), or equivalently we include the linearizations in the above data and ask for
$$(\underset{¯}{U}_{Hod},W,LGr^W):S_{Hod}F^{DP}.nGEOM$$
and
$$(\underset{¯}{U}_{B,𝐑},W,LGr^W):S_{B,𝐑}F^{DP}.nGEOM,$$
again with an analytic equivalence
$$\zeta _{\underset{¯}{𝒰}}:(\underset{¯}{U}_{DR},W,LGr^W)^{\mathrm{an}}\zeta _S^{}(\underset{¯}{U}_{B,𝐑},W,LGr^W)^{\mathrm{an}}.$$
Suppose $`\underset{¯}{𝒰}`$ is a pre-vnamhs (resp. linearized pre-vnamhs) over $`S`$. As indicated above, if $`sS`$ is a point then the restriction of $`𝒰`$ to the various objects $`\{s\}_{Hod}`$ or $`\{s\}_{B,𝐑}`$ yields a pre-namhs (resp. linearized pre-namhs) $`\underset{¯}{𝒰}(s)`$.
We think that the following definition is adequate. We say that a variation of nonabelian mixed Hodge structure $`\underset{¯}{𝒰}`$ over $`S`$ is a linearized pre-vnamhs $`\underset{¯}{𝒰}`$ such that for every point $`sS`$ the restriction $`\underset{¯}{𝒰}(s)`$ is a nonabelian mixed Hodge structure (i.e. satisfies the axioms of p. Nonabelian mixed Hodge structures), and such that the family of geometric $`n`$-stacks $`Tot^W(\underset{¯}{U}_{Hod})S_{Hod}`$, pulls back via $`SS_{Hod}`$ to a flat family over $`S`$. It is possible that it would also be necessary to apply condition $`\mathrm{𝐀𝟏}`$ “in a family” over $`S`$, but we conjecture that that is a consequence of the simple flatness hypothesis on the whole family.
Relative section stacks
Suppose $`YXB`$ is a pair of morphisms of $`n`$-stacks (with at least $`X`$ and $`B`$ being stacks of groupoids). We can define the “sections of $`Y/X`$ relative to $`B`$” which will be an $`n`$-stack over $`B`$, by the formula
$$\mathrm{\Gamma }(X/B,Y):=$$
$$\underset{¯}{Hom}(\frac{X}{B},\frac{Y}{B})\times _{\underset{¯}{Hom}(\frac{X}{B},\frac{X}{B})}B$$
where the first morphism in the fiber product is the composition with $`YX`$ and the second morphism is the section
$$1_{X/B}:B\underset{¯}{Hom}(\frac{X}{B},\frac{X}{B})$$
corresponding to the identity in each fiber of $`X/B`$.
This has the following universal property. If $`B^{}B`$ is a morphism of $`n`$-stacks, let $`X^{}:=X\times _BB^{}`$ and $`Y^{}:=Y\times _BB^{}`$. Then a lifting
$$B^{}\mathrm{\Gamma }(X/B,Y)B$$
is the same thing as a section $`X^{}Y^{}`$ i.e. a morphism composing to the identity of $`X^{}`$.
We can do the same thing from a slightly different viewpoint. Suppose $`X`$ and $`B`$ are $`n+1`$-stacks (of groupoids) and suppose that $`Y`$ is a cartesian family of $`n`$-stacks over $`X`$, i.e. a functor
$$Y:XnSTACK.$$
Then we again would like to obtain an object $`\mathrm{\Gamma }(X/B,Y)`$ which, this time, should be a cartesian family of $`n`$-stacks over $`B`$. We sketch the construction here; it runs up against a limitation in , so a rigorous treatment would require further work. The reader may skip this discussion; it is only intended to record for future reference what types of things we would like to know in the direction of future developpment of the category-theory side of things.
Recall from that for any $`n+1`$-stack $`A`$ we have an arrow family which is a morphism of $`n+1`$-stacks
$$Arr_A:A^o\times AnSTACK.$$
The further information we need is that this object is functorial in $`A`$. To phrase this precisely, let
$$P,S:(n+1)STACK(n+1)STACK$$
be the functors which respectively associate to an $`n+1`$-stack $`A`$,
$$P(A):=A^o\times A$$
and
$$S(A):=nSTACK$$
(thus $`S`$ is the constant $`n+1`$-stack with values $`nSTACK`$). The statement we need is that the arrow family construction provides a natural transformation $`Arr:PS`$.
Now suppose that $`f:AB`$ is a morphism of $`n+1`$-stacks and $`\alpha ,\beta :BA`$ are sections of $`f`$. Then we obtain the “morphisms in $`A`$ from $`\alpha `$ to $`\beta `$, relative to $`B`$” which is a cartesian family
$$(A/B)_{1/}(\alpha ,\beta ):BnSTACK,$$
constructed as follows using the previous functoriality of $`Arr`$. Think of $`A/B`$ as a cartesian family which we denote
$$\underset{¯}{A}:B(n+1)STACK,$$
and which could be heuristically denoted $`bA(b)`$. Composing with the functors $`P`$ (resp. $`S`$) we obtain cartesian families
$$P\underset{¯}{A}:B(n+1)STACK,bA(b)^o\times A(b),$$
$$S\underset{¯}{A}:B(n+1)STACK,bnSTACK,$$
and $`Arr\underset{¯}{A}`$ is a natural transformation from $`P\underset{¯}{A}`$ to $`S\underset{¯}{A}`$. On the other hand $`_B`$ is the cartesian family which is constant with values the one-point stack, and $`\alpha `$ and $`\beta `$ correspond to natural transformations
$$\alpha ,\beta :_B\underset{¯}{A}.$$
We obtain a natural transformation $`(\alpha ,\beta ):_BP\underset{¯}{A}`$ and composing with the arrow family gives a natural transformation
$$Arr(\alpha ,\beta ):_BS\underset{¯}{A}.$$
Now go back from the point of view of cartesian families of $`n+1`$-stacks over $`B`$, to the point of view of $`n+1`$-stacks mapping to $`B`$: the cartesian family $`_B`$ corresponds to the identity $`BB`$ and the cartesian family $`S\underset{¯}{A}`$ (which is constant with values $`nSTACK`$) corresponds to $`nSTACK\times BB`$. Our natural transformation $`Arr(\alpha ,\beta )`$ corresponds to a morphism of $`n+1`$-stacks from $`B`$ to $`nSTACK\times B`$, and composing with the first projection we obtain the desired
$$(A/B)_{1/}(\alpha ,\beta ):BnSTACK.$$
This construction seems somewhat complicated, for an object which is “supposed to exist” by general principles; and our proposed treatment rests upon the functoriality of $`Arr`$ which remains to be proven. On the other hand, it does seem to be useful for what we are planning to do below. Thus it would seem worthwhile to take a closer look at this situation in the future.
Now we get back to our proposed second construction of the relative sections stack. If $`XB`$ is a morphism of $`n+1`$-stacks (of groupoids) then a cartesian family $`Y:XnSTACK`$ corresponds to a section
$$B\stackrel{Y}{}A:=\underset{¯}{Hom}(\frac{X}{B},\frac{nSTACK\times B}{B})B.$$
The one-point cartesian family $`_X:XnSTACK`$ similarly corresponds to a section of $`A/B`$, and we can put
$$\mathrm{\Gamma }(X/B,Y):=(A/B)_{1/}(_X,Y):BnSTACK$$
using the above construction.
There should, in fact, be a third construction of $`\mathrm{\Gamma }(X/B,Y)`$ in the case where $`B`$ is an $`n+2`$-stack, $`X`$ is a cartesian family over $`B`$, and $`Y/X`$ is a family over $`B`$ of cartesian families over $`X`$. This situation isn’t formalized yet.
As usual, one would like to have a compatibility between these constructions of $`\mathrm{\Gamma }(X/B,Y)`$, but we don’t do that here. For the present, use the first construction which is mathematically sound. The only problem with that is that it takes place when all the stacks are $`n`$-stacks; but in the Betti case we had to use an $`n+1`$-stack $`X_{B,𝐑}`$ for the base of a vnamhs, and technically speaking if we want to take the direct image along $`XS`$ then $`S_{B,𝐑}`$ should be an $`n+2`$-stack. It suffices to think of everybody as $`n+2`$-stacks and apply the first construction.
Extra structure
Suppose now that $`Y`$ has extra structure relative to $`X`$, for example a filtration or bifiltration, eventually with linearization of one of the filtrations. Then $`\mathrm{\Gamma }(X/B,Y)`$ will have the same extra structure. If $`(Y,W,F)`$ is a bifiltered $`n`$-stack over $`X`$ for example, then we obtain a bifiltered section stack denoted either of two ways,
$$(\mathrm{\Gamma }(X/B,Y),W,F)=\mathrm{\Gamma }(X/B,(Y,W,F)):BF.F.nSTACK.$$
If $`Y:XnGEOM`$ is a geometric $`n`$-stack over $`X`$ then the relative section stack might not necessarily be geometric. Under some circumstances it will be geometric. For example if $`X/B`$ comes from a formal category which is projective and of smooth type over $`B`$, then sections 8-10 give some results saying when $`\underset{¯}{Hom}(X/B,Y/B)`$ is geometric. The treatment of concerns the case where $`Y`$ comes from a family over $`B`$; one would like to improve this to treat similar cases where $`Y`$ is a family over $`X`$. This corresponds to the case of local systems along the fibers of $`X/B`$.
The result of this lack of knowledge is that we can’t say that the cartesian families which will appear in our proposed construction of the relative section object of a vnamhs, will be geometric or not.
Relative sections of a vnamhs
We now come to the situation which interests us. Suppose $`f:XS`$ is a projective morphism of smooth varieties. We obtain a morphism of objects $`X_MS_M`$, in other words morphisms between all of the various objects that enter into $`X_M`$ and $`S_M`$, with the appropriate homotopies of compatibility between them.
We can now state the variational version of our basic Conjecture 1.
###### Conjecture 5
Suppose $`\underset{¯}{𝒰}`$ is a (linearized) pre-vnamhs over $`X`$. Then we obtain a (linearized) pre-namhs
$$\underset{¯}{𝒢}=\mathrm{\Gamma }(X_M/S_M,\underset{¯}{𝒰})$$
over $`S`$. The objects going into $`\underset{¯}{𝒢}`$ are supposed to be the relative section stacks of the corresponding objects in $`\underset{¯}{𝒰}`$, for example
$$(\underset{¯}{G}_{Hod},W):=\mathrm{\Gamma }^{\mathrm{se}}(X_{Hod}/S_{Hod},(\underset{¯}{U}_{Hod},W)),$$
and
$$(\underset{¯}{G}_{B,𝐑},W):=\mathrm{\Gamma }(X_{B,𝐑}/S_{B,𝐑},(\underset{¯}{U}_{B,𝐑},W)).$$
In the first case the superscript ‘se’ indicates that we need to restrict to semistable sections over the Dolbeault points $`S_{Dol}S_{Hod}`$.
If $`\underset{¯}{𝒰}`$ is a vnamhs over $`X`$ then $`\underset{¯}{𝒢}`$ should be a vnamhs over $`S`$.
Part III: Computations
Now that we have made our main definitions and described the basic conjectures, we get to stating what we will actually do in the last part of the paper. Our work breaks into two parts:
(A) First we will construct a nonabelian mixed Hodge structure $`𝒱`$ whose underlying homotopy type is the complexified $`2`$-sphere $`S_𝐂^2`$ (cf ). This construction is inspired by the idea that we would like to have
$$𝒱=𝒮(𝐏^1),$$
but we don’t actually prove this latter fact. A little bit more precisely, we can obtain a morphism $`𝐏_M^1𝒱`$ and a number of the properties required in Conjecture 3 will hold, however we don’t treat the main universal property.
(B) We show the first basic Conjecture 1 for the case of $`\underset{¯}{Hom}(X_M,𝒱)`$ for any smooth projective variety $`X`$ and the namhs $`𝒱`$ constructed in (I) above. We hope that this proof will provide some ideas for proving Conjecture 1 in the general case.
Some morphisms between Eilenberg-MacLane pre-namhs
In the next section we will construct a namhs $`𝒱`$. It will be the fiber of a morphism between Eilenberg-MacLane pre-namhs’s. Thus, in this section we will do a general study of morphisms between Eilenberg-MacLane pre-namhs’s. This is basically the “Breen calculations” (see , , ) for pre-namhs’s.
In this section, which prepares for the construction of $`𝒱`$ in the next section, we study some basic morphisms between “abelian” pre-namhs’s. Recall that a pre-mhs is a real vector space $`U_𝐑`$ with real weight filtration $`W`$ and with a “Hodge filtration” on the complexification $`(𝒱_𝐂,F)`$. By applying the construction $`\xi `$ of we obtain a pre-namhs (for $`n0`$)
$$𝒰=\{(U_𝐂,W,F),(U_𝐑,W),\mathrm{\hspace{0.33em}\hspace{0.33em}1}_{U_𝐂}\}.$$
Furthermore, this object is an abelian group object in $`0PNAMHS`$, which allows us to “deloop” it and define for any $`mn`$ the pre-namhs denoted $`K(𝒰,m)`$. One way of constructing this is to let $`𝒰[m]`$ be the complex with $`𝒰`$ in degree $`m`$ and zeros elsewhere; this complex is a pre-mhc, and
$$K(𝒰,m):=DP(𝒰[m]).$$
Recall (p. 1) that $`𝒰`$ is a shifted mixed Hodge structure (of shift $`s`$) if $`F`$ and $`\overline{F}`$ are $`k+s`$-opposed on $`Gr_k^W(U)`$. Recall that $`𝒰[m]`$ will be a mixed Hodge complex (i.e. satisfies condition $`\mathrm{𝐌𝐇𝐂}`$) if and only if $`𝒰`$ is a shifted mhs of shift $`m`$. More generally, $`𝒰[m]`$ will be an $`s`$-shifted mhc if and only if $`𝒰`$ is an $`sm`$-shifted mhs.
We call pre-namhs’s which are formed out of pre-mhs’s in the above way, “Eilenberg-MacLane pre-namhs’s”. We need to know the “Breen calculations” for these guys, namely we would like to calculate the $`\pi _0`$ (or other $`\pi _i`$) of the space of morphisms from $`K(𝒰,m)`$ to $`K(𝒰^{},m^{})`$. Recall that the space of morphisms is denoted as
$$PNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{})).$$
A preliminary remark is that (because of the fact that $`K(,i)`$ represents the delooping) we have
$$\pi _iPNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{}))=$$
$$\pi _0PNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{}i)).$$
Thus we only need to calculate the $`\pi _0`$.
###### Theorem 7
Suppose $`m,m^{}2`$, and suppose that $`m`$ divides $`m^{}`$ with $`m^{}=dm`$. Then for $`m`$ even we have
$$\pi _0PNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{}))=Hom_{MHS}(Sym^d(𝒰),𝒰^{})$$
whereas for $`m`$ odd we have
$$\pi _0PNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{}))=Hom_{MHS}(\stackrel{d}{}(𝒰),𝒰^{}).$$
Here $`Hom_{MHS}`$ denotes the set of morphisms in the category of real mixed Hodge structures.
Proof: We write the proof in the case $`m`$ even; for the case $`m`$ odd say the same thing but with $`Sym^d`$ replaced by $`^d`$.
The space of morphisms in question is a homotopy fiber product of the form
$$A\times _CB$$
where $`A`$ is the space of morphisms
$$Tot^{F,W}(K(𝒰_𝐂,m))Tot^{F,W}(K(𝒰_𝐂^{},m^{}))$$
relative to $`𝒜_{hod}\times 𝒜_{wt}`$, where $`B`$ is the space of morphisms
$$Tot^W(K(𝒰_𝐑,m))Tot^W(K(𝒰_𝐑^{},m^{}))$$
relative to $`𝒜_{wt,𝐑}`$, and where $`C`$ is the space of morphisms
$$Tot^W(K(𝒰_𝐂,m))Tot^W(K(𝒰_𝐂^{},m^{}))$$
relative to $`𝒜_{wt}`$. We claim that these spaces of morphisms have all components simply connected, and have $`\pi _0`$ equal to the spaces of morphisms from $`Sym^d(𝒰)`$ to $`𝒰^{}`$, respectively: for $`A`$, morphisms of bifiltered vector spaces; for $`B`$, morphisms of real filtered vector spaces; and for $`C`$, morphisms of complex filtered vector spaces.
Before proving the claim we recall the relative Breen calculations in general: they say that if $``$ and $`^{}`$ are vector bundles over a base $`B`$ then
$$\pi _i\underset{¯}{Hom}(\frac{K(/B,m)}{B},\frac{K(^{}/B,m^{})}{B})=H^{m^{}i}(K(/B,m)/B,^{})$$
$$=^{}_{𝒪_B}(Sym\text{or})^{\frac{m^{}i}{m}}(^{}).$$
Here the symmetric product is chosen if $`m`$ is even, the exterior product if $`m`$ is odd; and when the exponent is fractional the answer is taken as zero.
Now getting back to the proof of the claim, note first that $`Tot^{W,F}(𝒰)`$ is a vector bundle over $`𝒜_{hod}\times 𝒜_{wt}`$ and similarly for $`𝒰^{}`$; similarly $`Tot^W(𝒰_𝐑)`$ is a real vector bundle over $`𝒜_{wt,𝐑}`$ and the same for $`𝒰_𝐑^{}`$; and finally the previous phrase also holds after complexification of everything. The Breen calculations in the relative case over $`𝒜_{hod}\times 𝒜_{wt}`$, over $`𝒜_{wt,𝐑}`$ or over $`𝒜_{wt}`$, then say that the relative $`\underset{¯}{Hom}`$ from $`K(,m)`$ to $`K(^{},m^{})`$ relative to the appropriate base, has relative $`\pi _0`$ equal to the $`Hom`$ of vector bundles from the symmetric power of the first vector bundle to the second vector bundle; and the relative $`\pi _1`$ vanishes (this is because of the hypothesis $`m2`$ so the terms in the symmetric algebra are spaced out with at least one zero in between them); and the higher $`\pi _i`$ are again vector bundles. Now the bases in question, either $`𝒜_{hod}\times 𝒜_{wt}`$ or $`𝒜_{wt,𝐑}`$ or $`𝒜_{wt}`$, have trivial higher cohomology with coefficients in vector bundles. Thus only the relative $`\pi _0`$ contributes to the $`\pi _0`$ of the space of sections, and the trivial relative $`\pi _1`$ gives trivial $`\pi _1`$ of all components of the space of sections. Thus we get the simply-connectedness part of the claim, and the spaces of sections are the spaces of sections of the bundles given by the Breen calculations, i.e. the vector bundle $`Hom`$’s from the symmetric powers of the bundles for $`𝒰`$, to the bundles for $`𝒰^{}`$. These spaces of sections are indeed the claimed spaces of morphisms from $`Sym^d(𝒰)`$ to $`𝒰^{}`$. This proves the claim.
Now we have identified the space of morphisms in question, as a fiber product $`A\times _CB`$, and we have seen that $`A`$, $`B`$ and $`C`$ have simply connected components. Therefore we have
$$\pi _0(A\times _CB)=\pi _0(A)\times _{\pi _0(C)}\pi _0(B).$$
In view of the above descriptions: of $`\pi _0(A)`$ as the space of bifiltered morphisms from $`Sym^d(𝒰)`$ to $`𝒰^{}`$; of $`\pi _0(B)`$ as the space of real filtered morphisms from $`Sym^d(𝒰)`$ to $`𝒰^{}`$; and of $`\pi _0(C)`$ as the space of complex filtered morphisms from $`Sym^d(𝒰)`$ to $`𝒰^{}`$, we obtain that the fibered product is exactly the space of morphisms of pre-mhs from $`Sym^d(𝒰)`$ to $`𝒰^{}`$. This proves the theorem. $`|||`$
The following theorem could also be stated as a “conjecture” since we have only partially finished the proof; the remainder of the proof is left as an exercise.
###### Theorem 8
Suppose $`m,m^{}2`$, and suppose that $`m^{}=dm1`$. Then for $`m`$ even we have
$$\pi _0PNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{}))=Ext_{MHS}^1(Sym^d(𝒰),𝒰^{})$$
whereas for $`m`$ odd we have
$$\pi _0PNAMHS_{1/}(K(𝒰,m),K(𝒰^{},m^{}))=Ext_{MHS}^1(\stackrel{d}{}(𝒰),𝒰^{}).$$
Here $`Ext_{MHS}^1`$ denotes the $`Ext^1`$ in the category of real mixed Hodge structures. Outside of the present case and the case of the previous theorem, the answer is zero.
Sketch of Proof: Use the same proof as for Theorem 7. The only place where things change is at the line where it is stated that the relative $`\pi _0`$ is given by a certain formula and the relative $`\pi _1`$ vanishes. In the case treated by the last sentence, both the relative $`\pi _0`$ and the relative $`\pi _1`$ of that phrase vanish, so we get that the answer is zero. In the case $`m^{}=dm+1`$ the relative $`\pi _0`$ vanishes and the relative $`\pi _1`$ is given by the relative Breen calculations (it is the same as what occured as relative $`\pi _0`$ in the proof of Theorem 7). The remainder of the proof of Theorem 7 may be followed through, bearing in mind that the homotopy fiber product $`A\times _CB`$ is now a fiber product of connected but not simply connected spaces. This gives a certain formula. We leave as an exercise to the reader to identify the result given by this formula, with the $`Ext`$ group in the category of real mixed Hodge structures (see Beilinson ) as stated in the theorem. $`|||`$
Shifting the weight filtration
Something which will be useful in the next section is the following notation. Let $`D𝒜`$ denote the divisor defined by the origin. If need be we denote the two occurences of this as $`D_{hod}𝒜_{hod}`$ and $`D_{wt}𝒜_{wt}`$. Furthermore we also denote by $`D_{wt}`$ the divisor
$$D_{wt}\times 𝒜_{hod}𝒜_{wt}\times 𝒜_{hod}.$$
If $`𝒰=(U,W)`$ is a filtered vector bundle then we denote by $`𝒰(D)`$ the filtered bundle whose total bundle of the weight filtration is
$$Tot^W(𝒰(D)):=Tot^W(U)_𝒪𝒪_𝒜(D).$$
This corresponds to renumbering or “shifting” the filtration $`W`$, namely
$$W_{m+1}(𝒰(D))=W_m(𝒰).$$
To see that this is the way the renumbering works, note that $`uW_m(𝒰)`$ if and only if $`t^mu`$ is a section of $`T^W(𝒰)`$ (see p. Nonabelian mixed Hodge structures). We have the inclusion of locally free sheaves over $`𝐀^1`$
$$T^W(𝒰(D))T^W(𝒰),$$
and $`t^mu`$ is a section of $`T^W(𝒰)`$ if and only if $`t^{m+1}u`$ is a section of $`T^W(𝒰(D))`$, in other words
$$uW_m(𝒰)uW_{m+1}(𝒰(D)).$$
In the bifiltered situation we adopt a similar notation which gives rise to the same notation for pre-mhs’s: if $`𝒰`$ is a pre-mhs then we obtain a new pre-mhs $`𝒰(D_{wt})`$ where the weight filtration is shifted by one. Note that there is a canonical inclusion
$$𝒰(D_{wt})𝒰$$
which induces the zero map on the associated-graded $`Gr^W`$, and in fact $`𝒰(D_{wt})`$ can be characterized as the universal such object: any morphism of pre-mhs $`𝒰^{}𝒰`$ which induces zero on $`Gr^W`$, factors uniquely through the subobject $`𝒰(D_{wt})`$.
Construction of a namhs $`𝒱`$ of homotopy type $`S_𝐂^2`$
We would like to construct our $`𝒱`$ to have underlying homotopy type $`V_B`$ equal to the complexified $`2`$-sphere $`S_𝐂^2`$ (cf ). This means that we want $`𝒱`$ to fit into a Postnikov fibration sequence
$$K(𝒰^{},3)𝒱K(𝒰,2).$$
The twisting class should be nontrivial (as soon as this class is nontrivial the underlying homotopy type will be $`S_𝐂^2`$).
In fact, we would like $`𝒱`$ to model the homotopy type of $`𝐏^1`$ (since we know, a priori, that this should have a “mixed Hodge structure”).
We have to choose the shifted mixed Hodge structures $`𝒰`$ and $`𝒰^{}`$ appropriately; we look at how to do this in a moment. First note that the above sequence is obtained by specifying the classifying morphism
$$Q:K(𝒰,2)K(𝒰^{},4).$$
To be precise, $`𝒱`$ is the fiber of the map $`Q`$, thus it is the homotopy fiber product
$$𝒱=K(𝒰,2)\times _{K(𝒰^{},4)}.$$
By the Breen-type calculations of the previous section, this is equivalent to specifying a morphism of pre-mhs
$$Sym^2(𝒰)𝒰^{}.$$
Another thing to note is that we would like the resulting $`𝒱`$ to support a linearization; this means that the associated-graded of the morphism $`Q`$ for the weight filtration, should be zero. Another thing to notice is that we would like the base $`K(𝒰,2)`$ and the fiber $`K(𝒰^{},3)`$ to satisfy condition $`\mathrm{𝐌𝐇𝐂}`$. Finally, for our example we would like $`𝒱`$ to model the homotopy type of $`𝐏^2`$. Thus $`𝒰`$ and $`𝒰^{}`$ should be one-dimensional.
In our case of rank one structures, the real structure and the Hodge filtration determine a pure Hodge structure of some type $`(p,p)`$; on the other hand, the weight filtration has only one nontrivial step, at weight $`w`$. Thus the pre-mhs’s $`𝒰`$ and $`𝒰^{}`$ are determined respectively by specifying $`p`$ and $`w`$ (resp. $`p^{}`$ and $`w^{}`$).
The condition that the morphism $`Q`$ should be zero on the associated-graded, is equivalent to saying that $`Q`$ factors through a morphism
$$Q^{}:Sym^2(𝒰)𝒰^{}(D_{wt})$$
(see the end of the previous section p. Nonabelian mixed Hodge structures for the notation $`D_{wt}`$).
We would like to get a $`𝒱`$ which models the homotopy type of $`𝐏^1`$. During our original research on this question, we actually started out by looking at the homotopy type of $`𝐏^1`$ and then worked backwards from there to deduce what conditions needed to be required of a nonabelian mixed Hodge structure. In particular, one notices from the $`𝐏^1`$ example: the fact that one needs to include the “homotopy weight” in the picture (although this was known from another context of Hodge III ); and the condition that the Whitehead product induce zero on the associated-graded of the weight filtration, which suggested the general definition that the associated-graded of the weight filtration should have a structure of a “spectrum” or equivalently a perfect complex.
To get back to the arithmetic at hand, we need to specify $`p`$, $`w`$, $`p^{}`$ and $`w^{}`$. Since we would like to have $`𝒰`$ be $`\pi _2(𝐏^1)`$ and $`𝒰^{}`$ be $`\pi _3(𝐏^1)`$, we take $`p=1`$ and $`p^{}=2`$. This determines the weights, namely $`w=0`$ and $`w^{}=1`$.
This works out correctly for the condition $`\mathrm{𝐌𝐇𝐂}`$: note that $`LGr^W(𝒱)`$ will be the product of $`K(𝒰,2)`$ and $`K(𝒰^{},3)`$. We have (according to the choices of $`w`$ and $`w^{}`$)
$$H^i(LGr_k^W(𝒱))=\begin{array}{c}𝒰,k=w=0,i=2;\hfill \\ 𝒰^{},k=w^{}=1,i=3;\hfill \\ 0\text{otherwise}\hfill \end{array}.$$
The purity condition $`\mathrm{𝐌𝐇𝐂}`$ (cf p. 1) says that on $`H^i(LGr_k^W(𝒱))`$, the Hodge filtration $`F`$ and its complex-conjugate $`\overline{F}`$ (taken with respect to the real structure), are $`m`$-opposed with
$$m=ki.$$
Note that according to our choice of $`p`$ and $`p^{}`$, we obtain $`m=2p=2`$ for $`𝒰`$ (which works out as $`2=m=ki=02`$) and $`m=2p^{}=4`$ for $`𝒰^{}`$ (which works out as $`4=m=ki=13`$). Thus, with our above choices of $`p,w,p^{},w^{}`$ we do indeed obtain the condition $`\mathrm{𝐌𝐇𝐂}`$ for $`𝒱`$.
###### Theorem 9
The extension $`𝒱`$ constructed as the fiber of the morphism $`Q`$ (with $`p`$, $`w`$, $`p^{}`$ and $`w,`$ chosen as above), is a nonabelian mixed Hodge structure.
Proof: Note that all of the total spaces of filtrations involved in $`𝒱`$ are smooth, so the flatness and annihilator conditions are trivial. Our choice of $`p`$, $`w`$, $`p^{}`$ and $`w^{}`$ gives that $`K(𝒰,2)`$ and $`K(𝒰^{},3)`$ are both nonabelian mixed Hodge structures, i.e. they satisfy the condition $`\mathrm{𝐌𝐇𝐂}`$, as calculated immediately above. Note that they also satisfy the strictness condition $`\mathrm{𝐒𝐭𝐫}`$. The linearized associated graded $`LGr^W(𝒱)`$ is a direct sum of those for $`K(𝒰,2)`$ and for $`K(𝒰^{},3)`$ (because the morphism $`Q`$ is zero on the associated graded) so it satisfies condition $`\mathrm{𝐌𝐇𝐂}`$. $`|||`$
Remark: Note that $`K(𝒰^{},4)`$ does not satisfy the $`\mathrm{𝐌𝐇𝐂}`$ condition (since it is translated by one homotopical degree). In particular, we can’t say that $`Q`$ is a morphism of namhs. On the other hand, $`K(𝒰^{}(D_{wt}),4)`$ is a namhs, i.e. it does satisfy the condition $`\mathrm{𝐌𝐇𝐂}`$, so the morphism $`Q^{}`$ is a morphism of namhs which makes it “strict” in a certain sense. This phenomenon plays an important role in what goes on at the end of the proof in the next section.
The namhs on cohomology of a smooth projective variety with coefficients in $`𝒱`$
Suppose now that $`X`$ is a smooth projective variety. We consider the situation of Conjecture 1, for the particular namhs $`𝒱`$ constructed above (Theorem 9).
###### Theorem 10
If $`X`$ is a smooth projective variety and $`𝒱`$ is the nonabelian mixed Hodge structure constructed in Theorem 9, then the linearized pre-namhs
$$:=\underset{¯}{Hom}(X_M,𝒱)$$
constructed in Theorem 4, is a nonabelian mixed Hodge structure.
The proof occupies the remainder of the paper. Start by writing
$$=\underset{¯}{Hom}(X_M,𝒱)=$$
$$\underset{¯}{Hom}(X_M,K(𝒰,2))\times _{\underset{¯}{Hom}(X_M,K(𝒰^{},4))}\underset{¯}{Hom}(X_M,).$$
We have to analyse this fiber product. Note that $`\underset{¯}{Hom}(X_M,)=`$.
Put
$$:=\underset{¯}{Hom}(X_M,K(𝒰,2))$$
and
$$:=\underset{¯}{Hom}(X_M,K(𝒰^{},4)).$$
Also put
$$^{}:=\underset{¯}{Hom}(X_M,K(𝒰^{}(D_{wt}),4)),$$
where $`D_{wt}`$ was the divisor $`𝒜_{hod}\times [0]_{wt}𝒜_{hod}\times 𝒜_{wt}`$. These are all pre-mhc’s, and we have a morphism $`i:^{}`$ inducing zero on the associated-graded of the weight filtration. We have constructed above a morphism (of underlying pre-namhs’s) $`Q`$ from $`K(𝒰,2)`$ to $`K(𝒰^{},4)`$, which in turn yields
$$Q_{}:DPDP$$
factoring through $`Q_{}^{}:DPDP^{}`$. The fact that $`i`$ induces zero on the $`Gr^W`$ implies that $`Q_{}`$ extends to a morphism of linearized pre-namhs’s. On the other hand the zero-section is a morphism (again, of linearized pre-namhs’s)
$$z:DP.$$
We have
$$=\times _{},$$
in other words $``$ is the fiber of $`Q_{}`$ over the zero-section.
###### Lemma 26
In the above situation, $``$ is a mixed Hodge complex and $``$ is a $`1`$-shifted mhc.
Proof: This is a trivial case of , see also Theorem 5. The only difficulty is that $``$ and $``$ are truncations of the mixed Hodge complexes given by . Thus, a priori they are only truncated pre-mhc’s. However, in the present trivial case ($`X`$ being smooth projective), the weight filtration is strict too (its spectral sequence degenerates at $`E_1`$) which means that the cohomology objects are bundles over $`𝒜_{wt}`$; thus truncating yields again a complex whose cohomology objects are bundles, i.e. it is a perfect complex.
For the sign of the shift for $``$, note that $`K(𝒰^{},4)`$ has Hodge degree $`h=4`$, weight $`w=1`$, and homotopical degree $`i=4`$. Now $`h=(wi)+1`$ which means that $`K(𝒰^{},4)`$ is $`1`$-shifted; thus $``$ is also $`1`$-shifted. $`|||`$
###### Corollary 27
The pre-namhs $``$ satisfies the strictness condition $`\mathrm{𝐒𝐭𝐫}`$ as well as the condition $`\mathrm{𝐌𝐇𝐂}`$.
Proof: This was pointed out in Corollary 22; for the record we recall how this goes in this particular case. The connecting morphism $`Q`$ induces zero on the associated-graded $`Gr^W`$, so
$$LGr^W()=LGr^W()\times \mathrm{\Omega }LGr^W().$$
Here $`\mathrm{\Omega }`$ is the loop-space, which on the level of complexes means shifting to the right by one and truncating at $`0`$. As in the lemma, the mixed Hodge complex $``$ has a strict weight filtration in our easy situation, so taking the truncation of the shift again yields a pre-mhc. Recall that $``$ was a $`1`$-shifted mhc; taking the loop-space undoes this shift so $`\mathrm{\Omega }LGr^W()`$ is a mixed Hodge complex. Thus $`LGr^W()`$ is a mixed Hodge complex (which includes the strictness condition). $`|||`$
Some general theory
In order to analyze the annihilator ideals in the fiber product $`\times _{}`$ we discuss this general type of situation.
###### Lemma 28
Suppose $`f:XY`$ is a morphism of geometric $`n`$-stacks which induces an isomorphism on the level of $`\pi _0`$, and is surjective on $`\pi _1`$. Then $`f`$ is smooth.
Proof: A morphism between geometric $`n`$-stacks is smooth if and only if it is formally smooth (this remark was made e.g. in §7.3), thus it suffices to prove that $`f`$ is formally smooth. Suppose $`AB`$ are artinian local schemes. The problem is to show that the map
$$X(B)X(A)\times _{Y(A)}Y(B)$$
is surjective on $`\pi _0`$. Note that the component maps $`X(B)Y(B)`$ and $`X(A)Y(A)`$ are isomorphisms on $`\pi _0`$ and surjections on $`\pi _1`$. These mean that the fibers of these morphisms are connected. We obtain from the statement for $`A`$ that the fibers of the map
$$X(A)\times _{Y(A)}Y(B)Y(B)$$
are connected, in particular this map induces an isomorphism on $`\pi _0`$; but then from the statement for $`B`$ we obtain that the map
$$\pi _0(X(B))\pi _0(X(A)\times _{Y(A)}Y(B))=\pi _0(Y(B))$$
is an isomorphism. This yields the required lifting property. $`|||`$
###### Corollary 29
Suppose that $`X`$ is a geometric $`n`$-stack and suppose that $`\pi _0(X)`$ is represented by an algebraic space (i.e. it is a geometric $`0`$-stack). Then the morphism $`X\pi _0(X)`$ is smooth.
Proof: Indeed, it clearly satisfies the hypotheses of the previous lemma. $`|||`$
The same argument gives furthermore
###### Corollary 30
Suppose that $`X`$ is a geometric $`n`$-stack and suppose that there is $`kn`$ such that $`\tau _k(X)`$ is a geometric $`k`$-stack. Then the morphism $`X\tau _k(X)`$ is smooth.
Proof: Again, it satisfies the hypotheses of the previous lemma. $`|||`$
The above smooth morphisms are useful for the following reason.
###### Lemma 31
Suppose $`X\stackrel{g}{}Y𝐀^1`$ is a morphism of geometric $`n`$-stacks over $`𝐀^1`$. If $`g`$ is smooth, then the annihilator ideals for $`X`$ are the pullbacks of the annihilator ideals for $`Y`$.
Proof: This follows from the definition of annihilator ideals, since a smooth chart for $`X`$ gives by composing with $`g`$ a piece of a smooth chart for $`Y`$ (if $`g`$ wasn’t surjective it might be necessary to add on some other pieces to the chart). $`|||`$
###### Corollary 32
Suppose $`B`$ is a scheme, and suppose $`X`$ and $`Y`$ are geometric $`n`$-stacks over $`B`$. Suppose that $`U=\pi _0(X)`$ and $`V=\pi _0(Y)`$ are represented by schemes. Fix a section $`b:BY`$ and let $`F`$ denote the fiber of $`XY`$ over $`b`$, in other words,
$$F=X\times _YB.$$
Let $`G`$ be the fiber of $`UV`$ over the projection $`pb`$ (here $`p`$ denotes the projection from $`Y`$ to $`V`$). Then the morphism
$$FG$$
is smooth.
Proof: We can base change everything by the map $`BV`$, in particular we can suppose that $`V=B`$. Note then that $`G=U`$. With this hypothesis, our section $`b:BY`$ induces an isomorphism on $`\pi _0`$, so by Lemma 28 it is smooth. This implies that
$$F=X\times _YBX$$
is smooth. On the other hand, the morphism $`XU`$ is smooth, again by Lemma 28. Thus the composition $`FU=G`$ is smooth. $`|||`$
We apply this to the case where the base is $`B=𝐀^1`$. We conclude that the annihilator ideals for the fiber $`F`$ are the pullbacks of the annihilator ideals for the fiber $`G`$ of $`\pi _0(X)\pi _0(Y)`$ (under the hypothesis that these $`\pi _0`$ are schemes). The same remark holds for $`B=𝐀^1\times 𝐀^1`$ where we calculate the annihilator ideals with respect to one of the variables. Similarly, flatness of the fiber with respect to one of the variables is equivalent to flatness of the fiber with respect to the fiber of the map on the $`\pi _0`$.
To conclude, we find that in order to check the annihilator conditions and flatness conditions for the fiber of a morphism of linearized pre-namhs, if the $`\pi _0`$ are schemes (say for the total space of the Hodge and weight filtrations over $`𝐀^1\times 𝐀^1`$) then it suffices to check these conditions for the fiber of the induced morphism on $`\pi _0`$.
###### Lemma 33
Suppose $``$ and $`𝒞`$ are linearized pre-namhs’s, and suppose that $`f:𝒞`$ is a smooth morphism. Suppose that both $``$ and $`𝒞`$ satisfy the “mixed Hodge complex” condition $`\mathrm{𝐌𝐇𝐂}`$ and strictness $`\mathrm{𝐒𝐭𝐫}`$. Then $``$ is a namhs if and only if $`𝒞`$ is a namhs.
Proof: The point is to say that since $`f`$ is smooth, the annihilator ideals of $``$ are the pullbacks of those for $`𝒞`$ (and also flatness of the Hodge filtration is equivalent on $``$ or $`𝒞`$); thus the remaining conditions $`\mathrm{𝐀𝟏}`$, $`\mathrm{𝐀𝟐}`$, $`\mathrm{𝐅𝐥}`$ and $`\mathrm{𝐀𝟑}`$ for $``$ and for $`𝒞`$, are equivalent. $`|||`$
Continuation of the proof of Theorem 10
We analyze our fiber $``$ using the general theory. Note first of all,
###### Lemma 34
In the above situation, the relative zeroth-homologies $`H^0()`$, $`H^0()`$ and $`H^0(^{})`$ are themselves pre-mhs’s, in particular
$$Tot^{W,F}H^0()_{DR}=H^0(Tot^{W,F}(E_{DR})/𝒜_{hod}\times 𝒜_{wt})$$
is a vector bundle over $`𝒜_{hod}\times 𝒜_{wt}`$, and the same for $``$ and $`^{}`$.
Proof: This strictness comes from the fact that, exceptionally in this case, the spectral sequences for the weight filtration degenerate at $`E_1`$ rather than $`E_2`$, as in Lemma 26. $`|||`$
Note that
$$\pi _0(DP)=H^0(),\pi _0(DP)=H^0().$$
The morphism $`Q`$ is not linear so we represent it as a morphism between $`\pi _0`$.
###### Corollary 35
We get a morphism of linearized pre-namhs’s which are schemes
$$Q_{,0}:\pi _0(DP)\pi _0(DP).$$
Let $`𝒦`$ be the fiber of this morphism over the zero section of $`\pi _0(DP)`$. Then there is a smooth morphism of linearized pre-namhs’s
$$𝒦.$$
Proof: From above. $`|||`$
###### Corollary 36
In our situation of Corollary 35, the linearized pre-namhs $``$ is a nonabelian mixed Hodge structure if and only if $`𝒦`$ is.
$`|||`$
Thus to finish the proof of Theorem 10 it suffices to show that $`𝒦`$ is a namhs.
We now analyze the situation a bit further. Recall that $`𝒰`$ is a pure Hodge structure of Hodge type $`(1,1)`$ with a one-step weight filtration placed at weight $`0`$, and $`𝒰^{}`$ is a pure Hodge structure of type $`(2,2)`$ with weight filtration placed at weight $`1`$.
We have
$$H^0()=H^2(X,𝐂)𝐂^{1,1}\{0\}$$
and
$$H^0()=H^4(X,𝐂)𝐂^{2,2}\{1\}.$$
The brackets represent the weight of the pre-mhs in question. Note that $`H^0()`$ is a $`0`$-shifted mhs, pure of weight zero. On the other hand, $`H^0()`$ is a $`1`$-shifted mhs, where the Hodge filtration is pure of weight zero but the weight filtration is shifted and located at $`1`$.
The quadratic morphism $`Q_{,0}`$ is just cup-product, composed with the inclusion $`𝐂\{0\}𝐂\{1\}`$.
Recall that $`^{}=(D_{wt})`$. The morphism $`Q_{,0}`$ factors through a morphism
$$Q_{,0}^{}:\pi _0(DP)\pi _0(DP^{})$$
(this latter being just cup-product, without composition with the inclusion $`𝐂\{0\}𝐂\{1\}`$). Let $`𝒦^{}`$ denote the zero-subscheme of $`Q_{,0}^{}`$.
Note in the above terms that
$$\pi _0(DP^{})=H^0(^{})=H^4(X,𝐂)𝐂^{2,2}\{0\}.$$
It is a $`0`$-shifted mixed Hodge structure which is in fact pure of weight zero (see the renumbering of page Nonabelian mixed Hodge structures). Recall that the same was true of $`H^0()`$. In particular $`Q_{,0}^{}`$ is a quadratic morphism of mixed Hodge structures. We get back to $`Q_{,0}`$ by multiplying $`Q_{,0}^{}`$ by the coordinate $`t`$ defining the divisor $`D_{wt}𝒜_{wt}`$.
In the above particular case, the total spaces
$$Tot^{F,W}H^0()\text{and}Tot^{F,W}H^0(^{})$$
are bundles over $`𝒜_{hod}\times 𝒜_{wt}`$. The weight filtrations are one-step and in degree zero, so these bundles are actually pulled back from $`𝒜_{hod}`$. Furthermore, the Hodge filtrations have splittings by the Hodge decomposition, and the cup-product is compatible with the splitting. Thus on $`𝐀_{hod}^1`$ the bundles in question become trivial bundles, and the quadratic form $`Q^{}`$ is also trivial i.e. it is a product. Therefore the pullback of $`𝒦^{}`$ to $`𝐀_{hod}^1\times 𝐀_{wt}^1`$ has a structure of product
$$𝒦^{}\times _{𝒜_{hod}\times 𝒜_{wt}}𝐀_{hod}^1\times 𝐀_{wt}^1$$
$$𝐀_{hod}^1\times 𝐀_{wt}^1\times C$$
where $`C`$ is the quadratic cone in $`H^2(X,𝐂)`$, kernel of the cup-product map to $`H^4(X,𝐂)`$. This product decomposition is compatible with the inclusion into $`Tot^{F,W}H^0()`$ (or rather the restriction $`T()`$ of this latter over $`𝐀_{hod}^1\times 𝐀_{wt}^1`$). Finally, $`𝒦`$ is the subscheme of $`T()`$ defined by the equation of $`𝒦^{}`$, multiplied by $`t`$. Thus $`𝒦`$ is the scheme-theoretic union of $`𝒦^{}`$ and the inverse image of the divisor $`D_{wt}`$. In particular it is trivializable in the $`𝐀_{hod}^1`$-direction, so $`F`$ is flat. Also, the structure as union of something flat, with the inverse image of the reduced divisor $`D_{wt}`$, means that the annihilator ideal condition $`\mathrm{𝐀𝟏}`$ is satisfied. The annihilator ideal $`\mathrm{ker}(t)`$ is just the ideal defining $`𝒦^{}`$, and $`Ann^W(t;)`$ is the restriction of this over $`D_{wt}=[0]_{wt}\times 𝒜_{hod}`$. Condition $`\mathrm{𝐀𝟐}`$ is satisfied because $`𝒦^{}`$ is nonempty. Finally, the cup-product morphism is a morphism of mixed Hodge structures, so the ideal defined by it is a sub-mixed Hodge structure in the coordinate ring of $`H^2(X,𝐂)`$. This is condition $`\mathrm{𝐀𝟑}`$. $`|||`$
It would seem to be a reasonable project to try to replicate the above types of computations, for objects $`𝒱`$ supposed to represent the homotopy types of projective spaces $`𝐏^n`$ or Grassmanian varieties. These types of explicit computations would present an interest of their own, independant of any eventual general proof of Conjecture 1.
|
warning/0006/quant-ph0006106.html
|
ar5iv
|
text
|
# Entanglement, Information and Multiparticle Quantum Operations
## I Introduction
Many of the novel information-theoretic properties of quantum systems are attributable to the existence of entanglement. Entanglement is responsible for the non-local correlations which can exist between spatially separated quantum systems, as is revealed by the violation of Bell’s inequality. It also lies at the heart of several intriguing applications of quantum information, such as quantum teleportation, quantum computational speed-ups and certain quantum cryptographic protocols.
The central position of entanglement in quantum information theory, and its usefulness in applications, has led to considerable efforts being devoted to finding a suitable measure of how much entanglement a quantum system contains. This problem has been solved completely for bipartite pure states, and the accepted measure is the subsystem von Neumann entropy, conventionally taken to the base 2, so that a maximally-entangled state of a pair of two-level quantum systems, or qubits, possesses one unit of entanglement. This fundamental unit is known as an ebit.
The production of entanglement requires the transmission of quantum information between systems. Conversely, the transmission of quantum information between systems can be used to establish entanglement between them. Perhaps the most perfect expression of this duality is the fact that there are two equivalent definitions of the quantum capacity of a communications channel. According to one definition, it is the asymptotic maximum amount of quantum information that can be transmitted per use of the channel, measured in qubits. In the other, it is the asymptotic maximum number of ebits of entanglement that can be established between the sending and receiving stations, again per use of the channel. An important consequence of this equivalence is the fact that no entanglement can be created without the transmission of quantum information. That is, no entanglement can be created when only local quantum operations are allowed, and only classical information can be transmitted.
Collective quantum operations involving multiple quantum systems can create entanglement and be used to communicate classical information. Conversely, the use of entanglement shared by spatially-separated laboratories, in addition to facilities enabling classical communication and arbitrary local quantum operations, permit these laboratories to carry out collective operations upon a network of separated quantum systems. The ability to do this will have interesting implications for many potential applications of quantum information, such as distributed quantum computing, network quantum communication and the production of novel multiparticle entangled states.
This paper extends the analysis presented in , where we examined the entanglement resources required to carry out collective quantum operations upon $`N`$ qubits, in particular, for the case of even N. In addition to giving a fuller treatment of this problem, including an analysis of the odd case, we examine the classical communication resources required to carry out an arbitrary collective operation upon $`N`$ qubits, and also the amount of classical information that such an operation can be used to send. An intriguing issue highlighted by these considerations is that of how we might quantify the ‘inseparability’ of a quantum operation, rather than that of a quantum state. As we shall see, this inseparability has both classical and quantum aspects.
In section II, we examine the use of entanglement and classical communication to carry out arbitrary collective operations upon a pair of qubits. A simple protocol for achieving this, which uses quantum teleportation, is proposed. Two classical and two quantum measures of the inseparability of a quantum operation arise naturally from these considerations. The quantum measures are analogous to the entanglement of formation and distillation of quantum states. These are respectively the minimum amount of entanglement required to perform the operation, and the maximum amount of entanglement that the operation can establish. The classical measures of inseparability are respectively the minimum amount of classical information required to perform the operation, and the maximum amount of classical information that the operation can be used to communicate. The relationship between these measures leads to the conclusion that a maximally-inseparable quantum operation is the SWAP operation, or any other which can be obtained from it by local unitary transformations.
The remainder of this paper is concerned with collective operations upon $`N`$ qubits. The particular issues we address are: how much bipartite entanglement can an operation be used to establish and how much information can it be used to communicate? Also, how much bipartite entanglement and classical information are needed to perform an arbitrary operation?
In section III, we develop a graph-theoretic framework for the representation of bipartite entanglement and communication networks for $`N`$ laboratories. Using this framework, we generalise to the case of $`N`$ qubits our teleportation protocol. We show that this protocol is optimal in the class of protocols which operate by state teleportation. We also generalise our discussion of quantifying the inseparability of quantum operations to the $`N`$-particle case. As far as the ‘distillation’ measures are concerned, which quantifies the ability of a quantum operation to establish entanglement and communicate classical information, we find that permutation operations are maximally-inseparable. These operations can establish the largest amount of entanglement, and be used to communicate the largest amount of classical information.
In section IV, we are concerned with minimising the entanglement and communication resources required to perform an arbitrary quantum operation upon $`N`$ qubits. There are two distinct scenarios to consider here. On the one hand, we may wish to determine the minimum resources required to carry our an arbitrary operation just once. We refer to this as the ‘one-shot’ scenario. On the other hand, it may be the case that the $`N`$ laboratories share a very large amount of entanglement, and are able to communicate large amounts of classical information. They may wish to use these resources with maximum efficiency to carry out an arbitrary operation many times. The limit as both the resources and the number of repetitions of the operation tends to infinity is known as the asymptotic limit. In this scenario, the asymptotically minimum resources are the minimum entanglement and classical communication that must be used, on average, per run of the operation.
We find that in terms of both entanglement and communication, our teleportation protocol is optimal, in both the one-shot and asymptotic scenarios, for even $`N`$. We obtain lower bounds on the minimum resources for the odd case. We show that, for all $`N4`$, the classical communication and entanglement resources required to carry out an arbitrary operation are strictly greater than the amount of entanglement that can be established, and the amount of classical information that can be sent, by any particular operation. We also show that if the manipulation of these resources obeys the same efficiency restrictions as those found in entanglement swapping and indirect communication, then the teleportation protocol is optimal for all $`N12`$, and for all $`N4`$ for entanglement resources, in the one shot case if only integer resources are allowed.
## II Operations involving two qubits
We consider first the simple case of just two qubits. Suppose that two parties, by convention Alice and Bob, occupy laboratories $`A`$ and $`B`$ which contain qubits $`\alpha `$ and $`\beta `$ respectively. The Hilbert spaces of these systems are denoted by $`_\alpha `$ and $`_\beta `$, so that the Hilbert space of the collective system $`\alpha \beta `$ is the tensor product space $`_\alpha _\beta `$. In addition to these systems, Alice and Bob also possess auxiliary local quantum systems, shared entanglement and a two-way classical communication channel. This setup is illustrated in figure (1). Using these resources, Alice and Bob can perform any collective operation by carrying out the following four steps:
*
Step 1: Alice teleports the state of $`\alpha `$ to Bob in laboratory $`B`$. This costs 1 ebit of entanglement and 2 classical bits from $`A`$ to $`B`$.
*
Step 2: Bob, possibly making use of his auxiliary systems, carries out the operation locally upon the compound system.
*
Step 3: Bob teleports the final state of Alice’s qubit back to her. This costs 1 ebit of entanglement and 2 classical bits from $`B`$ to $`A`$.
*
Step 4: (Selective operations only) Bob transmits to Alice any classical information that he might have obtained at the end of his LQ operation. This step applies only to (generalised) measurements, in which case it would be information about the result.
*
Thus, the total CCSE resources required to perform an arbitrary collective operation on $`\alpha \beta `$ using teleportation, such that Alice and Bob share the same classical information at the end, are
$$2\mathrm{e}bits+2\mathrm{b}its(AB)+2\mathrm{b}its(BA)+C_S(BA).$$
(2.1)
The supplementary information $`C_S(BA)`$ is that which is conveyed by Bob to Alice in step 4. This additional information will be created when the operation, represented by a completely-positive, linear, trace-preserving map $``$, is selective. The most general kind of operation which gives rise to non-zero supplementary information is a generalised measurement. A generalised measurement with $`M`$ outcomes is described by $`M`$ positive, Hermitian operators $`E_r`$, where $`r=1,\mathrm{},M`$ and $`_rE_r=1`$. These operators form a positive, operator-valued measure (POVM) and each of them corresponds to a distinct outcome. If the initial state of $`\alpha \beta `$ is described by the density operator $`\rho `$, then the probability $`p_r`$ of obtaining outcome $`r`$ is given by $`\mathrm{T}r\rho E_r`$. The supplementary information generated at Bob’s laboratory is given by the Shannon entropy of this distribution
$$C_S=\underset{r=1}{\overset{M}{}}p_r\mathrm{log}_2p_r.$$
(2.2)
This quantity can take on any non-negative real value. Clearly, it is zero when the operation is non-selective. If, however, we consider an operation described by the POVM
$$E_r=\frac{1}{M},$$
(2.3)
where Bob records the outcome, then the supplementary information is equal to $`\mathrm{log}_2M`$, which diverges as $`M\mathrm{}`$. For this operation, one cannot decrease the supplementary information using any information that Alice may have about the initial state $`\rho `$, since the probability distribution is uniform regardless of what the initial state is.
For selective operations, the transmission of this supplementary information will have epistemological significance for Alice which may be important in some applications. She may, for example, wish to carry out some local operation upon her subsystem, depending on the supplementary information she receives from Bob. For the remainder of this paper however, we shall not be concerned with $`C_S`$, and when we speak of the classical information required to complete a quantum operation, we will mean that which is needed to carry it out non-selectively. In this paper, we shall be concerned largely with unitary operations anyway, which are non-selective.
Returning to the teleportation protocol, it may be the case that the CC and SE resources required to perform a particular operation, $``$, are less than those required to perform any operation, by this method. Let us denote by $`C_R(:AB)`$, $`C_R(:BA)`$ and $`E_R()`$ the number of classical bits transmitted in each direction and number of ebits of entanglement required to carry out $``$. These may be regarded respectively as classical and quantum measures of how nonlocal the operation is, and $`E_R()`$ is therefore somewhat analogous to the entanglement of formation of quantum states.
Alternative classical and quantum measures of inseparability arise naturally if we consider the fact that collective operations on quantum systems can be used to transmit classical information and establish entanglement between distant locations. Let us define the quantities $`C_C(:AB)`$, $`C_C(:BA)`$ and $`E_C()`$, respectively the maximum number of classical bits that the operation can be used to communicate in each direction, and the maximum number of ebits of entanglement that it can create between $`A`$ and $`B`$. $`E_C()`$ is correspondingly analogous to the entanglement of distillation of quantum states. We must have
$`C_C(:AB)`$ $``$ $`C_R(:AB),`$ (2.4)
$`C_C(:BA)`$ $``$ $`C_R(:BA),`$ (2.5)
$`E_C()`$ $``$ $`E_R().`$ (2.6)
The first two inequalities come from the fact that all classical information that the operation can be used to transmit must, in figure (1), be sent over the classical channel. Equivalently, no classical information can be transmitted using LQSE operations alone. Were this not the case, it would be possible to violate relativistic casuality. An intriguing argument for this has recently been described by Eisert et al. The third inequality comes from the fact that entanglement cannot increase under LQCC operations. For one-way classical communication, this has been shown by Horodecki and Horodecki, to be also equivalent to the impossibility of superluminal communication.
As a consequence of the teleportation protocol, the minimum CCSE resources required to perform any particular operation will not exceed 2 ebits of entanglement and 2 classical bits each way. The most nonlocal quantum operations with regard to the resource measures $`E_R`$ and $`C_R`$ are those for which the minimum values of these quantities are both equal to 2. Inequalities (2.4)-(2.6) imply that the maximum values of the $`E_C`$ and $`C_C`$ cannot exceed 2. Any operation which saturates the limits of 2 on the latter measures must also then saturate inequalities (2.4-2.6), and can be termed a maximally-inseparable operation.
One such operation is the SWAP operation. This is a unitary operation $`U_S`$ which, for any state $`|\psi _\alpha _\alpha `$ and any state $`|\psi _\beta _\beta `$, acts as follows:
$$U_S|\psi _\alpha |\psi _\beta =|\psi _\beta |\psi _\alpha ,$$
(2.7)
that is, it exchanges the states of $`\alpha `$ and $`\beta `$. The ability of SWAP to create 2 ebits of entanglement and transmit 2 classical bits each way is easily demonstrated. We shall now do this, with reference to figures (2) and (3). The remarkable properties of the SWAP operation are also described by Collins et al and Eisert et al.
In figure (2), Alice and Bob initially share 2 ebits of entanglement in the form of Bell states. Using superdense quantum coding, Alice and Bob can each manipulate one of their particles, those represented by hollow circles, to produce any of the 4 Bell states that they wish. The final shared Bell states are $`|B_\alpha `$ and $`|B_\beta `$. The SWAP operation is them performed on the states of the hollow qubits, resulting in each party being in possession of the entire Bell state which the other party created. Each then performs a Bell measurement, which has 4 possible outcomes and thus reveals 2 bits of information, showing how SWAP can transmit 2 classical bits each way.
Figure (3) shows how SWAP can be used to establish 2 ebits of entanglement between Alice and Bob. Each party initially possesses 1 local ebit of entanglement. If the SWAP operation is used to interchange the states of one particle from each entangled pair, the result is that Alice and Bob share 2 ebits of entanglement.
Notice that the SWAP operation cannot be used to create 2 ebits of entanglement, and communicate 2 classical bits each way, simultaneously. In fact, looking at figures (3) and (4), we can see that one of these processes is essentially the time-reverse of the other.
A broader class of maximally-inseparable operations on 2 qubits can be obtained by considering those which are equivalent to SWAP up to a bilateral local unitary operation. Specifically, any unitary operation $`T`$ of the form $`T=(U_{\alpha _2}U_{\beta _2})U_S(U_{\alpha _1}U_{\beta _1})`$ must require the same entanglement and communication resources as $`U_S`$. Here, $`U_{\alpha _i}`$ and $`U_{\beta _i}`$ are local unitary operations on $`\alpha `$ and $`\beta `$ respectively. The reason for this is simple: it is possible to convert this operation into the SWAP operation by just local-unitary transformations, that is, without any additional entanglement or classical communication resources. This follows from the simple observation that $`U_S=(U_{\alpha _2}^{}U_{\beta _2}^{})T(U_{\alpha _1}^{}U_{\beta _1}^{})`$.
## III Multiparticle systems, graphical representations and teleportation.
Let us now extend our discussion to the case of $`N`$-particle systems. Instead of just two spatially-separated laboratories, we now have $`N`$ of them, which we label $`A_j`$, where $`j=1,\mathrm{},N`$. In each of these laboratories is a qubit, and we label these $`q_j`$. We are interested in the CCSE resources required to perform an arbitrary collective quantum operation involving all $`N`$ qubits.
Each laboratory shares a certain number of ebits of entanglement with every other laboratory. In this paper, we shall, except where indicated, take all entanglement to be in pure, bipartite form. The $`N`$ laboratories are also linked by classical communication channels, so that each can communicate a certain number of classical bits to the others. Each laboratory also possesses auxiliary quantum systems allowing arbitrary local quantum operations to be performed.
The CCSE resources available to the network of laboratories are conveniently represented using the concepts of graph theory. Recall that a graph $`G=(V,E)`$ is a set $`V`$ of vertices connected by edges comprising a set $`E`$. If the edges have a sense of direction indicating an asymmetrical relationship between the vertices it connects, the graph is said to be a directed graph, or a digraph. If there is no preferred direction, the graph is undirected.
These resources can be represented by distinct entanglement and communication graphs. Both graphs are comprised of $`N`$ vertices, each of which represents one of the laboratories $`A_j`$. The resource entanglement graph $`G_E`$ represents the amount of bipartite entanglement shared between each pair of laboratories. Specifically, we write both the $`j`$th laboratory and its corresponding vertex as $`A_j`$. The weight of the edge joining vertices $`A_i`$ and $`A_j`$ is equal to the number of ebits of entanglement shared by these laboratories. The graph is characterised completely by the $`N\times N`$ resource entanglement matrix $`𝐄_R`$. The element $`E_R^{ij}`$ of this matrix is equal to the number of ebits of entanglement shared by $`A_i`$ and $`A_j`$. The diagonal elements of this matrix are zero.
Clearly, $`𝐄_R`$ is symmetric and the graph $`G_E`$ is undirected. These observations follow from the fact that entanglement is a shared, rather than a directed resource.
As an example, a resource entanglement graph for $`N=4`$ is depicted in figure (4). This corresponds to the following resource entanglement matrix:
$$𝐄_R=\left(\begin{array}{cccc}0& 3& 2& 6\\ 3& 0& 1& 0\\ 2& 1& 0& 0\\ 6& 0& 0& 0\end{array}\right).$$
(3.1)
Likewise, we can define a resource communication graph $`G_C`$. This represents the number of classical bits that the laboratories can communicate directly to each other. By directly, we mean that it is not relayed by a set of intermediate laboratories from origin to destination. The weight of the edge running from $`A_i`$ and $`A_j`$ represents the number of classical bits that $`A_i`$ can communicate directly to $`A_j`$. These weights are the elements of a correspondingly defined resource communication matrix $`𝐂_R`$. The $`ij`$ element of this matrix, $`C_R^{ij}`$, is equal to the number of classical bits that $`A_i`$ can communicate directly to $`A_j`$. The diagonal elements of this matrix are also zero. $`𝐂_R`$ is not necessarily symmetric and the graph $`G_C`$ is directed, which follows from the fact that communication operations have a natural sense of direction from sender to receiver. An example of a resource communication graph for $`N=4`$ is given in figure (5), which corresponds to the resource communication matrix
$$𝐂_R=\left(\begin{array}{cccc}0& 1& 4& 0\\ 2& 0& 0& 9\\ 0& 0& 0& 0\\ 5& 0& 0& 0\end{array}\right).$$
(3.2)
The fact that each pair of vertices may be joined by more than one edge means that $`G_C`$ is, strictly speaking, a multigraph, indeed a multidigraph since these edges are directed. We do not, however, wish to unduly proliferate terminology, so we shall simply use the term graph.
In either graph, an edge of weight zero is equivalent to no edge. Thus, if two vertices are not linked by an edge in the graph $`G_E`$, then the corresponding laboratories share no entanglement. Similarly, if there is no edge running from vertex $`A_i`$ to $`A_j`$ in the graph $`G_C`$, then $`A_i`$ cannot communicate any classical information directly to $`A_j`$.
Two quantities which will be of particular interest to us are the total shared entanglement and the total number of classical bits that can be communicated. Respectively, these are
$`E_R`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{ij}{}}E_R^{ij},`$ (3.3)
$`C_R`$ $`=`$ $`{\displaystyle \underset{ij}{}}C_R^{ij}.`$ (3.4)
The factor of $`1/2`$ in Eq. (3.3) occurs as a consequence of the shared nature of entanglement, which implies that the entanglement shared between each pair of laboratories is counted twice in the summation.
Having established the framework within which we will work, let us now see how such resources can be used to perform an arbitrary collective quantum operation upon the $`N`$ qubits $`q_j`$. The teleportation-based procedure for 2 qubits described in the preceding section admits a natural generalisation to the case of $`N`$ qubits, which we now describe.
We consider the situation in which all laboratories share entanglement and have the resources for two-way classical communication with one particular laboratory. Let this laboratory be $`A_1`$. It follows that the other laboratories can teleport the states of their qubits to $`A_1`$. The operation can then be carried out at $`A_1`$ as an LQ operation. The final states of the other qubits can then be teleported back to their original laboratories, completing the procedure.
This multiparticle protocol generalises the first 3 steps of the 2-qubit protocol described in the preceding section. It requires each of the laboratories $`A_2,\mathrm{},A_N`$ to share 2 ebits of entanglement with $`A_1`$ and for 2 bits of classical information to be communicated each way between each of them and $`A_1`$. The elements of the corresponding resource entanglement and communication matrices are
$$E_R^{ij}=C_R^{ij}=2|\delta _{i1}\delta _{1j}|.$$
(3.5)
The corresponding graphs $`G_E`$ and $`G_C`$ are depicted in figures (6) and (7). The total resource entanglement and communication are
$$E_R=\frac{C_R}{2}=2(N1).$$
(3.6)
The graph $`G_E`$ representing the entanglement resources required by the teleportation protocol is said to be a tree. Generally speaking, a tree is an connected, acyclic graph, that is, one where every pair of vertices is connected by at least one path, and where there are no closed paths.
Any quantum operation on $`N`$ qubits can be performed using this method and thus, at least for the topology of entanglement and communication in our protocol, the values of $`E_R`$ and $`C_R`$ in Eq. (3.6) are sufficient.
Much of the remainder of this paper will be concerned with the issue of whether or not this protocol is optimal: that is, whether or not there exists a procedure for carrying out any quantum operation on $`N`$ qubits which less resources than this protocol. Prior to doing so, it is of interest to determine whether or not this protocol is the most efficient among those that operate by teleportation of the states of the qubits concerned.
We have $`N`$ laboratories $`A_i`$, each of which possesses a corresponding qubit $`q_i`$. If we wish the $`N`$ laboratories to be able to carry out any collective operation upon the $`q_i`$ by teleporting single qubits, then, as we now show, at least $`2(N1)`$ such teleportations must take place.
To see why, suppose that the first teleportation is from $`A_1`$ to $`A_2`$. $`A_2`$ now has information about $`A_1`$. Secondly another lab $`A_r`$ teleports a state to $`A_3`$. If they are completely different labs from the first pairs of laboratories then $`A_3`$ can hold information only about one other lab, $`A_r`$. If, however, $`r=2`$, then $`A_3`$ can hold information about 3 qubits, $`q_1`$, $`q_2`$ and $`q_3`$.
The most efficient way to pass on information is for $`A_3`$ to teleport a state $`A_4`$ and so on. After $`N1`$ steps the best possible situation is that one lab $`A_N`$ can have information from all of the other labs. None of the other labs can have a complete set of information. So now there must be at least a further $`N1`$ communication events required so that each of the first $`N1`$ labs can get information from lab $`A_N`$. This gives a total of a least $`2(N1)`$ communications in all which costs $`2(N1)`$ ebits.
We saw in the preceding section that the total resource entanglement for an arbitrary operation upon two particles can be recovered if the operation in question is unitarily equivalent to SWAP. Also, for such an operation, the required classical communication facilities required to complete an arbitrary operation can be fully used to communicate useful information. An important question is, does there exist an operation, or class of operations which fulfills this role in the general $`N`$-particle case?
Let us denote the maximum total entanglement that can be established, and the maximum number of classical bits that can be sent by any operation by $`E_C`$ and $`C_C`$ respectively. To address this issue, it is helpful to partition the entire network of $`N`$ qubits into a single qubit and a compound system comprised of the remaining $`N1`$ qubits. How much entanglement can be established between the location of the isolated qubit and the rest of the network? Also, how much classical information can be transmitted in both directions between the location of this qubit and the remainder?
In the teleportation protocol, a special status was given to laboratory $`A_1`$. However, this choice was arbitrary, and clearly this role could have been assumed by any laboratory. It follows that any collective quantum operation upon $`N`$ qubits can be carried out with each laboratory sharing no more than 2 ebits of entanglement, and able to exchange no more than 2 classical bits each way, with the rest of the network. The reasoning which led us to inequalities (2.4-2.6) then implies that no operation can be used to establish more than 2 ebits of pure bipartite entanglement, or be used to exchange more than 2 classical bits each way, between any particular laboratory and the rest of the network.
The maximum total entanglement that can be established is then obtained by multiplication of 2 ebits by the number of laboratories and then dividing by 2, since entanglement is shared, giving
$$E_CN.$$
(3.7)
The maximum number of classical bits that any collective operation can be used to communicate is obtained by multiplying the maximum amount of information that one laboratory can communicate, namely 2 bits, by N, the number of laboratories, giving
$$C_C2N.$$
(3.8)
These bounds are tight, that is, they can be accessed by a specific class of quantum operations, the permutation operations.
A unitary permutation operator upon $`N`$ qubits is described by
$$U_P|\psi _1\mathrm{}|\psi _N=|\psi _{P(1)}\mathrm{}|\psi _{P(N)},$$
(3.9)
where $`P(i)`$ represents a permutation of the index $`i[1,N]`$. Here, we consider only permutation operations which satisfy $`P(i)ii[1,N]`$.
To see that $`N`$ ebits of entanglement can be established using a permutation operation, suppose that $`A_i`$ contains one local ebit, in the form of, for example, some standard Bell state. We shall denote this state by $`|B^{i,i}`$. The first and second indices denote the laboratories which possess the first and second qubits respectively. Suppose now that the second qubits’ states are permuted according to Eq. (3.9). This transforms $`|B^{i,i}`$ into $`|B^{i,P(i)}`$. Following this permutation, laboratories $`A_i`$ and $`A_{P(i)}`$ share the Bell state $`|B^{i,P(i)}`$. There are $`N`$ laboratories, and so $`N`$ shared ebits of entanglement in the form of Bell states have been established. This procedure is illustrated in figure (8).
To see that a permutation operation can be used to communicate $`2N`$ classical bits, suppose that $`A_{P^1(i)}`$ shares the Bell state $`|B^{i,P^1(i)}`$ with $`A_i`$. Locally, using superdense coding, $`A_{P^1(i)}`$ can manipulate the state of the second qubit in this Bell state so that it becomes any of the four possible Bell states. Figure (9) illustrates this scenario, where each second qubit is represented by a hollow circle. We may therefore write the state following this local manipulation as $`|B_{\mu (i)}^{i,P^1(i)}`$, where the integer $`\mu (i)[1,\mathrm{},4]`$. The permutation operation is then carried out on the set of locally-manipulated qubits, resulting in $`A_i`$ being in possession of the state $`|B_{\mu (i)}^{i,i}`$. By performing a Bell measurement, $`A_i`$ can read the two bits of information sent by $`A_{P^1(i)}`$, and in total $`2N`$ bits have been communicated.
As is the case with the SWAP operation for 2 qubits, the number of ebits that $`U_P`$ can establish is also the minimum amount of entanglement required to carry out this operation. The same is true of the classical communication resources involved. Suppose that $`A_i`$ shares one ebit of entanglement with $`A_{P(i)}`$ and can communicate 2 classical bits to this location. Then the permutation operation can be carried out using these resources to teleport the state of qubit $`q_i`$ from $`A_i`$ to $`A_{P(i)}`$. Permutation operations, including the SWAP operation, make maximally efficient use of the resources required to carry them out.
As is also the case with the SWAP operation, any operation which is equivalent to $`U_P`$ up to an $`N`$-partite local unitary transformation, that is, any unitary operation $`T`$ of the form
$$T=\left(\underset{i=1}{\overset{N}{}}U_i^2\right)U_P\left(\underset{j=1}{\overset{N}{}}U_j^1\right),$$
(3.10)
where $`U_i^1,U_j^2`$ are arbitrary local unitary operations on qubits $`q_i`$ and $`q_j`$, is also maximally-inseparable. This is a consequence of the fact that $`U_P`$ can be obtained from $`T`$ by the local unitary operation
$$U_P=\left(\underset{i=1}{\overset{N}{}}U_i^2\right)T\left(\underset{j=1}{\overset{N}{}}U_j^1\right).$$
(3.11)
Comparing (3.7) and (3.8) with (3.6), we see that the total amount of entanglement that can be established, and the total amount of classical information that can be sent is strictly less than that required to carry out an arbitrary operation using the teleportation protocol, with the exception of the case $`N=2`$. We have not, however, established the optimality of the teleportation protocol. We examine this issue in the following section.
## IV Resources Required to Perform Arbitrary Multiparticle Operations
### A Graph Symmetrisation
The teleportation-based method for performing an arbitrary collective quantum operation upon $`N`$ spatially separated qubits requires $`E_R=2(N1)`$ ebits of entanglement and $`C_R=4(N1)`$ classical bits. An obviously important question is: are these figures optimal, in the sense that no less entanglement and communication will suffice?
Unlike the case of $`N=2`$, for general $`N`$ we cannot answer this question by making use of the fact that the resource entanglement and communication required by the teleportation protocol can respectively be recovered and used to communicate messages, as can be done with the SWAP operation. For $`N>2`$, the values of $`E_R`$ and $`C_R`$ for the teleportation protocol, given by Eq. (3.6), are strictly greater than the upper bounds on $`E_C`$ and $`C_C`$ in (3.7) and (3.8). Another approach must be taken to resolve this issue. In this section, we show that, for even $`N`$, the resource entanglement and communication required to perform an arbitrary quantum operation upon $`N`$ qubits using the teleportation protocol are indeed the minimum possible values. We describe a novel proof technique, which we term graph symmetrisation, to establish this fact. The same method is then used to find lower bounds on the minimum values of $`E_R`$ and $`C_R`$ for odd $`N`$. We find that, for $`N4`$, these lower bounds are strictly greater than the upper bounds on $`E_C`$ and $`C_C`$ in Eqs. (3.7) and (3.8).
The problem we will investigate is the following. A network of laboratories $`A_i`$ possesses shared bipartite entanglement, described by the graph $`G_E`$, and facilities enabling limited classical communication between them, described by a graph $`G_C`$. If these graphs describe sufficient resources to enable any collective operation to be performed upon their respective resident qubits $`q_j`$, then what lower bounds must the corresponding values of $`E_R`$ and $`C_R`$ satisfy?
We commence by making the following observation: if the graphs $`G_E(V)`$ and $`G_C(V)`$ describe sufficient resources, then so does any other pair of graphs obtained from them by a permutation of the vertices. Note that we have written the dependence of the graphs on the vertex set explicitly here. This makes sense intuitively. Nevertheless, here we provide a short proof. Let $`G_E^{}(V)`$ and $`G_C^{}(V)`$ be the entanglement and communication graphs obtained from $`G_E(V)`$ and $`G_C(V)`$ by a permutation $`P`$ of the vertex set. We may write $`G_E^{}(V)=G_E(P[V])`$ and $`G_C^{}(V)=G_C(P[V])`$, where $`P[V]`$ is the permutation. The reversibility of permutation operations implies that $`G_E(V)=G_E^{}(P^1[V])`$ and $`G_C(V)=G_C^{}(P^1[V])`$. Consider now a quantum operation $``$ on the $`N`$ qubits. This also depends on the vertex set and so we write it as $`(V)`$. We can obtain another quantum operation $`^{}`$ from $``$ by applying the same permutation to the vertex set, that is, $`^{}(V)=(P[V])`$, and $`(V)=^{}(P^1[V])`$.
If there exists an operation $`^{}(V)`$ which cannot be performed using the resources described by the graphs $`G_E^{}(V)`$ and $`G_C^{}(V)`$, then by reversing the permutation $`P`$, it follows that $`(V)`$ cannot be carried out using $`G_E(V)`$ and $`G_C(V)`$, in contradiction with our premise. Thus, if the resources described by the graphs $`G_E(V)`$ and $`G_C(V)`$ can be used to carry out any quantum operation, then so do those described by $`G_E(P[V])`$ and $`G_C(P[V])`$ for any permutation P of the vertex set V.
Let us now consider the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$, defined by
$`\stackrel{~}{G}_E`$ $`=`$ $`{\displaystyle \underset{P[V]}{}}G_E(V),`$ (4.1)
$`\stackrel{~}{G}_C`$ $`=`$ $`{\displaystyle \underset{P[V]}{}}G_C(V).`$ (4.2)
These graphs are constructed by summing over all of the graphs obtained from $`G_E`$ and $`G_C`$ by permuting the vertices. By summing, we mean summing the entanglement and communication represented by the weights of the edges. The resource entanglement and communication matrices for these graphs are easily obtained. Their elements are
$`\stackrel{~}{E}_R^{ij}`$ $`=`$ $`{\displaystyle \underset{P[V]}{}}E_R^{P(i),P(j)},`$ (4.3)
$`\stackrel{~}{C}_R^{ij}`$ $`=`$ $`{\displaystyle \underset{P[V]}{}}C_R^{P(i),P(j)}.`$ (4.4)
These graphs are regular and complete. A complete graph is one where each pair of vertices is joined by an edge. In the case of the graph $`\stackrel{~}{G}_C`$, this means that each pair of vertices is connected by an edge in each direction. A regular graph is one where all edges have the same weight. In a network represented by these graphs, all pairs of laboratories share the same amount of entanglement, and can communicate the same amount of classical information, in both directions.
For the purposes of illustration, the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$ are shown in figures (10) and (11) corresponding to the particular graphs $`G_E`$ and $`G_C`$ in figures (4) and (5).
The regularity and completeness properties are easily proven, and follow immediately from the fact that the graphs $`\stackrel{~}{G}_C`$ and $`\stackrel{~}{G}_E`$, being defined as sums over all vertex permutations, are clearly themselves permutation invariant.
The total resource entanglement and communication for these graphs, $`\stackrel{~}{E}_R`$ and $`\stackrel{~}{C}_R`$, are easily evaluated in terms of the corresponding resources represented by the original graphs $`G_E`$ and $`G_C`$. Take the case of $`\stackrel{~}{E}_R`$: there are $`N!`$ permutations of the vertex set, so $`\stackrel{~}{G}_E`$ describes $`N!`$ times as much entanglement as $`G_E`$, that is
$$\stackrel{~}{E}_R=N!E_R.$$
(4.5)
Similarly,
$$\stackrel{~}{C}_R=N!C_R.$$
(4.6)
All edges in each of these graphs have the same weight, and it will be convenient to label these two weights. For $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$, we denote these edge weights simply by $`e`$ and $`c`$ respectively. These are
$`e`$ $`=`$ $`2(N2)!E_R,`$ (4.7)
$`c`$ $`=`$ $`(N2)!C_R,`$ (4.8)
which follows from Eqs. (4.5) and (4.6), and also from the fact that the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$ have $`N(N1)/2`$ and $`N(N1)`$ edges respectively. For the graphs in figures (10) and (11), we find that $`e=24`$ and $`c=42`$.
There are $`N!`$ permutations of the vertex set. The permutation invariance of the sufficiency condition then implies that the resources represented by the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$ can be used to perform any operation $`N!`$ times. By this, we mean the following: suppose that $`A_i`$ contains $`N!`$ qubits. We can then define $`N!`$ sets of qubits, where each contains one qubit from each laboratory. It will be possible to perform the same operation separately upon every one of these sets.
In the next two subsections, we will use the formalism developed here, together with inequalities (2.4-2.6) to establish lower bounds on the values of e and c. These lead to lower bounds on $`E_R`$ and $`C_R`$ through Eqs. (4.7) and (4.8). We shall treat the cases of even and odd N separately, since, for even $`N`$, it is possible to use this technique to solve for the minimum values of $`E_R`$ and $`C_R`$ which are sufficient to carry out any operation. These are those required to implement the teleportation protocol described in section III.
### B Necessary and Sufficient Resources for Even N
Using the formalism we have set up, we can obtain the minimum values of $`E_R`$ and $`C_R`$ exactly when $`N`$ is even. The network of $`N`$ laboratories is assumed to possess sufficient resources, described by the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$, to enable any operation to be carried out $`N!`$ times. Here, we consider one particular operation, which we will refer to as the pairwise-SWAP (PS) operation. This operation has the effect of swapping the state of a qubit at $`A_j`$ with that of one at $`A_{j+1}`$, for all odd $`j`$. If we write the two-particle SWAP operation exchanging the states of qubits at $`A_j`$ and $`A_{j+1}`$ as $`U_S^{j+1,j}`$, then the PS operation may be written as
$$U_{PS}=U_S^{N,N1}U_S^{N2,N3}\mathrm{}U_S^{2,1}.$$
(4.9)
This operation is illustrated in figure (12).
The PS operation is a permutation operation which leaves no vertex invariant, and so it can be used to establish $`N`$ ebits of entanglement and to communicate $`2N`$ classical bits. Performing this operation $`N!`$ times can then be used to create $`N!N`$ ebits of entanglement and to send $`2N!N`$ bits of classical information. From the assumption that the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$ represent sufficient resources to carry out the $`N!`$-fold PS operation, we can deduce the minimum values of $`e`$ and $`c`$, and using Eqs. (4.7) and (4.8), those of $`E_R`$ and $`C_R`$, required to do so.
To determine the minimum value of $`e`$, we will make use of the fact that entanglement cannot increase under LQCC operations. Consider the situation depicted in figure (13). We partition the entire network into two sets. One contains the even laboratories $`A_2,A_4,\mathrm{},A_N`$, and the other contains the odd ones $`A_1,A_3,\mathrm{},A_{N1}`$. We shall refer to these sets as $`S_{even}`$ and $`S_{odd}`$.
The total entanglement initially shared by these sets can be calculated in a straightforward manner. Each of the $`N/2`$ laboratories in $`S_{odd}`$ shares e ebits with each laboratory in $`S_{even}`$, that is, $`Ne/2`$ ebits with $`S_{even}`$ in total. Adding up the $`N/2`$ such contributions from the laboratories in $`S_{odd}`$ gives $`(N/2)^2e`$ ebits initially shared by $`S_{even}`$ and $`S_{odd}`$. The final entanglement they share is $`N!N`$ ebits. The total entanglement that $`S_{even}`$ and $`S_{odd}`$ share cannot increase, giving the inequality
$$\left(\frac{N}{2}\right)^2eN!N.$$
(4.10)
Making use of Eq. (4.7), we find that
$$E_R2(N1).$$
(4.11)
This lower bound on the total resource entanglement is precisely the amount which is required by the teleportation protocol. Thus, for even $`N`$, the teleportation protocol is optimal with regard to the required total resource entanglement.
This bound has been derived on the basis of the fact that, in a multiparticle system, the (bipartite) entanglement shared by two exhaustive subsets cannot increase under LQCC operations. Although the entanglement initially shared by each pair of laboratories is in pure, bipartite form, the transformation shown in figure (13) may, at some point, manipulate the resource entanglement into, possibly mixed, multiparticle entanglement. Our argument still holds under these circumstances. If the final entanglement is in multiparticle form, then in order to carry out the $`N!`$-fold PS operation, $`A_j`$ and $`A_{j+1}`$ will have to be able to distill $`2N!`$ ebits of pure, bipartite entanglement. The total distillable entanglement between $`S_{even}`$ and $`S_{odd}`$ cannot increase, which leads to inequality (4.10) and thus the teleportation bound in (4.11).
The nonincreasing of entanglement under LQCC operations is an asymptotic result. It follows that the teleportation protocol is asymptotically optimal for even $`N`$. By asymptotic, we mean that, given a very large number of sets of separated qubits, where the same, arbitrary operation is to be carried out on each set, the teleportation protocol uses the minimum average entanglement that is required per run of the operation.
In practical situations, it is often the resources required to carry out an operation successfully just once that will be of interest. For general information processing tasks, the resources required in the ‘one-shot’ scenario are at least equal to the resources required asymptotically. For the problem we have considered here, when $`N`$ is even, the entanglement resources required in both scenarios are equal. This is because the teleportation protocol, which requires $`2(N1)`$ ebits, can be used to carry out any collective operation on $`N`$ qubits once.
The proof that the $`N`$ laboratories must also be able to send $`4(N1)`$ classical bits proceeds similarly. The graph $`\stackrel{~}{G}_C`$ is assumed to represent sufficient CC resources to perform any operation $`N!`$ times. If this operation is the PS operation, then it should then be able to communicate $`2N!N`$ bits. Given this, and the fact that each laboratory can communicate $`c`$ classical bits to each other one, we can determine the minimum value of $`c`$, from which we can infer the minimum of $`C_R`$ through Eq. (4.8).
Again, we partition the vertex set into $`S_{even}`$ and $`S_{odd}`$. According to inequalities (2.4-2.5), the total amount of resource communication between the sets $`S_{even}`$ and $`S_{odd}`$ cannot be less than the amount of classical information that the $`N!`$-fold PS operation can be used to communicate between these two sets.
According to $`\stackrel{~}{G}_C`$, each of the $`N/2`$ laboratories in $`S_{even}`$ can communicate c classical bits to each one in $`S_{odd}`$. From this, we find that the maximum amount of classical information that can be sent in either direction between $`S_{even}`$ to $`S_{odd}`$ odd is $`(N/2)^2c`$ bits.
The $`N!`$-fold PS operation can be used to send $`N!N`$ bits in either direction $`S_{even}`$ to $`S_{odd}`$. Inequalities (2.4-2.5) imply that
$$\left(\frac{N}{2}\right)^2cN!N.$$
(4.12)
Making use of Eq. (4.8), we obtain
$$C_R4(N1),$$
(4.13)
which is the amount of resource communication required to implement the teleportation protocol. We have thus shown that, in terms of the total resource entanglement and communication, the teleportation protocol in section III is maximally efficient.
### C Necessary Resources for Odd N
Let us now examine the case of odd $`N`$. We have been unable to find a specific operation which proves that the minimum resources required to carry out any operation on an odd number of qubits are those employed by the teleportation protocol. However, using the graph symmetrisation technique, it is still possible to obtain lower bounds on these minimum resources. As before, we assume that the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$ represent sufficient resources to perform any operation $`N!`$ times. The specific operation we shall consider here is the PS operation upon the first $`N3`$ qubits, and a separate, cyclic permutation of the remaining three. For $`N=3`$, there is only this latter part of the operation. We shall refer to this as the PS+CP operation, and it is illustrated in figure (14).
The PS+CP operation is again a permutation operation which leaves no vertex invariant. It follows that it can be used to establish $`N`$ ebits of entanglement and to communicate $`2N`$ classical bits.
We shall now apply the same arguments as those used for the PS operation for even $`N`$ to obtain lower bounds on the resources required to carry out the PS+CP operation. Again, we divide the $`N`$ laboratories into two sets, $`S_{even}`$ and $`S_{odd}`$.
Our aim, as before, is to obtain lower bounds on the minimum values of $`c`$ and $`e`$ from the assumption that $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$ represent sufficient resources to perform this particular operation $`N!`$ times.
We begin by deriving a lower bound on the minimum sufficient resource entanglement $`E_R`$. In figure (15), the total initial entanglement between $`S_{even}`$ and $`S_{odd}`$ is depicted, as is the amount of entanglement that can be established by the $`N!`$-fold PS+CP operation. As before, these sets are divided by an imaginary partition, and the total entanglement across this partition cannot increase.
Initially, each of the $`(N1)/2`$ laboratories in $`S_{even}`$ shares e ebits with each of the $`(N+1)/2`$ laboratories in $`S_{odd}`$. The total amount of entanglement initially shared by $`S_{even}`$ and $`S_{odd}`$ is then $`(N^21)e/4`$ ebits. There are two contributions to the amount of entanglement that can be created by the $`N!`$-fold PS+CP operation. One is that created by the PS part of the operation on the qubits in the first $`N3`$ laboratories. This can create $`N!(N3)`$ ebits. The second contribution comes from the cyclic permutation on the remaining three laboratories. This gives an additional $`2N!`$ ebits. Inequality (2.6) then implies
$$\left(\frac{N^21}{4}\right)eN!(N1).$$
(4.14)
Making use of Eq. (4.8), we obtain a corresponding lower bound on $`E_R`$:
$$E_R2\left(\frac{N}{N+1}\right)(N1),$$
(4.15)
that is, the teleportation bound multiplied by a factor of $`N/(N+1)`$. The argument given for even $`N`$ that this bound cannot be improved upon by converting the initial bipartite resource entanglement into multiparticle entanglement also applies here.
Let us now obtain a lower bound on the minimum resource communication $`C_R`$. As with the even case, we will make use of the fact that the amount of classical information that the PS+CP operation can be used to communicate, in either direction between $`S_{even}`$ and $`S_{odd}`$, cannot exceed the amount of resource communication in this direction that must be consumed in order to implement the $`N!`$-fold PS+CP operation.
For the sake of concreteness, we shall consider communication from $`S_{even}`$ to $`S_{odd}`$. Initially, each of the $`(N1)/2`$ laboratories in $`S_{even}`$ can communicate c classical bits to each of the $`(N+1)/2`$ laboratories in $`S_{odd}`$. This implies that the total resource communication from $`S_{even}`$ to $`S_{odd}`$ is $`(N^21)c/4`$ bits. It is easy to show that it is the same in the opposite direction.
As with entanglement, the PS and CP parts of the $`N!`$-fold PS+CP operation make distinct contributions to the amount of information that this operation can be used to send from $`S_{even}`$ to $`S_{odd}`$. For a single implementation of PS+CP, the PS part can communicate $`(N3)`$ bits from $`S_{even}`$ to $`S_{odd}`$, while the CP part can be used to send 2 bits. Thus, the total amount of classical information that the $`N!`$-fold PS+CP operation can be used to send in either direction across the partition is $`N!(N1)`$ bits.
The impossibility of this exceeding the resource communication in either direction across the partition implies that
$$\left(\frac{N^21}{4}\right)cN!(N1).$$
(4.16)
and, making use of Eq. (4.8), we obtain the corresponding bound for $`C_R`$:
$$C_R4\left(\frac{N}{N+1}\right)(N1),$$
(4.17)
which, like the entanglement bound in (4.15), is the teleportation bound multiplied by $`N/(N+1)`$.
Like the bounds in (4.11) and (4.13) for the even case, the lower bounds we have obtained here for the minimum resource entanglement and communication for odd N are asymptotic results. However, the fact that the bounds (4.15) and (4.17) are not integers suggests that if the available resources are at these bounds, they may not be very useful in the one-shot case, where it is more desirable to be able to transmit whole bits of classical information, and to manipulate whole ebits of entanglement. With this in mind, let us return to the bound on $`E_R`$ in (4.15) and consider the inequality
$$\frac{2N(N1)}{N+1}=2(N1)2+\frac{4}{N+1}2(N1)2$$
(4.18)
where the equality is attained only in the limit as $`N\mathrm{}`$. If we are to round this bound up to the next integer, we obtain $`2(N1)1`$. Thus, the minimum number of integer ebits able to carry out an arbitrary operation on an odd number of qubits, in the one-shot case, is bounded from below by one ebit less than the teleportation bound.
By a similar calculation, one can show, using the bound in (4.17), that in the one-shot case, if classical information is to be transmitted in integer amounts, then the minimum number of bits needed to carry out an arbitrary operation on an odd number of qubits is bounded from below by 3 bits less than the teleportation bound. For $`N=3,5`$, a stronger bound of 2 bits less than resource communication for the teleportation protocol is obtained.
With these observations in mind, the case of $`N=3`$ appears to be particularly significant. For this case, in the one-shot scenario, we see that at least 3 ebits and 6 classical bits are required. However, we know that a permutation of 3 qubits can create 3 ebits or be used to send 6 classical bits. This implies that these bounds must also hold asymptotically.
It is important to compare the bounds in (4.15) and (4.17), which hold rigorously in both the one-shot and asymptotic scenarios, with the maximum amount of entanglement that can be created, and the maximum amount of classical information that can be sent, by an $`N`$ qubit operation. With this in mind, we note the following inequality, which holds for all $`N3`$:
$$2\left(\frac{N}{N+1}\right)(N1)N.$$
(4.19)
The equality is obtained only when $`N=3`$. This implies that, for all $`N4`$, the resources required to carry out an arbitrary operation exceed those that can be recovered, either by re-establishing consumed entanglement, or using the resource communication which was consumed to implement the operation to send useful messages.
As we saw in section II, this is not the case for $`N=2`$, which can be seen from the properties of the SWAP operation. The remaining case, that of $`N=3`$, is presently unsolved.
### D Transfer of Expendable Resources
In our derivation of the lower bounds on the minimum resource entanglement and communication needed to carry out any multiqubit operation, we used specific operations where certain pairs of laboratories needed to be able to establish large amounts of entanglement or communicate large amounts of classical information: more than is represented by the corresponding edges in the graphs $`\stackrel{~}{G}_E`$ and $`\stackrel{~}{G}_C`$. Thus, to carry out either the $`N!`$-fold PS or PS+CP operation, the resources from the other edges in these graphs must somehow be ‘transferred’ to the edges which must gain resources.
We can formalise this notion in the following way: consider a multiqubit operation $``$ on $`N`$ qubits. If $``$ is carried out $`N!`$ times, then depending on the initial conditions, there may be some pairs of laboratories which will end up sharing more that e ebits of entanglement, or exchanging more than c classical bits in either, or perhaps both directions. Let is define the target entanglement and communication graphs $`G_E^T`$ and $`G_C^T`$. These will represent either the number of ebits shared by each pair of laboratories or the number of classical bits communicated, following the $`N!`$-fold implementation of $``$. These graphs can be characterised by the target entanglement and communication matrices $`𝐄_T=\{E_T^{ij}\}`$ and $`𝐂_T=\{C_T^{ij}\}`$, in the same way as for the resource graphs.
For $`\stackrel{~}{G}_E`$ and $`G_E^T`$, we will define a pair of complementary subsets of the edge set, $`S_{E+}`$ and $`S_E`$, in the following way: $`S_{E+}`$ is the subset of the edge set, where each edge is denoted by the unordered pair $`(i,j)`$, such that $`E_T^{ij}>e`$. The set $`S_E`$ contains all edges for which $`E_T^{ij}e`$. These sets contain the edges which respectively gain, and do not gain entanglement.
Similarly, for the classical communication graphs $`\stackrel{~}{G}_C`$ and $`G_C^T`$, we will define the subsets $`S_{C+}`$ and $`S_C`$ of the edge set. $`S_{C+}`$ contains the edges, represented by ordered pairs $`[i,j]`$, for which $`C_T^{ij}>c`$, and $`S_C`$ contains all edges $`[i,j]`$ for which $`C_T^{ij}c`$.
Here, we shall be particularly interested in the edges which gain resources. In fact, the resources contained in the other edges, contained in the sets $`S_E`$ and $`S_C`$, will be considered expendable. The total expendable entanglement and communication are given by
$`E_E`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{(i,j)S_E}{}}\stackrel{~}{E}_R^{ij},`$ (4.20)
$`C_E`$ $`=`$ $`{\displaystyle \underset{[i,j]S_C}{}}\stackrel{~}{C}_R^{ij}.`$ (4.21)
The question we would like to answer is: how much of the expendable entanglement or communication can be transferred to the set $`S_{E+}`$ or $`S_{C+}`$? We have been unable to obtain the general solution to this problem, although the analysis of the PS operation suggests intuitively appealing upper bounds.
For the $`N!`$-fold PS operation, the values of $`E_E`$ and $`C_E`$ are easily calculated, where the sets $`S_{E+}`$ and $`S_{C+}`$ contain the edges linking laboratories whose qubits are to be swapped. We find that
$`E_E`$ $`=`$ $`{\displaystyle \frac{1}{2}}(N^22N)e,`$ (4.22)
$`C_E`$ $`=`$ $`(N^22N)c.`$ (4.23)
If each pair of swapped qubits generates 2 ebits of entanglement, then as we know, the $`N!`$-fold PS operation can by used to create $`N!N`$ ebits. The total amount of entanglement which has been added to the set $`S_{E+}`$ is then $`N!N(Ne/2)`$ ebits. From inequality (4.10), we see that
$$N!N\frac{Ne}{2}\frac{E_E}{2}$$
(4.24)
that is, at most half of the expendable entanglement can be added to the edges in $`S_{E+}`$. Whether or not this bound holds in general for all $`N`$, and when the initial resource entanglement is not described by a regular, complete graph, is currently unknown. However, we can prove that it holds in general for $`N=3`$. Consider 3 laboratories, $`A_1\mathrm{}A_3`$. Let their initial and final entanglement be described by the resource and target entanglement graphs $`G_E`$ and $`G_E^T`$, characterised by the corresponding matrices $`𝐄_R`$ and $`𝐄_T`$. The difference between the initial and final entanglement between each pair of laboratories can be represented by the matrix $`𝚫=𝐄_T𝐄_R`$.
The fact that the total amount of entanglement shared by one laboratory and the other pair cannot increase implies that the sum of the elements in each row or column of $`𝚫`$ cannot exceed zero. This also implies that the entanglement between at most one pair of laboratories can increase. Let this pair of laboratories be $`A_1`$ and $`A_2`$. Summing up the elements of $`𝚫`$ in rows 1 and 2, together with the nonincreasing property of the column sums, gives $`\mathrm{\Delta }_{12}|\mathrm{\Delta }_{13}+\mathrm{\Delta }_{23}|/2`$. The numerator on the right hand side is the total amount of entanglement lost. We see that the entanglement transferred to the edge (1,2) cannot exceed half of this loss.
This kind of entanglement loss was originally discovered in association with entanglement swapping. It would be useful to know whether or not it is an unavoidable feature of all operations which transfer entanglement, and for an arbitrary number of spatially separated systems. Any proof, or disproof, of this conjecture must take into account the possibility that the initial bipartite resource entanglement is converted into multiparticle entangled states. Some progress has recently been made towards developing a theory of conversion between bipartite and multiparticle entangled states. The study of certain particular situations has indicated that these conversions are typically lossy. Consequently, we do not believe that multiparticle entangled states will enable more efficient entanglement transfer.
Returning to the $`N!`$-fold PS operation, we will show how a similar result relating the expendable communication to the communication that can be added to the edges in $`S_{C+}`$ can be obtained. The $`N!`$-fold PS operation can by used to send $`2N!N`$ bits. The total amount of communication which has been added to the set $`S_{C+}`$ is then $`2N!NNc`$ bits. From inequality (4.12), we see that
$$N!N\frac{Nc}{2}\frac{C_E}{2}$$
(4.25)
that is, at most half of the expendable communication can be added to the edges in $`S_{C+}`$.
This restriction holds in general if the expendable communication is used to transmit information indirectly between pairs of laboratories. By ‘indirectly’, we mean the following. the weight of an edge in a resource communication graph is equal to the number of bits that one party can transmit along a channel to some other party, without passing through some intermediate laboratory. Clearly, the sender can transmit more information to the receiver if he sends some information via some intermediate laboratories. By indirect communication, we mean this relaying procedure.
Thus, if the sender wishes to send $`\kappa `$ bits indirectly, he will use up at least $`\kappa `$ bits of resource communication sending this information to the intermediate parties, who will in turn use up at least a further $`\kappa `$ bits of resource communication relaying it to the receiver. So, the $`\kappa `$ bits actually communicated from sender to receiver cannot exceed the lower bound of 2$`\kappa `$ bits depleted from the resource communication.
For the remainder of this section, we will make the hypothesis that at most half of the expendable resources can be transferred is of general validity, and explore its consequences for odd N and the PS+CP operation. We will show that this leads to tighter bounds than (4.15) and (4.17) on the amount of resource entanglement and communication needed to carry out an arbitrary operation on an odd number of qubits.
Let us return to the $`N!`$-fold PS+CP operation. Beginning with entanglement, the amount of entanglement transferred to the edges in $`S_{E+}`$ has two contributions. The entanglement transferred by the PS part of the operation is easily calculated to be $`(N3)(2N!e)/2`$ ebits. The amount transferred by the CP part is $`3(N!e)`$ ebits. The expendable entanglement is that initially represented by all other edges in the graph $`\stackrel{~}{G}_E`$, and is found to be
$$E_E=\left[\frac{N^2}{2}N\frac{3}{2}\right]e.$$
(4.26)
The assumption that at most half of the expendable entanglement can be transferred to the edges in $`S_{E+}`$ leads to the inequality
$$N!Ne\left[\frac{N3}{2}+3\right]+\frac{E_E}{2}.$$
(4.27)
From the relationship between e and the initial resource entanglement $`E_R`$, expressed in Eq. (4.7), we find
$$E_R\frac{2(N1)}{1+3/N^2}.$$
(4.28)
Similar reasoning can be applied to the required minimum classical communication. If at most half of the expendable communication is transferable, then the minimum number of classical bits required to perform the PS+CP operation with odd $`N`$ is bounded by
$$C_R\frac{4(N1)}{1+3/N^2}.$$
(4.29)
Assuming that inequalities (4.28) and (4.29) hold, let us deduce the minimum integer resources for the one shot case, as we did in the previous subsection. To this end, we note the inequality
$$\frac{2(N1)}{1+3/N^2}2(N1)1,$$
(4.30)
for $`N3`$, with the equality only being attained when $`N=3`$. From this inequality, we see that for all $`N4`$, the minimum integer resource entanglement is equal to that required to implement the teleportation protocol. By a similar calculation, one can show that for all $`N>3`$, the minimum integer resource communication is at least equal to one bit short of the teleportation bound, and that for all $`N12`$, it is the teleportation bound.
Figure (16) illustrates the main asymptotic bounds we have considered in this paper: the teleportation bound, the bound derived from the PS+CP operation derived in the previous subsection and the bound derived from the PS+CP operation based on the assumption that only half of the expendable resources are transferable.
To summarise this subsection, we have worked on the assumption that at most half of the expendable resources can be transferred. The analysis of the PS operation supports this conjecture that it is true for all $`N`$. If it indeed is true in general, then in the one-shot case, the teleportation protocol is optimal with regard to the resource entanglement for all odd $`N3`$, and also in terms of the resource communication for all odd $`N13`$, if only integer resources are permitted.
## V Discussion.
In this paper, we have examined the properties of collective quantum operations performed upon spatially separated quantum systems. We have considered a network of $`N`$ spatially separated laboratories, each of which contains one qubit. The network is equipped with facilities for classical communication and local quantum operations, and each pair of laboratories also shares bipartite pure entanglement.
This scenario we have considered helps to emphasise the fact that the final state of each system will depend upon the initial states of the others. The evolution thus requires information to be exchanged between the systems. In classical physics, this is simply classical information. If the systems are quantum mechanical, then the exchange of quantum information is necessary.
The transmission of quantum information from one location to another can be achieved by sending quantum systems, or by quantum teleportation. We have proposed a simple teleportation-based protocol which allows any quantum operation to be performed upon $`N`$ separated, identical quantum systems. Teleportation requires the transmission of classical information and existence of entanglement shared between the sending and receiving stations. In the case of $`N=2`$, one particular class of operations, namely those equivalent, up to bipartite local unitary transformations, to the SWAP operation, permits either the minimum classical communication or entanglement resources required to perform any operation to be ‘recovered’ for other tasks. These operations may be regarded as the most inseparable operations for $`N=2`$.
For $`N>2`$, the situation is more interesting. For $`N4`$, no operation can establish the entanglement or be used to communicate the information necessary to perform any operation. Whether or not this is also the case for $`N=3`$ is currently unknown. For all $`N`$ we have determined the maximum total amount of entanglement that can be established, and the maximum total number of classical bits that can be communicated, by any operation. Permutation operations attain these limits, which are also the minimum resources required to carry out these specific operations.
We have also examined the problem of finding the minimum resources required to perform an arbitrary operation. The scenario we considered was one where each pair of laboratories shares a certain amount of entanglement, and can communicate a certain number of classical bits to each other. The problem we addressed was: what are the minimum values of the total entanglement and communication required to carry out an a priori unknown operation, that is, unknown prior to the entanglement and communication resources being set up?
For even $`N`$, we have found these minimum resources exactly, and these can be used to perform an arbitrary operation using teleportation. We arrived at these bounds using a technique we refer to as graph symmetrisation. We have shown that the teleportation protocol is optimal for even N. Whether or not it is also optimal for odd N is an important outstanding problem. We have shown, in the even case, that the optimality of the teleportation protocol can be reinterpreted as coming from a restriction on the extent to which expendable resources can be transferred from one pair of laboratories to another. In particular, for any amount of resources transferred, at least as much are irrevocably lost. The assumption that this restriction always holds leads to tighter bounds on the resources required to carry out an arbitrary operation on an odd number of qubits. These bounds imply that if, in the one shot scenario, resources can only be consumed in integer amounts, then the teleportation protocol is optimal, for all $`N>3`$ for entanglement and for all $`\mathrm{N}12`$ for communication also. One clear conclusion from our work is that the case of $`N=3`$ is of particular interest, since many of our results which apply to all other $`N2`$ have not been established for this case. It could be that graph-theoretic techniques are not suitable for analysing the 3 qubit case, and that other tools must be employed.
## Acknowledgements.
We would like to thank Sandu Popescu, Noah Linden, Osamu Hirota and Masahide Sasaki for interesting discussions. We also wish to thank John Vaccaro for help with some of the figures. Part of this work was carried out at the Japanese Ministry of Posts and Telecommunications Communications Research Laboratory, Tokyo, and we would like to thank Masayuki Izutsu for his hospitality. This work was funded by the UK Engineering and Physical Sciences Research Council, and by the British Council.
|
warning/0006/hep-ph0006201.html
|
ar5iv
|
text
|
# Event-by-event fluctuations of the charged particle ratio from non-equilibrium transport theory
## Acknowledgements
This work was supported by the Director, Office of Science, Office of High Energy and Nuclear Physics, Division of Nuclear Physics, and by the Office of Basic Energy Sciences, Division of Nuclear Sciences, of the U.S. Department of Energy under Contract No. DE-AC03-76SF00098. M.B. was further supported by the A. v. Humboldt foundation. This research used resources of the National Energy Research Scientific Computing Center (NERSC).
|
warning/0006/math0006179.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The algebra of the Euclidean quantum space $`_q^3`$ has been discussed in . Its defining relations are:
$`X^3X^+q^2X^+X^3=0`$
$`X^3X^{}q^2X^{}X^3=0`$
$`X^{}X^+X^+X^{}=\lambda X^3X^3`$
$`\overline{X^3}=X^3,\overline{X^+}=qX^{}`$
$`\lambda =qq^1,q,q1.`$
This space is a $`so_q(3)`$-module algebra. The whole set of relations can be found in . The generators of the quantum Lie algebra $`so_q(3)`$ are interpretated as angular momentum operators.
In this paper we are going to show that the elements of $`_q^3`$ and $`so_q(3)`$ can be represented by differential operators in $`^3`$. We use polar coordinates $`(r,\theta ,\phi )`$, $`(\xi \mathrm{cos}\theta )`$ and find
$`X^3`$ $`=`$ $`r\xi `$
$`X^+`$ $`=`$ $`re^{i\phi }\sqrt{{\displaystyle \frac{1q^2\xi ^2}{1+q^2}}}q^{2\xi \frac{}{\xi }1}`$ (1.2)
$`X^{}`$ $`=`$ $`+re^{i\phi }\sqrt{{\displaystyle \frac{1q^2\xi ^2}{1+q^2}}}q^{2\xi \frac{}{\xi }+1}`$
for the coordinates. The generators of the $`q`$-deformed orbital angular momentum are represented as follows:
$`T_{orb}^3`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}\left(1q^{4i\frac{}{\phi }}\right)`$
$`T_{orb}^+`$ $`=`$ $`{\displaystyle \frac{e^{i\phi }}{q\lambda \xi }}\sqrt{1q^2\xi ^2}q^{2\xi \frac{}{\xi }}`$ (1.3)
$`{\displaystyle \frac{1}{q\lambda \xi }}\sqrt{1q^2\xi ^2e^{4i\frac{}{\phi }}}e^{i\phi }`$
$`T_{orb}^{}`$ $`=`$ $`{\displaystyle \frac{qe^{i\phi }}{\lambda \xi }}\sqrt{1q^2\xi ^2}q^{2\xi \frac{}{\xi }}`$
$`{\displaystyle \frac{q}{\lambda \xi }}\sqrt{1q^2\xi ^2e^{4i\frac{}{\phi }}}e^{i\phi }`$
All these operators (1), (1) have the expected limit (2), (5.5) for $`q1`$.
A similar result for $`so_q(3)`$ has been obtained in , where its generators have been constructed in terms of generators of $`so(3)`$.
In the representation theory of the algebra has been studied. It was found that the representation is unique (apart from a scaling factor for the radius) if we demand that the conjugation properties
$`\overline{X^3}=X^3,\overline{X^+}=qX^{}`$
$`\overline{T^3}=T^3,\overline{T^+}=q^2T^{}`$ (1.4)
are represented by the conjugation of linear operators in a Hilbert space and that the equal sign in (1) includes the domain on which the linear operators are defined. This domain is supposed to be dense in the Hilbert space.
The spectrum of the linear operator $`X^3`$ that was found in does not agree with the spectrum of the differential operator in (1) if we consider it as differential operator in $`L^2`$. We cannot expect the differential operators of (1) and (1) to have the desired conjugation properties as linear operators on $`L^2`$.
To obtain the representation found in we use the following strategy: we consider the space of $`C^{\mathrm{}}`$ functions on which the differential operators act. This space is made to an algebra by a convolutionary product. This algebra has an ideal that is left invariant under the action of the differential operators (1) and (1). We consider the factor space of the $`C^{\mathrm{}}`$ algebra with respect to this ideal. On this factor space we can define a scalar product making it a Hilbert space. This is the representation space where the operators (1) and (1) have the desired properties.
To achieve this we start from a basis in the $`C^{\mathrm{}}`$ space where the elements are the product of functions of $`r`$, $`\xi `$ and $`\phi `$.
The set of $`C^{\mathrm{}}`$ functions that vanish at $`r=r_0q^{4M+2}`$ for all $`M`$ forms an ideal $`_r^{r_0}`$ under pointwise multiplication in the set $`_r`$ of all $`C^{\mathrm{}}`$ functions $`f(r)`$, $`r_+`$. $`r_0`$ is an arbitrary positive parameter, the scaling factor mentioned above. Since the differential operators (1) and (1) do not change the radius, it would be enough, to consider functions that vanish at one fixed $`r`$. But if one wants to introduce real momenta, one needs a scaling operator $`\mathrm{\Lambda }`$, such that $`\mathrm{\Lambda }R=q^4R\mathrm{\Lambda }`$ . Therefore we consider a whole $`q`$-lattice in radial direction.
We introduce the factor space
$$H_r^{r_0}=\frac{_r}{_r^{r_0}}.$$
(1.5)
The scalar product that makes this space a Hilbert space, which we denote by $`_r^{r_0}`$, is formulated with the Jackson integral:
$$(g,f)=\underset{M=\mathrm{}}{\overset{\mathrm{}}{}}q^{4M}g^{}(q^{4M+2}r_0)f(q^{4M+2}r_0).$$
(1.6)
The eigenvectors of the multiplication operator $`r`$ with the eigenvalues $`r_0q^{4M+2}`$ form a basis in this Hilbert space. We denote these vectors by $`u_M`$:
$$ru_M=r_0q^{4M+2}u_M.$$
(1.7)
For the set $`_\xi `$ of functions $`f(\xi )`$ we proceed similarly. The product is again the pointwise product of the functions.
The ideal is generated by the functions that vanish at $`\xi =\pm q^{2m1}`$ for all $`m,m0`$, we call it $`_\xi `$. The representation space is the factor space
$$H_\xi =\frac{_\xi }{_\xi }.$$
(1.8)
The scalar product that makes it a Hilbert space $`_\xi `$ is again defined with the help of the Jackson integral:
$$(\psi ,\varphi )=\underset{\sigma =\pm 1}{}\underset{m=\mathrm{}}{\overset{0}{}}q^{2m}\psi ^{}(\sigma q^{2m1})\varphi (\sigma q^{2m1}).$$
(1.9)
The eigenfunctions of $`\xi `$ in this Hilbert space will be denoted by $`\chi _{\pm m_t}`$:
$$\xi \chi _{\pm m_t}=\pm q^{2m_t1}\chi _{\pm m_t}.$$
(1.10)
For the set $`\stackrel{~}{}_\phi `$ of functions $`f(\phi )`$ we define the product by the convolution:
$$(\stackrel{~}{fg})_m=\stackrel{~}{f}_m\stackrel{~}{g}_m,$$
(1.11)
where $``$ denotes the Fourier transformation. This defines an algebra. The functions for which $`\stackrel{~}{f}_m=0`$ for $`m<\underset{¯}{m}`$ form an ideal. We construct the factor space with the scalar product
$$(h,g)=\underset{m=\underset{¯}{m}}{\overset{\mathrm{}}{}}\stackrel{~}{h}_m^{}\stackrel{~}{g}_m.$$
(1.12)
This will lead to the representation space when we allow $`\underset{¯}{m}`$ to depend on $`m_t`$.
In this space we have the following basis:
$`\psi _{M,m_t,m}`$ $`=`$ $`u_M\chi _{m_t}e^{im\phi }`$ (1.13)
$`M`$ $`=`$ $`\mathrm{}\mathrm{}\mathrm{}`$
$`m_t`$ $`=`$ $`\mathrm{}\mathrm{}0`$
$`m`$ $`=`$ $`m_t\mathrm{}\mathrm{}`$
To define the scalar product for functions $`\psi (r,\xi ,\phi )`$ we fouriertransform with respect to $`\phi `$ to obtain $`\stackrel{~}{\psi }_m(r,\xi )`$.
$`(\varphi ,\psi )`$ $`=`$ $`{\displaystyle \underset{M=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{m_t=\mathrm{}}{\sigma =\pm 1}}{\overset{0}{}}}{\displaystyle \underset{m=m_t}{\overset{\mathrm{}}{}}}q^{4M}q^{2m_t}\times `$
$`\times \stackrel{~}{\varphi }_m^{}(q^{4M+2}r_0,\sigma q^{2m_t1})\stackrel{~}{\psi }_m(q^{4M+2}r_0,\sigma q^{2m_t1}).`$
## 2 The $`𝐗`$-Algebra
Our aim is to represent the algebra (1) in terms of differential operators acting on $`^3`$. We use polar coordinates $`(r,\theta ,\phi )`$ and $`\xi =\mathrm{cos}\theta `$. An operator that will play a major role in this attempt is:
$$𝒵_\xi =\frac{1}{2}\left(\xi \frac{}{\xi }+\frac{}{\xi }\xi \right)=\xi \frac{}{\xi }+\frac{1}{2}.$$
(2.1)
It is defined in such a way that when acting on $`L^2`$-functions in the common domain of $`𝒵_\xi `$ and $`𝒵_\xi ^{}`$,
$$𝒵_\xi ^{}=𝒵_\xi $$
(2.2)
holds. The property of $`𝒵_\xi `$ that we will use frequently is:
$`[𝒵_\xi ,\xi ]=\xi ,`$
$`e^{\alpha 𝒵_\xi }\xi e^{\alpha 𝒵_\xi }=e^\alpha \xi .`$ (2.3)
We now make an ansatz:
$`X^3`$ $`=`$ $`r\xi `$
$`X^{}`$ $`=`$ $`Arf(\xi )e^{2\alpha 𝒵_\xi },f(0)=1,`$ (2.4)
$`X^+`$ $`=`$ $`Brg(\xi )e^{2\beta 𝒵_\xi },g(0)=1.`$
From the first two equations of (1) follows:
$$e^{2\alpha }=e^{2\beta }=q^2,$$
(2.5)
and from the third equation:
$$AB\left\{f(\xi )g(q^2\xi )f(q^2\xi )g(\xi )\right\}=\lambda \xi ^2.$$
(2.6)
With the definition
$$\varphi (\xi )=ABf(\xi )g(q^2\xi )$$
(2.7)
this equation becomes:
$$\varphi (\xi )\varphi (q^2\xi )=\lambda \xi ^2$$
(2.8)
and has the solution:
$$\varphi (\xi )=\varphi (0)+\frac{q^3}{1+q^2}\xi ^2.$$
(2.9)
It is natural to identify the radius $`r`$ with the invariant length in $`_q^3`$:
$$r^2=R^2X^3X^3qX^+X^{}q^1X^{}X^+.$$
(2.10)
This determines $`\varphi (0)`$:
$$\varphi (0)=\frac{q}{1+q^2}.$$
(2.11)
To obtain $`f`$ as well as $`g`$ from $`\varphi `$ we have to use the conjugation property
$$\overline{X^+}=qX^{},$$
(2.12)
that leads to
$$Bg(q^2\xi )=q\overline{A}\overline{f}(\xi )$$
(2.13)
and, as a consequence of the definition (2.7) of $`\varphi `$:
$$\varphi =q|Af|^2.$$
(2.14)
Introducing $`\mathrm{\Lambda }_\xi q^{2𝒵_\xi }`$, such that $`\mathrm{\Lambda }_\xi \xi =q^2\xi \mathrm{\Lambda }_\xi `$, and combining (2.9), (2.11) and (2.14) leads to the result:
$`X^3`$ $`=`$ $`r\xi ,`$
$`X^+`$ $`=`$ $`re^{i\phi }\sqrt{{\displaystyle \frac{1q^2\xi ^2}{1+q^2}}}\mathrm{\Lambda }_\xi ^1,`$
$`X^{}`$ $`=`$ $`re^{i\phi }\sqrt{{\displaystyle \frac{1q^2\xi ^2}{1+q^2}}}\mathrm{\Lambda }_\xi .`$ (2.15)
We have found a representation of the $`X`$-algebra. In the limit $`q1`$ we obtain:
$`X^3`$ $``$ $`r\mathrm{cos}\theta `$
$`X^+`$ $``$ $`{\displaystyle \frac{1}{\sqrt{2}}}r\mathrm{sin}\theta e^{i\phi }`$ (2.16)
$`X^{}`$ $``$ $`+{\displaystyle \frac{1}{\sqrt{2}}}r\mathrm{cos}\theta e^{i\phi }.`$
## 3 The $`𝐭`$-algebra
There is a homomorphism of the $`T`$-algebra into the $`X`$-algebra :
$`t^+`$ $`=`$ $`{\displaystyle \frac{1}{\lambda q^3}}\sqrt{1+q^2}X^+(X^3)^1`$
$`t^{}`$ $`=`$ $`{\displaystyle \frac{q^2}{\lambda }}\sqrt{1+q^2}X^{}(X^3)^1`$ (3.1)
$`t^3`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}\left(1+R^2(X^3)^2\right).`$
With (2) this $`t`$-algebra can be represented by differential operators:
$`t^3`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}(1+\xi ^2)`$
$`t^+`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}e^{i\phi }\xi ^1\sqrt{1q^2\xi ^2}\mathrm{\Lambda }_\xi ^1`$ (3.2)
$`t^{}`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}e^{i\phi }\xi ^1\sqrt{1q^2\xi ^2}\mathrm{\Lambda }_\xi .`$
These are differential operators acting on $`C^{\mathrm{}}`$ functions. We cannot expect that they are defined on a dense set in $`L^2`$ such that the conjugation properties
$$\overline{t^+}=q^2t^{},\overline{t^3}=t^3$$
(3.3)
hold for the differential operators when the conjugation is identified with the conjugation of linear operators in the Hilbert space $`L^2`$:
$$(t^+)^{}=q^2t^{},(t^3)^{}=t^3.$$
(3.4)
From we actually know that this cannot be the case because we found that such a representation of the $`t`$-algebra is unique and leads to a spectrum of $`t_3`$ with the eigenvalues
$$\frac{1}{\lambda }\left(1+q^2q^{4m_t}\right),m_t0.$$
(3.5)
For $`\xi `$ this implies that only eigenvalues $`q^{2m1},m0`$ are allowed. Clearly, the Hilbert space of square integrable functions is not the Hilbert space that would lead to such a spectrum.
The spectrum of $`t^3`$ suggests that we should consider a factor space of the $`C^{\mathrm{}}`$ functions of the following type:
Consider the linear space of $`C^{\mathrm{}}`$ functions on the interval $`0<\xi <1`$
$$_\xi =\{f(\xi )|fC^{\mathrm{}}((0,1))\}$$
(3.6)
and the subspace generated by the functions
$$_\xi =\{hC^{\mathrm{}}([0,1])|h(\xi _m)=0\text{ for }\xi _m=q^{2m1},m0\}.$$
(3.7)
Under pointwise multiplication these functions form an algebra, which we also call $`_\xi `$ and $`_\xi `$, respectively. The algebra $`_\xi `$ is an ideal of $`_\xi `$ and we can define the factor space
$$H_\xi \frac{_\xi }{_\xi }.$$
(3.8)
On this factor space $`t^3`$ is defined and has the desired eigenvalues. Eigenvectors of $`t^3`$ with the eigenvalue $`\frac{1}{\lambda }\left(1+q^2q^{4m_t}\right)`$ we shall denote by $`\chi _{m_t}`$.
Next we show that the ideal $`_\xi `$ is left invariant by the action of $`t^+`$ and $`t^{}`$
$`t^+f(\xi )`$ $`=`$ $`{\displaystyle \frac{e^{i\phi }}{\lambda }}{\displaystyle \frac{1}{\xi }}\sqrt{1q^2\xi ^2}\mathrm{\Lambda }_\xi ^1f(\xi )`$ (3.9)
$`=`$ $`{\displaystyle \frac{e^{i\phi }}{\lambda }}{\displaystyle \frac{1}{\xi }}\sqrt{1q^2\xi ^2}q^1f(q^2\xi )`$
For $`f_\xi `$ follows $`t^+f_\xi `$.
Analogous:
$$t^{}f(\xi )=\frac{e^{i\phi }}{\lambda }\frac{1}{\xi }\sqrt{1q^2\xi ^2}qf(q^2\xi ).$$
(3.10)
$`t^{}`$ shifts the points $`\xi =q^{2m1}`$ to the points $`\xi =q^{2(m+1)1}`$. In the definition of $`_\xi `$ we have $`m0`$, but $`t^{}f(\xi )|_{\xi =q^1}=0`$, as can be seen from (3.10). Thus $`_\xi `$ is invariant under the action of $`t^{}`$ as well. It follows that $`t^3`$, $`t^+`$ and $`t^{}`$ are well defined on $`H_\xi `$.
We are now going to show that we can define a scalar product on $`H_\xi `$ to get an Hilbert space $`_\xi `$, such that $`(t^+)^{}=q^2t^{}`$:
$$(\psi ,\varphi )=\underset{m=\mathrm{}}{\overset{0}{}}\psi ^{}(\xi _m)\varphi (\xi _m)q^{2m},\xi _m=q^{2m1}.$$
(3.11)
We compute:
$`(\psi ,t^+\varphi )`$ $`=`$ $`{\displaystyle \underset{m=\mathrm{}}{\overset{0}{}}}q^{2m1}\psi ^{}(\xi _m){\displaystyle \frac{e^{i\phi }}{\lambda }}{\displaystyle \frac{\sqrt{1q^2\xi _m^2}}{\xi _m}}\varphi (\xi _{m1})`$ (3.12)
$`=`$ $`{\displaystyle \underset{m=\mathrm{}}{\overset{1}{}}}q^{2m+1}\psi ^{}(\xi _{m+1}){\displaystyle \frac{e^{i\phi }}{\lambda }}{\displaystyle \frac{\sqrt{1q^2\xi _{m+1}^2}}{\xi _{m+1}}}\varphi (\xi _m)`$
$`=`$ $`{\displaystyle \underset{m=\mathrm{}}{\overset{0}{}}}\left({\displaystyle \frac{e^{i\phi }}{\lambda }}{\displaystyle \frac{\sqrt{1q^2\xi ^2}}{\xi }}q^{2𝒵_\xi }\psi \right)^{}(\xi _m)\varphi (\xi _m)q^{2m2}`$
$`=`$ $`{\displaystyle \frac{1}{q^2}}(t^{}\psi ,\varphi ).`$
We have changed the summation index and extended the sum to include $`m=0`$ because the summand vanishes there.
There is a differential operator that commutes with the differential operators $`\stackrel{}{t}`$ of (3):
$`\widehat{\xi }\xi q^{2i\frac{}{\phi }}`$ (3.13)
$`[\widehat{\xi },\stackrel{}{t}]=0,[\widehat{\xi },R]=0,[\widehat{\xi },\stackrel{}{X}]=0.`$
We shall use this operator to represent the $`K`$-algebra of ref in terms of differential operators.
## 4 The $`𝐊`$-Algebra
The elements of the $`K`$-algebra as they were defined in ref all commute with the $`X`$ and the $`t`$ algebra. The $`K`$-relations are
$`q^2K^3K^+q^2K^+K^3`$ $`=`$ $`(q+q^1)K^+`$
$`q^2K^3K^{}+q^2K^{}K^3`$ $`=`$ $`(q+q^1)K^{}`$ (4.1)
$`q^1K^+K^{}qK^{}K^+`$ $`=`$ $`K^3={\displaystyle \frac{1}{\lambda }}(1\tau _k)`$
and
$$\overline{K^3}=K^3,\overline{K^+}=q^2K^{}.$$
(4.2)
The representation of the $`K`$-algebra that has to be used for orbital angular momentum has eigenvalues of $`\tau _k`$ of the form $`q^{4m_k2}`$, $`m_k0`$. This motivates the ansatz
$$\tau _k=\widehat{\xi }^2.$$
(4.3)
From
$$\widehat{\xi }e^{i\phi }=q^2e^{i\phi }\widehat{\xi }$$
(4.4)
follows that a promising ansatz for $`K^+`$ and $`K^{}`$ is:
$`K^+`$ $`=`$ $`h(\widehat{\xi })e^{i\phi }`$
$`K^{}`$ $`=`$ $`j(\widehat{\xi })e^{i\phi }.`$ (4.5)
It satisfies the first two relations of (4), the third one leads to a recursion formula for
$$J(\widehat{\xi })=h(\widehat{\xi })j(q^2\widehat{\xi }).$$
(4.6)
We find:
$$J(\widehat{\xi })q^2J(q^2\widehat{\xi })=\frac{q}{\lambda }(1+\widehat{\xi }^2)$$
(4.7)
with the solution
$$J(\widehat{\xi })=\frac{1}{\lambda ^2}\left\{1+\beta \widehat{\xi }q^2\widehat{\xi }^2\right\}$$
(4.8)
$`\beta `$ being a free parameter, not determined by (4.7).
From the conjugation property (4.2) follows
$$j(\widehat{\xi })=q^2\overline{h}(q^2\widehat{\xi })$$
(4.9)
if
$$\overline{\left(i\frac{}{\phi }\right)}=i\frac{}{\phi }.$$
(4.10)
The parameter $`\beta `$ is determined to be zero by the orbital angular momentum condition. This can be seen by a direct calculation following all the steps outlined in .
The result is:
$`K^+`$ $`=`$ $`{\displaystyle \frac{e^{i\vartheta }}{q^21}}\sqrt{1q^2\widehat{\xi }^2}e^{i\phi }`$
$`K^{}`$ $`=`$ $`{\displaystyle \frac{q^2e^{i\vartheta }}{q^21}}\sqrt{1q^2\widehat{\xi }^2}e^{i\phi }.`$ (4.11)
We now turn our attention to the representation space of the $`K`$-algebra. The operator $`\xi `$ is represented on the factor space $`_\xi `$ defined in (3.8). The eigenvectors of $`t^3`$ were denoted by $`\chi _{m_t}`$, from (3) we learn:
$$\xi \chi _{m_t}=q^{2m_t1}\chi _{m_t},m_t0.$$
(4.12)
The eigenvectors of $`\widehat{\xi }`$ (3.13) will be of the form $`\chi _{m_t}e^{im\phi }`$:
$$\widehat{\xi }\chi _{m_t}e^{im\phi }=q^{2(m_tm)1}\chi _{m_t}e^{im\phi }.$$
(4.13)
These are the eigenfunctions of $`\tau _k`$:
$$\tau _k\chi _{m_t}e^{im\phi }=q^{4(m_tm)2}\chi _{m_t}e^{im\phi }.$$
(4.14)
From we know that the eigenvalues of $`\tau _k`$ are $`q^{4m_k2},m_k0`$. It follows that
$$m=m_t+m_k,m_k0,m_t0.$$
(4.15)
For $`m_t`$ fixed we find the condition $`mm_t`$. This is in agreement with the expression of $`K^{}`$ in (4). The operator $`K^{}`$ changes the eigenvalue of $`K^3`$ from $`m_k`$ to $`m_k1`$. When applied to the eigenvector $`m_k=0`$, it should give zero.
$`K^{}\chi _{m_t}e^{im_t\phi }`$ $`=`$ $`q^2{\displaystyle \frac{e^{i\vartheta }}{q^21}}\sqrt{1q^2\widehat{\xi }^2}\chi _{m_t}e^{i(m_t1)\phi }`$ (4.16)
$`=`$ $`0.`$
To show this we use (4) and the action of $`\widehat{\xi }`$ (4.13). In this way we could have found the condition $`m_k0`$.
We see that for a given eigenvalue of $`\xi `$, the space of functions on which the $`K`$-algebra is represented is given by functions of $`\phi `$ with a truncated Fourier transformation:
$$g(\phi )=\frac{1}{\sqrt{2\pi }}\underset{m=m_t}{\overset{\mathrm{}}{}}c_me^{im\phi }.$$
(4.17)
This space of functions is invariant under the $`K`$-algebra, it was sufficient to show this for $`K^{}`$ (Eqn (4.16)) because $`K^+`$ shifts the eigenvalue from $`m`$ to $`m+1`$, and $`K^3`$ does not change the eigenvalue.
If we define the product of two functions as the convolution defined as product of the Fourier transformation we again have constructed an ideal by (4.17).
The factor space of the $`C^{\mathrm{}}`$ functions of $`\phi ,(0\phi 2\pi )`$ with respect to this ideal we call $`\stackrel{~}{H}_\phi ^{m_t}`$. On this space a scalar product is defined:
$$(h,g)=\underset{m=m_t}{\overset{\mathrm{}}{}}\stackrel{~}{h}_m^{}\stackrel{~}{g}_m,$$
(4.18)
where $`\stackrel{~}{h}`$ and $`\stackrel{~}{g}`$ stand for the Fourier transformation of $`h`$ and $`g`$. With this scalar product the conjugation property of the $`K`$-algebra (4.2) becomes
$$(K^3)^{}=K^3,(K^+)^{}=q^2K^{}.$$
(4.19)
This can easily be verified by a resummation and the use of (4.16).
## 5 Orbital angular momentum
Orbital angular momentum has been defined in :
$`\tau _{orb}`$ $`=`$ $`\tau _t\tau _k`$
$`T_{orb}^3`$ $`=`$ $`t^31+\tau _tK^3`$
$`T_{orb}^+`$ $`=`$ $`t^+1+\sqrt{\tau _t}K^+`$ (5.1)
$`T_{orb}^{}`$ $`=`$ $`t^{}1\sqrt{\tau _t}K^{}`$
For the differential operators this becomes:
$`\tau _{orb}`$ $`=`$ $`q^{4i\frac{}{\phi }}`$
$`T_{orb}^3`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}\left(1q^{4i\frac{}{\phi }}\right)`$
$`T_{orb}^+`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}e^{i\phi }{\displaystyle \frac{1}{\xi }}\sqrt{1q^2\xi ^2}\mathrm{\Lambda }_\xi ^1`$ (5.2)
$`+{\displaystyle \frac{1}{\xi }}{\displaystyle \frac{e^{i\vartheta }}{q^21}}\sqrt{1q^2\widehat{\xi }^2}e^{i\phi }`$
$`T_{orb}^{}`$ $`=`$ $`{\displaystyle \frac{1}{\lambda }}e^{i\phi }{\displaystyle \frac{1}{\xi }}\sqrt{1q^2\xi ^2}\mathrm{\Lambda }_\xi `$
$`+{\displaystyle \frac{1}{\xi }}{\displaystyle \frac{e^{i\vartheta }}{q^21}}q^2\sqrt{1q^2\widehat{\xi }^2}e^{i\phi }`$
$`𝒵_\xi =\xi {\displaystyle \frac{}{\xi }},\mathrm{\Lambda }_\xi =q^{2𝒵_\xi },\widehat{\xi }=\xi q^{2i\frac{}{\phi }}`$
The representation space with the proper conjugation properties has been constructed above.
It remains to show that in the limit $`q1`$ the operators in (5) tend to the generators of angular momentum. We take $`q=e^h`$ and study the limit $`h0`$. It is easy to see that:
$$T_{orb}^32i\frac{}{\phi }.$$
(5.3)
For $`T_{orb}^+`$ the limit is more involved, as the two parts of $`T_{orb}^+`$ have no individual limit.
$`T_{orb}^+`$ $`=`$ $`{\displaystyle \frac{1}{2h}}{\displaystyle \frac{e^{i\phi }}{\xi }}\sqrt{1q^2\xi ^2}e^{2h(\xi \frac{}{\xi }+\frac{1}{2})}`$ (5.4)
$`+e^{i\vartheta }{\displaystyle \frac{1}{2h}}{\displaystyle \frac{e^{i\phi }}{\xi }}\sqrt{1q^2\xi ^2e^{4ih\frac{}{\phi }}}.`$
For $`e^{i\vartheta }=1`$ the singular parts cancel and we obtain:
$`T_{orb}^+`$ $``$ $`e^{i\phi }\sqrt{1\xi ^2}\left\{{\displaystyle \frac{}{\xi }}+{\displaystyle \frac{\xi }{1q^2}}i{\displaystyle \frac{}{\phi }}\right\}`$ (5.5)
$``$ $`e^{i\phi }\left\{{\displaystyle \frac{}{\theta }}+\mathrm{cot}\theta i{\displaystyle \frac{}{\phi }}\right\}.`$
An analogous result is obtained for $`T_{orb}^{}`$. It is interesting to note that the phase $`e^{i\vartheta }`$ in the expression for the orbital angular momentum has been determined to be $`e^{i\vartheta }=1`$ by the requirement that the limit $`q1`$ exists. The condition that the differential operators have the correct $`q1`$ limit restricts the choise of the operators, without this condition more operators would satisfy the algebra.
|
warning/0006/hep-ph0006353.html
|
ar5iv
|
text
|
# I Introduction
## I Introduction
It has been more than a decade since the discovery of the Gourdin-Ellis-Jaffe sum rule (GEJ) violation in the polarized deep inelastic scattering (DIS) experiment by the European Muon Collaboration . The physics community was puzzled since the experimental data meant a surprisingly small contribution to the proton spin from the spins of the quarks, in contrast to the Gell-Mann-Zweig quark model in which the spin of the proton is totally provided by the spins of the three valence quarks. This gave rise to the proton “spin crisis” or “spin puzzle”, and triggered a vast number of theoretical and experimental investigations on the spin structure of the nucleon. Among them, there was an interesting contribution to understand the spin of the nucleon within a “minimal” simple quark model , where it was observed that the nucleon has only a small amplitude to be a bare three quark state $`|qqq`$, while the largest term in the wavefunction is $`|qqqQ\overline{Q}`$, in which $`Q\overline{Q}`$ denotes sea quark-antiquark pairs.
There was a prevailing impression that the proton spin structure is in conflict with the quark model. However, there has been an attempt to understand the proton spin puzzle within the quark model by using the Melosh-Wigner rotation effect , which comes from the relativistic effect of the quark intrinsic transversal motion inside the proton. It was pointed out that the quark helicity ($`\mathrm{\Delta }q`$) observed in polarized DIS is actually the quark spin defined in the light-cone formalism, and it is different from the quark spin ($`\mathrm{\Delta }q_{QM}`$) as defined in the quark model. Thus the small quark helicity sum observed in polarized DIS is not necessarily in contradiction with the quark model in which the proton spin is provided by the valence quarks . Recent progress has also been made on the Melosh-Wigner rotation effect in other physical quantities related to the spin structure of the nucleon, and the significance of the Melosh-Wigner rotation connecting the spin states between the light-front dynamics and the conventional instant-form dynamics has been widely accepted. Thus it is necessary to check what can obtained for the spin structure of the nucleon within the quark model, after we take into account the Melosh-Wigner rotation. Certainly our present understanding of the nucleon spin structure has been enriched from what we knew before the discovery of the GEJ sum rule violation, and we now know that both the sea quarks and the gluons play an important role in the spin structure of the nucleon. The purpose of this paper is to extend the simple Jaffe-Lipkin quark model to a more general framework, by including other necessary ingredients in the nucleon sea such as pseudoscalar mesons, whose addition is supported by available theoretical and experimental studies.
The paper is organized as follows. In Section II, we briefly review the Melosh-Wigner rotation effect in the quark model, and show that the introduction of an up($`u`$)-down($`d`$) quark flavor asymmetry in the Melosh-Wigner rotation factors can reproduce the present experimental data of the integrated spin structure functions for both the proton and the neutron , within a simple SU(6) quark model with only three valence quarks. In Section III, we introduce the contribution from the higher Fock states $`|BM=|qqqQ\overline{Q}`$ in which the quark and antiquark of an quark-antiquark pair are rearranged non-perturbatively with the three valence quarks into a pseudoscalar meson and a baryon, and we write the configuration as a baryon-meson (BM) fluctuation . It is shown that the consideration of the lowest $`p(uudD\overline{D})=n(udD)\pi ^+(u\overline{D})`$ fluctuation, which is supported by the observed Gottfried sum rule violation , introduces an $`u`$-$`d`$ flavor asymmetric term in the quark contributions to the nucleon and produces a reasonable $`u`$-$`d`$ Melosh-Wigner rotation asymmetry which is required to reproduce the data. In this section, we point out that the Jaffe-Lipkin term of quark-antiquark pairs (which are actually vector mesons in a baryon-meson fluctuation picture) will be important when there is need for negatively polarized antiquarks. Thus we present a new “miminal” quark model extension of Jaffe-Lipkin model, with three valence quarks, sea quark-antiquark pairs in terms of baryon-meson fluctuations where the mesons are either pseudoscalar or vector mesons, in order to understand the nucleon spin structure within the quark model framework. Finally, we present discussions and conclusions in Section IV.
## II The naive quark model and the Melosh-Winger rotation
The spin-dependent structure functions for the proton and the neutron, when expressed in terms of the quark helicity distributions $`\mathrm{\Delta }q(x)`$, should read
$$g_1^p(x)=\frac{1}{2}[\frac{4}{9}(\mathrm{\Delta }u(x)+\mathrm{\Delta }\overline{u}(x))+\frac{1}{9}(\mathrm{\Delta }d(x)+\mathrm{\Delta }\overline{d}(x))+\frac{1}{9}(\mathrm{\Delta }s(x)+\mathrm{\Delta }\overline{s}(x))],$$
(1)
$$g_1^n(x)=\frac{1}{2}[\frac{1}{9}(\mathrm{\Delta }u(x)+\mathrm{\Delta }\overline{u}(x))+\frac{4}{9}(\mathrm{\Delta }d(x)+\mathrm{\Delta }\overline{d}(x))+\frac{1}{9}(\mathrm{\Delta }s(x)+\mathrm{\Delta }\overline{s}(x))],$$
(2)
where the quantity $`\mathrm{\Delta }q(x)`$ is defined by the axial current matrix element
$$\mathrm{\Delta }q=p,\left|\overline{q}\gamma ^+\gamma _5q\right|p,.$$
(3)
By expressing the quark axial charge or the quark helicity related to the axial quark current $`\overline{q}\gamma ^\mu \gamma ^5q`$ by $`\mathrm{\Delta }Q=_0^1dx[\mathrm{\Delta }q(x)+\mathrm{\Delta }\overline{q}(x)]`$, we obtain
$$\mathrm{\Gamma }^p=_0^1dxg_1^p(x)=\frac{1}{2}(\frac{4}{9}\mathrm{\Delta }U+\frac{1}{9}\mathrm{\Delta }D+\frac{1}{9}\mathrm{\Delta }S),$$
(4)
$$\mathrm{\Gamma }^n=_0^1dxg_1^n(x)=\frac{1}{2}(\frac{1}{9}\mathrm{\Delta }U+\frac{4}{9}\mathrm{\Delta }D+\frac{1}{9}\mathrm{\Delta }S).$$
(5)
Two linear combinations of the axial charges, $`\mathrm{\Delta }Q^3=\mathrm{\Delta }U\mathrm{\Delta }D`$ and $`\mathrm{\Delta }Q^8=\mathrm{\Delta }U+\mathrm{\Delta }D2\mathrm{\Delta }S`$, are therefore given by
$$\mathrm{\Delta }Q^3=6(\mathrm{\Gamma }^p\mathrm{\Gamma }^n)=\mathrm{\Delta }U\mathrm{\Delta }D=G_A/G_V=1.261,$$
(6)
from neutron decay plus isospin symmetry, and by
$$\mathrm{\Delta }Q^8=\mathrm{\Delta }U+\mathrm{\Delta }D2\mathrm{\Delta }S=0.675$$
(7)
from strangeness-changing hyperon decays plus flavor SU(3) symmetry. Prior to the EMC experiment, the flavor singlet axial charge was evaluated, by Gourdin and Ellis-Jaffe , assuming $`\mathrm{\Delta }S=0`$, to be
$$\mathrm{\Delta }Q^0=\mathrm{\Sigma }=\mathrm{\Delta }U+\mathrm{\Delta }D+\mathrm{\Delta }S=\mathrm{\Delta }Q^8.$$
(8)
Then one obtains, neglecting small QCD corrections, the GEJ sum rule,
$$\mathrm{\Gamma }^p=\frac{1}{12}\mathrm{\Delta }Q^3+\frac{1}{36}\mathrm{\Delta }Q^8+\frac{1}{9}\mathrm{\Delta }Q^0=0.198,$$
(9)
which is significantly larger than the observed experimental result of 0.126 from the EMC experiment , but now revised to be 0.136 .
The discovery of the GEJ sum rule violation came as a big surprise to the physics community since the sum of the quark helicities $`\mathrm{\Sigma }`$ inferred from Eqs. (6) and (7) and the observed $`\mathrm{\Gamma }^p`$, by allowing $`\mathrm{\Delta }S0`$, gave the value
$$\mathrm{\Sigma }=\mathrm{\Delta }U+\mathrm{\Delta }D+\mathrm{\Delta }S=0.020$$
(10)
from the EMC data $`\mathrm{\Gamma }^p=0.126`$ , and
$$\mathrm{\Sigma }=\mathrm{\Delta }U+\mathrm{\Delta }D+\mathrm{\Delta }S0.30$$
(11)
from the revised results $`\mathrm{\Gamma }^p=0.136`$ and $`\mathrm{\Gamma }^n=0.03`$, assuming SU(3) symmetry . This is in conflict with the naive expectation that the spin of the proton is totally provided by the spins of the three valence quarks in the naive SU(6) quark model, if one interpreted the quark helicity $`\mathrm{\Delta }Q`$ as the quark spin contribution to the proton spin. Many theoretical and experimental investigations have been devoted to understanding this proton “spin puzzle” or “spin crisis” .
However, it has been pointed in Refs. that this puzzle can be easily explained within the naive SU(6) quark model if one properly considers the fact that the observed quark helicity $`\mathrm{\Delta }Q`$ is the quark spin defined in the light-cone formalism (infinite momentum frame), and it is different from the quark spin as defined in the rest frame of the nucleon (or in the quark model). In the light-cone or quark-parton descriptions, $`\mathrm{\Delta }q(x)=q^{}(x)q^{}(x)`$, where $`q^{}(x)`$ and $`q^{}(x)`$ are the probabilities of finding a quark or antiquark with longitudinal momentum fraction $`x`$ and polarization parallel or anti-parallel to the proton helicity in the infinite momentum frame. However, in the nucleon rest frame one finds ,
$$\mathrm{\Delta }q(x)=[\mathrm{d}^2𝐤_{}]M_q(x,𝐤_{})\mathrm{\Delta }q_{QM}(x,𝐤_{}),$$
(12)
with
$$M_q(x,𝐤_{})=\frac{(k^++m)^2𝐤_{}^2}{(k^++m)^2+𝐤_{}^2},$$
(13)
where $`M_q(x,𝐤_{})`$ is the contribution from the relativistic effect due to the quark transverse motion (or the Melosh-Wigner rotation effect), $`q_{s_z=\frac{1}{2}}(x,𝐤_{})`$ and $`q_{s_z=\frac{1}{2}}(x,𝐤_{})`$ are the probabilities of finding a quark and antiquark with rest mass $`m`$ and transverse momentum $`𝐤_{}`$ and with spin parallel and anti-parallel to the rest proton spin, $`\mathrm{\Delta }q_{QM}(x,𝐤_{})=q_{s_z=\frac{1}{2}}(x,𝐤_{})q_{s_z=\frac{1}{2}}(x,𝐤_{})`$, and $`k^+=x`$, where $`^2=_i\frac{m_i^2+𝐤_i^2}{x_i}`$. The Melosh-Wigner rotation factor $`M_q(x,𝐤_{})`$ ranges from 0 to 1; thus $`\mathrm{\Delta }q`$ measured in polarized deep inelastic scattering cannot be identified with $`\mathrm{\Delta }q_{QM}`$, the spin carried by each quark flavor in the proton rest frame or the quark spin in the quark model.
We now check whether it is possible to explain the observed data for $`\mathrm{\Gamma }^p`$ and $`\mathrm{\Gamma }^n`$ within the SU(6) naive quark model by taking into account the Melosh-Wigner rotation effect. Though we do not expect this to be the real situation, it is interesting since there existed a general impression that it is impossible to explain the proton “spin puzzle” within the SU(6) naive quark model. Also an early attempt for such purpose failed, by using the early EMC data $`\mathrm{\Gamma }^p=0.126`$ and $`\mathrm{\Gamma }^n`$ obtained from the Bjorken sum rule $`\mathrm{\Gamma }^p\mathrm{\Gamma }^n=\frac{1}{6}G_A/G_V`$. We start from the conventional SU(6) naive quark model wavefunctions for the proton and the neutron
$$|p^{}=\frac{1}{\sqrt{18}}(2|u^{}u^{}d^{}|u^{}u^{}d^{}|u^{}u^{}d^{})+(cyclicpermutation);$$
(14)
$$|n^{}=\frac{1}{\sqrt{18}}(2|d^{}d^{}u^{}|d^{}d^{}u^{}|d^{}d^{}u^{})+(cyclicpermutation).$$
(15)
One finds that the quark spin contributions $`\mathrm{\Delta }u_{QM}=\frac{4}{3}`$, $`\mathrm{\Delta }d_{QM}=\frac{1}{3}`$, and $`\mathrm{\Delta }s_{QM}=0`$ for the proton, and the exchange of $`ud`$ in the above quark spin contributions gives those for the neutron. Then we get the integrated spin structure functions for the proton and the neutron as
$$\mathrm{\Gamma }^p=\frac{1}{2}(\frac{4}{9}M_u\mathrm{\Delta }u_{QM}+\frac{1}{9}M_d\mathrm{\Delta }d_{QM}+\frac{1}{9}M_s\mathrm{\Delta }s_{QM});$$
(16)
$$\mathrm{\Gamma }^n=\frac{1}{2}(\frac{1}{9}M_u\mathrm{\Delta }u_{QM}+\frac{4}{9}M_d\mathrm{\Delta }d_{QM}+\frac{1}{9}M_s\mathrm{\Delta }s_{QM}),$$
(17)
where $`M_q`$ is the averaged Melosh-Wigner rotation factor for the quark $`q`$. From Eqs. (16) and (17) we obtain
$$M_u\mathrm{\Delta }u_{QM}=\frac{24\mathrm{\Gamma }^p6\mathrm{\Gamma }^nM_s\mathrm{\Delta }s_{QM}}{5};$$
(18)
$$M_d\mathrm{\Delta }d_{QM}=\frac{24\mathrm{\Gamma }^n6\mathrm{\Gamma }^pM_s\mathrm{\Delta }s_{QM}}{5},$$
(19)
from which we get the values
$$M_u\mathrm{\Delta }u_{QM}=0.689;$$
(20)
$$M_d\mathrm{\Delta }d_{QM}=0.307,$$
(21)
with the inputs $`\mathrm{\Gamma }^p=0.136`$, $`\mathrm{\Gamma }^n=0.03`$ , and $`\mathrm{\Delta }s_{QM}=0`$. Thus we get, for $`\mathrm{\Delta }u_{QM}=\frac{4}{3}`$ and $`\mathrm{\Delta }d_{QM}=\frac{1}{3}`$, that
$$M_u=0.517,M_d=0.921,\mathrm{a}ndr_{d/u}=M_d/M_u=1.78,$$
(22)
which means that we need a flavor asymmetry between the $`u`$ and $`d`$ quarks for the Melosh-Wigner rotation factors to reproduce the observed data $`\mathrm{\Gamma }^p`$ and $`\mathrm{\Gamma }^n`$ within the SU(6) naive quark model. The sum of quark helicities in this situation is
$$\mathrm{\Sigma }=M_u\mathrm{\Delta }u_{QM}+M_d\mathrm{\Delta }d_{QM}+M_s\mathrm{\Delta }s_{QM}0.38,$$
(23)
which is small and far from 1, the total quark spin contributions $`\mathrm{\Delta }u_{QM}+\mathrm{\Delta }d_{QM}+\mathrm{\Delta }s_{QM}`$ to the nucleon spin. We need to point out here that there is no mistake in calling the quark helicity $`\mathrm{\Delta }q=M_u\mathrm{\Delta }u_{QM}`$ the quark spin contribution as commonly accepted in the literature, if one properly understands it from a relativistic viewpoint. But in this case there should be also non-zero contribution to the relativistic orbital angular momentum even for the S-wave quarks in the naive SU(6) quark model. Detail illustrations concerning this point can be found in Ref. where the role played by the Melosh-Wigner rotation on the quark orbital angular momentum is studied.
We know that a symmetry between the valence $`u(x)`$ and $`d(x)`$ quark distributions would mean $`F_2^n(x)/F_2^p(x)\frac{2}{3}`$ for the unpolarized structure functions $`F_2(x)`$ in the whole $`x`$ region $`x=01`$, and this has been ruled out by the experimental observation that $`F_2^n(x)/F_2^p(x)<0.5`$ at $`x1`$. This indicates an asymmetry between the $`u(x)`$ and $`d(x)`$ valence quark distributions, and such an asymmetry, which can be reproduced in an SU(6) quark-spectator-diquark model , also implies an asymmetry between the Melosh-Wigner rotation factors for $`M_u`$ and $`M_d`$ . It is interesting to notice that the asymmetry ratio $`r_{d/u}=M_d/M_u`$ larger than 1 is in the right direction as predicted in the quark-spectator-diquark model , though the magnitude is not so big as that given in Eq. (22). This may imply that an additional source for a bigger $`ud`$ flavor asymmetry is needed for a more realistic description of the nucleon.
## III The intrinsic nucleon sea from the baryon-meson fluctuations
Though the proton spin problem raised doubt about the quark model at first, there has been a consistent attempt to understand the problem within the quark model framework on extended quark models , and also on the quark model in the light-cone formalism . For example, Jaffe and Lipkin found that both the EMC data and the $`\beta `$-decay data can be fitted using a “reasonable modification” of the standard quark model in which the only additional degrees of freedom are a single quark-antiquark pair in the lowest states of spin and orbital motion allowed by conservation laws. Keppler et al. pointed out that the $`5q`$ component should be dominated by pseudoscalar S-wave mesons. Qing, Chen, and Wang gave a numerical calculation of the coefficients of the total wavefunction in the non-relativistic quark potential model by including the Melosh-Wigner rotation effect , although in a different manner, and showed that the proton wavefunction is dominated by the bare $`3q`$ state.
In this section, we will perform a more detailed analysis of the spin structure in an extended quark model by taking into account the higher Fock states in the wavefunction of the proton, and check how these higher Fock states may influence the analysis in Section II, where we considered the effect of the Melosh-Wigner rotation with the three valence quark component only. In the higher Fock states, the quark and antiquark of a quark-antiquark pair are rearranged non-perturbatively with the three valence quarks into a meson and a baryon and we write the configuration as a baryon-meson fluctuation. In the “minimal” quark model of Jaffe-Lipkin , the quark-antiquark pairs are actually vector mesons in a baryon-meson fluctuation picture. The higher Fock state in the “minimal” quark model, which is referred to as Jaffe-Lipkin term, can be written as
$$|[JL]^{}=\mathrm{cos}\theta |[\mathrm{b}\epsilon ]^{}+\mathrm{sin}\theta |[\mathrm{bD}]^{},$$
(24)
where $`b`$ denotes the three quark $`qqq`$ component for a bare nucleon. The extra $`qqqQ\overline{Q}`$ component $`|[b\epsilon ]^{}`$ with the $`0^{++}`$ $`Q\overline{Q}`$ denoted by $`\epsilon `$ can be written as,
$$|\epsilon =\sqrt{\frac{1}{3}}|Y^{}X^{}+\sqrt{\frac{1}{3}}|Y^{}X^{}\sqrt{\frac{1}{3}}|Y^0X^0,$$
(25)
and the extra $`qqqQ\overline{Q}`$ component $`|[bD]^{}`$ with the $`1^{++}`$ $`Q\overline{Q}`$ denoted by $`D`$ can be written as,
$$|[bD]^{}=\sqrt{\frac{2}{3}}|b^{}D^{}\sqrt{\frac{1}{3}}|b^{}D^0,$$
(26)
with
$$|D^{}=\sqrt{\frac{1}{2}}|Y^{}X^0\sqrt{\frac{1}{2}}|Y^0X^{};$$
(27)
$$|D^0=\sqrt{\frac{1}{2}}|Y^{}X^{}\sqrt{\frac{1}{2}}|Y^{}X^{},$$
(28)
where $`D^{}`$, $`D^0`$, and $`D^{}`$ denote the $`J_3`$ states of the $`Q\overline{Q}`$ pair; $`Y^{}`$, $`Y^0`$, and $`Y^{}`$ denote the $`L_3`$ states of the $`Q\overline{Q}`$ spin; $`X^{}`$, $`X^0`$, and $`X^{}`$ denote the $`S_3`$ states of the $`Q\overline{Q}`$ spin; and $``$ denotes a $`J_3=1`$ spin contribution and $``$ denotes a $`J_3=1/2`$ spin contribution. With the above higher Fock states included, Jaffe and Lipkin found that the proton state has only a small amplitude to be a bare three-quark baryon state, in order to reproduce the large negative sea spin found in their analysis on the hyperon beta decay, baryon magnetic moments and the EMC result on the fraction of the spin of the nucleon carried by the spins of the quarks .
In the Jaffe-Lipkin term, only P-wave vector $`q\overline{q}`$ pairs have been taken into account. However, if we consider the $`qqqQ\overline{Q}`$ component as a baryon-meson fluctuation of the nucleon, then the dominant fluctuations should be the ones in which the baryon-meson has the smallest off-shell energy . Therefore energy considerations require that the $`qqqQ\overline{Q}`$ component should be dominated by pseudoscalar S-wave mesons, like the pion . In order to describe a nucleon state more realistically, we include these new higher Fock states in addition to the Jaffe-Lipkin states, and the nucleon state should be in principle extended to
$$|B^{}=\mathrm{cos}\alpha \mathrm{cos}\beta |\mathrm{b}^{}+\mathrm{sin}\alpha \mathrm{cos}\beta |[\mathrm{BM}]^{}+\mathrm{sin}\beta |[\mathrm{JL}]^{},$$
(29)
where $`\alpha `$ and $`\beta `$ are the mixing angles between the bare baryon state and the baryon-meson states $`|[BM]^{}`$ and $`|[JL]^{}`$, and the baryon-meson BM state can be written as
$$|[BM]^{}=\sqrt{\frac{2}{3}}|b^{}MY^{}\sqrt{\frac{1}{3}}|b^{}MY^0,$$
(30)
where $`M`$ denotes the spin contribution from the pseudoscalar meson (with spin zero but parity -1), and $`Y`$ denotes orbital angular momentum (with $`L=1`$) due to the relative motion between the baryon and the meson. We can also extend the BM term by including the $`b^{}=qqq`$ state with spin $`S=3/2`$ if higher order baryon-meson fluctuations need to be considered, and in this case we write
$$|[BM]^{}=A(bM)|[bM]^{}+A(b^{}M)|[b^{}M]^{},$$
(31)
where
$$|[bM]^{}=\sqrt{\frac{2}{3}}|b^{}MY^{}\sqrt{\frac{1}{3}}|b^{}MY^0,$$
(32)
as in Eq. (30), and
$$|[b^{}M]^{}=\sqrt{\frac{1}{2}}|b_{}^{}{}_{}{}^{}MY^{}\sqrt{\frac{1}{3}}|b_{}^{}{}_{}{}^{}MY^0+\sqrt{\frac{1}{6}}|b_{}^{}{}_{}{}^{}MY^{}.$$
(33)
The anti-quarks are unpolarized since they exit only in the pseudoscalar meson of the BM state.
Using the wavefunction (29), we now calculate the contributions $`\mathrm{\Sigma }_v`$, $`\mathrm{\Sigma }_s`$, and $`\mathrm{\Lambda }_s`$, of the valence quark spins, the spin of the sea, and the orbital angular momentum of the sea, to the spin of the proton, and we obtain
$$\mathrm{\Sigma }_v=\mathrm{cos}^2\alpha \mathrm{cos}^2\beta \frac{1}{3}\mathrm{sin}^2\alpha \mathrm{cos}^2\beta +\mathrm{sin}^2\beta \mathrm{cos}^2\theta \frac{1}{3}\mathrm{sin}^2\beta \mathrm{sin}^2\theta ;$$
(34)
$$\mathrm{\Sigma }_s=\frac{8}{3}\sqrt{\frac{1}{2}}\mathrm{sin}^2\beta \mathrm{sin}\theta \mathrm{cos}\theta +\frac{2}{3}\mathrm{sin}^2\beta \mathrm{sin}^2\theta ;$$
(35)
$$\mathrm{\Lambda }_s=\frac{8}{3}\sqrt{\frac{1}{2}}\mathrm{sin}^2\beta \mathrm{sin}\theta \mathrm{cos}\theta +\frac{2}{3}\mathrm{sin}^2\beta \mathrm{sin}^2\theta +\frac{4}{3}\mathrm{sin}^2\alpha \mathrm{cos}^2\beta ,$$
(36)
with
$$\mathrm{\Sigma }_v+\mathrm{\Sigma }_s+\mathrm{\Lambda }_s=1.$$
(37)
We can say alternatively that $`\mathrm{\Sigma }_v`$ comes from the spin sums of all $`b=qqq`$ terms, $`\mathrm{\Sigma }_s`$ from the spin sums of all $`Q\overline{Q}`$ terms ($`X`$ terms in the Jaffe-Lipkin term and $`M`$ terms in the BM term Eq. (30)), and $`\mathrm{\Lambda }_s`$ from the orbital angular momentum of all $`Y`$ terms in the nucleon state $`|B^{}`$.
It can be easily seen that the sea spin $`\mathrm{\Sigma }_s`$ comes entirely from the Jaffe-Lipkin term, since the spin contribution from the $`M`$ terms is zero. It is also interesting that $`\mathrm{\Sigma }_s`$ cannot be negative if there is no interference between the two components $`|b\epsilon `$ and $`|bD`$ in the Jaffe-Lipkin term Eq. (24). If we follow Ref. and adopt the two models for the sea spin $`\mathrm{\Sigma }_s`$, then we find that we must arrive at the conclusion of Jaffe-Lipkin term dominance. In the first model (called II in Ref. ), the sea is taken as $`\mathrm{SU}(3)_{\mathrm{flavor}}`$ symmetric, and $`\mathrm{\Sigma }_s(\mathrm{II})=0.69\pm 0.27`$. In the second model (called III in Ref. ), the sea is taken as $`\mathrm{SU}(2)_{\mathrm{flavor}}`$ symmetric, and $`\mathrm{\Sigma }_s(\mathrm{III})=0.56\pm 0.22`$. On the other hand, the data on hyperon and nucleon $`\beta `$-decays requires $`\mathrm{\Sigma }_v`$ to be approximately $`\frac{3}{4}`$. Of course, it is impossible for us to completely determinate $`\alpha `$, $`\beta `$ and $`\theta `$ using the values of $`\mathrm{\Sigma }_v`$ and $`\mathrm{\Sigma }_s`$ mentioned above. But, taking (34) and (35) as constraint conditions, we can give a range of values of these mixing angles. Selected values of mixing angles are shown in Tab. 1. The results in Tab. 1 show that a physically reasonable $`\mathrm{cos}\beta `$ can only have a very small value with the above $`\mathrm{\Sigma }_v`$ and $`\mathrm{\Sigma }_s`$, and this requires the Jaffe-Lipkin term dominance. The sea in the baryon-meson state (30) only provides the orbital angular momentum to the nucleon, and the Jaffe-Lipkin term (24) provides the negative polarized sea spin. Thus the importance of the Jaffe-Lipkin term depends on the constraints of the sea quark polarization of the nucleon.
Table 1 The mixing angles
| $`\mathrm{\Sigma }_s(\mathrm{II})=0.69`$ | $`\mathrm{\Sigma }_s(\mathrm{III})=0.56`$ |
| --- | --- |
| $`\mathrm{sin}\alpha `$ $`\mathrm{sin}\beta `$ $`\mathrm{sin}\theta `$ | $`\mathrm{sin}\alpha `$ $`\mathrm{sin}\beta `$ $`\mathrm{sin}\theta `$ |
| $`\pm 0.200`$ $`\pm 1.080`$ $`0.408`$ | $`\pm 0.200`$ $`\pm 0.947`$ $`0.452`$ |
| $`\pm 0.400`$ $`\pm 1.065`$ $`0.429`$ | $`\pm 0.400`$ $`\pm 0.956`$ $`0.436`$ |
| $`\pm 0.600`$ $`\pm 1.052`$ $`0.452`$ | $`\pm 0.600`$ $`\pm 0.966`$ $`0.419`$ |
| $`\pm 0.800`$ $`\pm 1.041`$ $`0.472`$ | $`\pm 0.800`$ $`\pm 0.975`$ $`0.405`$ |
From a strict sense, the sea spin $`\mathrm{\Sigma }_s`$ has not been measured directly, and also the Melosh-Wigner rotation factors should be introduced into the so called spin term $`\mathrm{\Sigma }_v`$ obtained from hyperon and nucleon $`\beta `$-decays, and the flavor asymmetry and SU(3) symmetry breaking should be important. Therefore the above analysis needs to be updated. It would be more practical to decompose the spin by the contributions from the quarks $`\mathrm{\Sigma }_q=\mathrm{\Sigma }_v+\frac{1}{2}\mathrm{\Sigma }_s`$, the antiquarks $`\mathrm{\Sigma }_{\overline{q}}=\frac{1}{2}\mathrm{\Sigma }_s`$, and the orbital angular momentum $`\mathrm{\Lambda }_s`$, which still meet the condition
$$\mathrm{\Sigma }_q+\mathrm{\Sigma }_{\overline{q}}+\mathrm{\Lambda }_s=1.$$
(38)
The antiquark helicity distributions extracted from semi-inclusive deep inelastic scattering experiments are consistent with zero , in agreement with the small antiquark polarization predicted in both the baryon-meson fluctuation model and a chiral quark model . There is still no direct evidence for a large negative antiquark polarization in experiments. We also point out here that there should be a quark-antiquark asymmetry for the spin of the sea when flavor decomposition is necessary .
Since new measurements on the polarized structure functions for both the proton and the neutron have become available, we will use the measured $`\mathrm{\Gamma }^p`$ and $`\mathrm{\Gamma }^n`$ as inputs to study the effects due to the inclusion of Jaffe-Lipkin and BM higher Fock state terms in the nucleon wavefunction. Another aspect that we need to take into account is that the $`u`$ and $`d`$ flavor asymmetries should exist in both the valence and sea contents of the nucleon. The observation of the Gottfried sum rule violation in several processes implies that there is an important contribution coming from the lowest baryon-meson fluctuation $`p(uudD\overline{D})=n(udD)\pi ^+(u\overline{D})`$ of the proton . This puts a constraint on the value of $`\alpha `$ for the BM mixing term. If one assumes an isospin symmetry between the proton and neutron , then the Gottfried sum rule violation implies an asymmetry between the $`u`$ and $`d`$ sea distributions inside the proton
$$_0^1dx\left[\overline{d}(x)\overline{u}(x)\right]=0.148\pm 0.039.$$
(39)
If we consider only the $`p(uudD\overline{D})=n(udD)\pi ^+(u\overline{D})`$ component inside the BM term and neglect flavor asymmetry in the Jaffe-Lipkin term, then we get the constraint,
$$\mathrm{sin}^2\alpha \mathrm{cos}^2\beta =0.148.$$
(40)
The $`u`$ and $`d`$ quark spins in the proton wavefunction should be
$`\mathrm{\Delta }u_{QM}=\mathrm{cos}^2\alpha \mathrm{cos}^2\beta \mathrm{\Delta }u_0{\displaystyle \frac{1}{3}}\mathrm{sin}^2\alpha \mathrm{cos}^2\beta \mathrm{\Delta }d_0+\mathrm{sin}^2\beta \mathrm{\Delta }u_{JL};`$ (41)
$`\mathrm{\Delta }d_{QM}=\mathrm{cos}^2\alpha \mathrm{cos}^2\beta \mathrm{\Delta }d_0{\displaystyle \frac{1}{3}}\mathrm{sin}^2\alpha \mathrm{cos}^2\beta \mathrm{\Delta }u_0+\mathrm{sin}^2\beta \mathrm{\Delta }d_{JL},`$ (42)
where $`\mathrm{\Delta }u_0=4/3`$ and $`\mathrm{\Delta }d_0=1/3`$ are the $`u`$ and $`d`$ quark spins for the bare $`qqq`$ proton, and $`\mathrm{\Delta }u_{JL}`$ and $`\mathrm{\Delta }d_{JL}`$ are the $`u`$ and $`d`$ quark spins for the Jaffe-Lipkin term Eq. (24) from $`b`$, $`\epsilon `$, and $`D`$,
$$\mathrm{\Delta }q_{JL}=(1\frac{4}{3}\mathrm{sin}^2\theta )\mathrm{\Delta }q_0+\frac{1}{4}\mathrm{\Sigma }_s$$
(43)
for $`q=u,d`$ in case of only charge neutral $`Q\overline{Q}`$’s with $`u`$ and $`d`$ flavors. Substituting the above $`\mathrm{\Delta }u_{QM}`$ and $`\mathrm{\Delta }d_{QM}`$ into Eqs. (20) and (21), we get
$$M_u=0.598,M_d=0.878,\mathrm{a}ndr_{d/u}=M_d/M_u=1.47,$$
(44)
for $`\beta =0`$ without the Jaffe-Lipkin term. We find that the $`u`$ and $`d`$ flavor asymmetry $`r_{d/u}`$ is reduced compared to Eq. (22) and this shows that the $`p(uudD\overline{D})=n(udD)\pi ^+(u\overline{D})`$ fluctuation produces a more reasonable $`d/u`$ Melosh-Wigner rotation asymmetry than in the naive picture with the bare nucleon state of only three valence quarks .
In fact, we should also include other baryon-meson fluctuations in a more realistic picture of intrinsic sea quarks , such as $`p(uudU\overline{U})=\mathrm{\Delta }^{++}(uuU)\pi ^{}(d\overline{U})`$ for the intrinsic $`U\overline{U}`$ quark-antiquark pairs and $`p(uudS\overline{S})=\mathrm{\Lambda }(udS)K^+(u\overline{S})`$ for the intrinsic strange quark-antiquark pairs. In this case we can write the baryon-meson term as
$`\mathrm{sin}\alpha \mathrm{cos}\beta |[BM]^{}=A(n\pi ^+)|n\pi ^++A(\mathrm{\Lambda }K^+)|\mathrm{\Lambda }K^++A(\mathrm{\Delta }^{++}\pi ^{})|\mathrm{\Delta }^{++}\pi ^{},`$ (45)
where we take the baryon-meson configuration probabilities $`P(p=BM)=[A(BM)]^2`$ as
$$P(p=n\pi ^+)15\%;P(p=\mathrm{\Lambda }K^+)3\%;P(p=\mathrm{\Delta }^{++}\pi ^{})1\%,$$
(46)
as estimated from a reasonable physical picture . With the above baryon-meson fluctuations considered, we find,
$$M_u=0.624,M_d=0.912,\mathrm{a}ndr_{d/u}=M_d/M_u=1.46,$$
(47)
which are close to Eq. (44), the case with only $`p=n\pi ^+`$ fluctuation. Thus our above analysis supports a reasonable picture of a dominant valence three quark component with a certain amount of the energetically-favored baryon-meson fluctuations , as a “minimal” quark model for the spin relevant observations in DIS processes and also for several phenomenological anomalies related to the flavor content of nucleons . Of course, we can also include the necessary other higher $`5q`$ Fock states approximated in terms of the BM state and the Jaffe-Lipkin state.
The gluon distribution of a hadron is usually assumed to be generated from the QCD evolution. However, it has been pointed in Ref. that there exist intrinsic gluons in the bound-state wavefunction. Therefore we could also consider the possibility of including a $`(qqqg)`$ Fock state in our description. Unfortunately the gluon is always a relativistic particle, and it is not easy to incorporate it in the present framework. We must use a relativistic approach from the start, such as the one given in Ref. .
## IV Summary and discussion
We investigated the spin structure of the nucleon in a simple quark model. First, we studied the possible effect due to the Melosh-Wigner rotation. We found that an introduction of an up-down quark flavor asymmetry in the Melosh-Wigner rotation factors can reproduce the present experimental data of the integrated spin structure functions for both the proton and the neutron within a simple SU(6) quark model with only three valence quarks. And then, we discussed the contributions to the nucleon spin from various components of the nucleon wavefunction. The calculated results indicate that the baryon-meson state of Jaffe-Lipkin with vector meson is important concerning the sea polarization of the nucleon regardless of the existence of states which include the pseudoscalar mesons.
The Melosh-Wigner rotation is one of the most important ingredients of the light-cone formalism. Its effect is of fundamental importance in the spin content of hadrons, and it is mainly due to the transverse momentum of quarks in the nucleon. Actually, it reflects some relativistic effects of a quark system. On the other hand, the simple quark model discussed here includes the baryon-meson fluctuation component in the nucleon wavefunction, which is a nonperturbative effect. The present investigation shows that relativistic and nonperturbative effects are very important in order to understand the spin structure of the nucleon. In the simple quark model, the bare three quark component and the baryon-meson state with a pesudoscalar meson, are still important concerning the proton spin problem in polarized structure functions, after we take into account the Melosh-Wigner rotation effect.
ACKNOWLEDGMENTS: This work is partially supported by National Natural Science Foundation of China under Grant Numbers 19975052 and 19875024, No. 19775051, and by Fondecyt (Chile) postdoctoral fellowship 3990048, by Fondecyt (Chile) grant 1990806 and by a Cátedra Presidencial (Chile).
|
warning/0006/hep-ph0006080.html
|
ar5iv
|
text
|
# DIS STRUCTURE FUNCTIONS FROM THE SATURATION MODELaafootnote aPresented at the 8th International Workshop on Deep Inelastic Scattering, Liverpool, 25-30 April 2000.
## Acknowledgments
This work was done in collaboration with Mark Wüsthoff. Supported by Deutsche Forschungsgemeinschaft and Polish KBN grant no. 2 P03B 089 13.
## References
|
warning/0006/physics0006028.html
|
ar5iv
|
text
|
# A collision-induced satellite in the Lyman 𝛽 profile due to H-H collisions
## 1 Introduction
In Allard et al. 1999 allard99 we derived a classical path expression for a pressure-broadened atomic spectral line shape that allows for a radiative electric dipole transition moment which depends on the position of the perturbers. This factor is not included in the more usual approximation of Anderson & Talman 1956 anderson and Baranger 1958 baranger58a ; baranger58b . We used this theory to study the influence of the variation of the dipole moment on the satellites present in the far wing profiles of the Lyman series lines of atomic hydrogen seen in stars and in laboratory plasmas.
Satellite features at 1600 Å and 1405 Å in the Lyman $`\alpha `$ wing associated with free-free quasi-molecular transitions of H<sub>2</sub> and H$`{}_{}{}^{+}{}_{2}{}^{}`$ have been observed in ultraviolet (UV) spectra of certain stars obtained with the International Ultraviolet Explorer (IUE) and the Hubble Space Telescope (HST) koester93 ; koester94 ; holweger94 ; bergeron95 . The stars which show Lyman $`\alpha `$ satellites are DA white dwarfs, old Horizontal Branch stars of spectral type A, and the $`\lambda `$ Bootis stars. The last two have the distinctive property of poor metal content, that is, low abundances of elements other than H and He. Satellites also have been observed in the laboratory spectra of laser-produced hydrogen plasmas kielkopf95 ; kielkopf98 .
Satellite features in hydrogen lines are not limited to Lyman $`\alpha `$, which is the only Lyman-series line accessible to IUE as well as HST. Observations with HUT (Hopkins Ultraviolet Telescope) and with ORFEUS (Orbiting Retrievable Far and Extreme Ultraviolet Spectrograph) of Lyman $`\beta `$ of DA white dwarfs with T<sub>eff</sub> close to 20000 K have revealed a line shape very different from the expected simple Stark broadening, with line satellites near 1078 and 1060 Å koester96 ; koester98 . The satellites in the red wing of Lyman $`\beta `$ are in the 905 to 1187 Å spectral region covered by the Far Ultraviolet Spectroscopic Explorer (FUSE) launched in June 1999. Furthermore, Lyman $`\beta `$ profiles are also the subject of an ongoing study of the far ultraviolet spectrum of dense hydrogen plasmas. The strengths of these satellite features and indeed the entire shape of wings in the Lyman series are very sensitive to the degree of ionization in the stellar atmosphere and laboratory plasmas, because that determines the relative importance of broadening by ion and neutral collisions.
In a previous work allard98a we presented theoretical profiles of Lyman $`\beta `$ perturbed solely by protons. The calculations were based on the accurate theoretical H$`{}_{}{}^{+}{}_{2}{}^{}`$ molecular potentials of Madsen and Peek madsen71 to describe the interaction between radiator and perturber, and dipole transition moments of Ramaker and Peek ramaker72 . The line profiles were included as a source of opacity in model atmospheres for hot white dwarfs, and the predicted spectra compared very well with the observed ORFEUS spectra koester98 .
The very recent ab initio calculations of Drira 1999 drira99 of electronic transition moments for excited states of the H<sub>2</sub> molecule and very accurate molecular potentials of Schmelcher schmelcher ; detmer98 now allow us to compute Lyman $`\beta `$ profiles simultaneously perturbed by neutral atomic hydrogen and by protons. The aim of our paper is to point out a collision-induced satellite correlated to the $`\mathrm{B}^{\prime \prime }\overline{B}^1\mathrm{\Sigma }_u^+\mathrm{X}^1\mathrm{\Sigma }_g^+`$ asymptotically forbidden transition of H<sub>2</sub>. We show that the shape of the wing in the region between Lyman $`\beta `$ and Lyman $`\alpha `$ is particularly sensitive to the relative abundance of the neutral and ion perturbers responsible for the broadening of the lines.
## 2 Theory
The classical path theory for the shape of pressure-broadened atomic spectral lines which takes into account the variation of the electric dipole moment during a collision is only briefly outlined here. The theory has been described in detail in allard99 . Our approach is based on the quantum theory of spectral line shapes of Baranger baranger58a ; baranger58b developed in an adiabatic representation to include the degeneracy of atomic levels royer74 ; royer80 ; allard94 .
### 2.1 General expression for the spectrum in an adiabatic representation
The spectrum $`I(\omega )`$ can be written as the Fourier transform of the dipole autocorrelation function $`\mathrm{\Phi }`$(s),
$$I(\omega )=\frac{1}{\pi }Re_0^+\mathrm{}\mathrm{\Phi }(s)e^{i\omega s}𝑑s.$$
(1)
Here,
$`\mathrm{\Phi }(s)`$ $`=`$ $`\mathrm{𝐓𝐫}\rho 𝐃^{}e^{\frac{is𝐇}{\mathrm{}}}𝐃e^{\frac{is𝐇}{\mathrm{}}}`$ (2)
$`=`$ $`𝐃^{}(0)𝐃(s)`$ (3)
is the autocorrelation function of $`𝐃(s)`$ , the dipole moment of the radiator in the Heisenberg representation (we use bold notation for operators) allard82 . $`𝐇`$ is the total Hamiltonian
$$𝐇=𝐓_{nucl}+𝐓_{elec}+V(𝐱,𝐑),$$
(4)
where $`𝐓_{nucl}`$ and $`𝐓_{elec}`$ are sums of nuclear and electronic kinetic energy operators respectively, and $`V(𝐱,𝐑)`$ is the interaction between particles. Here $`𝐱`$ denotes collectively the set of electronic coordinates (position and spin) plus spin coordinates of the nuclei, while $`𝐑`$ denotes the set of position coordinates of the nuclei. We assume that the radiating atom is immersed in a perturber bath in thermal equilibrium. The density matrix $`\rho `$ is
$$\rho \frac{e^{\beta 𝐇}}{\mathrm{𝐓𝐫}e^{\beta 𝐇}},$$
(5)
where $`\beta `$ is the inverse temperature ($`1/kT`$).
We use the notation
$$()\mathrm{𝐓𝐫}\rho (),$$
(6)
where $`\mathrm{𝐓𝐫}`$ denotes the trace operation.
The adiabatic or Born-Oppenheimer representation comprises expanding states of the gas in terms of electronic states $`\chi _e(x;R)`$ corresponding to frozen nuclear configurations. In the Schrödinger equation
$`(𝐓_{elec}+V(𝐱,𝐑))\chi _e(x;R)`$ $`=`$ $`𝐇_{elec}(R)\chi _e(x;R)`$ (7)
$`=`$ $`E_e(R)\chi _e(x;R).`$ (8)
$`R`$ appears as a parameter, and the eigenenergies $`E_e(R)`$ play the role of potential energies for the nuclei. Any total wave function $`\mathrm{\Psi }(x,R)`$ can be expanded as
$$\mathrm{\Psi }(x,R)=\underset{e}{}\psi _e(R)\chi _e(x;R).$$
(9)
As the nuclei get far from each other, which we denote by $`R\mathrm{}`$ , the electronic energies $`E_e(R)`$ tend to asymptotic values $`E_e^{\mathrm{}}`$ which are sums of individual atomic energies. Since atomic states are usually degenerate, there are in general several different energy surfaces which tend to a same asymptotic energy as $`R\mathrm{}`$ . We will consider specifically a single radiating atom, the radiator, immersed in a gas of optically inactive atoms, the perturbers. For a transition $`\alpha =(i,f)`$ from initial state $`i`$ to final state $`f`$, we have $`R`$-dependent frequencies
$`\omega _{e^{}e}(R)(E_e^{}(R)E_e(R))/\mathrm{},`$ $`e\epsilon _i,e^{}\epsilon _f`$ (10)
which tend to the isolated radiator frequency
$$\omega _\alpha \omega _{fi}(E_f^{\mathrm{}}E_i^{\mathrm{}})/\mathrm{}$$
(11)
as the perturbers get sufficiently far from the radiator:
$`\omega _{e^{}e}(R)\omega _{fi}`$ $`\mathrm{as}`$ $`R\mathrm{},e\epsilon _i,e^{}\epsilon _f.`$ (12)
Let us introduce projectors $`𝐏_e`$ which select the $`e^{th}`$ adiabatic component of any $`\mathrm{\Psi }(x,R)`$ according to royer80
$$𝐏_e\mathrm{\Psi }(x,R)=\psi _e(R)\chi _e(x;R).$$
(13)
We write the dipole moment as a sum over transitions
$$𝐃=\underset{\alpha }{}𝐃_\alpha ,$$
(14)
$$𝐃_\alpha \underset{e,e^{}}{\overset{(\alpha )}{}}𝐏_e^{}\mathrm{𝐃𝐏}_e.$$
(15)
In the Heisenberg representation
$`𝐃_\alpha (t)`$ $``$ $`{\displaystyle \underset{e,e^{}}{\overset{(\alpha )}{}}}e^{\frac{it𝐇}{\mathrm{}}}𝐏_e^{}\mathrm{𝐃𝐏}_ee^{\frac{it𝐇}{\mathrm{}}},`$ (16)
$``$ $`{\displaystyle \underset{e,e^{}}{\overset{(\alpha )}{}}}𝐃_{e^{}e}(t).`$ (17)
The sum $`_{e,e^{}}^{(\alpha )}`$ is over all pairs ($`e,e^{}`$) such that $`\omega _{e^{},e}(R)\omega _\alpha `$ as $`R\mathrm{}`$. Thus $`𝐃_\alpha `$ connects all pairs of adiabatic states whose electronic energy differences become equal to $`\omega _\alpha `$ as $`R\mathrm{}`$. In the absence of perturbers, $`𝐃_\alpha `$ would be the component of $`𝐃`$ responsible for the radiative transitions of frequency $`\omega _\alpha `$. We note that the projection operators account for the weighting factors discussed in Ref. allard94 .
Introducing the expansion Eq. (14) for $`𝐃`$ into the expression Eq. (3) for $`\mathrm{\Phi }(s)`$, we obtain
$$\mathrm{\Phi }(s)=\underset{\alpha ,\beta }{}\mathrm{\Phi }_{\alpha ,\beta }(s)$$
(18)
where
$`\mathrm{\Phi }_{\alpha ,\beta }(s)`$ $`=`$ $`\mathrm{𝐓𝐫}\rho 𝐃_\alpha ^{}e^{\frac{is𝐇}{\mathrm{}}}𝐃_\beta e^{\frac{is𝐇}{\mathrm{}}}`$ (19)
$`=`$ $`𝐃_\alpha ^{}(0)𝐃_\beta (s).`$ (20)
The line shape is then
$$I(\omega )=\underset{\alpha ,\beta }{}I_{\alpha ,\beta }(\omega ).$$
(21)
The terms $`I_{\alpha ,\beta }(\omega )`$ , $`\alpha \beta `$, represent interference between different spectral lines baranger58b . When these interference terms are neglected, we get
$$I(\omega )=\underset{\alpha }{}I_\alpha (\omega )$$
(22)
and
$$\mathrm{\Phi }(s)=\underset{\alpha }{}\mathrm{\Phi }_\alpha (s)$$
(23)
where
$$\mathrm{\Phi }_\alpha (s)=𝐃_\alpha ^{}(0)𝐃_\alpha (s).$$
(24)
The time dependence of $`\mathrm{\Phi }_\alpha (s)`$ is determined by $`𝐃_\alpha (s)`$, the part of the dipole moment which, in the absence of perturbers, oscillates at the frequency $`\omega _\alpha `$. Let us now denote
$$𝐝_\alpha (s)𝐃_\alpha (s)e^{i\omega _\alpha s}$$
(25)
wherein the free evolution $`e^{i\omega _\alpha s}`$ is factored out.
For an isolated line, such as Lyman $`\alpha `$, we have shown (Allard et al. allard99 ) that the normalized line shape $`J_\alpha (\mathrm{\Delta }\omega )`$, in the uncorrelated perturbers approximation, is given by
$$J_\alpha (\mathrm{\Delta }\omega )=\mathrm{𝐅𝐓}[e^{ng_\alpha (s)}]$$
(26)
In the classical path approximation, where we assume that the perturber follows a rectilinear trajectory at a single mean velocity $`\overline{v}`$, we get from allard94 ; allard99 that $`g_\alpha (s)`$ can be written as
$`g_\alpha (s)`$ $`={\displaystyle \frac{1}{_{e,e^{}}^{(\alpha )}|d_{ee^{}}|^2}}{\displaystyle \underset{e,e^{}}{\overset{(\alpha )}{}}}`$ (27)
$`{\displaystyle _0^+\mathrm{}}2\pi \rho 𝑑\rho {\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\stackrel{~}{d}_{ee^{}}[r(0)]`$
$`[e^{\frac{i}{\mathrm{}}_0^s𝑑tV_{e^{}e}[r(t)]}\stackrel{~}{d^{}}_{ee^{}}[r(s)]\stackrel{~}{d}_{ee^{}}[r(0)]].`$
The separation of the radiator and perturber is
$`r(t)=[\rho ^2+(x+\overline{v}t)^2]^{1/2}`$ with $`\rho `$ the impact parameter of the perturber trajectory and $`x`$ is the position of the perturber along its trajectory at time $`t=0`$. The total line strength of the transition is $`_{e,e^{}}^{(\alpha )}|d_{ee^{}}|^2`$. The potential energy for a state $`e`$ is
$$V_e[r(t)]=E_e[r(t)]E_e^{\mathrm{}};$$
(28)
the difference potential is
$$V_{e^{}e}[r(t)]=V_e^{}[r(t)]V_e[r(t)];$$
(29)
and we defined a modulated dipole allard99
$$\stackrel{~}{d}_{ee^{}}[r(t)]=d_{ee^{}}[r(t)]e^{\frac{\beta }{2}V_e[r(t)]},$$
(30)
where we denoted
$$d_{ee^{}}(\stackrel{}{𝐫})=\chi _e(\stackrel{}{𝐫})|𝐝|\chi _e^{}(\stackrel{}{𝐫}).$$
(31)
In the above, we neglected the influence of the potentials $`V_e(r)`$ and $`V_e^{}(r)`$ on the perturber trajectories, which remain straight lines. Although we should drop the Boltzmann factor $`e^{\beta V_e(r)}`$ for consistency with our straight trajectory approximation, by keeping it we improve the result in the wings. Note that over regions where $`V_e(r)<0`$, the factor $`e^{\beta V_e(r)}`$ accounts for bound states of the radiator-perturber pair, but in a classical approximation wherein the discrete bound states are replaced by a continuum; thus any band structure is smeared out.
## 3 Theoretical analysis
### 3.1 Formation of line satellites
Close collisions between a radiating atom and a perturber are responsible for transient quasi-molecules which may lead to the appearance of satellite features in the wing of an atomic line profile.
When the difference $`\mathrm{\Delta }V(R)`$ between the upper and lower interatomic potentials for a given transition goes through an extremum, a relatively wider range of interatomic distances contribute to the same spectral frequency, resulting in an enhancement, or satellite, in the line wing. The unified theory anderson ; allard82 predicts that there will be satellites centered periodically at frequencies corresponding to the extrema of the difference potential between the upper and lower states, $`\mathrm{\Delta }\omega =k\mathrm{\Delta }V_{\mathrm{ext}},(k=1,2,\mathrm{})`$ allard78 ; royer78 ; kielkopf79 . Here $`\mathrm{\Delta }\omega `$ is the frequency difference between the center of the unperturbed spectral line and the satellite feature, measured for convenience in the same units as the potential energy difference. This series of satellites is due to many-body interactions.
### 3.2 Diatomic potentials
The adiabatic interaction of the neutral hydrogen atom with a proton or another hydrogen atom is described by potential energies $`V_e(R)`$ for each electronic state of the H$`{}_{}{}^{+}{}_{2}{}^{}`$ or H<sub>2</sub> molecule ($`R`$ denotes the internuclear distance between the radiator and the perturber). For H-H<sup>+</sup> collisions we have used the potentials of H$`{}_{}{}^{+}{}_{2}{}^{}`$ calculated by Madsen and Peek madsen71 . For H-H collisions we have used the potentials of H<sub>2</sub> calculated by Schmelcher schmelcher .
In Fig. 1 we have plotted the H<sub>2</sub> potential differences $`\mathrm{\Delta }V(R)`$ for the singlet states which contribute to Lyman $`\beta `$. We have used letters here to label the states. B3 is the well known $`\mathrm{B}^{\prime \prime }\overline{\mathrm{B}}^1\mathrm{\Sigma }_u^+`$ state. At small internuclear separation, $`R`$, the state correlates with the $`\mathrm{B}^{\prime \prime }{}_{}{}^{1}\mathrm{\Sigma }_{u}^{+}`$ state, the third $`{}_{}{}^{1}\mathrm{\Sigma }_{u}^{+}`$ state of the Rydberg series. At larger internuclear separation the state has an ionic character until $`R`$= 19 Å where the potential energy curve of the $`\mathrm{H}^++\mathrm{H}^{}(1s)^2`$ state crosses the $`\mathrm{H}(n=1)+\mathrm{H}(n=3)`$ energy levels, because of an avoid crossing $`\overline{B}`$ looses its ionic character. Dabrowski and Herzberg dabrowski76 predicted the existence of the $`\overline{B}`$, the first calculations were done by Kolos kolos76 ; kolos81 . More recent calculations are reported in detmer98 ; reinhold99 .
We use B4 and B5 respectively to label the 4 $`{}_{}{}^{1}\mathrm{\Sigma }_{u}^{+}`$ and 5 $`{}_{}{}^{1}\mathrm{\Sigma }_{u}^{+}`$ states, and V and D to label the 3 $`{}_{}{}^{1}\mathrm{\Pi }_{u}^{}`$ and 2 $`{}_{}{}^{1}\mathrm{\Pi }_{u}^{}`$ states.
Each difference potential exhibits at least one extremum which, in principle, leads to a corresponding satellite feature in the wing of Lyman $`\beta `$ (see Sect. 3.1). The present approach now allows us also to take into account the asymptotically forbidden transitions of quasi-molecular hydrogen which dissociate into (1s,3s) and (1s,3d) atoms. The satellite amplitude depends on the value of the dipole moment through the region of the potential extremum responsible of the satellite and on the position of this extremum. We have shown allard99 ; allard98a ; allard98b that a large enhancement in the amplitude of a spectral line satellite occurs whenever the dipole moment increases through the region of internuclear distance where the satellite is formed.
The potential differences of the B3-X and B4-X transitions exhibit double wells. The maximum at 3.0 Å of the B3-X potential and minimum of the B4-X potential are due to an avoided-crossing.
The most significant characteristic of Fig. 1 is the existence of the deep outer well at 6 Å of the B3-X potential. The ionic interaction decays slowly making the potential energy difference very broad compared to the very steep wells of the other transitions. This is very important as the position of the extremum and the functional dependence of the potential difference on internuclear separation determine the amplitude and shape of the satellites allard94 .
### 3.3 Electronic transition dipole moments
The dipole moment taken between the initial and final states of a radiative transition determines the transition probability, but for two atoms in collision, the moment depends on their separation. This modifies relative contributions to the profile along the collision trajectory. Dipole moments for H$`{}_{2}{}^{}{}_{}{}^{+}`$ and H<sub>2</sub>, calculated as a function of internuclear distance respectively by Ramaker and Peek ramaker72 and by Drira drira99 , were used for the transitions contributing to Lyman $`\alpha `$ and Lyman $`\beta `$. For Lyman $`\beta `$, the four components which correspond to $`1s3p`$ atomic transitions are dipole allowed drira99 : the two singlet B4-X and D-X transitions, and the two triplet 4 $`{}_{}{}^{3}\mathrm{\Sigma }_{g}^{+}`$-$`{}_{}{}^{3}\mathrm{\Sigma }_{u}^{+}`$ and 2 $`{}_{}{}^{3}\mathrm{\Pi }_{g}^{}`$-$`{}_{}{}^{3}\mathrm{\Sigma }_{u}^{+}`$ transitions .
If electronic states $`i`$ and $`f`$ of an isolated radiator are not connected by the dipole moment operator, that is if $`D_{if}(R\mathrm{})=0`$, allowed radiative transitions cannot occur between these two states. This happens for the other transitions which correspond to $`1s3s`$ and $`1s3d`$ atomic transitions. Although these transitions should not contribute to the unperturbed line profile, $`D_{if}(R)`$ may differ from zero when a perturber passes close to the radiator. In this instance radiative transitions are induced by collisions, but not at the unperturbed line frequency.
It often arises that an extremum in the potential difference occurs when the final (or initial) potential energy curve exhibits an avoided crossing, the corresponding wavefunctions exchange their characteristics and the radiative dipole transition moment varies dramatically with $`R`$. This is exactly what happens for the B3-X transition. To point out the importance of variation of dipole moment on the formation of a collision-induced (CI) satellite, we have displayed in Fig. 2 $`D(R)`$ together with the corresponding $`\mathrm{\Delta }V(R)`$ for the $`B3X`$ asymptotically forbidden transition. The dipole transition is extremely small for the isolated radiating atom ( $`R\mathrm{}`$ ) but it goes through a maximum at the value of $`R`$ where the avoid-crossing occurs and remains quite important at the internuclear distance where the potential difference of the outer well goes through a minimum.
In such a case we expect a contribution from this transition in the wing and the formation of a CI line satellite, if it is not smeared out by larger dipole-allowed contributions.
## 4 Lyman $`\beta `$ profile
The Lyman profiles and satellites shown here are calculated at the low densities met in the atmospheres of stars. The typical particle densities from 10<sup>15</sup> to 10<sup>17</sup> cm<sup>-3</sup> allows us to use an expansion of the autocorrelation function in powers of density as described in allard94 ; royer71 . Line profiles are normalized so that over $`\omega `$ they integrate to 1.
### 4.1 Collisional profile perturbed by neutral H
We will consider the two following mechanisms which contribute to the Lyman $`\beta `$ wing.
* The far wing of allowed dipole lines, due to the free-free transition in a pair of colliding atoms.
* The collision-induced absorption due to the free-free transition involving the transient dipole moment existing during a binary collision.
The line profile calculation shown Fig. 3 has been done at a temperature of 10000 K for a perturber density of $`10^{16}`$ cm<sup>-3</sup> of neutral hydrogen. The only line feature is a broad CI satellite situated at 1150 Å in the far wing, due to the B3-X dipole forbidden transition. Normally such an effect would be overshadowed by the allowed transition wing, but in this case there is no large contribution of the dipole allowed transitions in this region, as can be easily predicted by the examination of Fig. 1. The extrema of the allowed B4-X and D-X transitions occur for very short distances, and are much smaller compared to the position and depth of the outer well in the B3-X transition (see Sect. 3.1 and 3.2).
The collision-induced absorption depends on the internuclear separation and produces very broad spectral lines with a characteristic width of the order of the inverse of the duration of the close collision. It is strongly dependent on
* the temperature
* the amplitude of the dipole for $`R_{\mathrm{min}}`$ when the potential difference presents a minimum.
This emphasizes the importance of the accuracy of both the potential energies and the dipole moments for the line shape calculations.
Oscillatory structures appear near the 1150 Å satellite as they appear between the line and the 1600 Å satellite in both theory and experiment allard99 . These oscillations were predicted by Royer royer71b and Sando et al. sando69 ; sando73 .
### 4.2 Simultaneous perturbations by H and H<sup>+</sup>
The complete Lyman $`\beta `$ profile perturbed by collisions with neutral hydrogen and protons is shown in Fig 3. We notice that the collision-induced H-H satellite is much broader than the allowed H-H<sup>+</sup> satellite since the dipole moment differs from zero only over a short range of internuclear distances (see Sect. 4.1). The CI satellite of Lyman $`\beta `$ is quite far from the unperturbed Lyman $`\beta `$ line center, actually closer to the Lyman $`\alpha `$ line. It is therefore necessary to take into proper account the total contribution of both the Lyman $`\alpha `$ and Lyman $`\beta `$ wings of H perturbed simultaneously by neutrals and protons and to study the variation of this part of the Lyman series with the relative density of ionized and neutral atoms.
In allard99 we evaluated both the Lyman $`\alpha `$ and Lyman $`\beta `$ wings of H perturbed by protons. However, we neglected interference terms between the two lines. Equation (51) of allard99 , which gives the profile for a pair of lines such as Lyman $`\alpha `$ and $`\beta `$, is
$`I(\omega )=\varphi _\alpha ^{(0)}e^{nf_\alpha (0)}J_\alpha (\omega \omega _\alpha )`$
$`+\varphi _\beta ^{(0)}e^{nf_\beta (0)}J_\beta (\omega \omega _\beta ).`$ (32)
The perturbed line strength $`\varphi _\alpha ^{(0)}e^{nf_\alpha (0)}`$ differs from the free line strength $`\varphi _\alpha ^{(0)}`$ by the factor $`e^{nf_\alpha (0)}`$. This density-dependent factor expresses the fact that the total intensity radiated increases or decreases when the dipole moment is increased or decreased, on average, by the proximity of perturbers. Because of the low densities we consider, we have neglected this factor here.
We show the sum of the profiles of Lyman $`\alpha `$ and Lyman $`\beta `$ in Fig. 4. We can see that a ratio of 5 between the neutral and proton density is enough to make the CI satellite appear in the far wing. The CI satellite appearance is then very sensitive to the degree of ionization and may be used as a temperature diagnostic.
## 5 Conclusions
In the case of Lyman $`\alpha `$ and the H-H<sup>+</sup> Lyman $`\beta `$ satellites, the potential shape played a dominant role in the large difference in the broadening of the quasi-molecular features allard98b . The width of the collision-induced absorption is determined, for the most part, by the short range over which the corresponding transition dipole moment is significant. In the CI satellite, it is the dipole moment which is very important and which is responsible of the shape of the satellite, the observation of such a satellite would be a test of the accuracy of the dipole moment calculation. Satellites due to allowed and forbidden transitions depend linearly on density. The CI satellite is very sensitive to the temperature of the absorption, and it may also be used as a diagnostic tool for temperature. We emphasize that the effect of finite collision duration does play a role in the shape of the far wing. The present calculations are done in an adiabatic approximation using a rectilinear trajectory. This should affect slightly the shape of the satellite, although no great error is expected. We are developing methods to include trajectory effects in the evaluation of the line shape.
## Acknowledgements
The computations of dipole transition moments were performed on the CRAY of the computer center IDRIS. The work at the University of Louisville is supported by a grant from the U.S. Department of Energy, Division of Chemical Sciences, Office of Basic Energy Sciences, Office of Energy Research.
|
warning/0006/gr-qc0006009.html
|
ar5iv
|
text
|
# The classical regime of a quantum universe obtained through a functional method.
## I Introduction.
One of the most important problems of theoretical physics in the last years was to answer the question: How and in what circumstances a quantum system becomes classical ? . In spite of the great effort made by the physicists to find the answer, the problem is still alive and we are far from a complete understanding of many of its most fundamental features. In fact the most developed and sophisticated theory on the subject: histories decoherence is not free of strong criticisms .
Nevertheless there is an almost unanimous opinion that the classical regime is produced by two phenomena:
i. Decoherence, that in quantum systems, restore the boolean statistic typical of quantum mechanics and
ii.-Correlations, that circumvent the uncertainty relation at the macroscopic level.
But the techniques to deal with these two phenomena are not yet completely developed. One of the main problems is to find a proper and unambiguous definition of the, so called, pointer basis, where, decoherence takes place.
Our contribution to solve this problem is based in old ideas of Segal and van Howe , reformulated by Antoniou et al. . We have developed these ideas in papers and where we have shown how the Riemann-Lebesgue theorem can be used to prove the destructive interference of the off-diagonal terms of the state density matrix yielding decoherence. Using this technique we have found decoherence and correlations in simple quantum systems where we have defined the pointer basis in an unambiguous way. <sup>*</sup><sup>*</sup>*The relation of our method with the histories decoherence is studied in paper . They turn out to be equivalent, but in our method the pointer basis is better defined..
On the other hand, the appearance of a classical universe in quantum gravity models is the cosmological version of the problem we are discussing. Then, decoherence and correlations must also appear in the universe . In this paper, using our method, we will solve this problem in a simple quantum-cosmological model and we will find:
i.-Decoherence in all the dynamical variables and in a well defined pointer basis.
ii.-Correlations, in such a way that the Wigner function $`F_W`$ of the asymptotic diagonal matrix $`\rho _{}`$ can be expanded as:
$$F_W(x,p)=p_{\{l\}[𝐚]}F_{W\{l\}[𝐚]}(x,p)d\{l\}d[𝐚]$$
(1)
where $`F_{W\{l\}[𝐚]}`$ is a classical density strongly peaked Precisely: peaked as allowed by the uncertainty principle. in a trajectory defined by the initial conditions a and the momenta $`l`$ and $`p_{\{l\}[𝐚]}`$ the probability of each trajectory. As the limit of quantum mechanics is not classical mechanics but classical statistical mechanics this is our final result: The density matrix is translated in a classical density, via a Wigner function, and it is decomposed as a sum of densities peaked around all possible classical trajectories, each one of these densities weighted by their own probability.
Thus our quantum density matrix behaves in its classical limit as a statistical distribution among a set of classical trajectories. Similar results are obtained in papers and .
## II The model.
Let us consider the flat Roberson-Walker universe (, ) with a metric:
$$ds^2=a^2(\eta )(d\eta ^2dx^2dy^2dz^2)$$
(2)
where $`\eta `$ is the conformal time and $`a`$ the scale of the universe. Let us consider a free neutral scalar field and let us couple this field with the metric, with a conformal coupling ($`\xi =\frac{1}{6})`$. The total action reads $`S=S_g+S_f`$ $`+S_i`$ and the gravitational action is:
$$S_g=M^2𝑑\eta [\frac{1}{2}\stackrel{2}{\stackrel{}{a}}V(a)]$$
(3)
where $`M`$ is the Planck mass, $`\stackrel{}{a}=da/d\eta ,`$ and the potential $`V`$ contains the a cosmological constant term and eventually the contribution of some form of classical mater. We suppose that $`V`$ has a bounded support $`0aa_1.`$ We expand the field $`\mathrm{\Phi }`$ as:
$$\mathrm{\Phi }(\eta ,𝐱)=f_𝐤e^{i𝐤𝐱}𝑑𝐤$$
(4)
where the components of $`𝐤`$ are three continuous variables.
The Wheeler De-Witt equation for this model reads:
$$H\mathrm{\Psi }(a,\mathrm{\Phi })=(h_g+h_f+h_i)\mathrm{\Psi }(a,\mathrm{\Phi })=0$$
(5)
where:
$`h_g={\displaystyle \frac{1}{2M^2}}_a^2+M^2V(a)`$
$`h_f={\displaystyle \frac{1}{2}}{\displaystyle (_𝐤^2k^2f_𝐤^2)𝑑𝐤}`$
$$h_i=\frac{1}{2}m^2a^2f_𝐤^2𝑑𝐤$$
(6)
being $`m`$ is the mass of the scalar field, $`𝐤/a`$ is the linear momentum of the field, and $`_𝐤`$ =$`/f_𝐤.`$
We can now go to the semiclassical regime using the WKB method (), writing $`\mathrm{\Psi }(a,\mathrm{\Phi })`$ as:
$$\mathrm{\Psi }(a,\mathrm{\Phi })=\mathrm{exp}[iM^2S(a)]\chi (a,\mathrm{\Phi })$$
(7)
and expanding $`S`$ and $`\chi `$ as:
$$S=S_0+M^1S_1+\mathrm{},\chi =\chi _0+M^1\chi _1+\mathrm{}$$
(8)
To satisfy eq. (5) at the order $`M^2`$ the principal Jacobi function $`S(a)`$ must satisfy the Hamilton-Jacobi equation:
$$\left(\frac{dS}{da}\right)^2=2V(a)$$
(9)
We can now define the (semi)classical time as a parameter $`\eta =\eta (a)`$ such that:
$$\frac{d}{d\eta }=\frac{dS}{da}\frac{d}{da}=\pm \sqrt{2V(a)}\frac{d}{da}$$
(10)
The solution of this equation is $`a=\pm F(\eta ,C),`$ where $`C`$ is an arbitrary integration constant. Different values of this constant and of the $`\pm `$ sign give different classical solutions for the geometry.
Then, in the next order of the WKB expansion, the Schroedinger equation reads:
$$i\frac{d\chi }{d\eta }=h(\eta )\chi $$
(11)
where:
$$h(\eta )=h_f+h_i(a)$$
(12)
precisely:
$$h(\eta )=\frac{1}{2}\left[\frac{^2}{f_𝐤^2}+\mathrm{\Omega }_𝐤^2(a)f_k^2\right]𝑑𝐤$$
(13)
where:
$$\mathrm{\Omega }_\varpi ^2=m^2a^2+k^2=m^2a^2+\varpi $$
(14)
where $`\varpi =k^2`$ and $`k=|𝐤|.`$ So the time dependence of the hamiltonian comes from the function $`a=a(\eta ).`$
Let us now consider a scale of the universe such that $`a_{out}a_1`$. In this region the geometry is almost constants. Therefore we have an adiabatic final vacuum $`|0`$ and adiabatic creation and annihilation operators $`a_𝐤^{}`$ and $`a_𝐤.`$ Then $`h=h(a_{out})`$ reads:
$$h=\mathrm{\Omega }_\varpi a_𝐤^{}a_𝐤𝑑𝐤$$
(15)
We can now consider the Fock space and a basis of vectors:
$$|𝐤_1,𝐤_2,\mathrm{},𝐤_n,\mathrm{}|\{k\}=a_{𝐤_1}^{}a_{𝐤_2}^{}\mathrm{}a_{𝐤_n}^{}\mathrm{}|0$$
(16)
where we have called $`\{k\}`$ to the set $`𝐤_1,𝐤_2,\mathrm{},𝐤_n,\mathrm{}`$ The vectors of this basis are eigenvectors of $`h:`$
$$h|\{k\}=\omega |\{k\}$$
(17)
where:
$$\omega =\underset{𝐤\{𝐤\}}{}\mathrm{\Omega }_\varpi =\underset{𝐤\{𝐤\}}{}(m^2a_{out}^2+\varpi )^{\frac{1}{2}}$$
(18)
We can now use this energy to label the eigenvector as:
$$|\{k\}=|\omega ,[𝐤]$$
(19)
where $`[𝐤]`$ is the remaining set of labels necessary to define the vector unambiguously. $`\{|\omega ,[𝐤]\}`$ is obviously an orthonormal basis so eq. (15) reads:
$$h=\omega |\omega ,[𝐤]\omega ,[𝐤]|𝑑\omega d[𝐤]$$
(20)
In the next section we will write this equation using a shorthand notation as:
$$h=\omega |\omega \omega |𝑑\omega $$
(21)
The dynamical variables $`[𝐤]`$ will reappear in section IV.
## III Energy decoherence.
As we are dealing with a system with a continuous spectrum in $`\omega ,`$ some care must be taken. If not the mathematical manipulations can contain multiplication of distribution and yield infinite meaningless results. In order to deal with this problem and to always work with usual functions (not distribution) a functional method was introduced in papers and , that we will now review. The method was used to show the decoherence and the existence of correlations in ordinary quantum mechanical systems and we will use it in our problem.
The physical basis of the method is the following: The states of the universe are only known and measure trough a measurement process where a space of observables $`𝒪`$ is used. For any observable $`O𝒪`$ we can only measure the mean value of this $`O`$ in a state $`\rho ,`$ namely:
$$O_\rho =Tr(\rho ^{}O)$$
(22)
Then we can consider that the states are linear functionals over the space of observables and write:
$$O_\rho =\rho [O]=(\rho |O)$$
(23)
Of course, the states would be endowed with extra some properties, so we will define a convex set of observables $`𝒮𝒪^{}`$ , being this last space the dual of $`𝒪`$ .
It is logical to ask that the hamiltonian $`h`$ would be contained in the space of observables $`𝒪`$ , then the observables must be defined generalizing eq. (21). This generalization, already used in papers and , reads:
$`O={\displaystyle O_\omega |\omega \omega |d\omega +O_{\omega \omega ^{}}|\omega \omega ^{}|𝑑\omega 𝑑\omega ^{}}=`$
$$O_\omega |\omega )d\omega +O_{\omega \omega ^{}}|\omega ,\omega ^{})d\omega d\omega ^{}$$
(24)
where we have introduced a $`basis`$ $`\{|\omega ),|\omega ,\omega ^{})\}`$of space $`𝒪`$ defined as:
$$|\omega )=|\omega \omega |,|\omega ,\omega ^{})=|\omega \omega ^{}|$$
(25)
The terms $`O_\omega `$ can be considered as the (singular) diagonal terms, while the terms $`O_{\omega \omega ^{}\text{ }}`$ can be considered as the (regular) off-diagonal terms.
We can now define the $`cobasis`$ $`\{(\omega |,(\omega ,\omega ^{}|\}`$ of space $`𝒪`$ (namely the basis of space $`𝒮𝒪^{}`$), that obviously satisfies the equations:
$$(\omega |\omega ^{})=\delta (\omega \omega ^{}),(\omega ,\omega ^{\prime \prime }|\omega ^{},\omega ^{\prime \prime \prime })=\delta (\omega \omega ^{})\delta (\omega ^{\prime \prime }\omega ^{\prime \prime \prime })$$
(26)
and all other $`(.|.)=0.`$
Then if $`\rho 𝒮`$ it can be expanded as:
$$\rho =\rho _\omega (\omega |d\omega +\rho _{\omega \omega ^{}}(\omega ,\omega ^{}|d\omega d\omega ^{}$$
(27)
where $`\rho _\omega 0,`$ $`\rho _{\omega \omega ^{}}=\rho _{\omega ^{}\omega }^{}.`$ Moreover, the ordinary functions $`O_\omega ,O_{\omega \omega ^{},}\rho _\omega ,`$ and $`\rho _{\omega \omega ^{}},`$ must be endowed with certain properties in order to make all the equations of the formalism well defined. These properties are listed in and they are assumed in this paper. Then:
$$(\rho |O)=\rho _\omega O_\omega 𝑑\omega +\rho _{\omega \omega ^{}}O_{\omega ^{}\omega }𝑑\omega 𝑑\omega ^{}$$
(28)
and from eq. (11):
$$(\rho (\eta )|O)=\rho _\omega O_\omega 𝑑\omega +\rho _{\omega \omega ^{}}O_{\omega ^{}\omega }e^{i(\omega \omega ^{})\eta }𝑑\omega 𝑑\omega ^{}$$
(29)
Then, when $`\eta \mathrm{},`$ essentially from the Riemann-Lebesgue theorem (see for details), we have:
$$\underset{\eta \mathrm{}}{lim}(\rho (\eta )|O)=\rho _\omega O_\omega 𝑑\omega =(\rho _{}|O)$$
(30)
where:
$$(\rho _{}|=\rho _\omega (\omega |d\omega $$
(31)
is the equilibrium time-asymptotic state, which only contains the diagonal term. So we have proved the existence of decoherence in the energy.
## IV Decoherence in the other dynamical variables.
If we reintroduce the other dynamical variables in eq. (31) we obtain:
$$(\rho _{}|=\rho _{\omega [𝐤][𝐤^{}]}(\omega ,[𝐤],[𝐤^{}]|d\omega d[𝐤]𝐝[𝐤^{}]$$
(32)
where {$`(\omega ,[𝐤],[𝐤^{}]|,(\omega ,\omega ^{},[𝐤],[𝐤^{}]\}`$ is the cobasis {$`(\omega |,(\omega ,\omega ^{}|\}`$ but now showing the hidden $`[𝐤].`$
Let us observe that if we would use polar coordinates for $`𝐤`$ eq.(4) reads:
$$\mathrm{\Phi }(x,n)=\underset{lm}{}\varphi _{klm}dk$$
(33)
where:
$$\varphi _{klm}=f_{k,l}(\eta ,r)Y_m^l(\theta ,\phi )$$
(34)
where $`k`$ is a continuous variable, $`l=0,1,\mathrm{},;`$ $`m=l,\mathrm{},l;`$ and $`Y`$ are spherical harmonic functions. So the indices $`k,l,m`$ contained in the symbol $`𝐤`$ are partially discrete and partially continuous.
As $`\rho _{}^{}=\rho _{}`$ then $`\rho _{\omega [𝐤^{}][𝐤]}^{}=`$ $`\rho _{\omega [𝐤][𝐤^{}]}`$ and therefore a set of vectors $`\{|\omega ,[𝐥]\}`$ exists such that:
$$\rho _{\omega [𝐤][𝐤^{}]}|\omega ,[𝐥]_{[𝐤^{}]}d[𝐤^{}]=\rho _{\omega [𝐥]}|\omega ,[𝐥]_{[𝐤]}$$
(35)
namely {$`|\omega ,[𝐥]\}`$ is the eigenbasis of the operator $`\rho _{\omega [𝐤][𝐤^{}]}.`$ Then $`\rho _{\omega [𝐥]}`$ can be considered as an ordinary diagonal matrix in the discrete indices like the $`l`$ and the $`m`$, and a generalized diagonal matrix in the continuous indices like $`k`$ E. g.: We can deal with this generalized matrix rigging the space $`𝒮`$ and using the Gel’fand-Maurin theorem , this procedure allows us to define a generalized state eigenbasis for system with continuous spectrum. It has been used to diagonalize hamiltonians with continuous spectra in , , , etc.. Under the diagonalization process eq. (32) is written as:
$$(\rho _{}|=U_{[𝐤]}^{[𝐥]}\rho _{\omega [𝐥][𝐥^{}]}U_{[𝐤^{}]}^{[𝐥^{}]}U_{[𝐤^{}]}^{[𝐥^{\prime \prime }]}(\omega ,[𝐥^{\prime \prime }],[𝐥^{\prime \prime \prime }]|U_{[𝐤]}^{[𝐥^{\prime \prime \prime }]}d\omega d[𝐤]𝐝[𝐤^{}]d[𝐥]𝐝[𝐥^{}]d[𝐥^{\prime \prime }]𝐝[𝐥^{\prime \prime \prime }]$$
(36)
where $`U_{[𝐤]}^{[𝐥]}`$ is the unitary matrix used to perform the diagonalization and:
$$\rho _{\omega [𝐥][𝐥^{}]}=\rho _{\omega [𝐥]}\delta _{[𝐥][𝐥^{}]}$$
(37)
where:
$$\rho _{\omega [𝐥][𝐥]}=\rho _{\omega [𝐥]}=U_{[𝐥]}^{[𝐤]}\rho _{\omega [𝐤][𝐤^{}]}U_{[𝐥]}^{[𝐤^{}]}d[𝐤]d[𝐤^{}]$$
(38)
so we can define:
$$(\omega ,[𝐥]|=(\omega ,[𝐥],[𝐥]|=U_{[𝐥]}^{[𝐤]}(\omega ,[𝐤],[𝐤^{}]|U_{[𝐥]}^{[𝐤^{}]}d[𝐤]d[𝐤^{}]$$
(39)
We can repeat the procedure with vectors $`(\omega ,\omega ^{},[𝐤],[𝐤^{}]|`$ and obtain vector $`(\omega ,\omega ^{},[𝐥]|.`$ In this way we obtain a diagonalized cobasis {$`(\omega ,[𝐥]|,(\omega ,\omega ^{},[𝐥]\}.`$ So we can now write the equilibrium state as:
$$\rho _{}=\rho _{\omega [𝐥]}(\omega ,[𝐥]|d\omega d[𝐥]$$
(40)
Since vectors $`(\omega ,[𝐥]|`$ can be considered as diagonals in all the variables we have obtained decoherence in all the dynamical variables. This fact will become more clear once we study the observables related with this vector and introduce the notion of pointer basis.
So, let us now consider the observable basis {$`|\omega ,[𝐥]),|\omega ,\omega ^{},[𝐥])\}`$ dual to the state cobasis $`\{(\omega ,[𝐥]|,(\omega ,\omega ^{},[𝐥]|\}.`$ From eq. (25) and as the $`\omega `$ does not play any role in the diagonalization procedure we obtain:
$$|\omega ,[𝐥])=|\omega ,[𝐥]\omega ,[𝐥]|,|\omega ,\omega ^{},[𝐥])=|\omega ,[𝐥]\omega ^{},[𝐥]|$$
(41)
So in the basis {$`|\omega ,[𝐥]),|\omega ,\omega ^{},[𝐥])\}`$ the hamiltonian reads:
$$h=\omega |\omega ,[𝐥])d\omega d[𝐥]=\omega |\omega ,[𝐥]\omega ,[𝐥]|d\omega d[𝐥]$$
(42)
Now, we can also define the operators:
$$𝐋=𝐥|\omega ,[𝐥])d\omega d[𝐥]=𝐥|\omega ,[𝐥]\omega ,[𝐥]|d\omega d[𝐥]$$
(43)
that can also be written:
$$L_i=l_i|\omega ,[𝐥])d\omega d[𝐥]=l_i|\omega ,[𝐥]\omega ,[𝐥]|d\omega d[𝐥]$$
(44)
where $`i`$ is an index such that it covers all the dimension of the $`𝐥`$ . Now we can consider the set $`(h,L_i),`$ which is a CSCO, since all the members of the set commute, because they share a common basis and find the corresponding eigenbasis of the set, precise $`|\omega ,[𝐥]`$ since <sup>§</sup><sup>§</sup>§In some occasions we will call $`h=L_0`$ and $`\omega =l_0.`$:
$$h|\omega ,[𝐥]=\omega |\omega ,[𝐥]$$
(45)
$$L_i|\omega ,[𝐥]=l_i|\omega ,[𝐥]$$
(46)
Of course the $`L_i`$ are constant of the motion because they commute with $`h.`$
From all these equations we can say that:
i.- $`(h,L_i)`$ is the pointer CSCO.
ii.- {$`|\omega ,[𝐥]),|\omega ,\omega ^{},[𝐥])\}`$ is the pointer observable basis.
iii.-$`\{(\omega ,[𝐥]|,(\omega ,\omega ^{},[𝐥]|\}`$ is the pointer states cobasis.
In fact, from eq. (40) we see that the final equilibrium state has only diagonal terms in this state (those corresponding to vectors $`(\omega ,[𝐥]|)`$ , it has not off-diagonal terms (those corresponding to vectors $`(\omega ,\omega ^{},[𝐥]|,(\omega ,[𝐤],[𝐤^{}]|,`$ or $`(\omega ,\omega ^{},[𝐤],[𝐤^{}]|),`$ and therefore we have decoherence in all the dynamical variables.
## V Correlations.
As it was explained in paper correlations are computed in the limit of small $`\mathrm{}.`$ Under this assumption it is demonstrated that for each observable (e. g.: momenta or energy) we can find a canonically conjugated dynamical variable (e. g.: configuration variables or the hand of a clock, namely time), if we neglect $`O(\mathrm{})`$ terms. So we will use these approximated canonically conjugated variables in this section, since, we repeat, we are only interested in observational conditions where $`\mathrm{}`$ can be considered very small.
Accordingly with this idea the canonically conjugated variable of $`h`$ would be essentially $`\eta ,`$ but since $`\rho _{}`$ is a $`\eta `$-constant, the time variable is completely unimportant in this section (we will discuss this matter further on section 6.3). Let us call $`a_i`$ the canonically conjugated variable (precisely ”conjugated variable up to $`O(\mathrm{})`$ terms”) of the observable $`L_i.`$ Then, $`(a_i)`$ will be our configuration variables and $`(L_i`$) our momentum variables. (We will call $`x,p`$ to the old variables of eq. (2) and $`\xi `$ or $`𝐚`$ to the new configuration variable and $`\pi `$ or $`𝐥`$ to the new momentum variables). Using these new variables we will compute the Wigner function corresponding to the operator $`\rho _.`$ We can also use the usual transcription rules:
$$hi\frac{}{\eta },L_ii\frac{}{a_i}$$
(47)
since the difference with respects to other transcription rules in other coordinates is just a $`O(\mathrm{}).`$ Then:
$$\eta ,[\mathrm{\Delta }𝐚+𝐚_0]|\omega [𝐥]=e^{i(\omega \eta +𝐥𝚫𝐚)}0,[𝐚_0]|\omega [𝐥]$$
(48)
But we will only consider the state of affairs for $`\eta =0.`$ We will call:
$$[𝐚]=[\mathrm{\Delta }𝐚+𝐚_0],\{l\}=(\omega ,[𝐥]),\{\pi \}=(\omega ,[\pi ])$$
(49)
where, in the second and third equations we have restored the notation of eq. (16) With this notation and for $`\eta =0`$, eq. (48) reads:
$$[𝐚]|\{l\}=e^{i\mathrm{\Delta }𝐚𝐥}[𝐚]|\{l\}$$
(50)
The Wigner function corresponding to matrix $`\rho _{}`$ reads:
$$F_W(x,p)=F_W(\xi ,\pi )_{\mathrm{}}^{\mathrm{}}\xi \eta |\rho _{}|\xi +\eta e^{2i\pi \eta }d[\eta ]$$
(51)
Then from eqs. (40) and (50) (and eq. (65), written for the spatial coordinates for the continuous indices, see details in ) we have:
$`F_W(x,p){\displaystyle }\mathrm{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}d[\eta ]{\displaystyle }d\{l\}\rho _{\{l\}}\xi \eta |(\{l\}||\xi +\eta e^{2i\pi \eta }=`$
$`{\displaystyle \mathrm{}_{\mathrm{}}^{\mathrm{}}d[\eta ]d\{l\}\rho _{\{l\}}e^{i(\xi \eta )l}(\{l\}|[𝐚_0])e^{i(\xi +\eta )l}e^{2i\pi \eta }}`$
$$d\{l\}\rho _{\{l\}}|\{l\}|[𝐚_0]|^2\delta ([\pi ][𝐥])$$
(52)
where the probability $`(\{l\}|[𝐚_0])`$ has been called with the more familiar (but not rigorous) symbol $`|\{l\}|[𝐚_0]|^2`$ (that turns out to be rigorous only for the discrete $`l)`$ We see that $`\xi `$ disappears from the equation, so $`F_W(\xi ,\pi )`$ is neither a function of the position $`\xi `$ nor of the conventional origin $`𝐚_0`$. This is a consequence of the spatial homogeneity of the model we are studying. Moreover, it can also be seen that if we use eq. (48) with $`\eta 0`$ the function $`F_W(\xi ,\pi )`$ is a constant of $`\eta `$ (as it should be). So, essentially, in all this section we are dealing with functions that are constants in time.. The $`\delta ([\pi ][𝐥])`$ does not contain the energy. But in the footnote of section 6.3 we will prove that a $`\delta `$term in the energy can also be added, so finally:
$$F_W(x,p)d\{l\}\rho _{\{l\}}|\{l\}|[𝐚_0]|^2\delta (\{\pi \}\{l\})=d\{l\}\rho _{\{l\}}|\{l\}|[𝐚_0]|^2\underset{i=0}{}\delta (\pi _il_i)$$
(53)
The last equation can be interpreted as follows:
i.- $`\delta (\{p\}\{l\})`$ is a classical density function, strongly peaked at certain values of the constants of motion $`\{l\},`$ corresponding to a set of trajectories, where the momenta are equal to the eigenvalues of eqs. (45) and (46), namely $`\pi _i=l_i`$ $`(i=0,1,2,\mathrm{})`$. This fact already shows the presence of correlations in our model. In fact: We can consider each set of trajectories labelled by $`\{l\}`$ (i.e. a ”history” obtained using some apparatuses that measure only the momenta) and prove that in these trajectories the usual coordinate $`x`$ and the usual momentum $`p`$ are correlated as it is allowed by the uncertainty principle (see ). For the conjugated variables $`𝐥`$ and $`𝐚,𝐥`$ is completely defined and $`𝐚`$ is completely undefined, satisfying also the uncertainty principle.
ii.- $`\rho _{\{l\}}`$ is the probability to be in one of these sets of trajectories labelled by $`\{l\}`$. Precisely: if some initial density matrix is given, from eq. (40) it is evident that its diagonal terms $`\rho _{\{l\}}`$ are the probabilities to be in the states $`(\omega ,[𝐥]|`$ and therefore the probability to find, in the corresponding classical equilibrium density function $`F_W(x,p)`$, the density function $`\delta (\{p\}\{l\}),`$ namely the set of trajectories labelled by $`\{l\}=(\omega ,[𝐥]).`$
iii.- The factor $`|[𝐚_0]|\{l\}|^2`$ corresponds to the probability that one of the trajectories $`\{l\}`$ would pass by $`𝐚_0`$ at time $`\eta =0`$ and it can easily be computed from the model From the spatial homogeneity of the problem and the usual normalization we have $`(\{l\}|[𝐚_0])=|[𝐚_0]|\{l\}|^2\omega ^n`$, being $`n`$ the particle number..
iv.- Therefore $`\rho _{\{l\}}|[𝐚_0]|\{l\}|^2=p_{\{l\}[𝐚_0]}`$ is the probability that, given an initial density matrix, a trajectory, with constant of the motion $`\{l\}`$ would pass by the point $`𝐚_0`$ at time $`\eta =0`$, and then it would follow the classical trajectory:
$$𝐚=𝐥\eta +𝐚_0$$
(54)
But, really $`p_{\{l\}[𝐚_0]}`$ is not a function of $`𝐚_0`$, it is simply a constant in $`𝐚_0`$ (as explained in a previous footnote) since this is only an arbitrary point and our model is spatially homogenous, we can write:
$$p_{\{l\}[𝐚_0]}=p_{\{l\}[𝐚_0]}\underset{i=1}{}\delta (\xi _ia_{0i})d[𝐚_0]$$
(55)
in this way we have changed the role of $`𝐚_0`$, it was a fixed (but arbitrary) point and it is now a variable that moves all over the space. Then eq. (53) reads:
$$F_W(x,p)p_{\{l\}[𝐚_0]}\underset{i=0}{}\delta (\pi _il_i)\underset{j=1}{}\delta (\xi _ja_{0j})d[𝐚_0]d\{l\}$$
(56)
So if we call :
$$F_{W\{l\}[𝐚_0]}(x,p)=\underset{i=0}{}\delta (\pi _il_i)\underset{j=1}{}\delta (\xi _ja_{0j})$$
(57)
we have:
$$F_W(x,p)p_{\{l\}[𝐚_0]}F_{W\{l\}[𝐚_0]}(x,p)d[𝐚_0]d\{l\}$$
(58)
From eq. (57) we see that $`F_{W\{l\}[𝐚_0]}(x,p)0`$ only in a narrow strip around the classical trajectory (54) defined by the momenta $`\{l\}`$ and passing through the point \[$`𝐚_0]`$ (really the density function is as peaked as it is allowed by the uncertainty principle, so its width is essentially a $`O(\mathrm{}),),`$ since the $`\delta `$functions of all the equation are really Dirac’s deltas when $`\mathrm{}0)`$ These trajectories explicitly show the presence of correlations in our model <sup>\**</sup><sup>\**</sup>\**Of course, our ”trajectories” are not only one trajectory for a one particle state, but they are $`n`$ trajectories (each one corresponding to a momenta $`(l_1,l_2,\mathrm{}l_n)=\{l\}`$ and passing by a point ($`𝐚_1,𝐚_2,\mathrm{},𝐚_n)=[𝐚])`$ for the n particle states. As $`p_{\{l\}\{a\}}\omega ^n`$ the probability decreases with the particle number and the energy.. So we have proved eq. (58) which, in fact, it is eq. (1) as announced <sup>††</sup><sup>††</sup>††In this section we have faced the following problem: $`F_W(\xi ,\pi )`$ is a $`\xi `$ constant that we want to decompose in functions $`F_{W\{l\}[𝐚_0]}(x,p)`$ which are different from zero only around the trajectory (54) and therefore are variables in $`\xi .`$Then,essentially we use the fact that if $`f(x,y)=g(y)`$ is a constant function in $`x`$ we can decompose it as:
$`g(y)={\displaystyle g(y)\delta (xx_0)𝑑x_0}`$
namely the densities $`\delta (xx_0)`$ are peaked in the trajectories $`x=x_0=const.,y=var.`$ and, therefore, are functions of $`x.`$ This trajectories play the role of those of eq. (55). As all the physics, including the correlations, is already contained in eq. (53) (as explained in point i) the reader may just consider the final part of this section, from eq. (55) to eq. (58) a didactical trick..
Then we have obtained the classical limit. When $`\eta \mathrm{}`$ the quantum density $`\rho `$ becomes a diagonal density matrix $`\rho _{}.`$ The corresponding classical distribution $`F_W(x,p)`$ can be expanded as a sum of classical trajectories density functions $`F_{W\{l\}[𝐚_0]}(x,p),`$ each one weighted by its corresponding probability $`p_{\{l\}[𝐚_0]}.`$ So, as the limit of our quantum model we have obtained a statistical classical mechanical model, and the classical realm appears.
## VI Discussion and comments.
### A Characteristic times.
The decaying term of eq. (29) (i. e. the second term of the r. h. s.) can be analytically continued using the techniques explained in papers , , and . In these papers it is shown that each pole $`z_i=\omega _ii\gamma _i,`$ of the S-matrix of the problem considered, originates a damping factor $`e^{\gamma _i\eta }.`$ Then if $`\gamma =\mathrm{min}(\gamma _i)`$ the characteristic decoherence time is $`\gamma ^1.`$ This computation is done in the specific models of papers . If $`\gamma 1,`$ even if the Riemann-Lebesgue theorem is always valid, there is no practical decoherence since $`\gamma ^11.`$
### B Sets of trajectories decoherence.
It is usual to say that in the classical regime there is decoherence of the set trajectories labelled by the constant of the motion $`\omega ,`$ $`[𝐥]`$. This result can easily be obtained with our method in the following way.
i.- Let us consider two different states $`|\omega [𝐥]`$ and $`|\omega ^{}[𝐥^{}]`$ that will define classes of trajectories with different constants of the motion $`(\omega ,[𝐥])(\omega ^{},[𝐥^{}]).`$ We must compute:
$$\omega [𝐥]|\rho _{}|\omega ^{}[𝐥^{}]=(\rho _{}||\omega \omega ^{}[𝐥][𝐥^{}])=[\rho _{\omega ^{\prime \prime }[𝐥^{\prime \prime }]}(\omega ^{\prime \prime }[𝐥^{\prime \prime }]|d\omega ^{\prime \prime }d[𝐥^{\prime \prime }]]|\omega \omega ^{}[𝐥][𝐥^{}])=0$$
(59)
due to the orthogonality of the basis $`\{(\omega ,[𝐥]|,(\omega ,\omega ^{},[𝐥]|\}`$ .
ii.- But if we compute:
$`\omega [𝐥]|\rho _{}|\omega [𝐥]=(\rho _{}||\omega [𝐥])=[{\displaystyle }\rho _{\omega ^{\prime \prime }[𝐥^{\prime \prime }]}(\omega ^{\prime \prime }[𝐥^{\prime \prime }]|d\omega ^{\prime \prime }d[𝐥^{\prime \prime }]]|\omega \omega [𝐥])=`$
$$\rho _{\omega ^{\prime \prime }[𝐥^{\prime \prime }]}\delta (\omega \omega ^{\prime \prime })\delta ([𝐥][𝐥^{\prime \prime }])𝑑\omega ^{\prime \prime }d[𝐥^{\prime \prime }]=\rho _{\omega [𝐥]}0$$
(60)
The last two equations complete the demonstration. We will discuss the problem of the decoherence of two trajectories, with the same $`\{l\}`$ but different $`[𝐚_0]`$ in subsection 6.4.
### C A discussion on time decoherence.
It is well known that one of the main problems of quantum gravity is the problem of the time definition (see ). A not well studied feature of this problem is that, there must be a decoherence process related with time, since time is as a classical variable. In this subsection, using the functional technique, we will give a model that shows that this is the case.(but we must emphasize that this subject is not completely developed).
We must compute $`\eta |\rho _{}|\eta ^{}`$ where $`|\eta `$ and $`|\eta ^{}`$ are two states of the system for different times that evolve as <sup>‡‡</sup><sup>‡‡</sup>‡‡Cf. eq. (48) and remember that therefore in this subsection we are dealing with equations only valid when $`\mathrm{}0.`$:
$$|\eta =e^{ih\eta }|0$$
(61)
$`|\eta \eta ^{}|`$ can be considered as an observable, then:
$$\eta ^{}|\rho _{}|\eta =(\rho _{}||\eta \eta ^{}|)$$
(62)
But:
$$(\omega ||\eta \eta ^{}|)=(\omega |e^{ih\eta }|00|e^{ih\eta ^{}})=[e^{ih\eta ^{}}(\omega |e^{ih\eta }]||00|)$$
(63)
Now, for any observable $`O`$ we have:
$`[e^{ih\eta ^{}}(\omega |e^{ih\eta }]||O)=[e^{ih\eta ^{}}(\omega |e^{ih\eta }]|[{\displaystyle }O_\omega ^{}|\omega ^{})d\omega ^{}+{\displaystyle }{\displaystyle }O_{\omega ^{}\omega ^{\prime \prime }}|\omega ^{},\omega ^{\prime \prime })d\omega ^{}d\omega ^{\prime \prime })=`$
$`[e^{ih\eta ^{}}(\omega |e^{ih\eta }]|[{\displaystyle }O_\omega ^{}|\omega ^{})d\omega ^{}+\mathrm{}=(\omega |[{\displaystyle }O_\omega ^{}e^{i\omega ^{}\eta }|\omega ^{})e^{i\omega ^{}\eta ^{}}d\omega ^{}]=`$
$$e^{i\omega (\eta ^{}\eta )}(\omega |O)$$
(64)
Thus <sup>\**</sup><sup>\**</sup>\**Considering this equation and repeating the procedure done, from eq. (51) to eq. (53), we can see that there is an extra $`\delta `$factor $`\delta (\pi _0l_0)`$ related with the energy. Therefore the trajectories described in section 5 conserve, not only the momenta $`l`$ , but also the energy $`h.`$:
$$(\omega ||\eta \eta ^{}|)=e^{i\omega (\eta ^{}\eta )}(\omega ||00|)$$
(65)
So now we can compute the following two cases:
i.-
$$\eta ^{}|\rho _{}|\eta =(\rho _{}||\eta \eta ^{}|)=[\rho _\omega (\omega |d\omega ]||\eta \eta ^{}|)=\rho _\omega e^{i\omega (\eta ^{}\eta )}(\omega ||00|)d\omega 0$$
(66)
when$`|`$ $`\eta ^{}\eta |\mathrm{}`$ , due to the Riemann-Lebesgue theorem.
ii.- Analogously:
$$\eta |\rho _{}|\eta =\rho _\omega (\omega ||00|)d\omega 0$$
(67)
So we have time decoherence for two times $`\eta `$ and $`\eta ^{}`$ if they are far enough.
This result is important for the problem of time definition, since in order to have a reasonable classical time this variable must first decohere. The result above shows that this is the case for $`\eta `$ and $`\eta ^{}`$ far enough, but also that, for closer times (namely such that their difference is smaller than Planck’s time) there is no decoherence and time cannot be considered as a classical variable. Classical time is a familiar concept but the real nature of the non-decohered quantum time is opened to discussion.
### D Decoherence in the space variables.
Now that we know that there is time decoherence we can repeat the reasoning for the rest of the variables $`\xi `$ at time $`\eta =0`$ and changing eq. (61) by:
$$|\xi =e^{i\xi 𝐥}|\mathrm{𝟎}$$
(68)
and we will reach to the conclusions:
i.-
$$\xi |\rho _{}|\xi ^{}0$$
(69)
when $`|`$ $`\xi \xi ^{}|\mathrm{}.`$
ii,-
$$\xi |\rho _{}|\xi 0$$
(70)
therefore there is also decoherence between two trajectories with the same $`\{l\}`$ but different $`[𝐚_0]`$.
These facts complete the scenario about decoherence and correlations.
## VII Conclusion.
After the WKB expansion and the decoherence and correlations process our quantum model has:
i.-A defined classical time $`\eta `$ and a defined classical geometry related by eq. (11).
ii.- Decoherence has appeared in a well defined pointer basis.
iii.- The quantum field has originated a classical final distribution function (eq. (58)) that is a weighted average of some set densities, each one related to a classical trajectory. The weight coefficients are the probabilities of each trajectory.
We can foresee that if instead of a spinless field we would coupled the geometry with a spin 2 metric fluctuation field the result would be more or less the same. Then the corresponding quantum fluctuations would become classical fluctuations that would correspond to matter inhomogeneities (galaxies, clusters of galaxies, etc.) that will move along the trajectories described above. But this subject will be treated elsewhere with great detail.
## VIII Acknowledgments.
This work was partially supported by grants CI1-CT94-0004 and PSS$`{}_{}{}^{}0992`$ of the European Community, PID 3183/93 of CONICET, EX053 of the Buenos Aires University, and also grants from Fundación Antorchas and OLAM Foundation.
|
warning/0006/astro-ph0006342.html
|
ar5iv
|
text
|
# Dependence of Halo Properties on Interaction History, Environment and Cosmology
## 1 Introduction
The majority of spiral galaxies do not exist in isolation, but rather in an environment which sustains repeated close encounters and merging processes with both small and large companions (Schwarzkopf & Dettmar (2000)). Increasing evidence links disturbed or lopsided features of these galaxies to recent dynamical interactions. HST observations at high redshift have shown a preponderance of “train wreck” morphologies arising from major interactions. More recently, minor mergers have begun to be linked to $`m=1`$ mode disturbances as observed in the Fourier azimuthal decomposition of the galaxy’s surface brightness, generally denoted as “lopsidedness” (Rudnick, Rix, & Kennicut (2000); Rudnick & Rix (1998)). Specifically, studies by Richter & Sancisi (1994) have estimated that about half of late-type spirals exhibit significantly lopsided HI profiles, and half of these are field galaxies removed from cluster environments. Furthermore, Zaritsky & Rix find that 20% of all late-type spirals are significantly lopsided at optical and near-infrared wavelengths, yet have no obvious interactors. Rudnick & Rix (1998) find that one fifth of all disk galaxies, regardless of Hubble type, have azimuthal asymmetries in their stellar mass distribution. They also show that this observed asymmetry is in fact due to the underlying mass distribution and not dust obscuration or other non-dynamical effects. Dynamical profiles of this nature have been shown to be consistent with the accretion of a satellite galaxy of a larger system (Levine & Sparke (1998); Schwarzkopf & Dettmar (2000); and references therein). Consequently, it is probable that a substantial fraction of disk galaxies are experiencing interaction due to minor mergers.
Field spheroidals also have a number of properties such as shells, counterrotating inner disk, and bluer colors, which are believed to also be associated with recent merger events (de Carvalho & Djorgovski (1992); Rose et al. (1994); Longhetti et al. (1998); Abraham, et al. (1999)). By counting major mergers in large-scale simulations, Governato et al. (1999) show that a substantial fraction of the field elliptical population could be the result of major merging activity.
Hence, even in the field, observational evidence exists to support the hierarchical clustering scenario (HCS, see Peebles 1993) wherein larger objects are formed through the progressive assemblage of smaller mass collapsed structures. In short, the galaxy formation and evolution process is ubiquitously influenced by interactions. Therefore, to fully understand galactic formation and evolution and their consequences (both observational and theoretical), one must understand link between galaxy properties and interaction history, and how both of these may be modified by environmental effects.
The Press & Schechter formalism (1974, PS) is an analytical model to describe the hierarchical clustering scenario. The PS theory allows one to derive, in a relatively simple and straightforward way, the mass distribution of collapsed objects and its evolution with time in different FRW universes.
Despite its simple derivation, which relies uniquely on the linear theory, the PS formula captures the major feature of the hierarchical evolution as shown in the good fit of the PS mass function to N-body simulations (Frenk et al. (1988); Lacey & Cole (1994); Gardner et al. (1997); Borgani et al. (1999) and references therein). Recently it has become possible to pinpoint quantitative differences between this picture and the more complex N–body description. At cluster-sized masses, it is possible to match PS and simulated mass functions reasonably well (Governato et al. (1999); Borgani et al. (1999)), albeit with some tuning. However, at lower masses, evidence suggests that PS may provide a systematic overestimate of the halo number density (Bryan & Norman (1998); Gross et al. (1998); Tormen (1998)).
The PS formalism can be extended to give the probability that a halo of mass $`M_0`$ at redshift $`z_0`$ was also in a halo of mass $`M_1`$ at time $`z_1`$ (Bond et al. (1991)). This Extended Press-Schechter (EPS) can be further instrumented to construct halo formation times, mass histories, and merger rates in Monte-Carlo techniques (Lacey & Cole (1993)). Additional enhancements can be made which follow the detailed evolution and interaction of halos (Baugh et al. (1998) and references therein) in the form of semi-analytic merger trees (SAM) which are far less costly than N-body simulations. The SAM formalism has been used to predict a wide range of galaxy properties such as colors, morphology, merging history, and formation time (Baugh et al. (1996) and references therein). It has been used successfully to predict the same qualitative behavior found in N-body simulations. For instance, both SAM studies and N-body simulations find that the typical redshift of last major interaction of field ellipticals is lower than the cluster elliptical population (Governato et al. (1998); Baugh et al. (1996)). Furthermore, SAM results also generally reproduce observed behavior such as the morphology-density relation, color-magnitude diagrams, the Butcher-Oemler effect, and the luminosity evolution of color-selected galaxy samples (Kauffmann (1995); Kauffmann (1996); Kauffmann & Charlot 1998a ; Kauffmann & Charlot 1998b )
Given the great utility of PS methods, it is crucial to understand how the PS expression of structure formation relates to N-body simulations. PS predictions began to be tested against N-body results a number of years ago (Efstathiou et al. (1988); Lacey & Cole (1994)) and generally found, within the dynamic range of the simulations, to compare relatively well, although soon results were found that hinted at discrepancies between PS and simulations (Gelb & Bertschinger (1994)). As the range in masses accessible in the N-body method has increased, a number of other studies have been performed (Somerville et al. (2000); Tormen (1998); Sheth, Mo, & Tormen (1999); Borgani et al. (1999); Governato et al. (1999); Grammann & Suhhonenko (1999); Bryan & Norman (1998); Gross et al. (1998)). Somerville et al. (2000) compare the extended Press-Schechter model and find that, at high masses, discrepancies as high as 50% can exist between SAM and N-body predictions of progenitor numbers. They find discrepancies in such quantities as the PS mass function and conditional mass function, as well as SAM number-mass distributions. On the other hand, they find that a number of more relative quantities, such as the ratio of progenitor masses, are reproduced fairly well. Tormen (1998) also finds that EPS underpredicts the number of high-mass progenitors of cluster members at high redshift and concludes that caution should be exercised when modeling clustering properties of halo progenitors.
PS provides a representation of halos over a cosmologically representative range of environment. How applicable is the PS mass function in environmental extremes? Lemson & Kauffmann (1999) use N-body simulations to examine a number of halo parameters as function of environment. They find environmental dependence only for the mass distribution of halos, with angular momentum, concentration parameter, and formation time remaining unaffected. Consequently, one would expect PS to maintain as accurate a representation of environmental extremes as it does more representative regimes. In this work, I compare PS output to N-body simulations in both a representative volume and an extremely underdense region.
One quantity which cannot be derived from semi-analytic modeling is the angular momentum properties of a halo. Workers who study structure within galactic halos may derive halo mass and interaction histories from PS based methods, but spin parameters are typically assigned from the prescriptions found by previous N-body work. For example, Dalcanton, Spergel, & Summers (1997) use PS in combination with the spin parameter distribution of Warren et al. (1992) in order to study the distributions of surface brightness and scale length in disk galaxies. Van den Bosch (1998) uses this same distribution to examine the formation of disks and bulges within dark matter halos. Mo, Mao & White (1998) examine the formation of disk galaxies by incorporating an a priori distribution of dark matter halo angular momentum. Any SAM investigation of galaxy evolution and interaction (e.g. Somerville & Primack (1999)) must also assign spin parameters in this manner. When a halo experiences a major merging event, a new spin parameter can be assigned. However, should it be taken from the same general distribution found in N-body simulations or are the angular momenta of merger remnants systematically different from the quiescent population?
This paper is divided into two major thrusts. In the first, I examine the distribution of halo spin parameters in the overall halo population as well as merger remnants. I do this for several cosmologies, and for a large cosmologically representative volume (“uniform volume”) as well as a large, underdense void. In the second thrust, I compare semi-analytic halo progenitor masses and halo formation times to N-body simulations. I examine this relation in 3 uniform volumes and two voids, which collectively sample 3 different cosmologies.
## 2 Methods
### 2.1 The PS formalism
The standard PS formula predicts the number density $`N(M,z)dM`$ of high contrast structure of a given mass $`M`$ at a given epoch. The counting of such structures is based on the hypothesis that the overdense regions in the initial density field filtered on the scale of interest, will eventually collapse to form high contrast structure. Then to obtain the fraction of the mass collapsed on a certain scale, it needs to single out volumes of the filtered density field, where the (linearly) evolved density contrast $`\delta `$ exceeds a nominal threshold $`\delta _c`$ depending on the epoch, which guarantees that the associated volume is collapsed into a virialized object. The PS formula states:
$$N(M,z)dM=\sqrt{\frac{2}{\pi }}\frac{\rho _0}{M}\frac{\delta _c}{\sigma ^2(M)}e^{\delta _c^2/2\sigma ^2(M)}\frac{d\sigma (M)}{dM}dM,$$
(1)
where $`\sigma ^2(M)`$ is the mass variance of the perturbation field on the mass scale $`M`$:
$$\sigma ^2(M)=\frac{1}{2\pi ^2}P(k)W^2(kR)k^2𝑑k,$$
(2)
where $`P(k)`$ is the power spectrum and $`W(kR)`$ the Fourier transform if the real-space top-hat function. The conventional way to normalize the power spectrum is $`\sigma _8`$, the rms value of the linear fluctuation in the mass distribution on scales of $`R=8h^1`$ Mpc. The field is taken to be Gaussian, and then the variance $`S=\sigma ^2(M)`$ completely describes the density perturbation field. The threshold $`\delta _c`$ refers to the collapse of homogeneous spherical volumes and is found to be $`\delta _c(z=0)=1.68`$ for $`\mathrm{\Omega }_0=1`$ models. The time variation of $`\delta _c(z)`$ is proportional to the linear growth factor $`D(t)`$ (see e.g. Peebles 1993) and contains the dependence from the specific FRW model. The overall normalization is such that at any $`z`$ all the available mass density $`\rho `$ resides in collapsed objects. For a detailed derivation and the analysis of the underlying statistics, see Bond et al (1991).
Paulo Tozzi graciously furnished the author with a copy of his code, an implementation of the Monte-Carlo method given in Lacey & Cole (1993) which is in turn based on the EPS “excursion set” formalism developed by Bond et al. (1991). In this method, the linear density field is smoothed at progressively larger mass scales and the mass of the halo containing a given particle at time $`t_1`$ is equal to the mass $`M_1`$ of the largest sphere which in which the average overdensity $`\delta _c(t_1)`$ exceeds the collapse threshold. This ensures that a halos of mass $`M`$ has not yet been subsumed by a larger mass halo. $`\delta _c(t)`$, of course, changes with time and thus at later times, our halo of mass $`M_1`$ and time $`t_1`$ may ultimately become part of a larger collapsed structure $`M_2`$ which does not reach critical density until time $`t_2`$. By examining the growth of linear modes in the same region, one can statistically construct halo growth and interaction histories. Tozzi’s method differs from that of Lacey & Cole (1993) in the specifics of how it calculates the relationship between mass ($`M_1,M_2`$, …) and collapse threshold ($`\delta _c(t_1),\delta _c(t_2)`$, …) but yields essentially the same results (Tozzi, private communication). The method is described in detail by Governato & Tozzi (2001).
### 2.2 Simulations
I present 6 simulations which are evolved from 3 cosmological models: a critical universe (SCDM) ($`\mathrm{\Omega }_0=1`$, $`hH_0/100`$ km s<sup>-1</sup> Mpc$`{}_{}{}^{1}=0.5`$, $`\sigma _8=0.7`$), an open (OCDM) universe ($`\mathrm{\Omega }_0=0.3`$, $`h=0.75`$, $`\sigma _8=1`$) and a tilted critical model (TCDM) with h$`=0.5`$, $`\sigma _8=0.6`$, a primordial index n$`=0.8`$ and a gamma factor $`\mathrm{\Gamma }=`$ 0.37. The transfer function of Bardeen et al. (1986) was used in all cases. The simulated volume was 100 Mpc on a side (h already included) in all six runs. The parameters of the TCDM model have been chosen to satisfy both the cluster abundance and the COBE normalization at very large scales (Cavaliere, Menci, & Tozzi (1998)). This model differs from the SCDM in having less power and a steeper power spectrum at scales under 8h<sup>-1</sup> Mpc in comparison with the other two models. Each simulation was performed using PKDGRAV (Stadel & Quinn, in preparation) a parallel N–body treecode supporting periodic boundary conditions. All runs were started at redshifts sufficiently high to ensure that the absolute maximum density contrast $`|\delta |<1`$. All simulations were evolved using between 500 and 1000 major time steps in which the forces on all particles were calculated. Each major time step is divided into a number of smaller substeps, the exact number of which is adaptively chosen by PKDGRAV such that all particles are moved on steps consonant with their dynamical times.
One simulation in each cosmology (the “uniform volume”) was run beginning from a spatially uniform grid of $`144^3`$ ($``$ 3 million) equal–mass particles with spline softening set to 60 kpc, allowing us to resolve individual halos with present-day circular velocities $`V_c`$ as low as 130 km s<sup>-1</sup> with 100 particles in OCDM and 170 km s<sup>-1</sup> with 100 particles in SCDM and TCDM. The particle masses were $`1.57\times 10^{10}`$ M in OCDM and $`2.32\times 10^{10}`$ M in SCDM and TCDM, while the opening angles of the SCDM and OCDM simulations were $`\theta =0.5`$ for redshifts $`z>2`$ and $`\theta =0.7`$ thereafter. TCDM was run with $`\theta =0.5`$ at all redshifts.
The remaining three simulations, one in each cosmology, were of a large void approximately 40 Mpc in diameter, where I defined the void region as the largest sphere that could be drawn in the simulation volume with mean overdensity $`\overline{\delta }=0.9`$. The radii of the resultant spheres were 18.5 Mpc, 17 Mpc, and 13.5 Mpc for SCDM, OCDM and TCDM cosmologies respectively. The void runs were performed using the hierarchical grids technique in which the void region is represented with large numbers of low-mass particles, while the surrounding environment is composed of small numbers of high-mass particles. This allows much greater mass resolution to be achieved while ensuring that cosmological context is maintained. I used approximately 5 million particles in each void simulation, with an effective resolution of $`432^3`$ across the 100 Mpc box. $`\theta `$ was set to 0.6 for $`z>2`$ and 0.8 thereafter. Particles were spline-softened to 6.67 kpc, being $`8.6\times 10^8`$ M in mass in critical models and $`5.8\times 10^8`$ in OCDM. Thus, structure in the voids was resolved in mass at a factor of 27 superior to the uniform volume runs. For convenience, I shall adopt the notation of SCDMu, OCDMu and TCDMu for the uniform volumes and SCDMv, OCDMv and TCDMv for the voids.
Throughout the paper, I identify halos using two primary methods. The first is the classic friends-of-friends (FOF) method (Davis et al. (1985)) using a linking length that corresponds to the mean interparticle separation at the density contour defining the virial radius of an isothermal sphere: $`b=(n\rho _{vir}(z)/\overline{\rho }(z)/3.0)^{1/3}`$ where $`n`$ is the particle number density and $`\rho _{vir}(z)`$ is the virial density which depends on cosmology and is given in Kitayama & Suto (1996). The second is a variant of the spherical overdensity (SO) technique (Lacey & Cole (1994)) where I determine the most bound particle in a FOF group and find the radius from that particle at which the enclosed density crosses the virial threshold $`\rho _{vir}(z)`$.
### 2.3 Halo Mass Histories
In the simplest test between PS and N–body simulations, one assumes as input the spectral parameter $`P(k)`$ and $`\sigma _8`$, and compares the mass distribution at a given epoch. This essentially decides whether the PS formula correctly states the shapes and the normalization of the mass distribution of collapsed objects. Work has been done comparing the PS mass function to N-body simulations (Lacey & Cole (1994), Borgani et al. (1999), Governato et al. (1999), Bryan & Norman (1998), Gross et al. (1998), Tormen (1998)), and I do not use this method in my analysis.
I focus on another test which examines the evolution of the linear density field and compares it to the hierarchical clustering paradigm (cf. Kaiser (1986)) as realized in the N-body method. The parameter $`\delta _c`$ (cf. equation (1)) is the value of linear overdensity above which a perturbation is considered to be collapsed (evolved to the present day). Thus, $`\delta _c`$ determines the mapping between the linear perturbation field and the mass of a halo. By demanding that the $`z=0`$ PS mass function matches the z=0 mass function of the simulation, one is effectively anchoring $`\delta _c`$. This corresponds to the method used by Tormen (1998) in examining the conditional mass function of the calibrated EPS method. Thus there are no remaining free parameters to the evolution of the density perturbation field unless one invokes a $`\delta _c(z)`$ which varies non-linearly with redshift such as the one examined by Governato et al (1999).
In the uniform volume simulations, I identified all halos at z=0 of at least mass $`M_{res}=2.318\times 10^{12}`$ M, which corresponds to 100 particles in the flat models and roughly 150 in OCDM. With a mass-based cutoff, it is more straightforward to compare time evolution of the different cosmologies. Having at least 100 particles also insures that the halos are substantially larger than their softening lengths, guarding against the influence of numerical effects on global halo properties. I then “traced” each identified halo backwards in time, following the mass of the most massive progenitor as a function of redshift. In the case of the void runs, I identified all groups with at least 50 particles and traced them in a manner identical to the uniform volume counterparts. Unfortunately, the TCDM cosmology forms structure extremely late, and the void had only 12 halos in it greater than this mass cutoff. I therefore drop TCDMv from the following analysis.
I now wish to construct a Monte-Carlo realization which most accurately compares against the simulated results. Comparing individual halos is fruitless, given the extreme variation of mass histories even between halos of the same mass. Instead I wish to characterize the mean mass history of the sample, i.e. the average mass of a halo in the sample as a function of redshift. This was easily done by drawing the mass distribution of the halos to be modeled from the simulations themselves. Specifically, for each $`M>M_{res}`$ halo in a simulation, 100 Monte-Carlo realizations of the mass history of that halo were constructed. Thus, I can directly compare the aggregate mass history of the sample of simulated halos to the collective mass history of the Monte-Carlo realizations.
Errors in both simulated and Monte-Carlo samples were estimated using the bootstrap method. For each sample, 1000 identically-sized subsamples of halos were drawn from the z=0 distribution. Then, for each subsample, the mean mass history was calculated and used in the bootstrap estimate.
### 2.4 Halo Angular Momentum
For SCDMu and OCDMu uniform volumes and for the OCDMv void, I determine the distribution of halo spin parameters $`\lambda =LE^{1/2}/GM^{5/2}`$ for recent merger remnants versus non-remnants, where $`E`$ is the total energy of the halo, $`L`$ the angular momentum, and $`M`$ the mass. I select halos for this analysis based on particle number in the $`z=0`$ FOF group: 100 particles for SCDMu and OCDMu and 50 particles for OCDMv. The OCDMu particle cutoff is lower than in the section 2.3 for more favorable number statistics. To identify mergers, I use the groups determined by SKID, a halo-finding algorithm based on local density maxima. The SKID algorithm is similar to the DENMAX scheme (Gelb & Bertschinger, 1994) as it groups particles by moving them along the density gradient to the local density maximum. The density field and density gradient is defined everywhere by smoothing each particle with a cubic spline weighting function of size determined by the distance encompassing the nearest 32 neighbors. At a given redshift, only particles with local densities greater than one-third of 177 times the critical density are “skidded” to the local density maximum. This threshold corresponds roughly to the local density at the virial radius. The final step of the process is to remove all particles that are not gravitationally bound to their parent halo. SKID was originally designed to find high contrast density structures within larger halos (see Ghigna et al. (1998) for a more complete description). For halos with no resolved internal substructure (i.e. no halos within halos) it gives results very similar to FOF. However, SKID does not suffer of the well known pathology of FOF of linking together close binary systems (Governato et al. (1997)). Instead, it only combines two “clumps” of substructure when they are mutually gravitationally bound and thus is well suited for the task of identifying binary systems in the process of merging.
I denote a halo as a merger remnant if, at some time during $`0<z<z_m`$ it was classified as a single SKID group in one output, but two separate groups with a mass ratio $`3:1`$ in the preceding output. This method is essentially independent of output spacing except in the extreme case where the halo mass does grows quiescently by order its own mass between outputs. $`z_m`$ was chosen to be as close to $`z=0.5`$ as the outputs would allow. In SCDMu this is $`z_m=0.4`$, and in both OCDM runs $`z_m=0.5`$.
I use SKID to identify halos before the merging event and spherical overdensity to identify the descendant halo at the present time. The SO method has the advantage of culling out pathologically shaped groups which depart substantially from the isothermal sphere model. The most common example is a “dumbbell” halo, which consists of two spherical halos that FOF links together by a thin bridge of particles spaced close to $`\rho _{vir}(z)/3`$. The total spin parameter of such a configuration is irrelevant to this study which is concerned only with cases where the angular momentum of a galaxy is linked to its parent halo. The spin parameter $`\lambda `$ is calculated by considering the contributions of all particles within the SO virial radius. If I use the FOF group membership rather than the SO, the angular momenta for non-pathological cases (about the FOF center of mass) is generally the same. Strangely-shaped groups are in the minority and hence even with their addition in the FOF sample, the the results remain qualitatively similar.
## 3 Results
### 3.1 Angular Momentum of Merger Remnants
The halo spin parameter distribution in simulations is found to be well approximated by the lognormal function
$$p(\lambda )\mathrm{d}\lambda =\frac{1}{\sigma _\lambda \sqrt{2\pi }}\mathrm{exp}\left(\frac{\mathrm{ln}^2(\lambda /\overline{\lambda })}{2\sigma _\lambda ^2}\right)\frac{\mathrm{d}\lambda }{\lambda }$$
(3)
(van den Bosch (1998); Barnes & Efstathiou (1987); Ryden (1998); Cole & Lacey (1996); Warren et al. (1992)). Figure 1 shows $`p(\lambda )\mathrm{d}\lambda `$ of simulated halos above $`M_{res}`$ at z=0. The solid histograms denote the merger-remnant population and the dotted histograms are halos which have not experienced a major merger since $`z=z_m0.5`$. The smooth curves are fits of the form in equation 3 to the simulated distribution. Parameters for the fits are given in Table 1. Since the OCDM void model contains so few halos, it is easier to see the details of the distributions by viewing them in log space in Figure 1d, where the fitting function is a Gaussian. It should be noted that Lemson & Kauffmann (1999) find no environmental dependence of spin parameter on environment. Hence, any differences in OCDMu vs. OCDMv arise from the fact that the simulations sample substantially different ranges in mass and not from environmental effects.
N-body results are often used to provide an angular momentum distributions for galaxy halos in semi-analytic models (Dalcanton et al. 1997; Mo et al. 1998; van den Bosch (1998)). Warren et al. (1992) find that the general distribution of halos peaks around $`\overline{\lambda }=0.05`$ with $`\sigma _\lambda =0.7`$, although later N-body researchers find distributions more consistent with $`0.04<\overline{\lambda }<0.05`$ and $`\sigma _\lambda 0.5`$ (Cole & Lacey (1996); Mo, Mao, & White (1998); see also Steinmetz & Bartelmann (1995); Catelan & Theuns (1996)). I find $`\sigma _\lambda `$ of all halos (in the uniform volumes) lower than 0.6 with the merger remnant population tending lower still, while the peak of the all three simulations presented is $`\overline{\lambda }0.035`$. Hence, my results for the global halo population are in concordance with previous findings.
I find that in all cases the distribution of spin parameters of merger remnants is greater than non-merger remnants. The K-S probabilities that the two sets were drawn from the same distribution are given as $`P_{KS}`$ in Table 1 and are quite low. The K-S test is most certain in SCDMv, given the large number of major mergers. Even though both OCDM runs contain fewer resolved merger remnants, the K-S test differentiates the merger and non-merger remnant populations at substantial confidence. To determine the robustness of the dissimilar means $`\overline{\lambda }`$ of distributions, I use the Wilcoxon test (cf. Lupton 1993). The value $`P_{Wil}`$ given in Table 1 is the probability that the difference between the mean of the merger population the mean of the non-merger population could have occurred by chance. In all cases, the probability is quite high that $`\overline{\lambda }`$ of the two distributions are indeed significantly different. Consequently, spin parameters of halos which have experienced a recent major merger are distributed differently than halos which have not undergone a recent significant interaction.
Angular momentum has been found to be relatively insensitive to cosmology and collapse anisotropy for halos formed by less violent collapse, with the spread in $`\overline{\lambda }`$ arising mainly from tidal effects (Huss, Jain, & Steinmetz (1999)). This view is supported by Nagashima & Gouda (1998) who, using semi-analytics (the “merging-cell model”), find that the orbital angular momentum of major mergers is not important for the remnant, and that major merging activity as a whole has little effect on the halo angular momentum distribution. Given the large spread in $`\overline{\lambda }`$ in my results, it seems tempting to conclude that on a halo-by-halo basis, tidal effects are more important than merger history in determining the ultimate angular momentum of a halo. Merging may be looked at simply as adding a small upward bias. However, one should note that the dispersion $`\sigma _\lambda `$ in the merger remnants is typically lower than the non-merger remnants. This can imply one of two possibilities: A) the angular momentum of mergers is largely independent of tidal torques, depending more on the mergers themselves or B) merging adds selects for specific environments. In B), by considering mergers, we are simply biasing ourselves toward examining halos which are in a region that promotes merging, nurturing halos with a narrower range in $`\lambda `$ and a higher mean. In A), it is a combination of the orbital angular momentum of the merger and the spin-alignment of the halos that determines a remnant’s resultant angular momentum. However, if the spin-alignment were dominant, one would expect the dispersion in $`\lambda `$ to be larger than the non-merger distribution, as two average halos can either add their spins or have them cancel completely. Consequently in case A), the orbital angular momentum of merging halos is the more likely contributor to the final remnant momentum. To summarize, tidal effects can dominate the merger remnant angular momentum distribution only if the lower dispersion of the merger remnant population can be explained. A likelier explanation is that by examining mergers, one is merely selecting a specific environment and hence only a subset of the possible tidal torques. The equally likely alternative is that the orbital angular momentum of the pre-merger halos dominates the spin parameters of merger remnants.
How relevant is the contribution of mergers to the overall spin parameter distribution? Figure 2 plots the fraction of halos in my sample with spin parameters $`\lambda `$ that are major merger remnants. The result is strongly dependent on cosmology, given that in open universes, merging activity at low redshift is substantially less likely than in $`\mathrm{\Omega }_0=1`$ models (Governato et al. (1998)). Consequently, the global effect of merging on halo angular momentum is simply determined by the rate of major mergers. This information is, however, of marginal utility since galaxy history or properties are rarely used to infer angular momentum. In general, the distribution $`p(\lambda )d\lambda `$ is employed in an a priori manner. In this case, researchers may wish to adopt a different distribution for galaxies which are merger remnants and those which are not.
The overall implication of these results is that merging activity is correlated with higher angular momentum. Orbital angular momentum of interaction halos would seem to be the cause most consistent my data. However, on galaxy scales, the angular momentum of ellipticals is generally much lower than spirals (Nagashima & Gouda (1998)). This is troublesome if ellipticals are to be considered merger remnants and spirals the results of long periods of quiescence. The quiescent accretion required to form a higher-$`\lambda `$ spiral disk would appear to take place in halos with systematically lower spin parameters. Differences in angular momenta between galaxy morphologies may be more due to formation epoch (Huss, Jain, & Steinmetz (1999)) than merger history.
### 3.2 Halo Mass Histories and Formation Times
Here I compare the Monte-Carlo approximation to simulated data in differing cosmologies and environment. Figure 3 shows the mean mass histories of halos selected in accordance with section 2.3 for the uniform volumes in all three cosmologies, and the voids in SCDM and OCDM. This is essentially a measurement of the conditional mass function examined by other authors (Tormen (1998); Somerville et al. (2000); Bower (1991)). It has been established in previous studies that the Press-Schechter model underestimates the number of halos relative to simulations at mass scales which lessen with increasing redshift (Somerville et al. (2000), Gross et al. (1998), Tormen (1998)). Somerville et al. (2000) find that for their tilted CDM model, EPS underpredicts the masses of the most massive progenitor at redshifts $`z>0.5`$. In the flat uniform volumes, I also find that EPS systematically predicts lower-mass principle progenitors than simulations. The discrepancy in the average progenitor mass is as high as 20% in SCDM and TCDM. Interestingly enough, things are different in the OCDMu open model in contrast to the findings of Somerville et al. (2000). In this case, the semi-analytic predictions appear to be roughly in line with the simulations. Also, EPS appears to reproduce the mean mass histories of both voids reasonably well, trailing outside the 1-$`\sigma `$ dispersion bars for SCDMv only at very high redshift. It should be noted that the void simulations sample halos of a much lower mass range resulting in higher typical formation redshifts and making a direct comparison between uniform volume and void runs somewhat hazardous. The difference in mass history between SCDMu vs. SCDMv and OCDMu vs. OCDMv is most likely caused by the difference in mass range. The important comparison is between Monte-Carlo and simulated results in each regime. The masses of the halos in the void runs are somewhat lower than in the uniform volumes and thus farther from $`M_{}`$ of the mass function. Since the shape of the mass function changes little in this low-mass regime, it may be possible for $`10^{11}M_{}`$ halos to match while $`10^{13}M_{}`$ halos do not. Somerville et al. (2000) also find that EPS progenitor masses match better for halos of mass $`10^{12}M`$ than for halos $`M>10^{13}M_{}`$. Hence, these results are reasonably consistent with their findings. Given the small number of halos in the void, however, these results are not conclusive and should be reexamined with a larger sample size.
The difference in results can be examined more quantitatively by comparing halo formation times. A halo is said to have formed when the mass of the most massive progenitor is 50% of the z=0 mass. Given the discreteness of the simulation outputs, the formation time of simulated halos can only be bounded between adjacent outputs where the halo transitions the 50% formation threshold. Figure 4 plots the formation times of halos in simulations vs. their respective Monte-Carlo realizations. The points from the simulations (connected by lines) denote the number of halos formed in a bin $`\mathrm{\Delta }z`$ where the bin boundaries correspond to simulation output times. In some cases multiple simulation outputs are combined into a single bin in order to increase statistics. In order to make a meaningful comparison, the formation time for each Monte-Carlo realization is binned in exactly the same manner as the simulation output. The bin boundaries are given by the horizontal bars in the Monte-Carlo data in Figure 4. To determine confidence limits, I bin 1000 bootstrap samples of the Monte-Carlo realization and show the regime in which 90% of the realizations fall as vertical errorbars. Consequently, for the simulation to match the EPS prediction, 90% of the simulation points should fall within the Monte-Carlo errorbars. As we expect from the mass history results, the EPS results are clearly different from SCDMu and TCDMu simulations at the 90% confidence level. The median formation time is a good measure for the “error” in formation time between simulated and EPS realizations. In SCDMu, the median simulated formation time in units of the $`z=0`$ age of an SCDM universe is approximately 0.20 while the Monte-Carlo realizations find a median formation time of 0.38. For OCDMu, 4 out of 10 simulated points are outside the errorbars, meaning that EPS also fails to match at the 90% confidence level. In the case of the void runs, the sampling errors are too large to rule out either Monte-Carlo realization, consistent with the interpretation from Figure 3. Consequently, the EPS appears to diverge notably from simulated realizations in flat models, as well as maintain a measurable difference in open models. The difference in the results between open and flat-matter cosmologies is most likely related to the growth of fluctuations in each. Since structure grows faster in $`\mathrm{\Omega }_0=1`$ models, the differences between EPS and simulations is most likely exaggerated with respect to low-$`\mathrm{\Omega }_0`$.
Note again that these results allow no freedom of choice of the critical density parameter $`\delta _c`$ as the PS and simulated mass functions are constrained to match at redshift $`z=0`$. Governato et al. (1999) provide evidence that a better fit can be attained by evolving $`\delta _c`$ in a nonlinear fashion. Hence, one is effectively renormalizing the mass function at every epoch, thus changing the mapping between perturbation spectrum and mass. If I use their fitting formula $`\delta _c(z)=1.686\left[(0.7/\sigma _8)(1+z)\right]^{0.125}`$ (valid for $`\mathrm{\Omega }_0=1`$) for SCDMu, I find that this does indeed change the Monte-Carlo mass history of Figure 3b substantially. While the Monte-Carlo history is lower for small redshifts, it crosses the simulated curve and eventually diverges $`2\sigma `$ above the simulated histories for $`z>1.4`$. Governato et al. construct this relation from cluster-mass halos, well over an order of magnitude more massive than this study, so it is not surprising that using this fit does not induce EPS to match my simulated data. Still, the simulated results are bracketed with static $`\delta _c`$ producing too rapid halo growth and redshift-variant $`\delta _c`$ producing halos which grow slower than in N-body simulations. It may be possible to fine-tune a prescription for $`\delta _c(z)`$ to simulated data. As with any fine tuning of semi-analytic parameters, however, it is essential to understand the physical mechanisms underlying such parameters lest their predictive power be compromised. A study of such scope is best left to a future paper.
## 4 Conclusions
I present results from numerical N-body simulations regarding the effect of merging events on the angular momentum distribution of galactic halos as well as a comparison of halo growth in Monte-Carlo vs. N-body methods. In all three models tested and both in a uniform sample and in an underdense region, the distribution of halo spin parameters of merger remnants is greater than non-merger remnants. The spin parameters of the global halo population are consistent with other recent results Cole & Lacey (1996); Mo, Mao, & White (1998); Steinmetz & Bartelmann (1995); Catelan & Theuns (1996)). However, the K-S probabilities of merger and non-merger samples being from the same sample are quite low, and the mean spin parameters of the two samples $`\overline{\lambda }`$ are significantly different. The actual contribution of merging on the global spin parameter distribution depends on the fraction of halos that are merger-remnants and hence is greatly affected by cosmology. The overall effect of merging is to increase the mean and decrease the dispersion of the spin parameter distribution. I offer the tentative conclusion that this is consistent with the orbital angular momentum of the merging halos playing a major role in establishing the remnant’s resultant net angular momentum. The causes and time-evolution of this behavior should be investigated in greater detail in future studies using simulations of superior dynamic range. It is also possible that angular momentum distributions could vary as a function of redshift, but I leave this for a future paper.
I trace the most massive progenitor of halos of mass $`M_i10^{13}M_{}`$ at redshift $`z=0`$ back to $`z=5`$ for “uniform volumes” in all three cosmologies as well as mass $`M_i10^{11}M_{}`$ halos for large voids in the $`\mathrm{\Omega }_0=1`$ and $`\mathrm{\Omega }_0=0.3`$ cosmologies. I find that in the uniform volumes, the Press-Schechter based method reproduces the general trend of the simulated mass histories relatively well, although in $`\mathrm{\Omega }_0=1`$ models it underpredicts the average mass history of halos by roughly 20%. This leads to a further discrepancy in the estimation of halo formation times. For example, the median formation time of halos in a flat universe (SCDMu) is 20% the age of the Universe according to simulations, but 38% the age of the Universe in Monte-Carlo realizations. Because the PS $`z=0`$ mass function is constrained to match the simulated halo mass function, $`\delta _c`$ is no longer a free parameter. It may be possible fine-tune PS to match simulated results by the use of a $`\delta _c`$ which evolves non-linearly with redshift, although such a study is beyond the scope of this paper.
My results are consistent with work by Somerville et al. (2000) and Tormen (1998). Somerville et al. find that EPS reproduces relative properties of the progenitor halo distribution, e.g. their mass ratios, quite well. However, the absolute progenitor masses and the overall conditional mass function exhibits discrepancies of up to 50%. Tormen concludes that the EPS formulation is inadequate to describe clustering in a constrained environment and may possibly be inadequate in a general context. I find that the latter is indeed true, as the conditional halo mass function in a cosmologically representative sample does not agree with simulated data. Consequently, care should be taken when comparing the raw masses of halo progenitors, or their mass relative to their eventual “children” of the present day. Formation times of halos from the EPS conditional mass function are systematically biased. Furthermore, care should be taken when examining the redshift evolution of any absolute quantity which depends on the evolution of mass, such luminosity or color. Halo properties should be compared in a relative sense at each redshift and not in an absolute sense with respect to their parents or children at other times.
Recently, alternative expressions for the PS mass function have been proposed and investigated. Sheth & Tormen (1999) modify the Press-Schechter formalism, producing a much better match to the mass function, especially at the low-mass end, of simulations. Mo, Sheth, & Tormen (1999) have shown that significant improvements to the standard PS approach can be made by modeling the formation of bound structures as an ellipsoidal rather than a spherical collapse. Jenkins et al. (2000) derive an empirical mass function by fitting the results numerous N-body simulations. My results further demonstrate the need for using these alternative approaches when generating Monte-Carlo mass histories.
The author gratefully acknowledges the generosity of Paolo Tozzi for providing his semi-analytic code for this study. The author also thanks Fabio Governato, without whose guidance and support this paper would have been impossible, and Tom Quinn who never ceases to provide useful advice and feedback. Furthermore, it is a pleasure to acknowledge frequent and useful discussions with Frank van den Bosch and Julianne Dalcanton. The author was supported by NASA Grant NGT5-50078 for the duration of this work. The simulations were performed on the Cray T3D/E at the Arctic Region Supercomputing Center.
|
warning/0006/hep-th0006065.html
|
ar5iv
|
text
|
# References
Non-Abelian (2,0)-Super-Yang-Mills
Coupled to Nonlinear $`\sigma `$-Models
M. S. Góes-Negrão <sup>1</sup><sup>1</sup>1 negrao@gft.ucp.br and negrao@cbpf.br, J. A. Helayël-Neto <sup>2</sup><sup>2</sup>2helayel@gft.ucp.br and helayel@cbpf.br and M. R. Negrão <sup>3</sup><sup>3</sup>3 guida@gft.ucp.br and guida@cbpf.br
Universidade Católica de Petrópolis(UCP-GFT)
Centro Brasileiro de Pesquisas Físicas(CBPF-CCP)
## Abstract
Following a previous work on Abelian (2,0)-gauge theories, one reassesses here the task of coupling (2,0) relaxed Yang-Mills super-potentials to a (2,0)-nonlinear $`\sigma `$-model, by gauging the isotropy or the isometry group of the latter. One pays special attention to the extra “chiral-like” component-field gauge potential that comes out from the relaxation of constraints.
PACS numbers: 11.10.Lm, 11.15.-q and 11.30.Pb
In a previous paper , one has investigated the dynamics and the couplings of a pair of Abelian vector potentials of a class of $`(2,0)`$-gauge super multiplets (-) whose symmetry lies on a single U(1) group. Since a number of interesting features came out, it was a natural question to ask how these fields would behave if the non-Abelian version of the theory was to be considered.
We can see that some subtle changes indeed occur. As we wish to make a full comparison between the two aspects(Abelian and non-Abelian) of the same sort of theory, all the general set up of the original formulation was kept. The coordinates we choose to parametrise the $`(2,0)`$-superspace are given by:
$`z^A(x^{++},x^{};\theta ,\overline{\theta }),`$ (1)
where $`x^{++}`$, $`x^{}`$ denote the usual light-cone variables, whereas $`\theta `$, $`\overline{\theta }`$ stand for complex right-handed Weyl spinors. The supersymmetry covariant derivatives are taken as:
$`D_+_\theta +i\overline{\theta }_{++}`$ (2)
and
$`\overline{D}_+_{\overline{\theta }}+i\theta _{++},`$ (3)
where $`_{++}`$ (or $`_{}`$) represents the derivative with respect to the space-time coordinate $`x^{++}`$ (or $`x^{}`$). They fulfill the algebra:
$`D_+^2=\overline{D}_+^2=0\{D_+,\overline{D}_+\}=2i_{++}.`$ (4)
With these definitions for $`D`$ and $`\overline{D}`$, one can check that:
$`e^{i\theta \overline{\theta }_+}D_+e^{i\theta \overline{\theta }_+}=_\theta ,`$ (5)
$`e^{i\theta \overline{\theta }_+}\overline{D}_+e^{i\theta \overline{\theta }_+}=_{\overline{\theta }}.`$ (6)
The fundamental non-Abelian matter superfields are the scalar and left-handed spinor superfields, both subject to the chirality constraint; their respective component-field expressions are given by:
$`\mathrm{\Phi }^i(x;\theta ,\overline{\theta })`$ $`=`$ $`e^{i\theta \overline{\theta }_{++}}(\varphi ^i+\theta \lambda ^i),`$
$`\mathrm{\Psi }^I(x;\theta ,\overline{\theta })`$ $`=`$ $`e^{i\theta \overline{\theta }_{++}}(\psi ^I+\theta \sigma ^I);`$ (7)
the fields $`\varphi ^i`$ and $`\sigma ^I`$ are scalars, whereas $`\lambda ^i`$ and $`\psi ^I`$ stand respectively for right- and left-handed Weyl spinors. The indices $`i`$ and $`I`$ label the representations where the corresponding matter fields are set to transform under the Yang-Mills group.
We present below the gauge transformations for both $`\mathrm{\Phi }`$ and $`\mathrm{\Psi }`$, assuming that we are dealing with a compact and simple gauge group, $`𝒢`$, with generators, $`G_a`$, that fulfill the algebra $`[G_a,G_b]`$=$`if_{abc}G_c`$. The transformations read as below:
$`\mathrm{\Phi }^i=R(\mathrm{\Lambda })_j^i\mathrm{\Phi }^j,\mathrm{\Psi }^I=S(\mathrm{\Lambda })_J^I\mathrm{\Psi }^J,`$ (8)
where $`R`$ and $`S`$ are matrices that respectively represent a gauge group element in the representations under which $`\mathrm{\Phi }`$ and $`\mathrm{\Psi }`$ transform. Taking into account the chiral constraints on $`\mathrm{\Phi }`$ and $`\mathrm{\Psi }`$, and bearing in mind the exponential representation for $`R`$ and $`S`$ in terms of the group generators, we find that the gauge parameter superfields, $`\mathrm{\Lambda }^a`$, must satisfy the same sort of constraint, namely, they are chiral scalar superfields. They can therefore be expanded as follows:
$`\mathrm{\Lambda }^a(x;\theta ,\overline{\theta })`$ $`=`$ $`e^{i\theta \overline{\theta }_{++}}(\alpha ^a+\theta \beta ^a),`$ (9)
where $`\alpha ^a`$ are scalars and $`\beta ^a`$ are right-handed spinors.
The kinetic action for $`\mathrm{\Phi }^i`$ and $`\mathrm{\Psi }^I`$ can be made invariant under the local transformations (8) by minimally coupling the gauge potential superfields, $`\mathrm{\Gamma }_{}^a(x;\theta ,\overline{\theta })`$ and $`V^a(x;\theta ,\overline{\theta })`$, according to the minimal coupling prescriptions:
$`S_{inv}={\displaystyle d^2x𝑑\theta 𝑑\overline{\theta }\{i[\overline{\mathrm{\Phi }}e^{hV}(_{}\mathrm{\Phi })(\overline{}_{}\overline{\mathrm{\Phi }})e^{hV}\mathrm{\Phi }]+\overline{\mathrm{\Psi }}e^{hV}\mathrm{\Psi }\}},`$ (10)
where the gauge-covariant derivatives are defined in the sequel.
The Yang-Mills supermultiplets are introduced by means of the gauge-covariant derivatives which, according to the discussion of ref. , can be expressed as below:
$`_+`$ $``$ $`D_++\mathrm{\Gamma }_+,\overline{}_+\overline{D}_+,`$ (11)
$`_{++}`$ $``$ $`_{++}+\mathrm{\Gamma }_{++}and_{}_{}ig\mathrm{\Gamma }_{},`$ (12)
with the gauge superconnections $`\mathrm{\Gamma }_+`$, $`\mathrm{\Gamma }_{++}`$ and $`\mathrm{\Gamma }_{}`$ being all Lie-algebra-valued. Note that $`\mathrm{\Gamma }_{++}`$ does not enter the Lagrangian density of eq.(10). The gauge couplings, $`g`$ and $`h`$, can in principle be taken different; nevertheless, this would not mean that we are gauging two independent symmetries. There is a single simple gauge group, $`𝒢`$, with just one gauge-superfield parameter, $`\mathrm{\Lambda }`$. It is the particular form of the $`(2,0)`$-minimal coupling (realised by the exponentiation of $`V`$ and the connection present in $`_{}`$) that opens up the freedom to associate different coupling parameters to the gauge superfields $`V`$ and $`\mathrm{\Gamma }_{}`$. The superpotentials $`\mathrm{\Gamma }_+`$ and $`\mathrm{\Gamma }_{++}`$ can be both expressed in terms of the real scalar superfield, $`V(x;\theta ,\overline{\theta })`$, according to :
$`\mathrm{\Gamma }_+=e^{gV}(D_+e^{gV})`$ (13)
and
$`\mathrm{\Gamma }_{++}={\displaystyle \frac{i}{2}}\overline{D}_+[e^{gV}(D_+e^{gV})].`$ (14)
To establish contact with a component-field formulation and to actually identify the presence of an additional gauge potential, we write down the $`\theta `$-expansions for $`V^a`$ and $`\mathrm{\Gamma }_{}^a`$:
$`V^a(x;\theta ,\overline{\theta })=C^a+\theta \xi ^a\overline{\theta }\overline{\xi }^a+\theta \overline{\theta }v_{++}^a`$ (15)
and
$`\mathrm{\Gamma }_{}^a(x;\theta ,\overline{\theta })`$ $`=`$ $`{\displaystyle \frac{1}{2}}(A_{}^a+iB_{}^a)+i\theta (\rho ^a+i\eta ^a)`$ (16)
$`+`$ $`i\overline{\theta }(\chi ^a+i\omega ^a)+{\displaystyle \frac{1}{2}}\theta \overline{\theta }(M^a+iN^a).`$
$`A_{}^a`$, $`B_{}^a`$ and $`v_{++}^a`$ are the light-cone components of the gauge potential fields; $`\rho ^a,\eta ^a,\chi ^a`$ and $`\omega ^a`$ are left-handed Majorana spinors; $`M^a,N^a`$ and $`C^a`$ are real scalars and $`\xi ^a`$ is a complex right-handed spinor.
The infinitesimal gauge transfomations for $`V^a`$ and $`\mathrm{\Gamma }^a`$ are given by
$`\delta V^a={\displaystyle \frac{i}{h}}(\overline{\mathrm{\Lambda }}\mathrm{\Lambda })^a{\displaystyle \frac{1}{2}}f^{abc}(\overline{\mathrm{\Lambda }}+\mathrm{\Lambda })_bV_c`$ (17)
and
$`\delta \mathrm{\Gamma }_{}^a=f^{abc}\mathrm{\Lambda }_b\mathrm{\Gamma }_c+{\displaystyle \frac{1}{g}}_{}\mathrm{\Lambda }_a.`$ (18)
No derivative acts on the $`\mathrm{\Lambda }^a`$’s in eq.(17); this suggests the possibility of choosing a Wess-Zumino gauge for $`V^a`$. If such a choice is adopted, it can be shown that the gauge transformations of the $`\theta `$-component fields above read as shown:
$`\delta C`$ $`=`$ $`{\displaystyle \frac{2}{h}}\mathrm{}m\alpha ,`$
$`\delta \xi `$ $`=`$ $`{\displaystyle \frac{i}{h}}\beta ,`$
$`\delta v_{++}^a`$ $`=`$ $`{\displaystyle \frac{2}{h}}_{++}\alpha ^af_{abc}\alpha ^bv_{++}^c,`$
$`\delta A_{}^a`$ $`=`$ $`{\displaystyle \frac{2}{g}}_{}\alpha ^af_{abc}\alpha ^bA_{}^c,,`$
$`\delta B_{}^a`$ $`=`$ $`f_{abc}\alpha ^bB_{}^c,`$
$`\delta \eta ^a`$ $`=`$ $`f_{abc}\alpha ^b\eta ^c,`$
$`\delta \rho ^a`$ $`=`$ $`f_{abc}\alpha ^b\rho ^c,`$
$`\delta M^a`$ $`=`$ $`f_{abc}\alpha ^bM^c+f_{abc}_{++}\alpha ^bB_{}^c,`$
$`\delta N^a`$ $`=`$ $`{\displaystyle \frac{2}{g}}_{++}_{}\alpha ^af_{abc}\alpha ^bN^cf_{abc}_{++}\alpha ^bA_{}^c,`$
$`\delta \chi ^a`$ $`=`$ $`f_{abc}\alpha ^b\chi ^c,`$
$`\delta \omega ^a`$ $`=`$ $`f_{abc}\alpha ^b\omega ^c.`$ (19)
This gauge variations suggest that we should take $`h=g`$, so that the $`v_{++}^a`$-component could be identified as the light-cone partner of $`A_{}^a`$,
$`v_{++}^aA_{++}^a;`$ (20)
this procedure yields a pair of component-field gauge potentials: $`A^\mu (A^0,A^1)=(A^{++};A^{})`$ and $`B_{}`$; the latter without the $`B_{++}`$ partner.
It is interesting to point out that at this stage the first remarkable difference between the Abelian and the non-Abelian versions of the theory arises. In the Abelian case , it was shown that both fields $`\chi `$ and $`\omega `$ were gauge invariant and the fields $`M`$ and $`N`$ could be identified with a combination of $`A_{}`$ and $`B_{}`$. This combination, which was naturally dictated by the form of the gauge transformations, ensured the symmetry of the Lagrangian. In the present situation, the gauge transformations do not undertake that we may express $`M`$ and $`N`$ in terms of $`A_{}`$ and $`B_{}`$, as it was done before. On the other hand, the $`\chi `$-and $`\omega `$-fields are no longer auxiliary fields, contrary to what happens in the Abelian version.
To discuss the field-strength superfields, we start by analysing the algebra of the gauge covariant derivatives. The former are defined such that:
$`\{_+,_+\}`$ $``$ $`=2D_+\mathrm{\Gamma }_+,`$
$`\{_+,\overline{}_+\}`$ $``$ $`2i_{++},`$
$`[_+,_{++}]`$ $``$ $`W_{}=D_+\mathrm{\Gamma }_{++}_{++}\mathrm{\Gamma }_+,`$
$`[_+,_{}]`$ $``$ $`W_+=igD_+\mathrm{\Gamma }_{}_{++}\mathrm{\Gamma }_+ig[\mathrm{\Gamma }_+,\mathrm{\Gamma }_{}],`$
$`[\overline{}_+,_{++}]`$ $``$ $`U_+,`$
$`[\overline{}_+,_{}]`$ $``$ $`U_{}=ig\overline{D}_+\mathrm{\Gamma }_{},`$
$`[_{++},_{}]`$ $``$ $`𝒵_+=ig_{++}\mathrm{\Gamma }_{}_{}\mathrm{\Gamma }_{++}ig[\mathrm{\Gamma }_+,\mathrm{\Gamma }_{}].`$ (21)
The results obtained for the field-strengths are consistent with the Bianchi identities. The identity for $`U_+`$,
$`[\overline{}_+,\{_+,\overline{}_+\}]+[_+,\{\overline{}_+,\overline{}_+\}]+[\overline{}_+,\{\overline{}_+,_+\}]=0`$ (22)
gives immediately that $`U_+=0`$. The Bianchi identity for $`Z_+`$,
$`[_{},\{_+,\overline{}_+\}]+\{_+,[\overline{}_+,_{}]\}\{\overline{}_+,[_{},\overline{}_+]\}=0,`$ (23)
allows us to express $`Z_+`$ as
$`Z_+={\displaystyle \frac{i}{2}}_+U_{}{\displaystyle \frac{i}{2}}\overline{}_+W_{};`$ (24)
and, finally, the Bianchi identity
$`[\overline{}_+,\{_+,_+\}]+[_+,\{_+,\overline{}_+\}]+[_+,\{\overline{}_+,_+\}]=0`$ (25)
leads to
$`W_+={\displaystyle \frac{i}{4}}\overline{D}_+.`$ (26)
These are the relevant results yielded by pursuing an investigation of the Bianchi identities.
The gauge field, $`A_\mu `$, has its field strength, $`F_{\mu \nu }`$, located at the $`\theta `$-component of the combination $`\mathrm{\Omega }W_{}+\overline{U}_{}`$. This suggests the following kinetic action for the Yang-Mills sector:
$`S_{YM}`$ $`=`$ $`{\displaystyle \frac{1}{8g^2}}{\displaystyle d^2x𝑑\theta 𝑑\overline{\theta }Tr\mathrm{\Omega }\overline{\mathrm{\Omega }}}`$ (27)
$`=`$ $`{\displaystyle d^2xTr[\frac{1}{4}F_{\mu \nu }F^{\mu \nu }+\frac{i}{4}\mathrm{\Sigma }\underset{++}{\overset{}{}}\overline{\mathrm{\Sigma }}+\frac{1}{4}M^2]},`$
where $`\mathrm{\Sigma }=\rho +i\eta +\overline{\chi }i\overline{\omega }`$ and $`A\stackrel{}{}B(A)BA(B)`$.
Choosing now a supersymmetry-covariant gauge-fixing, instead of the Wess-Zumino, we propose the following gauge-fixing term in superspace:
$`S_{gf}`$ $`=`$ $`{\displaystyle \frac{1}{2\alpha }}{\displaystyle d^2x𝑑\theta 𝑑\overline{\theta }Tr[\mathrm{\Pi }\overline{\mathrm{\Pi }}]}`$ (28)
$`=`$ $`{\displaystyle \frac{1}{2\alpha }}{\displaystyle }d^2x\{[(_\mu A^\mu )^2+(_\mu A^\mu )N+{\displaystyle \frac{1}{4}}N^2]`$
$`+`$ $`{\displaystyle \frac{1}{4}}[M^22M_{++}B_{}+(_{++}B_{})^2]`$
$``$ $`i(\rho +i\eta )\underset{++}{\overset{}{}}(\overline{\rho }i\overline{\eta })\},`$
where $`\mathrm{\Pi }=iD_+\mathrm{\Gamma }_{}+\frac{1}{2}D_+_{}V`$.
So, the total action reads as follows:
$`S`$ $`=`$ $`{\displaystyle }d^2xTr\{{\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }{\displaystyle \frac{1}{2\alpha }}(_\mu A^\mu )^2{\displaystyle \frac{1}{2\alpha }}(_\mu A^\mu )N{\displaystyle \frac{1}{8\alpha }}N^2`$ (29)
$`+`$ $`{\displaystyle \frac{1}{4}}(1{\displaystyle \frac{1}{2\alpha }})M^2+{\displaystyle \frac{1}{4\alpha }}M(_{++}B_{}){\displaystyle \frac{1}{8\alpha }}(_{++}B_{})^2`$
$`+`$ $`{\displaystyle \frac{i}{2\alpha }}(\rho +i\eta )\underset{++}{\overset{}{}}(\overline{\rho }i\overline{\eta })+{\displaystyle \frac{i}{4}}\mathrm{\Sigma }\underset{++}{\overset{}{}}\overline{\mathrm{\Sigma }}\}.`$
Using eq.(29), we are ready to write down the propagators for $`A^a`$, $`B_{}^a`$, $`N^a`$, $`M^a`$, $`\rho ^a`$, $`\eta ^a`$, $`\chi ^a`$ and $`\omega ^a`$:
$`AA`$ $`=`$ $`{\displaystyle \frac{i}{2\mathrm{}(1\mathrm{})}}\omega ^{\mu \nu },`$
$`BB`$ $`=`$ $`{\displaystyle \frac{i(2\alpha 1)}{4\alpha (1\alpha )}}{\displaystyle \frac{_{}^2}{\mathrm{}^2}},`$
$`AN`$ $`=`$ $`{\displaystyle \frac{i\alpha }{\mathrm{}(1\mathrm{})}}(1\mathrm{}+\alpha )^\mu ,`$
$`NA`$ $`=`$ $`{\displaystyle \frac{i}{(1\mathrm{})\mathrm{}}}^\nu `$
$`NN`$ $`=`$ $`{\displaystyle \frac{2i\alpha ^2}{(1\mathrm{})}}`$
$`MM`$ $`=`$ $`{\displaystyle \frac{i}{16\alpha }}{\displaystyle \frac{1}{(1\alpha )}}`$
$`MB`$ $`=`$ $`BM={\displaystyle \frac{i}{8\alpha (1\alpha )}}{\displaystyle \frac{_{}}{\mathrm{}}}`$
$`(\rho +i\eta )(\overline{\rho }i\overline{\eta })`$ $`=`$ $`{\displaystyle \frac{2\alpha }{(\alpha 1)}}{\displaystyle \frac{\underset{}{\overset{}{}}}{4\mathrm{}}}`$
$`(\rho +i\eta )(\chi +i\omega )`$ $`=`$ $`{\displaystyle \frac{\alpha }{4}}{\displaystyle \frac{\underset{}{\overset{}{}}}{\mathrm{}}}`$
$`(\overline{\chi }i\overline{\omega })(\overline{\rho }i\overline{\eta })`$ $`=`$ $`+{\displaystyle \frac{\alpha }{4(\alpha 1)}}{\displaystyle \frac{\underset{}{\overset{}{}}}{\mathrm{}}}`$
$`(\overline{\chi }i\overline{\omega })(\chi +i\omega )`$ $`=`$ $`+{\displaystyle \frac{(\alpha +2)}{4}}{\displaystyle \frac{\underset{}{\overset{}{}}}{\mathrm{}}}.`$ (30)
Expressing the action of equation (10) in terms of component fields, and coming back to the $`(2,0)`$-version of the Wess-Zumino gauge, the matter-gauge sector Lagrangian reads:
$`_{mattergauge}`$ $`=`$ $`2\varphi ^i\mathrm{}\varphi _iig[\varphi ^iA_{}^a(G_a)_i^j_{++}\varphi _jc.c]+\overline{\sigma }^i\sigma _i+`$ (31)
$``$ $`ig[\varphi ^iA_{++}^a(G_a)_i^j_{}\varphi _jc.c]g\varphi ^iM^a(G_a)_i^j\varphi _j+`$
$``$ $`{\displaystyle \frac{i}{2}}g^2\varphi ^iA_{++}^aA_{}^b\varphi _id_{abc}G_cg\overline{\lambda }^iA_{}^a(G_a)_i^j\lambda _j+`$
$``$ $`{\displaystyle \frac{1}{2}}\varphi ^iA_{++}^aB_{}^b\varphi _if_{abc}G_c+2i\overline{\lambda }^i_{}\lambda _i+`$
$``$ $`ig\varphi ^i[(\chi ^a+\overline{\rho }^a+i\omega ^ai\overline{\eta }^a)(G_a)_i^j\lambda _jc.c]+`$
$``$ $`2i\overline{\psi }^i_{++}\psi _ig\overline{\psi }^iA_{++}^a(G_a)_i^j\psi _j,`$
where $`d_{abc}`$ are the(representation-dependent) symmetric coefficients associated to $`\{G_a,G_b\}`$.
One immediately checks that the extra gauge field, $`B_{}`$, does not decouple from the matter sector. Our point of view of keeping the superconnection $`\mathrm{\Gamma }_{}`$ as a complex superfield naturally introduced this extra gauge potential in addition to the usual gauge field, $`A_\mu `$: $`B_{}`$ behaves as a second gauge field. The fact that it yelds a massless pole of order two in the spectrum may harm unitarity. However, the mixing with the $`M`$-component of $`\mathrm{\Gamma }_{}`$, which is a compensating field, indicates that we should couple them to external currents and analyse the imaginary part of the current-current amplitude at the pole. In so doing, this imaginary part turns out to be positive-definite, and so no ghosts are present. It is very interesting to point out that, in the Abelian case, $`B_{}`$ showed the same behaviour . It coupled to $`C`$ instead of $`M`$, but these two fields show the same kind of behaviour: $`C`$ (in the Abelian case) and $`M`$ (in the non-Abelian case) are both compensating fields. This ensures us to state that $`B_{}`$ behaves as a physical gauge field: it has dynamics and couples both to matter and the gauge field $`A^\mu `$. Its only peculiarity regards the presence of a single component in the light-cone coordinates. The $`B`$-field plays rather the rôle of a “chiral gauge potential”. Despite the presence of the pair of gauge fields, a gauge-invariant mass term cannot be introduced, since $`B`$ does not carry the $`B_{++}`$-component, contrary to what happens with $`A^\mu `$.
Let us now turn to the coupling of the two gauge potentials, $`A_\mu `$ and $`B_{}`$, to a nonlinear $`\sigma `$-model, always keeping a supersymmetric scenario. It is our main purpose henceforth to carry out the coupling of a $`(2,0)`$ $`\sigma `$-model to the relaxed gauge superfields of the ref. , and show that the extra vector degrees of freedom do not decouple from the matter fields (that is, the target space coordinates). The extra gauge potential, $`B_{}`$, obtained upon relaxing constraints can therefore acquire a dynamical significance by means of the coupling between the $`\sigma `$-model and the Yang-Mills fields of ref.. To perform the coupling of the $`\sigma `$-model to the Yang-Mills fields we reason along the same considerations as i ref. and find out that:
$`_\xi `$ $`=`$ $`_i[K(\mathrm{\Phi },\stackrel{~}{\mathrm{\Phi }})\xi (\mathrm{\Phi })\stackrel{~}{\xi }(\stackrel{~}{\mathrm{\Phi }})]_{}\mathrm{\Phi }^i+`$ (32)
$``$ $`\stackrel{~}{}_i[K(\mathrm{\Phi },\stackrel{~}{\mathrm{\Phi }})\xi (\mathrm{\Phi })\stackrel{~}{\xi }(\stackrel{~}{\mathrm{\Phi }})]_{}\stackrel{~}{\mathrm{\Phi }}^i,`$
where $`\xi (\mathrm{\Phi })`$ and $`\overline{\xi }(\overline{\mathrm{\Phi }})`$ are a pair of chiral and antichiral superfields, $`\stackrel{~}{\mathrm{\Phi }}_iexp(i𝐋_{V.\overline{k}})\overline{\mathrm{\Phi }}_i`$ and $`_{}\mathrm{\Phi }^i`$ and $`_{}\stackrel{~}{\mathrm{\Phi }}^i`$ are defined in perfect analogy to what is done in the case of the bosonic $`\sigma `$-model:
$`_{}\mathrm{\Phi }_i_{}\mathrm{\Phi }_ig\mathrm{\Gamma }_{}^\alpha k_\alpha ^i(\mathrm{\Phi })`$ (33)
and
$`_{}\stackrel{~}{\mathrm{\Phi }}_i_{}\stackrel{~}{\mathrm{\Phi }}_ig\mathrm{\Gamma }_{}^\alpha \overline{k}_{\alpha i}(\stackrel{~}{\mathrm{\Phi }}).`$ (34)
The interesting point we would like to stress is that the extra gauge degrees of freedom accommodated in the component-field $`B_{}(x)`$ of the superconnection $`\mathrm{\Gamma }_{}`$ behave as a genuine gauge field that shares with $`A^\mu `$ the feature of coupling to matter and to $`\sigma `$-model . This result can be explicitly read off from the component-field Lagrangian projected out from the superfield Lagrangian $`_\xi `$.
We therefore conclude that our less constrained $`(2,0)`$-gauge theory yields a pair of gauge potentials that naturally transform under the action of a single compact and simple gauge group and may be consistently coupled to matter fields as well as to the $`(2,0)`$ nonlinear $`\sigma `$-models by means of the gauging of their isotropy and isometry groups.
Relaxing constraints in the $`N=1`$\- and $`N=2D=3`$ supersymmetric algebra of covariant derivatives may lead to a number of peculiar features of the gauged $`O(3)`$-$`\sigma `$-model in the presence of Born-Infeld terms for the pair of gauge potentials which share the same symmetry group; of special interest are the self-dual equations that may stem from this model .
The authors would like to thank M. A. De Andrade, A. L. M. A. Nogueira and O. Del Cima for enlightening discussions; CNPq and Capes are acknowledged for the invaluable financial support.
|
warning/0006/astro-ph0006132.html
|
ar5iv
|
text
|
# Linear dynamics of the solar convection zone: excitation of waves in unstably stratified shear flows
## INTRODUCTION
The excitation and propagation of waves are important for understanding the dynamics of the sun and stars. It is believed that most of the solar mechanical energy is accumulated in the turbulent motions in its convection zone. In the convection zone the gravitational stratification drives the convective instability providing the dynamical activity of this relatively thin region. The dynamics of the solar convection is studied to explain many observational features of the Sun. Notably, it is thought that the solar acoustic oscillations are excited by the turbulence in the convection zone \[1-5\].
Lighthill’s ideas of aerodynamic sound generation form the basis of the theoretical investigation of the wave excitation in a hydrodynamic medium . This theory of wave excitation by a free turbulence has been generalized for stratified fluids by Stein . From a physical point of view, Lighthill’s theory of wave generation employs the concept of stochastic excitation of oscillations (waves). In Lighthill’s theory perturbations are described by an inhomogeneous wave equation, with linear terms forming the oscillatory part and the inhomogeneous terms standing for the source function. The source terms, which may be classified by their multipole order, are stochastically created by the turbulent perturbations. The amplifying effect of a sheared mean flow on the fluctuations of the Reynolds stress (nonlinear source term) and thus on the wave production has been noted by Lighthill . However, this effect has not received further attention within the context of stochastic excitation.
Significant advances in the investigation of the dynamics of flows with velocity shear have been achieved together with the disclosure of specific features of shear flow phenomena . Operators arising in the mathematical formalism of the canonical modal analysis in the study of the linear dynamics of shear flows are not self-adjoint. Consequently eigenmode interference introduces principal complications. The nonmodal approach has proved to be an alternative successful route for exploring the dynamics of shear flows. This approach employs the study of temporal evolution of the spatial Fourier harmonics of perturbations.
Impressive progress has been made by use of the nonmodal analysis (see e.g., \[11-16\]). This approach has led to the discovery of new channels of energy exchange between different modes in shear flows. Resonant phenomena of wave transformations have been studied in \[17-23\]. The nonresonant phenomenon of the conversion of vortices into acoustic waves has been described in . The same mechanism is found to operate for magnetosonic as well as for plasma Langmuir oscillations .
In this report we introduce a new dynamical source of acoustic waves in unstably stratified shear flows. Namely, the linear nonresonant conversion of convective into acoustic wave modes in a stratified shear flows. Convectively unstable exponentially growing buoyancy perturbations generate acoustic wave oscillations in presence of a sheared mean flow. We identify this linear conversion of modes in shear flows as a new excitation mechanism of the solar oscillations and waves. It differs in principle from the stochastic excitation mechanism and should significantly contribute to the process of acoustic wave generation in the solar convection zone.
## Physical approach
The equations governing the dynamics of a compressible stratified flow are:
$$\left[_t+(𝐕)\right]\rho +\rho (𝐕)=0,$$
$`(1.a)`$
$$\left[_t+(𝐕)\right]𝐕=P/\rho +𝐠,$$
$`(1.b)`$
$$\left[_t+(𝐕)\right]P=(\gamma P/\rho )\left[_t+(𝐕)\right]\rho .$$
$`(1.c)`$
We consider the hydrodynamic situation where a horizontal shear flow $`\text{V}_0=(Ay,0,0)`$ occurs in a vertically stratified medium $`\text{g}=(0,0,g)`$. For simplicity we assume that $`A=const`$ and $`g=const`$. This yields the stratified equilibrium state:
$$P_0(z)/P_0(0)=\rho _0(z)/\rho _0(0)=\mathrm{exp}(zk_H),$$
$`(2)`$
where $`k_H\gamma g/c_s^2`$ and $`c_s^2\gamma P_0/\rho _0`$. We introduce the linear perturbations in the following way:
$$\text{V}=\text{V}_0+\text{V}^{}\rho _0(0)/\rho _0(z),P=P_0+P^{},\rho =\rho _0+\rho ^{}.$$
$`(3)`$
Here the velocity perturbations are normalized to exclude the exponential height dependence due to the vertical stratification of the background flow. We use the Cowling approximation and neglect the perturbations of the gravitational acceleration. Following the standard method of nonmodal analysis (see for a rigorous mathematical interpretation) we introduce the spatial Fourier harmonics (SFH) of the perturbations with time dependent phases:
$$\mathrm{\Psi }(\text{r},t)=\psi (\text{k}(t),t)\mathrm{exp}(\mathrm{i}k_xx+\mathrm{i}k_y(t)y+\mathrm{i}\stackrel{~}{k}_zz),$$
$`(4.a)`$
$$k_y(t)=k_y(0)Ak_xt,$$
$`(4.b)`$
where $`\stackrel{~}{k}_zk_z+\mathrm{i}k_H/2`$. For compactness of notation we introduce the generalized vector of perturbations and their SFHs as follows: $`\mathrm{\Psi }(\text{V}^{},p^{},\rho ^{})`$ and $`\psi (\text{u},p,\rho )`$. To avoid complex coefficients in the dynamical equations, we construct the normalized entropy and vertical velocity perturbation SFHs in the following way:
$$s(\mathrm{i}c_s^2\stackrel{~}{k}_z^{}/g1)(pc_s^2\rho )/(\gamma 1),$$
$`(5.a)`$
$$v(c_s^2\stackrel{~}{k}_z^{}+\mathrm{i}g)u_z,$$
$`(5.b)`$
where $`\stackrel{~}{k}_z^{}=k_z\mathrm{i}k_H/2`$. From Eqs. (1-5) we obtain, by the use of straightforward manipulations, the following set of differential equations that govern the SFH of the linear perturbations in stratified shear flow:
$$\stackrel{\text{.}}{p}(t)=c_s^2(k_xu_x+k_y(t)u_y)+v,$$
$`(6.a)`$
$$\underset{x}{\overset{\text{.}}{u}}(t)=Au_yk_xp,$$
$`(6.b)`$
$$\underset{y}{\overset{\text{.}}{u}}(t)=k_y(t)p,$$
$`(6.c)`$
$$\stackrel{\text{.}}{v}(t)=(N_\mathrm{B}^2c_s^2\overline{k}_z^2)pN_\mathrm{B}^2s,$$
$`(6.d)`$
$$\stackrel{\text{.}}{s}(t)=v.$$
$`(6.e)`$
$`N_\mathrm{B}^2`$ is the square of the frequency of the Brunt-Väisälä: $`N_\mathrm{B}^2gk_H(\gamma 1)/\gamma `$ and $`\overline{k}_z^2=|\stackrel{~}{k}_z|^2=k_z^2+k_H^2/4`$. In an unstably stratified flow negative buoyancy ($`N_\mathrm{B}^2<0`$) requires that the adiabatic index $`\gamma <1`$. Such an effective value may be assigned to this parameter under a certain thermodynamic approach . However, in Eqs (6.a-e) we retain only $`N_\mathrm{B}^2`$ and argue that these equations are more general than the underling $`\gamma `$ prescription.
Further we note that vorticity is conserved in the wave-number space: $`I=k_xu_yk_y(t)u_x(A/c_s^2)(ps)`$. The spectral energy of the perturbations can be defined as follows:
$$E=\rho _0/(2c_s^2)\left(E_\mathrm{K}+E_\mathrm{P}+E_\mathrm{T}\right),$$
$`(7.a)`$
$$E_\mathrm{K}=c_s^2(u_x^2+u_y^2)+v^2/(c_s^2\overline{k}_z^2N_\mathrm{B}^2),$$
$`(7.b)`$
$$E_\mathrm{P}=p^2,E_\mathrm{T}=N_\mathrm{B}^2s^2/(c_s^2\overline{k}_z^2N_\mathrm{B}^2).$$
$`(7.c,d)`$
where $`E_\mathrm{K}`$, $`E_\mathrm{P}`$ and $`E_\mathrm{T}`$ correspond to the kinetic, elastic and thermobaric energies of the perturbations, respectively. Formally the perturbation energy is conserved in the shearless limit: $`\stackrel{\text{.}}{E}=Ac_s^2u_xu_y`$. The instability of the convective eddies corresponds to a negative value of the thermobaric energy.
### Linear modes in the shearless limit
The linear modes may be classified explicitly in the shearless limit ($`A=0`$). In this case the full Fourier expansion of the linear perturbations $`\mathrm{\Psi }(\mathrm{r},t)\stackrel{~}{\psi }(\mathrm{k},\omega )`$ yields the dispersion equation:
$$\omega (\omega ^4c_s^2k^2\omega ^2+N_\mathrm{B}^2c_s^2k_{}^2)=0,$$
$`(8)`$
where $`k_{}^2k_x^2+k_y^2`$ and $`k^2=k_{}^2+\overline{k}_z^2`$. The solutions of Eq. (8) describe the stability and characteristic temporal variation scales of the existing modes:
$$\omega _\mathrm{v}=0,$$
$`(9.a)`$
$$\omega _{s,c}^2=\frac{1}{2}c_s^2k^2\left\{1\pm \left(1\frac{4N_\mathrm{B}^2k_{}^2}{c_s^2\overline{k}^4}\right)^{1/2}\right\},$$
$`(9.b)`$
where the subscripts v, s, c define the frequencies of the vortex, acoustic and convective modes, respectively. In an unstably stratified flow, i. e., when $`N_\mathrm{B}^2<0`$, i$`\omega _c`$ defines the growth rate of the buoyancy perturbations.
Obviously the $`I=constant`$ law demonstrates the existence of the stationary ($`\omega =0`$) vortex mode in the linear spectum. The conserved vorticity $`I`$ may be considered as the vortex mode measure. The physical eigenfunctions of the acoustic $`\mathrm{\Phi }_s(t)`$ and convective $`\mathrm{\Phi }_c(t)`$ modes may be rigorously defined in this limit:
$$\mathrm{\Phi }_s(t)p(t)+N_\mathrm{B}^2\frac{\mathrm{\Omega }_s^2\omega _s^2}{\mathrm{\Omega }_c^4}\left(s(t)\frac{\overline{k}_z^2}{k^2}p(t)\right)$$
$`(10.a)`$
$$\mathrm{\Phi }_c(t)s(t)\frac{\overline{k}_z^2}{k^2}p(t)\frac{\mathrm{\Omega }_c^2\omega _c^2}{N_B^2}p(t)$$
$`(10.b)`$
where $`\mathrm{\Omega }_s^2c_s^2k^2`$ and $`\mathrm{\Omega }_c^2N_\mathrm{B}^2k_{}^2/k^2`$. Hence the equations governing the dynamics of the perturbations of the different modes may be decoupled as follows:
$$\underset{s}{\overset{\text{..}}{\mathrm{\Phi }}}(t)+\omega _s^2\mathrm{\Phi }_s(t)=0,$$
$`(11.a)`$
$$\underset{c}{\overset{\text{..}}{\mathrm{\Phi }}}(t)+\omega _c^2\mathrm{\Phi }_c(t)=0.$$
$`(11.b)`$
Starting from this simple situation we study the velocity shear effects on the perturbation modes.
### Effects of a sheared flow
To study the effects of the velocity shear on the linear modes we introduce the small-scale approximation: $`k_z^2k_H^2`$. This approximation strongly simplifies the mathematical formulation and is justified for the following two reasons. Firstly, our analysis needs constant vertical gravity, an assumption that may be adopted for perturbations with vertical height scales shorter than the stratification scale. Secondly, this approximation is necessary for our assumption of a constant linear shear of the flow velocity, especially in the turbulent flows. Using Eq. (7) this approximation may be represented by the following condition: $`c_s^2k_z^2N_\mathrm{B}^2`$. In terms of the frequencies it yields $`(\mathrm{\Omega }_s^2\omega _s^2)(\mathrm{\Omega }_c^2\omega _c^2)0`$, which strongly simplifies the characteristic physical quantities of the perturbation modes: $`\mathrm{\Phi }_s(t)p(t)`$ and $`\mathrm{\Phi }_c(t)(s(t)\overline{k}_z^2p(t)/k^2(t))`$. To analyze the dynamics of acoustic oscillations in the shear flow we rewrite Eqs. (6.a-e) in the form of coupled second order differential equations for the variables $`p(t)`$ and $`y(t)`$:
$$\stackrel{\text{..}}{p}(t)+f(t)\stackrel{\text{.}}{p}(t)+\mathrm{\Omega }_1^2(t)p(t)=\lambda _1(t)\stackrel{\text{.}}{y}(t)+\lambda _2(t)y(t),$$
$`(12.a)`$
$$\stackrel{\text{..}}{y}(t)+\mathrm{\Omega }_2^2(t)y(t)=0,$$
$`(12.b)`$
where the convection variable $`y(t)`$ is introduced as follows:
$$y(t)\frac{k_{}(t)}{k(t)}\left(s(t)\frac{\overline{k}_z^2}{k^2(t)}p(t)\right)$$
$`(13)`$
and
$$\mathrm{\Omega }_1^2(t)=c_s^2k^2(t)+2A^2\frac{k_x^2}{k^2(t)}4A^2\frac{k_x^2k_y^2(t)\overline{k}_z^2}{k_{}^2(t)k^4(t)},$$
$`(14.a)`$
$$\mathrm{\Omega }_2^2(t)=N_\mathrm{B}^2\frac{k_{}^2(t)}{k^2(t)}+2A^2\frac{k_x^2k_z^2}{k_{}^4(t)k^4(t)}\left[3k_{}^2(t)k^2(t)4k_y^2(t)k_{}^2(t)k_y^2(t)\overline{k}_z^2\right],$$
$`(14.b)`$
$$f(t)=2A\frac{k_xk_y(t)}{k^2(t)},$$
$`(14.c)`$
$$\lambda _1(t)=2A\frac{k_xk_y(t)}{k_{}(t)k(t)},$$
$`(14.d)`$
$$\lambda _2(t)=2A^2\frac{k_x^2k_{}(t)}{k^3(t)}\left(1\frac{k_y^2(t)\overline{k}_z^2}{k_{}^4(t)}\right).$$
$`(14.e)`$
In deriving Eqs. (12.a-b) we have used the following two simplifications. Firstly, we have retained only the terms describing the effect of the buoyancy perturbations on the acoustic waves, and we have neglected the effect of the acoustic pressure perturbations on the evolution (exponential amplification) of the buoyancy perturbations in the right hand side (rhs) of Eq. (12.b). Secondly, we have neglected the source terms in the rhs of the two dynamical equations that describe the shear induced coupling between the vortex and acoustic wave modes (in Eq. 12.a) and vortex and buoyancy modes (in Eq. 12.b). In fact, the coupling of the vortex and acoustic wave is a process that has been studied to reveal the mean flow shear induced nonresonant mode conversion phenomenon in . However, in the present case, the source terms of the acoustic waves that are proportional to the vortex mode measure, conserved quantity $`I`$, are dominated by the source terms, associated with the exponentially amplifying convective modes: $`y(t)`$ and $`\stackrel{\text{.}}{y}(t)`$. It should be emphasized that the present approach is justified only for a convectively unstable medium with $`N_\mathrm{B}^2<0`$, so that the buoyancy modes undergo exponential amplification in the linear regime.
The dynamics of the acoustic waves in the absence of the buoyancy perturbations is described by the homogeneous part of Eq. (12.a). The acoustic wave frequency and amplitude variations are described by the parameters $`\mathrm{\Omega }_1^2(t)`$ and $`f(t)`$ (see for a detailed study). The dynamics of the convective mode is described by Eq. (12.b). Eq. (16.b) shows the transient stabilization effect of the sheared mean flow in an unstably stratified medium. The stabilization occurs at times, when $`|k_y(t)/k_x|<1`$ and reaches its maximum at $`t=t^{}`$, when $`k_y(t^{})=0`$ (see Eq. 16.b).
The terms $`\lambda _1(t)\stackrel{\text{.}}{y}(t)`$ and $`\lambda _2(t)y(t)`$ in the rhs of Eq. (14.a) describe the coupling between the convective and acoustic waves modes. The shear flow origin of these source terms is obvious from Eqs. (16.d,e). Hence, Eqs. (14.a,b) describe the mean flow shear induced buoyancy – acoustic wave mode conversion in a convectively unstable medium. Some specific features of this phenomenon are due to its linear nature; SFH of the exponentially growing buoyancy perturbations are able to generate SFH of the acoustic waves with the same wave-numbers. The amplitude of the excited wave mode depends on the values of the source terms $`\lambda _1(t)`$ and $`\lambda _2(t)`$. So, convective modes with $`k_x=0`$ can not generate acoustic waves at all ($`\lambda _1=\lambda _2=0`$). While maximal efficiency of the mode conversion phenomenon should occur at $`k_z=0`$, or in a realistic physical approximation (see Eq. 12) at $`k_z^2N_\mathrm{B}^2/c_s^2`$. Naturally, acoustic wave emission from convection should generally increase when the mean flow shear parameter $`A`$ increases.
We numerically analyze Eqs. (6.a-e) to verify the analytical results obtained from the approximate equations (12.a,b). We select the initial perturbations in a specific manner, which enables us to excite the convective and acoustic wave modes individually at the initial moment of time. It appears that exponentially growing buoyancy perturbations instantly excite the acoustic wave mode harmonics at a given point in time, when the perturbation wave-number along the flow velocity shear is zero: $`t=t^{}`$, $`k_y(t^{})=0`$. The generation of acoustic waves is clearly traced from the pressure variation, as well as the compression of the perturbations. Numerical analysis shows that the efficiency of this mode conversion phenomenon increases with the flow shear parameter.
## DISCUSSION AND CONCLUSIONS
We have presented a study of compressible convection in shear flows. In particular we have focused on linear small-scale perturbations in unstably stratified flows with constant shear of velocity. The linear character of the system enables us to identify the perturbation modes and to study their dynamics individually. We find a mode conversion that originates from the velocity shear of the flow: exponentially growing perturbations of convection are able to excite acoustic waves. This process offers a novel approach to the hydrodynamic problem of the acoustic wave generation.
This wave excitation phenomenon can be important for the acoustic oscillations of the sun. Being responsible for the wave generation in high shear regions of a stratified turbulent flow, this nonresonant phenomenon can contribute to the production of sound in the solar convection zone. Moreover, the process of the wave excitation should be triggered by a weak vertical magnetic field. In this case we anticipate the production of high frequency compressional MHD waves. The latter process will considerably increase the extraction of the mechanical energy of the convection by waves.
Specific to this phenomenon is that perturbations of buoyancy are able to excite acoustic waves with similar wave-numbers. This property makes it clearly distinct from stochastic excitation, where the generated frequencies are similar to the life-times of the source perturbations. In contrast, frequencies of the oscillations generated by the mean flow velocity shear induced mode conversion may be qualitatively higher than the temporal variation scales of the perturbations in the source flow of a compressible convection. The frequency spectrum of the excited acoustic waves should be intrinsically correlated to the velocity field of the turbulent source flow. Shear flow induced wave excitation in stratified flows offers a natural explanation of the fact, that the solar acoustic oscillation are mainly excited in the high shear regions of the convection, intergranular dark lanes . It also explains the puzzling wave-number dependence of the observed mode energies at fixed frequencies (see and references therein). A detailed comparison with observational data requires a more realistic physical model. The simplicity of our model is used to demonstrate the basic features of this excitation phenomenon.
Finally we note that in the present formalism we have focused on the waves with frequencies higher than the characteristic cut-off frequency for the acoustic waves in the convection zone. Shear flow initiates the qualitative change of the temporal variation scales of perturbations and the excitation of the waves that are not trapped in the convective envelope. Hence, this mode conversion presents a new significant contribution into the channel of energy transfer from the dynamically active interior to the atmosphere of the Sun.
## ACKNOWLEDGMENTS
A. G. Tevzadze would like to acknowledge the financial support as ”bursaal” of the ”FWO Vlaanderen”, project G.0335.98. This work was supported in part by the INTAS grant GE97-0504.
|
warning/0006/hep-ph0006049.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The search for supersymmetry (SUSY) in high energy physics relies both on high energy colliders as well as on experiments based on perturbative corrections to various experimentally measurable quantities. Traditionally, the measurement of an electron’s anomalous magnetic moment has been highly effective in verifying the prediction of quantum electrodynamics (QED) to a very high order. Probing Beyond the Standard Model physics with supersymmetry is seen to be possible with a precision $`(g2)`$ measurement of the muon. The ongoing muon $`(g2)`$ measurement E821 at Brookhaven National Laboratory (BNL), designed to verify the results of Standard Model (SM) electroweak corrections, has already provided a more accurate result than the previous CERN experiment, by a factor of 2 or so. With improved design and state of the art technology, it is expected that within a few years from now, the accuracy of the BNL result will be increased by a factor of 20, or even more, compared to the same of the previous CERN measurement.
Supersymmetric electroweak corrections to muon $`(g2)`$ can be as large as the SM electroweak correction, and this fact has been seen in a number of references ranging from the minimal supersymmetric standard model (MSSM), Supergravity based models, and Gauge Mediated Supersymmetry Breaking scenarios. In the recent past, considerable interest has been seen in a different type of SUSY breaking mechanism, other than Supergravity and Gauge Mediated SUSY Breaking, which at its dominating scenario, is generically known as Anomaly Mediated Supersymmetry Breaking (AMSB) \[ReferencesReferences\]. This effect originates from the existence of a super Weyl anomaly while considering SUSY breaking. As we will discuss latter, the problem of the resulting tachyonic sleptons which arises within the AMSB sparticle spectrum, is avoided in the minimal definition \[ReferencesReferences\] of the model, via adding a common scalar mass $`m_0`$ with all the scalars of the theory, at a given scale.
Large SUSY contributions $`a_\mu ^{\mathrm{SUSY}}\frac{1}{2}(g2)_\mu ^{\mathrm{SUSY}}`$ in the minimal AMSB framework have already been seen in Ref. , in which the authors discussed in detail a broad range of interesting phenomenological implications involving colliders, as well as various low energy signatures within the model, in addition to showing that the minimal model may remain natural even for a superheavy $`m_0`$. In this work, we will analyze the constraint coming from $`a_\mu ^{\mathrm{SUSY}}`$ with regard to the high precision Brookhaven experiment, within the minimal AMSB framework. Considering $`a_\mu ^{\mathrm{expt}}a_\mu ^{\mathrm{SM}}=a_\mu ^{\mathrm{SUSY}}`$ and taking into account the associated error limits in quadratures, we will constrain the SUSY parameter space of the model.
Our work will be organized as follows. In Sec. 2, we will discuss the SM result for $`a_\mu `$ for its different parts of contributions. We will see the existence of a large error associated with the lowest order hadronic vacuum polarization contribution and sources of its possible improvement in the near future, via low energy $`e^+e^{}hadron`$ data from various experiments. We will describe the AMSB framework and the necessity of defining a minimal scenario in Sec. 3. In Sec. 4, we will use Ref. for the result of $`a_\mu ^{\mathrm{SUSY}}`$, where the analysis was performed in the minimal supergravity (mSUGRA) model to see the constraint from the high precision BNL experiment. We will analyze the constraint from $`a_\mu ^{\mathrm{SUSY}}`$ on the parameter space of the minimal AMSB model, due to the present SM and experimental results of $`a_\mu `$. We will also investigate the potential for constraining the minimal AMSB model using the predicted level of the uncertainty of the E821 BNL experiment. We will do so in the minimal no-deviation scenario, which here means seeing no disagreement from the SM result within the experimental and the theoretical uncertainties, once the measurement is complete at the desired level of accuracy. We will also quote the result of the $`bs+\gamma `$ constraint for both signs of $`\mu `$ and the positive role of an $`a_\mu ^{\mathrm{SUSY}}`$ analysis in this regard. In Sec. 5 , we will comment on the disallowed regions which appear due to reasons other than $`a_\mu ^{\mathrm{SUSY}}`$ limits. Primarily, we will examine the disallowed regions of parameters space due to the combined effect of the scale invariant part of the scalar mass relations of sleptons with gauge and Yukawa couplings within the minimal AMSB model and the nature of the associated renormalization group evolutions. We will also see the effect of SUSY-QCD corrections to the bottom-quark mass on the minimal AMSB spectra, a large $`\mathrm{tan}\beta `$ effect, which also has specific features within AMSB models. These regions, when combined with $`a_\mu ^{\mathrm{SUSY}}`$ eliminated parameter space, provide simpler and definite predictions for the lower bounds of the masses of relevant the supersymmetric particles, as well as of the input parameters of the model.
## 2 Standard Model result $`a_\mu ^{\mathrm{SM}}`$ and sources of uncertainty
The Standard Model result for $`a_\mu `$ is
$$\frac{1}{2}(g2)_\mu ^{\mathrm{SM}}=a_\mu ^{\mathrm{SM}}=11659162.8(6.5)\times 10^{10}(11659162.8\pm 6.5)\times 10^{10}$$
(1)
In contrast, the latest data from the ongoing E821 BNL experiment amounts to:
$$a_\mu ^{\mathrm{expt}}=11659210(46)\times 10^{10}$$
(2)
The uncertainty amount is expected to be reduced to $`\delta a_\mu ^{\mathrm{BNL}}\begin{array}{c}<\hfill \\ \hfill \end{array}4\times 10^{10}`$.
The SM result is broken up into
$$a_\mu ^{\mathrm{SM}}=a_\mu ^{\mathrm{QED}}+a_\mu ^{\mathrm{hadronic}}+a_\mu ^{\mathrm{EW}}.$$
(3)
Here $`a_\mu ^{\mathrm{QED}}`$ is the pure QED contribution computed up to five loops in electromagnetic coupling. The quantity $`a_\mu ^{\mathrm{hadronic}}`$ refers to the total hadronic contribution including the lowest order and the next to lowest order hadronic vacuum polarizations and the light-by-light hadronic contribution. The electroweak part $`a_\mu ^{\mathrm{EW}}`$ is the SM electroweak contribution up to two loops. The amounts from the individual parts with corresponding references are summarized in Table (1).
Within $`a_\mu ^{\mathrm{hadronic}}`$, the contribution $`a_\mu ^{had1}`$ arising from the $`\alpha ^2`$ level of hadronic vacuum polarization diagram is the least accurate quantity, and its uncertainty is almost the same as the overall error of $`a_\mu ^{\mathrm{SM}}`$. Hence an accurate determination of $`a_\mu ^{had1}`$ will be increasingly important for compatibility with the high-precision measurement at BNL, which has an expected level of uncertainty of $`\delta a_\mu ^{\mathrm{BNL}}\begin{array}{c}<\hfill \\ \hfill \end{array}4\times 10^{10}`$. However, the present uncertainty in $`a_\mu ^{\mathrm{SM}}`$ is quite small compared to the experimental uncertainty level, a situation which is going to be changed within a few years. The largest hadronic contribution $`a_\mu ^{had1}`$ is obtained from the total Born cross section (lowest order in QED) for hadron productions in $`e^+e^{}`$ annihilation, a result found via dispersion theory and optical theorem. Accurate low energy $`e^+e^{}hadrons`$ data hence become necessary to lower the uncertainty level. Measurements of low energy hadron production cross sections in BES, CMD-II and DA$`\mathrm{\Phi }`$NE, will significantly improve the result of $`a_\mu ^{\mathrm{SM}}`$. In the recent past, the authors of Ref. used ALEPH data from hadronic $`\tau `$ decay, and QCD sum-rule techniques to evaluate $`a_\mu ^{had1}`$ and this improved the hadronic error estimate very significantly. Our analysis uses this result. Further prospects for improving the result of $`a_\mu ^{had1}`$, as well as a critical evaluation of the above estimate may be seen in Ref. .
## 3 Anomaly Mediated Supersymmetry Breaking in the Minimal Scenario
In a supersymmetric theory, additionally, soft SUSY breaking parameters are also contributed via Super-Weyl anomaly, which is a generic feature if SUSY is broken in a supergravity framework. In anomaly mediated supersymmetry breaking scenario, the Super-Weyl anomaly contributions dominate, because the SUSY breaking sector and the visible sector reside in different parallel 3-branes in a string theoretical perspective, and there are no tree-level couplings between the two branes. The form of soft parameters thus generated, are renormalization group (RG) invariant, and at any desired scale, the soft parameters are determined by the appropriate gauge and Yukawa couplings for the same scale. This is particularly interesting to avoid a SUSY flavor problem, because the scale invariance of soft parameters which is provided by special RG trajectories, eliminates the effect of any flavor violating unknown physics possibly existing at a higher scale.
In spite of many desirable features of anomaly mediation, within the framework of the MSSM, one finds that sleptons become tachyonic. In the minimal AMSB model such tachyonic sleptons are avoided by introducing an additional common mass parameter $`m_0`$ for all the scalars of the theory. But, this obviously violates the scale invariance of the model, whereas preserving the same would be desirable in regard to the flavor changing neutral current (FCNC) constraint. Tachyonic sleptons are avoided differently in non-minimal AMSB models, which have appropriate scale invariant scalar mass combinations within RG evolutions, but these are outside the scope of our present work. However, via the existence of focus point of the renormalization group equation (RGE) of $`m_{H_u}^2`$, the minimal model allows the possibility of multi-TeV squarks and sleptons without increasing the fine-tuning measure, and this is an important feature to address many phenomenological issues.
Scalar masses are hence determined via renormalization group equations of the MSSM starting from their respective values at the unification scale ($`M_G1.5`$ to $`2.0\times 10^{16}`$ GeV) and leading up to the electroweak scale $`M_Z`$, the scale for mass of the Z-boson. However, for the first two generations of scalars, because of the negligible first two generation Yukawa couplings, the effect translates to having simply an overall additive constant $`m_0`$ at $`M_G`$, which would have minimal changes due to RG evolutions.
In this analysis, the evolutions of scalar masses, gauge and Yukawa couplings are computed at two loop level of RGE, and trilinear couplings are evolved via one-loop level of the same. Unification of gauge couplings is incorporated with having $`\alpha _3(M_Z)0.118`$. For the Higgsino mixing, $`\mu ^2`$ is computed radiatively via electroweak symmetry breaking condition at the complete one loop level of the effective potential, while optimally choosing a renormalization scale $`Q=\sqrt{(m_{\stackrel{~}{t}_1}m_{\stackrel{~}{t}_2})}`$ for minimization. The analysis also includes supersymmetric QCD correction to bottom quark mass, which is considerable for large $`\mathrm{tan}\beta `$ regions also having its important features in AMSB scenarios.
With a high degree of predictivity the model is described by the following parameters: the gravitino mass $`m_{3/2}`$, the common scalar mass parameter $`m_0`$, the ratio of Higgs vacuum expectation values $`\mathrm{tan}\beta `$, and $`sgn(\mu )`$. Following Ref. , with having the same sign conventions for $`\mu `$ and $`A`$-parameters in this work, we see that the masses are given via the couplings as follows.
Gauginos:
$$\stackrel{~}{m}_1=\frac{33}{5}\frac{g_1^2}{16\pi ^2}m_{3/2},\stackrel{~}{m}_2=\frac{g_2^2}{16\pi ^2}m_{3/2},\stackrel{~}{m}_3=3\frac{g_3^2}{16\pi ^2}m_{3/2}$$
(4)
Higgs and Third generation scalars:
$$\stackrel{~}{m}_i^2=C_i\frac{m_{3/2}^2}{(16\pi ^2)^2}+m_0^2$$
(5)
where $`i(Q,U,D,L,E,H_u,H_d)`$, with $`C_i`$’s being given as
$`C_Q=\frac{11}{50}g_1^4\frac{3}{2}g_2^4+8g_3^4+h_t\widehat{\beta }_{h_t}+h_b\widehat{\beta }_{h_b}`$, $`C_U=\frac{88}{25}g_1^4+8g_3^4+2h_t\widehat{\beta }_{h_t}`$
$`C_D=\frac{22}{25}g_1^4+8g_3^4+2h_b\widehat{\beta }_{h_b}`$, $`C_L=\frac{99}{50}g_1^4\frac{3}{2}g_2^4+h_\tau \widehat{\beta }_{h_\tau }`$, $`C_E=\frac{198}{25}g_1^4+2h_\tau \widehat{\beta }_{h_\tau }`$
$`C_{H_u}=\frac{99}{50}g_1^4\frac{3}{2}g_2^4+3h_t\widehat{\beta }_{h_t}`$, and $`C_{H_d}=\frac{99}{50}g_1^4\frac{3}{2}g_2^4+3h_b\widehat{\beta }_{h_b}+h_\tau \widehat{\beta }_{h_\tau }`$
Here, Q and L are the superpartners of quark and lepton doublet fields, respectively. The superpartners for singlet quark fields for up and down type are U and D, and the same for singlet lepton is E.
Trilinear couplings:
$$A_t=\frac{\widehat{\beta }_{h_t}}{h_t}\frac{m_{3/2}}{16\pi ^2},A_b=\frac{\widehat{\beta }_{h_b}}{h_b}\frac{m_{3/2}}{16\pi ^2},\mathrm{and}A_\tau =\frac{\widehat{\beta }_{h_\tau }}{h_\tau }\frac{m_{3/2}}{16\pi ^2},$$
(6)
where $`\widehat{\beta }`$’s are defined by
$`\widehat{\beta }_{h_t}=h_t\left(\frac{13}{15}g_1^23g_2^2\frac{16}{3}g_3^2+6h_t^2+h_b^2\right)`$, $`\widehat{\beta }_{h_b}=h_b\left(\frac{7}{15}g_1^23g_2^2\frac{16}{3}g_3^2+h_t^2+6h_b^2+h_\tau ^2\right)`$, and $`\widehat{\beta }_{h_\tau }=h_\tau \left(\frac{9}{5}g_1^23g_2^2+3h_b^2+4h_\tau ^2\right)`$.
The quantities for the first two generations can be obtained similarly by considering appropriate Yukawa couplings, which, however, are neglected in our analysis. We note that, Eq. (4) and Eq. (6) are scale invariant. Hence, having no intrinsic RG evolutions of their own, the masses and the trilinear couplings may be computed at any scale once the appropriate gauge and Yukawa couplings are known. However, because of the addition of the $`m_0^2`$ term, which rescues sleptons from being tachyonic, Eq. (5) is not scale invariant. Here the scale for obtaining the mass values of Eq. (5) is chosen as $`M_G`$. Thereafter, RG evolutions of soft scalar parameters and use of the electroweak radiative breaking condition at the complete one loop level produce the sparticle mass spectra. One of the important features of the minimal AMSB model is that the resulting $`SU(2)`$ gaugino mass $`\stackrel{~}{m}_2`$ is quite smaller than $`\stackrel{~}{m}_1`$ as well as $`|\mu |`$. Here, we have also incorporated the non-negligible next to leading order (NLO) corrections for gaugino masses. As a result, the lighter chargino $`\stackrel{~}{\chi }_1^\pm `$ and lightest neutralino $`\stackrel{~}{\chi }_1^0`$ are wino dominated; indeed they are almost degenerate, with the latter becoming the lightest supersymmetric particle (LSP). This has interesting phenomenological aspects, like what is seen most recently in Ref. , where a definite signal in a linear $`e^+e^{}`$ collider could be predicted as a possible minimal AMSB signature. Compared to other SUSY scenarios where the lightest neutralino has a distinctly smaller mass, here this similar mass value of $`m_{\stackrel{~}{\chi }_1^\pm }`$ and $`m_{\stackrel{~}{\chi }_1^0}`$ effectively decreases $`|a_\mu ^{\mathrm{SUSY}}|`$, although a weakly contributing effect. Another striking result of the minimal AMSB model is the strong mass degeneracy between left and right sleptons. Consequently, the third and the second generation L-R mixing angles become significantly larger, going up to the maximal limit for a large $`\mathrm{tan}\beta `$. We will see the strong effect of large smuon L-R mixing on the neutralino loop contributions of $`a_\mu ^{\mathrm{SUSY}}`$ in Sec. 4.
## 4 Results
The diagrams to compute the supersymmetric contributions to muon $`(g2)`$, as shown in Figs. (1a) and (1b), are divided into the chargino-sneutrino loop and the neutralino-smuon loop. We only quote the chargino part of the result here, which dominates in $`a_\mu ^{\mathrm{SUSY}}`$. The neutralino contribution may be seen in Refs. <sup>1</sup><sup>1</sup>1The most general result of the SUSY electroweak contribution to muon $`(g2)`$ in the MSSM, where CP violating phases are considered, can be seen in Ref. .. Separating the chargino contributions into chirality diagonal and nondiagonal parts we have
$$a_{\mu }^{\mathrm{SUSY}}{}_{}{}^{\chi ^\pm }=a_{\mu }^{\mathrm{SUSY}}{}_{}{}^{\chi ^\pm }(\mathrm{nondiag})+a_{\mu }^{\mathrm{SUSY}}{}_{}{}^{\chi ^\pm }(\mathrm{diag}),$$
(7)
where
$$a_{\mu }^{\mathrm{SUSY}}{}_{}{}^{\chi ^\pm }(\mathrm{nondiag})=\frac{m_\mu ^2\alpha _{em}}{4\sqrt{2}\pi m_W\mathrm{sin}^2\theta _W\mathrm{cos}\beta }\underset{i=1}{\overset{2}{}}\frac{1}{m_{\chi _i^\pm }}(U_{i2}V_{i1})F\left(\frac{m_{\stackrel{~}{\nu _\mu }}^2}{m_{\chi _i^\pm }^2}\right)$$
(8)
and
$$a_{\mu }^{\mathrm{SUSY}}{}_{}{}^{\chi ^\pm }(\mathrm{diag})=\frac{m_\mu ^2\alpha _{em}}{24\pi \mathrm{sin}^2\theta _W}\underset{i=1}{\overset{2}{}}\frac{1}{m_{\chi _i^\pm }^2}\left(\frac{m_\mu ^2}{2m_W^2\mathrm{cos}^2\beta }U_{i2}^2+V_{i1}^2\right)G\left(\frac{m_{\stackrel{~}{\nu _\mu }}^2}{m_{\chi _i^\pm }^2}\right).$$
(9)
Here, in general, U and V are unitary $`2\times 2`$ matrices, which diagonalize the chargino mass matrix $`M_{\stackrel{~}{\chi }^\pm }`$ as shown below, via a bi-unitary transformation, $`U^{}M_{\stackrel{~}{\chi }^\pm }V^1=\mathrm{diag}(m_{\stackrel{~}{\chi }_1^\pm },m_{\stackrel{~}{\chi }_2^\pm })`$:
$$M_{\stackrel{~}{\chi }^\pm }=\left(\begin{array}{cc}\stackrel{~}{m}_2& \sqrt{2}m_W\mathrm{sin}\beta \\ \sqrt{2}m_W\mathrm{cos}\beta & \mu \end{array}\right).$$
(10)
With $`M_{\stackrel{~}{\chi }^\pm }`$ being real U and V are orthogonal matrices. The functions $`F(x)`$ and $`G(x)`$ arising from loop integrations are given by: $`F(x)=(3x^24x+12x^2\mathrm{ln}x)/(x1)^3`$, and $`G(x)=(2x^3+3x^26x+16x^2\mathrm{ln}x)/(x1)^4`$.
Using the complete $`a_\mu ^{\mathrm{SUSY}}`$ result for numerical computations, Figs. (2a) and (2b) show the dominance of chargino contributions over the neutralino parts. Here, the two types of contributions are plotted along the axes for $`\mathrm{tan}\beta `$=25, when $`m_{3/2}`$ and $`m_0`$ are varied over a broad range of values ($`m_{3/2}<100`$ TeV and $`m_0<1`$ TeV). It is indeed the chirality nondiagonal term involving the lighter chargino and sneutrino part which dominates over the other contributions in $`a_\mu ^{\mathrm{SUSY}}`$. Because of the same reason, as explained further in a similar mSUGRA analysis of Ref. , there is a definite sign dependence between $`a_\mu ^{\mathrm{SUSY}}`$ and $`\mu `$, namely, $`a_\mu ^{\mathrm{SUSY}}>0`$ for $`\mu >0`$, and $`a_\mu ^{\mathrm{SUSY}}<0`$ for $`\mu <0`$, an important result for $`a_\mu ^{\mathrm{SUSY}}`$. Thus, the lighter chargino mass ($`m_{\stackrel{~}{\chi }_1^\pm }`$) has a significant role in $`a_\mu ^{\mathrm{SUSY}}`$. On the other hand, for a given $`m_{\stackrel{~}{\chi }_1^\pm }`$ value, a heavier $`m_{\stackrel{~}{\nu }_\mu }`$ decreases $`|a_\mu ^{\mathrm{SUSY}}|`$. Because of the presence of Yukawa coupling ($`\frac{1}{\mathrm{cos}\beta }`$) within the chirality nondiagonal terms for both the chargino (see Eq. 8) and neutralino results, we see that $`|a_\mu ^{\mathrm{SUSY}}|`$ is almost proportional to $`\mathrm{tan}\beta `$.
The special signature of AMSB due to an extremely large smuon L-R mixing angle, as mentioned before, affects the neutralino results to diminish strongly via partial cancellation between the terms. The particular neutralino term (see Ref. ), which involves smuon mixing, becomes almost comparable to the significantly contributing chirality nondiagonal neutralino term associated with Yukawa coupling. Both terms depend on $`\mathrm{tan}\beta `$, as well as on the sign of $`\mu `$. A detail numerical investigation shows that the two terms always come in opposite signs, giving a large cancellation within the neutralino result. On a relative scale, we have seen that, for a naturalness favored region of SUSY spectra with a given $`\mathrm{tan}\beta `$, the ratio of neutralino to chargino contributions within $`a_\mu ^{\mathrm{SUSY}}`$ in minimal AMSB is typically smaller by 50% or so, compared to the same within a similar natural mSUGRA spectra.
### 4.1 Constraints from present values of $`a_\mu ^{\mathrm{expt}}`$ and $`a_\mu ^{\mathrm{SM}}`$
Figures (3a) to (3d) show the constraints arising from $`a_\mu ^{\mathrm{SUSY}}`$ when the present experimental data from Brookhaven is compared with the Standard Model result. Here we consider the residual amount $`a_\mu ^{\mathrm{expt}}a_\mu ^{\mathrm{SM}}`$ to limit $`a_\mu ^{\mathrm{SUSY}}`$ within the 2$`\sigma `$ level of combined error estimates, added in quadratures ( $`43.0\times 10^{10}<a_\mu ^{\mathrm{SUSY}}<142.8\times 10^{10}`$). Considering the largest possible $`|a_\mu ^{\mathrm{SUSY}}|`$ within the model, we see that, essentially, a constraint exists only for $`\mu <0`$. The regions excluded by $`a_\mu ^{\mathrm{SUSY}}`$ bounds when combined with the disallowed regions ( labeled by X ) characteristic of the minimal AMSB model itself, along with the experimental constraints on various sparticle masses, a value of $`m_0`$ below 275 GeV is completely eliminated for any value of $`m_{3/2}`$ (see Fig. (3a)). The nature of the excluded region as marked by X principally originates from sleptons turning into the LSP and then becoming tachyonic at the electroweak scale ( see Sec. 5 ). The same for $`\mathrm{tan}\beta =40`$ as shown in Fig. (3c) amounts to $`m_0375`$ GeV. Additionally, as seen in the same displays, significantly larger $`m_0`$ values than what are mentioned above are excluded for a limited region of $`m_{3/2}`$.
Constraining the minimal AMSB model via $`a_\mu ^{\mathrm{SUSY}}`$ can be further effective in the $`(m_{\stackrel{~}{\chi }_1^\pm },m_{\stackrel{~}{\nu }_\mu })`$ plane (see Figs.(3b) and (3d)) because, as noted earlier, the chirality nondiagonal lighter chargino terms dominate over the other contributions. A value of $`m_{\stackrel{~}{\nu }_\mu }`$ below 225 (325) GeV for $`\mathrm{tan}\beta =25(40)`$ is explicitly ruled out via the current limit on $`a_\mu ^{\mathrm{SUSY}}`$. Here we note that in situations similar to the minimal AMSB model, where $`m_{\stackrel{~}{\chi }_1^\pm }`$ and $`m_{\stackrel{~}{\chi }_1^0}`$ masses are almost degenerate and sneutrinos are light, the present experimental lower bound of $`m_{\stackrel{~}{\chi }_1^\pm }`$ is 56 GeV. The white regions denoted by X in the bottom of the $`a_\mu ^{\mathrm{SUSY}}`$ allowed and disallowed zones of Figs.(3b) and (3d) are disallowed for the same reasons as mentioned before and we will further comment on them in Sec. 5 while discussing similar regions in Figs.(4) and (5).
### 4.2 Probing the minimal AMSB scenario further via $`a_\mu ^{\mathrm{SUSY}}`$ and the BNL experiment at its predicted level of accuracy
The uncertainty level of $`\delta a_{\mu }^{}{}_{}{}^{BNL}=4\times 10^{10}`$, which is going to be achieved at Brookhaven within a few years, will, at least, significantly constrain the parameter space of a theory beyond the Standard Model. Considering this predicted level of accuracy, we constrain $`a_\mu ^{\mathrm{SUSY}}`$ within the $`2\sigma `$ limit (see Figs.(4) and (5)), where $`\sigma `$ is obtained from the predicted uncertainty level of BNL experiment and the error associated with the SM result, added in quadratures. The assumed nondiffering central estimates of the experimental and theoretical results would be the limiting scenario of seeing no deviation from the Standard Model result. This analysis would be valid in the situation when the experiment is complete and no deviation from the Standard Model is seen within the error limits. This is in a similar line to what have been seen in Refs. for supersymmetric as well as nonsupersymmetric theories. On a further note, we assume that the hadronic error in $`a_\mu ^{had1}`$ would be staying at its present level. A reduction, on the other hand, which will occur in the near future, would further constrain a similar analysis.
The lower triangular regions in the ($`m_{3/2},m_0`$) plane of Figs. (4) and (5) are the disallowed zones, where $`|a_\mu ^{\mathrm{SUSY}}|`$ exceeds the 2$`\sigma `$ limit of the combined uncertainty. The same result within the $`a_\mu ^{\mathrm{SUSY}}`$-relevant mass pairs ($`m_{\stackrel{~}{\chi }_1^\pm },m_{\stackrel{~}{\nu }_\mu }`$) is presented in the right hand sides of Figs.(4) and (5). We note that, corresponding to $`\mathrm{tan}\beta `$=10 and 25, the minimal AMSB satisfied parameter space is reasonably identical, with respect to the sign of $`\mu `$. In Figs.(4e) and (5e), however, the regions differ in this aspect, and we will come back to them in Sec. 5.
Fig.(4) and (5) indicate that $`m_0<`$ 275 (475) GeV domains will be entirely eliminated for $`\mathrm{tan}\beta =`$10 (25) for both signs of $`\mu `$ . For $`\mathrm{tan}\beta =40`$, the corresponding limits are 625 GeV for $`\mu <0`$ and 800 GeV for $`\mu >0`$. The limit of $`m_0800`$ GeV for $`\mu >0`$ appears because of reasons which we will discuss soon. Within the ($`m_{\stackrel{~}{\chi }_1^\pm },m_{\stackrel{~}{\nu }_\mu }`$) planes of the right hand side of Figs.(4) and (5), we find that for $`\mu <0`$ and $`\mathrm{tan}\beta =`$ 10, 25, and 40, values of $`m_{\stackrel{~}{\nu }_\mu }`$ less than 210, 400, and 560 GeV are excluded. The situation for $`\mu >0`$ is identical for $`\mathrm{tan}\beta =`$ 10 and 25. A significant difference between the signs of $`\mu `$ can be seen now, switching to $`\mu >0`$ and $`\mathrm{tan}\beta =`$40. A similar disallowed range will be very stringent here; namely, $`m_{\stackrel{~}{\nu }_\mu }`$ below 780 GeV will be excluded.
Interestingly, an important result is found, when this analysis of $`a_\mu ^{\mathrm{SUSY}}`$ is combined with the $`bs+\gamma `$ constraint. The constraint from $`bs+\gamma `$ within the minimal AMSB scenario as analyzed in Ref. is somewhat complementary to what we find here from $`a_\mu ^{\mathrm{SUSY}}`$. This is because the $`bs+\gamma `$ calculation, which has many special features in AMSB models, puts severe mass limits for $`\mu >0`$ and much smaller limits for $`\mu <0`$. On the other hand, within the above scenario of seeing no deviation from the SM result once the experiment is performed at the predicted level of accuracy, $`a_\mu ^{\mathrm{SUSY}}`$ limits in the minimal AMSB model impose a very significant constraint for both $`\mu <0`$ and $`\mu >0`$ cases.
## 5 Generally disallowed parameter zones
A discussion about the generally eliminated parameter space may be useful in studying the supersymmetric contribution to muon $`(g2)`$ in a given SUSY model, because a combined constraint from $`a_\mu ^{\mathrm{SUSY}}`$, as well as from any generic disallowedness, results in simpler and definite predictions. Restricting the stau from becoming tachyonic corresponds to a significant constraint in AMSB models. In this section, by allowedness we mean valid input parameters from the model, in addition to satisfying various experimental lower bounds of sparticle masses, without reference to any $`a_\mu ^{\mathrm{SUSY}}`$ constraint.
We will first describe a few observations as revealed from our numerical analysis. For a given $`m_0`$, the larger $`m_{3/2}`$ values falling within the region labeled by X in Figs. (4) and (5) ( also in Fig. (3) ) are eliminated because of a decreasing stau mass ($`m_{\stackrel{~}{\tau }_1}`$), which either goes below the experimental lower limit of 70 GeV or becomes the LSP, hence discarded in our R-parity conserved scenario. Thereafter, with a further increase of $`m_{3/2}`$, the stau becomes tachyonic. We also see that the maximum possible $`m_{3/2}`$ for a given $`m_0`$, as allowed by the minimal AMSB model, is larger for a smaller $`\mathrm{tan}\beta `$. Thus, for $`\mu <0`$ and $`m_0=`$ 400 GeV, comparing Figs. (4a), (4c), and (4e) we find that such maximum possible values of $`m_{3/2}`$ are approximately 67, 52, and 41 TeV for $`\mathrm{tan}\beta =`$ 10, 25, and 40, respectively. On the other hand, for a given $`\mathrm{tan}\beta `$, an allowed $`m_{3/2}`$ increases for an increase in $`m_0`$. Besides, smaller $`m_{3/2}`$ regions below the origins of the displays are eliminated via the experimental constraint of $`m_{\stackrel{~}{\chi }_1^\pm }\begin{array}{c}>\hfill \\ \hfill \end{array}`$ 56GeV.
We will try to explain, qualitatively, the behavior of stau mass with variations of the basic parameters of the model. The effect of $`\stackrel{~}{\tau }_1`$ becoming tachyonic, as described above, is best explained via $`\stackrel{~}{m}_L^2`$ \[see Eq. (5\] for $`iL`$) assuming smaller left-right slepton mixing for convenience. There are two effects in $`\stackrel{~}{m}_L^2`$ due to a change in $`\mathrm{tan}\beta `$, which may support or oppose each other. The first one arises from the scale invariant part of $`\stackrel{~}{m}_L^2`$ and the other one originates from RG evolution of the same.
The Yukawa term in the scale invariant part of $`\stackrel{~}{m}_L^2`$ in Eq. (5) is intrinsically negative, itself being also gauge coupling dominated within the corresponding $`\widehat{\beta }`$ function, for the range of $`\mathrm{tan}\beta `$ considered in this analysis. Hence, the value of $`\stackrel{~}{m}_L^2`$ at $`M_G`$ decreases if $`\mathrm{tan}\beta `$ is larger. We consider here moderate values of $`m_0`$ for a simpler discussion. Until $`\mathrm{tan}\beta `$ is in a smaller domain, so that $`\tau `$-Yukawa coupling ($`h_\tau `$) within the scalar mass RG equation may be neglected compared to the gauge terms, the RGE effect due to running from $`M_G`$ to the electroweak scale always increases $`\stackrel{~}{m}_L^2`$ because of gauge domination. Thus the two effects oppose each other. But within smaller $`\mathrm{tan}\beta `$ domains, regarding the value of $`\stackrel{~}{m}_L^2`$ at the electroweak scale for an increase in $`\mathrm{tan}\beta `$, the effect of the AMSB specified decrease in $`\stackrel{~}{m}_L^2`$ at $`M_G`$ is stronger than the increase due to the RGE effect. We have also verified this numerically in a broad domain of parameter ranges. As a result, $`\stackrel{~}{m}_L^2`$ and consequently $`m_{\stackrel{~}{\tau }_1}`$ decrease with an increase in $`\mathrm{tan}\beta `$. This in turn means that, for a given $`m_0`$, the upper limit of $`m_{3/2}`$ is reached sooner for a larger $`\mathrm{tan}\beta `$. This we may see from Figs. (4) and (5), as well as from the values quoted above within this section.
For a further increase in $`\mathrm{tan}\beta (40`$ in our analysis), instead of opposing, the two effects may go in the same direction, although with varying strengths, because the $`\tau `$-Yukawa term may now start to dominate within the RG evolution. In fact, this may also be true when $`m_0`$ is large, with $`\mathrm{tan}\beta `$ in a moderately larger domain ($``$ 25).
On the other hand, corresponding to the lowest $`m_{3/2}`$ values satisfying the lighter chargino experimental bound, a gradual increase of the lower limit of $`m_0`$ with an increase in $`\mathrm{tan}\beta `$ is found (see left side displays of Fig.(4) and (5)), because, as explained before, the scale invariant part of $`\stackrel{~}{m}_L^2`$ turns further negative for increasing $`\mathrm{tan}\beta `$, and larger $`m_0`$ values are hence needed to compensate. However, there is a marked difference between the $`\mu <0`$ and $`\mu >0`$ cases for $`\mathrm{tan}\beta =40`$ \[see Fig.(4e) and (5e)\]. The upper limit of the $`m_{3/2}`$ for $`\mu >0`$ as allowed by the model, is much smaller compared to the same for $`\mu <0`$. This happens due to a generic large $`\mathrm{tan}\beta `$ effect, the effect of large SUSY-QCD loop corrections of bottom quark mass. Significantly, this correction has special features in the AMSB scenario , because within the same the $`SU(3)`$ gaugino mass $`\stackrel{~}{m}_3`$ comes with a negative sign. Consequently, for $`\mu >0`$ and large $`\mathrm{tan}\beta `$, as a result of a large SUSY QCD loop correction, a very large $`h_b`$ ($`h_t`$) causes $`m_{H_d}^2`$ for the Higgs scalar to turn sufficiently negative so that the CP-odd Higgs particle becomes tachyonic. However, here $`\stackrel{~}{\tau }_1`$ can still remain nontachyonic. A further increase of $`m_{3/2}`$ causes $`\stackrel{~}{\tau }_1`$ to become tachyonic, as usual.
Considering now the combined effect of the model specified disallowed space, as well as the constraint from $`a_\mu ^{\mathrm{SUSY}}`$, we find that a region below $`m_0=800`$ GeV for $`\mu >0`$ will be completely eliminated within the scenario of seeing no deviation from the SM. The right hand side displays of Figs.(4) and (5) also show model specified eliminated regions, as identified by X in the ($`m_{\stackrel{~}{\chi }_1^\pm },m_{\stackrel{~}{\nu }_\mu }`$) plane. Obviously, the masses are not independent, because they are derived from the basic set of input parameters of the minimal AMSB model. The same reason for $`\stackrel{~}{\tau }_1`$ becoming tachyonic eliminates a large region within the zones X. The disallowed X-zone is large for $`\mu >0`$ and $`\mathrm{tan}\beta =40`$ in Fig.(5f), compared to the same for $`\mu <0`$, as shown in Fig.(4f). This occurs because of the same reason for which the upper limit of $`m_{3/2}`$ is smaller in Fig.(5e), which we have explained before, and due to the fact that the lighter chargino is wino dominated within AMSB, thus $`m_{\stackrel{~}{\chi }_1^\pm }`$ being almost proportional to $`m_{3/2}`$.
## 6 Conclusion
We have computed the supersymmetric contribution to the anomalous magnetic moment of the muon within the minimal Anomaly Mediated Supersymmetry Breaking model. There are one-loop contributions involving chargino-sneutrino and neutralino-smuon parts. The chiral interference term involving the lighter chargino is seen to contribute the most to $`a_\mu ^{\mathrm{SUSY}}`$ than the other chargino and neutralino terms, and this also results in a definite sign relationship between $`a_\mu ^{\mathrm{SUSY}}`$ and $`\mu `$. In addition, this also gives an almost proportional relationship of $`|a_\mu ^{\mathrm{SUSY}}|`$ with $`\mathrm{tan}\beta `$. We have also seen the effect of large smuon L-R mixing which causes strong partial cancellations between the various terms of the neutralino result of $`a_\mu ^{\mathrm{SUSY}}`$. This is a significantly important result of $`a_\mu ^{\mathrm{SUSY}}`$ within minimal AMSB.
We have analyzed the constraint coming from current values of $`a_\mu `$ from the Standard Model and the ongoing experiment at Brookhaven, assuming that the difference appears due to SUSY. The constraint which exists only for $`\mu <0`$ shows that, for $`\mathrm{tan}\beta =`$ 25 (40), regions with $`m_0<`$ 275 (375) GeV and, correspondingly, $`m_{\stackrel{~}{\nu }_\mu }<`$ 225 (325) GeV are eliminated. We have also investigated the constraint from $`a_\mu ^{\mathrm{SUSY}}`$ that would result if the Brookhaven experiment with its already predicted level of accuracy finds no deviation from the Standard Model result. In this scenario, one finds that for $`\mu <0`$ and $`\mathrm{tan}\beta =`$ 10, 25, and 40, the lower bounds of $`m_0`$ would be 275, 475, and 625 GeV, while the corresponding lower limits for $`m_{\stackrel{~}{\nu }_\mu }`$ would be 210, 400, and 560 GeV, respectively. The lower bounds for $`\mu >0`$ are identical to $`\mu <`$ 0, except for $`\mathrm{tan}\beta =`$40, where $`m_0<`$ 800 GeV and correspondingly $`m_{\stackrel{~}{\nu }_\mu }<`$ 780 GeV regions would be excluded. This happens due to a large SUSY-QCD correction to the bottom-quark mass, a large $`\mathrm{tan}\beta `$ effect.
We have also compared our constraint with the same obtained from $`bs+\gamma `$ from Ref. . We found that the high accuracy level of the BNL experiment will be very useful to constrain the model for $`\mu <0`$, because the $`bs+\gamma `$ limit is effective only for $`\mu >`$ 0. Furthermore, we have also analyzed the generically disallowed zones within the parameter space of the model, because a combined constraint from $`a_\mu ^{\mathrm{SUSY}}`$ and such invalid parameter ranges lead to a stronger prediction.
Acknowledgments: DKG wishes to acknowledge the hospitality provided by the Theory Division, CERN, where part of this work was completed.
|
warning/0006/hep-th0006160.html
|
ar5iv
|
text
|
# Untitled Document
hep-th/0006160 CALT-68-2283 CITUSC/00-027
Quantum Field Theories With Compact Noncommutative Extra Dimensions
Jaume Gomis, Thomas Mehen and Mark B. Wise
Department of Physics, California Institute of Technology, Pasadena, CA 91125
and
Caltech-USC Center for Theoretical Physics
University of Southern California
Los Angeles, CA 90089
gomis, mehen, wise@theory.caltech.edu
We study field theories on spaces with additional compact noncommutative dimensions. As an example, we study $`\varphi ^3`$ on $`R^{1,3}\times T_\theta ^2`$ using perturbation theory. The infrared divergences in the noncompact theory give rise to unusual dynamics for the mode of $`\varphi `$ which is constant along the torus. Correlation functions involving this mode vanish. Moreover, we show that the spectrum of Kaluza-Klein excitations can be very different from the analogous commuting theory. There is an additional contribution to the Kaluza-Klein mass formula that resembles the contribution of winding states in string theory. We also consider the effect of noncommutativity on the four dimensional Kaluza-Klein excitations of a six dimensional gauge field.
June 2000
1. Introduction
Quantum field theories on noncommutative spaces – usually referred to as noncommutative field theories – generalize the familiar structure of conventional (commutative) local quantum field theories. Field theories on these spaces can be studied using conventional techniques by representing the underlying noncommutative structure that defines the noncommutative space in terms of more familiar entities in a commutative space. For example, noncommutative Minkowski space $`R_\theta ^{1,d}`$ is defined in terms of space-time coordinates $`x^\mu ,\mu =0,\mathrm{},d`$, which satisfy the following commutation relations
$$[x^\mu ,x^\nu ]=i\theta ^{\mu \nu }\mu ,\nu =0,\mathrm{},d,$$
where $`\theta ^{\mu \nu }`$ is a real antisymmetric matrix. $`R_\theta ^{1,d}`$ is a noncommutative associative algebra, with elements given by ordinary continuous functions on $`R^{1,d}`$ and whose product is given by the Moyal bracket or $``$-product of functions
$$(fg)(x)=e^{\frac{i}{2}\theta ^{\mu \nu }\frac{}{\alpha ^\mu }\frac{}{\beta ^\nu }}f(x+\alpha )g(x+\beta )|_{\alpha =\beta =0}=fg+\frac{i}{2}\theta ^{\mu \nu }_\mu f_\nu g+𝒪(\theta ^2).$$
The generalization of conventional field theories on $`R^{1,d}`$ to $`R_\theta ^{1,d}`$ is achieved by replacing the usual multiplication of fields in the action by the $``$-product of fields.
The presence of $``$-products in the action leads to theories that are not maximally Lorentz invariant nor local. The nonlocality of the theory is apparent from the infinite number of derivatives that appear in the action. The noncommutativity of the space-time coordinates (1.1) gives rise to a space-time uncertainty relation
$$\mathrm{\Delta }x^\mu \mathrm{\Delta }x^\nu \frac{1}{2}|\theta ^{\mu \nu }|.$$
A dramatic consequence of this relation is the absence of decoupling of scales in these theories -. For instance, if $`\theta ^{12}0`$, short distance scales in the $`x^1`$ direction correspond to very large distance scales in the $`x^2`$ direction. Therefore, these theories exhibit a very peculiar mixing between the ultraviolet and the infrared, such that very high energy modes have drastic effects on low energy processes. In spite of this, the theories are apparently renormalizable -. Morever, field theories with only space noncommutativity (that is, $`\theta ^{0i}=0`$) have a unitary S-matrix. On the other hand, theories with only space-time noncommutativity (that is $`\theta ^{0i}0`$) are not unitary .
A strong motivation for understanding these field theories is the appearance they make in string theory -. For instance, noncommutative gauge theories with space noncommutativity describe the low energy excitations of open strings on D-branes in a background Neveu-Schwarz two-form field $`B`$ . The excited open string states and the closed strings decouple and the noncommutative field theory is the proper description of the physics. The lack of covariance of noncommutative field theory arises from the expectation value of $`B`$. The appearance of noncommutative field theory in a definite limit of string theory strongly suggests that these field theories are sensible quantum field theories. This motivates generalizing the conventional framework of local quantum field theory in order to understand these theories.
In nature we observe three spatial dimensions. However, it is possible that there are additional spatial dimensions. Such is the case in string theory and recently there has been speculation that extra dimensions may play a role in understanding the hierarchy puzzle or the cosmological constant problem . If extra dimensions exist the simplest way to explain why they have not been observed is to assume they are compact. It is possible that such additional compact spatial dimensions are noncommuting. In this paper we explore this possibility in some simple field theories that contain two additional compact noncommuting dimensions. We find that the UV-IR mixing that exists for infinite noncommuting dimensions leads to some interesting properties of the compactified theories.
In this paper we study in perturbation theory scalar and gauge quantum field theories with compact noncommutative extra dimensions. We examine the low energy dynamics of these theories and their spectrum of Kaluza-Klein excitations. For concreteness, we consider six dimensional field theories, such that the two extra dimensions are compact and noncommutative and four noncompact dimensions are commutative. The two extra dimensions are taken to correspond to a noncommutative two torus $`T_\theta ^2`$ whose coordinates satisfy
$$[x^4,x^5]=i\theta .$$
This system can be realized in string theory by wrapping a five-brane on a two-torus $`T^2`$ with a constant $`B`$-field along the torus. The low energy effective four dimensional theory resulting from compactification on a noncommutative space is local and Lorentz invariant, hence it can be relevant phenomenologically.
In section $`2`$ we consider scalar $`\varphi ^3`$ theory on $`R^{1,3}\times T_\theta ^2`$. The noncommutative theory on a noncompact space contains infrared divergences. In the compact theory these appear as additional divergences in two and three point functions involving the mode of $`\varphi `$ that is constant along the torus, $`\varphi _0`$. The correct interpretation of these divergences is unusual dynamics for $`\varphi _0`$. Since the mass of $`\varphi _0`$ diverges as the ultraviolet cutoff is taken to infinity, $`\varphi _0`$ no longer appears as a propagating degree of freedom in the low energy effective theory.
In section 3 we discuss the Kaluza-Klein mass formula for the scalar field theory introduced in section 2. There are one loop corrections which are singular as $`\theta `$ goes to zero. For small $`\theta `$, they lead to a four dimensional spectrum of states which differs qualitatively from the Kaluza-Klein spectrum of a commutative theory. The one loop correction to the dispersion relation for the Kaluza-Klein excitations resembles that of winding states in string theory .
In section 4 we briefly discuss similar issues in a $`U(1)`$ gauge theory compactified on $`R^{1,3}\times T_\theta ^2`$. Again we find that modes of the gauge field which are constant along the torus disappear from the low energy theory. The six dimensional gauge field contains Kaluza-Klein excitations that are either four-dimensional vectors or four-dimensional scalars. One loop corrections modify the spectrum of the four-dimensional scalars, however, the spectrum of the four-dimensional vectors is unchanged.
Concluding remarks are given in section 5.
2. Perturbation Theory for $`\varphi ^3`$ on $`R^{1,3}\times T_\theta ^2`$
The noncommutative scalar $`\varphi ^3`$ theory in six noncompact dimensions is defined by the following action:
$$S=d^6x\left(\frac{1}{2}(\varphi )^2\frac{1}{2}m^2\varphi ^2\frac{\lambda }{3!}\varphi \varphi \varphi \right).$$
The coordinates of commutative four dimensional Minkowski space are $`x^0,x^1,x^2,x^3`$. The coordinates $`x^4`$ and $`x^5`$ are noncommuting :
$$[x^4,x^5]=i\theta .$$
Since the theory defined by (2.1) contains an infinite number of higher derivative operators, naively one would expect it to be nonrenormalizable. However, it is a renormalizable theory because all ultraviolet divergences can be cancelled by counterterms of the form appearing in the action (2.1). Unlike conventional field theories, the counterterms are not local operators but the nonlocality is of the same type present in the tree level action. The S-matrix elements of this theory are plagued with infrared divergences. For instance, the one loop correction to the two point function leads to a term in the effective action of the form
$$d^6x\frac{\lambda ^2}{2(4\pi )^3}\varphi \frac{1}{}\varphi ,$$
where we have introduced the inner product $`pp=p_M(\theta ^2)^{MN}p_N=\theta ^2(p_4^2+p_5^2)`$.
The action in (2.1) must be augmented by cutoff dependent counterterms. In this paper we work at lowest nontrivial order in perturbation theory (i.e. order $`\lambda ^2`$ in 1PI diagrams) and ignore the obvious classical instability in (2.1). The one loop counterterms in the noncompact theory are
$$S_{ct}=d^6x\left(\frac{\lambda ^2}{8(4\pi )^3}(\mathrm{\Lambda }^2m^2\mathrm{ln}\mathrm{\Lambda }^2)\varphi ^2+\frac{\lambda ^2}{48(4\pi )^3}\mathrm{ln}\mathrm{\Lambda }^2(\varphi )^2+\frac{\lambda ^3}{48(4\pi )^3}\mathrm{ln}\mathrm{\Lambda }^2\varphi \varphi \varphi \right).$$
The explicit form of the ultraviolet regulator $`\mathrm{\Lambda }`$ will be discussed later.
In this paper, we will be concerned with the theory on a space with two noncommuting compact directions. The divergence structure of the noncommutative theory on a compact space is different from that of the theory on a noncompact space. The coordinates $`x^4`$ and $`x^5`$ are compactified on a rectangular torus with $`0x^4,x^52\pi R`$. Momenta along the compact directions are quantized, $`\stackrel{}{p}=\stackrel{}{n}/R`$, where $`\stackrel{}{n}=(n_4,n_5)`$ are integers. For the mode with $`\stackrel{}{n}0`$, the term in the one particle irreducible action shown in (2.1) is finite. It gives rise to an interesting modification of the Kaluza-Klein spectrum that will be discussed in section 3. For the mode with $`\stackrel{}{n}=0`$, (2.1) is infinite.
Given the special role played by $`\stackrel{}{n}=0`$ mode, it is very useful to make the following definitions
$$\begin{array}{cc}\hfill \varphi _0& =\frac{1}{(2\pi R)^2}𝑑x^4𝑑x^5\varphi \hfill \\ \hfill \overline{\varphi }& =\varphi \varphi _0.\hfill \end{array}$$
The $`\varphi _0`$ field contains the mode with $`\stackrel{}{n}=0`$, $`\overline{\varphi }`$ contains all modes with nonvanishing $`\stackrel{}{n}`$. In terms of these fields the action (2.1) is:
$$S=d^6x\left(\frac{1}{2}(\overline{\varphi })^2\frac{1}{2}m^2\overline{\varphi }^2+\frac{1}{2}(\varphi _0)^2\frac{1}{2}m^2\varphi _0^2\frac{\lambda }{3!}\overline{\varphi }\overline{\varphi }\overline{\varphi }\frac{\lambda }{2}\overline{\varphi }^2\varphi _0\frac{\lambda }{3!}\varphi _0^3\right).$$
The products involving $`\varphi _0`$ are ordinary products. This is because after integration the $``$-product of three fields reduces to an ordinary product whenever one of the fields has vanishing momentum along noncommuting dimensions. No operator of the form $`d^6x\overline{\varphi }\varphi _0^2`$ appears because it is forbidden by momentum conservation.
Fig. 1: Feynman rules for noncommutative $`\varphi ^3`$ on $`R^{1,3}\times T_\theta ^2`$. Solid lines are $`\overline{\varphi }`$ quanta, dotted lines are $`\varphi _0`$ quanta. Momenta along noncompact directions are denoted by $`p`$, while $`\stackrel{}{n}/R,\stackrel{}{k}/R`$ and $`\stackrel{}{m}/R`$ are momenta along compact directions. The wedge product is defined to be $`\stackrel{}{n}\stackrel{}{k}n_4k_5n_5k_4`$.
The Feynman rules derived from the action (2.1) are given in fig. 1. Noncommutativity introduces oscillatory factors in the vertices of the theory. These oscillatory factors never make a graph more ultraviolet divergent than the naive power counting estimate. Therefore if a graph is ultraviolet finite in the commutative theory it will also be ultraviolet finite in the noncommutative theory. Since $`\varphi ^3`$ in six dimensions is renormalizable, only the two and three point functions of this theory can have ultraviolet divergences.
Fig. 2: One loop corrections to the two point functions for $`\varphi _0`$ and $`\overline{\varphi }`$.
The one loop contributions to the two point functions are shown in fig. 2. We begin with diagram $`a)`$, which is a contribution to the two point function of $`\overline{\varphi }`$.
$$a)=\frac{\lambda ^2}{2}\frac{1}{(2\pi R)^2}_\stackrel{}{k}\frac{d^4l}{(2\pi )^4}\frac{\mathrm{cos}^2(\theta \stackrel{}{n}\stackrel{}{k}/(2R^2))(1\delta _{\stackrel{}{k},0})(1\delta _{\stackrel{}{n}+\stackrel{}{k},0})}{(l^2\stackrel{}{k}^2/R^2m^2)((l+p)^2(\stackrel{}{n}+\stackrel{}{k})^2/R^2m^2)}.$$
In (2.1), $`l`$ denotes the loop momenta along the noncompact directions, while $`\stackrel{}{k}/R`$ is loop momenta along compact directions. Similarly, $`p(\stackrel{}{n}/R)`$ is the external momenta along the noncompact (compact) directions. In (2.1), $`\stackrel{}{n}\stackrel{}{k}n_4k_5n_5k_4`$. Since the external field is $`\overline{\varphi }`$, $`\stackrel{}{n}0`$. This implies $`\delta _{\stackrel{}{k},0}\delta _{\stackrel{}{n}+\stackrel{}{k},0}=0`$. We can simplify the numerator by using the half angle formula for the cosine and the fact that $`\stackrel{}{n}\stackrel{}{k}=0`$ for $`\stackrel{}{k}=0,\stackrel{}{n}`$ to obtain:
$$a)=\frac{\lambda ^2}{4}\frac{1}{(2\pi R)^2}_\stackrel{}{k}\frac{d^4l}{(2\pi )^4}\frac{12\delta _{\stackrel{}{k},0}2\delta _{\stackrel{}{n}+\stackrel{}{k},0}+\mathrm{cos}(\theta \stackrel{}{n}\stackrel{}{k}/R^2)}{(l^2\stackrel{}{k}^2/R^2m^2)((l+p)^2(\stackrel{}{n}+\stackrel{}{k})^2/R^2m^2).}$$
The oscillatory factor $`\mathrm{cos}(\theta \stackrel{}{n}\stackrel{}{k}/R^2)`$ makes the last term of (2.1) ultraviolet finite. The first term is quadratically divergent, while the second and third terms are logarithmically divergent. Only the divergent terms will be evaluated in this section.
We first represent the propagator using Schwinger parameters then perform the Gaussian integral over the loop momentum to obtain
$$\begin{array}{cc}\hfill a)=\frac{i\lambda ^2}{4(4\pi )^3\pi R^2}_0^{\mathrm{}}\frac{d\alpha }{\alpha }_0^1dx(\underset{\stackrel{}{k}}{}\mathrm{exp}[\alpha (m^2x(1x)p^2+\frac{\stackrel{}{k}^2}{R^2}+x\frac{\stackrel{}{n}^2+2\stackrel{}{n}\stackrel{}{k}}{R^2})]& \\ \hfill 4\mathrm{exp}[\alpha (m^2x(1x)p^2+x\frac{\stackrel{}{n}^2}{R^2})]).& \end{array}$$
The sum over $`\stackrel{}{k}`$ is performed using the definition of the Jacobi theta function
$$\vartheta (\nu ,\tau )=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\mathrm{exp}(\pi in^2\tau +2i\pi n\nu ),$$
and the modular transformation $`\vartheta (\nu ,\tau )=(i\tau )^{1/2}\mathrm{exp}(\pi i\nu ^2/\tau )\vartheta (\nu /\tau ,1/\tau )`$. The result is
$$\begin{array}{cc}\hfill \frac{i\lambda ^2}{4(4\pi )^3}\{_0^{\mathrm{}}\frac{d\alpha }{\alpha ^2}_0^1dx\mathrm{exp}[\alpha (m^2+x(1x)(p^2+\frac{\stackrel{}{n}^2}{R^2}))]\widehat{\vartheta }(xn_4)\widehat{\vartheta }(xn_5)& \\ \hfill \frac{4}{\pi R^2}_0^{\mathrm{}}\frac{d\alpha }{\alpha }_0^1dx\mathrm{exp}[\alpha (m^2x(1x)p^2+x\frac{\stackrel{}{n}^2}{R^2})]\}.& \end{array}$$
In this formula we have defined $`\widehat{\vartheta }(\nu )=\vartheta (\nu ,i\pi R^2/\alpha )`$. The ultraviolet divergent contribution comes from the $`\alpha 0`$ region and in this limit we can set the $`\widehat{\vartheta }`$ functions to unity. Then the first term is proportional to the quadratically divergent one loop correction in the commutative theory in the $`R\mathrm{}`$ limit. Corrections to this limit come from the $`n0`$ terms in the expansion of (2.1) and these are finite. There are also corrections arising because of the absence of the $`\stackrel{}{n}=0`$ mode in the loop integral. These corrections depend explicitly on the size of the compact dimension and are logarithmically divergent since the loop integral is effectively four dimensional for the zero mode.
The integrals can be regulated by inserting into the integral $`\mathrm{exp}(1/(\mathrm{\Lambda }^2\alpha ))`$, where $`\mathrm{\Lambda }`$ is the ultraviolet cutoff. The divergent contribution from diagram $`a)`$ in fig. 2 is:
$$a)=\frac{i\lambda ^2}{4(4\pi )^3}[\mathrm{\Lambda }^2(m^2+\frac{1}{6}(p^2+\frac{\stackrel{}{n}^2}{R^2})+\frac{4}{\pi R^2})\mathrm{ln}\mathrm{\Lambda }^2].$$
The divergences in the other graphs in fig. 2 can be computed in the same way. We find
$$\begin{array}{cc}\hfill b)=& \frac{i\lambda ^2}{(4\pi )^3}\frac{\mathrm{ln}\mathrm{\Lambda }^2}{\pi R^2},\hfill \\ \hfill c)=& \frac{i\lambda ^2}{2(4\pi )^3}\left[\mathrm{\Lambda }^2\left(m^2\frac{p^2}{6}+\frac{1}{\pi R^2}\right)\mathrm{ln}\mathrm{\Lambda }^2\right],\hfill \\ \hfill d)=& \frac{i\lambda ^2}{2(4\pi )^3}\frac{\mathrm{ln}\mathrm{\Lambda }^2}{\pi R^2}.\hfill \end{array}$$
Fig. 3: Divergent one loop corrections to the three point functions.
Next we consider the one loop divergences in the three point functions of the theory. If a zero mode appears in the loop then the loop integration is effectively four dimensional and the graph is finite by power counting. Therefore, it is only necessary to consider loop graphs with internal $`\overline{\varphi }`$ quanta. The relevant Feynman graphs are shown in fig. 3. The leading divergences are logarithmic and these can be obtained from a calculation treating the loop momentum as continuous. The divergences in the graphs of fig. 3 are
$$\begin{array}{cc}\hfill a)=& \frac{i\lambda ^3}{8(4\pi )^3}\mathrm{cos}\left(\frac{\theta }{2R^2}\stackrel{}{n}\stackrel{}{k}\right)\mathrm{ln}\mathrm{\Lambda }^2,\hfill \\ \hfill b)=& \frac{i\lambda ^3}{4(4\pi )^3}\mathrm{ln}\mathrm{\Lambda }^2,\hfill \\ \hfill c)=& \frac{i\lambda ^3}{2(4\pi )^3}\mathrm{ln}\mathrm{\Lambda }^2.\hfill \end{array}$$
The two and three point functions for $`\overline{\varphi }`$ are rendered finite by adding the counterterms in (2.1). These counterterms do not render correlation functions involving $`\varphi _0`$ finite in the compactified theory. After adding the counterterms in (2.1), the divergent part of the $`\varphi _0`$ self energy is
$$\mathrm{\Sigma }_0=\frac{\lambda ^2}{4(4\pi )^3}\left[\mathrm{\Lambda }^2\left(m^2\frac{p^2}{6}\right)\mathrm{ln}\mathrm{\Lambda }^2\right].$$
A similar uncancelled divergence appears in three point functions involving $`\varphi _0`$. The counterterm in (2.1) regulates three point involving three $`\overline{\varphi }`$ but not the divergences from diagrams $`b)`$ and $`c)`$ in fig. 3.
Fig. 4: Resummation of divergent terms in $`\varphi _0`$ two-point function. The shaded blob represents $`\mathrm{\Sigma }_0`$ in (2.1).
In the noncompact theory the two and three point functions contain infrared divergences. In the compactified theory, these infrared divergences appear as uncancelled cutoff dependence in two and three point functions involving $`\varphi _0`$. The correct interpretation of the uncancelled dependence on $`\mathrm{\Lambda }`$ is obtained by considering renormalization from the Wilsonian point of view. Instead of trying to introduce new counterterms, we interpret the ultraviolet cutoff as a physically significant scale. Then it is clear that the $`\varphi _0`$ mode, unlike the $`\overline{\varphi }`$ modes, develops a mass of order $`\mathrm{\Lambda }`$ and hence decouples from the low energy theory as $`\mathrm{\Lambda }\mathrm{}`$. This decoupling can be seen diagramatically. As $`\mathrm{\Lambda }\mathrm{}`$, $`\mathrm{\Sigma }_0\mathrm{}`$. When $`\mathrm{\Sigma }_0`$ becomes large it is necessary sum insertions of $`\mathrm{\Sigma }_0`$ in the $`\varphi _0`$ propagator to all orders, as shown in fig. 4. The two-point correlation function $`0|T(\varphi _0(x)\varphi _0(y))|0`$ is of order $`1/\mathrm{\Lambda }^2`$ and hence vanishes as $`\mathrm{\Lambda }^2\mathrm{}`$. Other correlation functions involving $`\varphi _0`$ also vanish.
In any correlation function, the resummation shown in fig. 4 is to be performed whenever $`\varphi _0`$ is an external line or for graphs which fall apart when a $`\varphi _0`$ propagator is cut. Furthermore, at the order in perturbation theory we are working this resummation is only performed for $`\varphi _0`$ propagators which do not appear inside 1PI subgraphs. For example, this resummation is not performed for internal $`\varphi _0`$ in diagrams $`b)`$ and $`d)`$ of fig. 2. This procedure gives finite answers for all correlation functions at this order in perturbation theory. There are uncancelled divergences proportional to $`\mathrm{ln}\mathrm{\Lambda }^2`$ in the 1PI three point functions with one or three external $`\varphi _0`$. However, the resummed propagators for the $`\varphi _0`$ fall off as $`1/\mathrm{\Lambda }^2`$.
Because of the sign of the correction in (2.1), the mass of the zero mode appears to be driven to negative, perhaps indicating that $`\varphi _0`$ is tachyonic. Unfortunately, it is impossible to address this issue in the one loop approximation we are employing in this paper. This is because higher order corrections to the self energy are of the form
$$\delta \mathrm{\Sigma }_0\lambda ^{2n+2}\mathrm{\Lambda }^2\mathrm{ln}^n\mathrm{\Lambda }.$$
Perturbation theory breaks down for $`\mathrm{\Lambda }\mathrm{}`$ and the sign of the quadratic divergence cannot be determined by lowest order perturbation theory. The zero mode mass is likely to be driven to $`\pm \mathrm{}`$, but the sign is not known from the one loop calculation we have performed in this section. At the next order in perturbation theory $`\mathrm{\Sigma }_0`$ will occur in loops that renormalize the masses of the $`\overline{\varphi }`$ modes. The cutoff dependence of these corrections must be cancelled by the counterterms if the low energy effective theory has dynamics for the modes of $`\overline{\varphi }`$ that is close to what is expected based on the one loop analysis.
This completes the analysis of the cutoff dependence of the compactified theory at the one loop level. All necessary counterterms are present in (2.1) and the additional divergences remove $`\varphi _0`$ from the low energy theory. An important outstanding issue is to see how ultraviolet divergences cancel at higher orders.
3. Kaluza-Klein Spectra on Noncommutative Tori
In this section we compute the spectrum of Kaluza-Klein modes that arises from (2.1) compactified on $`T_\theta ^2`$. In the classical limit, the $`\varphi `$ quanta of the commutative and noncommutative theories have the same dispersion relation. However, the nonlocality of noncommutative field theories leads to one loop corrections to the dispersion relation. In the noncompact case the one loop dispersion relation is of the form
$$p^2m^2+\frac{\lambda ^2}{(4\pi )^3}\frac{1}{pp}+\mathrm{}=0,$$
where the ellipsis denote terms that are less singular as $`pp0`$. The nonanalytic dependence in $`\theta `$ of (3.1) results in corrections that diverge for soft momenta along the noncommutative directions. After compactification, the discretization of momenta isolates the zero momentum mode of the field, such that only for this component of the field one obtains a divergence in the one loop dispersion relation (3.1). As shown in section $`2`$ this quantum correction drives the $`\varphi _0`$ mass to infinity, effectively removing $`\varphi _0`$ from the theory.
It is straightforward to compute the one loop dispersion relation for the non-zero modes. The relevant Feynman graphs that need to be evaluated are diagrams $`a)`$ and $`b)`$ of fig. 2. As shown in section $`2`$, the divergences in these graphs are absorbed by counterterms. We are interested in the finite corrections coming from the nonplanar piece of diagram $`a)`$, i.e., the term proportional to $`\mathrm{cos}(\theta \stackrel{}{n}\stackrel{}{k}/R^2)`$ in (2.1). This piece gives the leading behavior for small $`\theta `$. The one loop self energy is
$$\begin{array}{cc}\hfill \mathrm{\Sigma }=\frac{\lambda ^2}{4(4\pi )^3}_0^1𝑑x_0^{\mathrm{}}𝑑\alpha \alpha ^2\mathrm{exp}\left[\alpha \left(m^2+x(1x)\left(p^2+\frac{\stackrel{}{n}^2}{R^2}\right)\right)\frac{\theta ^2\stackrel{}{n}^2}{4R^2\alpha }\right]& \\ \hfill \times \left\{\frac{1}{2}\widehat{\vartheta }\left(xn_4+i\frac{\theta n_5}{2\alpha }\right)\widehat{\vartheta }\left(xn_5i\frac{\theta n_4}{2\alpha }\right)+\frac{1}{2}\widehat{\vartheta }\left(xn_4i\frac{\theta n_5}{2\alpha }\right)\widehat{\vartheta }\left(xn_5+i\frac{\theta n_4}{2\alpha }\right)\right\}.& \end{array}$$
Setting $`p^2\stackrel{}{n}^2/R^2=m^2`$ in the above formula, the leading terms for small $`\theta `$ are
$$\mathrm{\Sigma }=\frac{\lambda ^2}{(4\pi )^3}\left(\frac{R^2}{\theta ^2\stackrel{}{n}^2}+\frac{5}{24}m^2\mathrm{ln}\left(\frac{m^2\theta ^2\stackrel{}{n}^2}{R^2}\right)+\mathrm{}\right),$$
so that the formula for the Kaluza-Klein masses is
$$m_\stackrel{}{n}^2=m^2+\frac{\stackrel{}{n}^2}{R^2}\frac{\lambda ^2}{(4\pi )^3}\left(\frac{R^2}{\stackrel{}{n}^2\theta ^2}+\frac{5}{24}m^2\mathrm{ln}\left(\frac{m^2\theta ^2\stackrel{}{n}^2}{R^2}\right)+\mathrm{}\right),$$
Spatial noncommutativity in the compact directions gives a correction to the Kaluza-Klein mass formula which resembles that of winding states in string theory. Note that the mass correction from the nonplanar diagram is negative. For sufficiently small $`\theta `$, this correction induces a negative mass squared for some Kaluza-Klein modes. The ellipses in Eqs. (3.1) and (3.1) denote terms that are less important as $`\theta 0`$.
4. Noncommutative $`U(1)`$ Gauge Theory in Six Dimensions
In a noncommutative space-time a pure $`U(1)`$ gauge theory is interacting. The action in six dimensions is
$$S=\frac{1}{4}d^6xF_{MN}F^{MN},$$
where the field strength tensor
$$F_{MN}=_MA_N_NA_Mig(A_MA_NA_NA_M),$$
transforms as
$$\delta _\alpha F_{MN}=ig(\alpha F_{MN}F_{MN}\alpha )$$
under the $`U(1)`$ gauge transformation
$$\delta _\alpha A_M=_M\alpha +ig(\alpha A_MA_M\alpha ).$$
Terms in the action quadratic in $`A`$ are the same as in the commutative theory so the tree level spectrum of Kaluza-Klein excitations follows that case. These are massive four vectors, labelled by $`A_\mu ^{(\stackrel{}{n})}`$, and scalars, labeled by $`h^{(\stackrel{}{n})}=(n_5A_4^{(\stackrel{}{n})}n_4A_5^{(\stackrel{}{n})})/\sqrt{\stackrel{}{n}^2}`$, with masses
$$m_\stackrel{}{n}^2=\frac{\stackrel{}{n}^2}{R^2}.$$
Note that the orthogonal scalar component, $`g^{(\stackrel{}{n})}=(n_4A_4^{(\stackrel{}{n})}+n_5A_5^{(\stackrel{}{n})})/\sqrt{\stackrel{}{n}^2}`$, can be gauged away. It has become the longitudinal component of the massive four vectors $`A_\mu ^{(\stackrel{}{n})}`$ through the standard Higgs mechanism.
As in the case of scalars discussed in Section 2 the $`\stackrel{}{n}=0`$ modes need to be treated differently. They consist of a four dimensional gauge field and two massless scalars. In this case the zero mode fields do not interact with each other nor do they couple to the $`\stackrel{}{n}0`$ modes. There are no one loop graphs contributing to the two point functions of these fields in the compactified theory. However, these fields receive a divergent wavefunction renormalization from the counterterms of the noncompact theory. This divergence can be removed simply by making a field redefinition. If currents charged under the U(1) gauge symmetry were added to the theory, the field redefinition would result in a suppression of the couplings of the gauge field to these currents. The $`\stackrel{}{n}=0`$ modes decouple from the low energy effective field theory, not because their masses are driven to infinity (as in the case of the scalar theory), but because their couplings to other fields are driven to zero.
The perturbative corrections to the masses of the Kaluza-Klein modes with $`\stackrel{}{n}0`$ follow from the two point function $`\mathrm{\Pi }_{MN}`$ of the gauge field. As in the case of the scalar field theory considered in the last section, the most important corrections for $`\theta <R^2`$ can be obtained by replacing the discrete sum over compact loop momenta with a continuum integral. This yields
$$\mathrm{\Pi }_{MN}=\mathrm{\Pi }_{MN}^p+\mathrm{\Pi }_{MN}^{np},$$
where <sup>1</sup> $`\mathrm{\Pi }_{MN}`$ is calculated in background field gauge. This amounts to replacing $`A_M`$ in Eq. action by $`Q_M+A_M`$ and adding the gauge fixing term $`1/2d^6xG^2`$, where $`G=_MQ^Mig(A_MQ^MQ_MA^M)`$. Note that we are using the convention $`g_{\mu \nu }=\mathrm{diag}(+,,,,,)`$.
$$\mathrm{\Pi }_{MN}^p=\frac{ig^2}{(4\pi )^3}_0^{\mathrm{}}\frac{d\alpha }{\alpha ^2}_0^1𝑑x\mathrm{exp}[p^2\alpha x(1x)]\left[(p^2g_{MN}p_Mp_N)(84(12x)^2)\right],$$
and
$$\begin{array}{cc}\hfill \mathrm{\Pi }_{MN}^{np}=\frac{ig^2}{(4\pi )^3}_0^{\mathrm{}}\frac{d\alpha }{\alpha ^2}_0^1𝑑x& \mathrm{exp}\left(p^2\alpha x(1x)\frac{pp}{4\alpha }\right)\hfill \\ & \times \left[(p^2g_{MN}p_Mp_N)(84(12x)^2)\frac{4}{\alpha ^2}\stackrel{~}{p}_M\stackrel{~}{p}_N\right].\hfill \end{array}$$
Here $`\stackrel{~}{p}_M=\theta _{MN}p^N`$ and $`p`$ is the six dimensional external momentum. The components of external momentum along the compact directions, $`p_4`$ and $`p_5`$, are quantized.
The first term inside the brackets in (4.1) gives rise to wavefunction renormalization for the massive vector modes $`A_\mu ^{(\stackrel{}{n})}`$ and does not change their masses. However, the term proportional to $`\stackrel{~}{p}_M\stackrel{~}{p}_N`$ can affect the masses of the scalars $`h^{(\stackrel{}{n})}`$. Using the same approximations that were used in the scalar case we find that the Kaluza-Klein spectrum for the $`h^{(\stackrel{}{n})}`$ at one loop is
$$m_\stackrel{}{n}^2=\frac{\stackrel{}{n}^2}{R^2}\frac{8g^2R^4}{\pi ^3\theta ^4(\stackrel{}{n}^2)^2}+\mathrm{}.$$
Again for small $`\theta `$ the second term in (4.1) can dominate over the first and drives $`m_\stackrel{}{n}^2`$ negative. Higher order terms in the effective potential for $`h^{(\stackrel{}{n})}`$ must be computed to determine whether this results in a vacuum expectation value for the $`h^{(\stackrel{}{n})}`$ or signals a true instability of the theory.
5. Conclusion
It is a widely held belief that there are additional compactified spatial dimensions. Extra dimensions occur in string theory and it has even been conjectured that their presence may lead to a resolution of the hierarchy puzzle or cosmological constant problem. If extra spatial dimensions do exist then it is possible that they are noncommuting. Motivated by this possibility we have explored the properties of some simple theories with noncommuting compact extra dimensions.
We considered, at lowest order perturbation theory, a $`\varphi ^3`$ theory compactified on a rectangular noncommuting two torus, $`T_\theta ^2`$. The UV-IR mixing that occurs in the uncompactified case has important implications for the compactified theory. It causes the mode of the $`\varphi `$ field that is constant on the torus (i.e. the zero mode) to have different dynamics than the other modes. Quantum corrections drive correlation functions of this mode to zero. It does not appear as a propagating degree of freedom in the effective four dimensional theory.
The mass spectrum of Kaluza-Klein excitations was calculated in this theory. At the one loop level the masses get a contribution that resembles that of winding states in string theory. This contribution to the square of the mass is negative and for small values of the noncommuting parameter it could dominate over the tree level term.
We also examined a U(1) gauge theory compactified on $`T_\theta ^2`$. Similar phenomena to those that occur in the scalar case were found.
Acknowledgments
We would like to thank Hirosi Ooguri for reading an earlier draft of this paper and Edward Witten for useful comments. J.G. would like to thank CERN for hospitality during the final stages of this work. J.G. and T.M. are supported in part by the DOE under grant no. DE-FG03-92-ER 40701.
References
relax S. Minwalla, M.V. Raamsdonk and N. Seiberg, “Noncommutative Perturbative Dynamics”, hep-th/9912072. relax M.V. Raamsdonk and N. Seiberg, “ Comments on Noncommutative Perturbative Dynamics”, hep-th/0002186. relax M. Hayakawa,“Perturbative analysis on infrared and ultraviolet aspects of noncommutative QED on $`R^4`$,” hep-th/9912167. relax W. Fischler, E. Gorbatov, A. Kashani-Poor, S. Paban, P. Pouliot and J. Gomis, “Evidence for winding states in noncommutative quantum field theory,” hep-th/0002067. relax A. Matusis, L. Susskind and N. Toumbas, “The IR/UV connection in the non-commutative gauge theories,” hep-th/0002075. relax T. Filk, “Divergences in a Field Theory on Quantum Space”, Phys. Lett. B376 53 (1996). relax J.C. Varilly and J.M. Gracia-Bondia, “On the ultraviolet behaviour of quantum fields over noncommutative manifolds”, Int. J. Mod. Phys. A14 (1999) 1305, hep-th/9804001. relax M. Chaichian, A. Demichev and P. Presnajder, “Quantum Field Theory on Noncommutative Space-times and the Persistence of Ultraviolet Divergences”, hep-th/9812180; “Quantum Field Theory on the Noncommutative Plane with E(q)(2) Symmetry”, hep-th/9904132. relax C.P. Martin, D. Sanchez-Ruiz, “The One-loop UV Divergent Structure of U(1) Yang-Mills Theory on Noncommutative $`R^4`$”, Phys. Rev. Lett. 83 (1999) 476-479,hep-th/9903077 relax M. Sheikh-Jabbari, “One Loop Renormalizability of Supersymmetric Yang-Mills Theories on Noncommutative Torus”, JHEP 06 (1999) 015, hep-th/9903107; “Noncommutative Super Yang-Mills Theories with 8 Supercharges and Brane Configurations”, hep-th/0001089. relax I.Ya. Areféva, D.M. Belov and A.S. Koshelev, “Two-Loop Diagrams in Noncommutative $`\varphi _4^4`$ theory”, hep-th/9912075; “A Note on UV/IR for Noncommutative Complex Scalar Field”, hep-th/0001215; “UV/IR Mixing for Noncommutative Complex Scalar Field Theory, II (Interaction with Gauge Fields)”, hep-th/0003176. relax H. Grosse, T. Krajewski and R. Wulkenhaar, “Perturbative quantum gauge fields on the noncommutative torus”, hep-th/9903187; “Renormalization of noncommutative Yang-Mills theories: A simple example”, hep-th/0001182. relax S. Cho, R. Hinterding, J. Madore and H. Steinacker, “Finite Field Theory on Noncommutative Geometries”, hep-th/9903239. relax E. Hawkins,“Noncommutative Regularization for the Practical Man”,hep-th/9908052. relax I. Chepelev and R. Roiban, “Renormalization of Quantum Field Theories on Noncommutative $`R^d`$, I. Scalars,” hep-th/9911098. relax H.O. Girotti, M. Gomes, V.O. Rivelles and A.J. da Silva, “A Consistent Noncommutative Field Theory: the Wess-Zumino Model”, hep-th/0005272. relax S.S. Gubser and S.L. Sondhi, “Phase structure of non-commutative scalar field theories”, hep-th/0006119. relax J. Gomis and T. Mehen, ”Space-Time Noncommutative Field Theories and Unitarity”, hep-th/0005129. relax A. Connes, M.R. Douglas and A. Schwarz, “Noncommutative Geometry and Matrix Theory: Compactification on Tori”, JHEP 9802(1998) 003, hep-th/9711162. relax M.R. Douglas and C. Hull, “D-branes and the Noncommutative Torus”, JHEP 9802 (1998) 008, hep-th/9711165. relax N. Seiberg and E. Witten, “String Theory and Noncommutative Geometry”, JHEP 9909 (1999) 032, hep-th/9908142. relax Y.-K. E. Cheung and M. Krogh, “Noncommutative Geometry From 0-Branes In A Background B Field”, Nucl. Phys. B528 (1998) 185, hep-th/ 98030031. relax 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. relax V. Schomerus, “D-Branes And Deformation Quantization”, JHEP 9906:030 (1999), hep-th/9903205. relax F. Ardalan, H. Arfaei and M.M. Sheikh-Jabbari, “Mixed Branes and M(atrix) Theory on Noncommutative Torus”, hep-th/9803067; “Noncommutative Geometry From Strings and Branes”, JHEP bf 9902 (1999), hep-th/9810072; “Dirac Quantization of Open Strings and Noncommutativity in Branes”, hep-th/9906161. relax N. Arkani-Hamed, S. Dimopoulos and G. Dvali, “The Hierarchy Problem and New Dimensions at a Millimeter”, hep-ph/98033115, Phys. Lett. B429 (1998) 263; I. Antoniadis, N. Arkani-Hamed, S. Dimopoulos and G. Dvali, “New Dimensions at a Millimeter to a Fermi and Superstrings at a TeV”, Phys. Lett. B436 (1998) 257; I. Antoniadis, “A Possible New Dimension at a Few TeV”, Phys. Lett. B246 (1990) 377; N. Arkani-Hamed, S. Dimopoulos and J. March-Russell, “Stabilization of Submillimeter Dimension: The New Guise of the Hierarchy Problem”, hep-th/9809124; L. Randall and R. Sundrum, “A Large Mass Hierarchy from a Small Extra Dimension”, hep-th/99050221, Phys. Rev. Lett. 83 (1999) 3370. relax N. Arkani-Hamed, S. Dimopoulos, N. Kaloper and R. Sundrum, “A Small Cosmological Constant from a Large Extra Dimension”, hep-th/00031197, Phys. Lett. B480 (2000) 193; S. Kachru, M. Schulz and E. Silverstein, “Self Tuning Flat Domain Walls in 5-D Gravity and String Theory”, hep-th/0001206. relax J.Gomis, M. Kleban, T. Mehen, M. Rangamani and S. Shenker, “Noncommutative Gauge Dynamics from the String World Sheet”, hep-th/0003215.
|
warning/0006/cond-mat0006327.html
|
ar5iv
|
text
|
# Has the FFLO state been observed in the organic superconductor limit-from𝜅-(BEDT-TTF)₂Cu(NCS)₂ ?
## References
|
warning/0006/hep-th0006036.html
|
ar5iv
|
text
|
# Untitled Document
hep-th/0006036 LPTHE-00-20 NYU-TH-30.5.00
Gauge and Topological Symmetries in the Bulk Quantization of Gauge Theories
Laurent Baulieu<sup>⋆†§</sup>, Pietro Antonio Grassi and Daniel Zwanziger
baulieu@lpthe.jussieu.fr, Daniel.Zwanziger@nyu.edu, pag5@nyu.edu
LPTHE, Universités P. & M. Curie (Paris VI) et D. Diderot (Paris VII), Paris, France,
Enrico Fermi Inst. and Dept. of Physics, University of Chicago, Chicago, IL 60637, USA
Physics Department, New York University, New-York, NY 10003, USA
A gauge theory with 4 physical dimensions can be consistently expressed as a renormalizable topological quantum field theory in 5 dimensions. We extend the symmetries in the 5-dimensional framework to include not only a topological BRST operator $`s`$ that encodes the invisibility of the “bulk” (the fifth dimension), but also a gauge BRST operator $`w`$ that encodes gauge-invariance and selects observables. These symmetries provide a rich structure of Ward identities which assure the renormalizability of the theory, including non-renormalization theorems. The 5-dimensional approach considerably simplifies conceptual questions such as for instance the Gribov phenomenon and fermion doubling. A confinement scenario in the 5-dimensional framework is sketched. We detail the five-dimensional mechanism of anomalies, and we exhibit a natural lattice discretization that is free of fermion doubling.
1. Introduction
In we gave arguments for the relevance of a five-dimensional representation of gauge theories in four-physical dimensions. We proved the existence of a local quantum field theory that is perturbatively renormalizable by power counting and free of the Gribov ambiguity. We also gave its lattice formulation, and suggested that the 5-dimensional framework naturally avoids fermion doubling. A possible interpretation is that the fifth-dimension is the stochastic time that Parisi and Wu proposed a long time ago for stochastically quantizing the Yang–Mills theory , and the 4-dimensional physical theory lives in any given chosen time-slice of the space with five dimensions. However the beauty of the resulting theory suggests that the fifth time plays a more fundamental and more general role for example in elucidating the Gribov and fermion doubling problems. In particular we consider the 5-dimensional functional integral to be more fundamental than the Langevin or Fokker-Planck equations.
After the discovery of the four-dimensional Yang–Mills topological quantum field theories , it was realized that the supersymmetric formulation of stochastic quantization also determines a topological field theory . This was an early example of a physical theory that lives in the “boundary” of a space with an additional dimension, independently of the process in the “bulk” that determines the many possible ways the a Fokker–Planck distribution converges to an equilibrium distribution. One recognizes a holographic phenomenon at work here, and various connections between different theories have been exhibited in this way, with the idea that stochastic quantization is analogous to a Stokes theorem for the path integral . With the inclusion of fermions, and BRST ghosts, the local 5-dimensional formulation transcends its purely stochastic origin. From now on, we call it bulk quantization.
One must establish the consistency of quantization with an additional time, and the four-dimensional Yang–Mills theory is an important and challenging case. Long before the invention of topological field theory, it was shown that the Parisi–Wu conjecture is compatible at the perturbative level with the Faddeev–Popov method . For reviews of stochastic quantization, see , and . Further development of stochastic quantization, particularly gauge-invariant stochastic regularization, may be found in , and , and for renormalization in . For renormalization of non-gauge theories in the 5-dimensional formulation, see .
The hope of enriching our perspectives is of course at the non-perturbative level. The 5-dimensional theory possesses a supersymmetry of the topological type which ensures that the expectation-value, once equilibrium is achieved, is independent of the details of the initial conditions. The existence of an unobservable fifth time considerably simplifies many conceptual problems that perplex the sole and too narrow four-dimensional perspective. For example the local quantum field theory in 5 dimensions avoids the question of Gribov copies that jeopardizes the Faddeev–Popov prescription in four dimensions . As shall be discussed elsewhere, the introduction of the additional time allows one to replace problematic gauge fixing by an appropriate gauge transformation in such a way that the Gribov question becomes irrelevant. The idea of looking at gauge theories from five dimensions can be put in correspondence with the description of conformal theories from the Chern–Simons action in three dimensions.
Our definition of observables in was not entirely satisfactory to the extent that it was not based on an invariance principle. Here we will fill this hole in our presentation, and introduce, in addition to the supersymmetry operator $`s`$ that expresses quantization with a fifth time , another BRST-symmetry operator, $`w`$, that implements gauge symmetry in 5 dimensions and which is compatible with $`s`$. This necessitates introducing an additional field, in a way that is inspired by an idea originally due to Horne in the context of the topological Yang–Mills theory in four dimensions, and emphasized in subsequent works using the idea of equivariant cohomology . The 5-dimensional action of the Yang–Mills theory is entirely defined by the requirement of $`w`$ and $`s`$ symmetry. We will define observables by the cohomology of $`w`$ taken at a fixed but arbitrary time-slice.
There are two features of the 5-dimensional formulation of 4-dimensional quantum field theory – whether of $`\varphi ^4`$ or of Yang-Mills type – which we wish to emphasize. The first is that even though the action is $`s`$-exact, $`I=sX`$, where $`s`$ has all possible characteristics of a topological BRST operator, we are not in the context of a topological theory of the usual type, as was noted in . The reason is that the observables $`O`$ are not required to be $`s`$-invariant, $`sO0`$. Indeed the cohomology of $`s`$ is empty, so if the observables were $`s`$-invariant, they would be $`s`$-closed, $`O=sY`$, and would have vanishing expectation-value, because our theory has no zero modes that would allow $`<sY>0`$. Rather, in the case of a scalar theory, the observables are all possible correlators taken at equal time, and in the case of gauge theories, they are also required to be in the cohomology of another BRST operator $`w`$, such that $`w^2=sw+ws=0`$. Beyond technical details and subtleties, $`w`$ is the expression of the gauge symmetry in the five-dimensional framework. The five-dimensional theory is thus not a topological theory, although its $`s`$-exact action looks topological. Rather, our interpretation is that the latter property is the simplest way to assure that possible renormalization constants are the same as in the 4-dimensional theory. The second feature, which holds both for gauge and scalar theories, is that the correlators are distributions in 5 dimensions, and in general they have singularities at equal times. As a result, physical observables, which are restricted to a time slice, may not be well-defined. Indeed, as shown in Appendix E, the correlator of three chiral currents, $`j(x_1,t_1)j(x_2,t_2)j(x_3,t_3)`$, is ambiguous in the equal-time limit, and this is the origin of the triangle anomaly in the 5-dimensional formulation. We expect that this kind of obstruction to consistent equal-time limits for the observables occurs only when the gauge theory is anomalous.
We will explain in a separate paper that the condition of fixed time relies on the correspondence of Schwinger–Dyson equations in 4 and 5 dimensions.
The 5-dimensional formulation accommodates in a local description the gauges that are actually used at present in lattice gauge theory for numerical evaluation. These gauges, such as the minimal Landau, minimal Coulomb or maximal Abelian gauges, are fixed by minimizing an appropriately chosen functional. They cannot be correctly described by the Faddeev–Popov method that is characterized by a local gauge condition, such as $`_\lambda A_\lambda =0`$, that does not distinguish between minima and saddle points of the minimizing functional. Nevertheless these gauges are represented by a local action in 5 dimensions. We would like to emphasize that the 5-dimensional local theory does not reproduce the gauge-non-invariant part of the standard 4-dimensional Faddeev–Popov formulation. The latter can only be reached from the 5-dimensional theory by means of a non-local action , and then only at the perturbative level.
Instead of a local gauge-fixing in 4 dimensions which is known not to exist , the 4-dimensional probability distribution is obtained from the solution of a Fokker–Planck equation in 5 dimensions that preserves the 4-dimensional probability at each instant. The solution of the Fokker–Planck equation, which determines all gauge-invariant correlation functions in 4 dimensions is represented by a local 5-dimensional gauge theory of topological type. Its path integral formula is valid non-perturbatively, and we have previously given its BRST-invariant lattice regularization .
The organization of the paper is as follows. The basic formulation of the theory is presented in sec. 2. Here the field content is explained, the $`s`$\- and $`w`$-symmetries are defined, and the most general renormalizable $`s`$ and $`w`$-invariant action is exhibited. It is important in this regard to keep in mind that the fifth time has engineering dimenison double of that of ordinary space-time coordinates $`[/t]=2[/x_\mu ]=2`$. In sec. 3, we briefly sketch a physical interpretation of the theory. A number of further developments of the theory and various issues are explored in the Appendices. In Appendix A, we derive the Ward identities which assure the stability of the theory under renormalization and fix a number of renormalization constants. In particular it is shown that $`gA_5`$ is invariant under renormalization in the minimal Landau gauge, and thus may be used to define an invariant charge in QCD. In Appendix B, properties of the minimal Landau-gauge are derived. This gauge provides the frame for a confinement scenario that is described in Appendix C, in which long-range “forces” are transmitted by $`A_5`$. In Appendix D, the theory is extended to Dirac spinor fields. The important topic of anomalies is addressed in Appendix E. Here a 5-form is found which is a candidate for an obstruction in the 5-dimensional theory. We show however that the familiar triangle anomaly of the 4-dimensional theory has another origin: it is a singularity that appears when fifth times are set equal, as is necessary to obtain the physical 4-dimensional correlators. In Appendix F, we show that fermion doubling may be avoided by lattice discretization of the 5-dimensional theory.
2. Field content and symmetries
We first focus on the pure Yang–Mills case, and will introduce coupling to spinors later. The 5-dimensional action used in
$$\begin{array}{cc}\hfill I=𝑑x^\mu 𝑑x^5s\mathrm{Tr}\left(\overline{\mathrm{\Psi }}^\mu (F_{5\mu }D_\lambda F_{\lambda \mu }\frac{1}{2}b_\mu )+\mathrm{}\right)& \end{array}$$
$`(2.1)`$
is $`s`$-exact, and $`\mathrm{}`$ will be specified shortly. (2.1) looks like a topological action. Indeed, $`s`$ is a topological BRST operator that will be defined shortly, as well as all relevant fields. Greek indices denote 4-dimensional Euclidean components, $`\lambda ,\mu =1,\mathrm{}4`$. The first term in (2.1) is invariant under 5-dimensional gauge transformations. Roughly speaking, it concentrates the path integral around the solutions of the equation $`F_{5\mu }D_\lambda F_{\lambda \mu }=0`$. The five-dimensional gauge symmetry of the action will be broken in a BRST-invariant way by means of
$$\begin{array}{cc}\hfill aA_5=_\lambda A_\lambda ,& \end{array}$$
$`(2.2)`$
where $`a`$ is a gauge parameter. This condition should not be interpreted as a Faddeev–Popov gauge-fixing. As will be discussed elsewhere, $`A_5=\dot{g}g^1`$ is the generator of a time-dependent gauge transformation $`g(x,t)`$ that acts on $`A_\mu `$ for $`\mu =1,\mathrm{}4`$, , and as such it can be fixed arbitrarily, apart from the constraints imposed by renormalizability. Indeed the role of $`A_5`$ as the generator of a time-dependent gauge transformation survives the algebraic quantization of the 5-dimensional theory, as is shown in Appendix A. The successful avoidance of the Gribov problem is reflected in the fact that the ghost propagators are parabolic when (2.2) is enforced in a BRST-invariant way, so the ghost propagators are retarded, $`G(x,t)=0`$ for $`t<0`$, and consequently closed ghost loops vanish and the Faddeev–Popov determinant is trivial. Heuristically, one recognizes that (2.2) is like an axial gauge in $`A_5`$, for which there is no Gribov ambiguity, and the infinite range of the variable $`t=x_5`$ avoids the Singer theorem .
When the ghost and auxiliary fields are integrated out, one obtains the action of stochastically quantized gauge theory ,
$$\begin{array}{cc}\hfill I=& 𝑑x^\mu 𝑑x^5\mathrm{Tr}\left(F_{5\mu }^2+(D_\lambda F_{\lambda \mu })^2\right)\hfill \end{array}$$
$`(2.3)`$
with stochastic gauge-fixing .
The observables of the theory were defined in in the following intuitive way: they are correlation functions of gauge-invariant functions of the gauge field components, $`A_\mu `$, for $`\mu =1,\mathrm{}4`$, taken at equal values of $`t=x_5`$. The Green functions can be first computed and renormalized at different values of time, and then one takes the limit of equal times. From the point of view of quantum field theory the fifth time appears as a regulator and from a geometrical point of view of topology, it appears as a variable that enlarges the space and simplifies topological properties. We shall refine the definition of observables here, and define observables from a symmetry principle, in order to have a better control of their renormalization properties. This will lead us to refine our knowledge of the symmetry of (2.1). We want to make this notion precise, and end up with the definition of all observables in the cohomology of a certain symmetry operator, $`w`$, which is compatible with $`s`$.
We have shown in that the action (2.1) gives a perturbatively renormalizable theory. Although the action (2.1) is not $`SO(5)`$ invariant, its BRST symmetry is $`SO(5)`$ invariant, which is sufficient for a consistent description, since the five-dimensional description is holographic, and the only things that matters is to recovers the $`SO(4)`$ invariance for the observables. Power-counting constrains the way the invariance under $`SO(5)`$ symmetry is reduced down to $`SO(4)`$: indeed, the canonical dimension in mass units of each one of the four-dimensional components of the gauge field is unity whereas $`A_5`$ has dimension two. (The Yang–Mills coupling constant has dimension zero).
Let us now define all the fields, which depend on $`x^\mu `$ and $`t=x_5`$ and are Lie algebra valued. New fields will occur, as compared to because of the $`w`$-symmetry. The bose and fermi fields are in one-to-one correspondence, and are related by the topological BRST operator $`s`$, as in a supersymmetric theory. (This suggests the possibility of a link between the topological BRST operator $`s`$ and some kind of Poincaré twisted supersymmetry.) In addition to $`A_\mu `$ and $`A_5`$ there are: $`c`$, which has the quantum numbers the ordinary Faddeev–Popov ghost but plays a different role; $`\mathrm{\Psi }_\mu `$ and $`\mathrm{\Psi }_5`$, the topological Fermi vector ghosts corresponding to $`A_\mu `$ and $`A_5`$; and $`\mathrm{\Phi }`$, a commuting ghost of ghost. These fields have respectively ghost number $`N_s=1,1,1`$ and $`2`$. We also have corresponding anti-ghosts $`\overline{c}`$, $`\overline{\mathrm{\Psi }}_\mu `$ and $`\overline{\mathrm{\Phi }}`$, with ghost number $`N_s=1,1`$ and $`2`$ respectively, and Lagrange multipliers $`b_\mu `$, $`l`$ and $`\overline{\eta }`$, with ghost number $`N_s=0,0`$ and $`1`$ respectively. (In we used $`\overline{c}=\overline{\mathrm{\Psi }}_5`$ and $`l=b_5`$.) See the diagram below for the relations of these fields and their quantum numbers.
There are five degrees of freedom for the choice of a dynamics, one for each component of $`A`$. The Lagrange multiplier field $`l`$ serves to impose the condition (2.2), and the $`b_\mu `$’s enforce the Langevin equation. Finally $`\overline{\eta }`$ is a fermion Lagrange multiplier for the gauge fixing of the longitudinal modes in $`\mathrm{\Psi }`$.
The topological BRST operator $`s`$ is not relevant for defining observables. Indeed, the cohomology of $`s`$ with ghost number zero is empty. Although it involves the gauge symmetry with a ghost of ghost phenomenon to ensure its nilpotency, it only represents the irrelevance of the details of the process in the bulk.
To encode gauge-invariance and distinguish gauge-invariant observables, we need another BRST operator, which we call $`w`$. We thus introduce a second ghost number $`N_w`$ and new fields. The total grading is the sum $`N_s+N_w+p`$, where $`p`$ is the ordinary form degree. We need a second Faddeev–Popov ghost $`\lambda `$ and, as will be explained shortly, also its ghost of ghost $`\mu `$. The ghost-field $`\lambda `$ plays the role of the ordinary Faddeev–Popov ghost, and in the “effective” theory on the boundary, the $`w`$ operator induces the ordinary BRST symmetry. One actually has a quartet of additional ghosts and anti-ghosts $`\lambda `$, $`\mu `$, $`\overline{\lambda }`$, $`\overline{\mu }`$, which maintains supersymmetry. All the fields and ghosts besides $`\lambda `$, $`\mu `$, $`\overline{\lambda }`$ and $`\overline{\mu }`$ have ghost number $`N_w=0`$. The latter have respectively ghost number $`N_w=1,1,1,1`$, and $`N_s=0,1,0,1`$. It is convenient to use a bigrading notation $`\varphi ^{(N_s,N_w)}`$ that indicates the ghost numbers of any given field $`\varphi `$. With this notation one has $`(s\varphi )^{(N_s+1,N_w)}`$ and $`(w\varphi )^{(N_s,N_w+1)}`$, $`(ws\varphi )^{(N_s+1,N_w+1)}`$. All ghost numbers can be read from the following diagram:
$$\begin{array}{cc}\hfill \begin{array}{c}\begin{array}{ccccc}& & & & A_\mu ^{(00)},A_5^{(00)}\\ & & & & & \\ & & \mathrm{\Psi }_\mu ^{(1,0)},\mathrm{\Psi }_5^{(1,0)}& & & & \overline{\mathrm{\Psi }}_\mu ^{(1,0)}\\ & & c^{(1,0)},\lambda ^{(0,1)}& & & & \overline{c}^{(1,0)}\\ & & & & & & \overline{\lambda }^{(0,1)}\\ & & & & & & & \\ \mathrm{\Phi }^{(2,0)},\mu ^{(1,1)}& & & & b_\mu ^{(0,0)}& & & & \overline{\mathrm{\Phi }}^{(2,0)},\overline{\mu }^{(1,1)}\\ & & & & l^{(0,0)}& & & & \\ & & & & & & & \\ \\ & & & & & & \overline{\eta }^{(1,0)}\end{array}\end{array}& \end{array}$$
$`(2.4)`$
The engineering dimensions (in mass units) of the fields are assigned according to:
$$\begin{array}{cc}& [c]=[\lambda ]=[\mu ]=[\mathrm{\Phi }]=0\hfill \\ & [\frac{}{x^\mu }]=[A_\mu ]=[\mathrm{\Psi }_\mu ]=1\hfill \\ & [\frac{}{x^5}]=[A_5]=[\mathrm{\Psi }_5]=2\hfill \\ & [b_\mu ]=[\overline{\mathrm{\Psi }}_\mu ]=3\hfill \\ & [l]=[\overline{c}]=[\overline{\mu }]=[\overline{\lambda }]=[\overline{\mathrm{\Phi }}]=[\overline{\eta }]=4.\hfill \end{array}$$
$`(2.5)`$
These values will play a key role in the determination of the five-dimensional action. The asymetry in the dimensions of the fields with ordinary labels $`\mu `$ and $`5`$ is absolutely crucial, and will explain how the good properties of the power counting in the conventional four-dimensional formulation of gauge theories are still present in the five-dimensional formulation. Moreover, since the anti-ghosts have dimension 3 or 4 and $`/t`$ has dimension 2, all ghost actions will be parabolic and all ghost propagators will be retarded.
In (2.4), the arrows conveniently relate the fields that can be transformed into each other by $`s`$\- or $`w`$-transformations that we will give shortly, in accordance with the separate conservation of the ghost numbers $`N_s`$ and $`N_w`$. The diagram (2.4) exhibits the boson-fermi pairing which always occurs in theories with a “topological” symmetry. The subtlety is the way the symmetry is expressed, with a delicate separation of the gauge symmetry transformations from the topological transformations, expressed respectively by the $`w`$\- and $`s`$-transformation. By convention, since the 2 ghost numbers are conserved, we assign dimension in such a way that both $`s`$ and $`w`$ leave dimension unchanged.
We now define the action of $`s`$ and $`w`$.<sup>1</sup> A more symmetric diagram would involve the idea of anti-$`s`$ and anti-$`w`$ symmetries We require
$$\begin{array}{cc}\hfill w^2=s^2=sw+ws=0.& \end{array}$$
$`(2.6)`$
This property is automatically assured by defining the symmetry in the following geometrical way, which is by now a standard in the BRST paradigm:
$$\begin{array}{cc}& (s+w+d)(A+c+\lambda )+\frac{1}{2}[A+c+\lambda ,A+c+\lambda ]=F+\mathrm{\Psi }_\mu dx^\mu +\mathrm{\Psi }_5dx^5+\mathrm{\Phi }\hfill \\ & (s+w+d)(F+\mathrm{\Psi }_\mu dx^\mu +\mathrm{\Psi }_5dx^5+\mathrm{\Phi })=[A+c+\lambda ,F+\mathrm{\Psi }_\mu dx^\mu +\mathrm{\Psi }_5dx^5+\mathrm{\Phi }]\hfill \end{array}$$
$`(2.7)`$
The separate conservation of the ghost numbers $`N_s`$ and $`N_w`$ determines the action of $`s`$ and $`w`$ on all fields except $`s\lambda `$ and $`wc`$, since the geometrical equation only determines
$$\begin{array}{cc}\hfill s\lambda +wc=[\lambda ,c].& \end{array}$$
$`(2.8)`$
To resolve this degeneracy we introduce the ghost $`\mu `$, with $`N_s=N_w=1`$, and impose $`s\lambda =\mu `$. The result of the decomposition is:
$$\begin{array}{cc}\hfill sA_\mu & =\mathrm{\Psi }_\mu +D_\mu c,s\mathrm{\Psi }_\mu =D_\mu \mathrm{\Phi }[c,\mathrm{\Psi }_\mu ],sc=\mathrm{\Phi }\frac{1}{2}[c,c],\hfill \\ \hfill sA_5& =\mathrm{\Psi }_5+D_5c,s\mathrm{\Psi }_5=D_5\mathrm{\Phi }[c,\mathrm{\Psi }_5],s\mathrm{\Phi }=[c,\mathrm{\Phi }].\hfill \\ & s\lambda =\mu ,s\mu =0\hfill \end{array}$$
$`(2.9)`$
$$\begin{array}{cc}\hfill wA_\mu & =D_\mu \lambda ,w\mathrm{\Psi }_\mu =[\lambda ,\mathrm{\Psi }_\mu ],wc=\mu [\lambda ,c],\hfill \\ \hfill wA_5& =D_5\lambda ,w\mathrm{\Psi }_5=[\lambda ,\mathrm{\Psi }_5],w\mathrm{\Phi }=[\lambda ,\mathrm{\Phi }]\hfill \\ & w\lambda =\frac{1}{2}[\lambda ,\lambda ]w\mu =[\lambda ,\mu ]\hfill \end{array}$$
$`(2.10)`$
Observe that $`w`$ acts on $`A_\mu `$ and $`A_5`$ like an infinitesimal gauge transformation with gauge parameter $`\lambda `$, and observables will be required to be $`w`$-invariant. Because of the inhomogeneous term $`\mu `$ in $`wc=\mu [\lambda ,c]`$, $`w`$-invariance of a quantity with ghost-number zero assures that it is independent of $`c`$. Moreover $`\mu `$ is a topological ghost for $`c`$ and consequently $`w`$-invariant observables are independent of $`c`$.
For the anti-ghosts of the $`s`$-symmetry, $`\overline{\mathrm{\Psi }}_\mu `$, $`\overline{\mathrm{\Phi }}`$, and the corresponding Lagrange multipliers $`b_\mu `$ and $`\overline{\eta }`$, one has:
$$\begin{array}{cc}\hfill (s+w)\overline{\mathrm{\Psi }}_\mu +[c+\lambda ,\overline{\mathrm{\Psi }}_\mu ]& =b_\mu \hfill \\ \hfill (s+w)b_\mu +[c+\lambda ,b_\mu ]& =[\mathrm{\Phi },\overline{\mathrm{\Psi }}_\mu ]\hfill \end{array}$$
$`(2.11)`$
$$\begin{array}{cc}\hfill (s+w)\overline{\mathrm{\Phi }}+[c+\lambda ,\overline{\mathrm{\Phi }}]& =\overline{\eta }\hfill \\ \hfill (s+w)\overline{\eta }+[c+\lambda ,\overline{\eta }]& =[\mathrm{\Phi },\overline{\mathrm{\Phi }}]\hfill \end{array}$$
$`(2.12)`$
(Notice that $`(s+w)^2=0`$ amounts to $`(s+w+[c=\lambda ,.])^2=[\mathrm{\Phi },.]`$) This definition implies:
$$\begin{array}{cc}\hfill s\overline{\mathrm{\Psi }}_\mu =[c,\overline{\mathrm{\Psi }}_\mu ]+b_\mu ,sb_\mu =[c,b_\mu ]+[\mathrm{\Phi },\overline{\mathrm{\Psi }}_\mu ]& \end{array}$$
$`(2.13)`$
$$\begin{array}{cc}\hfill w\overline{\mathrm{\Psi }}_\mu =[\lambda ,\overline{\mathrm{\Psi }}_\mu ],wb_\mu =[\lambda ,b_\mu ]]& \end{array}$$
$`(2.14)`$
$$\begin{array}{cc}\hfill s\overline{\mathrm{\Phi }}=[c,\overline{\mathrm{\Phi }}]+\overline{\eta },s\overline{\eta }=[c,\overline{\eta }]+[\mathrm{\Phi },\overline{\mathrm{\Phi }}]& \end{array}$$
$`(2.15)`$
$$\begin{array}{cc}\hfill w\overline{\mathrm{\Phi }}=[\lambda ,\overline{\mathrm{\Phi }}],w\overline{\eta }=[\lambda ,\overline{\eta }]& \end{array}$$
$`(2.16)`$
Finally we give the action of $`w`$ and $`s`$ on the anti-ghosts and Lagrange multipliers $`\overline{c}`$, $`l`$, $`\overline{\mu }`$ and $`\overline{\lambda }`$. We take:
$$\begin{array}{cc}\hfill (s+w)\overline{\mu }=\overline{c}+\overline{\lambda }& \\ \hfill (s+w)(\overline{c}+\overline{\lambda })=0& \end{array}$$
$`(2.17)`$
In order to break the degeneracy of the geometrical equation $`w\overline{\lambda }+s\overline{c}=0`$, we introduce the new Lagrange multiplier field $`l`$ , and impose $`s\overline{c}=l`$. After expansion in the ghost numbers $`N_s`$ and $`N_w`$ we obtain,
$$\begin{array}{cc}\hfill s\overline{c}=l,sl=0& \\ \hfill s\overline{\mu }=\overline{\lambda },s\overline{\lambda }=0& \end{array}$$
$`(2.18)`$
$$\begin{array}{cc}\hfill w\overline{\mu }& =\overline{c},w\overline{c}=0\hfill \\ \hfill w\overline{\lambda }& =l,wl=0\hfill \end{array}$$
$`(2.19)`$
We are now ready to ask the standard question, within the BRST paradigm, of determining the gauge-fixed action, the possible counter-terms, and the anomalies of a theory that is gauge and Lorentz invariant and renormalizable by power counting. Here we have the two symmetries $`w`$ and $`s`$, with $`(w+s)^2=0`$, and we must classify the local functions of the fields that are both $`w`$\- and $`s`$-invariant. The important question of anomalies is addressed in Appendix E, where we display an intriguing new cocycle stemming from (2.19). In what follows we consider the question of determining the action in 5 dimensions from the requirement of $`s`$\- and $`w`$-symmetry.
We wish to find the most general solution to the equations $`sI=0`$ and $`wI=0`$, where $`I=𝑑td^4x`$, and $``$ is a local Lagrangian density of engineering dimension 6, and $`I=I^{(0,0)}`$. Here we use the notation defined above to indicate the $`N_s`$ and $`N_w`$ ghost quantum numbers. We rely on the fact that the cohomology of $`s`$ with ghost number zero is empty. Thus the action $`I^{(0,0)}`$ must be an $`s`$-exact term, $`I^{(0,0)}=sI^{(1,0)}`$. We shall use the strategy of descent equations, with $`d`$ and $`s`$ replaced respectively by $`s`$ and $`w`$, because our $`s`$, like $`d`$, has empty cohomology. By $`w`$-invariance of $`I^{(0,0)}`$ we have $`0=wI^{(0,0)}=wsI^{(1,0)}=swI^{(1,0)}`$. Since the cohomology of $`s`$ is empty, this gives
$$\begin{array}{cc}\hfill wI^{(1,0)}=sI^{(2,1)}.& \end{array}$$
$`(2.20)`$
Upon multiplying this equation by $`w`$, we obtain $`0=wsI^{(2,1)}=swI^{(2,1)}`$, and so, since the cohomology of $`s`$ is empty, we have
$$\begin{array}{cc}\hfill wI^{(2,1)}=sI^{(3,2)}.& \end{array}$$
$`(2.21)`$
Now we use the fact that we have assigned the engineering dimensions to the fields in such a way that the operators $`s`$ and $`w`$ preserve engineering dimension, so the $`I^{(i,j)}`$ that we have introduced all have engineering dimension $`[I^{(i,j)}]=0`$, and all the corresponding densities have engineering dimension 6. Since $`I^{(3,2)}`$ has ghost number $`N_s=3`$, it must contain either at least 3 factors of the anti-ghost fields with $`N_s=1`$, or else at least one power of the anti-ghost field $`\overline{\mathrm{\Phi }}^{(2,0)}`$ and another anti-ghost field. In the first case the corresponding density has engineering dimensions 9 or greater because all anti-ghost fields have dimension 3 or 4, as one sees from (2.5). In the second case the corresponding density has engineering dimension 7 or greater. We conclude that $`I^{(3,2)}=0`$, so from (2.21), we have $`wI^{(2,1)}=0`$. Thus $`I^{(2,1)}`$ is of the form
$$\begin{array}{cc}\hfill I^{(2,1)}=I_{\mathrm{inv}}^{(2,1)}+wI^{(2,0)},& \end{array}$$
$`(2.22)`$
where $`I_{\mathrm{inv}}^{(2,1)}`$ is an element of the cohomology of $`w`$. However there is no local density in the cohomology of $`w`$, of engineering dimension 6, with these ghost quantum numbers. For $`I_{\mathrm{inv}}^{(2,1)}`$ must contain at least one power of either $`\mu ^{(1,1)}`$ or $`\lambda ^{(0,1)}`$. If it contains $`\mu ^{(1,1)}`$, then it must contain so many anti-ghost fields that its engineering dimension is too high. On the other hand if it contains $`\lambda ^{(0,1)}`$, then there is no such element of the cohomology of $`w`$. (For example $`(D_5\overline{\mathrm{\Phi }})^a\lambda ^a`$ has the right dimension, but is not in the cohomology of $`w`$.) We conclude that $`I_{\mathrm{inv}}^{(2,1)}=0`$, so from (2.22) we obtain $`I^{(2,1)}=wI^{(2,0)}`$. We substitute this into (2.20) and obtain $`wI^{(1,0)}=swI^{(2,0)}`$, or
$$\begin{array}{cc}\hfill w(I^{(1,0)}+sI^{(2,0)})=0.& \end{array}$$
$`(2.23)`$
Thus the quantity in parenthesis is of the form
$$\begin{array}{cc}\hfill I^{(1,0)}+sI^{(2,0)}=I_{\mathrm{inv}}^{(1,0)}+wI^{(1,1)},& \end{array}$$
$`(2.24)`$
where $`I_{\mathrm{inv}}^{(1,0)}`$ is an element of the cohomology of $`w`$. Upon substitution of this equation into $`I^{(0,0)}=sI^{(1,0)}`$, we obtain
$$\begin{array}{cc}\hfill I^{(0,0)}=sI_{\mathrm{inv}}^{(1,0)}+swI^{(1,1)}.& \end{array}$$
$`(2.25)`$
It follows that the Lagrangian density must be of the form
$$\begin{array}{cc}\hfill s\mathrm{Tr}[\overline{\mathrm{\Psi }}^\mu K_\mu ^{(0,0)}(A,b_\mu )+\overline{\mathrm{\Phi }}K^{(1,0)}(\mathrm{\Psi },A)]+ws\mathrm{Tr}[\overline{\mu }L^{(0,0)}(A)],& \end{array}$$
$`(2.26)`$
where $`K_\mu ^{(0,0)}(A,b_\mu )`$, $`K^{(1,0)}(\mathrm{\Psi },A)`$ and $`L^{(0,0)}(A)`$ have dimension 3, 2 and 2 respectively. Here we have used the fact that the fields $`c`$, $`\overline{\lambda }`$ and and $`\overline{\mu }`$ undergo transformations of topological type under the infinitesimal gauge transformation $`w`$, so they cannot appear in the cohomological term $`I_{\mathrm{inv}}^{(1,0)}`$. In fact $`I_{\mathrm{inv}}^{(1,0)}`$ must be constructed out of combinations of local fields that transform covariantly under $`w`$. Note that the only possible w-exact term that is $`s`$-invariant and thus $`s`$-exact is $`sw`$-exact. To obtain this result we used ghost number conservation and power counting arguments for a local Lagrangian density of engineering dimension 6.
One gets by inspection that the two first term in (2.26), up to multiplicative renormalization constants, must be:
$$\begin{array}{cc}\hfill I_{\mathrm{inv}}=𝑑x^\mu 𝑑x^5s\mathrm{Tr}[\overline{\mathrm{\Psi }}^\mu (F_{5\mu }D_\lambda F_{\lambda \mu }\frac{1}{2}b_\mu )+\overline{\mathrm{\Phi }}(a^{}\mathrm{\Psi }_5D_\lambda \mathrm{\Psi }_\lambda )].& \end{array}$$
$`(2.27)`$
This determines the gauge-invariant part of the action, or more precisely the part of the action that is in the non-trival part of the cohomology of $`w`$ with ghost number zero and dimension 6. The first term, corresponds to the 5-dimensional action of stochastic quantization that one can guess from the Langevin equation for the four-dimensional classical Yang–Mills action. It may be written as:
$$\begin{array}{cc}\hfill I_1=𝑑x^\mu 𝑑x^5s\mathrm{Tr}[\overline{\mathrm{\Psi }}^\mu (F_{5\mu }+\frac{\delta S}{\delta A_\mu }\frac{1}{2}b_\mu )],& \end{array}$$
$`(2.28)`$
where $`S=d^4x(1/4)F_{\mu \nu }^2`$ is the standard Yang–Mills action. The second term of (2.27), that is linear in $`\overline{\mathrm{\Phi }}`$, fixes in a gauge-covariant way the internal gauge invariance of $`\mathrm{\Psi }_\mu `$ and $`\mathrm{\Psi }_5`$. These terms, that we derived in a straightforward way from the requirement of locality and compatibility with power counting, show the conceptual limitation of the idea of stochastic quantization in its original formulation. We refer to for more details concerning the meaning of each term that occurs in the expansion of (2.27).
For a local action of dimension 6, the only possibility for the last term in (2.26) is
$$\begin{array}{cc}\hfill I_{\mathrm{gf}}=dx^\mu dx^5ws\mathrm{Tr}([\overline{\mu }(aA_5A)],& \end{array}$$
$`(2.29)`$
and the total action is given by
$$\begin{array}{cc}\hfill I=I_{\mathrm{inv}}+I_{\mathrm{gf}}.& \end{array}$$
$`(2.30)`$
Remarkably, the only possible gauge-fixing term with dimension 6 provides a linear gauge-fixing. (If the coupling to matter were included, the generalization of the Feynman–’tHooft gauge for spontaneously broken symmetries is easily obtained, $`aA_5=_\mu A_\mu +v\varphi `$.) The derivation of (2.29) using $`w`$-invariance is a significant improvement compared to the derivation in . The term $`I_{\mathrm{inv}}=sI_{\mathrm{inv}}^{(1,0)}`$ is the same as in , and its expansion may be found there.
To see in detail how (2.29) solves the question of raising the degeneracy with respect to ordinary gauge transformations of (2.27), we expand
$$\begin{array}{cc}\hfill ws[\overline{\mu }(aA_5_\nu A_\nu )]=& w\{\overline{\lambda }(aA_5_\nu A_\nu )\hfill \\ & +\overline{\mu }[a\mathrm{\Psi }_5+aD_5c_\nu (\mathrm{\Psi }_\nu +D_\nu c)]\}.\hfill \end{array}$$
$`(2.31)`$
The first term gives
$$\begin{array}{cc}\hfill l(aA_5_\nu A_\nu )\overline{\lambda }(aD_5\lambda _\nu D_\nu \lambda ),& \end{array}$$
$`(2.32)`$
which fixes the gauge for $`A_5`$ and $`\lambda `$. The second term in (2.31) gives:
$$\begin{array}{cc}& \overline{c}[a\mathrm{\Psi }_5+aD_5c_\nu (\mathrm{\Psi }_\nu +D_\nu c)]\hfill \\ & +\overline{\mu }\{aD_5\mu _\nu D_\nu \mu )+a[\lambda ,\mathrm{\Psi }_5+D_5c]_\nu [\lambda ,(\mathrm{\Psi }_\nu +D_\nu c)]\}.\hfill \end{array}$$
$`(2.33)`$
The equation of motion of $`\overline{c}`$ determines a certain linear combination of $`c`$ and $`\mathrm{\Psi }_5`$. But a similar situation occurs for $`\overline{\eta }`$. Using its equation of motion from $`I_{\mathrm{inv}}`$, $`\mathrm{\Psi }_5`$ is determined, so both $`c`$ and $`\mathrm{\Psi }_5`$ are determined. This can be seen by doing the translation $`\overline{\eta }\overline{\eta }\overline{c}`$. Finally, the equation of motion of $`\overline{\mu }`$ determines $`\mu `$, because its equation of motion is parabolic.
An essential feature is that the equations of motion of all ghosts is parabolic, of the type $`(_ta^1_\nu ^2+\mathrm{})\xi =0`$. It follows that the free ghost propagators $`G_0(x,t)=\theta (t)(\frac{a}{4\pi t})^2\mathrm{exp}(\frac{ax^2}{4t})`$ are all retarded, and consequently all closed ghost loops vanish. The only possible exception is the tadpole which gives a purely local contribution to the action which is a renormalization counter-term. However even the tadpole contribution vanishes with dimensional regularization. With dimensional regularization in mind, we conclude that the determinant of each ghost is unity.
Observables $`O`$ are required to be in the cohomology of $`w`$, thus $`wO=0`$ and $`Ow(X)`$, with ghost number 0, and lying in a fixed but arbitrary time slice. However, as noted in the Introduction, they are not $`s`$-invariant, $`sO0`$. We may impose additional conditions on the class of observables. The most conservative policy would be to allow only functions of the variables that are present in the 4-dimensional theory, namely the $`A_\mu `$ for $`\mu =1,\mathrm{}4`$, and Dirac spinor fields. Expectation-values of observables are independent of the parameters of the $`w`$-exact term, that is on the parameter $`a`$. Their independence of the other parameter $`a^{}`$ is quite clear, as shown in : closed ghost loops vanish, and so cannot contribute to the expectation values of observables, and the later cannot depend on $`a^{}`$. Since observables are $`w`$-invariant, which implies that they are gauge invariant, we do get the observables of the 4-dimensional theory.
As for a detailed perturbative proof of the stability of renormalization, it can be done quite rigorously, using the Ward identities derived in Appendix A. If one uses dimensional regularization by computing correlators in $`5ϵ`$ dimensions, one finds that the action $`I`$ undergoes a multiplication renormalization for all its fields and parameters, as a result of imposing the Ward identities of both $`w`$\- and $`s`$ -invariance in the gauge determined by (2.32), that is with the gauge condition (2.2). It should be understood that the correspondence between equal-time correlators in 5 dimensions and those in 4 dimensions is for the physical observables only (defined just above, from the $`w`$-cohomology).<sup>2</sup> The comparison for gauge non-invariant gauge functions would require a non-local theory in five dimensions, , and presents no interest since locality is a key tool for mastering perturbative renormalizability.
3. Interpretation of the action
One way to exhibit the physical content of the theory is to integrate out all ghost and auxiliary fields. We will do this in two steps, in order to get an intermediate result which will be useful later. The first step is to integrate out all ghost fields except $`\lambda `$ and $`\overline{\lambda }`$. As explained above, closed ghost loops vanish because ghost propagators are retarded, $`D(x,t)=0`$ for $`t<0`$, so the ghost determinant which results from this integration is a constant, and the action (2.30) reduces to
$$\begin{array}{cc}\hfill I=d^5x[& ib_\mu (F_{5\mu }D_\lambda F_{\lambda \mu })+(1/2)b_\mu ^2\hfill \\ & +il(aA_5_\mu A_\mu )+\overline{\lambda }(aD_5_\mu D_\mu )\lambda ].\hfill \end{array}$$
$`(3.1)`$
This action possesses the ordinary BRST invariance, here implemented by $`w`$, that encodes gauge invariance in five dimensions. The $`s`$-invariance that stabilizes the action (2.30) has already been exploited, and (3.1) is supposed to be already renormalized. Notice also that the elimination of $`\mathrm{\Psi }`$ and $`\overline{\mathrm{\Psi }}`$ deprived us of the Ito term that is useful non-perturbatively .
We next translate $`b_\mu b_\mu iD_\lambda F_{\lambda \mu }`$. The cross-term $`F_{5\mu }D_\lambda F_{\lambda \mu }`$ is an exact derivative because of the Bianchi identity, and we obtain
$$\begin{array}{cc}\hfill I=d^5x[& ib_\mu F_{5\mu }+(1/2)b_\mu ^2+(1/2)(D_\lambda F_{\lambda \mu })^2\hfill \\ & +il(aA_5_\mu A_\mu )+\overline{\lambda }(aD_5_\mu D_\mu )\lambda ].\hfill \end{array}$$
$`(3.2)`$
If one identifies $`b_\mu `$ as a 4-component color-“electric” field, this resembles Faddeev–Popov theory in first-order formalism, with a particular gauge-fixing term, except that the magnetic energy $`(1/2)B_i^2`$ is replaced by $`(1/2)(D_\lambda F_{\lambda \mu })^2`$. This action still has $`w`$-invariance, but $`s`$-invariance is lost.
As a second step we integrate out the remaining ghost and auxiliary fields $`\lambda ,\overline{\lambda },b_\mu `$, and $`b_5`$. Noting that the ghost determinant is again a constant, we obtain the 5-dimensional non-Abelian gauge action in second-order formalism
$$\begin{array}{cc}\hfill I[A]=d^5x[(1/2)(_5A_\mu a^1D_\mu _\lambda A_\lambda )^2+(1/2)(D_\lambda F_{\lambda \mu })^2].& \end{array}$$
$`(3.3)`$
Here $`A_5`$ is gauge fixed to $`A_5=a^1_\lambda A_\lambda `$, a gauge condition often referred to as stochastic gauge-fixing . Indeed the functional integral associated with the above 5-dimensional action represents the solution of a Fokker-Planck or Langevin equation. The latter describe a stochastic process that may be simulated numerically, and in which the variable $`t=x_5`$ counts sweeps over a 4-dimensional lattice.
The striking feature of the present approach is that the ghost determinant is unity as a result of the parabolic ghost equation. This is a strong indication that the Gribov ambiguity is not a difficulty in the present formulation. However we have not directly addressed here the issue of establishing that the action (3.3) is a valid quantization of the 4-dimensional Yang–Mills theory. We shall return to this topic on another occasion.
4. Conclusion
We have reconsidered the quantization of 4-dimensional gauge theories in 5 dimensions with particular attention to the invariance needed to characterize physical observables. We implemented this invariance by a BRST-operator $`w`$, with $`w^2=0`$, that encodes ordinary gauge invariance, and physical operators $`O`$ are required to satisfy $`wO=0`$, and $`Ow(X)`$. The operator $`w`$ is to be distinguished from the topological BRST-operator $`s`$, with $`s^2=0`$, that will occur in the quantization of all theories with an additional time, including gauge and non-gauge theories. The two operators are compatible in the sense that $`(s+w)^2=0`$. Together, these two symmetries are extremely restrictive, and we have constructed the most general action $`I`$, eqs. (2.27) and (2.29), that is invariant under both symmetries $`sI=wI=0`$, and that is renormalizability by power counting. In Appendix A we derived the Ward identities associated with these 2 symmetries. In the other Appendices we have examined various aspects and extensions of the theory such as the Landau-gauge limit, a confinement scenario, spinor fields, anomalies and lattice discretization without fermion doubling.
The esthetics of the formulation in five dimensions, and its attractive geometric interpretation, strongly suggest that it should be adopted as a starting point for defining a gauge theory. As we have seen in our discussion of anomalies, Appendix E, the fifth time acts as a regulator. Indeed the theory produces correlation functions in five dimensions, and the physical limit requires taking a slice in the fifth time. However new divergences appear when the times coincide (as is obvious in momentum space where setting the times equal corresponds to additional integrations over the conjugate momenta). This is our interpretation of the chiral anomaly.
The 5-dimensional formulation provides the stage for a simple confinement scenario in the minimal Landau gauge. In this scenario, the long-range confining “force” is transmitted by $`A_5`$, the fifth component of the gluon field. It is suggestive that the imposition of the analog of Gauss’s law in the fifth dimension can produce a confinement scenario similar to the one in the Coulomb gauge and , where the long-range confining force is transmitted by $`A_4`$. However the scenario is compatible with manifest 4-dimensional Lorentz invariance.
We use the $`s`$ and $`w`$ invariance to improve the argument that we gave in to prove renormalizability of the action in 5 dimensions. This requires that we also address possible anomalies. We exhibit an interesting cocycle in 5 dimensions (solution of the Wess–Zumino consistency conditions for $`s`$ and $`w`$ symmetries). However its coefficient appears to be 0 in general, because of the topological nature of the action. Consequently the Noether currents $`K_M`$ ($`M=1,\mathrm{},5`$) of the 5-dimensional action are strictly conserved. The origin of the 4-dimensional anomaly may be understood when one recognizes that the currents of physical interest are not the $`K_M`$, but rather the Noether currents $`J_\mu `$ ($`\mu =1,\mathrm{},4`$) of the 4-dimensional action. These two different currents are not related in any obvious way. Indeed the $`J_\mu `$ are not conserved in the 5-dimensional theory in general. However consistency requires that the $`J_\mu `$ generate the appropriate 4-dimensional Ward identities when inserted into correlation functions at a fixed fifth time. This may fail due to singularities that appear in the correlation functions when the times are set equal. Indeed we have verified by explicit computation that the ABBJ triangle anomaly appears as a discontinuity of the 3-point $`J`$-current correlator as equal times are approached. Power counting in 5-dimensions leads us to expect that there are no other such breakdowns.
The principles of locality, gauge symmetry and power counting also determine the form of the topological action for spinors, whether chiral or not. It is gratifying that this yields a convergent functional integral that gives well-defined correlation functions. It should be emphasized that the stochastic interpretation of the fifth coordinate must be abandoned when the local five-dimensional theory is extended to include fermions. Remarkably, the 5-dimensional spinor action allows a natural lattice discretization that does not suffer from fermion doubling, as is shown in Appendix F. It is not necessary to introduce a domain wall for this purpose , , .
In summary, the five-dimensional formulation of renormalizable gauge theories appears as a very powerful tool for investigating the non perturbative questions relative to the Yang–Mills theory.
Acknowledgments
The research of Daniel Zwanziger was partially supported by the National Science Foundation under grant PHY-9900769. P.A. Grassi is supported by National Science Foundation under grant PHY-9722083.
Appendix A. Ward Identities, Renormalization, and Stability of the Action
We shall show that with the action (3.1) the combination $`g\lambda `$ is invariant under renormalization,
$$\begin{array}{cc}\hfill g\lambda =g_r\lambda _r& \end{array}$$
$`(\text{A.}1)`$
Moreover in the Landau gauge limit $`a0`$, the quantity $`gA_5`$ is invariant also invariant under renormalization.
In and it was shown that $`gA_4`$ is invariant under renormalization in the Coulomb gauge in the 4-dimensional formulation. For the same reason $`gA_5`$ is invariant under renormalization in 5-dimensional formulation in the Landau gauge. \[The result depends on the following argument. The component $`j_5`$ of the conserved Noether current of the $`w`$-symmetry is given by $`j_5=/(_5A_\mu )D_\mu \lambda =ib_\mu D_\mu \lambda `$, because all derivatives with respect to $`t=x_5`$ are contained in $`F_{5,\mu }=_5A_\mu D_\mu A_5`$. Moreover the left-hand side of Gauss’s law is given by $`\delta I/\delta A_5=iD_\mu b_\mu `$. This allows us to write the conserved Noether charge $`Q=d^4xj_5`$ as $`Q=d^4x\lambda \delta I/\delta A_5`$. No product of fields appears in this expression, so it provides a Ward identity that is special to this gauge, and it leads to the condition $`Z_{A_5}Z_g=1`$ on the renormalization constants $`Z_{A_5}`$ and $`Z_g`$ of $`A_5`$ and $`g`$.\]
The renormalization of the model contains some interesting features that we discuss in the present appendix. First, we will show that the rich content of the symmetry (the $`s`$-symmetry and the $`w`$-symmetry (2.10)) provides strong constraints on the form of the gauge fixed action; second, we will show that the particular gauge choice (2.29) implies non-renormalization properties for the ghost fields.
In order to discuss the set of functional equations which implement the symmetries at the quantum level, we introduce the sources coupled to the variations of the fields. The latter are needed to renormalize theories with non-linear transformations as in our case. Given $`\varphi `$, a field of the set $`\{A_\mu ,A_5,\mathrm{\Psi }_\mu ,\mathrm{\Psi }_5,\mathrm{\Phi },c,\overline{\mathrm{\Psi }}_\mu ,b_\mu ,\overline{\mathrm{\Phi }},\overline{\eta }\}`$, we introduce its ($`ws`$)-source $`\varphi ^{}`$, its $`w`$-source $`\varphi ^{}`$, and its $`s`$-source $`\varphi ^{\prime \prime }`$. The statistic of $`\varphi ^{},\varphi ^{\prime \prime }`$ is opposite to that of $`\varphi `$, and the statistic of $`\varphi ^{}`$ is the same of $`\varphi `$. On the other hand, the ghost numbers of $`\varphi ^{},\varphi ^{\prime \prime },\varphi ^{}`$ are easily fixed by their couplings. In particular for a field $`\varphi `$ with ghost numbers $`(N_s,N_w)`$, we have
$$\begin{array}{cc}\hfill I_{\mathrm{sources}}[\varphi ]& =d^5x\underset{\varphi }{}w\left[s(\varphi ^{}\varphi )\right]=d^5x\underset{\varphi }{}w\left[(s\varphi ^{})\varphi +()^{N_s+N_w}\varphi ^{}s\varphi \right]=\hfill \\ \hfill =\underset{\varphi }{}& d^5x\left[(ws\varphi ^{})\varphi +()^{N_s+N_w+1}(s\varphi ^{})w\varphi +()^{N_s+N_w}(w\varphi ^{})s\varphi +\varphi ^{}ws\varphi \right].\hfill \end{array}$$
$`(\text{A.}2)`$
Adding a $`ws`$-trivial term does not modify either the $`s`$\- or the $`w`$-cohomology of the classical theory. However, at the quantum level, we have to translate the $`s`$\- and $`w`$-transformations in terms of nilpotent functional operators, and therefore, we also require
$$\begin{array}{cc}\hfill s\varphi ^{}& =()^{1+N_s+N_w}\varphi ^{},w\varphi ^{}=()^{N_s+N_w}\varphi ^{\prime \prime },ws\varphi ^{}=sw\varphi ^{}=0,\hfill \\ \hfill s\varphi ^{}& =0,w\varphi ^{}=0,s\varphi ^{\prime \prime }=0,w\varphi ^{\prime \prime }=0.\hfill \end{array}$$
$`(\text{A.}3)`$
Consequently, the eq. (A.2) becomes
$$\begin{array}{cc}\hfill I_{\mathrm{sources}}[\varphi ]=\underset{\varphi }{}d^5x\left[\varphi ^{}(w\varphi )+\varphi ^{\prime \prime }(s\varphi )+\varphi ^{}(ws\varphi )\right].& \end{array}$$
$`(\text{A.}4)`$
As concerns the fields $`\lambda ,\mu `$ and the anti-ghosts $`\overline{\lambda },\overline{\mu },\overline{c}`$, due to the fact that their $`s`$ and $`w`$-transformations are almost trivial, we can spare several sources, but we cannot avoid the following two terms
$$\begin{array}{cc}\hfill I_{\mathrm{ghost}}=d^5x\left[\lambda ^{}\left(\frac{1}{2}[\lambda ,\lambda ]\right)\lambda ^{}[\lambda ,\mu ]\right],& \end{array}$$
$`(\text{A.}5)`$
where we identify $`\lambda ^{}`$ with $`\mu ^{}`$. More specifically, we do not introduce sources for the antighost fields $`\overline{\lambda },\overline{c}`$ and $`\overline{\mu }`$, and for the fields $`\lambda `$ and $`\mu `$, we have $`s\lambda ^{}=\lambda ^{}`$ and $`w\lambda ^{}=0`$ which is consistent with the identification $`\lambda ^{}=\mu ^{}`$. In the following, the symbol $`I`$ denotes the action including the source terms.
Following the conventional procedure and given the source terms (A.4) and (A.5) , we can establish the functional identities for $`s`$\- and $`w`$-symmetry for the generating functional $`\mathrm{\Gamma }`$ of irreducible Green functions
$$\begin{array}{cc}\hfill 𝒮(\mathrm{\Gamma })& =d^5x\left\{\underset{\varphi }{}\left[\frac{\delta \mathrm{\Gamma }}{\delta \varphi }\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{\prime \prime }}()^{N_s+N_w}\varphi ^{}\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}\right]+\mu \frac{\delta \mathrm{\Gamma }}{\delta \lambda }\lambda ^{}\frac{\delta \mathrm{\Gamma }}{\delta \lambda ^{}}+\overline{\lambda }\frac{\delta \mathrm{\Gamma }}{\delta \overline{\mu }}l\frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}\right\}=0,\hfill \\ \hfill 𝒲(\mathrm{\Gamma })& =d^5x\left\{\underset{\varphi }{}\left[\frac{\delta \mathrm{\Gamma }}{\delta \varphi }\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}()^{N_s+N_w}\varphi ^{\prime \prime }\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}\right]+\frac{\delta \mathrm{\Gamma }}{\delta \lambda }\frac{\delta \mathrm{\Gamma }}{\delta \lambda ^{}}+\frac{\delta \mathrm{\Gamma }}{\delta \mu }\frac{\delta \mathrm{\Gamma }}{\delta \mu ^{}}+\overline{c}\frac{\delta \mathrm{\Gamma }}{\delta \overline{\mu }}+l\frac{\delta \mathrm{\Gamma }}{\delta \overline{\lambda }}\right\}=0,\hfill \end{array}$$
$`(\text{A.}6)`$
where the sum is extended to the entire set of fields $`\varphi `$ of the present model. The nilpotency and the commutation relations between the $`s`$-symmetry and the $`w`$-symmetry are expressed by the following equations
$$\begin{array}{cc}\hfill 𝒮_\mathrm{\Gamma }^2=0,𝒲_\mathrm{\Gamma }^2=0,\{𝒲_\mathrm{\Gamma },𝒮_\mathrm{\Gamma }\}=0,& \end{array}$$
$`(\text{A.}7)`$
where $`𝒮_\mathrm{\Gamma }`$ and $`𝒲_\mathrm{\Gamma }`$ are the linearized version of the functional operators involved in (A.6) and are defined by $`𝒮_\mathrm{\Gamma }(\mathrm{\Xi })\frac{}{ϵ}𝒮(\mathrm{\Gamma }+ϵ\mathrm{\Xi })|_0`$. Besides eqs. (A.6), one has to take into account the equations of motion for the Lagrangian multiplier $`l`$ (cf. eq. (2.29))
$$\begin{array}{cc}\hfill \frac{\delta \mathrm{\Gamma }}{\delta l}=aA_5_\mu A^\mu ,& \end{array}$$
$`(\text{A.}8)`$
where $`a`$ is a gauge parameter. From the five-dimensional point of view, the gauge fixing (A.8) appears like a Landau gauge fixing type, but from the four-dimensional point of view, $`A_5`$ plays the rôle of the Lagrangian multiplier and the gauge fixing (A.8) is a truly Lorentz gauge type. In the following we will show that, for a generic $`a`$ the theory satisfies a set of functional equations for the ghost fields $`\lambda ,c,\mu `$. Those equations imply the non-renormalization of the ghost fields themselves and, in the case $`a=0`$, a further equation implies the non-renormalization of the combination $`gA_5`$ field.
The existence of a solution to the system of equations (A.8), (A.6) is based on the Fröbenius theorem whose main hypothesis is the existence of an algebra of vector fields on the manifold spanned by functionals like $`\mathrm{\Gamma }`$. The algebra is constructed by computing the commutation relations between the eqs (A.6) (cf. (A.7)) and with the eq. (A.8). For a generic functional $``$, they read
$$\begin{array}{cc}\hfill 𝒲_{}\left(\frac{\delta }{\delta l}aA_5+_\mu A^\mu \right)\frac{\delta }{\delta l}𝒲()& =\frac{\delta }{\delta \overline{\lambda }}a\frac{\delta }{\delta A_5^{}}+_\mu \frac{\delta }{\delta A_\mu ^{}},\hfill \\ \hfill 𝒮_{}\left(\frac{\delta }{\delta l}aA_5+_\mu A^\mu \right)\frac{\delta }{\delta l}𝒮()& =\frac{\delta }{\delta \overline{c}}a\frac{\delta }{\delta A_5^{\prime \prime }}+_\mu \frac{\delta }{\delta A_\mu ^{\prime \prime }},\hfill \end{array}$$
$`(\text{A.}9)`$
and, by commuting again the resulting equations with the operators $`𝒮_{},𝒲_{}`$, we have
$$\begin{array}{cc}\hfill 𝒲_{}\left(\frac{\delta }{\delta \overline{c}}a\frac{\delta }{\delta A_5^{\prime \prime }}+_\mu \frac{\delta }{\delta A_\mu ^{\prime \prime }}\right)\left(\frac{\delta }{\delta \overline{c}}a\frac{\delta }{\delta A_5^{\prime \prime }}+_\mu \frac{\delta }{\delta A_\mu ^{\prime \prime }}\right)𝒲()& =\frac{\delta }{\delta \overline{\mu }}a\frac{\delta }{\delta A_5^{}}+_\mu \frac{\delta }{\delta A_\mu ^{}},\hfill \\ \hfill 𝒮_{}\left(\frac{\delta }{\delta \overline{\lambda }}a\frac{\delta }{\delta A_5^{}}+_\mu \frac{\delta }{\delta A_\mu ^{}}\right)\left(\frac{\delta }{\delta \overline{\lambda }}a\frac{\delta }{\delta A_5^{}}+_\mu \frac{\delta }{\delta A_\mu ^{}}\right)𝒮()& =\frac{\delta }{\delta \overline{\mu }}a\frac{\delta }{\delta A_5^{}}+_\mu \frac{\delta }{\delta A_\mu ^{}}.\hfill \end{array}$$
$`(\text{A.}10)`$
All the other commutation relations are trivial and therefore (A.6), (A.9), (A.10) describe all the possible non-vanishing commutation relations and the algebra is, in fact, closed. We notice that the commutation relations between the ghost equations (A.9) with the operators $`𝒮_{},𝒲_{}`$ generate the same functional equation for the ghost $`\mu `$. This fact is a consequence of the anti-commutation relations between $`𝒮_{}`$ and $`𝒲_{}`$. Finally, the equations (A.9), (A.10) expressed in term of $`\mathrm{\Gamma }`$ can be easily integrated. This amounts to a redefinition of the sources $`A_M^{},A_M^{\prime \prime }`$, and $`A_M^{}`$. We will not discuss these details since they are common to the conventional procedure of the BRST quantization with linear gauge fixings.
More interesting and specific to the present model are further functional equations for the antighost fields $`\overline{\lambda },\overline{c}`$ and $`\overline{\mu }`$. These can be derived by analyzing the corresponding tree level equations and observing that the composite operators which appear in those formulae are already present in the extended action $`I`$. From eqs. (2.31), (2.32), (2.33), and (A.4), we have
$$\begin{array}{cc}\hfill \frac{\delta I}{\delta \lambda (x)}& =\left(aD_5\overline{\lambda }D_\mu ^\mu \overline{\lambda }\right)a[\mathrm{\Psi }_5+D_5c,\overline{\mu }][\mathrm{\Psi }_\mu +D_\mu c,^\mu \overline{\mu }]+[\lambda ^{},\lambda ]+[\lambda ^{},\mu ]\hfill \\ & +\underset{\varphi }{}d^5y\left[()^{1+N_s+N_w}\varphi ^{}\frac{\delta }{\delta \lambda (x)}(w\varphi )(y)+()^{N_s+N_w}\varphi ^{}\frac{\delta }{\delta \lambda (x)}(ws\varphi )(y)\right],\hfill \\ \hfill \frac{\delta I}{\delta c(x)}& =\left(aD_5\overline{c}D_\mu ^\mu \overline{c}\right)+aD_5[\overline{\mu },\lambda ]D_\mu [^\mu \overline{\mu },\lambda ]+[c^{},\lambda ]\hfill \\ & +\underset{\varphi }{}d^5y[()^{N_s+N_w+1}\varphi ^{\prime \prime }\frac{\delta }{\delta c(x)}(s\varphi )(y)+()^{N_s+N_w}\varphi ^{}\frac{\delta }{\delta c(x)}(ws\varphi )(y)\hfill \\ & +()^{N_s+N_w+1}\varphi ^{}\frac{\delta }{\delta c(x)}(w\varphi )(y)],\hfill \\ \hfill \frac{\delta I}{\delta \mu (x)}& =\left(aD_5\overline{\mu }D_\mu ^\mu \overline{\mu }\right)[\lambda ^{},\lambda ]c^{}+\underset{\varphi }{}d^5y\left[\varphi ^{}\frac{\delta }{\delta \mu (x)}(ws\varphi )(y)\right].\hfill \end{array}$$
$`(\text{A.}11)`$
It is important to note that the terms proportional to the sources $`\varphi ^{},\varphi ^{\prime \prime }`$ are linear in the quantum fields and therefore they do not require an independent renormalization besides the usual field renormalization. As an example, some of those terms are explicitly shown in the last equation. On the other hand, the last terms proportional to the sources $`\varphi ^{}`$ are not linear in quantum fields and therefore they require more care. Eqs. (A.11) are not suitable for quantization since they involve the renormalization of operators like $`\left(aD_5\overline{\lambda }D_\mu ^\mu \overline{\lambda }\right)`$ which, unfortunately, do not belong to the set of those coupled to the sources (A.4) and (A.5). However, considering the integrated (over the 5-dimensional manifold) version of eqs. (A.11) and integrating by parts, we have the following equations for $`\mathrm{\Gamma }`$:
$$\begin{array}{cc}\hfill d^5x\frac{\delta \mathrm{\Gamma }}{\delta \lambda (x)}=d^5x& \{[\overline{\lambda },\frac{\delta \mathrm{\Gamma }}{\delta l}]+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}]+[\mu ^{},\mu ]+[\lambda ^{},\lambda ]\hfill \\ & +\underset{\varphi }{}[\varphi ^{},\varphi ]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{\prime \prime }}]\},\hfill \\ \hfill d^5x\frac{\delta \mathrm{\Gamma }}{\delta c(x)}=d^5x& \{[\overline{c},\frac{\delta \mathrm{\Gamma }}{\delta l}]+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta \overline{\lambda }}]+[\lambda ,\frac{\delta \mathrm{\Gamma }}{\delta \mu }]+[c^{},\lambda ]\hfill \\ & +\underset{\varphi }{}[\varphi ^{\prime \prime },\varphi ]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}]\},\hfill \\ \hfill d^5x\frac{\delta \mathrm{\Gamma }}{\delta \mu (x)}=d^5x& \left\{[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta l}][\lambda ^{},\lambda ]c^{}+\underset{\varphi }{}()^{N_s+N_w}[\varphi ,\varphi ^{}]\right\}.\hfill \end{array}$$
$`(\text{A.}12)`$
Notice that, by using the sources $`\varphi ^{}`$ and $`\varphi ^{\prime \prime }`$, we are able to translate from the tree level approximation (A.11) to the quantum level also the terms coming from the $`sw`$ variations. As a check of the procedure, we can compute the anti-commutation relations between the anti-ghost equations (A.12) and the ghost equations (A.9) and (A.10), to see that they close on the gauge-fixing equation (A.8) and they are compatible with each other. This supports the hypothesis of the Fröbenius theorem which allows us to integrate the complete system of equations. In particular, it is important to compute the commutation relation between the $`𝒲_\mathrm{\Gamma }`$ and the first equation of (A.12) or that of $`𝒮_\mathrm{\Gamma }`$ with the first one. This gives a new functional equation
$$\begin{array}{cc}\hfill d^5x& \{[\overline{\lambda },\frac{\delta \mathrm{\Gamma }}{\delta \overline{\lambda }}]+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta \overline{\mu }}]+[\overline{c},\frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}]+[\lambda ,\frac{\delta \mathrm{\Gamma }}{\delta \lambda }]+[\mu ,\frac{\delta \mathrm{\Gamma }}{\delta \mu }]\hfill \\ & +\underset{\varphi }{}[\varphi ,\frac{\delta \mathrm{\Gamma }}{\delta \varphi }]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}]+[\varphi ^{\prime \prime },\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{\prime \prime }}]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}]+[c^{},\mu ]\}=0,\hfill \end{array}$$
$`(\text{A.}13)`$
which implements the invariance under the rigid gauge transformations of the model. As in the case of Yang–Mills, quantized with the conventional BRST technique, in the Landau gauge, rigid gauge invariance is a by-product of the dynamics of the ghost fields .
Equations (A.12) control the renormalization of the ghost fields $`\lambda ,c,\mu `$ and, in particular, by means of those equations the ghost fields have no independent renormalization. This is a common feature of Landau type gauge fixing .
To derive these non-renormalization properties, and for pedagogical purposes, we consider a model without chiral fermions and we assume an invariant regularization scheme. In that framework, we can renormalize the model multiplicatively. That is, all the fields will be renormalized by a suitable wave function renormalization $`\varphi Z_\varphi \varphi `$. Requiring that eqs. (A.8), (A.9), (A.10), and (A.12) are preserved by the renormalization procedure, we immediately get
$$\begin{array}{cc}\hfill Z_a& =Z_AZ_5^1,Z_{\overline{\mu }}=Z_{\overline{c}}=Z_{\overline{\lambda }}=Z_l=Z_A^1,\hfill \\ \hfill Z_\mu & =Z_lZ_{\overline{\mu }}^1=1,Z_\lambda =Z_lZ_{\overline{\lambda }}^1=Z_{\overline{\mu }}^1Z_{\overline{c}}=1,\hfill \\ \hfill Z_c& =Z_lZ_{\overline{c}}^1=Z_{\overline{\mu }}^1Z_{\overline{\lambda }}=Z_\mu Z_\lambda ^1=1,\hfill \end{array}$$
$`(\text{A.}14)`$
where we used the rescaled gauge fields $`A_MgA_M`$. Due to the relations $`Z_\lambda =Z_c=Z_\mu =1`$, it is direct to conclude that the products $`g\lambda ,gc`$ and $`g\mu `$ do not renormalize.
Finally we can switch off the gauge parameter $`a`$ by letting it go to zero. In this case, the analysis can be repeated obtaining the same results for the non-renormalization of the ghost fields. However, instead of integrating the anti-ghost equations (A.11) over the five-dimensional space, we can also integrate only over the four-dimensional Lorentz invariant manifold. This implies that the anti-ghost equations (A.12), rewritten with integrals over four dimensions, are
$$\begin{array}{cc}\hfill d^4x\frac{\delta \mathrm{\Gamma }}{\delta \lambda (x)}=d^4x& \{_5A_5^{}+[\overline{\lambda },\frac{\delta \mathrm{\Gamma }}{\delta l}]+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}]+[\mu ^{},\mu ]+[\lambda ^{},\lambda ]\hfill \\ & +\underset{\varphi }{}[\varphi ^{},\varphi ]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{\prime \prime }}]\},\hfill \\ \hfill d^4x\frac{\delta \mathrm{\Gamma }}{\delta c(x)}=d^4x& \{_5A_5^{\prime \prime }+[\overline{c},\frac{\delta \mathrm{\Gamma }}{\delta l}]+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta \overline{\lambda }}]+[\lambda ,\frac{\delta \mathrm{\Gamma }}{\delta \mu }]+[c^{},\lambda ]\hfill \\ & +\underset{\varphi }{}[\varphi ^{\prime \prime },\varphi ]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}]\},\hfill \\ \hfill d^4x\frac{\delta \mathrm{\Gamma }}{\delta \mu (x)}=d^4x& \left\{_5A_5^{}+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta l}][\lambda ^{},\lambda ]c^{}+\underset{\varphi }{}()^{N_s+N_w}[\varphi ,\varphi ^{}]\right\}.\hfill \end{array}$$
$`(\text{A.}15)`$
They are local in the fifth component of the 5-dimensional space and satisfy all the commutation relations among themselves and with the other functional operators with the proper obvious modifications. However, due to the presence of new terms like $`_5d^4xA_5^{}`$, not killed by the integration over the 5-dimensional space, the commutation relation of the first equation of (A.15) with $`𝒲_\mathrm{\Gamma }`$, or the second one with $`𝒮_\mathrm{\Gamma }`$, generates the new functional equation
$$\begin{array}{cc}\hfill _5& d^4x\frac{\delta \mathrm{\Gamma }}{\delta A_5}+d^4x\{[\overline{\lambda },\frac{\delta \mathrm{\Gamma }}{\delta \overline{\lambda }}]+[\overline{\mu },\frac{\delta \mathrm{\Gamma }}{\delta \overline{\mu }}]+[\overline{c},\frac{\delta \mathrm{\Gamma }}{\delta \overline{c}}]+[\lambda ,\frac{\delta \mathrm{\Gamma }}{\delta \lambda }]+[\mu ,\frac{\delta \mathrm{\Gamma }}{\delta \mu }]\hfill \\ & +\underset{\varphi }{}[\varphi ,\frac{\delta \mathrm{\Gamma }}{\delta \varphi }]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}]+[\varphi ^{\prime \prime },\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{\prime \prime }}]+[\varphi ^{},\frac{\delta \mathrm{\Gamma }}{\delta \varphi ^{}}]+[c^{},\mu ]\}=0.\hfill \end{array}$$
$`(\text{A.}16)`$
This equation implements the invariance under gauge transformations, local in the fifth component, and, automatically, implies that the fifth component of the gauge field $`A_5`$ is not renormalized, or equivalently, by rescaling properly the gauge fields by the gauge coupling $`g`$, the combination $`gA_5`$ is not renormalized. We may use the $`gA_5`$ propagator to define an invariant charge in QCD. A similar situation happens for the abelian sector
of the Standard Model quantized in the background gauge .
To conclude the present section, we would like to stress that for a generic $`a`$ the content of the symmetry is so rich that it constrains the form of the action strongly. Indeed, by a tedious algebra, one can easily prove that the only solution of the system which satisfies the power-counting constraints is given by the action $`(2.27)`$.
Appendix B. Landau Gauge Limit
We wish to consider the important limiting case in which the gauge parameter $`a0`$, and the 4-vector potential $`A_\mu `$ becomes transverse $`_\mu A_\mu =0`$. Like the Coulomb gauge condition $`_iA_i=0`$ in the 4-dimensional formulation, this gauge condition is a singular in the present 5-dimensional formulation because it leaves $`t`$-dependent but $`x`$-independent gauge-transformations $`g(t)`$ unfixed (cf. eq. (A.16)). However it is expected that the limit $`a0`$ is finite in the sense that the renormalized correlation functions calculated at positive $`a`$ have a finite limit as $`a0`$. Indeed this property has been established in the 4-dimensional formalism when the Coulomb gauge is approached from an interpolating gauge . Without attempting here to establish the finiteness of the limit $`a0`$ in the present case, we shall instead derive some of its properties.
The transversality condition $`_\mu A_\mu =0`$ is generally called the Landau gauge. However this is not a well-defined gauge because of the Gribov ambiguity, and there is really an infinite class of Landau gauges. We shall show that in the limit $`a0`$ we end up in a gauge, which we call the “minimal” Landau gauge, in which the additional condition
$$\begin{array}{cc}\hfill M(A)_\mu D_\mu (A)0,& \end{array}$$
$`(\text{B.}1)`$
is satisfied. This states that the Faddeev–Popov operator $`M(A)`$ is non-negative. It is a condition on $`A_\mu `$ that defines the (first) Gribov region. By contrast, the Faddeev–Popov method in 4 dimensions does not restrict the functional integral in the Landau gauge to the Gribov region. Indeed, if one attributes a non-perturbative significance to the Faddeev–Popov formula by BRST quantization, then a signed sum over the entire region both inside and outside the Gribov horizon is implied.
To study the limit $`a0`$, we rescale $`tat`$ in (3.3) and obtain
$$\begin{array}{cc}\hfill I=d^5x[(2a)^1(_5A_\mu D_\mu A)^2+(a/2)(D_\lambda F_{\lambda \mu })^2],& \end{array}$$
$`(\text{B.}2)`$
where $`A_\mu A_\mu `$. As $`a`$ approaches 0, the first term dominates and the probability gets concentrated at its absolute minimum namely at
$$\begin{array}{cc}\hfill _5A_\mu =D_\mu A.& \end{array}$$
$`(\text{B.}3)`$
With $`t=x_5`$ as the time variable, this equation defines a flow that is tangent to the gauge orbit, with infinitesimal generator $`\omega =A`$. The global, non-perturbative character of this flow will be deduced from the fact that it describes steepest descent of the functional
$$\begin{array}{cc}\hfill F=(1/2)d^4xA_\mu ^2& \end{array}$$
$`(\text{B.}4)`$
restricted to the gauge orbit. For under an arbitrary infinitesimal gauge transformation $`\delta A_\mu =D_\mu \omega `$, the functional undergoes the infinitesimal variation
$$\begin{array}{cc}\hfill \delta F=d^4xA_\mu \delta A_\mu =d^4xA_\mu D_\mu \omega =d^4x_\mu A_\mu \omega ,& \end{array}$$
$`(\text{B.}5)`$
so steepest descent of this functional restricted to the gauge orbit is indeed achieved with $`\omega =_\mu A_\mu `$. Since the functional is bounded below, the steepest descent necessarily approaches a local minimum.<sup>3</sup> This argument is rigorous in the 5-dimensional compact lattice-gauge formulation . The classical descent could in principle end at a saddle-point but quantum fluctuations prevent this outcome. At a minimum, the functional is stationary, which yields the Landau gauge condition $`A=0`$. Moreover its second variation is non-negative at a minimum, which, from (B.5), gives the additional condition,
$$\begin{array}{cc}\hfill d^4x\omega (D)\omega 0\mathrm{for}\mathrm{all}\omega ,& \end{array}$$
$`(\text{B.}6)`$
which establishes (B.1). Thus in the limit $`a0`$, the gauge-fixed functional integral (B.2), which is local in 5 dimensions, has the 4-dimensional property of concentrating the probability within the Gribov horizon.<sup>4</sup> Because there are in general more than one relative minimum on a gauge orbit, the restriction to the Gribov region (B.1) is not a complete gauge fixing. However the validity of the present formulation does not in any way depend on how the probability may be distributed among the various possible relative minima in the limit $`a0`$, nor, for that matter, does it depend on taking the limit $`a0`$ at all, but rather is valid for any value of $`a0`$.
Remarkably, the Landau gauge limit just obtained from the local 5-dimensional action coincides with the “minimal” Landau gauge that is used in numerical studies of lattice gauge theory. Indeed in these studies, the gauge is fixed by numerically minimizing a lattice analog of the functional $`F=(1/2)d^4xA_\mu ^2`$ restricted to the gauge orbit, so that the configuration is likewise brought to a local minimum of the functional, namely to a point within the first Gribov region. Thus the 5-dimensional formulation provides a BRST-invariant, renormalizable, continuum description of the minimal Landau gauge that is accessible numerically. Moreover other numerically accessible gauges such as the minimal Coulomb gauge or maximal abelian gauge may be introduced in the 5-dimensional formalism by imposing the 5-dimensional gauge condition $`aA_5=G(x)F`$, where $`F`$ is an appropriately chosen functional, and $`G(x)=D_\mu \delta F/\delta A_\mu `$ is the gradient in the gauge orbit direction. These are not in the class of Faddeev–Popov gauges in which at the non-perturbative level a signed sum over the entire region outside the Gribov region is implied. Needless to say, if one wishes to compare numerically gauge-fixed quantities such as propagators with analytic predictions such as, for example, the Nielsen identities , they should be derived in the same gauge.
Appendix C. Confinement Scenario
The present 5-dimensional formulation provides a simple confinement scenario in the Landau gauge limit. The basic idea is that $`gA_5`$, which is a Lorentz scalar, provides a vehicle for the transmission of long-range correlations that correspond to a confining force. In this respect it resembles the component $`gA_4`$ of the gluon field in the Coulomb gauge. In fact the scenario was originally developed for the Coulomb gauge by Gribov and elaborated more recently . However in the present formulation, the mechanism respects manifest 4-dimensional Lorentz invariance. The discussion that follows is non-perturbative, and could be based on the lattice-gauge formulation of the 5-dimensional formalism presented in .
It is shown in Appendix A that in the Landau gauge limit, the field $`gA_5`$ is invariant under renormalization
$$\begin{array}{cc}\hfill gA_5=g_rA_{5r},\mathrm{for}a=0.& \end{array}$$
$`(\text{C.}1)`$
This implies that in the Landau gauge limit all correlation functions of $`gA_5`$ are renormalization-group invariants. As such they are finite and independent of the ultra-violet cut-off, and depend only on the QCD scale $`\mathrm{\Lambda }_{QCD}`$. Thus the statement that they are long range refers to the QCD scale.
To derive the properties of the Landau-gauge limit, we return to the first-order action (3.2), and integrate out the ghosts $`\lambda `$ and $`\overline{\lambda }`$, which again gives a constant determinant. We pose $`a=0`$, and obtain
$$\begin{array}{cc}\hfill I=d^5x[ib_\mu F_{5\mu }+(1/2)b_\mu ^2+(1/2)(D_\lambda F_{\lambda \mu })^2il_\mu A_\mu ].& \end{array}$$
$`(\text{C.}2)`$
Apart from the substitution $`B_i^2(D_\lambda F_{\lambda \mu })^2`$, this is the Coulomb-gauge action in 5 dimensional space-time, and we shall solve the constraints just as is in the Coulomb gauge. We integrate out the Lagrange multiplier field $`l`$, which gives $`\delta (_\mu A_\mu )`$, and imposes the gauge constraint $`A_\mu =A_\mu ^{\mathrm{tr}}`$. Integration on $`A_5`$ which appears in the action only in the term $`b_\mu F_{5\mu }=b_\mu (_5A_\mu D_\mu A_5)`$ gives $`\delta (D_\mu b_\mu )`$ which imposes a form of Gauss’s law,
$$\begin{array}{cc}\hfill D_\mu b_\mu =0& \end{array}$$
$`(\text{C.}3)`$
on the 4-dimensional color-electric field $`b_\mu `$. Indeed if we write this as
$$\begin{array}{cc}\hfill _\mu b_\mu =\rho ,& \end{array}$$
$`(\text{C.}4)`$
the color-charge density $`\rho =j_5=[A_\mu ,b_\mu ]`$ is the $`j_5`$-component of the conserved Noether current $`(j_\mu ,j_5)`$ of global gauge invariance. To solve Gauss’s law, we write $`b_\mu =b_\mu ^{\mathrm{tr}}_\mu \varphi `$, where $`\varphi `$ plays the role of a Coulomb potential for the 4-dimensional color-electric field $`b_\mu `$. Gauss’s law is solved by
$$\begin{array}{cc}\hfill \varphi =M^1\rho _l,& \end{array}$$
$`(\text{C.}5)`$
where $`\rho _l[A_\mu ^{\mathrm{tr}},b_\mu ^{\mathrm{tr}}]`$ is the color-charge density of the transverse gluons. Here $`M=M(A^{\mathrm{tr}})=D_\mu (A^{\mathrm{tr}})_\mu `$, with $`\mu =1,\mathrm{}4`$, is the 4-dimensional Faddeev–Popov operator. Its inverse $`M^1(A^{\mathrm{tr}})`$ is an integral operator that acts instantaneously in the 5-dimensional space-time and transmits the long-range force that is expected to be confining.
When this expression for $`\varphi `$ is substituted back into the action, it assumes the canonical form<sup>5</sup> One might expect that the integral over $`\varphi `$ will produce the inverse of the Faddeev–Popov determinant $`det^1(_\mu D_\mu )`$ which, with $`\mu =1,\mathrm{}4`$, is ill-defined in a 5-dimensional functional integral. However the integral over $`\varphi `$ should be performed before setting gauge parameter $`a=0`$, which gives instead $`det^1(aD_5_\mu D_\mu )`$. This is a constant for the same reason that the ghost determinant is a constant, namely the propagator of the parabolic operator $`a_5_\mu ^2`$ is retarded, so all closed loops vanish.
$$\begin{array}{cc}\hfill I(A^{\mathrm{tr}},b^{\mathrm{tr}})=𝑑t[ib_\mu ^{\mathrm{tr}}_5A_\mu ^{\mathrm{tr}}+H(A^{\mathrm{tr}},b^{\mathrm{tr}})],& \end{array}$$
$`(\text{C.}6)`$
with “hamiltonian”
$$\begin{array}{cc}\hfill H=& d^4x[(1/2)(b_\mu ^{\mathrm{tr}})^2+(1/2)(D_\lambda F_{\lambda \mu })^2]\hfill \\ & +(1/2)d^4xd^4y\rho _l(x)[M^1(^2)M^1](x,y)\rho _l(y).\hfill \end{array}$$
$`(\text{C.}7)`$
This is the Coulomb hamiltonian in one extra space dimension. The last term represents the instantaneous color-Coulomb interaction of separated color charge $`\rho _l`$.
The preceding derivation may also be used to show that the propagator
$$\begin{array}{cc}\hfill D_{55}(x,t)=gA_5(x,t)gA_5(0,0).& \end{array}$$
$`(\text{C.}8)`$
develops an instantaneous part in the Landau gauge limit,
$$\begin{array}{cc}\hfill D_{55}(x,t)=V(x)\delta (t)+(\mathrm{non}\mathrm{instantaneous}).& \end{array}$$
$`(\text{C.}9)`$
Here $`V(x)`$ is an analog of the instantaneous color Coulomb potential in one extra dimension that is given by
$$\begin{array}{cc}\hfill V(xy)=[M^1(^2)M^1](x,y).& \end{array}$$
$`(\text{C.}10)`$
As shown in Appendix A, the correlation functions of $`gA_5`$ are renormalization-group invariants including in particular the instantaneous part $`V(x)`$, and its fourier transform $`\stackrel{~}{V}(k)`$. It is thus of the form $`\stackrel{~}{V}(k)=g^2(k/\mathrm{\Lambda }_{QCD})/k^2`$, where $`g(k/\mathrm{\Lambda }_{QCD})`$ is a running coupling constant that depends only the QCD mass scale.
Thus the Landau gauge limit of the 5-dimensional formulation exhibits all the features of the Coulomb gauge in one extra space dimension, with the additional advantage of maintaining manifest 4-dimensional Lorentz invariance. Consequently the arguments for confinement in the Coulomb gauge , , may be taken over wholesale to the present case. Without reproducing these arguments here we recall that the restriction to the first Gribov region $`M(A^{\mathrm{tr}})0`$, demonstrated in the previous section, produces 2 related effects. (i) Configurations corresponding to small eigenvalues of $`M(A^{\mathrm{tr}})`$ are favored by entropy. This makes the Green function $`[M^1(A^{\mathrm{tr}})](xy)`$ long range, and consequently also the instantaneous color-Coulomb potential in (C.7)and (C.10). (ii) The low frequency components of $`A_\mu ^{\mathrm{tr}}`$ are suppressed, thereby eliminating the physical gluons from the physical spectrum. Indeed it has been proven in lattice gauge theory in the Landau or Coulomb gauge that at infinite lattice volume the gluon propagator $`D_{\mu \nu }(k)`$ of transverse gluons vanishes for any probability distribution that is restricted to the first Gribov region . It would require detailed dynamical arguments to determine whether the instantaneous color-Coulomb potential is sufficiently long range to confine all colored objects. However the mechanism of a long-range force that couples universally to color charge (C.7), is clearly present.
Appendix D. Spinor action and propagator
The five-dimensional formalism can accommodate the coupling of the gauge field to spinors, and moreover we observed heuristically in that this extended point of view naturally introduces many of the ingredients that look rather ad hoc in the genuine four-dimensional formulation when one tries to answer non-perturbative questions. We wish to address these issues in more detail, particularly the question of spinors and chirality.
The action that defines the propagators and the interactions of the four-dimensional spinors in the five-dimensional formalism was introduced in . Our postulate is that one must extend the definition of the topological BRST $`s`$-operator to spinors (in a way that does not necessitates the introduction of a Langevin equation with a fermionic noise.) Naturally this must be done in a way which is consistent with the symmetries such as the gauge invariance implemented by $`w`$-symmetry, the chirality, the conservation of the fermionic number, etc. Once this is done, the principle of locality and renormalizability must determine the form of a renormalizable local action in five dimensions. As we will see shortly, this simple five-dimensional point of view turns out to be surprisingly predictive.
Let $`q(x)`$ be a spinor in four dimensions. For concreteness we suppose it describes a quark, with spinor and color indices suppressed as usual. Extend it to a five-dimensional object $`q(x)q(x,t)`$, without affecting the spinorial index. In accordance with our postulate, $`q`$ is a member of a spinor quartet, made up of $`q,\mathrm{\Psi }_q,\overline{\mathrm{\Psi }}_q`$, and $`b_q`$, which are, besides the anti-commuting spinor $`q`$, its commuting topological ghost and anti-ghost, and $`b_q`$, the anti-commuting Lagrange multiplier. They each carry unit quark charge, and $`s`$ and $`w`$ act on them according to
$$\begin{array}{cc}\hfill (s+w)q+(c+\lambda )q=\mathrm{\Psi }_q(s+w)\mathrm{\Psi }_q+(c+\lambda )\mathrm{\Psi }_q=\mathrm{\Phi }q& \\ \hfill (s+w)\overline{\mathrm{\Psi }}_q+(c+\lambda )\overline{\mathrm{\Psi }}_q=b_q(s+w)\beta _q+(c+\lambda )b_q=\mathrm{\Phi }\overline{\mathrm{\Psi }}_q,& \end{array}$$
$`(\text{D.}1)`$
where $`c,\lambda `$ and $`\mathrm{\Phi }`$ act on the spinors in (say) the fundamental representation, $`cc^at^a`$, where $`t^a`$ are the Gell-Mann matrices. By separate conservation of $`N_s`$ and $`N_w`$, this gives $`sq=\mathrm{\Psi }_qcq`$, $`s\mathrm{\Psi }_q=c\mathrm{\Psi }_q\mathrm{\Phi }q`$ and $`s\overline{\mathrm{\Psi }}_q=c\overline{\mathrm{\Psi }}_q+b_q`$, $`sb_q=c,b_q\mathrm{\Phi },\overline{\mathrm{\Psi }}_q`$, and the $`w`$-symmetry acts on the field as an infinitesimal gauge transformation with a parameter equal to the ghost $`\lambda `$. There is a corresponding independent quartet for the anti-quark which consists of $`q^{\mathrm{}},\mathrm{\Psi }_q^{\mathrm{}},\overline{\mathrm{\Psi }}_q^{\mathrm{}}`$, and $`b_q^{\mathrm{}}`$, each with quark charge $`1`$. <sup>6</sup> Since we already use “bar” notation to designate an anti-ghost, such as $`\overline{\mathrm{\Psi }}`$, we unconventionally use “dagger” notation, such as $`q^{}`$ instead of $`\overline{q}`$, to designate Dirac conjugation.
Five-dimensional Lorentz covariance of spinors $`q,\mathrm{\Psi }_q,\overline{\mathrm{\Psi }}_q,b_q`$ does not concern us since we only have in view an $`SO(4)\times R`$-invariant theory rather than an $`SO(5)`$-invariant one, namely the 5-dimensional formulation of the standard Dirac action,
$$\begin{array}{cc}\hfill S=d^4xq^{\mathrm{}}(\gamma _\mu D_\mu m)q.& \end{array}$$
$`(\text{D.}2)`$
The method of stochastic quantization is not of direct relevance for the case of anticommuting fields. Therefore, to determine the 5-dimensional action for spinors, we postulate that it must be invariant under the topological BRST symmetry (D.1). Using power counting and locality requirements, this must give:
$$\begin{array}{cc}\hfill I_q=d^5xs\{\overline{\mathrm{\Psi }}_\mu \frac{\delta S}{\delta A_\mu }+& \overline{\mathrm{\Psi }}_q^{\mathrm{}}[D_5qK(\frac{\delta S}{\delta q^{\mathrm{}}}\frac{1}{2}b_q)]\hfill \\ & +[q^{\mathrm{}}D_5(\frac{\delta S}{\delta q}+\frac{1}{2}b_q^{\mathrm{}})K]\overline{\mathrm{\Psi }}_q\},\hfill \end{array}$$
$`(\text{D.}3)`$
that is,
$$\begin{array}{cc}\hfill I_q=d^5xs(\overline{\mathrm{\Psi }}_\mu q^{\mathrm{}}\gamma _\mu q+& \overline{\mathrm{\Psi }}_q^{\mathrm{}}\{[D_5K(\gamma _\mu D_\mu m)]q+\frac{K}{2}b_q\}\hfill \\ & +\{q^{\mathrm{}}[D_5(\gamma _\mu D_\mu m)K]b_q^{\mathrm{}}\frac{K}{2}\}\overline{\mathrm{\Psi }}_q).\hfill \end{array}$$
$`(\text{D.}4)`$
The first term contains the anti-ghost $`\overline{\mathrm{\Psi }}_\mu `$ of $`A_\mu `$, and is an additional contribution to the gluon action (2.27).
Here is $`K`$ is a kernel that in principle is at our disposal, because the expectation value of observables calculated on a given time-slice at equilibrium is independent of $`K`$ as can be seen by Ward identities. However we must choose it to obtain a well-defined functional integral. In particular one easily sees that it is necessary that the combination $`K(\gamma _\mu D_\mu +m)`$ be bounded below. If we give the standard canonical dimension $`3/2`$ to $`q`$ and its ghost $`\mathrm{\Psi }_q`$, and canonical dimension $`5/2`$ to the anti-ghost $`\overline{\mathrm{\Psi }}_q`$ and Lagrange multiplier $`b_q`$, and similarly for the $`q^{\mathrm{}}`$ quartet, and impose gauge-invariance in the form of $`w`$-invariance, then power counting, $`SO(4)`$-invariance and gauge invariance imply that
$$\begin{array}{cc}\hfill K=\gamma ^\mu D_\mu +M.& \end{array}$$
$`(\text{D.}5)`$
This is the most general action which one may write down with these properties, to within renormalization constants, apart from the fact that the same kernel $`K`$ multiplies both $`\gamma _\mu D_\mu `$ and $`b_q`$. As a matter of convenience we shall set the mass $`m`$ in the kernel equal to the mass of the Dirac spinor, $`M=m`$ so the combination
$$\begin{array}{cc}\hfill QK(\gamma _\mu D_\mu +m)=KK^{}=K^{}K=(\gamma _\mu D_\mu +m)K=(\gamma _\mu D_\mu )^2+m^2& \end{array}$$
$`(\text{D.}6)`$
is hermitian and positive and chirally even. (For well-defined perturbation theory, $`mM0`$ is sufficient.)
We may consider Weyl spinors in four dimensions, with chiral coupling to gauge fields. The signal that we deal with Weyl spinors is minimally contained in the range of values over which the spinor index runs, together with the various $`V\pm A`$ coupling that may arise in the interactions. The four-dimensional matrix $`\gamma ^5`$ extends trivially to the fifth $`\gamma `$ matrix in five dimensions.
The extension to chiral gauge coupling is automatic in the 5-dimensional fermion action (D.3), if the original 4-dimensional Dirac action (D.2) is chiral. In particular, one sees that if $`q`$ is, say, left-handed, then $`\mathrm{\Psi }_q`$ is also left-hand, whereas $`\overline{\mathrm{\Psi }}_q`$ and $`b_q`$ are right-handed, and oppositely for the anti-quark quartet.
Let us find the form of the free propagators. Because of the presence of the kernel there is a mixing of $`q`$ and $`baq`$. The quadratic approximation of the action is:
$$\begin{array}{cc}\hfill I_q=d^4xdt[& b_q^{}(_5_\mu _\mu +m^2)qq^{}(_5_\mu _\mu +m^2)b_q\hfill \\ & +b_q^{}(\gamma _\mu _\mu +m)b_q+\mathrm{ghost}\mathrm{terms}]\hfill \end{array}$$
$`(\text{D.}7)`$
This gives the following matrix propagator, between the independent pairs $`(q^{},b_q^{})`$ and $`(q,b_q)`$:
$$\begin{array}{cc}\hfill \left(\begin{array}{cc}\frac{m+i\gamma _\mu p_\mu }{\omega ^2+(p^2+m^2)^2}& \frac{1}{i\omega +p^2+m^2}\\ & \\ \frac{1}{i\omega +p^2+m^2}& 0\end{array}\right)& \end{array}$$
$`(\text{D.}8)`$
If we integrate the $`q^{}q`$ matrix element over $`\omega `$, we obtain the usual free Dirac propagator
$$\begin{array}{cc}\hfill \frac{1}{\pi }𝑑\omega \frac{m+i\gamma _\mu p_\mu }{\omega ^2+(p^2+m^2)^2}=\frac{m+i\gamma _\mu p_\mu }{p^2+m^2}.& \end{array}$$
$`(\text{D.}9)`$
There are closed loops of fields $`q`$ and $`q^{}`$, but the associated ghosts $`\mathrm{\Psi }_q`$ and $`\overline{\mathrm{\Psi }}_q`$ have retarded propagators so are no closed ghost loops, just as in the bose case. The topological quantum field theory in five dimensions operates for the spinors and their topological ghosts and Lagrange multipliers in the same way as it does for gauge fields. We have a well-defined perturbation theory for the system of spinors and gauge fields in five dimensions, and the familiar arguments about renormalizability apply.
Appendix E. Anomalies
The five-dimensional theory is a local quantum quantum field theory that is renormalizable by power counting. Our aim is that the generating functional of Green functions must be constructed from the requirement of $`s`$ and $`w`$ invariances. Then, observables are defined from the operators that are the cohomology with ghost number zero of of $`w`$, with expectation values taken at equal times.
The possibility of renormalizing the theory in five dimensions, while maintaining the two symmetries can be jeopardized by the presence of anomalies. Such anomalies can destroy the theory in two different ways, since the $`s`$-symmetry ensures the existence of a Fokker–Plank equation and the $`w`$-symmetry ensures that the drift forces give gauge invariant observables. Our knowledge of the four-dimensional theory indicates that some kind of breakdown must occur in the five-dimensional approach when one couples a chiral fermion to the theory, or a set of chiral fermions without couplings that ensure anomaly compensations.
A primary check is thus to verify whether radiative corrections can break the Ward identities of $`s`$ and $`w`$ symmetries. This can be done by computing the possible obstructions of the Ward identity in five dimensions and the values of their coefficient. But, as already mentioned, the currents and their conservation laws in four and five dimensions are not related in an obvious way. For instance, if one calls $`J_\mu (x)`$ a conserved current of the four-dimensional theory, the insertion of $`_\mu J_\mu (x,t)`$ in correlation functions with fields and/or operators taken at different values of the fifth time $`t`$ has no reason to give zero.
Since we cannot rely on the above approach to understanding the effect of the anomaly, we are led to investigate the possibility that, when one computes insertions of four-dimensionally conserved currents in Green functions, their conservation is not fulfilled in the limit of equal fifth times for all arguments of the Green function.
The search of anomalies relies therefore on an indirect argument. Knowing from general principles that the symmetry of the theory in a given slice is the ordinary four-dimensional gauge symmetry, we can select and compute in five dimensions the possible anomalous vertices the four-dimensional theory. We will do so, and find an ambiguity in the equal all fifth time limit for the external legs. We will see that this ambiguity is rooted in the usual problem of computing Feynman diagrams with superficial linear divergences.
Before doing this important calculation, we will nevertheless show the existence of a cocycle for the combined $`w`$ and $`s`$ symmetries. Its overall coefficient vanishes perturbatively (which indicates that no anomaly occurs in five dimensions, for all Green functions). However, this cocycle has an interesting form, and its role in quantum field theory could be important. We could not find its interpretation, but it is striking that it is related in a formal way to the usual four-dimensional consistent anomaly by descent equations.
The possible obstructions to the Ward identity for the $`w`$ and $`s`$ symmetries in five dimensions must be 5-forms, with ghost number (1,0) and (0,1) respectively. We call them $`\mathrm{\Delta }_5^{(1,0)}`$ and $`\mathrm{\Delta }_5^{(0,1)}`$. The ghost unification allows us to define $`\mathrm{\Delta }_5^1=\mathrm{\Delta }_5^{(1,0)}+\mathrm{\Delta }_5^{(0,1)}`$ Since $`(s+w)^2=0`$, $`\mathrm{\Delta }_5^1`$ must satisfies the consistency equation for the $`s`$ and $`w`$ symmetries:
$$\begin{array}{cc}\hfill (s+w)\mathrm{\Delta }_5^1+d\mathrm{\Delta }_4^2=0& \end{array}$$
$`(\text{E.}1)`$
If we assume that only exterior products play a role, we can easily manipulate this equation. We use that $`(s+w)d\mathrm{\Delta }_5^1=0`$, and thus $`d\mathrm{\Delta }_5^1+(s+w)\mathrm{\Delta }_6^0=0`$. Thus $`(s+w)d\mathrm{\Delta }_6^0=0`$. Since $`d\mathrm{\Delta }_6^0`$ has ghost number 0, it is a 7-form that can only depend on $`A`$ and $`dA`$, and since the symmetry is $`(s+w)A=\mathrm{\Psi }D(c+\lambda )`$, the only possibility is that $`d\mathrm{\Delta }_6^0=0`$, which implies in turn that $`\mathrm{\Delta }_6^0`$ is an invariant polynomial in the Yang–Mills curvature $`F=dA+AA`$, that is, $`\mathrm{\Delta }_6^0=\mathrm{Tr}(FFF)=dQ_5(A,F)`$, where $`Q_5(A,F)`$ is a Chern class of rank 5. We thus have that the solution of the Wess and Zumino consistency equation (E.1) is obtained from the piece with ghost number 2 in the following identity satisfied by $`\mathrm{\Delta }_6(F+\mathrm{\Psi }+\mathrm{\Phi })=\mathrm{Tr}[(F+\mathrm{\Psi }+\mathrm{\Phi })^3]`$:
$$\begin{array}{cc}\hfill (s+w+d)\mathrm{\Delta }_6(F+\mathrm{\Psi }+\mathrm{\Phi })=0& \end{array}$$
$`(\text{E.}2)`$
We thus have:
$$\begin{array}{cc}& w\mathrm{Tr}(\mathrm{\Psi }FF)=0\hfill \\ & (s+w)\mathrm{Tr}(\mathrm{\Psi }FF)+d\mathrm{Tr}(\mathrm{\Psi }\mathrm{\Psi }F+FF\mathrm{\Phi })=0\hfill \end{array}$$
$`(\text{E.}3)`$
and thus
$$\begin{array}{cc}\hfill \mathrm{\Delta }_5^{1,0}& =\mathrm{Tr}(\mathrm{\Psi }FF)\hfill \\ \hfill \mathrm{\Delta }_5^{0,1}& =0\hfill \end{array}$$
$`(\text{E.}4)`$
is a candidate for the anomaly.
On the other hand, due to the Chern–Simons formula,
$$\begin{array}{cc}\hfill \mathrm{Tr}[(F+\mathrm{\Psi }+\mathrm{\Phi })^3]=(d+s+w)Q_5(A+c+\lambda ,F+\mathrm{\Psi }+\mathrm{\Phi })& \end{array}$$
$`(\text{E.}5)`$
we have
$$\begin{array}{cc}\hfill \mathrm{Tr}(\mathrm{\Psi }FF)=d\mathrm{Tr}\left[c\frac{\delta }{\delta A}|_FQ_5(A,F)+\mathrm{\Psi }\frac{\delta }{\delta F}|_AQ_5(A,F)+sQ_5(A,F)\right]& \end{array}$$
$`(\text{E.}6)`$
Therefore, in five dimensions, $`\mathrm{Tr}(\mathrm{\Psi }FF)`$ is locally the sum of $`d`$\- and $`s`$\- exact terms. However, although it is $`w`$-invariant, it not $`w`$-exact, up to $`d`$\- and $`s`$-exact terms, and therefore we can identify $`\mathrm{Tr}(\mathrm{\Psi }FF)`$ as part (and probably the unique element) of the cohomology with ghost number one for the $`w`$-symmetry. Due to this fact, we can probably safely call $`\mathrm{Tr}(\mathrm{\Psi }FF)`$ the consistent five dimensional anomaly.
One recognizes among all terms in the right hand side of (E.6) the interesting piece: $`_t\mathrm{\Delta }_4^1`$, where
$$\begin{array}{cc}\hfill \mathrm{Tr}\mathrm{\Delta }_4^1=\mathrm{Tr}\left[c\frac{\delta }{\delta A}|_FQ_5(A,F)\right]& \end{array}$$
$`(\text{E.}7)`$
is nothing else that the consistent four-dimensional anomaly that can be directly related to the ABBJ triangle anomaly once it is inserted on the right hand side of the Ward identity in four dimensions. All this suggests that adding $`Q_5(A,F)`$ could become an interesting issue.
Although, it is certainly natural to interpret the existence of the cocycle $`\mathrm{Tr}(\mathrm{\Psi }FF)`$ as the origin of the anomaly that must occur when the theory is coupled to chiral four-dimensional spinors, we have not been able yet to see the way $`\mathrm{Tr}(\mathrm{\Psi }FF)`$ plays a role in the five-dimensional theory. This does not mean however that the cocycle doesn’t play a role, for instance in Fujikawa type manipulations when on reduces the theory in four dimensions.
Consider now practical computations, to understand how the anomaly will manifest itself. We can take the case of a single spinor $`q`$, and introduce the vector and axial currents $`J^\mu =\overline{q}\gamma ^\mu q`$ and $`J_5^\mu =\overline{q}\gamma ^5\gamma ^\mu q`$. The anomaly questions amounts to compute the form factor of the 1PI vertex in four dimensions $`T^{\alpha \beta }=<K_\mu J_5^\mu (K),J^\alpha (k),J^\beta (k^{})>`$ that is proportional to $`ϵ_{\alpha \beta \gamma \gamma }k^\alpha k_{}^{}{}_{}{}^{\beta }`$. Here we assume that the vector current is conserved, and $`K=k+k^{}`$.
The free propagators of spinors in five dimensions are given by (D.8). $`T^{\alpha \beta }(k,k^{},T,t,t^{})`$ is thus given by the following 5-dimensional Feynman integral
$$\begin{array}{cc}\hfill & d\mathrm{\Omega }d\omega d\omega ^{}d^4p\mathrm{exp}i(\mathrm{\Omega }T+\omega t+\omega ^{}t^{})\hfill \\ & \mathrm{tr}(K_\mu \gamma ^\mu \gamma ^5\frac{(pk)_\rho \gamma ^\rho }{\omega ^2+((pk)_\rho (pk)^\rho )^2}\hfill \\ & \gamma ^\alpha \frac{p_\rho \gamma ^\rho }{\mathrm{\Omega }^2+(p_\rho p^\rho )^2}\gamma ^\beta \frac{(p+k^{})_\rho \gamma ^\rho }{\omega _{}^{}{}_{}{}^{2}+((p+k^{})_\rho (p+k^{})\rho )^2})\hfill \end{array}$$
$`(\text{E.}8)`$
We can use the translation invariance along the fifth time direction, and set $`T`$ at the origin of time. It is thus sufficient to compute the integral at $`T=0`$. We can thus investigate carefully the various way one can approach the value $`T=t=t^{}=0`$ that must give the result of the four-dimensional theory.
We first observe, that if we set brutally $`T=t=t^{}=0`$ the integral (E.8) is just equal to the ordinary four-dimensional triangle diagram. This is easily seen by performing the integration over $`d\mathrm{\Omega }`$, $`d\omega `$ and $`d\omega `$ by using Cauchy theorem for picking out the poles in $`\mathrm{\Omega }`$, $`\omega `$ and $`\omega ^{}`$, which gives:
$$\begin{array}{cc}\hfill T^{\alpha \beta }(k,k^{},T=t=t^{})=𝑑\mathrm{\Omega }𝑑\omega 𝑑\omega ^{}d^4p& \mathrm{tr}(K_\mu \gamma ^\mu \gamma ^5\frac{(pk)_\rho \gamma ^\rho }{\omega ^2+((pk)_\rho (pk)^\rho )^2}\gamma ^\alpha \hfill \\ & \frac{p_\rho \gamma ^\rho }{\mathrm{\Omega }^2+(p_\rho p^\rho )^2}\gamma ^\beta \frac{(p+k^{})_\rho \gamma ^\rho }{\omega _{}^{}{}_{}{}^{2}+((p+k^{})_\rho (p+k^{})\rho )^2})\hfill \end{array}$$
$`(\text{E.}9)`$
When one extracts the term proportional to $`ϵ_{\alpha \beta \mu \nu }k^\mu k_{}^{}{}_{}{}^{\nu }`$, this expression has the well known linear divergence which provides a non-vanishing value for the anomaly.
If, on the other hand, we set $`t`$ and $`t^{}`$ different from zero, the divergence in the integration over $`p`$ is regularized due to the exponentials in $`t`$. We can compute the integral explicitly for small values of $`t`$ and $`t^{}`$, and one finds that the anomaly coefficient is proportional to:
$$\begin{array}{cc}\hfill ϵ_{\alpha \beta \mu \nu }k^\mu k_{}^{}{}_{}{}^{\nu }\frac{|tT|+|t^{}T|}{|tT|+|t^{}T|+|tt^{}|}.& \end{array}$$
$`(\text{E.}10)`$
This expression exhibits an ambiguity in the equal-time limit $`t=t^{}=T`$. The ratio of absolute values varies between $`1/2`$ and $`1`$. It equals unity if one first sets $`t=t^{}`$ and then $`t=T`$; but it equals $`1/2`$ if one first sets $`t=T`$ and then $`t^{}=T`$. This contrasts with the tree-level property that the correlation functions in four dimensions are obtained as a smooth limit of their counterparts in five dimensions.
The ambiguity of the expression (E.10) is related to the property that first setting $`t=t^{}=T`$ causes superficially linear divergences. We thus foresee that only the triangle can lead to an ambiguity in the limit $`t=t^{}=T`$; otherwise the fifth time acts an invariant regulator.
Appendix F. Absence of fermion doubling
We have seen in our analysis of anomalies that they do not appear in the 5-dimensional theory per se, but rather when the fifth time is restricted to a slice. This suggests that the absence of fermion doubling in the 5-dimensional formulation is practically automatic, and does not require the introduction of domain walls , , .
Indeed, consider the matrix form of the free Dirac action (D.7) between the independent pairs $`(q^{},b_q^{})`$ and $`(q,b_q)`$,
$$\begin{array}{cc}\hfill \left(\begin{array}{cc}0& _5+^2m^2\\ & \\ _5^2+m^2& \gamma _\mu _\mu +m\end{array}\right),& \end{array}$$
$`(\text{F.}1)`$
where $`^2_\mu _\mu `$ for $`\mu =1,\mathrm{}4`$, is the 4-dimensional lattice Laplacian. We shall use a standard lattice discretization of the operators that appear here, $`_5(_5)_d`$ etc., where the discretized operators are defined by
$$\begin{array}{cc}\hfill (_5)_dq(x,t)& \frac{1}{2}[q(x,t+1)q(x,t1)]\hfill \\ \hfill (^2)_dq(x,t)& \underset{\mu =1}{\overset{4}{}}[q(x+\widehat{\mu },t)+q(x\widehat{\mu },t)2q(x,t)]\hfill \\ \hfill (_\mu )_dq(x,t)& \frac{1}{2}[q(x+\widehat{\mu },t)q(x\widehat{\mu },t)],\hfill \end{array}$$
$`(\text{F.}2)`$
where $`\widehat{\mu }`$ is a unit vector in the $`+\mu `$-direction. We have preserved hermiticity properties, so $`(_5)_d`$ and $`(_\mu )_d`$ are anti-symmetric whereas $`(^2)_d`$ is symmetric. The trick is that we discretized the 4-dimensional lattice laplacian $`^2`$ and the lattice derivatives $`_\mu `$ independently, so $`(^2)_d_{\mu =1}^4[(_\mu )_d]^2`$. It will turn out that we never have to invert the Dirac operator but only the even chirality operators that appear in the off-diagonal matrix elements of the matrix $`(\text{F.}1)`$. This simplifying property results from our choice of the kernel, $`K=m+\gamma _\mu D_\mu `$, corresponding to $`M=m`$.
All these operators are diagonalized by lattice Fourier transformation, and the matrix (F.1) becomes in terms of lattice momenta $`\theta _\mu `$ and $`\theta _5`$,
$$\begin{array}{cc}\hfill \left(\begin{array}{cc}0& i\mathrm{sin}\theta _5Q_0\\ & \\ i\mathrm{sin}\theta _5+Q_0& K_0\end{array}\right),& \end{array}$$
$`(\text{F.}3)`$
where $`Q_0m^2+_{\mu =1}^42(1\mathrm{cos}\theta _\mu )`$, and $`K_0i\mathrm{sin}\theta _\mu \gamma _\mu +m`$.
The propagator is given by the inverse matrix,
$$\begin{array}{cc}\hfill \left(\begin{array}{cc}\frac{K_0}{\mathrm{sin}^2\theta _5+Q_0^2}& \frac{1}{i\mathrm{sin}\theta _5+Q_0}\\ & \\ \frac{1}{i\mathrm{sin}\theta _5Q_0}& 0\end{array}\right).& \end{array}$$
$`(\text{F.}4)`$
The free 4-dimensional lattice $`q^{}q`$ propagator in momentum space is obtained at equal fifth time, namely
$$\begin{array}{cc}\hfill S(\theta _\mu )=\frac{1}{2\pi }_\pi ^\pi 𝑑\theta _5\frac{K_0}{\mathrm{sin}^2\theta _5+Q_0^2}=\frac{K_0}{Q_0(1+Q_0^2)^{1/2}}.& \end{array}$$
$`(\text{F.}5)`$
In the continuum limit, this integral gets contributions from the neighborhood of $`\theta _5=0`$ and $`\theta _5=\pi `$ which reflects fermion doubling in $`\theta _5`$. However there is no doubling of the physical 4-dimensional propagator. Indeed in the continuum limit, $`Q_0^2=O(a^4)`$ is negligible compared to $`1`$, where $`a`$ is the lattice spacing, and we obtain
$$\begin{array}{cc}\hfill S(\theta _\mu )\frac{K_0}{Q_0}=\frac{m+i\underset{\mu =1}{\overset{4}{}}\mathrm{sin}\theta _\mu \gamma _\mu }{m^2+_{\mu =1}^42(1\mathrm{cos}\theta _\mu )}.& \end{array}$$
$`(\text{F.}6)`$
No fermion doubling occurs here. As asserted, the Dirac operator is never inverted, but only the operator $`i\mathrm{sin}\theta _5+m^2+_{\mu =1}^42(1\mathrm{cos}\theta _\mu )`$ that has even chirality. The denominator of the last expression cannot be factorized on the lattice, $`Q_0K_0K_0^{}`$, but it does factorize in the continuum limit, when $`\mathrm{sin}\theta _\mu q_\mu `$, and $`2(1\mathrm{cos}\theta _\mu )q_\mu ^2`$.
As regards practical numerical simulation, a possible advantage of the lattice discretization described here with respect to domain-wall fermions is that every hyperplane $`x_5=\mathrm{const}`$ may be used for 4-dimensional fermions, not just the one domain wall , , .
References
relax L. Baulieu and D. Zwanziger, QCD<sub>4</sub> From a Five-Dimensional Point of View, hep-th/9909006. relax G. Parisi, Y.S. Wu Sci. Sinica 24 (1981) 484. relax E. Witten, Topological Quantum Field Theory, hep-th/9403195, Commun. Math. Phys. 117 (1988)353. relax L. Baulieu, B. Grossman, A topological Interpretation of Stochastic Quantization Phys. Lett. B 212 (1988) 351; L. Baulieu Stochastic and Topological Field Theories, Phys. Lett. B 232 (1989) 479; Topological Field Theories And Gauge Invariance in Stochastic Quantization, Int. Jour. Mod. Phys. A 6 (1991) 2793. relax L. Baulieu, B. Grossman, Monopoles and Topological Field Theory, Phys. Lett. B214 (1988) 223. L. Baulieu, Chern-Simons Three-Dimensional and Yang–Mills-Higgs Two-Dimensional Systems as Four-Dimensional Topological Quantum Field Theories, Phys. Lett. B232 (1989) 473. relax L. Baulieu, D. Zwanziger, Equivalence of Stochastic Quantization and the-Popov Ansatz, Nucl. Phys. B 193 (1981) 163. relax P.H. Daamgard, H. Huffel, Phys. Rep. 152 (1987) 227. relax P.H. Daamgard, H. Huffel, Eds., Stochastic Quantization, World Scientific (1988). relax M. Namiki and K. Okano, Eds, Prog. Theor. Phys. Suppl 111 (1993). relax H.S. Chan, M.B. Halpern, Phys. Rev. D 33 (1986) 540. relax Yue-Yu, Phys. Rev. D 33 (1989) 540. relax J. Zinn-Justin, D. Zwanziger, Nucl. Phys. B 295 (1988) 297. relax J. Zinn-Justin, Nucl. Phys. B 275 (1986) 135. relax D. Zwanziger, Covariant Quantization of Gauge Fields without Gribov Ambiguity, Nucl. Phys. B 192, (1981) 259. relax J.H. Horne, Superspace versions of Topological Theories, Nucl.Phys. B 318 (1989) 22. relax S. Ouvry, R. Stora, P. Van Baal , Phys. Lett. B 220 (1989) 159; R. Stora, Exercises in Equivariant Cohomology, In Quantum Fields and Quantum Space Time, Edited by ’t Hooft et al., Plenum Press, New York, 1997 relax I.M. Singer, Comm. of Math. Phys. 60 (1978) 7. relax V.N. Gribov, Nucl. Phys. B 139 (1978) 1. relax D. Zwanziger, Renormalization in the Coulomb gauge and order parameter for confinement in QCD, Nucl.Phys. B 538 (1998) 237. relax D.B. Kaplan, A Method for Simulating Chiral Fermions on the Lattice, Phys. Lett. B 288 (1992) 342; Chiral Fermions on the Lattice, Nucl. Phys. B, Proc. Suppl. 30 (1993) 597. relax H. Neuberger, Chirality on the Lattice, hep-lat/9808036. relax Y. Shamir, Lattice Chiral Fermions Nucl. Phys. Proc. Suppl. 47 (1996) 212, hep-lat/9509023; V. Furman, Y. Shamir, Nucl.Phys. B 439 (1995), hep-lat/9405004. relax L. Baulieu, D. Zwanziger, Renormalizable Non-Covariant Gauges and Coulomb Gauge Limit, Nucl.Phys. B 548 (1999) 527. hep-th/ 9807024. relax O. Piguet and A. Rouet, Symmetries In Perturbative Quantum Field Theory, Phys. Rept. 76 (1981) 1; C. Becchi, ‘Lectures On The Renormalization Of Gauge Theories, In Les Houches 1983, Proceedings, Relativity, Groups and Topology, II, 787-821; L. Baulieu, Perturbative Gauge Theories, Phys. Rept. 129 (1985) 1; O. Piguet and S. P. Sorella, Algebraic renormalization: Perturbative renormalization, symmetries and anomalies, Berlin, Germany: Springer (1995) (Lecture notes in physics: m28). relax A. Blasi, O. Piguet and S. P. Sorella, Landau gauge and finiteness, Nucl. Phys. B356 (1991) 154. relax P. A. Grassi, The Abelian anti-ghost equation for the standard model in the ’t Hooft background gauge, Nucl. Phys. B537 (1999) 527. relax N. K. Nielsen, On The Gauge Dependence Of Spontaneous Symmetry Breaking In Gauge Theories, Nucl. Phys. B101, 173 (1975), H. Kluberg-Stern and J. B. Zuber, Renormalization Of Nonabelian Gauge Theories In A Background Field Gauge. 2. Gauge Invariant Operators, Phys. Rev. D12, 3159 (1975); Renormalization Of Nonabelian Gauge Theories In A Background Field Gauge: 1. Green Functions, Phys. Rev. D12, 467 (1975); Ward Identities And Some Clues To The Renormalization Of Gauge Invariant Operators, D12, 482 (1975); O. Piguet and K. Sibold, Gauge Independence In Ordinary Yang-Mills Theories, Nucl. Phys. B 253, 517 (1985); O. M. Del Cima, D. H. Franco and O. Piguet, Gauge independence of the effective potential revisited Nucl. Phys. B 551, 813 (1999), P. Gambino and P. A. Grassi, The Nielsen identities of the SM and the definition of mass, hep-ph/9907254, to appear in Phys. Rev. D. relax D. Zwanziger, Vanishing of zero-momentum lattice gluon propagator and color confinement, Nucl.Phys. B 364 (1991) 127.
|
warning/0006/hep-th0006246.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Gauge theories on non-commutative spaces cannot be formulated with Lie algebra valued infinitesimal transformations and consequently not with Lie algebra valued gauge fields. In the composition of the infinitesimal transformations commutators and anticommutators of the generators of the gauge group will appear, eventually generating all the higher powers of the generators. Thus the enveloping algebra of the Lie algebra seems to be a proper setting for such a gauge theory. This, however, is not very attractive because the enveloping algebra is infinite-dimensional, requiring an infinite number of coordinate dependent transformation parameters and an infinite number of gauge fields as a consequence.
In this paper we show that enveloping algebra valued infinitesimal transformations as well as enveloping algebra valued gauge fields can be restricted such that they depend on the Lie algebra valued parameters and the Lie algebra valued gauge fields and their space-time derivatives only. This renders the number of independent parameters and gauge fields to be the same as for the Lie algebra valued gauge theories. The coefficient functions of all the higher powers of the generators of the gauge group are functions of the coefficients of the first power. The construction of the dependent coefficients is based on the Seiberg-Witten map . The existence of this map can be proven in general , here we demonstrate this map by explicitely calculating the expansion to first order in a parameter that characterizes the deviation from commuting coordinates.
As a method we use the $``$-product formulation of the algebra . The objects are functions of commuting variables, the algebraic non-commutative properties are encoded in the $``$-product. In the following chapter we introduce this formalism, it can be used for all algebras that have the Poincare-Birkhoff-Witt property. In this paper we restrict the algebra to the algebra of the non-commuting coordinates and the Lie algebra of the gauge group.
For non-commuting spaces the concept of a gauge theory can already be introduced by defining covariant coordinates without speaking about derivatives . In general the algebraic setting of the theory will require an extension of the algebra by derivatives. This formalism has been developed for quantum planes .
For the canonical structure of the non-commuting coordinates it is shown in this paper that the derivatives can be obtained from the coordinates. Derivatives do not have to be introduced separately. For several examples of quantum planes this is true as well .
For the canonical structure integration can be defined as well. This is shown in chapter 4. This allows us to formulate the dynamics with an action. A gauge invariant action for the gauge field can be constructed from the gauge covariant tensors that agrees with the usual gauge invariant action in the limit of commuting spaces.
The approach generalizes to other non-commutative spaces and it is in particular possible to choose another non-commutative internal space instead of the Lie algebra. For an internal canonical structure this scenario has been studied in .
## 2 Non-commutative Spaces and the $``$-Product Formalism
The coordinates $`\widehat{z}^i`$, $`(i=1,\mathrm{}N)`$ of a non-commutative space structure are subject to relations. We have in mind the relations for a canonical structure
$$[\widehat{z}^i,\widehat{z}^j]=i\theta ^{ij},\theta ^{ij},$$
(2.1)
for a Lie structure
$$[\widehat{z}^i,\widehat{z}^j]=if^{ij}{}_{k}{}^{}\widehat{z}_{}^{k},f^{ij}{}_{k}{}^{},$$
(2.2)
or for a quantum plane structure
$$[\widehat{z}^i,\widehat{z}^j]=iC_{kl}^{ij}\widehat{z}^k\widehat{z}^l,C_{kl}^{ij}.$$
(2.3)
The non-commutative space can be defined as the associative algebra over $``$, which consists of the algebra freely generated by the coordinates and then divided by the ideal $``$ generated by the relations:
$$𝒜_z=\frac{[[\widehat{z}^1,\mathrm{},\widehat{z}^N]]}{}.$$
(2.4)
Formal power series are accepted. Among these algebras we will restrict our attention to those that have a basis with the Poincare-Birkhoff-Witt property (PBW). This means that when considered as a graded algebra the subspace of polynomials of fixed degree has the same dimension as the corresponding subspace of the polynomials of commuting variables. In this case any element of $`𝒜_z`$ is defined by its coefficient function and vice versa.
$`\widehat{f}={\displaystyle \underset{L=0}{\overset{\mathrm{}}{}}}f_{i_1,\mathrm{},i_L}:\widehat{z}^{i_1}\mathrm{}\widehat{z}^{i_L}:`$
$`\widehat{f}\{f_i\}.`$ (2.5)
$`:\widehat{z}^{i_1}\mathrm{}\widehat{z}^{i_L}:`$ denotes an element of the basis defined by some prescribed ordering, e.g., normal order $`i_1i_2\mathrm{}i_L`$ or, e.g., totaly symmetric. The product of two elements will have its own coefficient function, this defines the diamond product
$$\widehat{f}\widehat{g}=\widehat{h}\{f_i\}\{g_i\}=\{h_i\}.$$
(2.6)
The algebraic properties are now all encoded in the $``$ product.
Next we associate a function $`f`$ of commuting variables with an element of the algebra, say $`\widehat{f}`$, by substituting the commuting variable $`z^1,\mathrm{},z^N`$ for the non-commuting variables in (2)
$$\widehat{f}=f_{i_1\mathrm{}i_L}:\widehat{z}^{i_1}\mathrm{}\widehat{z}^{i_L}:f(z)=f_{i_1\mathrm{}i_L}z^{i_1}\mathrm{}z^{i_L}$$
(2.7)
The diamond product leads to a bilinear $``$-product of functions:
$$\{f_i\}\{g_i\}=\{h_i\}(fg)(z)=h(z).$$
(2.8)
This star product has been discussed in reference .
For the canonical structure it is the Moyal-Weyl product :
$$fg=e^{\frac{i}{2}\frac{}{z^i}\theta ^{ij}\frac{}{z^j}}f(z)g(z^{})|_{z^{}z}.$$
(2.9)
For the Lie structure we have:
$$fg=e^{\frac{i}{2}x^lg_l(i\frac{}{z^{}},i\frac{}{z^{\prime \prime }})}f(z^{})g(z^{\prime \prime })|_{\genfrac{}{}{0pt}{}{z^{}z}{z^{\prime \prime }z}},$$
(2.10)
where $`g_l`$ is defined by group multiplication:
$$e^{ik_l\widehat{z}^l}e^{ip_l\widehat{z}^l}=e^{i\{k_l+p_l+\frac{1}{2}g_l(k,p)\}\widehat{z}^l}.$$
(2.11)
The first terms are easily calculated from the Baker-Campbell-Hausdorff formula:
$`e^Ae^B`$ $`=`$ $`e^{A+B+\frac{1}{2}[A,B]+\frac{1}{12}([A,[A,B]]+[B,[B,A]])+\mathrm{}}`$
$`g_l(k,p)`$ $`=`$ $`k_ip_jf^{ij}{}_{l}{}^{}+{\displaystyle \frac{1}{6}}k_ip_j(p_kk_k)f^{ij}{}_{n}{}^{}f_{}^{nk}{}_{l}{}^{}+\mathrm{}.`$ (2.12)
For the quantum plane structure we have as an example the $``$-product for the Manin plane:
$`xy`$ $`=`$ $`qyx`$
$`fg`$ $`=`$ $`q^{\frac{1}{2}(x^{}\frac{}{x^{}}y\frac{}{y}+x\frac{}{x}y^{}\frac{}{y^{}})}f(x,y)g(x^{},y^{})|_{\genfrac{}{}{0pt}{}{x^{}x}{y^{}y}}`$ (2.13)
## 3 Enveloping algebra valued connection
A non-abelian gauge theory on a non-commutative space carries two algebraic structures, the algebra $`𝒜_x`$ discussed above and the non-abelian Lie algebra $`𝒜_T`$ of the gauge group with the generators $`T^1,\mathrm{},T^M`$ and the relations:
$$[T^a,T^b]=if^{ab}{}_{c}{}^{}T_{}^{c}.$$
(3.1)
It is natural to treat both algebras on the same footing and to denote the generating elements of the big algebra by $`\widehat{z}^i`$:
$`\widehat{z}^i`$ $`=`$ $`\{\widehat{x}^1,\mathrm{},\widehat{x}^N,T^1,\mathrm{},T^M\}`$
$`𝒜_z`$ $`=`$ $`{\displaystyle \frac{[[\widehat{z}^1,\mathrm{},\widehat{z}^{N+M}]]}{}}.`$ (3.2)
The $``$-product formalism as developed in the previous chapter can now be applied to the algebra $`𝒜_z`$ as well.
We study functions of the commuting variables $`x^\nu `$, $`(\nu =1,\mathrm{},N)`$ and $`t^a`$, $`(a=1,\mathrm{},M)`$ and define the star product reflecting the algebraic properties of the algebra $`𝒜_z`$.
In the case of a canonical structure for the space variables $`x^\nu `$ we have
$`(FG)(z)=`$ (3.3)
$`e^{\frac{i}{2}\left(\theta ^{\mu \nu }\frac{}{x^\mu }\frac{}{x^{\prime \prime \nu }}+t^ag_a(i\frac{}{t^{}},i\frac{}{t^{\prime \prime }})\right)}`$
$`\times F(x^{},t^{})G(x^{\prime \prime },t^{\prime \prime })|_{\genfrac{}{}{0pt}{}{x^{}x,x^{\prime \prime }x}{t^{}t,t^{\prime \prime }t}}.`$
To exemplify the formalism we shall concentrate on this structure in what follows.
To define gauge theories we first define fields. These are elements of the algebra $`𝒜_x`$ that form a representation of the $`T`$-algebra. Under a gauge transformation they transform as follows:
$$\delta \widehat{\psi }=i\widehat{\alpha }\widehat{\psi },\widehat{\psi }𝒜_x,\widehat{\alpha }𝒜_z.$$
(3.4)
The action of the generators $`T`$ of the Lie algebra on $`\widehat{\psi }`$ is defined as $`\widehat{\psi }`$ is supposed to form a representation of $`𝒜_T`$. Thus $`\delta \widehat{\psi }𝒜_x`$ despite $`\widehat{\alpha }𝒜_z`$.
Independent of a representation we have defined $`\widehat{\alpha }`$ as an element of the enveloping algebra of the gauge group and not as Lie algebra-valued, as we would have done it for commuting spaces. We say $`\widehat{\alpha }`$ is enveloping algebra-valued. The same will be true for the connection that we introduce to define covariant coordinates
$$\widehat{X}^\nu =\widehat{x}^\nu +\widehat{A}^\nu ,\widehat{A}^\nu 𝒜_z.$$
(3.5)
We demand that $`\widehat{X}^\nu \widehat{\psi }`$ transforms covariantly:
$$\delta \widehat{X}^\nu \widehat{\psi }=i\widehat{\alpha }\widehat{X}^\nu \widehat{\psi }$$
(3.6)
and find that this defines the transformation law of the enveloping algebra-valued connection $`\widehat{A}^\nu `$:
$`\delta \widehat{A}^\nu =i[\widehat{x}^\nu ,\widehat{\alpha }]+i[\widehat{\alpha },\widehat{A}^\nu ],`$
$`\widehat{A}^\nu 𝒜_z,\widehat{\alpha }𝒜_z,\delta \widehat{A}^\nu 𝒜_z.`$ (3.7)
At first sight it seems that an enveloping algebra-valued connection has infinitely many component fields and thus is not very useful. We shall show, however, that all the component fields can be obtained from a Lie algebra-valued connection by a Seiberg-Witten map . This was also observed in , where a result in this direction has been obtained for SO(n) and Sp(n). To show this we cast the algebraic setting into the $``$-product formalism. The transformation of the connection is then
$$\delta A^\nu =i[x^\nu \stackrel{}{,}\alpha ]+i[\alpha \stackrel{}{,}A^\nu ].$$
(3.8)
We treat the canonical case in more detail, the $``$-product in this case is given in Equation (3.3).
For the first term in the variation of $`A^\nu `$ we obtain
$$i[x^\nu \stackrel{}{,}\alpha ]=\theta ^{\nu \rho }\frac{}{x^\rho }\alpha .$$
(3.9)
The variation of $`A^\nu `$ itself starts with a linear term in $`\theta `$, we therefore assume, as in reference , that $`A^\nu `$ starts with a linear term in $`\theta `$ as well:
$`A^\nu `$ $`=`$ $`\theta ^{\nu \rho }V_\rho `$
$`\delta V_\rho `$ $`=`$ $`{\displaystyle \frac{}{x^\rho }}\alpha +i[\alpha \stackrel{}{,}V_\rho ].`$ (3.10)
As in reference we expand in $`\theta `$, but not in $`g_a`$:
$`fg=\left\{1+{\displaystyle \frac{i}{2}}{\displaystyle \frac{}{x^\nu }}\theta ^{\nu \mu }{\displaystyle \frac{}{x^\mu }}+\mathrm{}\right\}f(x,t^{})g(x^{},t^{\prime \prime })|_{\genfrac{}{}{0pt}{}{x^{}x}{t^{}t,t^{\prime \prime }t}}`$
$`f(x,t^{})g(x^{},t^{\prime \prime })=e^{\frac{i}{2}t^ag_a(i\frac{}{t^{}},i\frac{}{t^{\prime \prime }})}f(x,t^{})g(x^{},t^{\prime \prime }).`$ (3.11)
We first treat Equation (3) to zeroth order in $`\theta `$ and show that it can be solved by assuming $`V`$ and $`\alpha `$ to be linear in $`t`$. This has to be expected as $`\theta =0`$ corresponds to the usual gauge theory on commuting spaces where the infinitesimal transformation and the connection are Lie algebra-valued. To zeroth order:
$`\alpha `$ $`=`$ $`\alpha _a^1t^a,`$
$`V_\rho `$ $`=`$ $`a_{\rho ,a}^1t^a.`$ (3.12)
From (3) we obtain, as expected :
$$\delta a_{\rho ,a}^1=\frac{\alpha _a^1}{x^\rho }f^{bc}{}_{a}{}^{}\alpha _{b}^{1}a_{\rho ,c}^1.$$
(3.13)
We turn to first order in $`\theta `$ in the variation of $`V_\rho `$, Equation (3). The contributions that come from the zero order terms of $`\alpha `$ and $`V_\rho `$ are at most of second order in $`t`$. This is the case because $`t^ag_a(i\frac{}{t^{}},i\frac{}{t^{\prime \prime }})`$ reduces the power of $`t`$ by at least one. The terms of order zero in $`g_a`$ actually contribute exactly in order $`t^2`$. Their contribution to $`\delta V_\rho `$ is
$$\delta V_\rho =\theta ^{\nu \mu }_\nu \alpha _a^1_\mu a_{\rho ,b}^1t^at^b+\mathrm{}.$$
(3.14)
If we now assume that the terms of $`\alpha `$ and $`V_\rho `$ linear in $`\theta `$ are all of second order in $`t`$ we get a consistent set of equations. We define to first order in $`\theta `$:
$`\alpha `$ $`=`$ $`\alpha _a^1t^a+\alpha _{ab}^2t^at^b+\mathrm{}`$
$`V_\rho `$ $`=`$ $`a_{\rho ,a}^1t^a+a_{\rho ,ab}^2t^at^b+\mathrm{}`$ (3.15)
$`\alpha ^1`$ and $`a^1`$ are of order zero in $`\theta `$ and $`\alpha ^2`$ and $`a^2`$ of first order. This expansion in $`t`$ leads to an expansion in $`g_a`$ of the $``$-product, because higher order $`t`$-derivatives vanish.
In the calculation that now follows we have to use $`g_a`$ to the order given in (2). The term with three derivatives, however, vanishes on commutators of $``$-products because it is symmetric under the exchange of $`k`$ and $`p`$.
The result of the calculation is:
$`\delta a_{\rho ,ab}^2t^at^b`$ $`=`$ $`_\rho \alpha _{ab}^2t^at^b`$ (3.16)
$`\theta ^{\nu \mu }_\nu \alpha _a^1_\mu a_{\rho ,b}^1t^at^b`$
$`2f^{bc}{}_{a}{}^{}\{\alpha _b^1a_{\rho ,cd}^2+\alpha _{bd}^2a_{\rho ,c}^1\}t^dt^a.`$
This can be brought closer to the form of reference . We introduce the Lie algebra valued $`\epsilon `$ and the enveloping algebra-valued $`G_\rho `$:
$$\epsilon =\alpha _b^1T^b,G_\rho =a_{\rho ,cd}^2T^cT^d$$
(3.17)
and compute the commutator
$$i[\epsilon ,G_\rho ]=\alpha _b^1a_{\rho ,cd}^2f^{bc}{}_{l}{}^{}\{T^lT^d+T^dT^l\}.$$
(3.18)
This is true because $`a_{\rho ,cd}^2`$ is symmetric in $`c`$ and $`d`$. Now we remember that we have used a star product that corresponds to a completely symmetrical version of the monomials of the bases. Thus we have to replace
$$\frac{1}{2}\{T^lT^d+T^dT^l\}t^lt^d$$
(3.19)
and obtain from (3.18) the corresponding term in (3.16). The other term derives from the commutator
$`i[\gamma ,a_\rho ],\text{with}`$
$`\gamma =\alpha _{bd}^2T^bT^d,a_\rho =a_{\rho ,l}^1T^l.`$ (3.20)
This is exactly the structure as in reference :
$`\delta G_\nu `$ $`=`$ $`_\nu \gamma {\displaystyle \frac{1}{2}}\theta ^{\kappa \lambda }\{_\kappa \epsilon ,_\lambda a_\nu \}`$ (3.21)
$`+i[\epsilon ,G_\nu ]+i[\gamma ,a_\nu ].`$
It has the solution already found in .
$`\alpha _{ab}^2t^at^b`$ $`=`$ $`{\displaystyle \frac{1}{2}}\theta ^{\nu \mu }_\nu \alpha _a^1a_{\mu ,b}^1t^at^b`$
$`a_{\rho ,ab}^2t^at^b`$ $`=`$ $`{\displaystyle \frac{1}{2}}\theta ^{\nu \mu }a_{\nu ,a}^1(_\mu a_{\rho ,b}^1+F_{\mu \rho ,b}^1)t^at^b,`$ (3.22)
where $`F_{\nu \rho ,b}^1=_\nu a_{\mu ,b}^1_\mu a_{\nu ,b}^1+f^{cd}{}_{b}{}^{}a_{\nu ,c}^{1}a_{\mu ,d}^1`$.
The procedure can now be generalized to all the higher powers of $`\theta `$. We always assume that $`\alpha ^n`$ and $`a_\rho ^n`$ are of power $`\theta ^{n1}`$ and polynomials in $`t`$ of degree $`n`$. This amounts to a power series expansion in $`g_a`$ as well and the existence of the Seiberg-Witten map to all orders follows from the existence of the Seiberg-Witten map in the general setting of the previous chapter. This is discussed in detail in reference , where the full non-abelian Seiberg-Witten map is constructed.
It is straightforward to generalize this formalism to a $``$-product, with a coordinate dependent $`\theta `$ underlying the non-commutative, e.g. Lie or quantum plane, structure. We give the result:
$`\alpha `$ $`=`$ $`\alpha _a^1t^a+{\displaystyle \frac{1}{2}}\theta ^{\nu \mu }_\nu \alpha _a^1a_{\mu ,b}^1t^at^b`$
$`A^\nu `$ $`=`$ $`\theta ^{\nu \mu }a_{\mu ,a}^1t^a`$ (3.23)
$`{\displaystyle \frac{1}{2}}\theta ^{\sigma \mu }a_{\sigma ,a}^1\left(_\mu (\theta ^{\nu \rho }a_{\rho ,b}^1)+\theta ^{\nu \rho }F_{\mu \rho ,b}^1\right)t^at^b`$
For the Lie structure:
$$\theta ^{\mu \nu }=f^{\mu \nu }{}_{\kappa }{}^{}x_{}^{\kappa }.$$
(3.24)
For the quantum plane:
$$\theta ^{\mu \nu }=ihxy,h=\mathrm{ln}q.$$
(3.25)
Let us consider once more the gauge transformations (3.4). We see that $`\alpha `$ is not an arbitrary element of the algebra $`𝒜_z`$. There are only $`M`$ (dimension of the Lie group to be gauged) free parameters $`\alpha _l^1(x)`$, all the higher-order terms in the enveloping algebra can be expressed in terms of these parameters and the gauge field $`a_{\rho ,l}^1`$ and their derivatives. This can be achieved by the Seiberg-Witten map.
To summarize, the Lie algebra-valued term $`\epsilon =\alpha _a^1(x)T^a`$ of $`\alpha `$ determines all the other terms in the enveloping algebra:
$$\alpha =\alpha _a^1T^a+\frac{1}{4}\theta ^{\nu \mu }_\nu \alpha _a^1a_{\mu ,b}^1(T^aT^b+T^bT^a)+\mathrm{}$$
(3.26)
Thus a gauge transformation is determined by $`\alpha ^1(x)`$ and $`a_\rho ^1(x)`$, it is defined by:
$$\delta _{\alpha ^1}\psi =i\alpha (\alpha ^1,a_\rho ^1)\psi .$$
(3.27)
The composition of two transformations is defined for arbitrary enveloping algebra-valued transformations as follows:
$$\delta _\alpha \delta _\beta \delta _\beta \delta _\alpha =\delta _{i(\beta \alpha \alpha \beta )}.$$
(3.28)
We show that this is also true for the restricted transformation defined by $`\alpha ^1`$:
$$\delta _{\alpha ^1}\delta _{\beta ^1}\delta _{\beta ^1}\delta _{\alpha ^1}=\delta _{i(\beta \alpha \alpha \beta )^1}.$$
(3.29)
This reflects the compositon of standard Lie algebra-valued gauge transformations. We show this to first order in $`\theta `$, using the $``$-formalism.
$`\alpha `$ $`=`$ $`\alpha _a^1t^a+{\displaystyle \frac{1}{2}}\theta ^{\nu \mu }_\nu \alpha _a^1a_{\mu ,b}^1t^at^b`$
$`\beta `$ $`=`$ $`\beta _a^1t^a+{\displaystyle \frac{1}{2}}\theta ^{\nu \mu }_\nu \beta _a^1a_{\mu ,b}^1t^at^b`$ (3.30)
We compute $`[\alpha \stackrel{}{,}\beta ]`$ to first order in $`\theta `$ from Equation (3):
$`[\alpha \stackrel{}{,}\beta ]`$ $`=`$ $`i\alpha _a^1\beta _b^1f^{ab}{}_{c}{}^{}t_{}^{c}`$ (3.31)
$`+\theta ^{\nu \mu }\{{\displaystyle \frac{i}{2}}_\nu (\alpha _a^1\beta _b^1f^{ab}{}_{d}{}^{})a_{\mu ,c}`$
$`+{\displaystyle \frac{i}{2}}(\alpha _a^1_\nu \beta _d^1\beta _a^1_\nu \alpha _d^1)a_{\mu ,b}^1f^{ab}_c`$
$`+i_\nu \alpha _d^1_\mu \beta _c^1\}t^dt^c.`$
Now we compute from (3.4)
$$(\delta _{\beta ^1}\delta _{\alpha ^1}\delta _{\alpha ^1}\delta _{\beta ^1})\widehat{\psi }=(\widehat{\alpha }\widehat{\beta }\widehat{\beta }\widehat{\alpha })\widehat{\psi }+i((\delta _{\beta ^1}\widehat{\alpha })(\delta _{\alpha ^1}\widehat{\beta }))\widehat{\psi }.$$
(3.32)
The second term arises because $`\alpha `$ depends on the gauge fields $`a_\mu ^1`$ that transforms under gauge transformations
$$\delta _{\beta ^1}\alpha =\frac{1}{2}\theta ^{\rho \sigma }_\rho \alpha _a^1(_\sigma \beta _b^1f^{cd}{}_{b}{}^{}\beta _{c}^{1}a_{\sigma ,d}^1)t^dt^b.$$
(3.33)
We are now ready to compute $`\delta _{\beta ^1}\delta _{\alpha ^1}\delta _{\alpha ^1}\delta _{\beta ^1}`$ and obtain
$`\delta _{\beta ^1}\delta _{\alpha ^1}\delta _{\alpha ^1}\delta _{\beta ^1}=i(\delta _{\beta ^1}\alpha \delta _{\alpha ^1}\beta )[\alpha \stackrel{}{,}\beta ]`$ (3.34)
$`=`$ $`(i\alpha _a^1\beta _b^1f^{ab}{}_{c}{}^{}t_{}^{c}+{\displaystyle \frac{1}{2}}\theta ^{\nu \mu }_\nu (i\alpha _a^1\beta _b^1f^{ab}{}_{d}{}^{})a_{\mu ,c}^1t^dt^c).`$
This is exactly the formula (3.29) that we obtain if we start from the Lie algebra-valued part of $`[\alpha \stackrel{}{,}\beta ]`$
$$[\alpha \stackrel{}{,}\beta ]=i\alpha _a^1\beta _b^1f^{ab}{}_{c}{}^{}T_{}^{c}+\mathrm{}$$
(3.35)
This shows that our restricted enveloping algebra-valued form of the parameters is respected by the commutator of two transformations.
In reference we have introduced tensors
$$\widehat{T}^{\mu \nu }=[\widehat{X}^\mu ,\widehat{X}^\nu ]i\widehat{\theta }^{\mu \nu },$$
(3.36)
that transform
$$\delta \widehat{T}^{\mu \nu }=i[\widehat{\alpha },\widehat{T}^{\mu \nu }]$$
(3.37)
under the general enveloping algebra valued gauge transformation. $`\widehat{\theta }^{\mu \nu }`$ is the respective term for all the three structures, canonical, Lie and quantum plane. Equation (3.37) will also be true for our restricted gauge transformation (3.27).
For the canonical case to exemplify:
$`T^{\mu \nu }`$ $`=`$ $`i\theta ^{\mu \kappa }_\kappa A^\nu i\theta ^{\nu \lambda }_\lambda A^\mu +A^\mu A^\nu A^\nu A^\mu `$ (3.38)
$`=`$ $`\theta ^{\mu \kappa }\theta ^{\nu \lambda }\left\{_\kappa V_\lambda _\lambda V_\kappa +V_\kappa V_\lambda V_\lambda V_\kappa \right\}.`$
It is natural to introduce the field strength
$$F_{\kappa \lambda }=_\kappa V_\lambda _\lambda V_\kappa +V_\kappa V_\lambda V_\lambda V_\kappa .$$
(3.39)
It is easy to compute the first order correction to the classical field strength $`F^1`$ defined after Equation (3):
$`F_{\kappa \lambda ,a}t^a`$ $`=`$ $`F_{\kappa \lambda ,a}^1t^a`$ (3.40)
$`+\theta ^{\mu \nu }\{F_{\kappa \mu ,a}^1F_{\lambda \nu ,b}^1`$
$`{\displaystyle \frac{1}{2}}a_{\mu ,a}^1((D_\nu F_{\kappa \lambda }^1)_b+_\nu F_{\kappa \lambda ,b}^1)\}t^at^b.`$
In this formula a covariant derivative of $`F^1`$ is used: $`(D_\nu F_{\kappa \lambda }^1)_b=_\nu F_{\kappa \lambda ,b}^1+a_{\nu ,c}F_{\kappa \lambda ,d}^1f^{cd}_b`$. This expression can also be obtained from reference .
Using the transformation law for $`\alpha `$ (3.33) and $`a^1`$ (3.13) we find as expected
$$\delta _{\alpha ^1}F_{\kappa \lambda }=i[\alpha \stackrel{}{,}F_{\kappa \lambda }],$$
(3.41)
with the restricted form of $`\alpha `$.
## 4 Gauge covariant dynamics
The $``$-formalism can be used to formulate a dynamics on non-commutative spaces. The coefficient functions $`f(x)`$ are the objects for which dynamical laws can be defined. In general we have to enlarge the algebra by derivatives for this purpose. For the quantum plane structure such derivatives have been introduced in in a purely algebraic approach. For the algebra extended by derivatives the formalism developed in the second chapter can be used.
For the canonical structure we can define derivates following the same strategy as for the quantum plane structure. Derivatives have to be defined in such a way, that they do not lead to new relations for the coordinates. Proceeding this way we can define a Leibniz rule:
$$\widehat{}_\mu \widehat{x}^\nu =\delta _\mu ^\nu +d_{\mu \sigma }^{\nu \rho }\widehat{x}^\sigma \widehat{}_\rho ,$$
(4.1)
where the coefficients $`d_{\mu \sigma }^{\nu \rho }`$ have to be choosen in such a way that
$`\widehat{}_\rho \left\{[\widehat{x}^\mu ,\widehat{x}^\nu ]i\theta ^{\mu \nu }\right\}=`$ (4.2)
$`=`$ $`\delta _\rho ^\mu \widehat{x}^\nu \delta _\rho ^\nu \widehat{x}^\mu +d_{\rho \kappa }^{\mu \nu }\widehat{x}^\kappa d_{\rho \kappa }^{\nu \mu }\widehat{x}^\kappa `$
$`+\widehat{x}^\kappa \widehat{x}^\beta \left\{d_{\rho \kappa }^{\mu \sigma }d_{\sigma \beta }^{\nu \alpha }d_{\rho \kappa }^{\nu \sigma }d_{\sigma \rho }^{\mu \alpha }\right\}\widehat{}_\alpha `$
$`i\theta ^{\mu \nu }\widehat{}_\rho `$
does not lead to new relations when $`\widehat{}`$ is brought to the right hand side. This is the case if we define
$$d_{\mu \sigma }^{\nu \rho }=\delta _\sigma ^\nu \delta _\mu ^\rho ,$$
(4.3)
or simply
$$\widehat{}_\rho \widehat{x}^\mu =\delta _\rho ^\mu +\widehat{x}^\mu \widehat{}_\rho .$$
(4.4)
If we now compare this with (2.1) we see that
$$\widehat{x}^\alpha i\theta ^{\alpha \rho }\widehat{}_\rho $$
(4.5)
commutes with all coordinates. This allows us to divide the algebra by the ideal generated by the relation (see also )
$$\widehat{x}^\alpha i\theta ^{\alpha \rho }\widehat{}_\rho =0.$$
(4.6)
The star product, known for the coordinates is now defined for the derivations as well
$`_\rho f`$ $`=`$ $`i\theta _{\rho \sigma }^1x^\sigma f`$ (4.7)
$`=`$ $`i\theta _{\rho \sigma }^1\left([x^\sigma \stackrel{}{,}f]+fx^\sigma \right)`$
$`=`$ $`{\displaystyle \frac{}{x^\rho }}f+f_\rho .`$
We have used (3.9).
For the canonical structure an integral can be defined:
$$\widehat{f}=d^Nxf(x^1,\mathrm{}x^N).$$
(4.8)
For the moment it is simpler to consider just functions of $`x^i`$ that do not depend on the variables $`t^a`$ as well. The $``$-product simply is the one of (2.9):
$`fg`$ $`=`$ $`e^{\frac{i}{2}\frac{}{x^i}\theta ^{ij}\frac{}{x^j}}f(x)g(x^{})|_{x^{}x}`$ (4.9)
$`=`$ $`{\displaystyle d^Nx^{}\delta (\stackrel{}{x}\stackrel{}{x}^{})e^{\frac{i}{2}\frac{}{x^i}\theta ^{ij}\frac{}{x^j}}f(x)g(x^{})}`$
$`=`$ .
$`\delta (\stackrel{}{x}\stackrel{}{x}^{})`$ is the product of $`N`$ $`\delta `$-functions:
$$\delta (\stackrel{}{x}\stackrel{}{x}^{})=\delta (x^1x^1)\mathrm{}\delta (x^Nx^N).$$
(4.10)
We now show that
$$\widehat{f}\widehat{g}=\widehat{g}\widehat{f}.$$
(4.11)
We use the $``$-formalism:
$$\widehat{f}\widehat{g}=d^Nxd^Nx^{}\delta (\stackrel{}{x}\stackrel{}{x}^{})e^{\frac{i}{2}\frac{}{x^i}\theta ^{ij}\frac{}{x^j}}f(x)g(x^{}).$$
(4.12)
Partial integration:
$$\widehat{f}\widehat{g}=d^Nxd^Nx^{}f(x)g(x^{})e^{\frac{i}{2}\frac{}{x^i}\theta ^{ij}\frac{}{x^j}}\delta (\stackrel{}{x}\stackrel{}{x}^{}).$$
(4.13)
The $`\delta `$-function depends on $`\stackrel{}{x}\stackrel{}{x}^{}`$. Thus
$$\frac{}{x^l}\delta (\stackrel{}{x}\stackrel{}{x}^{})=\frac{}{x^l}\delta (\stackrel{}{x}\stackrel{}{x}^{}).$$
(4.14)
This leads to
$$\widehat{f}\widehat{g}=d^Nxd^Nx^{}f(x)g(x^{})e^{\frac{i}{2}\frac{}{x^i}\theta ^{ij}\frac{}{x^j}}\delta (\stackrel{}{x}\stackrel{}{x}^{}).$$
(4.15)
Partial integration again:
$`{\displaystyle \widehat{f}\widehat{g}}`$ $`=`$ $`{\displaystyle d^Nxd^Nx^{}\delta (\stackrel{}{x}\stackrel{}{x}^{})e^{\frac{i}{2}\frac{}{x^i}\theta ^{ij}\frac{}{x^j}}g(x^{})f(x)}`$ (4.16)
$`=`$ $`{\displaystyle \widehat{g}\widehat{f}}.`$
We also find Stokes theorem
$$[\widehat{}_l,\widehat{f}]=d^Nx[_l\stackrel{}{,}f]=d^Nx\frac{}{x^l}f=0$$
(4.17)
from (4.7) for functions that vanish at the boundary. Partial integration of the $``$-derivative follows now from (4.17) and the Leibniz rule
$$[_l\stackrel{}{,}(fg)]=([_l\stackrel{}{,}f])g+f([_l\stackrel{}{,}g]),$$
(4.18)
which is true, because $`\theta `$ is $`x`$ independent.
With this integral we can define an action. A tensor that transforms as in (3.37):
$$\delta \widehat{L}=i[\widehat{\alpha },\widehat{L}]$$
(4.19)
will lead to a gauge invariant action:
$$W=d^Nx\text{Tr}\widehat{L},\delta W=0.$$
(4.20)
The trace has to be taken for the group generators. A proper action would be
$$L=\frac{1}{4}F_{\kappa \lambda }F^{\kappa \lambda },$$
(4.21)
where $`F_{\kappa \lambda }`$ has been defined in (3.39). The first correction term in $`\theta `$ to the classical field strength has been computed in (3.40).
We have thus formulated dynamics on a non-commutative space entirely whithin the standard framework of quantum field theory. The method can be extended to the treatment of matter fields aswell.
|
warning/0006/hep-ph0006172.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Although isospin violation in nonleptonic weak interactions has been investigated many times in the past systematic treatments have appeared only rather recently . The topic is both of general interest and of considerable phenomenological relevance. Precise determinations of weak decay amplitudes are needed for many purposes, in particular for a reliable calculation of CP violation in the $`K^0\overline{K^0}`$ system. In the standard model, isospin violation arises from the quark mass difference $`m_um_d`$ and from electromagnetic corrections. Although these effects are expected to be small in general they are amplified in nonleptonic weak transitions. Because of the suppression of amplitudes with $`\mathrm{\Delta }I>1/2`$, isospin violation in the dominant $`\mathrm{\Delta }I=1/2`$ amplitudes leads to significantly enhanced corrections for the sub-dominant amplitudes. In fact, a quantitative analysis of the $`\mathrm{\Delta }I=1/2`$ rule is only possible with the inclusion of isospin-violating effects.
At first order in a systematic low-energy expansion, isospin breaking in the leading octet amplitudes of nonleptonic kaon decays is of order $`G_8(m_um_d)`$ and $`G_8e^2`$ where $`G_8`$ denotes the strength of the effective octet coupling. The corrections appear in the mass differences of charged and neutral mesons, via $`\pi ^0\eta `$ mixing and through electromagnetic penguins in the effective nonleptonic weak Hamiltonian. However, there are good reasons to believe that the problem cannot be understood at lowest order only. For instance, the resulting (tree-level) corrections do not produce a $`\mathrm{\Delta }I=5/2`$ component for which there is some phenomenological evidence .
The chiral realization of isospin violation due to the light quark mass difference is available also at next-to-leading order. The purpose of this paper is to close the gap in the electromagnetic sector by
* completing the construction of the effective chiral Lagrangian of $`𝒪(G_8e^2p^2)`$ and
* performing the complete renormalization at the one-loop level for nonleptonic weak transitions including electromagnetic corrections.
As our notation indicates, we only consider corrections to the leading octet part of the nonleptonic weak Hamiltonian. The results are applicable to the analysis of both $`K2\pi `$ and $`K3\pi `$ decays.
We start in Sec. 2 by recalling the ingredients for the construction of effective theories of strong, electromagnetic and nonleptonic weak interactions. In Sec. 3 we review the effective Lagrangian of lowest order. For this Lagrangian, we evaluate the one-loop divergence functional by standard heat-kernel techniques in Sec. 4. The new parts are terms of $`𝒪(G_8e^2p^2)`$ which arise also from using the equations of motion to transform to the standard bases for the nonleptonic weak Lagrangian of $`𝒪(G_8p^4)`$ and for the electromagnetic Lagrangian of $`𝒪(e^2p^2)`$ . In the following section we construct the complete and minimal Lagrangian of $`𝒪(G_8e^2p^2)`$ making use of CPS symmetry , Cayley-Hamilton relations, partial integration in the action and of the equations of motion. We order the terms in the effective Lagrangian according to their physical relevance: $`K\pi \pi `$ amplitudes receive contributions from the first 12 operators, the next two appear in $`K3\pi `$ and the rest turns out not to be relevant phenomenologically. In Sec. 6, we present the divergences for the three $`K2\pi `$ amplitudes and compare with the results of direct one-loop calculations . We summarize our findings in Sec. 7. Various quantities appearing in the heat-kernel expansion of the generating functional are collected in the Appendix.
## 2 Symmetries
For a complete treatment of isospin-breaking effects in nonleptonic kaon decays, an appropriate effective Lagrangian with the pseudoscalar octet and the photon as dynamical degrees of freedom has to be used. The symmetries of the standard model are serving as the basic guiding principles for its construction. Starting with QCD in the chiral limit $`m_u=m_d=m_s=0`$, the resulting symmetry under the chiral group $`G=SU(3)_L\times SU(3)_R`$ is spontaneously broken to $`SU(3)_V`$. The pseudoscalar mesons $`(\pi ,K,\eta )`$ are interpreted as the corresponding Goldstone fields $`\phi _i`$ ($`i=1,\mathrm{},8`$) acting as coordinates of the coset space $`SU(3)_L\times SU(3)_R/SU(3)_V`$. The coset variables $`u_{L,R}(\phi )`$ are transforming as
$`u_L(\phi )`$ $`\stackrel{G}{}`$ $`g_Lu_L(\phi )h(g,\phi )^1,`$
$`u_R(\phi )`$ $`\stackrel{G}{}`$ $`g_Ru_R(\phi )h(g,\phi )^1,`$
$`g=(g_L,g_R)`$ $``$ $`SU(3)_L\times SU(3)_R,`$ (2.1)
where $`h(g,\phi )`$ is the nonlinear realization of $`G`$ .
The photon field $`A_\mu `$ is introduced in
$$u_\mu =i[u_R^{}(_\mu ir_\mu )u_Ru_L^{}(_\mu il_\mu )u_L]$$
(2.2)
by adding appropriate terms to the usual external vector and axial-vector sources $`v_\mu `$, $`a_\mu `$:
$`l_\mu `$ $`=`$ $`v_\mu a_\mu eQ_LA_\mu ,`$
$`r_\mu `$ $`=`$ $`v_\mu +a_\mu eQ_RA_\mu .`$ (2.3)
The $`3\times 3`$ matrices $`Q_{L,R}`$ are spurion fields with the transformation properties
$$Q_L\stackrel{G}{}g_LQ_Lg_L^{},Q_R\stackrel{G}{}g_RQ_Rg_R^{}$$
(2.4)
under the chiral group. We also define
$$𝒬_L:=u_L^{}Q_Lu_L,𝒬_R:=u_R^{}Q_Ru_R$$
(2.5)
transforming as
$`𝒬_L`$ $`\stackrel{G}{}`$ $`h(g,\phi )𝒬_Lh(g,\phi )^1,`$
$`𝒬_R`$ $`\stackrel{G}{}`$ $`h(g,\phi )𝒬_Rh(g,\phi )^1.`$ (2.6)
At the end, one identifies $`Q_{L,R}`$ with the quark charge matrix
$$Q=\left[\begin{array}{ccc}2/3& 0& 0\\ 0& 1/3& 0\\ 0& 0& 1/3\end{array}\right].$$
(2.7)
External scalar and pseudoscalar sources are combined in
$$\chi =s+ip.$$
(2.8)
For the construction of the effective Lagrangian, it is convenient to introduce the quantities
$$\chi _\pm =u_R^{}\chi u_L\pm u_L^{}\chi ^{}u_R$$
(2.9)
with the same chiral transformation properties as $`𝒬_L,𝒬_R`$ in (2.6).
After integrating out the heavy degrees of freedom, the $`\mathrm{\Delta }S=1`$ weak interactions can be described in terms of an effective four-fermion Hamiltonian . With respect to the chiral group $`G`$, this effective Hamiltonian transforms as the direct sum
$$(8_L,1_R)+(27_L,1_R)+(8_L,8_R),$$
(2.10)
where the first piece, contributing only to $`\mathrm{\Delta }I=\frac{1}{2}`$ transitions, is largely dominant. In this work we shall consider only the electromagnetic corrections induced by the dominant octet part of the effective Hamiltonian. To this end we introduce a weak spurion $`\lambda `$ that is finally taken at
$$\lambda =\frac{\lambda _6i\lambda _7}{2}=\left[\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ 0& 1& 0\end{array}\right],$$
(2.11)
where $`\lambda _{6,7}`$ are Gell-Mann matrices. In analogy to (2.5) we also define
$$\mathrm{\Delta }:=u_L^{}\lambda u_L,$$
(2.12)
transforming again as in (2.6) under chiral transformations.
Although CP is broken by the weak interactions, the $`\mathrm{\Delta }S=1`$ transitions are still invariant under the so-called CPS symmetry : a CP transformation followed by a subsequent interchange of $`d`$ and $`s`$ quarks. This symmetry is also obeyed by strong and electromagnetic interactions, provided the 2-3 indices of the external fields are also exchanged appropriately (this implies, in particular, the exchange $`m_sm_d`$ in the mass terms). The explicit CPS transformation properties of the several building blocks introduced so far are given by
$`u_\mu (x)`$ $`\stackrel{CPS}{}`$ $`ϵ(\mu )Su_\mu ^T(\stackrel{~}{x})S,`$
$`\chi _\pm (x)`$ $`\stackrel{CPS}{}`$ $`\pm S\chi _\pm ^T(\stackrel{~}{x})S,`$
$`𝒬_{L,R}(x)`$ $`\stackrel{CPS}{}`$ $`S𝒬_{L,R}^T(\stackrel{~}{x})S,`$
$`\mathrm{\Delta }(x)`$ $`\stackrel{CPS}{}`$ $`S\mathrm{\Delta }^T(\stackrel{~}{x})S,`$ (2.13)
with
$$\stackrel{~}{x}=(x^0,\stackrel{}{x}),ϵ(0)=1,ϵ(1)=ϵ(2)=ϵ(3)=1,$$
(2.14)
and
$$S=\left[\begin{array}{ccc}1& 0& 0\\ 0& 0& 1\\ 0& 1& 0\end{array}\right].$$
(2.15)
## 3 The effective Lagrangian at lowest order
With the building blocks introduced in the previous section we may now assemble our effective Lagrangian. We adopt an expansion scheme where the n-th order is related to terms of order $`p^n`$ in the strong and weak sector and to terms of order $`e^2p^{n2}`$ in the electromagnetic sector where $`p`$ denotes a typical meson momentum. Terms of $`𝒪(e^4)`$ will be neglected throughout.
To lowest order $`(n=2)`$, our effective theory consists of the following parts: the strong sector is represented by the nonlinear sigma model in the presence of the external sources $`v_\mu ,a_\mu ,\chi `$ and the photon coupling introduced in (2):
$$\frac{F^2}{4}u_\mu u^\mu +\chi _+.$$
(3.1)
The symbol $``$ denotes the trace in three-dimensional flavour space and $`F`$ is the pion decay constant in the chiral limit. Explicit chiral symmetry breaking by the non-vanishing masses of the light quarks is achieved by evaluating the generating functional at
$$\chi =2B_{\mathrm{quark}}=2B\left[\begin{array}{ccc}m_u& 0& 0\\ 0& m_d& 0\\ 0& 0& m_s\end{array}\right].$$
(3.2)
The quantity $`B`$ is related to the quark condensate in the chiral limit by $`0|\overline{q}q|0=F^2B`$.
The $`(8_L,1_R)`$ piece of the nonleptonic weak interactions is represented by the well-known Cronin Lagrangian ,
$$F^2\mathrm{\Xi }u_\mu u^\mu ,\mathrm{\Xi }=G_8F^2\mathrm{\Delta }+\mathrm{h}.\mathrm{c}..$$
(3.3)
At lowest order, the parameter $`G_8`$ can be determined from $`K2\pi `$ decays to be $`|G_8|9\times 10^6\mathrm{GeV}^25(G_F/\sqrt{2})|V_{ud}V_{us}|`$.
Now also the electromagnetic interaction has to be included. Apart from the necessary modification in (2.2), we have to add a kinetic term for the photon field,
$$\frac{1}{4}F_{\mu \nu }F^{\mu \nu },F_{\mu \nu }=_\mu A_\nu _\nu A_\mu ,$$
(3.4)
and a strangeness-conserving term of $`𝒪(e^2p^0)`$ ,
$$e^2F^4Z𝒬_L𝒬_R.$$
(3.5)
The numerical value of the parameter $`Z`$ can be determined from the mass difference of charged and neutral pions. The relation $`M_{\pi ^\pm }^2M_{\pi ^0}^2=2e^2ZF^2`$ implies $`Z0.8`$.
Finally, we have to introduce a weak-electromagnetic term characterized by a coupling constant $`g_{\mathrm{ewk}}`$,
$$e^2F^4\mathrm{{\rm Y}}𝒬_R,\mathrm{{\rm Y}}=g_{\mathrm{ewk}}G_8F^2\mathrm{\Delta }+\mathrm{h}.\mathrm{c}..$$
(3.6)
Note that to lowest order only a single (linear independent) term of this type can be constructed once the relation
$$𝒬_L\mathrm{\Delta }=\mathrm{\Delta }𝒬_L=\frac{1}{3}\mathrm{\Delta }$$
(3.7)
is taken into account. This term is the lowest-order chiral realization of electromagnetic penguins and transforms as $`(8_L,8_R)`$ under $`G`$ when the $`Q_R`$ spurion field is “frozen” to the constant value (2.7). By chiral dimensional analysis we expect the coupling constant $`g_{\mathrm{ewk}}`$ to be of $`𝒪(1)`$. A recent estimate in Ref. corresponds in fact to $`g_{\mathrm{ewk}}=1.0\pm 0.3`$ (see also Ref. ).
Summing up all these contributions, our lowest-order effective Lagrangian assumes the form
$`_2`$ $`=`$ $`{\displaystyle \frac{F^2}{4}}u_\mu u^\mu +\chi _++F^2\mathrm{\Xi }u_\mu u^\mu `$ (3.8)
$`{\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }+e^2F^4Z𝒬_L𝒬_R+e^2F^4\mathrm{{\rm Y}}𝒬_R.`$
Using (2), one easily verifies that (3.8) is CPS invariant.
## 4 One-loop divergences
For the construction of the one-loop functional, we first add a gauge-breaking term (we are using the Feynman gauge) and external sources to (3.8):
$$_2_2\frac{1}{2}(_\mu A^\mu )^2J_\mu A^\mu .$$
(4.1)
Then we expand the lowest-order action associated with (4.1) around the solutions $`\phi _{\mathrm{cl}}`$, $`A_{\mathrm{cl}}^\mu `$ of the classical equations of motion. In the standard “gauge” $`u_R(\phi _{\mathrm{cl}})=u_L(\phi _{\mathrm{cl}})^{}=:u(\phi _{\mathrm{cl}})`$, a convenient choice of the pseudoscalar fluctuation variables $`\xi _i`$ ($`i=1,\mathrm{},8`$) is given by
$$u_R=u_{\mathrm{cl}}e^{i\xi _i\lambda _i/2F},u_L=u_{\mathrm{cl}}^{}e^{i\xi _i\lambda _i/2F},\xi _i(\phi _{\mathrm{cl}})=0,$$
(4.2)
with the Gell-Mann matrices $`\lambda _i`$ ($`i=1,\mathrm{},8`$). The photon field is decomposed as
$$A^\mu =A_{\mathrm{cl}}^\mu +\epsilon ^\mu $$
(4.3)
with a fluctuation field $`\epsilon ^\mu `$. In the following formulas, we shall drop the subscript “$`\mathrm{cl}`$” for simplicity. The classical equations of motion take the form
$`_\mu u^\mu `$ $`=`$ $`{\displaystyle \frac{i}{2}}\left(\chi _{}{\displaystyle \frac{1}{3}}\chi _{}\right)+2ie^2F^2Z[𝒬_R,𝒬_L]`$ (4.4)
$`+i[u_\mu u^\mu ,\mathrm{\Xi }]2(_\mu \{u^\mu ,\mathrm{\Xi }\}{\displaystyle \frac{1}{3}}_\mu \{u^\mu ,\mathrm{\Xi }\})`$
$`+2ie^2F^2[𝒬_R,\mathrm{{\rm Y}}],`$
$`\mathrm{}A_\mu `$ $`=`$ $`J_\mu +{\displaystyle \frac{eF^2}{2}}u_\mu (𝒬_R𝒬_L)+eF^2\mathrm{\Xi }\{𝒬_R𝒬_L,u_\mu \},`$ (4.5)
where
$`_\mu `$ $`=`$ $`_\mu +[\mathrm{\Gamma }_\mu ,],`$
$`\mathrm{\Gamma }_\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}[u^{}(_\mu ir_\mu )u+u(_\mu il_\mu )u^{}].`$ (4.6)
The solutions of (4.4) and (4.5) are uniquely determined functionals of the external sources $`v_\mu `$, $`a_\mu `$, $`\chi `$, $`J_\mu `$. (Note that the usual Feynman boundary conditions are always implicitly understood.)
Expanding (4.1) up to terms quadratic in the fields $`\xi _i`$, $`\epsilon _\mu `$, we obtain the second-order fluctuation Lagrangian $`^{(2)}`$. The one-loop functional $`W_{L=1}`$ is then given by the Gaussian functional integral
$`e^{iW_{L=1}}={\displaystyle [d\xi _id\epsilon _\mu ]e^{i{\scriptscriptstyle d^dx^{\left(2\right)}}}}.`$ (4.7)
In our case, $`^{(2)}`$ reads
$`^{(2)}`$ $`=`$ $`{\displaystyle \frac{F^2}{4}}_\mu \xi ^\mu \xi +{\displaystyle \frac{1}{2}}u_\mu \xi u^\mu \xi {\displaystyle \frac{1}{2}}(u_\mu u^\mu +\chi _+)\xi ^2`$ (4.8)
$`+e^2F^4Z\xi 𝒬_L\xi 𝒬_R{\displaystyle \frac{1}{2}}\xi ^2\{𝒬_L,𝒬_R\}`$
$`+{\displaystyle \frac{F^2}{4}}4\mathrm{\Xi }(_\mu \xi ^\mu \xi {\displaystyle \frac{1}{4}}\{u_\mu ,\{u^\mu ,\xi ^2\}\}+{\displaystyle \frac{1}{4}}\{u_\mu ,\xi \}\{u^\mu ,\xi \})2i[\xi ,\mathrm{\Xi }]\{_\mu \xi ,u^\mu \}`$
$`+e^2F^4\mathrm{{\rm Y}}(\xi 𝒬_R\xi {\displaystyle \frac{1}{2}}\{\xi ^2,𝒬_R\})`$
$`+{\displaystyle \frac{1}{2}}\epsilon _\mu \mathrm{}\epsilon ^\mu +{\displaystyle \frac{e^2F^2}{4}}(𝒬_R𝒬_L)^2\epsilon _\mu \epsilon ^\mu +e^2F^2\mathrm{\Xi }(𝒬_R𝒬_L)^2\epsilon _\mu \epsilon ^\mu `$
$`{\displaystyle \frac{ieF^2}{4}}[u_\mu ,𝒬_R+𝒬_L]\xi \epsilon ^\mu +{\displaystyle \frac{eF^2}{2}}(𝒬_R𝒬_L)_\mu \xi \epsilon ^\mu `$
$`{\displaystyle \frac{ieF^2}{2}}\mathrm{\Xi }\{[𝒬_R+𝒬_L,\xi ],u_\mu \}\epsilon ^\mu {\displaystyle \frac{ieF^2}{2}}[\xi ,\mathrm{\Xi }]\{𝒬_R𝒬_L,u_\mu \}\epsilon ^\mu `$
$`+eF^2\mathrm{\Xi }\{𝒬_R𝒬_L,_\mu \xi \}\epsilon ^\mu ,`$
where
$$\xi =\xi _i\lambda _i/F.$$
(4.9)
In the next step, we perform the field transformation
$$\xi \xi \{\mathrm{\Xi },\xi \}+\frac{2}{3}\mathrm{\Xi }\xi \mathrm{𝟏}.$$
(4.10)
Because of $`\mathrm{\Delta }=0`$, we do not pick up an additional contribution from the Jacobi determinant and the fluctuation Lagrangian (4.8) assumes the form
$$^{(2)}=\frac{1}{2}\xi _i(d_\mu d^\mu +\sigma )_{ij}\xi _j+\frac{1}{2}\epsilon _\mu (\mathrm{}+\kappa )\epsilon ^\mu +\epsilon _\mu a_i^\mu \xi _i+\epsilon _\mu b_id_{ij}^\mu \xi _j,$$
(4.11)
where
$$d_{ij}^\mu =\delta _{ij}^\mu +\gamma _{ij}^\mu .$$
(4.12)
The explicit expressions for $`\gamma _{ij}^\mu `$, $`\sigma _{ij}`$, $`\kappa `$, $`a_i^\mu `$, $`b_i`$ are given in the Appendix.
The divergent part of the one-loop functional,
$$W_{L=1}^{\mathrm{div}}=d^dx_{L=1}^{\mathrm{div}},$$
(4.13)
is determined by
$`_{L=1}^{\mathrm{div}}`$ $`=`$ $`{\displaystyle \frac{1}{(4\pi )^2(d4)}}[\mathrm{tr}({\displaystyle \frac{1}{12}}\gamma _{\mu \nu }\gamma ^{\mu \nu }+{\displaystyle \frac{1}{2}}\sigma ^2)a_i^\mu a_{\mu i}+a_i^\mu (d_\mu b)_i`$ (4.14)
$`+{\displaystyle \frac{1}{2}}(b_ib_i)^2b_i\sigma _{ij}b_j\kappa b_ib_i+2\kappa ^2],`$
where
$$\gamma _{\mu \nu }=_\mu \gamma _\nu _\nu \gamma _\mu +[\gamma _\mu ,\gamma _\nu ].$$
(4.15)
This formula can easily be derived from the well-known second Seeley-deWitt coefficient for bosonic systems .
## 5 The chiral Lagrangian at next-to-leading order
We are now in the position to construct the most general local action at next-to-leading order which will also renormalize the one-loop divergences discussed in the previous section.
The strong part of the local action of $`𝒪(p^4)`$ is, of course, generated by the well-known Gasser-Leutwyler Lagrangian associated with the low-energy constants $`L_1,\mathrm{},L_{12}`$. In the presence of virtual photons, the structure of the operators given in remains unchanged. The only necessary modification is the inclusion of the dynamical photon field in the generalized “sources” $`\mathrm{}_\mu `$ and $`r_\mu `$ (see (2)). The divergences corresponding to the strong sector of (4.14) are absorbed by the divergent parts of the $`L_i`$ . In the relevant case of chiral $`SU(3)`$, the strong terms generated by (4.14) can be written immediately as a linear combination of the $`𝒪(p^4)`$ operators of the Gasser-Leutwyler basis without using the equations of motion (4.4) or (4.5). Consequently, no additional (weak-)electromagnetic terms are induced at this point.
The strangeness-conserving terms of $`𝒪(e^2p^2)`$ have been constructed by Urech . His list of electromagnetic counterterms is associated with the coupling constants $`K_1,\mathrm{},K_{14}`$. In this case, (4.14) leads to that canonical basis only after the use of the equation of motion (4.4). In this way, also some divergent weak-electromagnetic contributions of $`𝒪(G_8e^2p^2)`$ are generated.
For the octet part of the nonleptonic weak Lagrangian of $`𝒪(G_Fp^4)`$ we refer to the standard form of Ecker, Kambor and Wyler with couplings $`N_1,\mathrm{},N_{37}`$. Again, because of the mismatch between (4.14) and the standard basis, the equation of motion has to be used and the (purely) electromagnetic piece in (4.4) induces divergent terms of $`𝒪(G_8e^2p^2)`$.
Finally, we have to construct the most general weak-electromagnetic Lagrangian of $`𝒪(G_8e^2p^2)`$. Some parts of this Lagrangian have appeared before in the literature . The complete minimal Lagrangian of $`𝒪(G_8e^2p^2)`$ takes the form
$$_{G_8e^2p^2}=G_8e^2F^4\underset{i=1}{\overset{32}{}}Z_iQ_i+\mathrm{h}.\mathrm{c}.,$$
(5.1)
with operators $`Q_i`$ of $`𝒪(p^2)`$ and dimensionless coupling constants $`Z_i`$. A linear independent set of operators is given by
$`Q_1`$ $`=`$ $`\mathrm{\Delta }\{𝒬_R,\chi _+\},`$
$`Q_2`$ $`=`$ $`\mathrm{\Delta }𝒬_R\chi _+,`$
$`Q_3`$ $`=`$ $`\mathrm{\Delta }𝒬_R\chi _+𝒬_R,`$
$`Q_4`$ $`=`$ $`\mathrm{\Delta }\chi _+𝒬_L𝒬_R,`$
$`Q_5`$ $`=`$ $`\mathrm{\Delta }u_\mu u^\mu ,`$
$`Q_6`$ $`=`$ $`\mathrm{\Delta }\{𝒬_R,u_\mu u^\mu \},`$
$`Q_7`$ $`=`$ $`\mathrm{\Delta }u_\mu u^\mu 𝒬_L𝒬_R,`$
$`Q_8`$ $`=`$ $`\mathrm{\Delta }u_\mu 𝒬_Lu^\mu ,`$
$`Q_9`$ $`=`$ $`\mathrm{\Delta }u_\mu 𝒬_Ru^\mu ,`$
$`Q_{10}`$ $`=`$ $`\mathrm{\Delta }u_\mu \{𝒬_L,𝒬_R\}u^\mu ,`$
$`Q_{11}`$ $`=`$ $`\mathrm{\Delta }\{𝒬_R,u_\mu \}𝒬_Lu^\mu ,`$
$`Q_{12}`$ $`=`$ $`\mathrm{\Delta }\{𝒬_R,u_\mu \}𝒬_Ru^\mu ,`$
$`Q_{13}`$ $`=`$ $`\mathrm{\Delta }𝒬_Ru_\mu u^\mu ,`$
$`Q_{14}`$ $`=`$ $`\mathrm{\Delta }𝒬_Ru_\mu u^\mu 𝒬_R,`$
$`Q_{15}`$ $`=`$ $`\mathrm{\Delta }𝒬_Ru_\mu u^\mu (𝒬_L𝒬_R),`$
$`Q_{16}`$ $`=`$ $`\mathrm{\Delta }\chi _+,`$
$`Q_{17}`$ $`=`$ $`{\displaystyle \frac{2}{3}}\mathrm{\Delta }\chi _++\mathrm{\Delta }\{𝒬_R,\chi _+\}+\mathrm{\Delta }[𝒬_R,\chi _{}],`$
$`Q_{18}`$ $`=`$ $`\mathrm{\Delta }\{𝒬_R,\chi _+\}{\displaystyle \frac{1}{3}}\mathrm{\Delta }[𝒬_R,\chi _{}]+\mathrm{\Delta }(\chi _{}𝒬_L𝒬_R𝒬_R𝒬_L\chi _{})`$
$`{\displaystyle \frac{4}{3}}\mathrm{\Delta }𝒬_R\chi _+\mathrm{\Delta }𝒬_R\chi _+𝒬_R+\mathrm{\Delta }\chi _+𝒬_L𝒬_R,`$
$`Q_{19}`$ $`=`$ $`\mathrm{\Delta }𝒬_R\chi _+(𝒬_L𝒬_R),`$
$`Q_{20}`$ $`=`$ $`i(\widehat{}_\mu \mathrm{\Delta })[𝒬_L,u^\mu ],`$
$`Q_{21}`$ $`=`$ $`i(\widehat{}_\mu \mathrm{\Delta })[𝒬_R,u^\mu ],`$
$`Q_{22}`$ $`=`$ $`i(\widehat{}_\mu \mathrm{\Delta })(𝒬_Lu^\mu 𝒬_R𝒬_Ru^\mu 𝒬_L),`$
$`Q_{23}`$ $`=`$ $`i(\widehat{}_\mu \mathrm{\Delta })(u^\mu 𝒬_L𝒬_R𝒬_R𝒬_Lu^\mu ),`$
$`Q_{24}`$ $`=`$ $`i\mathrm{\Delta }(u_\mu (\widehat{}^\mu 𝒬_L)𝒬_R𝒬_R(\widehat{}^\mu 𝒬_L)u_\mu ),`$
$`Q_{25}`$ $`=`$ $`i(\widehat{}_\mu \mathrm{\Delta })(u^\mu 𝒬_R𝒬_L𝒬_L𝒬_Ru^\mu ),`$
$`Q_{26}`$ $`=`$ $`i\mathrm{\Delta }(u_\mu 𝒬_R(\widehat{}^\mu 𝒬_R)(\widehat{}^\mu 𝒬_R)𝒬_Ru_\mu ),`$
$`Q_{27}`$ $`=`$ $`i\mathrm{\Delta }(𝒬_Ru_\mu (\widehat{}^\mu 𝒬_R)(\widehat{}^\mu 𝒬_R)u_\mu 𝒬_R),`$
$`Q_{28}`$ $`=`$ $`(\widehat{}_\mu \mathrm{\Delta })(\widehat{}^\mu 𝒬_L),`$
$`Q_{29}`$ $`=`$ $`(\widehat{}_\mu \mathrm{\Delta })(\widehat{}^\mu 𝒬_R),`$
$`Q_{30}`$ $`=`$ $`\mathrm{\Delta }(\widehat{}_\mu 𝒬_R)(\widehat{}^\mu 𝒬_R),`$
$`Q_{31}`$ $`=`$ $`(\widehat{}^\mu \mathrm{\Delta })\{\widehat{}_\mu 𝒬_L,𝒬_R\},`$
$`Q_{32}`$ $`=`$ $`(\widehat{}^\mu \mathrm{\Delta })\{\widehat{}_\mu 𝒬_R,𝒬_L\},`$ (5.2)
where
$`\widehat{}_\mu \mathrm{\Delta }`$ $`=`$ $`_\mu \mathrm{\Delta }+{\displaystyle \frac{i}{2}}[u_\mu ,\mathrm{\Delta }]=u(D_\mu \lambda )u^{},`$
$`\widehat{}_\mu 𝒬_L`$ $`=`$ $`_\mu 𝒬_L+{\displaystyle \frac{i}{2}}[u_\mu ,𝒬_L]=u(D_\mu Q_L)u^{},`$
$`\widehat{}_\mu 𝒬_R`$ $`=`$ $`_\mu 𝒬_R{\displaystyle \frac{i}{2}}[u_\mu ,𝒬_R]=u^{}(D_\mu Q_R)u,`$ (5.3)
with
$`D_\mu \lambda `$ $`=`$ $`_\mu \lambda i[l_\mu ,\lambda ],`$
$`D_\mu Q_L`$ $`=`$ $`_\mu Q_Li[l_\mu ,Q_L],`$
$`D_\mu Q_R`$ $`=`$ $`_\mu Q_Ri[r_\mu ,Q_R].`$ (5.4)
For the construction of the list of local terms (5) we have used CPS invariance, the relations (3.7) and
$$𝒬_{L,R}^2=\frac{2}{9}\mathrm{𝟏}+\frac{1}{3}𝒬_{L,R},$$
(5.5)
the Cayley-Hamilton formula, partial integration and the equations of motion (4.4).
If the spurion fields $`Q_{L,R}`$ and $`\lambda `$ are fixed to the constant values in (2.7) and (2.11), respectively, then $`_{G_8e^2p^2}`$ transforms under $`G`$ as
$$(8_L,1_R)+(8_L,8_R)+(27_L,1_R)+(27_L,8_R)+(8_L,27_R).$$
(5.6)
This structure is richer than the one of the $`𝒪(G_8)`$ terms in $`_2`$ and also of the weak four-fermion effective Hamiltonian . The last two pieces, in particular, which are responsible for $`\mathrm{\Delta }I=5/2`$ transitions, have no analog in the effective Hamiltonian of dimension six.
The operator $`Q_{16}`$ does not contribute to on-shell matrix elements . The terms $`Q_{17}`$, $`Q_{18}`$, $`Q_{19}`$ vanish for electrically neutral (pseudo)scalar sources,
$$[\chi ,Q]=0,$$
(5.7)
which is, of course, the case for all realistic physical processes. Also the operators $`Q_{20},\mathrm{}Q_{32}`$ are irrelevant for practical purposes. Because of (5) and (5), they contribute only in the presence of non-vanishing external (axial-)vector sources.
The coupling constants $`Z_1,\mathrm{},Z_{12}`$ appear in the amplitudes of $`K2\pi `$ decays. The operators $`Q_{13}`$ and $`Q_{14}`$ do not contribute to $`K2\pi `$ but they enter for $`K3\pi `$. $`Q_{15}`$ involves at least five pseudoscalar fields and is therefore irrelevant for $`K`$ decays. A few linear combinations of the operators in (5) were already given some time ago by de Rafael . His list was restricted to terms contributing to $`K2\pi `$, neglecting contributions $`M_\pi ^2`$ and those renormalizing $`G_8`$. A more recent extension of de Rafael’s list can be found in Ref. . However, their Lagrangian is still incomplete even for the $`K2\pi `$ amplitudes, as we shall discuss in the following section. There is in addition an obvious misprint in the operator multiplied by $`s_6`$ in , which would be in conflict with chiral symmetry. Some of the operators in (5) have also appeared in attempts to bosonize the $`\mathrm{\Delta }S=1`$ four-fermion effective Hamiltonian.
The low-energy couplings $`Z_i`$ are in general divergent. They absorb the divergences of the one-loop graphs via the renormalization
$`Z_i`$ $`=`$ $`Z_i^r(\mu )+z_i\mathrm{\Lambda }(\mu ),i=1,\mathrm{},32,`$
$`\mathrm{\Lambda }(\mu )`$ $`=`$ $`{\displaystyle \frac{\mu ^{d4}}{(4\pi )^2}}\left\{{\displaystyle \frac{1}{d4}}{\displaystyle \frac{1}{2}}[\mathrm{ln}(4\pi )+\mathrm{\Gamma }^{}(1)+1]\right\},`$ (5.8)
in the dimensional regularization scheme. The coefficients $`z_1,\mathrm{}z_{32}`$ are determined in such a way that the divergences generated by (4.14) are cancelled:
$$\begin{array}{cccc}z_1=\frac{17}{12}3Z+\frac{3}{2}g_{\mathrm{ewk}},\hfill & z_2=1+\frac{16}{3}Z+g_{\mathrm{ewk}},\hfill & z_3=\frac{3}{4}+7Z,\hfill & z_4=\frac{3}{4}7Z,\hfill \\ z_5=2,\hfill & z_6=\frac{7}{2}+5Z+\frac{3}{2}g_{\mathrm{ewk}},\hfill & z_7=\frac{3}{2}+5Z,\hfill & z_8=\frac{1}{2},\hfill \\ z_9=\frac{11}{6}+\frac{4}{3}Z+2g_{\mathrm{ewk}},\hfill & z_{10}=\frac{3}{2}Z,\hfill & z_{11}=\frac{3}{2}2Z,\hfill & z_{12}=\frac{3}{2},\hfill \\ z_{13}=\frac{35}{12}3Z+g_{\mathrm{ewk}},\hfill & z_{14}=3+15Z,\hfill & z_{15}=\frac{3}{2}+15Z,\hfill & z_{16}=\frac{4}{9}\frac{4}{3}Z,\hfill \\ z_{17}=\frac{2}{3}+2Z,\hfill & z_{18}=\frac{3}{4}+3Z,\hfill & z_{19}=4Z,\hfill & z_{20}=\frac{1}{2},\hfill \\ z_{21}=\frac{1}{6},\hfill & z_{22}=3+6Z,\hfill & z_{23}=39Z,\hfill & z_{24}=0,\hfill \\ z_{25}=3Z,\hfill & z_{26}=1,\hfill & z_{27}=0,\hfill & z_{28}=\frac{1}{2},\hfill \\ z_{29}=\frac{1}{2},\hfill & z_{30}=0,\hfill & z_{31}=\frac{3}{2}+6Z,\hfill & z_{32}=\frac{3}{2}+6Z.\hfill \end{array}$$
(5.9)
As already discussed above, the values in this list depend on our conventions for the basis systems in the strong, electromagnetic and weak parts of the next-to-leading order Lagrangian. The $`z_i`$ given in (5.9) have to be used together with the divergent parts of the coupling constants $`L_i`$ , $`K_i`$ and $`N_i`$ , respectively. The divergences involving the electroweak penguin coupling $`g_{\mathrm{ewk}}`$ are independent of this choice of basis and they agree with a recent calculation of Cirigliano and Golowich . Note that $`g_{\mathrm{ewk}}`$ appears only in the couplings of $`(8_L,8_R)`$ operators. This is because the lowest-order term proportional to $`g_{\mathrm{ewk}}`$ is already of $`𝒪(G_8e^2)`$. Therefore, the $`𝒪(G_8e^2p^2)`$ terms proportional to $`g_{\mathrm{ewk}}`$ arise from the product of the lowest-order $`(8_L,8_R)`$ weak operator times the $`𝒪(p^2)`$ invariant part of the strong Lagrangian.
The renormalized low-energy constants $`Z_i^r(\mu )`$ are in general scale dependent. The coefficients $`z_i`$ govern this scale dependence through the renormalization group equations
$$\mu \frac{dZ_i^r(\mu )}{d\mu }=\frac{z_i}{(4\pi )^2}.$$
(5.10)
By construction, the complete generating functional at next-to-leading order is then scale independent.
## 6 $`K\pi \pi `$
In the modern framework of chiral perturbation theory, electromagnetic corrections for $`K\pi \pi `$ decays to $`𝒪(G_8e^2p^2)`$ were discussed by de Rafael and have been treated in more detail by Cirigliano, Donoghue and Golowich . Together with corrections of $`𝒪(G_8(m_um_d)p^2)`$ , the complete isospin-breaking effects of next-to-leading order have obvious phenomenological implications, from the $`\mathrm{\Delta }I=1/2`$ rule to CP violation .
In this section, we present the tree-level contributions to the $`K\pi \pi `$ amplitudes from the Lagrangian (5.1). We compare those amplitudes and in particular their divergent parts with the results of Ref. . Using our own one-loop calculation of isospin-breaking corrections and the heat-kernel results (5.9), we find that the complete amplitudes of $`𝒪(G_8e^2p^2)`$ are indeed finite. We demonstrate the cancellation of divergences explicitly for the subset of amplitudes proportional to the electromagnetic penguin coupling $`g_{\mathrm{ewk}}`$ defined in (3.6).
From the Lagrangian (5.1) of $`𝒪(G_8e^2p^2)`$, we obtain the following amplitudes in units of $`C_{\mathrm{ewk}}:=iG_8e^2F`$:
$`A(K^0\pi ^+\pi ^{})`$ $`=`$ $`C_{\mathrm{ewk}}\sqrt{2}[(M_K^2M_\pi ^2)(2Z_1+4Z_24/3Z_3+4Z_4Z_51/3Z_6`$
$`2/3Z_7)+M_\pi ^2(6Z_1+6Z_2Z_6)],`$
$`A(K^0\pi ^0\pi ^0)`$ $`=`$ $`C_{\mathrm{ewk}}\sqrt{2}(M_K^2M_\pi ^2)(Z_5+2/3Z_62/3Z_7+Z_8+Z_9+2/3Z_{10}`$
$`2/3Z_{11}2/3Z_{12}),`$
$`A(K^+\pi ^+\pi ^0)`$ $`=`$ $`C_{\mathrm{ewk}}[(M_K^2M_\pi ^2)(2Z_1+4Z_24/3Z_3Z_6Z_8Z_92/3Z_{10}`$ (6.1)
$`4/3Z_{11}4/3Z_{12})+M_\pi ^2(6Z_1+6Z_2Z_6)].`$
These amplitudes agree with Ref. for $`Z_3=3Z_2,Z_{10}=Z_{11}=0`$. In addition, the coefficients $`s_8,s_9`$ in Eq. (35) of should be multiplied by 2/3.
In the $`SU(3)`$ limit for the mass matrix (3.2), the amplitudes (6) satisfy the relations
$`A(K^0\pi ^+\pi ^{})_{SU(3)}`$ $`=`$ $`\sqrt{2}A(K^+\pi ^+\pi ^0)_{SU(3)},`$
$`A(K^0\pi ^0\pi ^0)_{SU(3)}`$ $`=`$ $`0,`$ (6.2)
in accordance with a general theorem on $`K\pi \pi `$ transitions in the presence of electromagnetism .
The divergent parts of the $`Z_i`$ in (5.9) give rise to the following divergent tree-level amplitudes, with $`\mathrm{\Lambda }(\mu )`$ and $`Z`$ defined in (5.8) and (3.5), respectively:
$`A(K^0\pi ^+\pi ^{})_{\mathrm{div}}`$ $`=`$ $`C_{\mathrm{ewk}}\sqrt{2}\mathrm{\Lambda }(\mu )[M_K^2(327Z+13/2g_{\mathrm{ewk}})`$
$`+M_\pi ^2(3+36Z+7g_{\mathrm{ewk}})],`$
$`A(K^0\pi ^0\pi ^0)_{\mathrm{div}}`$ $`=`$ $`C_{\mathrm{ewk}}\sqrt{2}\mathrm{\Lambda }(\mu )(M_K^2M_\pi ^2)(2Z+3g_{\mathrm{ewk}}),`$ (6.3)
$`A(K^+\pi ^+\pi ^0)_{\mathrm{div}}`$ $`=`$ $`C_{\mathrm{ewk}}\mathrm{\Lambda }(\mu )\left[M_K^2(3Z+7/2g_{\mathrm{ewk}})+M_\pi ^2(6+6Z+10g_{\mathrm{ewk}})\right].`$
The (ultraviolet) divergences in (6) arise from three different sources:
* Photon loops proportional to $`G_8e^2`$;
* Loops involving the electromagnetic coupling (3.5) proportional to $`G_8e^2Z`$;
* Loops involving the coupling (3.6) proportional to $`G_8e^2g_{\mathrm{ewk}}`$.
Strong and electromagnetic wave function renormalization is included in all three categories.
We have performed a complete calculation of $`K\pi \pi `$ amplitudes to $`𝒪(G_8e^2p^2)`$ and $`𝒪(G_8(m_um_d)p^2)`$ . For $`m_u=m_d`$, we find that the explicit loop divergences are exactly cancelled by the divergent tree-level amplitudes (6). We exhibit those cancellations in detail for the divergences proportional to $`g_{\mathrm{ewk}}`$. Divergences arise both in loops with an electromagnetic penguin vertex shown in Fig. 1 and from (strong) wave function renormalization of tree diagrams from the Lagrangian (3.6).
In the exponential parametrization, the divergences due to the diagrams of Fig. 1 take the form
$`A(K^0\pi ^+\pi ^{})_{\mathrm{loops}}`$ $`=`$ $`{\displaystyle \frac{\sqrt{2}}{2}}C_{\mathrm{ewk}}g_{\mathrm{ewk}}\mathrm{\Lambda }(\mu )(7M_K^2+8M_\pi ^2),`$
$`A(K^0\pi ^0\pi ^0)_{\mathrm{loops}}`$ $`=`$ $`3\sqrt{2}C_{\mathrm{ewk}}g_{\mathrm{ewk}}\mathrm{\Lambda }(\mu )(M_K^2M_\pi ^2),`$
$`A(K^+\pi ^+\pi ^0)_{\mathrm{loops}}`$ $`=`$ $`C_{\mathrm{ewk}}g_{\mathrm{ewk}}\mathrm{\Lambda }(\mu )(M_K^2/2+7M_\pi ^2).`$ (6.4)
Wave function renormalization (again in exponential parametrization) leads to
$`A(K^0\pi ^+\pi ^{})_{\mathrm{wfr}}`$ $`=`$ $`3\sqrt{2}C_{\mathrm{ewk}}g_{\mathrm{ewk}}\mathrm{\Lambda }(\mu )(M_K^2+M_\pi ^2),`$
$`A(K^0\pi ^0\pi ^0)_{\mathrm{wfr}}`$ $`=`$ $`0,`$
$`A(K^+\pi ^+\pi ^0)_{\mathrm{wfr}}`$ $`=`$ $`3C_{\mathrm{ewk}}g_{\mathrm{ewk}}\mathrm{\Lambda }(\mu )(M_K^2+M_\pi ^2).`$ (6.5)
The sum of (6) and (6) is parametrization independent and it is exactly cancelled by the terms in (6) proportional to $`g_{\mathrm{ewk}}`$.
We have exhibited (part of) the loop divergences explicitly also because we do not completely agree with the results of Ref. . Although the divergences due to photon loops are identical, we obtain different results for some of the other divergences<sup>1</sup><sup>1</sup>1V. Cirigliano has informed us that they now agree with the divergences (6); see forthcoming erratum for Ref. .. Only for the channel $`K^0\pi ^+\pi ^{}`$, there is complete agreement for all three types of divergences.
The complete amplitudes of $`𝒪(G_8e^2p^2)`$ and $`𝒪(G_8(m_um_d)p^2)`$ together with a phenomenological analysis will be presented elsewhere .
## 7 Conclusions
We have supplied the missing ingredients for a complete analysis at next-to-leading order of the combined strong, nonleptonic weak and electromagnetic interactions of mesons. The main results are:
1. The complete and minimal Lagrangian (5.1) of $`𝒪(G_8e^2p^2)`$ contains 32 operators $`Q_i`$ and associated dimensionless coupling constants $`Z_i`$. Of these 32 operators, only 14 are of immediate phenomenological relevance. We have ordered the terms in a way most suitable for applications: the first 12 operators contribute to $`K2\pi `$ decays whereas the remaining two enter in $`K3\pi `$ amplitudes.
2. The one-loop divergence functional (4.13) determines the renormalization of the effective theory. Together with the previously known divergences, the new terms (5.9) in the coupling constants $`Z_i`$ ensure that the complete amplitudes for strong, nonleptonic weak and electromagnetic interactions of mesons at next-to-leading order are finite.
As a first application, we have presented the tree-level amplitudes of $`𝒪(G_8e^2p^2)`$ for $`K\pi \pi `$ decays. The associated divergent parts cancel with the explicit one-loop divergences to yield finite and scale independent decay amplitudes.
## Acknowledgements
We thank J. Gasser and E. de Rafael for having started this project with two of us (G.E., A.P.) some years ago. We are also grateful to J. Bijnens, V. Cirigliano and J. Gasser for helpful correspondence. H.N. acknowledges financial support from the University of Valencia through a “Visiting Professorship” and the IFIC Department of Theoretical Physics where part of this work was done for hospitality.
## Appendix
The quantities occurring in (4.14) can be decomposed with respect to (explicit<sup>2</sup><sup>2</sup>2Note that $`e`$ also appears in the vielbein $`u_\mu `$ (2.2) and in the connection $`\mathrm{\Gamma }_\mu `$ (4.6) via (2).) powers of $`e`$ and $`G_8`$ in the following way:
$`\sigma _{ij}`$ $`=`$ $`\sigma _{ij}|_{e^0G_8^0}+\sigma _{ij}|_{e^2G_8^0}+\sigma _{ij}|_{e^0G_8}+\sigma _{ij}|_{e^2G_8},`$
$`\gamma _\mu `$ $`=`$ $`\gamma _\mu |_{e^0G_8^0}+\gamma _\mu |_{e^0G_8},`$
$`a_i^\mu `$ $`=`$ $`a_i^\mu |_{eG_8^0}+a_i^\mu |_{eG_8},`$
$`b_i`$ $`=`$ $`b_i|_{eG_8^0}+b_i|_{eG_8},`$
$`\kappa `$ $`=`$ $`\kappa |_{e^2G_8^0}+\kappa |_{e^2G_8}.`$ (A.1)
The explicit expressions for the various terms are given by
$`\sigma _{ij}|_{e^0G_8^0}`$ $`=`$ $`{\displaystyle \frac{1}{8}}(u_\mu u^\mu +\chi _+)\{\lambda _i,\lambda _j\}{\displaystyle \frac{1}{4}}u_\mu \lambda _iu^\mu \lambda _j,`$ (A.2)
$`\sigma _{ij}|_{e^2G_8^0}`$ $`=`$ $`e^2F^2Z{\displaystyle \frac{1}{2}}\{𝒬_R,𝒬_L\}\{\lambda _i,\lambda _j\}\lambda _i𝒬_R\lambda _j𝒬_L\lambda _j𝒬_R\lambda _i𝒬_L,`$ (A.3)
$`\sigma _{ij}|_{e^0G_8}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\{\mathrm{\Xi },\lambda _i\}u_\mu \lambda _ju^\mu +{\displaystyle \frac{1}{4}}\{\mathrm{\Xi },\lambda _j\}u_\mu \lambda _iu^\mu `$ (A.4)
$`{\displaystyle \frac{1}{8}}(u_\mu u^\mu +\chi _+)(\{\lambda _i,\{\mathrm{\Xi },\lambda _j\}\}+\{\lambda _j,\{\mathrm{\Xi },\lambda _i\}\})`$
$`+{\displaystyle \frac{1}{6}}\mathrm{\Xi }\lambda _i\chi _+\lambda _j+{\displaystyle \frac{1}{6}}\mathrm{\Xi }\lambda _j\chi _+\lambda _i`$
$`+{\displaystyle \frac{1}{4}}\mathrm{\Xi }\{u_\mu ,\{u^\mu ,\{\lambda _i,\lambda _j\}\}\}`$
$`{\displaystyle \frac{1}{4}}\mathrm{\Xi }(\{u_\mu ,\lambda _i\}\{u^\mu ,\lambda _j\}+\{u_\mu ,\lambda _j\}\{u^\mu ,\lambda _i\})`$
$`+{\displaystyle \frac{i}{4}}[u_\mu ,^\mu \mathrm{\Xi }]\{\lambda _i,\lambda _j\}+{\displaystyle \frac{i}{4}}[^\mu u_\mu ,\mathrm{\Xi }]\{\lambda _i,\lambda _j\}`$
$`{\displaystyle \frac{1}{2}}\{\lambda _i,\lambda _j\}_\mu ^\mu \mathrm{\Xi },`$
$`\sigma _{ij}|_{e^2G_8}`$ $`=`$ $`e^2F^2Z\mathrm{\Xi }(\lambda _i𝒬_R\lambda _j𝒬_L+\lambda _j𝒬_R\lambda _i𝒬_L+𝒬_R\lambda _i𝒬_L\lambda _j+𝒬_R\lambda _j𝒬_L\lambda _i`$ (A.5)
$`+\lambda _i𝒬_L\lambda _j𝒬_R+\lambda _j𝒬_L\lambda _i𝒬_R+𝒬_L\lambda _i𝒬_R\lambda _j+𝒬_L\lambda _j𝒬_R\lambda _i`$
$`\lambda _i\{𝒬_L,𝒬_R\}\lambda _j\lambda _j\{𝒬_L,𝒬_R\}\lambda _i{\displaystyle \frac{1}{2}}\{\{𝒬_L,𝒬_R\},\{\lambda _i,\lambda _j\}\})`$
$`+e^2F^2\mathrm{{\rm Y}}({\displaystyle \frac{1}{2}}\{\{\lambda _i,\lambda _j\},𝒬_R\}\lambda _i𝒬_R\lambda _j\lambda _j𝒬_R\lambda _i),`$
$`\gamma _{ij}^\mu |_{e^0G_8^0}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{\Gamma }^\mu [\lambda _i,\lambda _j],`$ (A.6)
$`\gamma _{ij}^\mu |_{e^0G_8}`$ $`=`$ $`{\displaystyle \frac{i}{4}}[\lambda _i,\mathrm{\Xi }]\{u^\mu ,\lambda _j\}{\displaystyle \frac{i}{4}}[\lambda _j,\mathrm{\Xi }]\{u^\mu ,\lambda _i\},`$ (A.7)
$`a_i^\mu |_{eG_8^0}`$ $`=`$ $`{\displaystyle \frac{ieF}{4}}u^\mu [𝒬_R+𝒬_L,\lambda _i],`$ (A.8)
$`a_i^\mu |_{eG_8}`$ $`=`$ $`{\displaystyle \frac{ieF}{4}}\mathrm{\Xi }(u^\mu \lambda _i𝒬_R𝒬_R\lambda _iu^\mu )+{\displaystyle \frac{3ieF}{4}}\mathrm{\Xi }(u^\mu \lambda _i𝒬_L𝒬_L\lambda _iu^\mu )`$ (A.9)
$`{\displaystyle \frac{ieF}{2}}\mathrm{\Xi }(u^\mu 𝒬_R\lambda _i\lambda _i𝒬_Ru^\mu )+{\displaystyle \frac{ieF}{4}}\mathrm{\Xi }(\lambda _iu^\mu 𝒬_R𝒬_Ru^\mu \lambda _i)`$
$`+{\displaystyle \frac{ieF}{4}}\mathrm{\Xi }(\lambda _iu^\mu 𝒬_L𝒬_Lu^\mu \lambda _i){\displaystyle \frac{eF}{2}}(𝒬_R𝒬_L)\{^\mu \mathrm{\Xi },\lambda _i\},`$
$`b_i|_{eG_8^0}`$ $`=`$ $`{\displaystyle \frac{eF}{2}}(𝒬_R𝒬_L)\lambda _i,`$ (A.10)
$`b_i|_{eG_8}`$ $`=`$ $`{\displaystyle \frac{eF}{2}}\mathrm{\Xi }\{𝒬_R𝒬_L,\lambda _i\},`$ (A.11)
$`\kappa |_{e^2G_8^0}`$ $`=`$ $`{\displaystyle \frac{e^2F^2}{2}}(𝒬_R𝒬_L)^2,`$ (A.12)
$`\kappa |_{e^2G_8}`$ $`=`$ $`2e^2F^2\mathrm{\Xi }(𝒬_R𝒬_L)^2.`$ (A.13)
The expressions (A.12) and (A.13) are included for completeness only; $`\kappa `$ does not contribute to the order we are concerned with.
|
warning/0006/hep-ph0006008.html
|
ar5iv
|
text
|
# 1 Effect of mass correction and higher order in the mean thrust value. The solid and the dashed lines are respectively the first and second order calculations of without mass corrections. The dashed-dotted and dotted lines are complete and approximate (𝒪(𝑚)) first order calculations with quark mass effects.
One of the cleanest signatures of perturbative QCD comes from jet cross sections in $`e^+e^{}`$ annihilation. In such processes, it is possible to define infra-red safe event shape variables which can be calculated order by order in perturbative QCD and compared subsequently with experiment. However in order to carry out these comparisons, a method has to be evolved to parametrise non-perturbative effects which though expected to be small at present $`Q^2`$ values at LEP, actually turn out to be substantial ($``$ 25%) even at $`Qm_\mathrm{Z}`$. One of the reasons for this is that these non-perturbative effects are actually suppressed by a single power of $`Q`$ rather than $`Q^2`$. In addition, it is also possible that these power corrections could be comparable to $`𝒪(\alpha _\mathrm{s}^2)`$ at present LEP energies.
In order to address these issues the Milan group of Dokshitzer et al. drawing on the earlier work of Webber , Korchemsky and Sterman and others, presented a systematic approach for handling power corrections using perturbation theory. Very briefly, they studied the consequences of assuming that $`\alpha _s`$ has a low energy effective form which does not grow at low scales but has an infra-red regular form. The moments of $`\alpha _\mathrm{s}`$ are integrated only over the infra-red region. Various non-perturbative parameters are then parametrised and the form and magnitude of power corrections are determined.
However before one uses the approach of the Milan group in order to get a handle on power corrections and subsequently determine $`\alpha _\mathrm{s}`$ by a fit to the data, it is important to isolate power corrections coming from a purely perturbative region. The Milan approach neglects the masses of all the quarks but instead uses a gluon mass as a ‘trigger’ to differentiate the perturbative from the non-perturbative region. We find however that the masses of the quarks, particularly the c and the b quarks, even at present LEP energies, can contribute significantly (of the order of about 25%). In fact, if we go beyond the top quark threshold (which is expected, perhaps in the future NLC) the perturbative contribution to power contributions due the top quark mass is even larger. We will have more to say on this later in the paper.
In this paper, we consider the example of one such event shape variable \- the thrust - and show the significance of the effect of quark masses which need to be folded in before estimating the non-perturbative contribution to power corrections. We present explicit expressions to $`𝒪(\alpha _\mathrm{s})`$ of quark mass corrections expanded to $`𝒪(m)`$. We also show the effect of keeping the full mass contribution to $`𝒪(\alpha _\mathrm{s})`$ which unfortunately does not have a simple analytic form like the former and needs to be calculated numerically. Using these expressions we then fold in the power corrections of the Milan type and use this full expression to estimate both $`\alpha _0`$ and $`\alpha _\mathrm{s}`$ and compare it with estimates that exist in the literature without taking quark masses into account.
The first paper which calculated the effect of quark masses to $`𝒪(\alpha _\mathrm{s})`$ was published about 16 years ago by one of the authors . For completeness, in what follows, we quote those results from that paper which we need for our analysis here. The thrust, as defined traditionally, is given by
$$T=2\frac{\mathrm{max}_{iϵh}(p_i\widehat{n})}{_i|p_i|},$$
(1)
where the denominator runs over all observed particles and the numerator runs over all particles in a hemisphere. $`\widehat{n}`$ is a unit vector chosen in a direction that maximises the numerator and defines the jet axis.
While this definition is appropriate for all massless particles, to include mass effects in the definition of the thrust, we modify the above definition slightly and write
$$T=2\frac{\mathrm{max}_{iϵh}(p_i\widehat{n})}{W},$$
(2)
where $`W^2=s`$. Of course the denominator equals $`_i|p_i|`$ when all the particles are massless. This normalisation with the total energy is also what is used by the Milan group in their analysis though in their case the massive gluon eventually decays into massless quarks and gluons.
For a three particle final state, the thrust, as we define it, is given by
$$T=\mathrm{max}[(x_1^2\xi )^{1/2},(x_2^2\xi )^{1/2},x_3],$$
(3)
where $`x_i=2E_i/W`$, $`E_i`$ being the energy of the $`i`$th particle in the final state in the c.m. frame and $`\xi =4m^2/W^2`$, $`m`$ being the mass of the quarks. Note that in the two-jet limit $`T=T_0\sqrt{1\xi }`$.
The average value of the thrust is defined by
$$<T>=\frac{\left[T\frac{d\sigma }{dT}𝑑T\right]}{\left[\frac{d\sigma }{dT}𝑑T\right]}.$$
(4)
The numerator of the above is given up to $`𝒪(\alpha _\mathrm{s})`$ and to $`𝒪(\xi )`$ by ($`\sigma _0=(4\pi \alpha _2/s)e_i^2`$ is the total cross section for $`e^+e^{}q_i\overline{q}_i`$)
$`{\displaystyle \frac{1}{\sigma _0}}{\displaystyle T\frac{d\sigma }{dT}}`$ $`=`$ $`1{\displaystyle \frac{\xi }{2}}+{\displaystyle \frac{4\alpha _\mathrm{s}}{3\pi }}\{{\displaystyle \frac{137}{16}}\xi +{\displaystyle \frac{5}{4}}\xi \mathrm{ln}2{\displaystyle \frac{1}{2}}\xi ^{1/2}+{\displaystyle \frac{7}{9}}+{\displaystyle \frac{1}{4}}\xi \mathrm{ln}^2\xi \xi \mathrm{ln}2\mathrm{ln}\xi +{\displaystyle \frac{\pi ^2}{6}}`$ (5)
$`{\displaystyle \frac{\xi \pi ^2}{6}}{\displaystyle \frac{1}{8}}\xi \mathrm{ln}\xi {\displaystyle \frac{1}{2}}\xi \mathrm{ln}3\mathrm{ln}2{\displaystyle \frac{9}{2}}\xi \mathrm{ln}3\mathrm{ln}^23+{\displaystyle \frac{3}{8}}\mathrm{ln}3`$
$`{\displaystyle \frac{1}{3}}\xi \left[Li_2(1\xi ^{1/2}+\xi /2)Li_2({\displaystyle \frac{1}{3}}+{\displaystyle \frac{1}{2}}\xi )\right]2Li_2({\displaystyle \frac{2}{3}}{\displaystyle \frac{1}{2}}\xi )`$
$`+\xi Li_2({\displaystyle \frac{2}{3}}{\displaystyle \frac{1}{2}}\xi ){\displaystyle \frac{1}{2}}\xi Li_2({\displaystyle \frac{1}{3}}{\displaystyle \frac{1}{4}}\xi )+\xi \mathrm{ln}^22\},`$
where $`Li_2(x)`$ is the dilogarithm function. In the $`\xi 0`$ limit this gives, for the average thrust,
$$<T>=1+\frac{4\alpha _\mathrm{s}}{3\pi }\left[\frac{1}{36}+\frac{\pi ^2}{6}\mathrm{ln}^23+\frac{3}{8}\mathrm{ln}32Li_2(\frac{2}{3})\right],$$
(6)
which works out to, for the perturbative thrust in the massless limit,
$$<1T>=1.05\frac{\alpha _\mathrm{s}}{\pi },$$
(7)
as quoted in numerous places in the literature.
Several points here are worthy of note. The leading term in the $`O(m)`$ expansion above is $`\xi ^{1/2}`$. Thus the leading mass correction goes as $`1/Q`$. To the best of our knowledge, this fact was noticed for the first time in and subsequently in and who have traced it to appear from the soft phase space boundary. We would like to stress that this $`1/Q`$ behaviour is a pure perturbative higher twist effect to the thrust and not related to any non-perturbative contribution. Thus, it seems clear, that the coefficient of $`1/Q`$ in the full expression for the thrust would include contributions both from the perturbative as well as the non-perturbative sectors. This aspect will become more quantitative, when we do our fits later.
The second point to note is a calculational one. Since, in the two jet limit, the thrust is equal to $`T_0=\sqrt{1\xi }`$, in order to make the virtual contributions vanish we need to calculate, not as in the usual case $`<1T>`$, but $`<T_0T>`$. It is then a trivial matter to add a term $`<1T_0>`$ to obtain $`<1T>`$ to compare with experiment.
In order to compare with experiment, however, and to redo the fits for $`\alpha _\mathrm{s}`$ and $`\alpha _0`$ we have used not only the $`𝒪(m)`$ contribution above but also the full massive contribution to $`𝒪(\alpha _\mathrm{s})`$ albeit evaluated numerically. In addition we have also compared the $`𝒪(\alpha _\mathrm{s}^2)`$ massless corrections to the $`𝒪(\alpha _\mathrm{s})`$ massive correction to try and estimate how much of the $`1/Q`$ corrections can be mimicked by higher orders in the coupling constant.
Figure 1 shows $`<1T>`$ as a function of centre of mass energy computed with $`\alpha _\mathrm{s}(m_\mathrm{Z})`$ = 0.12. As one sees from the curve the contribution of the second order terms are large over the entire energy region ($``$ 55% at $`Q`$ = 12 GeV going down to 33% at 200 GeV). On the other hand the effect of quark masses, evaluated only to first order in $`\alpha _\mathrm{s}`$, is even larger at low centre of mass energy ($``$ 76% at $`Q`$ = 12 GeV). This is clearly a $`1/Q`$ power law effect and hence dies off faster, becoming 2.5% at $`Qm_\mathrm{Z}`$ and negligible at 200 GeV. It is clear from the figure that one needs to take the full massive correction rather than the $`𝒪(m)`$ contribution, because it accounts only 60% (30%) of the mass correction at 20 GeV (12 GeV).
In order to compare the theoretical predictions with the measurements done at different centre of mass energies at PETRA, PEP, TRISTAN, SLC and LEP, we add the non-perturbative contribution a la the Milan group to the perturbative contribution. In this paper, we use only the $`𝒪(\alpha _\mathrm{s})`$ calculation of $`<1T>`$ and a more detailed comparison with a $`𝒪(\alpha _\mathrm{s}^2)`$ calculation is under preparation . We will have more to say on this later. The non-perturbative contribution, as is well known, is given by an additive contribution $`<1T>_{\mathrm{pow}}`$:
$$<1T>_{\mathrm{pow}}=2\frac{4C_F}{\pi ^2}\frac{\mu _I}{Q}\left[\alpha _0(\mu _I)\alpha _\mathrm{s}(Q)\beta _0\frac{\alpha _\mathrm{s}^2(Q)}{2\pi }\left(\mathrm{ln}\frac{Q}{\mu _I}+\frac{K}{\beta _0}+1\right)\right]$$
(8)
where $`\mu _I`$ is an infra-red matching scale (taken as 2 GeV), $`K`$ = (67/18 $``$ $`\pi ^2`$/6)$`C_A`$ $``$ $`5\mathrm{N}_\mathrm{f}/9`$ and $``$ is the Milan factor (determined to be 1.49) .
Figure 2 shows the experimental values of $`<1T>`$ together with the two fits which use respectively the massless and the massive forms (both to $`𝒪(\alpha _\mathrm{s})`$ for the perturbative contribution). These fits have been carried out with two free parameters $`\alpha _\mathrm{s}(m_\mathrm{Z})`$ and $`\alpha _0`$. Both massless and massive formulation of the perturbative component give reasonable fits to the data with $`\chi ^2`$ of 90.2 and 69.2 respectively for 48 degrees of freedom corresponding to confidence levels of $`0.22\times 10^3`$ and $`0.24\times 10^1`$. However, they do differ in the final values of $`\alpha _\mathrm{s}(m_\mathrm{Z})`$ and $`\alpha _0`$ as can be seen in Table 1. In these fits the scale parameter is chosen to be 1.0.
The errors quoted in the Table 1 are experimental errors obtained from the minimisation procedure. We can also estimate the theoretical uncertainties on these quantities by varying the scale parameter. If we vary the scale parameter between 0.5 and 2.0, we obtain uncertainties in $`\alpha _\mathrm{s}`$ and $`\alpha _0`$ to be $`\pm 0.010`$ and $`\pm 0.12`$ respectively. The value of $`\alpha _\mathrm{s}(m_\mathrm{Z})`$, obtained from the fits, when quark mass effects are included or ignored, differ by 0.008 which is much larger than the experimental uncertainty of about .001 on the $`\alpha _\mathrm{s}`$ value and comparable in fact to the theoretical uncertainty.
It is thus clear from the preceding analysis that an estimate of the power corrections due to the non-zero masses of the quarks is crucial in getting better and more realistic estimates on the strong coupling constant and indeed, in general, on power corrections. The next obvious step would be to calculate mass corrections to $`𝒪(\alpha _\mathrm{s}^2)`$. Some results in this direction have been obtained by Nason and Oleari which could be used to carry out a similar analysis to the one presented above. We are, at present, in the process of extending our analysis to second order in the strong coupling using the results of .
In various projected Linear Collider scenarios (like, for example, the NLC) energies upwards of 500 GeV are expected. In such a region, the effect of the top quark would be dramatic and significant. The combination of the large mass of the top quark and a charge squared of 4/9 implies that the usual massless expressions for the thrust would not work. We have estimated that the difference between choosing a massless formula for describing the thrust beyond the top quark threshold and using the (more appropriate) massive formula changes the value of the thrust by about a factor of 5 near the threshold. Most of this contribution comes, in fact, from the top quark mass. In the table below we give an estimate of the change that would occur between choosing all quarks massless and massive above the top quark threshold. It is obvious that the effect is spectacularly large, particularly near the threshold. Mass effects in the resummation of event shape variables are also expected to be significant and this is presently being studied.
Thus, it is imperative that in order that reliable estimates be made of the thrust at these energies, we have available, calculations to higher orders in $`\alpha _\mathrm{s}`$ of $`e^+e^{}`$ scattering with massive quarks in the final state. This would also give us a handle on the relative magnitudes of power corrections to the thrust to a particular order in $`\alpha _\mathrm{s}`$ and the magnitude of the next order term in $`\alpha _\mathrm{s}`$ . For example NNLO effects might be capable of mimicking the $`1/Q`$ behavior. Mass effects in the resummation of event shape variables are also expected to be significant and this is presently being studied.
Acknowledgements: One of us (R.B.) would like to thank George Sterman for many illuminating discussions. This project was started at the Sixth Workshop on High Energy Physics Phenomenology (WHEPP-6), held at Chennai, India in January 2000 and we would like to thank all the funding agencies which made this workshop possible.
|
warning/0006/hep-th0006144.html
|
ar5iv
|
text
|
# References
Problems of consistent equations of motion for interacting higher spin fields deserve studying due to many reasons. First of all, string theory includes an infinite tower of massive excitations with all possible spins and thus should allow some consistent effective description of arbitrary spin fields interaction. Second, composite resonance particles with higher spins do exist and one should be able to describe their interaction (for example, with external electromagnetic and gravitational fields) in terms of some effective local field theory. At last, investigation of higher spin fields is interesting on its own from the general point of view. It would be surprising if nature admits description of free fields with arbitrary spins but stops, say, at spin 1 in case of interacting massive fields. Even if it is really the case, one should try to understand why this is the way the nature works.
This note is devoted to investigation of the massive spin 2 field interacting with external gravity which represents one of the simplest higher spin models. It has been studied in numerous papers -<sup>2</sup><sup>2</sup>2For analysis of the corresponding massless models see e.g. but careful and general analysis of the consistency and causality of the theory in arbitrary curved spacetime was still absent.
There are at least two ways the interaction may spoil the consistency of a higher spin theory. Firstly, interaction may change the number of dynamical degrees of freedom. For example, a massive field with spin $`s`$ in $`D=4`$ Minkowski spacetime is described by a rank $`s`$ symmetric traceless transverse tensor $`\varphi _{(\mu _1\mathrm{}\mu _s)}`$ satisfying the mass shell condition:
$$(^2m^2)\varphi _{\mu _1\mathrm{}\mu _s}=0,^\mu \varphi _{\mu \mu _1\mathrm{}\mu _{s1}}=0,\varphi ^\mu {}_{\mu \mu _1\mathrm{}\mu _{s2}}{}^{}=0.$$
(1)
To reproduce all these equations from a single lagrangian one needs to introduce auxiliary fields $`\chi _{\mu _1\mathrm{}\mu _{s2}}`$, $`\chi _{\mu _1\mathrm{}\mu _{s3}}`$, …, $`\chi `$ . These symmetric traceless fields vanish on shell but their presence in the theory provides lagrangian description of the conditions (1). In higher dimensional spacetimes there appear fields of more complex tensor structure but general situation remains the same, i.e. lagrangian description always requires presence of unphysical auxiliary degrees of freedom.
Namely these auxiliary fields create problems when one tries to turn on interaction in the theory. Arbitrary interaction makes the auxiliary fields dynamical thus increasing the number of degrees of freedom. Usually these extra degrees of freedom are ghostlike and should be considered as pathological. Requirement of absence of these extra dynamical degrees of freedom imposes severe restrictions on the possible interaction.
To construct a consistent massive field theory one often starts with a corresponding consistent massless theory which should be invariant with respect to gauge symmetry and then breaks this symmetry by introducing mass terms into lagrangian. In this case invariance of the kinetic part of lagrangian guarantees the correct number of degrees of freedom in both massless and massive theories. For example, spin 2 field possesses a gauge invariant massless lagrangian only if external gravity satisfies vacuum Einstein equations and so it is usually believed that the massive spin 2 field can consistently propagate also only in Einstein spacetimes.
In this paper which is a sequel to we show that this belief is not true and that in massive case there exists a number of possibilities of providing the correct number of degrees of freedom. Of course, invariance of kinetic term does always provide the correct number of degrees of freedom but this is not the only possibility. In particular, we describe below an example when external spacetime does not fulfill Einstein equations but the theory is consistent with the flat spacetime limit.
The other problem that may arise in higher spin fields theories is connected with possible violation of causal properties. This problem was first noted in the theory of spin 3/2 field in external fields (see also the review and a recent discussion in )
In general, when one has a system of differential equations for a set of fields $`\varphi ^B`$ (to be specific, let us say about second order equations)
$$M_{AB}{}_{}{}^{\mu \nu }_{\mu }^{}_\nu \varphi ^B+\mathrm{}=0,\mu ,\nu =0,\mathrm{},D1$$
(2)
the following definitions are used. A characteristic matrix is the matrix function of $`D`$ arguments $`n_\mu `$ built out of the coefficients at the second derivatives in the equations: $`M_{AB}(n)=M_{AB}{}_{}{}^{\mu \nu }n_{\mu }^{}n_\nu .`$ A characteristic equation is $`detM_{AB}(n)=0.`$ A characteristic surface is the surface $`S(x)=const`$ where $`_\mu S(x)=n_\mu `$.
If for any $`n_i`$ ($`i=1,\mathrm{},D1`$) all solutions of the characteristic equation $`n_0(n_i)`$ are real then the system of differential equations is called hyperbolic and describes propagation of some wave processes. The hyperbolic system is called causal if there is no timelike vectors among solutions $`n_\mu `$ of the characteristic equations. Such a system describes propagation with a velocity not exceeding the speed of light. If there exist timelike solutions for $`n_\mu `$ then the corresponding characteristic surfaces are spacelike and violate causality.
Turning on interaction in theories of higher spin fields in general changes the characteristic matrix and there appears possibility of superluminal propagation. Such a situation also should be considered as pathological.
Both these problems arise in the theory of massive spin 2 field coupled to external gravitational field. To provide consistency of the interaction we should conserve the same number of physical degrees of freedom and constraints that the theory possesses in flat spacetime. To find the complete set of constraints we will use the general lagrangian scheme which is equivalent to the Dirac-Bergmann procedure in hamiltonian formalism but for our purposes is simpler. In the case of second class constraints (which is relevant for massive higher spin fields) it consists in the following steps. If in a theory of some set of fields $`\varphi ^A(x)`$, $`A=1,\mathrm{},N`$ the original lagrangian equations of motion define only $`r<N`$ of the second time derivatives (“accelerations”) $`\ddot{\varphi }^A`$ then one can build $`Nr`$ primary constraints, i.e. linear combinations of the equations of motion that does not contain accelerations. Requirement of conservation in time of the primary constraints either define some of the missing accelerations or lead to new (secondary) constraints. Then one demands conservation of the secondary constraints and so on, until all the accelerations are defined and the procedure closes up.
Before considering the theory in external gravitational field we analyze the structure of equations of motion in Minkowski spacetime. The purpose of this analysis is twofold. First, we illustrate the general scheme of calculating the constraints within covariant lagrangian framework. In addition, building a consistent and causal theory in curved spacetime we use these flat constraints as a reference point.
Free spin 2 field is known to be described by the Fierz-Pauli action (we consider arbitrary spacetime dimension):
$`S`$ $`=`$ $`{\displaystyle }d^Dx\{{\displaystyle \frac{1}{4}}_\mu H^\mu H{\displaystyle \frac{1}{4}}_\mu H_{\nu \rho }^\mu H^{\nu \rho }{\displaystyle \frac{1}{2}}^\mu H_{\mu \nu }^\nu H+{\displaystyle \frac{1}{2}}_\mu H_{\nu \rho }^\rho H^{\nu \mu }`$ (3)
$`{\displaystyle \frac{m^2}{4}}H_{\mu \nu }H^{\mu \nu }+{\displaystyle \frac{m^2}{4}}H^2\}.`$
Here the role of auxiliary field is played by the trace $`H=\eta ^{\mu \nu }H_{\mu \nu }`$. The equations of motion
$`E_{\mu \nu }`$ $`=`$ $`^2H_{\mu \nu }\eta _{\mu \nu }^2H+_\mu _\nu H+\eta _{\mu \nu }^\alpha ^\beta H_{\alpha \beta }_\sigma _\mu H^\sigma {}_{\nu }{}^{}_\sigma _\nu H^\sigma _\mu `$ (4)
$`m^2H_{\mu \nu }+m^2H\eta _{\mu \nu }=0`$
contain $`D`$ primary constraints (expressions without second time derivatives $`\ddot{H}_{\mu \nu }`$):
$`E_{00}`$ $`=`$ $`\mathrm{\Delta }H_{ii}_i_jH_{ij}m^2H_{ii}\phi _0^{(1)}0`$ (5)
$`E_{0i}`$ $`=`$ $`\mathrm{\Delta }H_{0i}+_i\dot{H}_{kk}_k\dot{H}_{ki}_i_kH_{0k}m^2H_{0i}\phi _i^{(1)}0.`$ (6)
The remaining equations of motion $`E_{ij}=0`$ allow to define the accelerations $`\ddot{H}_{ij}`$ in terms of $`\dot{H}_{\mu \nu }`$ and $`H_{\mu \nu }`$. The accelerations $`\ddot{H}_{00}`$, $`\ddot{H}_{0i}`$ cannot be expressed from the equations directly.
Conditions of conservation of the primary constraints in time $`\dot{E}_{0\mu }0`$ lead to $`D`$ secondary constraints. On-shell they are equivalent to
$$\phi _\nu ^{(2)}=^\mu E_{\mu \nu }=m^2_\nu Hm^2^\mu H_{\mu \nu }0$$
(7)
Conservation of $`\phi _i^{(2)}`$ defines $`D1`$ accelerations $`\ddot{H}_{0i}`$ and conservation of $`\phi _0^{(2)}`$ gives another one constraint. It is convenient to choose it in the covariant form by adding suitable terms proportional to the equations of motion:
$`\phi ^{(3)}=^\mu ^\nu E_{\mu \nu }+{\displaystyle \frac{m^2}{D2}}\eta ^{\mu \nu }E_{\mu \nu }=Hm^4{\displaystyle \frac{D1}{D2}}0`$ (8)
Conservation of $`\phi ^{(3)}`$ gives one more constraint on initial values
$$\phi ^{(4)}=\dot{H}_{00}+\dot{H}_{kk}=\dot{H}0$$
(9)
and from the conservation of this last constraint the acceleration $`\ddot{H}_{00}`$ is defined.
Altogether there are $`2D+2`$ constraints on the initial values of $`\dot{H}_{\mu \nu }`$ and $`H_{\mu \nu }`$. The lagrangian theory is equivalent to the system of the equations
$$(^2m^2)H_{\mu \nu }=0,^\mu H_{\mu \nu }=0,H^\mu {}_{\mu }{}^{}=0.$$
(10)
and describes traceless and transverse symmetric tensor field of the second rank.
Obviously, the equations of motion (10) are causal because the characteristic equation
$$detM(n)=(n^2)^{D(D+1)/2}$$
(11)
has 2 multiply degenerate roots
$$n_0^2+n_i^2=0,n_0=\pm \sqrt{n_i^2}.$$
(12)
which correspond to real null solutions for $`n_\mu `$. Note that analysis of causality is possible only after calculation of all the constraints. Original lagrangian equations of motion (4) have degenerate characteristic matrix $`detM(n)0`$ and do not allow to define propagation cones of the field $`H_{\mu \nu }`$.
In the massless limit $`m^2=0`$ the structure of the theory (3) changes. Instead of the secondary constraints (7) conservation of the primary constraints lead to identities $`^\mu E_{\mu \nu }0`$ which mean that the theory becomes gauge invariant with respect to the local transformations $`\delta H_{\mu \nu }=_\mu \xi _\nu +_\nu \xi _\mu `$. Such a theory represents the quadratic part of the Einstein-Hilbert action for gravitational field and the gauge invariance is a linear counterpart of the general coordinate invariance.
Now if we want to construct a theory of massive spin 2 field on a curved manifold first of all we should provide the same number of propagating degrees of freedom as in the flat case. It means that new equations of motion $`E_{\mu \nu }`$ should lead to exactly $`2D+2`$ constraints and in the flat spacetime limit these constraints should reduce to their flat counterparts.
Generalizing (3) to curved spacetime we should substitute all derivatives by the covariant ones and also we can add non-minimal terms containing curvature tensor with some dimensionless coefficients in front of them. As a result, the most general action for massive spin 2 field in curved spacetime quadratic in derivatives and consistent with the flat limit should have the form :
$`S={\displaystyle }d^Dx\sqrt{G}\{{\displaystyle \frac{1}{4}}_\mu H^\mu H{\displaystyle \frac{1}{4}}_\mu H_{\nu \rho }^\mu H^{\nu \rho }{\displaystyle \frac{1}{2}}^\mu H_{\mu \nu }^\nu H+{\displaystyle \frac{1}{2}}_\mu H_{\nu \rho }^\rho H^{\nu \mu }`$
$`+{\displaystyle \frac{a_1}{2}}RH_{\alpha \beta }H^{\alpha \beta }+{\displaystyle \frac{a_2}{2}}RH^2+{\displaystyle \frac{a_3}{2}}R^{\mu \alpha \nu \beta }H_{\mu \nu }H_{\alpha \beta }+{\displaystyle \frac{a_4}{2}}R^{\alpha \beta }H_{\alpha \sigma }H_\beta {}_{}{}^{\sigma }+{\displaystyle \frac{a_5}{2}}R^{\alpha \beta }H_{\alpha \beta }H`$
$`{\displaystyle \frac{m^2}{4}}H_{\mu \nu }H^{\mu \nu }+{\displaystyle \frac{m^2}{4}}H^2\}`$ (13)
where $`a_1,\mathrm{}a_5`$ are so far arbitrary dimensionless coefficients, $`R^\mu {}_{\nu \lambda \kappa }{}^{}=_\lambda \mathrm{\Gamma }_{\nu \kappa }^\mu \mathrm{}`$, $`R_{\mu \nu }=R^\lambda _{\mu \lambda \nu }`$.
Equations of motion
$`E_{\mu \nu }`$ $`=`$ $`^2H_{\mu \nu }G_{\mu \nu }^2H+_\mu _\nu H+G_{\mu \nu }^\alpha ^\beta H_{\alpha \beta }_\sigma _\mu H^\sigma {}_{\nu }{}^{}_\sigma _\nu H^\sigma _\mu `$ (14)
$`+2a_1RH_{\mu \nu }+2a_2G_{\mu \nu }RH+2a_3R_\mu {}_{\nu }{}^{\alpha }{}_{}{}^{\beta }H_{\alpha \beta }^{}+a_4R_\mu {}_{}{}^{\alpha }H_{\alpha \nu }^{}+a_4R_\nu {}_{}{}^{\alpha }H_{\alpha \mu }^{}`$
$`+a_5R_{\mu \nu }H+a_5G_{\mu \nu }R^{\alpha \beta }H_{\alpha \beta }m^2H_{\mu \nu }+m^2HG_{\mu \nu }0`$
contain second time derivatives of $`H_{\mu \nu }`$ in the following way:
$`E_{00}`$ $`=`$ $`(G^{mn}G_{00}G^{00}G^{mn}+G_{00}G^{0m}G^{0n})_0_0H_{mn}+O(_0),`$
$`E_{0i}`$ $`=`$ $`(G_{0i}G^{00}G^{mn}+G_{0i}G^{0m}G^{0n}G^{0m}\delta _i^n)_0_0H_{mn}+O(_0),`$
$`E_{ij}`$ $`=`$ $`(G^{00}\delta _i^m\delta _j^nG_{ij}G^{00}G^{mn}+G_{ij}G^{0m}G^{0n})_0_0H_{mn}+O(_0).`$ (15)
So we see that accelerations $`\ddot{H}_{00}`$ and $`\ddot{H}_{0i}`$ again (as in the flat case) do not enter the equations of motion while accelerations $`\ddot{H}_{ij}`$ can be expressed through $`\dot{H}_{\mu \nu }`$, $`H_{\mu \nu }`$ and their spatial derivatives.
There are $`D`$ linear combinations of the equations of motion which do not contain second time derivatives and so represent primary constraints of the theory:
$$\phi _\mu ^{(1)}=E^0{}_{\mu }{}^{}=G^{00}E_{0\mu }+G^{0j}E_{j\mu }$$
(16)
Now one should calculate time derivatives of these constraints and define secondary ones. In order to do this in a covariant form we can add to the time derivative of $`\phi _\mu ^{(1)}`$ any linear combination of equations of motion and primary constraints. So we choose the secondary constraints in the following way:
$`\phi _\mu ^{(2)}`$ $`=`$ $`^\alpha E_{\alpha \mu }=\dot{\phi }_\mu ^{(1)}+_iE^i{}_{\mu }{}^{}+\mathrm{\Gamma }_{\alpha 0}^\alpha \phi _\mu ^{(1)}+\mathrm{\Gamma }_{\alpha i}^\alpha E^i{}_{\mu }{}^{}\mathrm{\Gamma }_{\mu 0}^\sigma \phi _\sigma ^{(1)}\mathrm{\Gamma }_{\mu i}^\sigma E^i_\sigma `$ (17)
$`=`$ $`(2a_1Rm^2)^\mu H_{\mu \nu }+(2a_2R+m^2)_\nu H+2a_3R^{\mu \alpha }{}_{\nu }{}^{}{}_{}{}^{\beta }_{\mu }^{}H_{\alpha \beta }+a_4R^{\mu \alpha }_\mu H_{\alpha \nu }`$
$`+(a_42)R^\alpha {}_{\nu }{}^{}_{}^{\mu }H_{\alpha \mu }+a_5R^{\alpha \mu }_\nu H_{\alpha \mu }+(a_5+1)R^\alpha {}_{\nu }{}^{}_{\alpha }^{}H`$
$`+(2a_1+{\displaystyle \frac{a_4}{2}})H_{\alpha \nu }^\alpha R+(2a_2+{\displaystyle \frac{a_5}{2}})H_\nu R`$
$`+H_{\alpha \beta }[(2a_3+a_5+1)_\nu R^{\alpha \beta }+(a_42a_32)^\alpha R^\beta _\nu ]`$
At the next step conservation of these $`D`$ secondary constraints should lead to one new constraint and to expressions for $`D1`$ accelerations $`\ddot{H}_{0i}`$. This means that the constraints (17) should contain the first time derivatives $`\dot{H}_{0\mu }`$ through the matrix with the rank $`D1`$:
$`\phi _0^{(2)}`$ $`=`$ $`A\dot{H}_{00}+B^j\dot{H}_{0j}+\mathrm{}`$
$`\phi _i^{(2)}`$ $`=`$ $`C_i\dot{H}_{00}+D_i{}_{}{}^{j}\dot{H}_{0j}^{}+\mathrm{}`$ (18)
$$\text{rank}\widehat{\mathrm{\Phi }}_\mu {}_{}{}^{\nu }\text{rank}\left|\right|\begin{array}{cc}A& B^j\\ C_i& D_i^j\end{array}\left|\right|=D1$$
(19)
In the flat spacetime we had the matrix
$$\widehat{\mathrm{\Phi }}_\mu {}_{}{}^{\nu }=\left|\right|\begin{array}{cc}0& 0\\ 0& m^2\delta _i^j\end{array}\left|\right|$$
(20)
In the curved case the explicit form of this matrix elements in the constraints (17) is:
$`A`$ $`=`$ $`RG^{00}(2a_1+2a_2)+R^{00}(a_4+a_5)+R^0{}_{0}{}^{}G_{}^{00}(a_4+a_51)`$
$`B^j`$ $`=`$ $`m^2G^{0j}+RG^{0j}(2a_1+4a_2)+2a_3R^{0j}{}_{0}{}^{}{}_{}{}^{0}+R^j{}_{0}{}^{}G_{}^{00}(a_42)`$
$`+R^{0j}(a_4+2a_5)+R^0{}_{0}{}^{}G_{}^{0j}(a_4+2a_5)`$
$`C_i`$ $`=`$ $`R^0{}_{i}{}^{}G_{}^{00}(a_4+a_51)`$
$`D_i^j`$ $`=`$ $`m^2G^{00}\delta _i^j+2a_1RG^{00}\delta _i^j+2a_3R^{0j}{}_{i}{}^{}{}_{}{}^{0}+a_4R^{00}\delta _i^j`$ (21)
$`+(a_42)R^j{}_{i}{}^{}G_{}^{00}+(a_4+2a_5)R_i^0G^{0j}`$
At this stage the restrictions that consistency imposes on the type of interaction reduce to the requirements that the above matrix elements give
$$det\widehat{\mathrm{\Phi }}=0,detD_i{}_{}{}^{j}0$$
(22)
When the gravitational background is arbitrary it is not clear how to fulfill this condition by choosing some specific values of non-minimal couplings $`a_1`$, …$`a_5`$. For example, requirement of vanishing of the elements $`A`$ and $`C_i`$ (21) would lead to contradictory equations $`a_4+a_5=0`$, $`a_4+a_51=0`$.
But the consistency conditions (22) can be fulfilled in a number of specific gravitational background. Namely, any spacetime which in some coordinates has
$$R^0{}_{i}{}^{}=0$$
(23)
provides such an example. In such a spacetime $`R^{00}=R^0{}_{0}{}^{}G_{}^{00}`$ and choosing coefficients $`a_1+a_2=0`$, $`2a_4+2a_5=1`$ we have the first column of the matrix $`\widehat{\mathrm{\Phi }}`$ vanishing and so the conditions (22) fulfilled.
As a first example where (23) holds let us consider an arbitrary static spacetime, i.e. a spacetime having a timelike Killing vector and invariant with respect to the time reversal $`x^0x^0`$. In such a spacetime one can always find coordinates where
$$_0G_{\mu \nu }=0,G_{0i}=0.$$
(24)
The matrix elements (21) in this case become
$`A=RG^{00}(2a_1+2a_2)+R^{00}(2a_4+2a_51),B^j=0,C_i=0,`$
$`D_i{}_{}{}^{j}=(m^2G^{00}+2a_1RG^{00}+a_4R^{00})\delta _i^j+(a_42)R^j{}_{i}{}^{}G_{}^{00}+2a_3R^{0j}{}_{i}{}^{}^0`$ (25)
and (22) lead to the following conditions:
$$2a_1+2a_2=0,2a_4+2a_51=0,detD_i{}_{}{}^{j}0$$
(26)
The last inequality may be violated in strong gravitational field and as we comment below this fact may lead to causal problems.
Suppose that all the conditions (26) are fulfilled. For simplicity we also choose $`a_3=0`$. Then we have the equations of the form (14) with the coefficients
$$a_1=\frac{\xi _1}{2},a_2=\frac{\xi _1}{2},a_3=0,a_4=\frac{1}{2}\xi _2,a_5=\xi _2$$
(27)
where $`\xi _1`$, $`\xi _2`$ are two arbitrary coupling parameters.
One of the secondary constraints
$`\phi _0^{(2)}`$ $`=`$ $`^\alpha E_{\alpha 0}=_0H_{ij}[G^{ij}(m^2\xi _1R)+(1+\xi _2)G^{ij}R^0{}_{0}{}^{}+\xi _2R^{ij}]`$ (28)
$`+_iH_{0j}[G^{ij}(\xi _1Rm^2)({\displaystyle \frac{3}{2}}+\xi _2)G^{ij}R^0{}_{0}{}^{}+({\displaystyle \frac{1}{2}}\xi _2)R^{ij}]`$
$`+({\displaystyle \frac{1}{4}}+\xi _1{\displaystyle \frac{\xi _2}{2}})H_{0i}^iR({\displaystyle \frac{3}{2}}+\xi _2)H_{0i}^iR^0_0`$
does not contain velocities $`\dot{H}_{00}`$, $`\dot{H}_{0i}`$ and so its conservation leads to a new constraint $`\phi ^{(3)}_0^\alpha E_{\alpha 0}`$. After exclusion from this expression the accelerations $`\ddot{H}_{ij}`$ we get this constraint as the following combination of the equations of motion:
$`\phi ^{(3)}`$ $`=`$ $`_0^\mu E_{\mu 0}\xi _2G_{00}R^{ij}E_{ij}+{\displaystyle \frac{1}{D2}}\left[m^2G_{00}+(\xi _2\xi _1)RG_{00}+R_{00}\right]G^{ij}E_{ij}`$ (29)
$`=`$ $`\{{\displaystyle \frac{D1}{D2}}m^4+{\displaystyle \frac{2\xi _22\xi _1(D1)}{D2}}m^2R+(2\xi _2+{\displaystyle \frac{D1}{D2}})m^2R^0{}_{0}{}^{}\xi _2^2R_{ij}R^{ij}`$
$`+\xi _1^2RR+({\displaystyle \frac{\xi _2\xi _1(D1)}{D2}}2\xi _1\xi _2)RR^0{}_{0}{}^{}+\xi _2(\xi _2+1)R_{00}R^{00}\}H_{00}+\mathrm{}`$
We did not write down the explicit form of this constraint because everything we should know about it is the way it contains the component $`H_{00}`$. Namely, $`\phi ^{(3)}`$ contains neither the acceleration $`\ddot{H}_{00}`$ nor the velocity $`\dot{H}_{00}`$. It means that its conservation in time leads to another new constraints
$$\phi ^{(4)}_0\phi ^{(3)}$$
(30)
and hence the total number of constraints is the same as in the flat spacetime provided that the expression in the braces in front of $`H_{00}`$ in $`\phi ^{(3)}`$ does not vanish.
Vanishing of this expression in braces as well as violation of the inequality (26) leads to local changing of the number of degrees of freedom and this fact is known to be related with acausal behavior in higher spin theories in external fields . In general, causality breaks in those cases when there are points in spacetime in which it is impossible to define all the accelerations from the conservation of constraints.
In our case it means that causality will hold everywhere only if $`detD^i{}_{j}{}^{}0`$ and the expression in braces in $`\phi ^{(3)}`$ also does not vanish. Obviously, in general case there are values of $`R_{\mu \nu }`$ that violate these requirements.
It is instructive to consider in more detail the Reissner-Nordstrom solution in $`D=4`$ as a simple example of non-trivial static spacetime
$$ds^2=(1\frac{2M}{r}+\frac{Q^2}{r^2})dt^2+\frac{dr^2}{1\frac{2M}{r}+\frac{Q^2}{r^2}}+r^2d\mathrm{\Omega }^2$$
(31)
In this case causality problems are absent when the expressions
$`detD_i^j`$ $``$ $`\left(m^2(1+2\xi _2){\displaystyle \frac{Q^2}{r^4}}\right)\left(m^2+2{\displaystyle \frac{Q^2}{r^4}}\right)^2`$
$`\phi ^{(3)}`$ $`=`$ $`\left\{{\displaystyle \frac{D1}{D2}}m^4+m^2\left(2\xi _2+{\displaystyle \frac{D1}{D2}}\right){\displaystyle \frac{Q^2}{r^4}}+\xi _2(12\xi _2){\displaystyle \frac{Q^4}{r^8}}\right\}H_{00}+\mathrm{}`$ (32)
do not vanish. Far enough from the horizon where all the terms containing $`r`$ in (32) are too small and so propagation is causal in this region. Causal problems may develop only for small values of $`r`$. This might be excluded if all terms in (32) were positive, that is if
$$1+2\xi _2<0,2\xi _2+\frac{D1}{D2}>0,\xi _2(12\xi _2)>0$$
(33)
but these three conditions are contradictory.
It means that for any value of the coupling parameter expressions in (32) vanish for some values of the coordinate $`r`$ and the massive spin 2 field propagate causally only in the regions near infinity but close to the horizon causality is lost. Of course, this example does not mean that there cannot exist other spacetimes where causality might be achieved everywhere for some special values of coupling parameters.
Another possible way to fulfill the consistency requirements (22) is to consider spacetimes representing solutions of vacuum Einstein equations with arbitrary cosmological constant:
$$R_{\mu \nu }=\frac{1}{D}G_{\mu \nu }R.$$
(34)
In this case the coefficients $`a_4`$, $`a_5`$ in the lagrangian (13) are absent and the matrix $`\widehat{\mathrm{\Phi }}`$ takes the form:
$$\widehat{\mathrm{\Phi }}_\mu {}_{}{}^{\nu }=\left|\right|\begin{array}{cc}RG^{00}(2a_1+2a_2\frac{1}{D})& RG^{0j}(2a_1+4a_2)+2a_3R^{0j}{}_{0}{}^{}{}_{}{}^{0}+m^2G^{0j}\\ & \\ & \\ & \\ 0& 2a_3R^{0j}{}_{i}{}^{}{}_{}{}^{0}+RG^{00}\delta _i^j(2a_1\frac{2}{D})m^2G^{00}\delta _i^j\end{array}\left|\right|$$
(35)
The simplest way to make the rank of this matrix to be equal to $`D1`$ is provided by the following choice of the coefficients:
$$2a_1+2a_2\frac{1}{D}=0,a_3=0,2R\left(a_1\frac{1}{D}\right)m^20.$$
(36)
As a result, we have one-parameter family of theories:
$`a_1={\displaystyle \frac{\xi }{D}},a_2={\displaystyle \frac{12\xi }{2D}},a_3=0,a_4=0,a_5=0`$
$`R_{\mu \nu }={\displaystyle \frac{1}{D}}G_{\mu \nu }R,{\displaystyle \frac{2(1\xi )}{D}}R+m^20.`$ (37)
with $`\xi `$ an arbitrary real number.
The action in this case takes the form
$`S={\displaystyle }d^Dx\sqrt{G}\{{\displaystyle \frac{1}{4}}_\mu H^\mu H{\displaystyle \frac{1}{4}}_\mu H_{\nu \rho }^\mu H^{\nu \rho }{\displaystyle \frac{1}{2}}^\mu H_{\mu \nu }^\nu H+{\displaystyle \frac{1}{2}}_\mu H_{\nu \rho }^\rho H^{\nu \mu }`$
$`+{\displaystyle \frac{\xi }{2D}}RH_{\mu \nu }H^{\mu \nu }+{\displaystyle \frac{12\xi }{4D}}RH^2{\displaystyle \frac{m^2}{4}}H_{\mu \nu }H^{\mu \nu }+{\displaystyle \frac{m^2}{4}}H^2\}.`$ (38)
and the corresponding equations of motion are
$`E_{\mu \nu }=^2H_{\mu \nu }G_{\mu \nu }^2H+_\mu _\nu H+G_{\mu \nu }^\alpha ^\beta H_{\alpha \beta }_\sigma _\mu H^\sigma {}_{\nu }{}^{}_\sigma _\nu H^\sigma _\mu `$
$`+{\displaystyle \frac{2\xi }{D}}RH_{\mu \nu }+{\displaystyle \frac{12\xi }{D}}RHG_{\mu \nu }m^2H_{\mu \nu }+m^2HG_{\mu \nu }=0`$ (39)
The secondary constraints built out of them are
$$\phi _\mu ^{(2)}=^\alpha E_{\alpha \mu }=(_\mu H^\alpha H_{\mu \alpha })\left(m^2+\frac{2(1\xi )}{D}R\right)$$
(40)
and the matrix $`\widehat{\mathrm{\Phi }}`$ looks like
$$\widehat{\mathrm{\Phi }}_\mu {}_{}{}^{\nu }=(m^2+\frac{2(1\xi )}{D}R)\left|\right|\begin{array}{cc}0& G^{0j}\\ & \\ 0& G^{00}\delta _i^j\end{array}\left|\right|$$
(41)
Just like in the flat case, in this theory the conditions $`\dot{\phi }_i^{(2)}0`$ define the accelerations $`\ddot{H}_{0i}`$ and the condition $`\dot{\phi }_0^{(2)}0`$ after excluding $`\ddot{H}_{0i}`$ gives a new constraint, i.e. the acceleration $`\ddot{H}_{00}`$ is not defined at this stage.
To define the new constraint in a covariant form we use the following linear combination of $`\dot{\phi }_\mu ^{(2)}`$, equations of motion, primary and secondary constraints:
$`\phi ^{(3)}`$ $`=`$ $`{\displaystyle \frac{m^2}{D2}}G^{\mu \nu }E_{\mu \nu }+^\mu ^\nu E_{\mu \nu }+{\displaystyle \frac{2(1\xi )}{D(D2)}}RG^{\mu \nu }E_{\mu \nu }=`$ (42)
$`=`$ $`H{\displaystyle \frac{1}{D2}}\left({\displaystyle \frac{2(1\xi )}{D}}R+m^2\right)\left({\displaystyle \frac{D+2\xi (1D)}{D}}R+m^2(D1)\right)0.`$
This gives tracelessness condition for the field $`H_{\mu \nu }`$ provided that parameters of the theory fulfill the conditions:
$$\frac{2(1\xi )}{D}R+m^20,\frac{D+2\xi (1D)}{D}R+m^2(D1)0$$
(43)
Requirement of conservation of $`\phi ^{(3)}`$ leads to one more constraint
$`\dot{\phi }^{(3)}\dot{H}\phi ^{(4)}=\dot{H}0.`$ (44)
The last acceleration $`\ddot{H}_{00}`$ is expressed from the condition $`\dot{\phi }^{(4)}0`$.
Using the constraints for simplifying the equations of motion we see that the original equations are equivalent to the following system:
$`^2H_{\mu \nu }+2R^\alpha {}_{\mu }{}^{\beta }{}_{\nu }{}^{}H_{\alpha \beta }^{}+{\displaystyle \frac{2(\xi 1)}{D}}RH_{\mu \nu }m^2H_{\mu \nu }=0,`$
$`H^\mu {}_{\mu }{}^{}=0,\dot{H}^\mu {}_{\mu }{}^{}=0,^\mu H_{\mu \nu }=0,`$ (45)
$`G^{00}_0_iH^i{}_{\nu }{}^{}G^{0i}_0_iH^0{}_{\nu }{}^{}G^{0i}_i_0H^0{}_{\nu }{}^{}G^{ij}_i_jH_\nu ^02R^{\alpha 0\beta }{}_{\nu }{}^{}H_{\alpha \beta }^{}`$
$`{\displaystyle \frac{2(\xi 1)}{D}}RH^0{}_{\nu }{}^{}+m^2H^0{}_{\nu }{}^{}=0.`$
The last expression represents $`D`$ primary constraints.
For any values of $`\xi `$ (except two degenerate values excluded by (43)) the theory describes the same number of degrees of freedom as in the flat case - the symmetric, covariantly transverse and traceless tensor. $`D`$ primary constraints guarantees conservation of the transversality conditions in time.
Let us now consider the causal properties of the theory. Again, if we tried to use the equations of motion in the original lagrangian form (39) then the characteristic matrix
$$M_{\mu \nu }{}_{}{}^{\lambda \kappa }(n)=\delta _{(\mu \nu )}{}_{}{}^{(\lambda \kappa )}n_{}^{2}G_{\mu \nu }G^{\lambda \kappa }n^2+G^{\lambda \kappa }n_\mu n_\nu +G_{\mu \nu }n^\lambda n^\kappa \delta _\nu ^{(\kappa }n^{\lambda )}n_\mu \delta _\mu ^{(\kappa }n^{\lambda )}n_\nu $$
(46)
would be degenerate. This fact can be seen from the relation
$$n^\mu M_{\mu \nu }{}_{}{}^{\lambda \kappa }(n)0$$
(47)
which means that any symmetric tensor of the form $`n_{(\mu }t_{\nu )}`$ (with $`t_\nu `$ an arbitrary vector) represents a “null vector” for the matrix $`M(n)`$ and therefore $`detM=0`$.
After having used the constraints we obtain the equations of motion written in the form (45) and the characteristic matrix becomes non-degenerate:
$$M_{\mu \nu }{}_{}{}^{\lambda \kappa }(n)=\delta _{\mu \nu }{}_{}{}^{\lambda \kappa }n_{}^{2},n^2=G^{\alpha \beta }n_\alpha n_\beta .$$
(48)
The characteristic cones remains the same as in the flat case. At any point $`x_0`$ we can choose locally $`G^{\alpha \beta }(x_0)=\eta ^{\alpha \beta }`$ and then
$$n^2|_{x_0}=n_0^2+n_i^2$$
(49)
Just like in the flat case the equations are hyperbolic and causal.
Now let us discuss the massless limit of the theory under consideration. There are several points of view on the definition of masslessness in a curved spacetime of an arbitrary dimension. We guess that the most physically accepted definition is the one referring to appearance of a gauge invariance for some specific values of the theory parameters (see e.g. for a recent discussion).
In our case it means that the real mass parameter $`M`$ for the field $`H_{\mu \nu }`$ in an Einstein spacetime is defined as
$$M^2=m^2+\frac{2(1\xi )}{D}R$$
(50)
When $`M^2=0`$ instead of $`D`$ secondary constraints $`\phi _\mu ^{(2)}`$ we have $`D`$ identities for the equations of motion $`^\mu E_{\mu \nu }0`$ and the theory acquires gauge invariance $`\delta H_{\mu \nu }=_\mu \xi _\nu +_\nu \xi _\mu `$. This explains the meaning of the first condition in (43), it just tells us that the theory is massive.
In fact, two parameters $`m^2`$ and $`\xi `$ enter the action (38) in a single combination $`M^2`$ (50). Since scalar curvature is constant in Einstein spacetime there is no way to distinguish between the corresponding terms $`\xi RHH`$, $`m^2HH`$ (with arbitrary $`\xi `$, $`m`$) in the action. The difference between the two will appear only if we consider Weyl rescaling of the metric. Note that the “massless” theory with $`M^2=0`$ is not Weyl invariant. In the case of dS/AdS spacetimes the difference between masslessness, conformal and gauge invariance and null cone propagation was discussed in detail in . In our case the theory obviously cannot possess Weyl invariance.
The second inequality (43) is more mysterious. If it fails to hold, i.e. if $`M^2=M_c^2\frac{D2}{D(D1)}R`$ then instead of the constraint $`\phi ^{(3)}`$ the scalar identity
$$^\mu ^\nu E_{\mu \nu }+\frac{R}{D(D1)}G^{\mu \nu }E_{\mu \nu }=0$$
(51)
with the corresponding gauge invariance
$$\delta H_{\mu \nu }=_\mu _\nu ϵ+\frac{R}{D(D1)}G_{\mu \nu }ϵ$$
(52)
arise.
Appearance of this gauge invariance with a scalar parameter was first found for the massive spin 2 in spacetime of constant curvature in and was further investigated in spacetimes with positive cosmological constant. Our analysis shows that this gauge invariance is a feature of more general spin 2 theories in arbitrary Einstein spacetimes. In this case we can simplify the equations of motion using the secondary constraints (40):
$$^2H_{\mu \nu }_\mu _\nu H+2R_\mu {}_{\nu }{}^{\alpha }{}_{}{}^{\beta }H_{\alpha \beta }^{}+\frac{2D}{D(D1)}RH_{\mu \nu }\frac{1}{D(D1)}RG_{\mu \nu }H=0.$$
(53)
After imposing the gauge condition<sup>3</sup><sup>3</sup>3It does not fix (52) completely and the residual symmetry with the prameter obeying $`\left(^2+\frac{R}{D1}\right)ϵ=0`$ remains. $`H=0`$ one can see that these equations describe causal propagation of the field $`H_{\mu \nu }`$ but the number of propagating degrees of freedom corresponds to neither massive nor massless spin 2 free field. It was argued in that appearance of the gauge invariance (52) leads to such pathological properties as violation of the classical Hamiltonian positiveness and negative norm states in the quantum version of the theory. One should expect similar problems in the general spin 2 theory in arbitrary Einstein spacetime described in this paper.
We demonstrated that correct number of degrees of freedom in the massive spin 2 theory (algebraic consistency) can be achieved in a large class of curved spacetimes which include as particular cases arbitrary static spacetimes and vacuum Einstein spacetimes. An analysis of the constraints structure shows that in case of a static spacetime there esxists a potential source of acausal behavior. However, as we see on the example of Reissner-Norsdtrom spacetime causal propagation is possible in the regions where gravitational field is weak enough. In Einstein spacetimes spin 2 massive field can be consistently described by a one-parameter family of theories (38). For any value of the parameter satisfying (43) the corresponding equations describe the correct number of degrees of freedom which propagate causally.
It is interesting to compare our approach with the paper which is devoted to investigation of consistency of higher rank spin-tensor fields in curved spacetime from a different point of view. The authors of considered equations for the fields carrying irreducible representations of Euclidian version of the four dimensional Lorentz group SO(4) and analyzed when irreducibility of these representations is preserved in curved space. In particular, they showed that symmetric second rank tensor equations are consistent in this sense in Einstein spaces. However, such an analysis (sufficient for the proof of index theorems in ) is not enough when one tries to build a consistent theory for a physical field on a curved manifold starting from an irreducible representation of the Poincare group in flat spacetime. Preservation of the correct number of degrees of freedom in such a theory is a requirement independent from the algebraic consistency considered in . Einstein spacetimes provide an example when both these conditions are fulfilled but as we saw in our analysis correct number of degrees of freedom can be preserved in much wider class of spacetimes. Besides, in physical theories for interacting higher spin fields we face a new problem of causality which should also be studied independently. In general there can be theories with correct number of degrees of freedom but acausal, and we really see such examples in case of spin 2 field in external gravity. It is worth to note that the authors of emphasized that their analysis has no direct relation to the problem of consistent propagation of higher spin physical fields and that they did not set this problem at all.
In case of Einstein spacetimes our lagrangian for the spin 2 field in curved spacetime is the most general known so far, in all previous works only the theories with specific values of the parameter $`\xi `$ were considered . Two degenerate values of the parameter $`\xi `$ describe the theories with different degrees of freedom. One of this degenerate values corresponds to massless spin 2 field in an Einstein spacetime, another one describes neither massive nor massless spin 2 field.
The next natural step would consist in building a theory describing dynamics of both gravity and massive spin 2 field. In such a theory in addition to dynamical equations for the massive spin 2 field one would have dynamical equations for gravity with the energy-momentum tensor constructed out of spin 2 field components. The analysis of consistency then changes and one needs to have correct number of constraints and causality for both fields interacting with each other .
The only known consistent system of a higher spin field interacting with dynamical gravity is the theory of massless helicity 3/2 field, i.e. supergravity (see also the book ). In that case consistency with dynamical gravity requires four-fermion interaction. If a consistent description of spin 2 field interacting with dynamical gravity exists it may also require some non-trivial modification of the lagrangian. At least, it is known that lagrangians quadratic in spin 2 field do not provide such a consistency . A possible way of consistent description of the spin 2 field on arbitrary gravitational background was recently proposed in . This was achieved by means of representation of the lagrangian in the form of infinite series in curvature and imposing the consistency condition perturbatively in each order (earlier similar construction was investigated for symmetric Einstein spacetime in ).
Further generalizations of our analysis may include theories of massive spins $`s3`$ fields (which would require more complex structure of auxiliary fields) and interaction with other background fields, e.g. with scalar dilaton and antisymmetric tensor that are relevant in string theory.
Acknowledgements. We are grateful to V. Krykhtin, S. Kuzenko, H. Osborn, B. Ovrut, A. Tseytlin, M. Vasiliev and G. Veneziano for useful discussions of some aspects of this work. The work of I.L.B. and V.D.P. was supported by GRACENAS grant, project 97-6.2-34 and RFBR grant, project 99-02-16617; the work of I.L.B. and V.D.P. was supported by RFBR-DFG grant, project 99-02-04022 and INTAS grant N 991-590. I.L.B. is grateful to FAPESP and D.M.G. is grateful to CNPq for support of the research.
|
warning/0006/gr-qc0006067.html
|
ar5iv
|
text
|
# A Gaussian Weave for Kinematical Loop Quantum Gravity
## I Introduction
One of the most important physical results of loop quantum gravity is the prediction that the spectra of geometric operators, corresponding to their classical analogues such as areas of surfaces and volumes of regions, are purely discrete . Some other intriguing features of this ‘Non-perturbative Quantum Geometry’ have emerged already at the kinematical level, for example, the non-commutativity of geometric operators , and the statistical derivation of Black Hole entropy . Despite the relative success of the canonical/loop quantum gravity approach, two major problems need to be successfully tackled: dynamics (that is, a complete and anomaly-free implementation of the quantum Hamiltonian constraint) and the recovery of general relativity as a low energy/macroscopic regime of the quantum theory. How could the two problems be related is an open question, but it is likely that a complete understanding of the quasi-classical regime of the theory at the kinematical level (in particular states which can approximate functions of connections) can give some insight into the dynamics of the theory .
In this approach, General Relativity (GR) has to arise from the semi-classical limit of the quantum theory through properly defined ‘coherent states’ and a suitable coarse-graining. Coherent states have been constructed only recently in loop quantum gravity . However, we shall argue that some important issues have to be addressed first in order to obtain a complete understanding of the problem. At the kinematical level, the first attempts in this direction were given by the so called ‘weave states’ in the old loop representation . As originally constructed, weave states were intended to be semi-classical states (initially, eigenstates of geometrical operators) that approximate smooth geometries on a background spatial (and compact) 3-manifold $`\mathrm{\Sigma }`$ for large scales. That is, given a fiducial metric on $`\mathrm{\Sigma }`$, the spatial average of expectation values of area and volume operators in this state should be given by their classical values as measured by the fiducial 3-metric on $`\mathrm{\Sigma }`$ above some macroscopic scale $`L`$. There is increasing hope that once one has a good candidate for a weave describing flat space, one would be able to develop quantum field theory (QFT) on these states. There are some evidences pointing out in the direction that QFT’s on weaves would be free of infinities and should not require any renormalization procedure .
Another important issue that deserves further attention is whether semi-classical states can be promoted to solutions to the Hamiltonian constraint of the theory. It is well known that eigenstates of 3-geometries are not suitable states describing 4-dimensional Minkowski space-time, as happens, for instance, in quantum electrodynamics (QED) where an eigenstate of the electric field operator with null eigenvalue has a very different physical meaning than that of QED vacuum . However, kinematical results could be important by themselves in situations where classical boundary conditions such as asymptotical flat space times and isolated black holes boundary conditions are imposed on the system .
Recent attempts in constructing weaves were given in with the use of ‘spin networks’, and have been recently used as a background ‘thermal’ state to construct new Hilbert spaces for non-compact spaces . Nevertheless, weaves are defined based on the behavior of geometrical operators that depend only in momentum variables and therefore, only provide information about half of the classical phase space while the behavior of some connection variable operators might not be under control in these states. In fact, the original weaves were expected to be ‘concentrated’ in geometry, so they are highly delocalized in the connection variable and the fluctuations of the configuration operators might be very large. For instance, it has been recently shown that the natural strategy for approximating connections by taking the elementary configuration variable, namely the $`SU(2)`$ holonomy, as the multiplicative operator, fails . Any attempt to reproduce their classical analogue is doomed since there are no quantum states that support such supposition and such that they satisfy additional minimum uncertainty relations . Instead, a more ambitious proposal has been given by the authors of and consists on a tentative definition of ‘quasi-classicity’, a latticization of the underlying manifold and the use of ‘magnetic flux’ type operators. This was the first step in attempting to construct within the loop approach, realistic semi-classical states at the kinematical level (in the simplified case of two spatial dimensions). More recently, a series of papers by Thiemann and Winkler have appeared in which semi-classical states are constructed from a (complex) coherent-state transform on phase space (see and references therein).
In this work we follow the same strategy as in , namely we construct an (invertible) weave state based on a cylindrical function, that might serve as background for the construction of new Hilbert spaces when the space is non-compact. There is however, an important difference between our weave state and the one originally proposed in : Our state depends in a ‘Gaussian way’, i.e. on a suitable exponential arrangement of square of group elements. The state constructed here is peaked around the trivial ‘flat’ connection that yields the identity element for the holonomy around each loop defining the state. With this prescription, one might hope to overcome the objection previously raised regarding large fluctuations for (every) holonomy operator. In this sense, if one is able to reproduce the flat (metric) behavior through the geometric operators, and the connection is also ‘peaked’ around the flat connection, one might hope to have a semi-classical state approximating Minkowski space-time.
It is worth mentioning that the exponential dependence on group elements is a common feature of coherent states in compact Lie groups . We will show that this one-parameter family of states has better behavior than the original ‘quasi-coherent’ weave, and that in this case, the states are peaked and centered around the fundamental representation, when written in the spin network basis. This result is particularly intriguing, since a similar behavior, i.e., a dominance of the fundamental representation, is also observed in the computation of the black hole entropy .
This paper is organized as follows: In Section II we give a very brief review of the most important facts of loop quantum gravity that are relevant for this work. Readers familiar with the loop formalism may want to skip to Section III where the notion of a weave state is recalled. In the Section III B we proceed to the construction of the ‘Gaussian’ weave which approximates flat space. We also take the opportunity to make some general observations and, in particular, we discuss some subtleties that arise when an eigenbasis of the area operator is introduced. We close with a discussion in Section IV.
## II Preliminaries
Canonical/Loop quantum gravity is a canonical approach for quantizing general relativity in a non-perturbative way. In this approach, general relativity is a purely constrained system, a feature for all diffeomorphism invariant theories. A rigorous kinematical framework to handle theories whose classical configuration space is given by connections moduli local gauge transformations, has been constructed in the past decade . General relativity can be written in this form using the so-called Ashtekar-Barbero variables . In this section we summarize the main results of this approach.
The classical phase space consists of canonical pairs of a $`SU(2)`$ valued connection $`A_a^i`$ on an orientable spatial 3-manifold $`\mathrm{\Sigma }`$ and its conjugate momentum variable, a densitized triad $`E_i^a`$ which takes values in the dual of the Lie algebra . The dynamics of the theory, for compact $`\mathrm{\Sigma }`$, is pure gauge and is encoded in the Gauss constraint which generates $`SU(2)`$ gauge transformations, the spatial 3-dimensional diffeomorphism constraint generating spatial 3d diffeomorphisms on $`\mathrm{\Sigma }`$, and the Hamiltonian constraint which generates the coordinate time evolution.
The classical configuration space $`𝒜`$, is given by the space of all smooth connections on a principal $`SU(2)`$ bundle over $`\mathrm{\Sigma }`$. Since $`𝒜`$ is infinite dimensional the quantum configuration space $`\overline{𝒜}`$ is a certain completion of the classical one which includes all ‘distribution-like’ connection, in a similar role played in free field theory in Minkowskian space by the tempered distributions. Hence $`\overline{𝒜}`$ is taken to be the space of generalized connections which have a well defined action on oriented paths over $`\mathrm{\Sigma }`$ and the group of generalized $`SU(2)`$ gauge transformations .
Since $`\overline{𝒜}`$ is compact it admits a regular diffeomorphism-invariant measure $`d\mu _o`$ and the Hilbert space can be taken to be the space $`=L^2(\overline{𝒜},d\mu _o)`$ of square-integrable functions on $`\overline{𝒜}`$. This space is actually non-separable, but we can associate quantum states of the gauge theory to finite graphs $`\gamma `$ <sup>*</sup><sup>*</sup>*For simplicity we will take the paths to be analytic embedded.. If we consider all possible graphs, we obtain a very large set of states which are dense on $``$.
These states defined on finite graphs are called cylindrical functions and depend on finite sets of holonomies along the edges of the graph. Therefore, given any complex valued function $`\psi :(SU(2))^n𝐂`$ we can associate a cylindrical function $`\mathrm{\Psi }_\gamma (A)𝒜_\gamma `$ as follows:
$$\mathrm{\Psi }_{\gamma ,\psi }(A)=\psi (h{}_{e_1}{}^{}(A),..,h{}_{e_n}{}^{}(A)).$$
(1)
$`h_{e_i}`$ denote the $`SU(2)`$ parallel propagator matrices of the connection $`A\overline{𝒜}`$ along the curve $`e_i`$. Exploiting the compactness of $`SU(2)`$ the space of cylindrical functions $`𝒞`$ is equipped with the inner product:
$$\mathrm{\Psi }_{\gamma _1,\psi _1}\mathrm{\Psi }_{\gamma _2,\psi _2}=_{SU(2)^n}\overline{\psi _1(h{}_{e_1}{}^{},..,h{}_{e_n}{}^{})}\psi _2(h{}_{e_1}{}^{},..,h{}_{e_n}{}^{})d\mu _H(h{}_{e_1}{}^{})\mathrm{}d\mu _H(h{}_{e_n}{}^{})$$
(2)
where $`d\mu _H(h{}_{e_1}{}^{})\mathrm{}d\mu _H(h{}_{e_n}{}^{})`$ is the n-copy Haar measure naturally induced from $`SU(2)`$. If the graphs $`\gamma _1`$ and $`\gamma _2`$ are different, the respective cylindrical functions can be viewed as being defined in the same graph, which is the union of both: $`\gamma _1\gamma _2`$.
The classical algebra of elementary observables $`𝒮`$ consist of the set of canonical variables $`A_a^i`$ and $`e_{abi}:=\eta _{abc}E_i^c`$ smeared against suitable fields which in this case are one dimensional for the connection, and two-dimensional for the triad field . The space of configuration variables is given by the space Cyl of functions of the type (1), when the graph $`\gamma `$ is varied over all arbitrary finite graph. The momentum variables are obtained by smearing the pseudo 2-forms $`e_{abi}`$ against a test field $`f^i`$ which takes values in the Lie-algebra of $`SU(2)`$.
$${}_{}{}^{2}E[S,f]=_Se_{abi}f^i𝑑S^{ab}.$$
(3)
Here, the surface $`S`$ is restricted to be of the type $`S=\overline{S}\overline{S}`$, where $`\overline{S}`$ is any compact, analytic 2-dimensional sub-manifold of $`\mathrm{\Sigma }`$ possibly with boundary. See for details and for an alternative choice of variables. The standard representation of the quantum algebra on $``$ is to have the cylindrical (configuration) functions act as multiplicative operators, and “electric flux” variables as certain derivative operators .
When the spatial slice $`\mathrm{\Sigma }`$ is non-compact, a quantum state that describes, say, an asymptotically-flat space requires an infinite volume (and a contribution to this volume coming from everywhere within $`\mathrm{\Sigma }`$). This in turn, implies that the state must have contributions from graphs with an infinite number of edges. This state would not, in particular, belong to our Hilbert space $``$. In order to have a formalism that would allow for these situations, Arnsdorf and Gupta have proposed an algebraic construction, a la GNS, to overcome this difficulty. The basic idea is to construct new Hilbert spaces where a particular, semi-classical, non-normalizable state would serve as a vacuum of the theory, and the ordinary normalizable states would be seen as ‘finite excitations’ of the geometry. For details of the construction see . For our purposes, it is enough to know that the formalism requires an invertible state, that is, a state that does not vanish for any point on $`\overline{𝒜}`$, and that has support on an infinite number of edges. In the next section we shall construct such an state.
## III Approximating flat space
This section has two parts. In the first one we recall the general definition of a weave state. In the second part, we construct a Gaussian weave and show that it satisfies the requirements for a decent weave state and that it has some interesting features that make it an acceptable candidate for a quantum theory on non-compact spaces.
### A General Framework
In any quantum field theory, the semi-classical regime is obtained or represented by two main ingredients: “coherent states” and a suitable coarse graining. However, in the case of the gravitational field additional problems arise since the notion of “vacuum state” is not provided a priori, unless one forces it by splitting the metric into a preferred vacuum metric (usually taken flat) and a perturbative term whose fluctuations define the quantum theory. In the other hand, in canonical/loop quantum gravity, as in any other non-perturbative approach to quantum gravity, the natural ‘vacuum state’ corresponds to the ‘no metric’ space $`g_{ab}=0`$. In this case flat space becomes a highly excited state of the 3-geometry, containing an infinite number of fundamental excitations. Weave states are kinematic states constructed to provide semi-classical metric information over the spatial 3-manifold $`\mathrm{\Sigma }`$, that is, they approximate flat space above some macroscopic scale $`L`$. However, since the spatial metric $`g_{ab}`$ is constructed completely from momentum variables (densitized triads), weaves are normally thought to be highly concentrated in geometry, and it is possible that certain configuration variables, when promoted to operators, would be wildly delocalized in these states. This would render the weave states as unsuitable candidates for quasi-classical states. However, the weave state that we shall construct here is peaked, in the connection representation, around the flat connection for each loop of the defining graph. Thus, one expects that certain configuration observables be localized on these states. It is important to recall that all relevant results in the semi-classical regime of the theory have emerged at the kinematicalwherein the diffeomorphism and Hamiltonian constraint are ignored. level since a complete and anomaly-free implementation of the constraints have not been found successfully.
The discrete nature of quantum geometry provides a suitable criterion for approximating flat space. There are geometric operators which are self-adjoint in the Hilbert space $``$ and that measure areas of surfaces and volumes of regions. Therefore, a classical spatial metric can be approximated by requiring that the expectation values of areas and volumes of macroscopic surfaces and regions correspond to their classical analogues.
Let suppose that we want to approximate the Euclidean 3-space $`R^3`$ with the weave state $`\mathrm{\Phi }_w`$. The strategy is as follows: Let consider a macroscopic region in $`R^3`$ (of size larger than $`L`$) with bulk $`R`$ and surface $`S`$. We perform successive measurements of the volume of the region $`\widehat{V}_R`$ and the area of the surface $`\widehat{A}_S`$ on the state $`\mathrm{\Phi }_w`$ at different positions in $`\mathrm{\Sigma }`$. The state $`\mathrm{\Phi }_w`$ is called a ‘weave’ state if it satisfies the following requirements:
1. The spatial average of the expectation values of the area $`S`$ and volume $`R`$ agree with their classical values.
$$\overline{\widehat{A}_S}_w=_S\sqrt{q_{ab}}+O(ϵ^2/L^2),$$
(5)
$$\overline{\widehat{V}_R}_w=_R\sqrt{g_{ab}}+O(ϵ^3/L^3),$$
(6)
where $`L`$ is the characteristic macroscopic size of the bulk and $`ϵ`$ is a microscopic scale, which is normally taken to be of the order of the Planck length $`\mathrm{}_P`$, $`q_{ab}`$ is the induced metric on $`S`$, the “over-line” indicates spatial average and the expectation values are taken in the state $`\mathrm{\Phi }_w`$.
2. The state $`\mathrm{\Phi }_w`$ must have small quantum fluctuations in the measurements with regard to the macroscopic scale $`L`$.
$`\delta _VL^3`$ (8)
$`\delta _AL^2`$ (9)
where $`\delta _V`$ and $`\delta _A`$ are the standard deviations of the repeated measurements of the volume and area of the corresponding region.
Actually, the conditions (2) can be satisfied by an infinitely number of quantum states and do not single out a preferred state. In general, one could identify two distinct steps in the construction of a candidate of a weave state. First, one specifies a graph on $`\mathrm{\Sigma }`$ where the state is supposed to have support (or a family of graphs in a statistical sum). For instance, one could provide a lattice that ‘covers’ the whole space $`\mathrm{\Sigma }`$ , or a ‘unit’ graph $`\gamma `$ that is going to be spread over $`\mathrm{\Sigma }`$ to cover it. The second step involves the specification of a cylindrical function to be living on the basic units of the construction, be it on plaquetes of the lattice in the first case above, or basic cylindric functions on the graph $`\gamma `$. Some explicit examples of both types have been constructed in the literature, the first ones being simultaneous eigenstates of the volume and area operators (,) and a recent ‘quasi-coherent’ state with some nice properties. An example of states based on a lattice is given in .
In this note, we construct a particular weave state similar in spirit to the one given in . Actually, one is only proposing a cylinder function (i.e., the second step in the construction outlined above), which could in principle be implemented on different graphs. We choose a graph construction first proposed in (and also considered in ) for simplicity, and to be able to compare the properties of our Gaussian ansatz with the other proposals. It should be clear from the discussion that one could take our Gaussian ansatz and implement it in other graph/lattice constructions. Our ansatz has the property that it shares some of the nice properties of earlier proposals, that is: it is an invertible state, is peaked in the connection and the spin-network basis. However, it has the particular feature that the peak in the spin-network basis is centered around the fundamental representation. We shall refer to this state as the ‘Gaussian’ weave because it is constructed as a Gaussian function of cylindrical factors.
### B The ‘Gaussian’ Weave
The aim of a weave is to reproduce the area and volume of any region in the slice $`\mathrm{\Sigma }`$. Therefore, the graph in which the state is based must fill all $`\mathrm{\Sigma }`$ and if the space $`\mathrm{\Sigma }`$ is non-compact, one is left with two options: The graph either consists of edges of infinite length or it is an infinite set of finite graphs. In particular, our weave state is based on a disjoint union of an infinite collection of finite graphs: $`𝒢_\rho ^r=_{i=1}^{\mathrm{}}\gamma _i`$ where $`\gamma _i`$ represent the graph of a system of two non-coplanar circles $`\alpha _i`$ and $`\beta _i`$ with the same radius $`r`$ respect to a fiducial metric, which intersect in a single vertex, see Fig. 1. These systems of circles are randomly distributed in $`\mathrm{\Sigma }`$ with average density $`\rho `$ measured with the same fiducial metric.
Let us define the cylindrical function based in the graph $`\gamma _i`$
$$g_i(A)=N\mathrm{exp}(\lambda \mathrm{Tr}[(h_e{}_{i}{}^{\alpha }(A)h_e{}_{i}{}^{\beta }(A)e)_{(1)}])^2,$$
(10)
where $`N`$ is a normalization factor, $`\lambda `$ is a positive parameter, $`h_e{}_{i}{}^{\alpha ,\beta }(A)`$ are the holonomies along the circles $`\alpha `$ and $`\beta `$ respectively and the subindex $`1`$ represents the fundamental representation in $`SU(2)`$ (Color 1 in the terminology of Penrose). It is worth noting the Gaussian dependence on group elements. The Gaussian weave can be written as an infinite product of functions (10).
$$𝒲=\underset{n\mathrm{}}{lim}\underset{i=1}{\overset{n}{}}g_i(A).$$
(11)
This state does not belong to the Hilbert space $``$ but is defined through the cylindrical function $`g_i`$ based on the graph $`\gamma _i`$ and can be considered as a candidate for a background (thermal) state for constructing new Hilbert spaces as in .
The state (11) has common features with that given in and can be considered as a ‘quasi-coherent’ state in the same footing. By definition, the cylindrical functions given by the finite product $`𝒲_n=_{i=1}^ng_i(A)`$ take on their maximum values when the holonomies along the circles in the graph $`\gamma _i`$ are the identity in $`SU(2)`$ and if $`A_e`$ denote the connection which give the trivial holonomy, as $`n\mathrm{}`$ the function $`𝒲_n`$ become increasingly peaked around $`A_e`$. The width of the peak is regulated by the Gaussian parameter $`\lambda `$.
We are interested in exploring the main features of the state $`𝒲`$ and in showing that it satisfies the weave conditions (1), and (2). For this purpose we begin expanding the cylindrical function $`g_i`$ into the De-Pietri-Rovelli spin-network basis. The only elements of the basis that have a non-vanishing contribution are given by,
$$\mathrm{\Phi }_j=\mathrm{Tr}[(h_e{}_{i}{}^{\alpha }(A)h_e{}_{i}{}^{\beta }(A))_{(j)}],$$
(12)
based on the graph $`\gamma _i`$, where the subindex $`j`$ denote the ‘color’ representation of $`SU(2)`$. This choice arises because $`\mathrm{\Phi }_j`$ can be written as a linear expansion of spin-network states some of which are eigenstates of the area operator for an arbitrary analytic surface $`S`$ that intersect the graph $`\gamma _i`$. The fact that these states are not eigenstates of the area operator for any surface is particularly interesting and deserves further explanation.
On the full Hilbert space $``$, and for gauge-invariant cylindrical functions $`N_\gamma `$We will restrict to gauge-invariant states on $`\overline{𝒜}`$. This will be useful because it permits the transition to the loop representation where from the beginning one restrict oneself to gauge invariant states (spin-network states)., the spectrum of the area operator $`\widehat{A}`$ is unbounded from above and the ‘area gap’, which is the smallest non-zero eigenvalue, is given in the special situation when the graph intersect the surface $`S`$ with an -up or down- edge and a tangential edge at a bivalent vertex $`v`$ . If a single edge intersect the surface (without crossing $`S`$), or if the edges that meets at the vertex $`v`$ are all ‘up’ or ‘down’, the Gauss law constraint implies that the corresponding eigenvalue be $`0`$. Therefore, to obtain non-trivial eigenvalues of the area operator is necessary that the edges of the graph ‘cross’ the surface $`S`$. This means that for the system of circles $`\gamma _i`$, the state (12) is an eigenstate of the area operator with non-trivial eigenvalues if at least one of the circles crosses $`S`$, that is, if the same circle intersect the surface in two points (vertices). If one or both of the circles only ‘touch’ the surface in one point, the arcs that meet that point will be at the same side of $`S`$ and the corresponding eigenvalue would be $`0`$ as was stated above. The maximum number of intersections between $`\gamma _i`$ and the surface is $`4`$ and is obtained when both circles cross $`S`$.
On the other hand, given the spin-network $`s=(\gamma _i,\stackrel{}{j},\stackrel{}{l})`$ which assigns the labelling $`\stackrel{}{j}=(j,j)`$ to the circles $`\alpha _i`$ and $`\beta _i`$ of the graph $`\gamma _i`$ with the $`j`$ representation of $`SU(2)`$, and a labelling $`\stackrel{}{l}=(l_1)`$ of the single tetravalent vertex of $`\gamma _i`$ with an invariant tensor $`c_1`$, that is, an intertwining tensor from the representations of the incoming edges to the representations of the outgoing edges. The spin-network states can be constructed by contracting the matrices associated to the $`j`$ representation of the circles on $`\gamma _i`$ with the intertwining tensor $`c_1`$ at the tetravalent vertex. The basis element (12) when written in the spin-network (state) expansion, correspond to the choice of a particular basis of the space of the invariant tensor of the tetravalent vertex, and is given by the Fig. 2 in the De Pietri-Rovelli notation .
Graphically, the choice of the basis in the space of invariant tensors is equivalent to the ‘decomposition’ of the tetravalent vertex into a trivalent graph with two vertices joined by a ‘virtual’ edge $`j^{}`$. Since the Clebsh-Gordan condition must hold in each of the trivalent vertices, the only allowed value for the virtual vertex is $`j^{}=0`$. Thus, we would be lead to the conclusion that the basis element (12) is an eigenstate of the area operator. This is true in general, but it remains the subtle case when the tetravalent vertex of the graph $`\gamma _i`$ lies on the surface $`S`$. There are two possible (non-generic) configurations in this special case:
1. The first configuration is obtained when the surface $`S`$ divides the graph $`\gamma _i`$ into two halfs. Each of them has a part of the circle $`\alpha _i`$ and a part of the circle $`\beta _i`$. In this situation, (12) has no contribution to the area coming from the vertex (since $`j^{}=0`$) and is then an eigenstate of the area operator with eigenvalues proportional to $`8\pi \mathrm{}_P^2\sqrt{\frac{j}{2}(\frac{j}{2}+1)}`$ for each intersection of the circles with $`S`$.
2. The other ‘non-generic’ configuration is realized when the surface $`S`$ divides the graph $`\gamma _i`$ in such way that one of the circles, say $`\alpha _i`$, is entirely placed on one side of the surface and the circle $`\beta _i`$ on the other. In this case the basis (12) is not an eigenstate of the area operator and it is indispensable the use of the recoupling theorem in order to perform the suitable change of basis in which the area operator can take a diagonal form. Even when the state is written in a diagonal form, there will be in general contributions from eigenstates with different eigenvalues for the area.
At first sight this seems to be a problem when expanding the Gaussian function (10) into the basis elements (12) but a closer look shows that the number of ‘pathological’ configurations represents a zero measure set in the collection of all possible configurations. However, even though the configurations with the tetravalent vertex on $`S`$ do not contribute in a significant way to the expansion of the function (10), they are at the same time, conceptually relevant. The fact that not every spin-network state is an eigenstate of the area operator, in the case when the vertex has valence four or more, was noted earlier , but seems not to be widely recognized in the literature. This particular feature leads to the conclusion that the area operators of two surfaces that intersect along a line fail to commute for states which have 4 or higher valent vertex in the intersection. Thus, the quantum Riemannian geometry that arises from loop gravity is intrinsically non-commutativity . Therefore, some questions arise concerning the semi-classical regime of the theory: Under what conditions do we expect to obtain a commutative geometry in the semi-classical approach? If the notion of the weave is a good candidate to describe classical space, What are the conditions on the weave states in order to reconcile our classical notion of a commutative geometry? These issues are under current study.
Once we have shown that it is plausible to expand the cylindrical function (10) into the basis (12) we proceed to determine the coefficients of the linear expansion $`g_i(A)=_jc_j\mathrm{\Phi }_j`$ to be able to evaluate the expectation values of the area $`\widehat{A}`$, the volume $`\widehat{V}`$ and their corresponding deviations. Let us start by defining the cylindrical function $`g_i`$ by its series expansion
$$g_i=Ne^{\lambda ^2(\mathrm{\Phi }_12)^2}=Ne^{4\lambda ^2}\underset{n=0}{\overset{\mathrm{}}{}}\frac{\lambda ^nH_n(2\lambda )}{n!}\mathrm{\Phi }_1^n.$$
(13)
$`H_n(x)`$ denote the Hermite polynomials of order $`n`$. This expression can be reduce noting that the following tensorial-algebra relation holds in theory of representations of $`SU(2)`$.
$$\mathrm{\Phi }_1^n=\underset{j}{}a_j^n\mathrm{\Phi }_j\mathrm{where}a_j^n=\frac{(j+1)n!}{(\frac{nj}{2})!(\frac{n+j}{2}+1)!}\mathrm{for}\frac{nj}{2}Z^+,$$
(14)
and $`a_j^n=0`$ otherwise. Defining $`𝒞(\lambda )=Ne^{4\lambda ^2}`$ the coefficients $`c_j`$ of the expansion can be written as
$$c_j=𝒞(\lambda )\underset{n=0}{\overset{\mathrm{}}{}}\frac{\lambda ^nH_n(2\lambda )}{n!}a_j^n,$$
(15)
Thus, using the relations (14) the cylindrical function (10) expanded on the basis $`\mathrm{\Phi }_j`$ is given by
$$g_i(A)=𝒞(\lambda )\underset{j}{}(j+1)\underset{n=0}{\overset{\mathrm{}}{}}\frac{\lambda ^nH_n(2\lambda )}{(\frac{nj}{2})!(\frac{n+j}{2}+1)!}\mathrm{\Phi }_j,$$
(16)
We would want to evaluate the sum in the index $`n`$ of the equation (16) but it can not be done exactly. Therefore, one is forced to use the definition of Hermite polynomials in its series expansion an evaluate the expression numerically. The coefficients $`c_j`$ are given by
$$c_j=𝒞(\lambda )(j+1)\underset{k=0}{\overset{\mathrm{}}{}}\frac{\lambda ^{2k+j}}{k!(k+j+1)!}\underset{m=0}{\overset{[\frac{2k+j}{2}]}{}}\frac{(1)^m(2k+j)!}{m!(2k+j2m)!}(4\lambda )^{2k+j2m},$$
(17)
where $`[\nu ]`$ denotes the larger integer $`\nu `$.
A numerical evaluation of equation (17) was performed and shows, with the normalization $`_{i=0}^{\mathrm{}}c_j^2=1`$, that the series converges in the interval $`0<\lambda <1`$ and for values above $`\lambda 0.5`$ a peak starts to grow at the fundamental representation. Fig. 3 shows some coefficients for some values of $`\lambda `$. It is important to stress out that for larger values of the ’color’ $`j`$ the expansion continues, but in this case a numerical evaluation turns out to be inappropriate due to float point errors.
The most important and intriguing difference between the ‘quasi-coherent’ weave and the ‘Gaussian’ weave is that, in the last case the peak of the coefficients (17) in the basis of spin-network states is centered around the fundamental representation $`j=1`$. The dominance of the fundamental representation in other physical situation such as the BH entropy seems to point to a deep role played by the fundamental representation in loop quantum gravity. The origin of this behavior deserves further attention.
One can get a closed formula for the expectation value for the area in the state $`g_i`$ using the machinery of recoupling theory in the loop representation
$$\widehat{A}=\sqrt{\frac{j}{2}(\frac{j}{2}+1)}\underset{k}{}\underset{j}{}c_k^{}c_j\text{ }\text{}\text{ }\text{ }\text{}\text{ }.$$
(18)
Where the bracket can be evaluated chromatically to show that the basis states are normalized.
Thus, the expectation value of the area operator is given by,
$$\widehat{A}=(8\pi \mathrm{}_P^2)\underset{j}{}|c_j|^2\sqrt{\frac{j}{2}(\frac{j}{2}+1)}.$$
(19)
In the case of the volume operator, a closed formula for the expectation value is not available. However, we know that the largest eigenvalue in the state $`\mathrm{\Psi }_j`$ increase as $`j^{3/2}`$ and we can make an estimation. For the particular value of the Gaussian width $`\lambda =0.75`$ we have the following expectation values for the geometric operators and their respective dispersions:
$`\widehat{A}=1.149(8\pi \mathrm{}_P^2);\mathrm{\Delta }_A:=\sqrt{\widehat{A}^2\widehat{A}^2}=0.432(8\pi \mathrm{}_P^2),`$ (21)
$`\widehat{V}=1.078(8\pi \mathrm{}_P^2)^{3/2};\mathrm{\Delta }_V:=\sqrt{\widehat{V}^2\widehat{V}^2}=2.435(8\pi \mathrm{}_P^2)^{3/2},`$ (22)
As expected, the mean values and dispersion of the basic geometric observable are of the order of $`\mathrm{}_P^2`$ in the case of the area and of $`\mathrm{}_P^3`$ in the case of the volume. Our state $`g_i`$ is peaked in area and volume as in a similar fashion as the proposal given in , but with better accuracy. It is easy to see that the state (11) satisfies the necessary conditions to be considered a weave state, because the expectation values (3) are of the order of suitable powers of the Planck length. For a formal demonstration we refer to .
## IV Discussion and outlook
In this work we have proposed a basic cylindrical function with a Gaussian dependence on the group element. This function can be used, in particular, to construct a weave state a la Grot-Rovelli. The particular proposal has the desired features of a weave state plus a peakedness property around the fundamental representation.
There exists a striking similarity between the main contribution of the state in the fundamental representation and the dominating states, in the statistical counting, that contributes to the entropy of an ‘isolated horizon’. This dominant contribution correspond to punctures all of which have labels in the fundamental representation (spin $`1/2`$) .
The Gaussian state proposed here can play an important role both as a background to construct new Hilbert spaces and as a suitable state in the recent proposal for lattice based semi-classical states in loop quantum gravity .
We finish by pointing out several issues that deserve further investigation:
1. One would like to understand the physical foundations for the fact that there is a dominance of the fundamental representation in at least two different situations.
2. It is necessary to investigate the role played by these (or other) ‘quasi-classical’ states in the recovery of a commutative classical geometry.
3. In the construction of new Hilbert spaces a la GNS, different choices of ‘vacuum states’ might lead to unitarily inequivalent Hilbert spaces. It is important to understand if the Gaussian ansatz presented here yields an equivalent or inequivalent quantum theory to the one constructed in . There are also coherent states constructed using ‘infinite tensor product Hilbert spaces’ . The relation of this method to the algebraic GNS construction remains unclear and deserves further attention.
4. Finally, we would like to understand the relation of the Gaussian ansatz to the coherent states constructed in , specially since both are constructed using coherent state-like states on the group.
Some of these issues are under investigation and shall be reported elsewhere.
###### Acknowledgements.
We would like to thank S. Gupta and J.A. Zapata for discussions, and a referee for helpful comments. This work was partially funded by DGAPA No. IN121298 and CONACyT No. J32754-E grants. JMR was supported in part by CONACyT scholarship No. 85976.
|
warning/0006/gr-qc0006036.html
|
ar5iv
|
text
|
# Nonthermal nature of extremal Kerr black holes
## 1 Introduction
Recent results in quantum field theory in curved spacetime have demonstrated fairly conclusively that extremal Reissner-Nordström (RN) black holes are of a fundamentally different class than their nonextremal counterparts. Liberati, Rothman and Sonego (LRS, ) examined spherical bodies collapsing into extremal Reissner-Nordström black holes and showed that the object’s particle-emission spectrum is nonthermal. Consequently a temperature is undefined for them. Furthermore, at no time during the history of the object do the properties of the stress-energy tensor of the radiation correspond to those of a nonextremal RN hole, and thus extremal holes never behave as thermal objects. Simultaneously with the LRS work, Anderson, Hiscock and Taylor demonstrated that static extremal RN solutions for “physically realistic” values of the curvature coupling constant simply do not exist to perturbative order in semi-classical gravity. Taken together, these results and others (see LRS for references) constitute strong evidence that the $`Q^2=M^2`$ RN solutions, which are the absolute-zero state in black hole dynamics, are for reasons as yet unclear disallowed by nature.
Of greater astrophysical interest than charged black holes are rotating black holes. For that reason it is important to extend the results to the extremal Kerr metric. It is well known (see, e.g. Hawking or DeWitt ) that the spectrum of the nonextremal Kerr solution can be obtained from the Schwarzschild-Droste (SD)<sup>1</sup><sup>1</sup>1Johannes Droste, a pupil of Lorentz, independently announced the “Schwarzschild” exterior solution within four months of Schwarzschild. See, “The field of a single centre in Einstein’s theory of gravitation, and the motion of a particle in that field,” Koninklijke Nederlandsche Akademie van Wetenschappen, Proceedings 19, 197 (1917). spectrum merely by taking into account the frequency shift induced by the rotation of the hole (below). One suspects that the extremal Kerr solution can be obtained from the extremal RN solution in the same way but because previous work has shown that extremality is dangerous territory, one might want to see this as explicitly as possible. In this letter I demonstrate that, to be sure, the behavior of the extremal Kerr solution is essentially identical to that of the RN case. Thus it appears that the nonthermal nature of extremal black holes is a generic property. In retrospect, this is to have been expected because the nature of the horizon changes completely at extremality .
LRS used the well known “moving mirror” technique to model the collapse. Rather than deal with the full four-dimensional problem one considers a one-dimensional mirror moving in two-dimensional Minkowski space. Massless scalar fields are propagated inwards from $`^{}`$ and bounced off the mirror to $`^+`$. Due to the mirror’s motion, one expects that the In and Out vacuum states will differ, leading to particle production. The spectrum is a function of the mirror’s trajectory, and so a key step in the calculation is to determine the worldline of the star’s center just before formation of the horizon. This late-time trajectory is taken in turn to correspond to the asymptotic trajectory of the mirror. Unfortunately, Kruskal corrdinates, which are typically employed in such calculations and which contain the surface gravity $`\kappa `$, break down in the extremal limit where $`\kappa `$ vanishes (below). LRS were able to provide a simple coordinate extension for this case, which allowed them to calculate the late-time history of the center of the collapsing object. Remarkably, this turned out to correspond precisely to that of a uniformly accelerated mirror in Minkowski space. The properties of such a system are well known . In particular, the emission spectrum is nonplanckian, demonstrating that temperature is undefined for these objects. I now show that the same applies to the Kerr solution.
## 2 Coordinate extension for the extremal Kerr solution
The Kerr metric in Kerr coordinates is
$`ds^2`$ $`=`$ $`[12Mr\rho ^2]dv^2+2drdv+\rho ^2d\theta ^2+\rho ^2[(r^2+a^2)^2`$ (2.1)
$`\mathrm{\Delta }a^2sin^2\theta ]sin^2\theta d\varphi _{}^22asin^2\theta d\varphi _{}dr4Mra\rho ^2sin^2\theta d\varphi _{}dv,`$
where as usual
$$dv=dt+\frac{(r^2+a^2)dr}{\mathrm{\Delta }};d\varphi _{}=d\varphi +\frac{adr}{\mathrm{\Delta }}$$
(2.2)
and
$$\rho ^2=r^2+a^2cos\theta ;\mathrm{\Delta }=r^22Mr+a^2.$$
(2.3)
Note that $`dv`$ represents an advanced null coordinate corresponding to the $`dv=dt+dr_{}`$ employed in SD or RN calculations, and that $`a`$ is the angular momentum parameter.
To find the late-time history of the collapsing body’s center, it is useful to employ double-null coordinates and so we define
$$du=dt\frac{(r^2+a^2)dr}{\mathrm{\Delta }}$$
(2.4)
Before rewriting the Kerr metric in terms of $`u,v`$, it is important to verify that these coordinates are continuous in the extremal limit $`a=M`$. This is easily done. Integrating the expression for $`dv`$ in (2.2) yields for the nonextremal case:
$$\begin{array}{ccc}v(a,M)\hfill & \hfill & t+r_{}(a,M)\hfill \\ & =\hfill & t+\frac{1}{2\sqrt{M^2a^2}}\{(r_+r_{})r+(r_+^2+a^2)\mathrm{ln}(rr_+)\hfill \\ & & (r_{}^2+a^2)\mathrm{ln}(rr_{})\}+\mathrm{const}.\hfill \end{array}$$
Here, $`r_\pm M\pm \sqrt{M^2a^2}`$.
The extremal value for $`v`$ is found by setting $`a=M`$ in the expression for $`dv`$ before integrating. Then $`\mathrm{\Delta }=(rM)^2`$ and
$$dv(M,M)=dt+\frac{(r^2+M^2)dr}{(rM)^2}.$$
(2.5)
Integrating gives
$`v(M,M)`$ $``$ $`t+r_{}(M,M)`$ (2.6)
$`=`$ $`t+r+2M\mathrm{ln}(rM){\displaystyle \frac{2M^2}{(rM)}}+\mathrm{const}.`$
Note that (2) appears to give $`0/0`$ when $`a=M`$, whereas (2.6) is well behaved except at $`r=M`$, the extremal horizon. Nevertheless, if one sets $`M=a+ϵ,ϵ<<a`$ and works to first order in $`ϵ`$, it is straightforward to show that (2.6) is the limit of (2) as $`ϵ0`$. Therefore, the coordinate $`v`$ is continous at extremality. The same holds for $`u`$ and $`\varphi _{}`$.
Since $`u`$ and $`v`$ are good at extremality, we now specialize to the case $`a=M`$. The metric (2.1) becomes
$`ds^2`$ $`=`$ $`[1{\displaystyle \frac{2Mr}{\rho ^2}}]dv^2+2drdv+\rho ^2d\theta ^2+{\displaystyle \frac{1}{\rho ^2}}[(r^2+M^2)^2(rM)^2M^2sin^2\theta ]sin^2\theta d\varphi _{}^2`$ (2.7)
$`2Msin^2\theta drd\varphi _{}{\displaystyle \frac{4M^2r}{\rho }}\mathrm{sin}^2\theta d\varphi _{}dv.`$
Eliminating $`dr`$ in favor of $`du`$ and $`dv`$ yields
$$\begin{array}{ccc}ds^2\hfill & =\hfill & [\frac{(rM)^2}{(r^2+M^2)}+(\frac{2M}{r}1)]dv^2\frac{(rM)^2}{(r^2+M^2)}dudv\hfill \\ & & +\frac{1}{\rho ^2}[(r^2+M^2)^2(rM)^2M^2sin^2\theta ]sin^2\theta d\varphi _{}^2\hfill \\ & & [\frac{4M^2r}{\rho ^2}+\frac{M(rM)^2}{(r^2+M^2)}]sin^2\theta d\varphi _{}dv+\frac{M(rM)^2}{(r^2+M^2)}sin^2\theta d\varphi _{}du+\rho ^2d\theta ^2.\hfill \end{array}$$
This metric is evidently degenerate at $`r=M`$ and so the coordinates $`u,v`$ are not good on the horizon at extremality. However, the $`(rM)^2`$ degeneracy is of exactly the same form LRS found for the extremal RN case. Thus, the same coordinate extension should work here as well. We define a transformation
$$\psi (\xi )=4M\left(\mathrm{ln}\xi \frac{M}{2\xi }\right)$$
(2.8)
and assume
$$\begin{array}{ccc}u=\psi (𝒰)\hfill & \hfill & 𝒰=\psi ^1(u)\hfill \\ v=\psi (𝒱)\hfill & \hfill & 𝒱=\psi ^1(u)\hfill \end{array}$$
(2.9)
where $`𝒰`$ and $`𝒱`$ are the new coordinates that should be regular on the horizon at extremality.
To motivate this choice of $`\psi `$, recall that $`v,u`$ are constructed adding or subtracting $`r_{}`$ to $`t`$. In the SD case, $`r_{}`$ has only a logarithmic divergence at $`r=2M`$, as can be seen by setting $`a=0`$ in (2). (The same holds for nonextremal RN). If one took only the first term in (2.8), $`\psi (u)=4M\mathrm{ln}(𝒰)`$, then one would have $`𝒰=e^{\kappa u};\kappa =1/4M`$, which is exactly the Kruskal transformation. However, for charged, rotating holes $`\kappa (M^2a^2Q^2)`$. Thus, in the extremal limit, the surface gravity vanishes and, as mentioned above, the Kruskal transformation becomes ill-defined. From (2.6), however, we observe that at extremality, $`r_{}`$ has not only the logarithmic divergence of the nonextremal case but the added pole $`(2M^2)/(rM)`$. We therefore pick the simplest possible generalization of the Kruskal transformation and add this term to $`\psi `$ to get (2.8). Note also that $`\psi ^{}(\xi )=4M/\xi +2M^2/\xi ^2>0`$, always, and so $`\psi `$ is monotonic; therefore (2.9) is a well-defined coordinate transformation. From (2.6) we have at extremality
$$r_{}(M,M)=r+\frac{1}{2}\psi (rM),$$
(2.10)
Near the horizon $`v`$ is constant, which implies $`tr_{}`$ and therefore $`u2r\psi (rM)`$. Thus, $`𝒰=\psi ^1(rM)`$ and the derivative $`\psi ^{}(𝒰)4M^2/(rM)^2`$. It is this factor of $`(rM)^2`$ in the denominator that will kill the $`(rM)^2`$ factors in the numerator of (2), for when we rewrite the metric in terms of the new variables $`𝒰,𝒱`$ we find
$$\begin{array}{cc}ds^2\hfill & [\frac{(rM)^2}{(r^2+M^2)}+(\frac{2M}{r}1)]\psi ^{}(𝒱)^2d𝒱^2\frac{4M^2}{(r^2+M^2)}\psi ^{}(𝒱)d𝒰d𝒱\hfill \\ & +\frac{1}{\rho ^2}[(r^2+M^2)2(rM)^2M^2sin^2\theta ]sin^2\theta d\varphi _{}^2\hfill \\ & [\frac{4M^2r}{\rho }+\frac{M(rM)^2}{(r^2+M^2)}]sin^2\theta d\varphi _{}\psi ^{}(𝒱)d𝒱+\frac{4M^3}{(r^2+M^2)}d\varphi _{}d𝒰+\rho ^2d\theta ^2\hfill \end{array}$$
Note that since $`𝒱`$ is everywhere nonzero and finite, $`\psi ^{}(𝒱)`$ is regular on the horizon. Thus nothing particular happens at $`r=M`$ and $`\psi `$ is a good coordinate transformation there. Note also that the degeneracy on the horizon is due solely to the behavior of the coordinate $`u`$. For this reason the coordinate extension $`\psi `$ is good for any value of $`\theta `$ and the results are completely general.
## 3 Late-time history
We now use $`\psi `$ to construct the late-time history of the collapsing star’s center in the coordinates $`u,v`$. LRS performed such a calculation for the case of spherical symmetry. Of course the Kerr metric is only axially symmetric, not spherically symmetric, but if we specialize to the equatorial plane $`\theta =\pi /2`$, then the problem becomes two-dimensional and can be treated in the same way. Moreover, as long as we are concerned solely with the center of the star (as opposed to, say, the surface), all angular coordinates become degenerate and cannot be relevant. Thus, even for arbitrary $`\theta `$, one evidently needs to consider only $`r`$ and $`t`$ (or $`u`$ and $`v`$), as LRS did for the RN case.
The coordinates $`u`$ and $`v`$ are valid only outside the collapsing star and so the trajectory of the center, we must extend them to the interior. This is accomplished in standard fashion . Ignoring the angular variables, the interior can be described by some interior null coordinates, $`U`$ and $`V`$ and the metric components $`g_{\mu \nu }(U,V)`$ can be chosen to be regular on the horizon. The center of the star can be taken at radial coordinate = 0, in which case $`V=U`$ and $`\mathrm{d}V=\mathrm{d}U`$ there.
Because the Kruskal-like coordinates $`𝒰`$ and $`𝒱`$ are regular everywhere as well, they can be matched to $`U`$ and $`V`$. In particular, if two nearby outgoing rays differ by $`\mathrm{d}U`$ inside the star, then they will also differ by $`\mathrm{d}U=\beta (𝒰)\mathrm{d}𝒰`$, with $`\beta `$ a regular function, outside. By the same token, since $`V`$ and $`v`$ are regular everywhere, we have $`\mathrm{d}V=\zeta (v)\mathrm{d}v`$, where $`\zeta `$ is another regular function. In fact, if we consider the last ray $`v=\overline{v}`$ that passes through the center of the star before the formation of the horizon, then to first order $`\mathrm{d}V=\zeta (\overline{v})\mathrm{d}v`$, where $`\zeta (\overline{v})`$ is now constant.
We can write near the horizon
$$\mathrm{d}U=\beta (0)\frac{\mathrm{d}𝒰}{\mathrm{d}u}\mathrm{d}u.$$
(3.1)
Since for the center of the star $`\mathrm{d}U=\mathrm{d}V=\zeta (\overline{v})\mathrm{d}v`$, this immediately integrates to
$$\zeta (\overline{v})(v\overline{v})=\beta (0)𝒰(u)=\beta (0)\psi ^1\left(u\right)2\beta (0)\frac{M^2}{u}.$$
(3.2)
The last approximation follows from Eq. (2.8) where $`\xi \psi ^1(2M^2/\xi )`$ near the horizon.
Thus the late-time worldline for the center of the star is, finally, represented by the equation
$$v\overline{v}\frac{A}{u},u+\mathrm{},$$
(3.3)
where $`A=2\beta (0)M^2/\zeta (\overline{v})`$ is a positive constant that depends on the details of the internal metric and consequently on the dynamics of collapse.<sup>2</sup><sup>2</sup>2One might wonder why we did not consider $`dV=\zeta (𝒱)d𝒱`$ instead of $`\mathrm{d}V=\zeta (v)\mathrm{d}v`$. In the former case we do not know the function $`\zeta (𝒱)`$ and so the above integration could not be carried out. This worldine is obviously hyperbolic and corresponds identically to that of a uniformly accelerated mirror in Minkowski space. This should be contrasted with the late-time worldline of a nonextremal hole, as can be found by using the “Kruskal-part” of the transformation $`\psi `$ in (3.2).
$$v\overline{v}B\mathrm{e}^{\kappa u},u+\mathrm{}.$$
(3.4)
As shown by LRS for the case of RN holes, the hyperbolic worldline cannot be obtained as an approximation of the nonextremal worldline (3.4) and so (3.3) represents a true discontinuity between extremal RN holes and their nonextremal counterparts. We now see the same conclusion applies to extremal Kerr holes.
With the benefit of hindsight this result is perhaps not so surprising, at least for the RN case. Consider the typical homework problem of dropping a rock into a black hole from some $`r=R_o`$. One generally solves this by starting from the condition on the four-velocity $`u_\mu u^\mu =1`$ and the condition $`u_o=\stackrel{~}{E}`$ = energy per unit mass = constant. For RN one has $`u_o=(12M/r+Q^2/r^2)u^o`$ and $`u_r=(12m/R+Q^2/r^2)^1u^r`$, with $`u^o=dt/d\tau `$ and $`u^r=dr/d\tau `$. Forming $`dr/dt`$ for the extremal RN metric, one gets after a few elementary substitutions
$$dt=\frac{Mz^3dz}{(z1)^3[z^2e^2]^{1/2}}$$
(3.5)
where $`e\stackrel{~}{E}^1`$ and $`z1+M/(rM)`$. The integral is cumbersome but can be evaluated analytically in terms of elementary functions. The result near the horizon ($`z>>1`$) is
$$t=k_1+Mz+2M\mathrm{ln}z+\frac{k_2}{z}+𝒪(\frac{1}{z^2})$$
(3.6)
where $`k_1`$ and $`k_2`$ are constants. For the extremal RN solution, $`r_{}=rM^2/(rM)+2M\mathrm{ln}(rM)`$. Near the horizon ($`r=M`$) the $`M^2`$ term dominates. Using this and the fact that in the same regime $`tr_{}u2r_{}`$, one brings the above equation into the form
$$vk_1+\frac{k_3}{u},$$
(3.7)
again a hyperbolic worldline, as distinct from the exponential form similar to (3.4) that one finds for a rock falling into a SD or nonextremal RN black hole. As mentioned in the Introduction, the horizon structure changes completely at extremality. Moreover, the horizon itself is moved in to $`r=M`$, and the shape of the effective potential is totally altered. Given all this, it should really come as no suprise that the trajectories of objects falling into extremal versus nonextremal holes should differ.
## 4 Nonthermal Spectrum
To implement the moving mirror analogy, one requires that the wave equation $`\varphi _{;\mu }^{;\mu }=0`$ be separable, which implies that $`\varphi =f(u)+g(v)`$. Then one move the problem to Minkowski space, where the null coordinates are now assumed to be $`u=tx`$ and $`v=t+x`$. The mirror trajectory is taken to be that of the center of the star, $`v=p(u)`$ . With the boundary conditions that one has pure ingoing solutions on $`^{}`$ and pure outgoing solutions on $`^+`$ and that there be total reflection at the mirror, i.e., $`\varphi (p(u),u)=0`$, it is easy to show that
$$\varphi _\omega ^{(\mathrm{in})}(u,v)=\frac{\mathrm{i}}{\sqrt{4\pi \omega }}\left(\mathrm{e}^{\mathrm{i}\omega v}\mathrm{e}^{\mathrm{i}\omega p(u)}\right)$$
(4.1)
and
$$\varphi _\omega ^{(\mathrm{out})}(u,v)=\frac{\mathrm{i}}{\sqrt{4\pi \omega }}\left(\mathrm{e}^{\mathrm{i}\omega u}\mathrm{\Theta }\left(\overline{v}v\right)\mathrm{e}^{\mathrm{i}\omega q(v)}\right).$$
(4.2)
Here $`q(v)=p^1(v)`$ and $`\omega >0`$. (See LRS for further details.)
The relevant Bogoliubov coefficient is defined as
$$\beta _{\omega \omega ^{}}=(\varphi _\omega ^{(\mathrm{out})},\varphi _\omega ^{}^{(\mathrm{in})})=\mathrm{i}_0^+\mathrm{}dx\left[\varphi _\omega ^{(\mathrm{out})}(u,v)_t^{^{^{}}}\varphi _\omega ^{}^{(\mathrm{in})}(u,v)\right]_{t=0}.$$
(4.3)
which gives a particle spectrum
$$N_\omega =_0^+\mathrm{}d\omega ^{}|\beta _{\omega \omega ^{}}|^2.$$
(4.4)
For SD and nonextremal RN, it is the evaluation of (4.3) and (4.4) that yields the black-body spectrum because the function $`p`$ is exponential at late times. For the extremal RN case, however, one uses the hyperbolic form, equivalent to a uniformly accelerated mirror. The properties of the uniformly accelerated mirror have been extensively studied and the Bogoliubov coefficients have been calculated . In the asymptotic regime for the trajectory (3.3)
$$\beta _{\omega \omega ^{}}\frac{\sqrt{A}}{\pi }\mathrm{e}^{\mathrm{i}\sqrt{A}(\omega \omega ^{})}K_1(2(A\omega \omega ^{})^{1/2}),$$
(4.5)
where $`K_1`$ is a modified Bessel function. For argument $`z`$, $`K_1(z)1/z`$ for $`z0`$, and $`K_1(z)\sqrt{\pi /(2z)}\mathrm{e}^z`$ when $`z+\mathrm{}`$
For the extremal Kerr black hole the analysis goes through almost unchanged. The solutions to wave equation for Kerr can in fact be separated into the above form , where now $`u`$ and $`v`$ are exactly the null coordinates we have been using. In the vicinity of the Kerr horizon, the radial wave function has the form $`Rexp(\pm i\stackrel{~}{\omega }r_{})`$. Here, $`\stackrel{~}{\omega }\omega m\mathrm{\Omega }`$, $`m`$ is the azimuthal quantum number and $`\mathrm{\Omega }=1/(2M)`$ is the angular velocity of the black hole on the horizon at extremality. The outgoing modes $`\varphi ^{(out)}exp(i\stackrel{~}{\omega }u)`$ get transported back to to become $`\varphi ^{(out)}exp(i\stackrel{~}{\omega }q(v))`$ on $`^{}`$. The Bogoliubov coefficient is essentially the Fourier transform of this quantity,
$$\beta _{\stackrel{~}{\omega }^{}\stackrel{~}{\omega }}exp\{i\stackrel{~}{\omega }^{}v+i\stackrel{~}{\omega }q(v)\}𝑑v$$
(4.6)
which is the relevant integral one encounters while evaluating (4.3). Thus, since we have established that $`p(u)`$ and hence $`q(v)`$ are the same for both extremal Kerr and extremal RN, the Bogoliubov coefficients are also the same with the replacement $`\omega \stackrel{~}{\omega }`$. This gives the frequency shift mentioned in the Introduction. Consequently, with the substitution $`\stackrel{~}{\omega }`$ for $`\omega `$, the extremal Kerr spectrum is found to be identical to the extremal RN spectrum.
The main point is that since the spectrum goes like the square of a $`K_1`$ Bessel function it is definitely not a black-body spectrum. Therefore temperature is simply undefined. Moreover, as shown by LRS the variance of the stress-energy tensor decays to zero as the power law
$$:T_{uu}^2:=\frac{A^2}{4\pi ^2\left(A+\left(v\overline{v}\right)u\right)^4}\frac{A^2}{4\pi ^2\left(v\overline{v}\right)^4u^4}.$$
(4.7)
as distinct from the nonextremal black hole, the variance of whose stress-energy tensor decays exponentially:
$$:T_{uu}^2:=\frac{1}{(4\pi )^2}(\frac{\kappa ^4}{48}\frac{4\kappa ^2B^2\mathrm{e}^{2\kappa u}}{\left(v\overline{v}+B\mathrm{e}^{\kappa u}\right)^4})\frac{\kappa ^4}{768\pi ^2}\frac{\kappa ^2B^2\mathrm{e}^{2\kappa u}}{4\pi ^2\left(v\overline{v}\right)^4}.$$
(4.8)
Contrary to the extremal case, the “nonextremal” variance $`\mathrm{\Delta }T_{uu}`$ tends not to zero as $`u+\mathrm{}`$, but to the value $`\kappa ^2\sqrt{2}/(48\pi )`$, which corresponds to thermal emission.
## 5 Conclusions
All these results, including those of , demonstrate that at no time in their history do extremal black holes act as thermal objects . The conclusion would seem to have significant implications in regard to string theory. From the point of view of thermodynamics, if temperature is undefined, entropy is also undefined. Yet string caclulations , which involve only state-counting arguments, indicate that extremal black hole solutions possess the usual Bekenstein-Hawking entropy. The results of semi-classical gravity and string theory would appear, then, to be in direct contradiction. The resolution to this dilemma is far from clear. The calculations discussed in this letter assume $`a=M`$ configurations exist, but ignore back-reaction to the metric (or recoil of the mirror). Evidently, extraordinary fine-tuning early on is required to produce extremal solutions and it may well be that, taking such backreaction into account, the formation of extremal objects is simply disallowed by nature. If true, string-theory calculations refer to impossible objects. On the other hand, if one cannot ignore back-reaction—at least at late times—then all semi-classical calculations to date (including Hawking’s) are incorrect. Presently, about all that can be said is that extremal black holes seem to represent the limits of our current theories of quantum gravity.
Acknowledgements It is a pleasure to thank George Ellis and Jeff Murugan for stimulating conversations and helpful comments on the manuscript.
|
warning/0006/cond-mat0006235.html
|
ar5iv
|
text
|
# Net Charge on a Noble Gas Atom Adsorbed on a Metallic Surface
## I Introduction
The problem of measuring and understanding the binding energy of noble gas atoms on metal substrates has always been of considerable interest . The atomic binding is that energy released when an atom from the vapor sticks to the surface. Several studies have related atomic binding to surface charge distributions. Induced surface electronic dipole moments near the adsorbed atom have been of particular interest. In the work which follows, the intimate relationship between binding energy and induced atomic charge will be considered in theoretical detail.
As the atom is lowered onto the surface and becomes adsorbed, the atomic dipole tends to be oriented with the positive side of the dipole pointing away from the metal. In reality, the negative side of the atomic dipole is better thought to be on average an electron (negative) charge $`Z_{eff}|e|`$ donated to the metal. This leaves the atom with a positive net mean charge $`+Z_{eff}|e|`$. The physical situation is pictured in FIG.1 below. Even in a situation often regarded as physisorption, one does not expect a noble gas atom to remain in perfect charge neutrality. When the atom is adsorbed on the substrate, the negative end of the atomic dipole moment neutralizes the positive end of the image dipole moment. This leaves a mean net positive charge on the atom and a mean negative electronic charge deposited in the metal. From an experimental viewpoint, the positive nature of the mean charge on an adsorbed noble gas atom is equivalent to the donation of a negative charge to the metal. The physical effect is made manifest by the diminution of the electronic work function as the first monolayer of atoms is deposited on the metallic substrate surface. This effect is quite large and is observed for all combinations of gases and metals. The magnitude of the reduction of the work function is proportional to the amount of adsorbate and for coverage up to a monolayer.
In Secs. II and III the long ranged part of Van der Waals interaction between a noble gas atom and a metal will be reviewed. The height of the atom, over and above the metal surface, is considered to be large compared with the atomic size. The interaction is between a quantum fluctuating dipole and its correlated image. Both the binding energy and the probability of electronic excitation will be computed numerically in Sec. IV. In Sec. V, non-perturbative expressions for the binding energy and the probability of electronic excitation are derived. In Sec. VI it is shown that the stronger the binding energy the larger the net atomic adsorptive charges. The physical importance of the adsorptive charge on a noble gas atom is discussed in the concluding Sec. VII.
## II Perturbation Theory
Suppose a Hamiltonian of the form
$$=H+V,$$
(1)
in which
$$H\psi _n=E_n\psi _n,$$
(2)
represents an unperturbed energy eigenvalue problem, and
$$\mathrm{\Psi }_n=_n\mathrm{\Psi }_n,$$
(3)
represents the perturbed energy eigenvalue problem. Without loss of generality, one may assume (for the unperturbed ground state wave function) that
$$(\psi _0,V\psi _0)=0.$$
(4)
Two quantities of importance are the exact ground state energy $`_0`$, and the probability that the interaction $`V`$ introduces an excitation in the unperturbed quantum states
$$P=\underset{n0}{}\left|(\psi _n,\mathrm{\Psi }_0)\right|^2=1\left|(\psi _0,\mathrm{\Psi }_0)\right|^2.$$
(5)
In lowest order perturbation theory, one then finds the energy shift
$$U=\left(_0E_0\right)=\underset{n0}{}\frac{\left|(\psi _n,V\psi _0)\right|^2}{\left(E_nE_0\right)}+\mathrm{},$$
(6)
The ground state wave function in first order perturbation theory is given by
$$\mathrm{\Psi }_0=\psi _0\underset{n0}{}\frac{(\psi _n,V\psi _0)\psi _n}{\left(E_nE_0\right)}+\mathrm{}.$$
(7)
The first order wave equation can be used to compute the excitation probability $`P`$ to second order in the perturbation potential; it is
$$P=\underset{n0}{}\left|\frac{(\psi _n,V\psi _0)}{\left(E_nE_0\right)}\right|^2+\mathrm{}.$$
(8)
Let us apply these ideas to the Casimir effect, i.e. the attractive force on an atom located at height $`h`$ above a metallic substrate. The final energy is well known. However, the excitation probability of the atom will also be calculated.
## III The Van der Waals Force
Let the unperturbed problem consist of an isolated atom and an isolated metal
$$H=H_{atom}+H_{metal}.$$
(9)
With $`n=(j,a)`$ as a double index referring to the atomic state $`j`$ and the metallic substrate state $`a`$, we may write
$$\psi _n=\psi _j^{atom}\psi _a^{metal},$$
(10)
with an unperturbed energy
$$E_n=E_j^{atom}+E_a^{metal}.$$
(11)
Further, let the interaction between the atom and metal be of the dipole form
$$V=\mu 𝐄$$
(12)
where $`\mu `$ represents the atomic electric dipole moment operator, and the electric field E is produced by the metal. The second order perturbation interaction energy between the atom and the metal is usually called the Casimir effect and is given by
$$U_C=\underset{(a,j)(0,0)}{}\frac{\left|\mu _{j0}𝐄_{a0}\right|^2}{\left(\mathrm{}\omega _{j0}+\mathrm{}\omega _{a0}^{}\right)},$$
(13)
where
$$\mu _{j0}=(\psi _j^{atom},\mu \psi _0^{atom}),$$
(14)
and
$$𝐄_{a0}=(\psi _a^{metal},𝐄\psi _0^{metal}).$$
(15)
The atomic Bohr frequencies are defined as
$$\mathrm{}\omega _{j0}=\left(E_j^{atom}E_0^{atom}\right).$$
(16)
and the metallic Bohr frequencies are defined as
$$\mathrm{}\omega _{a0}^{}=\left(E_a^{metal}E_0^{metal}\right).$$
(17)
For an isotropic atom, one may define a ground state dipole moment quantum noise spectral function
$$S(\omega )=\frac{1}{3}\underset{j0}{}\left|(\psi _j^{atom},\mu \psi _0^{atom})\right|^2\delta \left(\omega \omega _{j0}\right),$$
(18)
Similarly, one may define the electric field quantum zero point fluctuations due to the metal
$$S_𝐄(\omega ^{})=\underset{a0}{}\left|(\psi _a^{metal},𝐄\psi _0^{metal})\right|^2\delta \left(\omega ^{}\omega _{a0}^{}\right).$$
(19)
One may now compute the Casimir energy $`U_C`$ in terms of these spectral functions; i.e.
$$U_C=\left(\frac{1}{\mathrm{}}\right)_0^{\mathrm{}}_0^{\mathrm{}}\left(\frac{S(\omega )S_𝐄(\omega ^{}}{\omega +\omega ^{}}\right)𝑑\omega 𝑑\omega ^{}.$$
(20)
Finally, one may find the spectral functions by employing the zero point quantum fluctuation response theorems.
For example, if the atomic polarization response to an external electric field at complex frequency $`\zeta `$ (with $`\mathrm{}m(\zeta )>0`$),
$$\delta \mu =\alpha (\zeta )\delta 𝐄_{ext},$$
(21)
defines the ground state atomic polarizability $`\alpha (\zeta )`$ for a spherical atom, then the fluctuation response theorem theorem asserts
$$S(\omega )=\left(\frac{\mathrm{}}{\pi }\right)\mathrm{}m\alpha (\omega +i0^+).$$
(22)
The fluctuation spectral function for the electrostatic field $`𝐄=\varphi `$ produced by the metal is a bit more subtle.
If an external charge density at complex frequency $`\zeta `$ produces an electrostatic potential according to the rule
$$\delta \varphi (𝐫)=𝒢(𝐫,𝐫^{},\zeta )\delta \rho _{ext}(𝐫^{})d^3𝐫^{},$$
(23)
then the spectral function for zero point electrostatic potential fluctuations
$$S_\varphi (𝐫,𝐫^{},\omega )=$$
$$_{\mathrm{}}^{\mathrm{}}\mathrm{cos}(\omega t)\mathrm{}e0\left|\mathrm{\Delta }\varphi (𝐫,t)\mathrm{\Delta }\varphi (𝐫^{},0)\right|0\left(\frac{dt}{2\pi }\right)$$
(24)
obeys the fluctuation response theorem in the form
$$S_\varphi (𝐫,𝐫^{},\omega )=\left(\frac{\mathrm{}}{\pi }\right)\mathrm{}m𝒢(𝐫,𝐫^{},\omega +i0^+).$$
(25)
For the problem at hand, suppose that the metal is located in the half-space $`z<0`$; e.g. the substrate surface is the $`z=0`$ $`x`$-$`y`$ plane. If both $`𝐫`$ and $`𝐫^{}`$ are in the vacuum (i.e. $`z>0`$ and $`z^{}>0`$), then the “method of images” yields the Greens function
$$𝒢(𝐫,𝐫^{},\zeta )=\left(\frac{1}{|𝐫𝐫^{}|}\right)\left(\frac{\eta (\zeta )}{|𝐫𝐫_i^{}|}\right),$$
(26)
where
$$𝐫^{}=(x^{},y^{},z^{}),$$
(27)
has a corresponding “image” position
$$𝐫_i^{}=(x^{},y^{},z^{}),$$
(28)
and
$$\eta (\zeta )=\left(\frac{\epsilon (\zeta )1}{\epsilon (\zeta )+1)}\right).$$
(29)
The dielectric response function for the metal $`\epsilon (\zeta )`$ determines the conductivity $`\sigma (\zeta )`$ via
$$\epsilon (\zeta )=1+\left(\frac{4\pi i\sigma (\zeta )}{\zeta }\right).$$
(30)
Thus
$$S_\varphi (𝐫,𝐫^{},\omega )=\left(\frac{\mathrm{}}{\pi |𝐫𝐫_i^{}|}\right)\mathrm{}m\eta (\omega +i0^+).$$
(31)
With
$$𝐑=(0,0,h)$$
(32)
denoting the position of the atom at a height $`h`$ above the substrate surface, the electric field zero point fluctuations of the electric field at the atom may be computed via
$$S_𝐄(\omega )=\underset{𝐫𝐑}{lim}\underset{𝐫^{}𝐑}{lim}$$
$$\left(\frac{^2}{xx^{}}+\frac{^2}{yy^{}}+\frac{^2}{zz^{}}\right)S_\varphi (𝐫,𝐫^{},\omega ).$$
(33)
The differentiation is tedious but direct. It yields
$$S_𝐄(\omega )=\left(\frac{\mathrm{}}{2\pi }\right)\left(\frac{\mathrm{}m\eta (\omega +i0^+)}{h^3}\right).$$
(34)
Substituting Eqs.(22) and (34) into Eq.(20) yields the zero-point fluctuation Casimir effect potential
$$U_C=\left(\frac{C_{atom}}{h^3}\right)$$
(35)
where
$$C_{atom}=\left(\frac{\mathrm{}}{2\pi ^2}\right)\times $$
$$_0^{\mathrm{}}_0^{\mathrm{}}\left(\frac{\mathrm{}m\alpha (\omega +i0^+)\mathrm{}m\eta (\omega ^{}+i0^+)}{\omega +\omega ^{}}\right)𝑑\omega 𝑑\omega ^{}.$$
(36)
The Casimir form of the Van der Waals potential between an atom and a conducting surface is well known. The purpose for reviewing the Van der Waals result is that now we may also calculate the probability that the atom at a height $`h`$ is in an excited state.
The excited state probability in Eq.(5), when evaluated to the lowest order perturbation theory in Eq.(8), yields the Casimir excited state probability $`P_C`$. In the Casimir perturbation theory, the probability for the atom to be in an excited state contains an extra energy denominator. The final result is in the simple form
$$P_C=\left(\frac{v_{atom}}{h^3}\right).$$
(37)
The parameter $`v_{atom}`$ is a rough measure of the effective atomic volume; It is
$$v_{atom}=\left(\frac{1}{2\pi ^2}\right)\times $$
$$_0^{\mathrm{}}_0^{\mathrm{}}\left(\frac{\mathrm{}m\alpha (\omega +i0^+)\mathrm{}m\eta (\omega ^{}+i0^+)}{(\omega +\omega ^{})^2}\right)𝑑\omega 𝑑\omega ^{}.$$
(38)
The excited state probability in Eq.(37) is a new result. The noble gas atom above the metal, is not electronically inert. The atom in the vacuum is in the ground state, As the atom is lowered toward the metal surface, the atomic electronic state becomes excited with probability $`P>0`$. For heights $`h`$ such that $`h^3>>v_{atom}`$ we have $`PP_C`$ as in Eq.(37). Numerical results for the probability $`P_C`$ then follow (below) in a similar manner to numerical results for $`U_C`$ which have been previously computed by other workers.
## IV Numerical Evaluations
For the purpose of numerical evaluations of parameters associated with the Van der Waals interaction, we follow Rauber, Klein, Cole and Bruch; They choose a single pole approximation for both $`\alpha (\zeta )`$ and $`\eta (\zeta )`$. The pole in $`\alpha (\zeta )`$ is parameterized by an atomic frequency $`\omega _a`$, while the pole in $`\eta (\zeta )`$ is parameterized by a plasma frequency $`\mathrm{\Omega }_s`$.
In detail, the approximations read
$$\alpha ^{RKCB}(\zeta )\left(\frac{\omega _a^2\alpha _0}{\omega _a^2\zeta ^2}\right),$$
(39)
along with
$$\eta ^{RKCB}(\zeta )\left(\frac{\mathrm{\Omega }_s^2g_0}{\mathrm{\Omega }_s^2\zeta ^2}\right).$$
(40)
Eqs.(36), (38), (39) and (40) yield
$$C_{atom}^{RKCB}=\frac{\mathrm{}}{8}\left\{\frac{\alpha _0g_0\omega _a\mathrm{\Omega }_s}{\omega _a+\mathrm{\Omega }_s}\right\},$$
(41)
and
$$v_{atom}^{RKCB}=\frac{1}{8}\left\{\frac{\alpha _0g_0\omega _a\mathrm{\Omega }_s}{(\omega _a+\mathrm{\Omega }_s)^2}\right\}.$$
(42)
The values of the atomic parameters $`\alpha _0`$ and $`\omega _a`$ have been previously tabulated. and are listed in Table I. The values of $`g_0`$ and $`\mathrm{\Omega }_s`$, (for several metals) have also been previously tabulated and are here listed in Table II. In Table III, we have computed the values of $`C_{atom}`$ and $`v_{atom}`$.
The probability of atomic excitation in the Casimir regime $`P_C=(v_{atom}/h^3)`$ can be calculated in the limit of large $`h`$. When the height of the atom obeys $`h^3>>v_{atom}`$, the perturbation theory is reliable. However, for atoms adsorbed on a submonolayer of film the height $`h`$ is not large. Thus a non-perturbative method is required. We now turn our attention to this more detailed treatment.
## V Rigorous Results
The above considerations are true for atoms above a metal in the perturbative limit $`h\mathrm{}`$. When the atom is adsorbed on the metal surface, a non-perturbative viewpoint must be invoked. For example, consider the ground state matrix element analytic in the upper half complex energy plane $`\mathrm{}mz>0`$,
$$G(z)=(\psi _0,\left\{\frac{1}{z}\right\}\psi _0),$$
(43)
where $`\psi _0`$ is the unperturbed ground state,
$$H\psi _0=E_0\psi _0,$$
(44)
and the binding energy of the adsorbed atom,
$$U=B,B>0,$$
(45)
is determined by the full electronic ground state
$$\mathrm{\Psi }_0=(H+V)\mathrm{\Psi }_0=(E_0+U)\mathrm{\Psi }_0.$$
(46)
One may write Eq.(43) in the exact form
$$G(z)=\left\{\frac{1}{zE_0\mathrm{\Sigma }(z)}\right\},$$
(47)
where the self energy part $`\mathrm{\Sigma }(z)`$ determines the atomic binding energy
$$U=\mathrm{\Sigma }(E_0+U).$$
(48)
$`G(z)`$ has a simple pole at the exact ground state energy $`z_0=E_0+U=E_0B`$. The residue at the pole is the transition probability $`\psi _0\mathrm{\Psi }_0`$; i.e.
$$|(\mathrm{\Psi }_0,\psi _0)|^2=\left(\frac{1}{1\mathrm{\Sigma }^{}(E_0+U)}\right),$$
(49)
where $`\mathrm{\Sigma }^{}(z)=d\mathrm{\Sigma }(z)/dz`$. The exact expression for the self energy part reads
$$\mathrm{\Sigma }(z)=(\psi _0,V\left\{\frac{1}{z^{}}\right\}V\psi _0),$$
(50)
where $`^{}=\widehat{P}\widehat{P}`$ and where $`\widehat{P}`$ projects into the subspace normal to the unperturbed ground state. For example, from Eq.(49) and the definition of $`\widehat{P}`$ one finds the mean value $`P=(\mathrm{\Psi }_0,\widehat{P}\mathrm{\Psi }_0)`$ given by
$$P=1|(\mathrm{\Psi }_0,\psi _0)|^2=\left(\frac{\mathrm{\Sigma }^{}(E_0+U)}{1\mathrm{\Sigma }^{}(E_0+U)}\right).$$
(51)
The details of the mathematical derivation of Eqs.(47)-(51) are given in the appendix.
Suppose that the atom (located on the metal surface) starts out in the unperturbed state $`\psi _0`$. The atom will then decay into its renormalized ground state, giving up its excess binding energy $`U`$ to the bulk electrons in the metal. The transition rate per unit time to deposit an electronic energy $`W`$ to the metal is given by
$$\mathrm{\Gamma }(W)=\left(\frac{2\pi }{\mathrm{}}\right)(\psi _0,V\delta (E_0+W^{})V\psi _0).$$
(52)
From Eqs.(50) and (52) it follows that
$$\mathrm{\Sigma }(z)=\left(\frac{\mathrm{}}{2\pi }\right)_0^{\mathrm{}}\frac{\mathrm{\Gamma }(W)dW}{z(E_0+W)}.$$
(53)
Thus the binding energy $`U=B`$ is rigorously determined by Eqs.(48) and (53) to be
$$B=\left(\frac{\mathrm{}}{2\pi }\right)_0^{\mathrm{}}\frac{\mathrm{\Gamma }(W)dW}{W+B},$$
(54)
while Eqs.(51) and (53) imply
$$P=\left(\frac{\varpi }{1+\varpi }\right),$$
(55)
where
$$\varpi =\left(\frac{\mathrm{}}{2\pi }\right)_0^{\mathrm{}}\frac{\mathrm{\Gamma }(W)dW}{(W+B)^2},$$
(56)
In the perturbative limit,
$$\underset{h\mathrm{}}{lim}(P/P_C)=\underset{h\mathrm{}}{lim}(U/U_C)=1$$
(57)
where the Casimir values $`U_C`$ and $`P_C`$ have been defined, respectively, in Eqs.(35) and (37). For the non-perturbative limit of atomic adsorption for a finite height $`h=h_s`$, the exact results of Eqs.(54), (55) and (56) can be employed for making realistic estimates of the effective charge $`Z_{eff}|e|`$ of the atom.
## VI Net Charge on an Adsorbed Atom
For estimating the effective charge on the adsorbed atom, we note the following: (i) Eq.(54) is an implicit equation for the binding energy $`B=f(B)`$. (ii) If $`\overline{\mathrm{\Gamma }}=lim_{W0}\mathrm{\Gamma }(W)>0`$, then $`f(B0)(\mathrm{}\overline{\mathrm{\Gamma }}/2\pi )ln(const/B)`$, where a constant “cut-off” must be placed in the logarithm. (iii) We employ the following simple dispersion formula for $`\mathrm{\Gamma }(W)`$: Over a bandwidth $`0<W<E_a`$, we consider $`\mathrm{\Gamma }(W)=\overline{\mathrm{\Gamma }}`$ to be uniform. Outside this interval, i.e. for electronic energies above the atomic energy cut-off $`E_a=\mathrm{}\omega _a`$ in Table I, we consider $`\mathrm{\Gamma }(W)`$ to be negligible. Under this assumption, the binding energy Eq.(54) reads
$$B\left(\frac{\mathrm{}\overline{\mathrm{\Gamma }}}{2\pi }\right)ln\left(\frac{\mathrm{}\omega _a}{B}\right),$$
(58)
if $`B<<\mathrm{}\omega _a`$. Similarly, Eq.(56) reads
$$\varpi \left(\frac{\mathrm{}\overline{\mathrm{\Gamma }}}{2\pi B}\right),$$
(59)
The probability of excitation $`P=Z_{eff}`$ determines the effective charge via Eqs.(55), (58) and (59)
$$Z_{eff}=\left(\frac{1}{1+ln(E_a/B)}\right).$$
(60)
A plot of the effective charge versus the binding energy is shown in Fig.2.
In terms of experimental binding energies and mean atomic excitation energies listed Table I, the effective charges can be computed numerically from Eq.(60).
## VII Conclusions
A dispersion theory has been presented for computing the effective charge per noble gas atom adsorbed on metallic substrates. The general physical situation, shown in Fig.1, is complimentary to the charge density functional approach for computing atomic binding. Both approaches yield a picture in which there is an atomic dipole moment. The negative end of the dipole moment is better viewed as the donation of a mean negative electronic charge to the metal below. The positive end of the atomic dipole moment, is what has been called the net atomic charge $`Z_{eff}|e|`$.
The advantage to the dispersion method is that it fits smoothly into the power law limit of the well known Casimir force. In the density functional approach, the “image charges” are added in “by hand”. On the other hand, the charge density functional approach is intuitive on a microscopic level.
The evidence for $`Z_{eff}>0`$ is based on the change in metallic work function due to adding submonolayer adatoms, for thicker films the situation is complex. The work function can increase as the film thickness increases for film layers above the monolayer. This experimental fact has (thus far) no simple theoretical explanation.
APPENDIX
In this appendix, the details of the mathematical derivation of results in Sec. V will be made explicit. One may define two projection operators $`\widehat{Q}`$ and $`\widehat{P}`$ obeying
$$\widehat{Q}+\widehat{P}=1.$$
$`(A1)`$
The operator $`\widehat{Q}`$ projects a wave function onto the unperturbed ground state wave function $`\psi _0`$ and may be written as
$$\widehat{Q}=|\psi _0\psi _0|.$$
$`(A2)`$
If we define for any operator $`\widehat{A}`$, the projected operators $`A_{QQ}=\widehat{Q}\widehat{A}\widehat{Q}`$, $`A_{QP}=\widehat{Q}\widehat{A}\widehat{P}`$, $`A_{PQ}=\widehat{P}\widehat{A}\widehat{Q}`$ and $`A_{PP}=\widehat{P}\widehat{A}\widehat{P}`$, then the operator may be written in a partitioned matrix form
$$\widehat{A}=\left(\begin{array}{cc}A_{QQ}& A_{QP}\\ A_{PQ}& A_{PP}\end{array}\right).$$
$`(A3)`$
In particular, the resolvent operator
$$\widehat{}(z)=\left(\frac{1}{z}\right)$$
$`(A4)`$
may be written as
$$\widehat{}(z)=\left(\begin{array}{cc}_{QQ}(z)& _{QP}(z)\\ _{PQ}(z)& _{PP}(z)\end{array}\right),$$
$`(A5)`$
where
$$G(z)1_{QQ}=_{QQ}(z)=1_{QQ}\underset{n}{}\frac{|(\mathrm{\Psi }_n,\psi _0)|^2}{(z_n)}.$$
$`(A6)`$
From Eq.(A6), it follows that $`G(z)`$ has a ground state pole at $`z_0=_0`$ with a residue given by $`|(\mathrm{\Psi }_n,\psi _0)|^2`$; i.e.
$$G(z)\frac{|(\mathrm{\Psi }_0,\psi _0)|^2}{(zz_0)}\mathrm{as}zz_0.$$
$`(A7)`$
Eqs.(A4) and (A5) imply
$$\left(\begin{array}{cc}z_{QQ}& _{QP}\\ _{PQ}& z_{PP}\end{array}\right)\left(\begin{array}{cc}_{QQ}(z)& _{QP}(z)\\ _{PQ}(z)& _{PP}(z)\end{array}\right)$$
$$=\left(\begin{array}{cc}1_{QQ}& 0\\ 0& 1_{PP}\end{array}\right),$$
$`(A8)`$
from which
$$(z_{QQ})_{QQ}(z)_{QP}_{PQ}(z)=1_{QQ},$$
$`(A9)`$
and
$$_{PQ}_{QQ}(z)+(z_{PP})_{PQ}(z)=0.$$
$`(A10)`$
Eqs.(A9) and (A10) imply
$$\left(z_{QQ}_{QP}\frac{1}{z_{PP}}_{PQ}\right)_{QQ}(z)=1_{QQ}.$$
$`(A11)`$
Finally, if $`=H+V`$, $`(\psi _0,V\psi _0)=0`$ and $`^{}=_{PP}`$, then Eqs.(A6) and (A11) read
$$\left\{zE_0(\psi _0,V\left\{\frac{1}{z^{}}\right\}V\psi _0)\right\}G(z)=1;$$
$`(A12)`$
i.e.
$$G(z)=\left(\frac{1}{zE_0\mathrm{\Sigma }(z)}\right),$$
$`(A13)`$
with a self energy
$$\mathrm{\Sigma }(z)=(\psi _0,V\left\{\frac{1}{z^{}}\right\}V\psi _0),$$
$`(A14)`$
which completes our derivations.
|
warning/0006/nucl-th0006060.html
|
ar5iv
|
text
|
# Gluonic Correlations in Matter
## 1 INTRODUCTION
QCD Low energy theorems (LET) for the gluonic correlation functions have provided insights into the QCD dynamics both in the vacuum and in the nuclear medium. Indeed, the LET for the pseudoscalar gluon correlation function provides a powerful constraint on the contribution of the $`\eta ^{}`$ mass and the vacuum gluon condensate . We recall that a quantitative calculation of glueball masses using QCD sum rule analysis, require the use of LET to provide the necessary subtraction constants needed in the dispersion analysis. Recently, the extension of the LET to finite temperature and density have been derived for the correlation functions of the scalar gluon operators. These theorems provide strong constraints to be satisfied by lattice calculations or any effective model calculations of QCD in matter. Here, we will derive anew the LET for the scalar gluon correlation functions at low density and show the equivalence of our result to that of ref.. Then, we will derive similar LET for the pseudoscalar gluonic correlation functions in two ways : First, by assuming that the self-dual gauge configurations are dominant in vacuum and (dilute) matter; Second, by using a generic heavy-quark expansion. In both cases, we will derive general relations on the density dependence of the correlation function in terms of differential operators. We briefly comment on the relevance of our results to the QCD spectrum.
## 2 LET in vacuum: review
Let us start with the definition of the scalar and pseudo scalar gluonic two point function.
$`S(Q^2)=i{\displaystyle d^4xe^{iqx}T^{}\frac{3\alpha _s}{4\pi }G^2(x)\frac{3\alpha _s}{4\pi }G^2(0)}`$
$`P(Q^2)=i{\displaystyle d^4xe^{iqx}T^{}\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}(x)\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}(0)}.`$ (1)
Here, $`G^2=G_{\mu \nu }^aG_{\mu \nu }^a`$ and $`G\stackrel{~}{G}=G_{\mu \nu }^a\frac{1}{2}ϵ_{\mu \nu \alpha \beta }G_{\alpha \beta }^a`$. The low energy theorem for the scalar gluon operator follows from the formula,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}O=i{\displaystyle d^4xT^{}O(x)g_0^2G^2(0)},`$ (2)
where $`g_0`$ is the bare coupling constant of QCD. Now using renormalization group arguments, one notes that the bare coupling is related to the ultraviolet cut-off $`M_0`$ via,
$`O=\mathrm{const}\left[M_0exp({\displaystyle \frac{8\pi ^2}{bg_0^2}})\right]^d,`$ (3)
with $`b=11\frac{2}{3}N_f`$ to one-loop. Therefore, the left hand side of eq.(2) yields,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}O=d{\displaystyle \frac{32\pi ^2}{b}}O`$ (4)
Substituting $`O=\frac{3\alpha _s}{4\pi }G^2(0)`$ into eq.(2) and using eq.(4) we find the LET for the scalar gluon correlation function.
$`S(Q^2=0)={\displaystyle \frac{18}{b}}{\displaystyle \frac{\alpha _s}{\pi }}G^2`$ (5)
The situation is more subtle for the pseudoscalar gluonic correlation function. Indeed, in the presence of light quarks, the pseudoscalar gluon field is a total differential of the light quark axial-current, i.e. $`\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}=_q_\mu \overline{q}\gamma _\mu \gamma ^5q`$ ($`m_q0`$). Hence,
$`P(Q^2=0)=0.`$ (6)
This is not true in the absence of light quarks (NLQ). The approximate LET for this case was originally derived in ref. by assuming that self-dual field configurations dominate the functional integral over the gauge fields. To show this, consider the QCD functional integral with nonzero $`\theta `$,
$`Z={\displaystyle [dA]exp(iI)}`$ (7)
where
$`I={\displaystyle d^4x\left[\frac{1}{4}G^2\theta \frac{g_0^2}{32\pi ^2}G\stackrel{~}{G}\right]}`$ (8)
Differentiating eq.(7) twice with respect to $`\theta `$, we have,
$`{\displaystyle \frac{^2ϵ}{\theta ^2}}=i{\displaystyle d^4xT^{}\frac{g_0^2}{32\pi ^2}G\stackrel{~}{G}(x),\frac{g_0^2}{32\pi ^2}G\stackrel{~}{G}(0)}`$ (9)
where $`ϵ`$ is the vacuum energy density. To evaluate the left hand side of eq.(9), we note that the energy density is related to the trace of the energy momentum tensor $`\theta _{\mu \nu }`$,
$`ϵ={\displaystyle \frac{1}{4}}\theta _\mu ^\mu ={\displaystyle \frac{\beta (\alpha _s)}{16\alpha _s}}G^2\stackrel{1loop}{}{\displaystyle \frac{b\alpha _s}{32\pi }}G^2.`$ (10)
Now we assume that the self-dual (anti-self dual) field dominates the functional integral. This changes the euclidean action as follows,
$`I_E=i{\displaystyle d^4x\left[\frac{1}{4g_0^2}\overline{G}^2\theta \frac{i}{32\pi ^2}\overline{G}\overline{\stackrel{~}{G}}\right]}i{\displaystyle d^4x\left[\frac{1}{4}(\frac{1}{g_0^2}+\frac{i\theta }{8\pi ^2})\overline{G}^2\right]}`$ (11)
where $`\overline{G}`$ is obtained by redefining the gauge field $`A_\mu \frac{1}{g_0}\overline{A}_\mu `$. Eq.(11) implies that within this approximation, the ultraviolet cutoff $`M_0`$ is related to $`\theta `$ in a physical quantity by a simple replacement of eq.(3),
$`O=\mathrm{const}\left[M_0exp({\displaystyle \frac{8\pi ^2}{bg_0^2}}{\displaystyle \frac{i\theta }{b}})\right]^d,`$ (12)
for small $`\theta `$ <sup>1</sup><sup>1</sup>1We note that for large $`\theta `$, $`2\pi `$ periodicity is recovered through the branch-structure of $`ϵ`$.. With eq.(12) and eq.(10), it is straightforward to calculate the left hand side of eq.(9). We obtain the LET for the pseudoscalar gluonic correlations,
$`P_{NLQ}^{sd}(Q^2=0)=S(Q^2=0)={\displaystyle \frac{18}{b}}{\displaystyle \frac{\alpha _s}{\pi }}G^2,`$ (13)
where the superscript $`sd`$ refers to the fact that it is based on the self-dual assumption detailed above.
We recall that the difference between the case with light quarks and without light quarks is the contribution due to the $`\eta ^{}`$, which adds up to give the required zero, i.e.
$`P(Q^2=0)=P_{NLQ}(Q^2=0)+{\displaystyle \frac{|\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}|\eta ^{}|^2}{m_\eta ^{}^2}}=0.`$ (14)
One should note that similar LET for higher point gluonic correlations can be obtained by taking further derivatives either with respect to $`\theta `$ or $`1/4g_0^2`$.
## 3 LET for the pseudoscalar gluonic current: Heavy Quark Mass expansion
The above derivation for the pseudo scalar current was obtained within the self-dual approximation in the path integral. This approximation can be averted. Indeed, let us consider a regularization scheme with $`\theta 0`$ in the presence of a heavy quark with mass $`m_Q`$. Substituting the pseudoscalar gluon field for the operator $`O`$ in eq.(2), we have,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}{\displaystyle \frac{3\alpha _s}{4\pi }}G\stackrel{~}{G}=i{\displaystyle d^4xT^{}\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}(x)g_0^2G^2(0)}.`$ (15)
Now, we will make use of the heavy quark mass expansion,
$`m_Q\overline{Q}i\gamma _5Q`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{8\pi }}G\stackrel{~}{G}+O\left({\displaystyle \frac{1}{m_Q^2}}\right)`$
$`m_Q\overline{Q}Q`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{12\pi }}G^2+O\left({\displaystyle \frac{1}{m_Q^2}}\right).`$ (16)
Substituting eq.(3) into eq.(15) we obtain,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}6m_Q\overline{Q}i\gamma _5Q=i{\displaystyle d^4xT^{}\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}(x)48\pi ^2m_Q\overline{Q}Q(0)}.`$ (17)
In general, such substitution is not allowed when the right hand side of eq.(17) has an $`e^{iqx}`$ factor as one is calculating the operator product expansion. However, since we are looking at the LET dominated by non-perturbative space-like field configuration, such substitution is valid.
Consider now making a chiral rotation in $`\theta \theta +\frac{\pi }{2}`$. This will only affect the heavy quark mass by $`m_Qm_Qe^{i\gamma _5\pi /2}`$. As a result, the scalar quark condensate turns to the pseudoscalar quark condensate and vice versa. Substituting the heavy quark operators back to the gluon operators, we obtain the following,
$`{\displaystyle \frac{2}{32\pi ^2}}{\displaystyle \frac{d}{d(1/4g_0^2)}}{\displaystyle \frac{\alpha _s}{\pi }}G^2=i{\displaystyle d^4xT^{}\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}(x)\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}(0)}.`$ (18)
Making use of eq.(4), yields
$`P_{NLQ}^{hq}(Q^2=0)={\displaystyle \frac{8}{b}}{\displaystyle \frac{\alpha _s}{\pi }}G^2,`$ (19)
where, the superscript $`hq`$ refers to the heavy quark mass expansion used. This is different from eq.(13) as obtained within the self-dual approximation. Given the uncertainty in the value of the non perturbative gluon condensate, both results are consistent with the vacuum phenomenology in eq.(14).
## 4 LET for the scalar gluonic current in nuclear matter
Here we derive the LET for the scalar gluonic current at low density. The same derivation holds for finite temperature. The starting point is taking the expectation value of eq.(2) at finite density. Then we make a low density expansion of the left hand side,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}O_\rho ={\displaystyle \frac{d}{d(1/4g_0^2)}}\left(O_0+\rho N|O|N+𝒪(\rho )\right).`$ (20)
Here, $`_\rho `$ denotes the expectation value at finite density, $`\rho `$ is the nucleon density and $`|N`$ a nucleon state normalized as $`N||N=(2\pi )^3E_N/m_N\delta ^3(0)`$. This is the linear density approximation. The derivative of the nucleon expectation value is simply obtained by substituting $`dd3`$ in eq.(4). The derivative of the nucleon density can be obtained from the fact that the chemical potential $`\mu =p_f^2/2m_N`$ is an external parameter and is independent of $`g_0`$. That is, using
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}\left({\displaystyle \frac{p_f^2}{2m_N}}\right)=0`$ (21)
and
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}m_N={\displaystyle \frac{32\pi ^2}{b}}m_N,`$ (22)
we find,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}p_f^2=`$ $`{\displaystyle \frac{32\pi ^2}{b}}2m_N\mu ,`$ (23)
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}\rho =`$ $`\rho {\displaystyle \frac{32\pi ^2}{b}}{\displaystyle \frac{3}{2}}.`$ (24)
Therefore,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}O_\rho ={\displaystyle \frac{32\pi ^2}{b}}\left(dO_0+(d{\displaystyle \frac{3}{2}})\rho N|O|N+𝒪(\rho )\right).`$ (25)
From this, it follows that the changes of the correlator for the scalar gluonic current to leading order in density is,
$`\mathrm{\Delta }S(Q^2)=S(Q^2;\rho )S(Q^2;\rho =0)={\displaystyle \frac{45}{4b}}\rho N|{\displaystyle \frac{\alpha _s}{\pi }}G^2|N,`$ (26)
where $`S(Q^2,\rho )`$ is the correlation function defined in eq.(1) calculated at finite density $`\rho `$.
We note that eq.(25) is also consistent with the result of ref. applied to leading order in density. In ref. it was noted that a physical quantity of mass scale $`d`$ has the following dependence on the temperature or density.
$`O_{\rho ,T}=\mathrm{\Lambda }^df({\displaystyle \frac{T}{\mathrm{\Lambda }}},{\displaystyle \frac{\mu }{\mathrm{\Lambda }}}),`$ (27)
where $`_{\rho ,T}`$ is the expectation value at finite density and temperature, and the mass scale $`\mathrm{\Lambda }`$ depends on the coupling as in eq.(3) with $`d=1`$. Then, it follows that the differential operator acting on physical parameter can be replaced as,
$`{\displaystyle \frac{d}{d(1/4g_0^2)}}O_{\rho ,T}={\displaystyle \frac{32\pi ^2}{b}}\left(dT{\displaystyle \frac{}{T}}\mu {\displaystyle \frac{}{\mu }}\right)O_{\rho ,T}.`$ (28)
When we apply this differential operator to our eq.(20) and note that $`\mu \frac{\rho }{\mu }=\frac{3}{2}\rho `$, we recover eq.(25).
## 5 LET for the pseudo-scalar gluonic current in nuclear medium: self-dual approximation
The LET for the pseudo-scalar gluonic currents at low density can be similarly obtained in the self-dual approximation. Here, we consider the case with no $`sea`$ light quarks, stressing the effects of matter on the gauge configurations. The starting point is the generalization of eq.(9) to finite density. The difference compared to the scalar case is in taking the derivative with respect to $`\theta `$ instead of $`g_0^2`$. Again, using the linear density approximation, we have
$`{\displaystyle \frac{d^2}{d\theta ^2}}O_\rho ={\displaystyle \frac{d^2}{d\theta ^2}}\left(O_0+\rho N|O|N+𝒪(\rho )\right).`$ (29)
To evaluate the derivatives with respect to $`\theta `$, we only need the relation in eq.(12) and the fact that $`d\mu /d\theta =0`$. To leading order in density, we have
$`{\displaystyle \frac{d\rho }{d\theta }}`$ $`=`$ $`{\displaystyle \frac{3}{2}}({\displaystyle \frac{i}{b}})\rho `$
$`{\displaystyle \frac{d^2\rho }{d\theta ^2}}`$ $`=`$ $`{\displaystyle \frac{9}{4}}({\displaystyle \frac{i}{b}})^2\rho .`$ (30)
Therefore, we have
$`{\displaystyle \frac{d^2}{d\theta ^2}}O_\rho `$ $`=`$ $`({\displaystyle \frac{i}{b}})^2(d^2O_0+({\displaystyle \frac{9}{4}}+3(d3)+(d3)^2)\rho N|O|N`$ (31)
$`+𝒪(\rho )).`$
Substituting this into eq.(9) gives,
$`\mathrm{\Delta }P_{NLQ}^{sd}(Q^2=0)`$ $`=`$ $`P_{NLQ}^{sd}(Q^2=0;\rho )P_{NLQ}^{sd}(Q^2=0;0)`$ (32)
$`=`$ $`{\displaystyle \frac{225}{32b}}\rho N|{\displaystyle \frac{\alpha _s}{\pi }}G^2|N.`$
We can also derive a general formula for eq.(31) in terms of differential operators involving the derivatives with respect to the temperature and/or chemical potential. Again, we have the $`\theta `$ dependence of a physical parameter through eq.(27) and also eq.(12). Therefore, we have
$`{\displaystyle \frac{d^n}{d\theta ^n}}O_{\rho ,T}=\left({\displaystyle \frac{i}{b}}\right)^n\left(dT{\displaystyle \frac{}{T}}\mu {\displaystyle \frac{}{\mu }}\right)^nO_{\rho ,T},`$ (33)
which is consistent with eq.(31). In the linear density approximation eq.(33) can be written as,
$`{\displaystyle \frac{d^n}{d\theta ^n}}(O_0+\rho N|O|N)=\left({\displaystyle \frac{i}{b}}\right)^n(d^nO_0+(d{\displaystyle \frac{3}{2}})^n\rho N|O|N)).`$ (34)
## 6 LET for the pseudo-scalar gluonic current in nuclear medium: Heavy quark expansion
The generalization of the low energy theorem for the pseudo-scalar current based on the heavy quark expansion can be obtained in the same way as in the scalar case. The derivation provides for a check on the self-duality assumption. Starting from the expectation value of eq.(18) at finite density, and substituting eq.(25) into the left hand side, we obtain,
$`\mathrm{\Delta }P_{NLQ}^{hq}(Q^2=0)`$ $`=`$ $`P_{NLQ}^{hq}(Q^2=0;\rho )P_{NLQ}^{hq}(Q^2=0;0)`$ (35)
$`=`$ $`{\displaystyle \frac{5}{b}}\rho N|{\displaystyle \frac{\alpha _s}{\pi }}G^2|N.`$
Note that the overall factor is different from eq.(32), but the sign is the same. The relation in terms of differential operators can be obtained in the same way as in the scalar case by substituting eq.(28) into the left hand side of eq.(18).
## 7 Summary
We have derived the LET for scalar and pseudoscalar gluonic correlation functions at low density. Although, we have restricted our discussions to two point functions, we can easily generalize this method to higher point functions and to finite temperature. Since eq.(14) is still valid in the medium, we expect,
$`\mathrm{\Delta }\left({\displaystyle \frac{|\frac{3\alpha _s}{4\pi }G\stackrel{~}{G}|\eta ^{}|^2}{m_\eta ^{}^2}}\right)=\mathrm{\Delta }P_{NLQ}(Q^2=0)<0.`$ (36)
The magnitude of the right hand side depends on the approximation scheme we have used (self-duality or heavy quark). However, in both cases, the sign is negative. To leading order in the matter density and if we were to assume that residues do not change, then the $`\eta ^{}`$ mass will increase, while the scalar glueball mass will decrease. These observations are worth testing using QCD lattice simulations. The faith of the $`\eta ^{}`$ mass in matter may have interesting consequences on low mass dilepton emissions, pion thresholds as well as the strong CP problem in matter.
## 8 acknowledgments
We thank KIAS for hospitality during the completion of this work. The work of SHL was supported by the Brain Korea 21 project and by KOSEF grant number 1999-2-111-005-5. The work of IZ was supported in part by US DOE grant DE-FG02-88ER40388.
|
warning/0006/astro-ph0006218.html
|
ar5iv
|
text
|
# Halo Properties in Cosmological Simulations of Self-Interacting Cold Dark Matter
## 1. Introduction
The cold dark matter (CDM) family of cosmological models provides an excellent description of a wide variety of observational results, from the earlier observable epochs detected via microwave background fluctuations to present-day observations of galaxies and large-scale structure. A “concordance model” with roughly one-third matter and two-thirds vacuum energy, either a cosmological constant or quintessence (Caldwell, Dave & Steinhardt (1998)), is consistent with almost all current observations on scales $`1`$ Mpc (Bahcall et al. (1999)).
Recently, improving observations and numerical techniques have enabled a comparison of CDM scenarios to observations on galactic scales of $``$ few kpc. The results have not been encouraging. There are a number of distinct observations that may be in conflict with predictions of CDM:
* The density profile of galaxies in the inner few kiloparsecs appears to be much shallower than predicted by numerical simulations (Navarro, Frenk & White 1996, hereafter NFW ). For density profiles characterized by $`\rho (r)r^\alpha `$ as $`r0`$, CDM predicts $`\alpha 1.5`$ with little scatter (Moore et al. 1999), while current H$`\alpha `$ observations suggest $`\alpha 0.5`$ with significant scatter (Swaters, Madore & Trewhella (2000); Dalcanton & Bernstein (2000), though see van den Bosch & Swaters (2000)).
* The central density of dark matter halos is observed to be $`\rho _c0.02M_{}\mathrm{pc}^3`$ roughly independent of halo mass (Firmani et al. 2000b ), while CDM predicts halos with $`\rho _c1M_{}\mathrm{pc}^3`$ at dwarf galaxy masses, increasing to larger masses (Moore et al. 1999).
* The number of dwarf galaxies in the Local Group is an order of magnitude fewer than predicted by CDM simulations, with the discrepancy growing towards smaller masses (Moore et al. 1999a ; Klypin et al. (1999)).
* Hydrodynamic simulations produce galaxy disks that are too small and have too little angular momentum, yielding a Tully-Fisher relation whose zero-point is off by several magnitudes from observations (Navarro & Steinmetz (2000)).
* The robustness of rapidly rotating bars in high surface brightness spiral galaxies implies lower density cores than predicted by CDM (de Battista & Sellwood (1998)).
* Cluster CL 0024+1654 is nearly spherical with a large, soft core, while CDM typically predicts triaxial clusters with cuspy cores (Tyson, Kochanski & Dell’antonio (1998), though see Miralda-Escudé 2000 for a counterexample).
Each piece of evidence taken individually is perhaps not convincing enough to claim that CDM has failed on galactic scales. For instance, until recently there was controversy amongst simulators regarding inner profiles (Kravtsov et al. (1998)), but more careful simulations have converged on a consistent prediction (Klypin (2000)). Observationally, inner galactic profiles are uncertain due to beam smearing effects in H I observations (Swaters, Madore & Trewhella (2000); van den Bosch et al. (2000)), though samples of high-resolution H$`\alpha `$ observations continue to show shallower profiles than predicted by CDM (Dalcanton & Bernstein (2000)). The number of observed Local Group dwarf galaxies may be reconciled with CDM via plausible scenarios for suppressed galaxy formation (e.g. Bullock, Kravtsov & Weinberg (2000)), or else compact high-velocity clouds could represent the “missing satellites” that are seen in $`N`$-body simulations (Blitz et al. (1999)). Hydrodynamic simulations of disk galaxy formation are fraught with the usual concerns about the effects of feedback, artificial viscosity and resolution, though it appears the discrepancies above are due to the underlying dark matter distribution (Navarro & Steinmetz (2000)).
It becomes more interesting to consider alternatives to conventional CDM when one recognizes that all these discrepancies may be symptomatic of a single cause: Dark matter halos in CDM simulations appear to be more centrally concentrated than observed. Recognizing this, various authors have recently forwarded a plethora of alternative dark matter theories that suppress the central concentration of dark matter in galaxy halos. Among such theories are that the dark matter is warm (Sommer-Larsen & Dolgov (2000); Colín, Avila-Reese & Valenzuela (2000); Hannestad & Scherrer (2000)), repulsive (Goodman (2000)), fluid (Peebles (2000)), fuzzy (Hu, Barkana & Gruzinov (2000)), decaying (Cen (2000)), annihilating (Kaplinghat, Knox & Turner (2000)), and the alternative we investigate here, self-interacting (SIDM; Spergel & Steinhardt 2000). Interestingly, all theories may be tuned to solve the problems mentioned above (at least in analytic approximations), all theories may be motivated from particle physics considerations, and all theories retain the desirable properties of CDM on extragalactic scales (though warm dark matter is non-trivially constrained by this requirement; see Narayanan et al. (2000)).
SIDM is governed by a single free parameter, the cross section per unit mass $`\sigma _{\mathrm{DM}}`$ of the interacting dark matter particle. Spergel & Steinhardt suggested $`\sigma _{\mathrm{DM}}10^{22}10^{25}\mathrm{cm}^2\mathrm{GeV}^1`$ in order to reduce the central concentration of galaxy halos by a sufficient amount to alleviate the above problems. Intriguingly, this value is close to the cross section of ordinary hadrons, motivating some particular particle physics candidates for SIDM (Steinhardt et al., in preparation). If $`\sigma _{\mathrm{DM}}`$ is significantly smaller than this range, then the optical depth at galactic densities is much less than unity, implying that SIDM would have a negligible effect on the dark matter distribution in halos.
A qualitative picture of the evolution of an SIDM halo is as follows: At early times there is no difference between SIDM and CDM since the densities and peculiar velocities are sufficiently low that collisions are rare; hence SIDM makes identical predictions to CDM regarding cosmic microwave background fluctuations and the Lyman alpha forest. As the halo forms and grows via gravitational instability, the central density increases. Eventually, collisions are so frequent that dark matter particles scatter out of the center as fast as they are accreted, and the density growth is halted, forming a core. Such a limit is not present in the CDM model, where the central density grows unchecked. The SIDM core then begins to extend while retaining constant central density. Heat transfer from the outer parts of the halo raises the temperature in the halo core. If the halo is truly isolated, then eventually the core thermalizes with the exterior resulting in an isothermal halo with a steep density profile. This initiates gravothermal collapse, where the direction of heat transfer is reversed and the exterior begins to cool the halo center. However, in a realistic cosmological setting, galaxies constantly accrete material, keeping the outer halo hot and heat flowing inwards, thus delaying core collapse. The interplay between collisional heat transfer and accretion determines whether a halo will undergo core collapse in a Hubble time.
It is important to appreciate that the transport behavior does not change monotonically with $`\sigma _{\mathrm{DM}}`$. For small cross-sections, heat transfer increases with $`\sigma _{\mathrm{DM}}`$ since the frequency of collisions increases; however, for large cross-sections, the conductivity $`\kappa \sigma _{\mathrm{DM}}^10`$ and no heat transfer occurs. Thus, as we discuss in §6, the fluid approximation is a poor decription of SIDM in the moderate cross-section regime proposed by Spergel and Steinhardt. Furthermore, the behavior in the moderate cross-section regime cannot be surmised by interpolating between the fluid and the non-interacting CDM regimes. A proper treatment of the SIDM proposal, which includes the interplay of accretion and heat transfer, its non-monotonic dependence $`\sigma _{\mathrm{DM}}`$, and the effects of merging demands numerical simulations designed to explore the moderate cross-section regime.
In this paper we investigate the statistical properties of halos in SIDM and CDM in cosmological $`N`$-body simulations. Our spatial and mass resolutions are sufficient to probe the inner regions ($`1h^1\mathrm{kpc}`$) of small halos ($``$ few$`\times 10^9M_{}`$), while maintaining sufficient volume so as to have a significant sample of such halos. We use a Monte Carlo technique similar to Kochanek & White (2000) to model collisions. The primary difference between our simulations and prior investigations (discussed in more detail in §6) is that we model a cosmologically-significant random volume of the universe with self-interaction cross sections in the range favored by Spergel & Steinhardt , enabling us to characterize the statistical properties of halos as we vary $`\sigma _{\mathrm{DM}}`$.
§2 describes our initial conditions and simulation techniques using a Monte Carlo $`N`$-body approach. In §3 we compare the structural properties of halos in CDM versus SIDM models with several cross sections. In particular we examine their central densities, inner profile slopes $`\alpha `$, the mass dependence of $`\alpha `$, concentrations, phase space densities, and ellipticities, and where possible compare to observations. In §4 we use lower resolution simulations to test the effects of finite particle numbers in our Monte Carlo method. In §5 we examine the subhalo population around the largest halo in our simulations. In §6 we compare our findings to the simulations of other groups who have conducted numerical studies of self-interacting dark matter, and examine results from a wider range of $`\sigma _{\mathrm{DM}}`$. We summarize our results and discuss observational constraints in §7. We find that SIDM with $`\sigma _{\mathrm{DM}}10^{23}10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ produces halos that are in better agreement than collisionless CDM for a wide variety of observations.
## 2. Simulating SIDM
### 2.1. Code and Cosmology
We use a modified version of GADGET (Springel, Yoshida & White (2000)), a publically-available TreeSPH code for distributed-memory parallel machines. Here we only employ the gravitational $`N`$-body portion. We evolve a $`4h^1\mathrm{Mpc}`$ randomly-chosen volume of a $`\mathrm{\Lambda }`$CDM universe with $`\mathrm{\Omega }=0.3`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$, $`H_0=70\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$, and $`\sigma _8=0.8`$, similar to the “concordance model” in agreement with a wide variety of observations (Bahcall et al. (1999)). We generate initial conditions using COSMICS (Ma & Bertschinger (1995)) at $`z=49.7`$, where our particle distribution first becomes nonlinear, and evolve to $`z=0`$. We employ $`128^3`$ dark matter particles in each run, resulting in a dark matter particle mass of $`m_p=3.6\times 10^6M_{}`$, and a spline kernel softening of $`ϵ=1h^1\mathrm{kpc}`$ (i.e. force is Newtonian at $`2ϵ`$). To test resolution effects, we also run a suite of simulations with $`64^3`$ particles and $`ϵ=2h^1\mathrm{kpc}`$. Their initial conditions have an initial density field identical to the $`128^3`$ runs, constructed by sampling at alternate grid points.
While our $`4h^1\mathrm{Mpc}`$ box is small, well below the nonlinear scale at $`z=0`$, we are interested here in the behavior on scales of a few kpcs, and it is unlikely that the missing large-scale power would have a significant effect on the inner portions of halos. In addition, our primarily conclusions are based on a comparative study between collisionless and collisional dark matter for individual halos, so we expect these results to be robust to volume effects.
### 2.2. Modeling self-interactions
We have modified GADGET to include self-interactions using a Monte Carlo $`N`$-body technique to probabilistically incorporate collisions, along the same lines as Burkert (2000) and Kochanek & White , closer to the latter as we use $`\mathrm{\Delta }𝐯`$ from individual particles colliding rather than setting $`\mathrm{\Delta }𝐯`$ to be the particle’s velocity; see the discussion in Kochanek & White . Each pair of particles with positions and velocities $`(𝐫_1,𝐯_1)`$ and $`(𝐫_2,𝐯_2)`$, separated by $`\delta x|𝐫_1𝐫_2|/(2ϵ)`$ and $`\delta v|𝐯_1𝐯_2|`$, interact with a probability given by
$$P=f_{\mathrm{geom}}(\delta x)\frac{\delta v\mathrm{\Delta }t}{\lambda _X},$$
(1)
where $`\mathrm{\Delta }t`$ is the timestep,
$$\lambda _X=\frac{4\pi (2ϵ)^3}{3m_p}\frac{1}{\sigma _{\mathrm{DM}}},$$
(2)
and
$$f_{\mathrm{geom}}(\delta x)=N\frac{_0^1W(\delta x)W(\delta x+\delta x^{})d(\delta x^{})}{_0^1W^2(\delta x^{})d(\delta x^{})},$$
(3)
where $`W`$ is the cubic spline kernel used in GADGET. This geometrical factor weights the probability of interaction by the product of spline kernel-weighted density distributions of the two particles at their given separation. The normalization $`N`$ is set by requiring that
$$_0^1f_{\mathrm{geom}}(\delta x)\mathrm{\hspace{0.33em}4}\pi \delta x^2d(\delta x)=1,$$
(4)
which ensures that when a particle has interacted with all its neighbors within $`2ϵ`$, the resulting probability is equivalent to
$$P=\sigma _{\mathrm{DM}}\rho \delta v\mathrm{\Delta }t,$$
(5)
where $`\rho `$ is the local dark matter density.
In our code, the scatterings are performed between individual particles at the time that the acceleration between those particles is being computed (i.e. during the “treewalk”). In order to ensure that all possible scatterings are considered, a tree cell is opened whenever it is within $`2ϵ`$ of a particle, regardless of the opening criterion.
If two particles scatter, their velocities are randomly re-oriented, keeping the magnitudes of their velocities fixed. In practice, a running sum is kept of the change in velocity due to the interactions that a given particle undergoes on every processor, and at the end of the step the velocity change for each particle is summed over all processors and added to that particle’s velocity. In this way, energy and momentum are explicitly conserved, even if the scattered particles are on different processors, or a particle undergoes more than one scatter in a single timestep (which is very rare for the cross sections considered here).
We consider $`\sigma _{\mathrm{DM}}=0`$ (collisionless), $`10^{24}`$, and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$. We also examined $`\sigma _{\mathrm{DM}}=10^{25}\mathrm{cm}^2\mathrm{GeV}^1`$ and $`\sigma _{\mathrm{DM}}=10^{22}\mathrm{cm}^2\mathrm{GeV}^1`$ in a $`64^3`$ simulation, which we will examine in §6. The total number of collisions per particle in our simulations are 1.01 for $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, and 6.05 for $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, with slightly lower numbers (0.9 and 5.3) for the $`64^3`$ runs. Note that a factor of ten increase in $`\sigma _{\mathrm{DM}}`$ translates only to a factor of six increase in the number of collisions, since the lowered central densities (§3.2) partially compensates for the increase in $`\sigma _{\mathrm{DM}}`$. All runs were performed on Fluffy, a 32-processor Beowulf-class machine at Princeton, with each $`128^3`$ run taking approximately one week.
### 2.3. The Simulated Halo Sample
We identify dark matter halos using SKID <sup>1</sup><sup>1</sup>1http://www-hpcc.astro.washington.edu/tools/SKID/ (Spline Kernel Interpolative DENMAX; see Katz, Weinberg & Hernquist (1996)), with a linking length of $`2ϵ`$. We only consider halos containing 64 or more particles, to ensure a roughly complete sample of such halos in our simulations (Weinberg et al. (1999)). Table 1 lists the number of halos for identified in these simulations.
A specific resolution issue arises from the finite number of particles used to probabalistically model collisions in the SIDM simulations: The number of particles in a given halo must be high enough to properly Monte Carlo sample the distribution. As we will show in §4, halos with $`1000`$ particles at $`z=0`$ seem to be accurately represented with this technique for the simulations considered here. This is quite restrictive, but still permits a significant sample of halos (roughly 30 in each $`128^3`$ run) with which to compute statistics. We also use the full sample of halos to examine certain aspects, but we will be cautious about interpretations made from halos below this “Monte Carlo resolution limit”.
## 3. Halo Structure
### 3.1. Halo Profiles
We determine halo profiles $`\rho (r)`$ by spherical averages over radii $`r=ϵ30ϵ`$, in 20 equal intervals of $`\mathrm{log}r`$. A sample of 16 halo profiles from our $`128^3`$ simulations is shown in Figure 1. Each panel shows a halo profile for $`\sigma _{\mathrm{DM}}=0`$ (solid), $`10^{24}`$ (short dashed), and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (long dashed). Dotted line segments emanating from the innermost radius point of the $`\sigma _{\mathrm{DM}}=0`$ curve indicate slopes $`\alpha =0.5,1`$ and $`1.5`$ for comparison. The same corresponding halos are chosen from each simulation, allowing a case-by-case comparison of the effect of SIDM. The outer halo profiles ($`r10h^1\mathrm{kpc}`$) are virtually identical for each halo, showing that the effects of self-interactions are limited to the inner few kpcs of halos, and confirming that the same halos are being compared in the different simulations. The halos in the leftmost column are the four most massive ones in our simulations, while the halos in other columns are chosen randomly from a descending range of masses. Note that rightmost column shows halos with roughly 300 particles each, below our Monte Carlo resolution requirement of $`1000`$ particles (i.e. $`3.6\times 10^9M_{}`$), thus the effects of self-interactions are not necessarily accurately represented in these cases.
From Figure 1 it is immediately evident that SIDM produces halos that have enlarged central cores and shallower inner profiles. CDM halos are almost all cuspy ($`\alpha 1`$ typically), while most $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ cores are close to flat. $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ leads to profiles that are intermediate between these two. In some cases, non-cuspy CDM halos are seen, especially at lower masses. In these cases the halo may have undergone recent merging activity that temporarily lowers the central density, which is particularly effective in smaller mass halos. Additionally, recent mergers that have not relaxed make it difficult to unambiguously identify the halo center about which to compute profiles, typically making profiles appear shallower. We make no cut in regards to the merging history or “isolatedness” of halos, but we do note that the missing large-scale power in our simulations will tend to generate fewer mergers, and make the largest objects in our simulations appear more isolated.
Figure 1: 16 selected halo profiles for collisionless (solid), $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ (dashed), and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (dot-dashed). Corresponding halos from each simulation are presented, allowing a direct comparison of the effect of SIDM on a halo-by-halo basis. The total halo mass in $`M_{}`$ for the $`\sigma _{\mathrm{DM}}=0`$ halo is shown in the lower right; the SIDM halo masses are typically within 20%. The columns are ordered by mass, with the four highest mass halos shown in the leftmost column. Dotted lines from innermost point show reference slopes of $`\alpha =0.5,1.0,1.5`$.
At high masses, the effect of SIDM is very prominent. The upper leftmost halo is Milky Way-sized ($`6\times 10^{11}M_{}`$), and shows a large core of $`15h^1\mathrm{kpc}`$ for $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ ($`8h^1\mathrm{kpc}`$ for $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$). The core size and difference in inner slope become less prominent to lower masses, though this could be due to the increasing effects of unrelaxed halos, as well as the Monte Carlo resolution issues discussed earlier. We examine these issues quantitatively in §3.4.
No evidence is seen for SIDM halo profiles that are isothermal, as would be expected if the cross section was so large that core collapse would occur on timescales significantly shorter than a Hubble time. This supports the analytic estimates of Spergel & Steinhardt that core collapse on a Hubble time would not occur until $`\sigma _{\mathrm{DM}}10^{22}\mathrm{cm}^2\mathrm{GeV}^1`$.
Overall, there is a clear trend on a case-by-case basis that SIDM results in a reduced central density and shallower inner slope of the dark matter halo, with increasing $`\sigma _{\mathrm{DM}}`$ having a greater such effect.
### 3.2. Central Densities
Figure 2 shows the central density of dark matter halos $`\rho _c`$, taken to be the density at our innermost resolved radius $`ϵ`$, as a function of halo mass. Here we only consider halos with more than 1000 particles, where our Monte Carlo technique has sufficient numbers to represent the collisional behavior (as we will discuss in §4). The central halo density of galaxies is observed to be $`0.02M_{}\mathrm{pc}^3`$ (Firmani et al. 2000b ), and is consistent with being independent of halo mass. The observed range of halo densities is shown as the hatched region, with a majority of their data falling towards the lower end of that region. The arrow in the upper left indicates the increase in $`\rho _c`$ projecting the profile from $`1h^1\mathrm{kpc}`$ in to 500 pc, typical of observations of dwarf and low surface brightness galaxy central densities, using the slope shown.
SIDM $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ halos are in good agreement with these observations, while $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ produces inner densities that are a few times higher, but still marginally consistent with observations. In addition, $`\rho _c`$ in SIDM models show little trend with halo mass, in agreement with observations, because the core density is set by collisional physics.
Figure 2: $`\rho _c`$ vs. $`M_{\mathrm{halo}}`$ for CDM (left panel) and SIDM (right panel). For SIDM, crosses show $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, while open squares show $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$. Only halos with 1000 or more particles are shown. The hatched region indicates the range of observed $`\rho _c`$ compiled by Firmani et al. (2000a). Arrow in upper left indicates how much each value of $`\rho _c`$ would increase if measured at 500 pc (instead of $`1h^1\mathrm{kpc}`$), typical of observations, assuming a profile with the slope shown.
Conversely, the more massive halos in CDM have central densities that are too high by at least an order of magnitude already at $`1h^1\mathrm{kpc}`$, and because of their cuspy profile the disagreement would be much worse at smaller radii, as indicated by the arrow in the upper left. Moreover, CDM halos have central densities that increase with mass, in conflict with observations.
As such, SIDM halos appear to agree better with observations. Table 1 lists the median central density of halos with more than 1000 particles in our various models. This shows that our simulations reproduce the observed central halo densities for $`\sigma _{\mathrm{DM}}10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$.
### 3.3. Inner Profile Slopes
We estimate the inner halo profile slope $`\alpha `$ as the slope between the innermost resolved radii, $`r=11.5h^1\mathrm{kpc}`$. Figure 3 shows a histogram of this slope for the collisionless (solid line), $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ (dashed), and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (dot-dashed) cases, for all halos that have more than 1000 particles. The number of such halos in each simulation is indicated in the legend. The qualitative impression from Figure 1 that SIDM produces shallower inner profiles is quantified in Figure 3. The median values of $`\alpha `$ are indicated by the arrows from the upper $`x`$-axis, and are listed in Table 1.
CDM produces halos that have cuspy cores, with $`\alpha _{\mathrm{med}}1.5`$. This is consistent with the work of Moore et al. , among others. 25 of the 28 CDM halos have $`\alpha <1`$, indicating that cuspy cores are a common feature of CDM models. Conversely, the inner slopes in SIDM models are significantly shallower. For $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, $`\alpha _{\mathrm{med}}0.9`$, while for $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, $`\alpha _{\mathrm{med}}0.4`$, with no halos having $`\alpha <1`$. This latter case has a median $`\alpha `$ close to the value preliminarily suggested by H$`\alpha `$ observations of low surface brightness galaxies (Dalcanton & Bernstein (2000)), though a definitive value awaits a more thorough analysis of observational biases.
Figure 3: Histogram of inner slopes $`\alpha `$ for collisionless (solid), $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ (dashed), and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (dot-dashed). Only halos with $`M>3.6\times 10^9M_{}`$ ($`>1000`$ particles) are included; the number of such halos is indicated in parantheses in the legend. The median values of $`\alpha `$ are indicated by the arrows from the top edge of the plot, and are listed in Table 1.
The scatter in $`\alpha `$ is mostly real. There is some scatter due to discreteness effects in measuring the inner slope as we do. While we could obtain an inner slope from fitting a general halo shape (Hernquist (1990); Klypin (2000)), with 5 free parameters the inner slope would be poorly constrained by 20 correlated data points, thus we choose our simpler definition. Further scatter arises from recently merged halos that temporarily have shallower profiles until relaxed. However, neither of these effects is very significant for the large mass halos plotted in Figure 3. Still, we choose to quote the median $`\alpha `$ rather than the mean, in order to quantify “typical” halos in these models and reduce sensitivity to outliers, although the mean is similar. We note that a significant scatter in inner slopes is also seen in the observations (e.g. de Blok, McGaugh & van der Hulst (1996); Dalcanton & Bernstein (2000)).
Our CDM profiles are, at face value, in better agreement with the analytic profile of Moore et al. (1999), with an asymptotic slope of $`\alpha =1.5`$, rather than an NFW profile having $`\alpha (r0)=1`$. However, profile fitting is a tricky business (as discussed in Klypin (2000)). By reducing the scale radius in the NFW profile (i.e. increasing the concentration), one can push the transition to a slope of $`\alpha =1`$ to a radius smaller than $`1h^1\mathrm{kpc}`$ where we cannot resolve the profiles (the “cusp-core degeneracy”; see van den Bosch & Swaters (2000)). Thus we suspect that our CDM profiles can also be adequately fit by an NFW profile having a large concentration parameter. As such, we do not argue for or against either profile form. Our simulations can only predict the slope at $`r1h^1\mathrm{kpc}`$, and that is what should be compared to observations.
SIDM appears to be in better agreement with observations of the inner slopes of dark halo profiles than CDM. At face value, $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ is preferred, but given uncertainties in observations and simulation techniques, $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ is probably also consistent. A similar value of $`\sigma _{\mathrm{DM}}`$ also reproduces the observed central density of galaxies. Such a coincidence is not expected a priori, and may represent a significant success of the SIDM scenario.
### 3.4. Mass Dependence of Inner Slope
Figure 4 shows a plot of inner halo slope $`\alpha `$ vs. halo mass $`M_{\mathrm{halo}}`$ for all halos in our CDM (left panel) and SIDM (right panel, $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$) simulations. The curve shows the running median value of $`\alpha `$ in bins of $`\mathrm{\Delta }(\mathrm{log}M)=0.5`$.
Figure 4: $`\alpha `$ vs. $`M_{\mathrm{halo}}`$ for CDM (left panel) and SIDM ($`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$; right panel). All halos with 64 or more particles are shown. The line shows a running median of the $`\alpha `$ distribution, with a variance computed from all halos within each mass bin.
The CDM case shows almost no trend with mass, with the median slope always around $`1.21.5`$. The largest halo does have a slightly shallower slope, consistent with the trend seen in very high resolution CDM simulations of individual halos (Jing & Suto (2000)). The scatter increases to low mass due to discreteness and merging effects described in §3.3.
The SIDM case shows dramatically different behavior, suggesting at face value that smaller halos have steeper inner profiles. However, caution must be used in interpreting this result. First, smaller halos have smaller cores, meaning that a slope measured at a fixed radius (not scaled to the halo core size) will result in a steeper slope. For CDM this effect is less significant, since the slope remains similar from the outer to the inner halo. Second, the small numbers of particles in the low-mass halos makes the Monte Carlo technique less effective in modeling collisions, thereby making SIDM appear more like CDM; we investigate this issue further in §4. Thus we make no claim regarding a trend of $`\alpha `$ with $`M_{\mathrm{halo}}`$.
### 3.5. Mass Concentration Parameter
As seen in Figure 1, SIDM appears to have the desired effect of reducing the concentration of dark matter halos. In this section we quantify this effect using a concentration parameter, which we define differently than previous authors in order to facilitate a more direct comparison with observations.
The canonical definition of a concentration parameter is given by NFW as the ratio between the virial radius $`r_{200}`$ (taken to be the radius at which the halo density is $`200\times `$ the cosmic mean) and the scale radius of the halo $`r_s`$ in the NFW profile. This concentration parameter, however, is difficult to compute unambiguously in the case of non-isolated halos, and difficult to compare directly to observations that seldom extend out to $`r_{200}`$. Furthermore, $`r_s`$ is only defined within the context of the specific NFW model, and profile fits are typically degenerate between $`r_s`$ and concentration (Klypin (2000)). Colín et al. (2000) circumvent some of these issues by defining the concentration as the ratio of the minimum of $`r_{200}`$ and the halo radius to the radius that encloses 20% of the halo mass. However, this inner radius is dependent on knowing the total halo mass, something which is difficult to determine observationally.
Instead, we choose to define a mass concentration parameter $`c_M`$, based on enclosed mass rather than radii, and restrict the scales in our definition to those where observations are available, typically $`r20h^1\mathrm{kpc}`$. We define
$$c_M=27\frac{M(<r_{\mathrm{in}})}{M(<r_{\mathrm{out}})},$$
(6)
with
$$r_{\mathrm{in}}=\frac{1}{3}r_{\mathrm{out}}=8.5\mathrm{kpc}\frac{v_{\mathrm{circ}}}{220\mathrm{km}\mathrm{s}^1},$$
(7)
where $`v_{\mathrm{circ}}`$ is the circular velocity of the halo. The choice of $`r_{\mathrm{in}}`$ is arbitrary; here we base it on the Milky Way, as it is convenient and results in observationally accessible scales. The scaling with $`v_{\mathrm{circ}}`$ is that expected for self-similar halos following the Tully-Fisher relation. The normalization factor of 27 results in a uniform density distribution having a mass concentration of unity. A flat rotation curve between $`r_{\mathrm{in}}`$ and $`r_{\mathrm{out}}`$ implies $`M(r)r`$, resulting in $`c_M=9`$. In our simulations, we take $`v_{\mathrm{circ}}`$ to be the maximum circular velocity of the halo as output by SKID.
Figure 5: Histogram of mass concentrations $`c_M`$ for collisionless (solid), $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ (dashed), and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (dot-dashed). Only halos with $`M>3.6\times 10^9M_{}`$ ($`>1000`$ particles) are included. The median values of $`c_M`$ are indicated by the arrows from the top edge of the plot, and are listed in Table 1.
Figure 5 shows a histogram of $`c_M`$ for all halos with more than 1000 particles. As expected, there is a clear trend for CDM to have more concentrated halos than SIDM, with the amount of concentration decreasing with increasing $`\sigma _{\mathrm{DM}}`$. Note that the difference between SIDM with $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ is exaggerated relative to the difference between the inner slopes of those models (cf. Figure 3). This is because the concentration is increased in $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ relative to $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ both due to the increased inner slope, as well as the reduced core radius. We also examined the mass dependence of $`c_M`$ and found no obvious trend, but our range of masses is small.
For comparative purposes, our mass concentration parameter $`c_M`$ may be analytically related to the NFW concentration parameter $`c_{\mathrm{NFW}}`$. From NFW ,
$$\frac{v_\mathrm{c}^2(x)}{v_{200}^2}=\frac{1}{x}\frac{\mathrm{ln}(1+c_{\mathrm{NFW}}x)c_{\mathrm{NFW}}x/(1+c_{\mathrm{NFW}}x)}{\mathrm{ln}(1+c_{\mathrm{NFW}})c_{\mathrm{NFW}}/(1+c_{\mathrm{NFW}})},$$
(8)
where $`x=r/r_{200}`$, $`v_\mathrm{c}(x)`$ is the circular velocity at $`x`$, and $`v_{200}`$ is the circular velocity at $`r_{200}`$. Our $`v_{\mathrm{circ}}`$ is taken to be the maximum halo circular velocity, which may obtained by maximizing equation 8; this occurs at $`x_{\mathrm{max}}2/c_{\mathrm{NFW}}`$ (though we compute it exactly for the results shown below).
Let $`\widehat{v}_{\mathrm{circ}}v_\mathrm{c}(x_{\mathrm{max}})/v_{200}`$. In appropriate units, $`v_{200}=r_{200}`$ (see NFW , equation A2). Thus
$$x_{\mathrm{in}}r_{\mathrm{in}}/r_{200}=\frac{8.5}{220}\widehat{v}_{\mathrm{circ}},$$
(9)
implying $`x_{\mathrm{in}}`$ and $`x_{\mathrm{out}}=3x_{\mathrm{in}}`$ are solely functions of $`c_{\mathrm{NFW}}`$ (note that this arises because we defined $`r_{\mathrm{in}}v_{\mathrm{circ}}`$). Using $`M(<r)rv_\mathrm{c}^2(r)`$,
$$c_M=9\frac{v_c^2(x_{\mathrm{in}})}{v_c^2(x_{\mathrm{out}})},$$
(10)
which is purely a function of $`c_{\mathrm{NFW}}`$. The resulting relationship is shown in Figure 6.
Figure 6: Mass concentration parameter $`c_M`$, defined in equation 6, vs. NFW concentration parameter $`c_{\mathrm{NFW}}`$.
Figure 6 shows that $`c_M8`$, typical of halos in our CDM model, corresponds to $`c_{\mathrm{NFW}}23`$. This value is in agreement with expectations for dwarf galaxies in a $`\mathrm{\Lambda }`$CDM model. Conversely, $`c_M5.6`$, which is the median value for SIDM with $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, corresponds to $`c_{\mathrm{NFW}}11`$. Note that the minimum value of $`c_M`$ for an NFW halo is 3. Thus SIDM with $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, with $`c_{M,\mathrm{med}}2.6`$, produces halos that typically cannot be described properly by $`c_{\mathrm{NFW}}`$. This is because these halos have $`\alpha >1`$ typically, so NFW profiles with $`\alpha (r0)=1`$ are a poor fit. This further illustrates why $`c_{\mathrm{NFW}}`$ is a poor way to describe halos in general.
The largest halo in our simulations has a mass comparable to the Milky Way’s, $`6\times 10^{11}M_{}`$. The concentrations of this halo are 6.7, 4.4 and 2.3 in CDM, $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ and $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, respectively. $`c_M`$ of the Milky Way halo is somewhat uncertain, because of the uncertainty in the rotation curve outside the solar circle ($`R_{}`$) and the effect of baryonic mass within $`R_{}`$, but we make a rough estimate here. If the rotation curve is flat, $`c_M=9`$ as stated before. There are suggestions that the rotation curve rises somewhat beyond the solar circle (though this is uncertain; see Olling & Merrifield (2000)), in which case $`c_M`$ is reduced; let us take $`c_M=8`$ as a working estimate. The rotation curve measures the total mass, so we must correct for the baryons to compare with our simulated $`c_M`$. If we take the fraction of baryonic mass to be 50% inside $`R_{}`$ and 20% inside $`3R_{}`$, then $`c_M`$ reduces to 5. In addition, baryons adiabatically compress the dark matter as they dissipate, so we must correct the Milky Way $`c_M`$ further downwards to compare to our dissipationless halos. From the analysis of Avila-Reese, Firmani & Hernandez (1998), this reduction factor is $`1.52`$, resulting in the Milky Way halo having $`c_M3`$. Thus after reasonable corrections, the Milky Way mass concentration appears to be in better agreement with SIDM than CDM. The rapid rotation of bars also suggests a lower concentration for the Milky Way-sized galaxies than that predicted by CDM (de Battista & Sellwood (1998)).
A more direct comparison with simulations may be obtained from rotation curves of dark matter-dominated, low surface brightness galaxies, where baryonic corrections are smaller. We expect that this mass concentration measure $`c_M`$ will be relatively straightforward to compute from such rotation curves (e.g. Dalcanton & Bernstein (2000)), so we look forward to comparisons. $`c_M`$ has the advantage that it is independent of halo fitting parameters, as the enclosed mass can be obtained directly from the observed circular velocity with modest assumptions. In this sense, it is a more robust comparison than the inner slope and the NFW concentration parameter, which are degenerate and sensitive to scales outside those typically observed (van den Bosch & Swaters (2000)), and the central density, which depends on an uncertain contribution from baryons.
### 3.6. Phase Space Densities
A recently popularized measure of the concentration of dark matter halos is the central phase space density. Dalcanton & Hogan (2000) find that observed phase space densities $`Q\rho /\sigma ^3`$ scale as $`Q\sigma ^3\sigma ^4`$, where $`\sigma `$ is the velocity dispersion, from dwarf spheroidals up to clusters of galaxies. Observations compiled by Sellwood (2000) suggest a similar relation, albeit with a large scatter, and he uses them to argue against any form of collisionless dark matter (though see Madsen (2000)).
Figure 7 shows the phase space density $`Q`$ of dark matter within $`r_{\mathrm{in}}`$ as a function of $`\sigma `$, for halos with $`\sigma >30\mathrm{km}\mathrm{s}^1`$. We calculate $`\sigma `$ as the velocity dispersion around the group center of mass velocity, within $`r_{\mathrm{in}}`$ (cf. equation 7). Open circles show CDM halos, crosses indicate SIDM ($`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$) halos. The dashed and dotted lines show $`Q\sigma ^3`$ and $`Q\sigma ^4`$, respectively, that bracket the observations, reproduced from Dalcanton & Hogan (2000).
Figure 7: Phase space density $`Q`$ vs. velocity dispersion $`\sigma `$ for CDM (open circles), and SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (crosses). Dashed and dotted lines bracket observations, showing scalings of $`Q\sigma ^3`$ and $`Q\sigma ^4`$, respectively.
SIDM generally shows lower phase space densities than CDM. SIDM is in somewhat better agreement with observations, falling in the middle of the observed range. Measuring $`Q`$ in galaxies is a difficult task, because the stellar velocity dispersion is not necessarily that of the dark matter. Furthermore, in rotationally supported galaxies the dubious assumption of an isothermal spherical halo is used to relate circular velocity to dispersion. Thus $`Q`$ is perhaps not among the most useful observational discriminants between CDM and SIDM.
An interesting remark from Figure 7 is that the scaling of $`Q(\sigma )`$ is roughly the same in both models, roughly $`Q\sigma ^3`$. Dalcanton & Hogan (2000) argue that such a scaling results from the dynamical assembly of halos, and is not expected based on simple phase packing arguments. This further motivates simulations of SIDM that include the cosmological growth of halos via dynamical processes of merging and accretion.
### 3.7. Ellipticities
SIDM produces halos that are more spherical than CDM, because of the isotropic nature of the collisions (Spergel & Steinhardt ). This is a generic feature of SIDM with any significant cross section, since in the inner portions of halos where collisions are frequent, the velocity ellipsoid is quickly isotropized. Thus the shapes of dark matter halos provide an important observational discriminant between CDM and SIDM.
We compute axis ratios of our halos using the prescription outlined in Dubinski & Carlberg (1991). They define a tensor
$$M_{ij}=\frac{x_ix_j}{a^2},\mathrm{with}a\left(x_1^2+\frac{x_2^2}{q^2}+\frac{x_3^2}{s^2}\right)^{\frac{1}{2}},$$
(11)
where $`a`$ is the elliptical radius, and $`sq1`$ are the axis ratios, and the sum is over all particles with distances $`(x_1,x_2,x_3)`$ from the halo center along the axes of the ellipsoid. Then,
$$q=\left(\frac{M_{yy}}{M_{xx}}\right)^{\frac{1}{2}}\mathrm{and}s=\left(\frac{M_{zz}}{M_{xx}}\right)^{\frac{1}{2}},$$
(12)
where $`M_{xx}M_{yy}M_{zz}`$ are the eigenvalues of $`M`$. $`q`$ represents the axisymmetry of the halo, while $`s`$ measures the halo flattening. Since $`a`$ depends on $`q`$ and $`s`$, the calculation of $`M_{ij}`$ must be iterated until convergence, which we take to be better than 0.01 in $`q`$ and $`s`$. This scheme weights particles roughly equally regardless of distance from center, unlike a moment of inertia tensor which weights the outskirts heavily, and thus better represents the ellipticity of the density distribution, as shown in Dubinski & Carlberg (1991).
Figure 8: Axisymmetry $`q`$ and flattening $`s`$ vs. $`r`$ for a $`6\times 10^{11}M_{}`$ halo (left panel) and a $`7\times 10^{10}M_{}`$ halo (right panel). Solid lines are CDM, dashed lines are SIDM, $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, and dot-dashed lines are SIDM, $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$. Lower of two curves is $`s`$.
Figure 8 shows axis ratios as a function of radius in our two most massive halos, having masses $`6\times 10^{11}M_{}`$ (left panel) and $`7\times 10^{10}M_{}`$ (right panel). Solid line is the CDM halo, dashed line is SIDM with $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, and dot-dashed line is SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$. $`q`$ is the upper of two curves for a given model.
CDM halos are fairly triaxial, while SIDM produces halos that are much closer to spherical. The effect is dependent on radius, as in the outer regions SIDM and CDM become more similar, since the effect of self-interactions is confined to the inner parts of halos. Still, even at $`30h^1\mathrm{kpc}`$ there are significant differences between SIDM and CDM. While not stated, this trend with radius is also evident from Figure 1 of Yoshida et al. (2000b).
Figure 9 shows histograms of axis ratios $`q`$ and $`s`$ at $`2h^1\mathrm{kpc}`$ (top panels) and $`10h^1\mathrm{kpc}`$ (bottom panels) for all halos with more than 1000 particles. The median value for each model is indicated by the corresponding tick mark on the top axis. The difference between CDM and SIDM is more pronounced at small radii, where CDM produces significantly triaxial halos while SIDM halos remain spherical. At large radii there is a much milder trend to more spherical halos with increasing $`\sigma _{\mathrm{DM}}`$. $`s`$ also shows more differences than $`q`$.
Figure 9 shows that while CDM produces halos are typically more spherical, there is still significant non-sphericity in many SIDM halos. In particular, there is a tail in the distributions of both CDM and SIDM to smaller axis ratios. This may be due to asymmetric infall that temporarily distorts the shape of the density in some halos, particularly smaller ones. This also may just be an artifact of finite number of collisions in smaller halos. Note that the two largest halos shown in Figure 8 show greater differences at $`10h^1\mathrm{kpc}`$ than suggested by the statistics in Figure 9.
Figure 9: Histogram of axis ratios $`q`$ (left panels) and $`s`$ (right panels), at $`2h^1\mathrm{kpc}`$ (upper panels) and $`10h^1\mathrm{kpc}`$ (lower panels). Only halos with more than 1000 particles are included. Tick marks at the upper axis show median values.
A comparison with observations of halo shapes is as yet inconclusive. In the inner portions of dark halos the shape of the potential is likely to be dominated by baryons, so a comparison to these simulations is not straightforward. Further out, perhaps the most direct observations of axisymmetry are those for galaxies with HI rings such as IC 2006, which suggests a very axisymmetric halo, $`q0.93\pm 0.08`$ at $`13`$ kpc (Franx, van Gorkom & de Zeeuw (1994)). Other observations (see Sackett (1999)) are more dependent on observational and theoretical uncertainties such as viewing angle and potential modeling, but persistently suggest $`q0.8`$ at $`1520`$ kpc. Both CDM and SIDM halos are consistent with these observations. Lensing maps of galaxies and clusters offer the best hope for mapping the mass potential in the inner halos, which should place strong constraints on SIDM.
Conversely, observations of $`s`$ from polar ring galaxies (e.g. Sackett et al. (1994)) and X-ray isophotes (e.g. Buote & Canizares (1998)) suggest a substantial amount of flattening, $`s0.5\pm 0.2`$ at $`r15`$ kpc in the density distribution. Such a flattening, if confirmed, may prove troublesome for SIDM. The baryonic component would provide a flattened contribution, but is not expected to be significant at those radii. It is not immediately evident how these discrepancies may be resolved, but we note that the problem is almost as severe for CDM as SIDM in our simulations. It is worth mentioning that our small simulation volume results in significantly reduced tidal distortion of large halos, so our simulations may not accurately represent the ellipticities of the outer portions of halos.
## 4. The Monte Carlo Resolution Limit
Our spatial resolution and mass resolution are well-understood. However, another resolution issue arises due to the Monte Carlo modeling of self-interactions. A Monte Carlo method must be sufficiently well sampled, resulting in a separate criterion for the number of particles in a halo to be well-represented by our simulation technique. In this section, we determine this criterion using our suite of lower resolution simulations with $`64^3`$ particles described in §2.
Since we are most concerned with the inner parts of halos, we focus on the inner slope as a function of mass as the best measure for examining this Monte Carlo resolution limit. Figure 10 shows a plot similar to right panel of Figure 4, except for the $`64^3`$ simulation of the $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ SIDM model. The top axis shows the number of particles in these halos. Here we compute $`\alpha `$ at $`2h^1\mathrm{kpc}`$ since that is the spatial resolution of our $`64^3`$ runs. The dashed line is the running median $`\alpha `$ from the $`128^3`$ simulation, computed at $`2h^1\mathrm{kpc}`$.
Figure 10 shows that for halos with $`1000`$ particles in the $`64^3`$ run, the median value of $`\alpha `$ is within $`1\sigma `$ of that of the $`128^3`$ run, though consistently lower. By 300 particles, the value of $`\alpha `$ is significantly lower in the $`64^3`$ run. The reason it is lower is because with few particles, the Monte Carlo procedure results in too few interactions to make the profile depart significantly from the collisionless CDM case. Thus to lower masses, SIDM looks increasingly like CDM when modeled using this technique.
Figure 10: $`\alpha `$ vs. $`M_{\mathrm{halo}}`$ for SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, for our $`64^3`$ run. $`\alpha `$ here is measured at $`2h^1\mathrm{kpc}`$, the resolution of our $`64^3`$ runs. Solid line shows a running median of the $`\alpha `$ distribution. Dashed line with error bars shows a similar curve from $`128^3`$ run. Deviations between two at a level $`>1\sigma `$ occur for halos having somewhere between 300 and 1000 particles.
Note that this limit is specific to our simulation parameters, redshift, and $`\sigma _{\mathrm{DM}}`$, and is not a general statement about the Monte Carlo $`N`$-body technique. The limit becomes higher as $`\sigma _{\mathrm{DM}}`$ is lowered, since collisions become less frequent, but even for $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, $`1000+`$ particle halos also appear convergent. We therefore take our Monte Carlo resolution limit to be $`1000`$ particles.
We list the values of $`\alpha _{\mathrm{med}}`$ and $`c_{M,\mathrm{med}}`$ for the $`64^3`$ runs in Table 1 for comparison with the $`128^3`$ results, where the median values here are computed for all halos with more that 500 particles in these smaller runs (roughly 10 in each simulation). Note that $`\alpha _{\mathrm{med}}`$ is computed at $`2h^1\mathrm{kpc}`$ instead of $`1h^1\mathrm{kpc}`$, partially explaining the steeper slopes even at the highest masses. In general, the trends indicated by the $`128^3`$ runs are reproduced at this lower resolution, suggesting that discrete particle effects do not significantly affect our conclusions. We have also examined the $`128^3`$ statistics presented previously using a limit of 300 particles instead of 1000, and our overall conclusions remain the same.
## 5. Subhalo Population
Self-interacting dark matter is predicted to significantly lower the population of subhalos orbiting around large halos, thereby bringing simulation predictions into better agreement with observations of the Local Group dwarf population. There are two reasons why SIDM has this effect: (1) The lowered central concentration and larger core radius makes small halos more susceptible to tidal disruption, and (2) Dark matter is ram-pressure stripped out of small galaxies as they move through the large central halo. In our $`128^3`$ simulations, we have one halo that is roughly Milky Way sized, having $`M6.7\times 10^{11}M_{}`$. In this section we examine the subhalo population around this large halo.
Figure 11: Subhalo positions within $`500h^1\mathrm{kpc}`$ the largest object in our volume, in CDM (left panel) and SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (right panel). Central cross is the position of the large halo.
Figure 11 shows a projected plot of halos within $`500h^1\mathrm{kpc}`$ of the largest halo in our volume, indicated by the central cross. Circle sizes are scaled as $`\mathrm{log}(M_{\mathrm{halo}})`$, with the smallest circles representing halos with $`M5\times 10^8M_{}`$. Left panel shows CDM, right panel shows SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$. The positions of subhalos are different due to the accumulated differences of chaotic orbits within a highly nonlinear potential well. Thus a halo-by-halo comparison for these small halos is not possible. A careful examination reveals that SIDM has fewer subhalos than the CDM distribution, especially the smallest ones.
Figure 12: Number of halos within $`500h^1\mathrm{kpc}`$ of the largest halo, histogrammed by mass, for collisionless (solid), $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ (dashed), and $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ (dot-dashed). Vertical dotted line indicates our 1000-particle Monte Carlo resolution limit.
Figure 12 quantifies this effect, showing the mass function of halos within $`500h^1\mathrm{kpc}`$ of our largest halo, for our three $`128^3`$ simulations. There is a clear trend that SIDM suppresses the subhalo population at the smallest masses. For $`10^{8.5}<M<10^9M_{}`$, CDM has 56 neighboring halos, $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ has 40, and $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ has 29.
While SIDM does reduce the population of smallest halos, the reduction is not nearly by the order of magnitude required to obtain agreement with Local Group dwarf galaxy counts (Moore et al. 1999a ). However, the effect of subhalo suppression in these SIDM simulations should be regarded as a lower limit to the true strength of the effect. The reason is that almost all these subhalos are well below our Monte Carlo resolution limit (dotted line in Figure 12), therefore their concentrations are approaching those in CDM models. Thus tidal disruption of these halos is not much stronger in SIDM as in CDM, and is increasingly similar to lower masses. Furthermore, ram-pressure stripping is reduced in effectiveness for the same reason that the Monte Carlo technique is less effective in these small halos. Hence the numbers of small SIDM halos are not as significantly suppressed relative to CDM as they should be.
We conclude that the simulations considered here suggest a weak trend in reducing the number of subhalos with increasing $`\sigma _{\mathrm{DM}}`$, but due to resolution effects we can make no robust quantitative estimates. What is required is to simulate a large halo with incredibly high resolution, having subhalos containing thousands of particles to properly model the effects of self-interactions. Such a simulation is unfortunately beyond the scope of our current computational resources. Alternatively, more sophisticated algorithms are necessary to model self-interactions in small halos moving through large ones, which is an avenue we are currently pursuing.
## 6. Comparison with Previous Work
A number of authors have investigated SIDM using $`N`$-body simulations. The literature divides into two subsets: Those that model interactions in the fluid approximation, effectively employing a large cross section, and those that model self-interactions in the optically thin regime as suggested by Spergel & Steinhardt . In both cases, there is disagreement over whether SIDM makes halos less or more concentrated than CDM.
Moore et al. (2000) and Yoshida et al. (2000a) simulate a galaxy cluster within a cosmological context, using a treecode with SPH to model interactions in the fluid approximation. Both studies resulted in halos that had isothermal profiles and were more centrally concentrated than CDM halos. This may be because the large effective cross section increases heat transfer efficiency, though as mentioned in the Introduction, for a sufficiently large cross section one expects heat transfer to be diminished. Yoshida et al. (2000a) suggested that intermediate cross sections would likely yield results that were intermediate between steep CDM profiles and steeper isothermal profiles, and thereby argued against SIDM. However, this is contradicted by Yoshida et al. (2000b), as well as the results presented here, confirming that the intermediate case results in halos that have long-lived shallow profiles.
In contrast, Bryan (private communication) uses a adaptive mesh hydrodynamics code to model self-interactions in a cosmological volume, and finds that even in the fluid limit SIDM produces sizeable, long-lived cores. It is not clear why different hydrodynamic codes give different results when they should be operating in the same regime. Perhaps the effective cross section is larger in the adaptive mesh code due to algorithmic differences, reducing heat transfer. Another possibility is that a numerical effect in SPH in which cold clumps moving through hot halos have their drag significantly overestimated (Tittley, Couchman & Pearce (1999)) makes objects rapidly sink into a dense, isothermal core. It is beyond the scope of this paper to resolve these issues. We simply note that the highly optically thick limit is not the relevant scenario to test the cross section range proposed by Spergel & Steinhardt .
Burkert (2000) and Kochanek & White (2000) simulated isolated halos with SIDM having a cross section closer to the range of Spergel & Steinhardt . They begin with a fully formed cuspy galaxy halo and study the evolution after interactions are turned on. They both find that halos develop a shallow core for some length of time, and then undergo core collapse. Burkert and Kochanek & White disagree on the timescales of core collapse; Burkert finds $`t_c16t_{\mathrm{dyn}}`$ in agreement with estimates from two-body relaxation, while Kochanek & White finds a much shorter collapse timescale of $`t_c2t_{\mathrm{dyn}}`$ for the same dark matter cross section. Kochanek & White explain this difference by arguing that Burkert ’s method underestimates collisions of slow moving particles. We note that our method does not suffer from this concern, as it is more like Kochanek & White .
We suggest some possible reasons why the results of Kochanek & White and (to a lesser extent) Burkert are at odds with ours. The first is that they begin with a cuspy Hernquist profile. This halo evolves rapidly initially (as seen in Figure 3 of Kochanek & White ), resulting in an artificially large amount of heat transfer. Second, they simulate an isolated halo, ignoring the accretion of dynamically hot material during the formation process that would keep the outer halo hot and delay core collape. We note that their dimensionless cross section $`\widehat{\sigma }_{DM}=M_{\mathrm{halo}}\sigma _{\mathrm{DM}}/r_s^2`$ converts to ours by a factor of $`2\times 10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ for $`M_{\mathrm{halo}}2\times 10^{10}M_{}`$ and $`r_s5h^1\mathrm{kpc}`$ (cf. Figure 1). Kochanek & White ’s simulation with $`\widehat{\sigma }_{DM}=1`$ already produces halos that maintain cores over many dynamical times (cf. top panel of their Figure 2). We suggest that a somewhat smaller $`\widehat{\sigma }_{DM}`$ might be consistent with observations as well as their limits on core collapse timescales, even without considering the effects of accretion and merging.
Kochanek & White also point out, as we have, that the Monte Carlo $`N`$-body technique requires a large number of particles for accurate modeling, and show that $`10^5`$ particles is sufficient. We note that while most of our halos do not have that many particles, our largest (Milky Way sized) halo has roughly $`2\times 10^5`$ particles, and its properties are consistent with those of smaller halos.
Yoshida et al. (2000b) have now performed a cosmological cluster simulation with a cross sections $`\sigma _{\mathrm{DM}}2\times 10^{25}2\times 10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ using a Monte Carlo $`N`$-body method. They find shallower central slopes and less concentrated cores with SIDM. Our combined results span the range from dwarf galaxies to clusters, and are in broad agreement with each other if we focus purely on the numerical results. For instance, they find that $`\sigma _{\mathrm{DM}}2\times 10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ produces a cluster having a core size of $`160h^1\mathrm{kpc}`$. If we extrapolate our largest halo’s core size (in our $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ run) to cluster scales using the expected $`r_{\mathrm{core}}v_{\mathrm{circ}}`$, we predict a core size of $`150h^1\mathrm{kpc}`$ for their cluster. For $`\sigma _{\mathrm{DM}}2\times 10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, they get $`100h^1\mathrm{kpc}`$ core, while we would predict $`80h^1\mathrm{kpc}`$ for $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$, again in good agreement. They then proceed to scale down from clusters to dwarf galaxies using a simplistic argument based on number of collisions, but their scaled result is contradicted by our direct simulations. We estimate that this may be because they use a CDM value for the scale radius and $`c_{\mathrm{NFW}}`$ of dwarfs, and compare them to SIDM values for their cluster. We find that $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ produces NFW scale radii that are double that of CDM (cf. Figure 6 and discussion); such a factor would go a long way towards alleviating the discrepancy. Taking this into account, we find the simulations of Yoshida et al. (2000b) to be broadly consistent with ours.
Figure 13: Halo profile of the largest halo in our $`64^3`$ simulations, for a range of $`\sigma _{\mathrm{DM}}`$ values. Halos are progressively less concentrated and have larger cores with increasing $`\sigma _{\mathrm{DM}}`$.
In order to explore the high-$`\sigma _{\mathrm{DM}}`$ limit, we ran $`64^3`$ simulations of SIDM with $`\sigma _{\mathrm{DM}}=10^{25}10^{22}\mathrm{cm}^2\mathrm{GeV}^1`$. The most illustrative result is to compare the density profile of the largest halo in all our $`64^3`$ simulations, as shown in Figure 13. As seen in Figure 1, there is a smooth trend of increasing core radius with $`\sigma _{\mathrm{DM}}`$. SIDM with $`\sigma _{\mathrm{DM}}=10^{25}\mathrm{cm}^2\mathrm{GeV}^1`$ is quite similar to CDM, though it may also have a core below our $`2h^1\mathrm{kpc}`$ resolution limit. Increasing $`\sigma _{\mathrm{DM}}`$ to $`10^{22}\mathrm{cm}^2\mathrm{GeV}^1`$, we continue to see no evidence for the development of an isothermal core due to accelerated heat transfer. The reason is because the collisions are so frequent in the outer portion of the halo that a dense core cannot develop. Instead, collisions randomize the dark matter velocities and prevent a smooth radial inflow required to generate a dense core. As dynamically hot material accretes onto the halo, heat keeps flowing inward and a large core is maintained. Our results are in better agreement with Bryan as opposed to Moore et al. (2000) and Yoshida et al. (2000a). This also illustrates why simulating SIDM beginning with an isolated cuspy Hernquist profile may not be appropriate for large $`\sigma _{\mathrm{DM}}`$; one should at least begin with a halo profile that is self-consistently stable for a few dynamic times.
## 7. Summary
We present a set of cosmological self-interacting dark matter simulations having cross-sections in the range favored by Spergel & Steinhardt (2000). Our simulations include the growth of halos from linear fluctuations in a random volume of the universe, with sufficient volume and resolution to obtain a statistical sample of galactic halos resolved to $`1h^1\mathrm{kpc}`$. We compare the resulting halos on a case-by-case basis to those in a collisionless CDM simulation having the same initial conditions.
Overall, SIDM is remarkably successful at reproducing observations of the inner portions of dark matter halos where CDM appears to fail. In particular, we find:
1. The inner slopes of SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ typical halos have $`\alpha 0.4`$ at $`r1h^1\mathrm{kpc}`$, with some scatter in $`\alpha `$. Our CDM halos have $`\alpha 1.5`$, in agreement with previous studies (e.g. Moore et al. 1999). SIDM with $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ is intermediate between these cases, with median $`\alpha 0.9`$. SIDM is in better agreement with a preliminary analysis of H$`\alpha `$ rotation curves of low surface brightness galaxies (Dalcanton & Bernstein (2000)).
2. SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ produces central densities $`\rho _c0.01M_{}\mathrm{pc}^3`$ at $`1h^1\mathrm{kpc}`$, and shows no trend with halo mass. SIDM with $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ has somewhat higher $`\rho _c`$ values, but remains fairly independent of mass. Conversely, $`\rho _c`$ in CDM halos is much larger than observed, typically $`0.1M_{}\mathrm{pc}^3`$ at $`1h^1\mathrm{kpc}`$, and shows a strong trend with halo mass. With their steep profiles, CDM halos are in significantly worse agreement at smaller radii. SIDM is thus is in better agreement with observations, as has also been argued by Firmani et al. (2000a).
3. Simulations with SIDM having $`\sigma _{\mathrm{DM}}=10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ are intermediate between CDM and SIDM with $`\sigma _{\mathrm{DM}}=10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$, indicating a smooth increase in the effect of SIDM with cross section, a result that extends (using lower-resolution simulations) from $`\sigma _{\mathrm{DM}}=10^{25}10^{22}\mathrm{cm}^2\mathrm{GeV}^1`$. In particular, the generation of singular isothermal halos is not seen in any of the massive halos simulated, even for $`\sigma _{\mathrm{DM}}=10^{22}\mathrm{cm}^2\mathrm{GeV}^1`$. This suggests that the dynamical process of halo growth in a cosmological setting helps keep outer regions of halos hot and prevents core collapse in a Hubble time.
4. We introduce a new mass concentration parameter $`c_M`$ based on a more directly observable quantity, the enclosed mass within tens of kpc. This halo concentration is significantly lower in SIDM models as compared to CDM, providing an observationally accessible discriminant that is not dependent on fitting a particular profile form. A rough estimate of $`c_M`$ for the Milky Way, with large corrections for baryonic effects, favors SIDM over CDM.
5. The central phase space density is lower in SIDM vs. CDM mostly due to the reduction in $`\rho _c`$. The velocity dispersions in the inner regions are quite similar. Both SIDM and CDM are consistent with observations shown in Dalcanton & Hogan (2000), though SIDM is mildly favored.
6. SIDM produces halos that are more spherical, especially in their inner regions, as compared to CDM. In principle, this is one of the strongest tests of the SIDM paradigm, as near the center any value of $`\sigma _{\mathrm{DM}}`$ that has a non-negligible effect on the dark matter distribution will increase the core sphericity, while CDM cores are almost always significantly triaxial. However, baryons are likely to dominate the shapes of the inner parts of halos, complicating a direct comparison, and in the outer parts the differences between SIDM and CDM are less pronounced.
7. The number of subhalos around our largest (Milky Way-sized) halo is somewhat reduced with increasing $`\sigma _{\mathrm{DM}}`$, but due to discreteness effects in our Monte Carlo $`N`$-body technique, we cannot put robust quantitative estimates on the strength of this effect.
Based on these simulations, our currently favored value for $`\sigma _{\mathrm{DM}}`$ is somewhere between $`10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ and $`10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$. Such a cross section simultaneously reproduces both the observed central density and inner slope, as well as being consistent with various observations considered here, which is non-trivial. In contrast, for instance, warm dark matter has difficulty simultaneously reproducing the observed central densities, inner slopes and subhalo population (Colín, Avila-Reese & Valenzuela (2000)).
As stated before, inner halo shapes may provide a strong discriminant between CDM and SIDM. On galactic scales, they are difficult to observe and confused by baryonic contributions. Conversely, clusters provide a cleaner test because they have large cores that are not baryon-dominated, and their mass distributions are directly observable via lensing. Miralda-Escudé (2000) uses the asphericity of cluster MS 2137-23 to (analytically) argue that $`\sigma _{\mathrm{DM}}<10^{25.5}\mathrm{cm}^2\mathrm{GeV}^1`$, effectively ruling out SIDM as a solution to halo concentration problems. On the other hand, CL 0024+1654 is very spherical, much more so than CDM models generally predict (Tyson, Kochanski & Dell’antonio (1998)). Our simulations cannot directly address the shapes of clusters, as we have no cluster-sized objects in our volume. However, SIDM shows some range of halo shapes due to asymmetric infall and unrelaxed mass distributions, so it is unclear whether a single object can definitively rule out SIDM. Support for this statement is provided by Yoshida et al. (2000b), whose cluster has enough triaxiality to be consistent with MS 2137-23 even for $`\sigma _{\mathrm{DM}}2\times 10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ (their model S1Wb), contradicting Miralda-Escudé’s scaling argument. We note that halo shapes are unaffected by annihilating (Kaplinghat, Knox & Turner (2000)) or decaying (Cen (2000)) dark matter, thus they also provide a discriminant between these variants and SIDM.
The cluster core sizes in the simulations of Yoshida et al. (2000b) are larger than observed ($`3070h^1\mathrm{kpc}`$, Miralda-Escudé (1995); Tyson, Kochanski & Dell’antonio (1998)), certainly for $`\sigma _{\mathrm{DM}}2\times 10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ ($`160h^1\mathrm{kpc}`$), and probably even for $`\sigma _{\mathrm{DM}}2\times 10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ ($`100h^1\mathrm{kpc}`$). So it may be that SIDM has difficulty matching observations at both dwarf galaxy and cluster scales. However, adiabatic contraction of baryons during the formation of the cD galaxy will reduce the cluster core radius from $`N`$-body predictions, so the discrepancy may not be that large. In any case, SIDM with $`\sigma _{\mathrm{DM}}10^{24}10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ comes remarkably close to matching dwarf galaxies, $`L^{}`$ galaxies and clusters given the $`10^5`$ range in mass scales, so we reserve judgement pending a more careful comparison with observations. Yoshida et al. (2000b) mention that SIDM core sizes would be in better agreement with observations if $`\sigma _{\mathrm{DM}}v^1`$, which would result in the effects of self-interactions being diminished in hot cluster environments as compared to galaxies. Such a scenario occurs naturally if the dark matter–dark matter scattering has low-lying resonance or bound state contributions, as is the case for ordinary nucleons.
Another theoretical avenue explored in relation to SIDM has been modeling the Tully-Fisher relation. Hydrodynamic simulations show that the simulated Tully-Fisher zero point may be brought into agreement with observations only if halos are less centrally concentrated than predicted by CDM (Navarro & Steinmetz (2000)). Mo & Mao (2000) determine that a cross section of $`\sigma _{\mathrm{DM}}10^{23}\mathrm{cm}^2\mathrm{GeV}^1`$ would produce a correct Tully-Fisher relation for $`v_{\mathrm{circ}}100`$ km/s halos. Thus it is conceivable that the cross section preferred from halo structure constraints may also alleviate the Tully-Fisher discrepancies.
Ostriker (2000) suggested that dark matter interactions in the centers of halos naturally produce central black holes with a mass scaling $`M_{\mathrm{BH}}\sigma ^5`$, in agreement with observations (Magorrian et al. (1998); Ferrarese & Merritt (2000)), and he estimates $`\sigma _{\mathrm{DM}}10^{24}\mathrm{cm}^2\mathrm{GeV}^1`$ in order to avoid central black holes that are too large. However, this estimate is based on an $`\rho r^2`$ profile all the way in to the black hole. If such a dense center never arises because collisions inhibit its formation, then the limits on $`\sigma _{\mathrm{DM}}`$ are weakened considerably.
Observations of dark matter halos promise to improve significantly in the coming years, particularly constraints on halo core shapes from lensing and on inner profiles and concentrations of halos from H$`\alpha `$ rotation curves. If the inner parts of dark matter halos are found to be generically triaxial, this would be the high place of sacrifice for self-interacting dark matter; conversely, spherical halos would provide strong support for this scenario. The main modeling work yet to be is done is an improved examination of the subhalo populations in SIDM, as well as simulations of a larger range of mass scales. The $`N`$-body Monte Carlo approach has difficulty achieving a large dynamic range due to the stringent Monte Carlo resolution limit (i.e. discreteness effects in the probabalistic description of collisions), therefore a different approach may be necessary.
The SIDM simulations presented here are a first attempt at examining the effect of self-interacting dark matter within the context of a realistic halo formation scenario. The results are quite encouraging that this simple variant of the cold dark matter paradigm will alleviate a wide range of difficulties faced by CDM on galactic scales. We look forward to further investigations and comparisons with observations.
We thank Julianne Dalcanton, Lars Hernquist, Jerry Ostriker, Penny Sackett, Scott Tremaine, Martin White, and Naoki Yoshida for helpful discussions. We thank Greg Bryan for sharing his unpublished results. RD and DNS are supported by NASA ATP grant NAG5-7066. DNS and BDW are supported by the NASA MAP/MIDEX program. PJS is supported by United States Department of Energy grant DE-FG02-91ER40671.
|
warning/0006/hep-th0006239.html
|
ar5iv
|
text
|
# Untitled Document
Noncommutative Spheres and the AdS/CFT Correspondence
Antal Jevicki, Mihail Mihailescu, Sanjaye Ramgoolam
Brown University
Providence, RI 02912
antal,mm,ramgosk@het.brown.edu
We present direct arguments for non-commutativity of spheres in the AdS/CFT correspondence. The discussion is based on results for the $`S_N`$ orbifold SCFT. Concentrating on three point correlations (at finite $`N`$) we exhibit a comparison with correlations on a non-commutative sphere. In this manner an essential signature of non-commutativity is identified giving further support for the original proposal of hep-th/9902059.
06/2000
1. Introduction
The AdS/CFT correspondence $`[1],[2],[3]`$ provides a constructive approach to supergravity, and closed string theory in curved backgrounds. Its basis is the large $`N`$ expansion of Yang-Mills type theory which turns into a loop expansion of gravity. The existence of a deductive procedure has prompted questions concerning any modification that the emerging gravity could exhibit. In particular a proposal was made in $`[4]`$, $`[5]`$, that the curved SUGRA background $`AdS\times S`$ is non-commutative with the noncommutativity parameter given by $`\frac{1}{N}`$. This follows an earlier proposal for q-deformation given in a framework of a simple matrix model $`[6]`$. The extension to higher spaces was further discussed in $`[7]`$.
The non-commutativity of $`AdS\times S`$ space naturally incorporates the exclusion principle of $`[8]`$ and stands to have important implications on the physics of black holes. It also conforms to a general principle of $`[9]`$ that string/M-theory inherently contains a space-time uncertainty( see $`[10]`$ ). Recently, a physical argument for the noncommutativity and the associated cutoff was given in $`[11]`$ based on a study of brane motions on spheres. This discussion involves a mechanism $`[12]`$, $`[13]`$ by which gravitons are polarized into extended spherical membranes which lead to non-commutativity $`[14]`$. Other examples are given in $`[15]`$, $`[16]`$, $`[17]`$, $`[18]`$, $`[19]`$.
It is clearly important to further clarify the nature of non-commutativity in $`AdS\times S`$ spaces. In comparison with the noncommutativity induced by an external B-field for the origin of noncommutativity in $`AdS\times S`$ is less trivial to exhibit. Since it involves the closed string coupling $`\frac{1}{N}`$, it represents a nonperturbative phenomenon.
In this paper, we describe further evidence for the above noncommutativity in AdS/CFT. The discussion is based on the $`S_N`$ orbifold model that already served as the basis for arguments presented in $`[4]`$, $`[5]`$. In this model, one is able to perform explicit calculations of three point interactions at finite $`N`$ and study their behavior. In this way, we exhibit an explicit signature for non-commutativity of the corresponding spheres.
The content of the paper is as follows. In section 2, we summarize the finite $`N`$ results for three-point correlations in the orbifold CFT and discuss their properties. We then review the (super)gravity calculations in commutative space-time in section 3 and proceed to evaluate the modifications due to a non-commutative sphere in section 4.We exhibit certain agreement with the SCFT study.
2. Results from $`S_N`$ orbifold
In this section, we review the results of CFT dual to the gravity in the case of $`AdS_3\times S^3`$ obtained in $`[5]`$. We also use the extension of these results to the nonextremal case that can be read off from recent work of $`[20]`$.
The SCFT in question is defined on symmetric product $`S^N(M)`$, where $`M`$ is either $`T^4`$ or $`K3`$ . The field content of the theory consists of: $`4N`$ real free bosons $`X_I^{a\dot{a}}`$ representing the coordinates of the torus for example and their superpartners $`4N`$ the fermions $`\mathrm{\Psi }_I^{\alpha \dot{a}}`$, where $`I=1,..,N`$, $`\alpha ,\dot{\alpha }=\pm `$ are the spinorial $`S^3`$ indices, and $`a,\dot{a}=1,2`$ are the spinorial indices on $`T^4`$. In essence, the field content of the theory is determined to be $`4N`$ real free bosons and $`2N`$ Dirac free fermions, giving a central charge $`c=6N`$. One has left and right superconformal symmetry with the corresponding currents. The lowest modes of this currents ,namely $`\{L_{0,\pm 1},G_{\pm \frac{1}{2}}^{\alpha a},J_0^{\alpha \beta }\}`$ together with their right counterparts generate the $`SU(2|1,1)_L\times SU(2|1,1)_R`$ symmetry.These according to the $`AdS/CFT`$ correspondence translate into the superisometries of the $`AdS_3\times S^3`$ space-time. In addition, one has other symmetries commuting with the previous set for example the ones related to global $`T^4`$ rotations. Even though the underlying CFT on $`T^4`$ is very simple, the complexity is given by the non-trivial implementation of the $`S_N`$ symmetry of the orbifold.This symmetry is an analogue of U(N) gauge symmetry in this context .Physical observables analogous to traces are now given by $`S_N`$ invariants . In particular the complete set of chiral primary operators was given in $`[4]`$, $`[5]`$. A fundamental role is played by the twist operators that impose the twisted the boundary conditions :
$$X_I(ze^{2\pi i},\overline{z}e^{2\pi i})=X_{I+1}(z,\overline{z}),I=1..n1,X_n(ze^{2\pi i},\overline{z}e^{2\pi i})=X_1(z,\overline{z})$$
for the n-twisted sector of the theory.First a construction of $`Z_n`$ twist operators is given in terms of which the $`S_N`$ invariant chiral primary operators are constructed after appropriately averaging over $`S_N`$.
In the correspondence with gravity on $`ADS_3\times S^3`$ one achieves a one-one correspondence with single particle states.
The computation of two and three point functions (for extremal momenta) was given in $`[5]`$ and has recently been extended in $`[20]`$. We present the three-point function for chiral primaries $`O_n^{(0,0)}`$, where $`n`$ denotes the length of the cycle (twist) used to construct the operator. The index $`n`$ is also identified with $`l`$ of the angular momentum on $`S^3`$ in gravity as $`n=l+1`$ (let us recall that the isometry group for $`S^3`$ is $`SO(4)=SU(2)\times SU(2)`$ and that $`l`$ is an angular momentum in the diagonal $`SU(2)`$). The correlation functions in terms of $`n=l_1+1`$, $`k=l_2+1`$ and $`n+k1=l_3+1`$ for the case $`l_3=l_1+l_2`$ (extremal) are:
$$O_{n+k1}^{(0,0)}(\mathrm{})O_k^{(0,0)}(1)O_n^{(0,0)}(0)=\left(\frac{(Nn)!(Nk)!(n+k1)^3}{(N(n+k1))!N!nk}\right)^{\frac{1}{2}}$$
This expressions for the three point correlation functions obtained in $`[5]`$ contains two types of factors: one that is solely dependent on the angular momenta with no $`N`$ dependence and the other with explicit $`N`$ dependence . It is the latter type that we concentrate on as it contains direct implications on the non-commutative nature of the spacetime.
The above result stands for the extremal $`l_1+l_2=l_3`$ case. We would now like to extend its main features to the general non-extremal case. We are guided by the following two facts. First are the results of $`[20]`$ which were done for a simpler bosonic orbifold but are found for general $`l`$‘s. Furthermore, as can be seen from the original discussion of$`[5]`$ the important $`N`$ dependence relevant to the cutoffs essentially originates from combinatorial properties of permutation group. Correspondingly, for general angular momenta $`l_{1,2,3}`$, we can expect the factor coming from $`S_N`$ permutations of the form:
$$\frac{\left((N1l_1)!(N1l_2)!(N1l_3)!\right)^{\frac{1}{2}}}{(N1\frac{l_1+l_2+l_3}{2})!}$$
It is this factor that contains the signature of non-commutativity. Firstly, one sees the ”exclusion principle” that, for a non-zero correlation function, any individual angular momenta $`l`$ should not exceed the bound N. There is a further bound on the sum of angular momenta characteristic of fusion rules of WZW models or of $`SU_q(2)`$ at roots of unity. But as we will see below, the full factorial form will appear to be present in the correlations to be evaluated on a non-commutative sphere.
3. Field theory on $`AdS\times S`$
We will now perform two parallel calculations,one with the standard (commutative) sphere and the other with the non-commutative (fuzzy) sphere. The purpose is to directly compare the results and see that the later case exhibits the factorial terms that were featured in the orbifold calculation.
In evaluating the correlation functions in SUGRA one has the following two-step process. The $`AdS`$ dependence is projected to the boundary of the $`AdS`$ space-time using the bulk to boundary propagator. For the sphere one expands in terms of spherical harmonics. The signature of noncommutative space that we are going to exhibit is associated with the $`S`$ (sphere) part of this calculation. The $`AdS`$ part does not lead to such a characteristic behavior and in much of what will be presented can be ignored. Indeed we should expect the essential features of the cutoff seen in the boundary correlators to be associated with the sphere part, since the models of q-deformed $`ADS`$ considered in have the property that the boundary is commutative. Other aspects like space-time uncertainty are manifested by the non-commutative ADS part. We will in the present section summarize the calculation of correlation functions in the commuting case and then in the next section repeat the analogous calculations in the non-commutative case.
Consider for simplicity a massless scalar field with cubic interaction:
$$S=𝑑x\sqrt{g}((\mathrm{\Phi })^2+\lambda \mathrm{\Phi }^3+\mathrm{})$$
where the integral is over the $`AdS\times S`$ space, $`g`$ is the corresponding metric and $`\lambda `$ represents the coupling constant for the cubic interaction. In studying field theory on products $`AdS\times S`$ space, the wavefunctions factorize into $`AdS`$ and $`S`$ components
$$\mathrm{\Phi }=\mathrm{\Phi }(\rho ,t,\varphi )\mathrm{\Psi }_I(\stackrel{}{n})$$
where $`\mathrm{\Psi }_I(\stackrel{}{n})`$ denote the spherical functions on the sphere $`S`$. For the case $`S_3`$, one has the well known $`D_{mm^{}}^{\mathrm{}}(\theta ,\phi ,\psi )`$ functions. For three point functions, one has schematically
$$\mathrm{\Phi }\mathrm{\Phi }\mathrm{\Phi }_{AdS}\mathrm{\Psi }_\mathrm{}_1\mathrm{\Psi }_\mathrm{}_2\mathrm{\Psi }_\mathrm{}_3_S$$
The sphere contributions (on which we will concentrate) exhibit typically the Clebsch-Gordon coefficients
$$\left(\begin{array}{ccc}\mathrm{}_1& \mathrm{}_2& \mathrm{}_3\\ m_1& m_2& m_3\end{array}\right)\left(\begin{array}{ccc}\mathrm{}_1& \mathrm{}_2& \mathrm{}_3\\ m_1^{}& m_2^{}& m_3^{}\end{array}\right)F(\mathrm{}_1,\mathrm{}_2,\mathrm{}_3;N)$$
and a factor $`F(\mathrm{};N)`$ solely dependent on the magnitudes of angular momenta and $`N`$. In the non-commutative case, this is the structure of a star product
$$Y^\mathrm{}_1Y^\mathrm{}_2=\left(ClebschGordon\right)f(\mathrm{}_1\mathrm{}_2\mathrm{}_3;N)Y^\mathrm{}_3$$
For studying the form of the “fusion coefficient”, $`F(\mathrm{}_1\mathrm{}_2,\mathrm{}_3:N)`$ , it is sufficient to consider the reduction of the sphere $`S_3`$ to $`S_2`$. One knows that in particular
$$D_{mm^{}=0}^{\mathrm{}}=Y_\mathrm{}m(\theta ,\varphi )$$
i.e. we have spherical harmonics on $`S_2`$. The nature of non-commutativity is essentially the same for the $`AdS_3\times S_3`$ space or $`AdS_2\times S_2`$ which is its $`U(1)\times U(1)`$ coset. So in what follows for simplicity of the calculation, we will discuss the latter.
We use the following representation for spherical harmonics:
$$Y^I=\mathrm{\Omega }_{i_1\mathrm{}i_l}^I\frac{x_1^i\mathrm{}x^{i_l}}{\rho ^l},$$
where $`\mathrm{\Omega }^I`$ is a traceless, symmetric, $`l`$-index tensor with indices $`i_k=1\mathrm{}3`$, $`x^i`$ are the coordinates of the three dimensional flat space and $`\rho =\sqrt{(x^1)^2+(x^2)^2+(x^3)^2}`$. $`Y^I`$ is an eigenvector of the sphere laplacian with eigenvalue $`l(l+1)`$. We also assume that the tensors $`\mathrm{\Omega }^I`$ are normalized in the sense: $`\mathrm{\Omega }_{i_1\mathrm{}i_l}^I\mathrm{\Omega }_{i_1\mathrm{}i_l}^J=\delta ^{IJ}`$ for equal index number and $`0`$ otherwise. By straightforward computations presented in the appendix 1 we obtain the following expressions for the product of two harmonics ((I,$`l_1`$), (J,$`l_2`$) indices) integrated over the sphere:
$$Y^IY^J\frac{1}{4\pi }_{S^2}Y^IY^J=\frac{\pi ^{\frac{1}{2}}\mathrm{\Gamma }(l+1)}{2^{l+1}\mathrm{\Gamma }(l+\frac{3}{2})}\delta ^{IJ},$$
For the product of three harmonics ((I,$`l_1`$), (J,$`l_2`$), (K,$`l_3`$) indices) we obtain:
$$Y^IY^JY^K\frac{1}{4\pi }_{S^2}Y^IY^JY^K=\frac{\pi ^{\frac{1}{2}}\mathrm{\Gamma }(l_1+1)\mathrm{\Gamma }(l_2+1)\mathrm{\Gamma }(l_3+1)}{2^{\frac{\mathrm{\Sigma }}{2}+1}\mathrm{\Gamma }(\alpha _1+1)\mathrm{\Gamma }(\alpha _2+1)\mathrm{\Gamma }(\alpha _3+1)\mathrm{\Gamma }(\frac{\mathrm{\Sigma }+3}{2})}C^{IJK}$$
where $`\mathrm{\Sigma }=l_1+l_2+l_3`$, $`\alpha _1=\frac{1}{2}(l_1+l_2+l_3)`$, $`\alpha _2=\frac{1}{2}(l_1l_2+l_3)`$, $`\alpha _3=\frac{1}{2}(l_1+l_2l_3)`$ and:
$$C^{IJK}=\mathrm{\Omega }_{i_1\mathrm{}i_{\alpha _3}k_1\mathrm{}k_{\alpha _2}}^I\mathrm{\Omega }_{i_1\mathrm{}i_{\alpha _3}j_1\mathrm{}j_{\alpha _1}}^J\mathrm{\Omega }_{j_1\mathrm{}j_{\alpha _1}k_1\mathrm{}k_{\alpha _2}}^J$$
We multiply the spherical harmonics with appropriate factors in order to normalize them. After we integrate over the sphere we obtain the following $`AdS_2`$ action:
$$S=_{AdS_2}d^2x\sqrt{g}(\varphi _I\varphi _I+l(l+1)\varphi _I\varphi _I+\lambda A^{IJK}\varphi _I\varphi _J\varphi _K)$$
where $`g`$ is now the $`AdS_2`$ metric and:
$$A^{IJK}=\frac{\pi ^{\frac{1}{4}}2^{\frac{1}{2}}\left(\mathrm{\Gamma }(l_1+1)\mathrm{\Gamma }(l_1+1)\mathrm{\Gamma }(l_3+1)\mathrm{\Gamma }(l_1+\frac{3}{2})\mathrm{\Gamma }(l_2+\frac{3}{2})\mathrm{\Gamma }(l_3+\frac{3}{2})\right)^{\frac{1}{2}}}{\mathrm{\Gamma }(\frac{\mathrm{\Sigma }+3}{2})\mathrm{\Gamma }(\alpha _1+1)\mathrm{\Gamma }(\alpha _2+1)\mathrm{\Gamma }(\alpha _3+1)}.$$
The action in $`AdS_2`$ can be related to the boundary action using the procedure and expressions developed in $`[3]`$, $`[21]`$ (see also $`[22]`$). The two- and three- point expressions for the constants in the correlation functions of the boundary operators $`𝙾_I`$ (corresponding to $`\varphi _I`$) are:
$$\begin{array}{cc}& 𝙾_I𝙾_J=\delta _{IJ}\frac{\mathrm{\Gamma }(\mathrm{\Delta }_I+1)}{\pi ^{\frac{1}{2}}\mathrm{\Gamma }(\mathrm{\Delta }_I\frac{1}{2})},\hfill \\ & 𝙾_I𝙾_J𝙾_K=\frac{\lambda A^{IJK}}{2\pi }\frac{\mathrm{\Gamma }(\frac{\mathrm{\Delta }_I+\mathrm{\Delta }_J+\mathrm{\Delta }_K}{2})\mathrm{\Gamma }(\frac{\mathrm{\Delta }_I\mathrm{\Delta }_J+\mathrm{\Delta }_K}{2})\mathrm{\Gamma }(\frac{\mathrm{\Delta }_I+\mathrm{\Delta }_J\mathrm{\Delta }_K}{2})\mathrm{\Gamma }(\frac{\mathrm{\Delta }_I+\mathrm{\Delta }_J+\mathrm{\Delta }_K2}{2})}{\mathrm{\Gamma }(\mathrm{\Delta }_I\frac{1}{2})\mathrm{\Gamma }(\mathrm{\Delta }_J\frac{1}{2})\mathrm{\Gamma }(\mathrm{\Delta }_K\frac{1}{2})},\hfill \end{array}$$
where $`\mathrm{\Delta }`$ for each operator is $`l+1`$. We redefine the operators such that the constant in the two-point function is $`\delta _{IJ}`$. After introducing all the factors, those coming from normalization and $`A^{IJK}`$, we obtain the following expression for the three-point correlation functions:
$$\begin{array}{cc}& 𝙾_I𝙾_J𝙾_K=\frac{\lambda C^{IJK}}{(2\pi )^{\frac{1}{2}}}\left(\frac{(l_1+\frac{1}{2})(l_2+\frac{1}{2})(l_3+\frac{1}{2})}{(l_1+1)(l_2+1)(l_3+1)}\right)^{\frac{1}{2}}\frac{\mathrm{\Gamma }(\alpha _1+\frac{1}{2})\mathrm{\Gamma }(\alpha _2+\frac{1}{2})\mathrm{\Gamma }(\alpha _3+\frac{1}{2})}{\mathrm{\Gamma }(\alpha _1+1)\mathrm{\Gamma }(\alpha _2+1)\mathrm{\Gamma }(\alpha _3+1)}\times \hfill \\ & \times \frac{\mathrm{\Gamma }(\frac{\mathrm{\Sigma }1}{2})}{\mathrm{\Gamma }(\frac{\mathrm{\Sigma }+3}{2})}.\hfill \end{array}$$
We observe that in our case the scalar has a mass given by the sphere laplacian $`\sqrt{l(l+1)}`$. In the case of $`AdS_2\times S_2`$ gravity, much of the analysis is in terms of chiral primary fields. The chiral primary fields are combinations of fields coming from four dimensional gravity that have the lowest possible $`AdS_2`$ mass for a given $`l`$: $`\sqrt{l(l1)}`$. We can assume that we do not start with such a simple theory as in $`(3.1)`$, but with one that after sphere reduction leads to the lowest mass, appropriate for a chiral primary field $`[3]`$. For such a field, the corresponding $`\mathrm{\Delta }`$ is $`l`$. In such a theory, we also consider a simple cubic interaction and obtain a qualitative picture for the sphere reduction. In this case, the three-point correlation functions for chiral primaries operators obtained in the end are simpler:
$$𝙾_I𝙾_J𝙾_K=\frac{\lambda C^{IJK}}{(2\pi )^{\frac{1}{2}}}\frac{4\left((l_1^2\frac{1}{4})(l_2^2\frac{1}{4})(l_3^2\frac{1}{4})\right)^{\frac{1}{2}}}{\alpha _1\alpha _2\alpha _3(\mathrm{\Sigma }^21)}.$$
We observe that the correlation functions in the case of chiral primary $`(3.1)`$ is much simpler than $`(3.1)`$. This is one of the features observed in all gravity cases of $`AdS_p\times S^p`$ reduction for chiral primary operators. The cancellation of factors appears between those coming from sphere and those coming from $`AdS`$ spaces. We also note that $`(3.1)`$ gives a divergent result for extremal cases like $`l_1+l_2=l_3`$. The divergences are due to our simplified model and they disappear in a realistic model. In the case of gravity, both the quadratic and cubic terms contain higher derivatives of the fields and these are responsible for both lower mass and consistent three-point correlation functions.
4. Field theory on non-commutative $`AdS_2\times S^2`$
We will now repeat the calculation of the above section, replacing the sphere by its noncommutative counterpart. In our previous work we have argued for non-commutative (q-deformation of) space-time in the context of the AdS/CFT correspondence. The q-deformed two-sphere can be defined as a quotient of $`SU(2)_q`$ and belongs to the classification of . There are also transformations between q-spheres with manifest $`SU(2)_q`$ symmetry and spheres with manifest $`SU(2)`$ symmetry. For generic $`q`$ this takes the form of a connection between the classical sphere and the q-sphere. For roots of unity this takes the form of a connection between the q-sphere and the fuzzy sphere. The technical reason for these connections is essentially the deformation maps between $`U_q`$ generators and the generators of the classical symmetry discussed in . Applying these transformations to both the algebra of functions on the sphere and to the symmetry generators acting on the algebra gives a transformation between $`q`$-sphere with $`U_q`$ symmetry and classical sphere with $`U(SU(2))`$ symmetry for generic $`q`$. This can be expected to lead, at roots of unity, to a transformation between fuzzy sphere and $`q`$-sphere. Indeed $`q`$-spheres at roots of unity are know to admit finite $`U_q`$ covariant truncations . The transformation is a non-commutative version of a diffeomorphism which should be a symmetry in these applications of quantum spheres to non-commutative gravity. In the following we work with the fuzzy sphere.
We first review the definition and properties of the fuzzy sphere giving formulae for the integration in parallel with the commutative case. This will lead to calculation of all the relevant quantities such as the normalization constants $`(3.1)`$ and the three harmonic interaction $`(3.1)`$.
The definition of the fuzzy sphere (for reviews see $`[27]`$, $`[28]`$, $`[29]`$) is given in terms of an algebra of polynomials in $`X^i`$, $`i=1\mathrm{}3`$ subject to the following constraints:
$$\begin{array}{cc}& [X^i,X^j]=iϵ_{ijk}X^k,\hfill \\ & (X^1)^2+\mathrm{}+(X^3)^2=\rho ^2,\hfill \end{array}$$
where $`\rho ^2`$ is a constant equal to $`\frac{N^21}{4}`$ ($`N`$ is a positive integer measuring the fuzziness of the sphere). Such a deformation preserves the $`SO(3)`$ symmetry of the sphere. We can represent the $`X`$’s (and the algebra) as hermitian operators in the $`SO(3)`$ representation having spin $`\frac{N1}{2}`$. As such, the coordinates are now hermitian $`N\times N`$ matrices. In this representation we define the integral over the sphere as:
$$\frac{1}{4\pi }_{S^2}(\mathrm{})\frac{1}{N}Tr(\mathrm{})$$
where $`Tr`$ is the trace in the $`\frac{N1}{2}`$ representation and $`\mathrm{}`$ mean a function on the sphere. It is also straightforward to represent the $`su(2)`$ symmetry (generators $`J_i`$) in this algebra:
$$J_if(X)=[X^i,f(X)],i=1\mathrm{}3$$
The spherical harmonics are constructed in the same way as in commutative sphere $`(3.1)`$ replacing the commutative coordinates $`x^i`$ with the noncommutative ones $`X^i`$. It is straightforward to prove that the vector space of symmetric traceless polynomial of degree $`l`$ is left invariant by the $`su(2)`$ generators and that it forms an irreducible representation with highest weight $`l`$.
The symmetric polynomials in $`X`$ of any degree appear in the Taylor series of $`exp(iJX)=exp(iJ_iX^i)`$, where $`J`$’s are regular commutative numbers. In order to compute the normalization constants and the three point interaction we construct the following quantity:
$$I(J^1,J^2,J^3)=\frac{1}{N}Tr(e^{iJ^1X}e^{iJ^2X}e^{iJ^3X})$$
We can extract from this the trace of the product of two and three symmetric polynomials as:
$$\begin{array}{cc}& \frac{1}{N}Tr(X^{(i_1}\mathrm{}X^{i_{l_1})}X^{(j_1}\mathrm{}X^{j_{l_2})})=_{J_{i_1}^1}\mathrm{}_{J_{i_{l_1}}^1}_{J_{j_1}^2}\mathrm{}_{J_{j_{l_2}}^2}I(J^1,J^2,0)|_{J^{1,2}=0},\hfill \\ & \frac{1}{N}Tr(X^{(i_1}\mathrm{}X^{i_{l_1})}X^{(j_1}\mathrm{}X^{j_{l_2})}X^{(k_1}\mathrm{}X^{k_{l_3})})=\hfill \\ & _{J_{i_1}^1}\mathrm{}_{J_{i_{l_1}}^1}_{J_{j_1}^2}\mathrm{}_{J_{j_{l_2}}^2}_{J_{k_1}^3}\mathrm{}_{J_{k_{l_3}}^3}I(J^1,J^2,J^3)|_{J^{123}=0},\hfill \end{array}$$
where $`(\mathrm{})`$ means the symmetrized product of $`X`$’s.
The evaluation of $`I(J^1,J^2,J^3)`$ can be done if we note that the RHS of $`(4.1)`$ is the trace of a product of three $`SO(3)`$ rotations with parameters $`J^{1,2,3}`$. The product of three rotations is itself a rotation with a parameter $`J=J(J_1,J_2,J_3)`$:
$$e^{iJX}=e^{iJ^1X}e^{iJ^2X}e^{iJ^3X},$$
and the trace of this operator can be evaluated easily in a basis where $`JX`$ is diagonal as:
$$I(J)I(J^1,J^2,J^3)=\frac{1}{N}\frac{sin(\frac{JN}{2})}{sin(\frac{J}{2})}$$
The dependence of $`J`$ (or rather $`cos(\frac{J}{2})`$) on $`J^{1,2,3}`$ can be easy computed (see appendix 2) and we list here the result:
$$\begin{array}{cc}& cos(\frac{J}{2})=cos(\frac{J^1}{2})cos(\frac{J^2}{2})cos(\frac{J^3}{2})cos(\frac{J^1}{2})\frac{sin(\frac{J^2}{2})}{J^2}\frac{sin(\frac{J^3}{2})}{J^3}cos(\frac{J^2}{2})\frac{sin(\frac{J^3}{2})}{J^3}\frac{sin(\frac{J^1}{2})}{J^1}\hfill \\ & cos(\frac{J^3}{2})\frac{sin(\frac{J^1}{2})}{J^1}\frac{sin(\frac{J^2}{2})}{J^2}+\frac{sin(\frac{J^1}{2})}{J^1}\frac{sin(\frac{J^2}{2})}{J^2}\frac{sin(\frac{J^3}{2})}{J^3}(J^1\times J^2)J^3\hfill \end{array}$$
The cubic interaction in the case of fuzzy sphere introduces an additional subtlety, namely:
$$\frac{1}{N}Tr(Y^IY^JY^K)\frac{1}{N}Tr(Y^IY^KY^J)$$
Because of this, some of the properties of cubic interaction we find in commutative case, are not there in the noncommutative case. In particular, we loose the appearance of the Clebsch-Gordon coefficients in the cubic interaction. This asymmetry is not present in the case $`\varphi ^3`$ interaction and not even in the case of $`\varphi _1^2\varphi _2`$, but it is present in the case $`\varphi _1\varphi _2\varphi _3`$ type interaction, where $`\varphi _{1,2,3}`$ are three different fields. We like to preserve those properties of interaction, as they seem to be present in the CFT $`[5]`$, and we change the definition of the integration over the fuzzy sphere by replacing the trace over the product of harmonics with the trace over the symmetric product of harmonics. The change amounts in the end in dropping the last factor appearing in the expression of $`cos(\frac{J}{2})`$ $`(4.1)`$, the only one not symmetric in $`J^{1,2,3}`$. After this change, we remain with $`J`$ depending on the following variables only: $`|J^1|`$, $`|J^2|`$, $`|J^3|`$, $`J^1J^2`$, $`J^2J^3`$ and $`J^1J^3`$, where $`|J|=\sqrt{J^2}`$.
The method used for the evaluation of the two- and three- interaction for harmonics is given in $`(4.1)`$. Spherical harmonics come with polynomial in $`X`$’s that are both symmetric and traceless. The traceless property of polynomials is shifted to the traceless of partial derivatives in $`J^1`$, $`J^2`$ and $`J^3`$ $`(4.1)`$ and as such, leads after setting $`J`$’s to $`0`$ to the following expression (we denote by $`\stackrel{~}{A}^{(i)(j)}`$ and $`\stackrel{~}{A}^{(i)(j)(k)}`$ the LHS of the equations $`(4.1)`$ with the property that the indices $`i`$’s, $`j`$’s and $`k`$’s are also traceless):
$$\begin{array}{cc}& \stackrel{~}{A}^{(i)(j)}=\delta _{l_1l_2}\frac{1}{2^{2l_1}}(\frac{}{(cos(\frac{J}{2}))})^{l_1}I(J)|_{J=0}(\delta ^{i_1j_1}\mathrm{}\delta ^{i_{l_1}j_{l_1}}+\mathrm{}),\hfill \\ & \stackrel{~}{A}^{(i)(j)(k)}=\frac{1}{2^{l_1+l_2+l_3}}(\frac{}{(cos(\frac{J}{2}))})^{\frac{l_1+l_2+l_3}{2}}I(J)|_{J=0}(\delta ^{i_1j_1}\mathrm{}\delta ^{i_{\alpha _3+1}k_1}\mathrm{}+\mathrm{})\hfill \end{array}$$
The $`\mathrm{}`$ in the last equations mean all possible contractions between indices $`i`$’s and $`j`$’s, $`j`$’s and $`k`$’s and $`i`$’s and $`k`$’s. We have also to specify:
$$(\frac{}{(cos(\frac{J}{2}))})^lI(J)|_{J=0}=\frac{(N+l)!}{N(N1l)!(2l+1)!!},$$
and we give a derivation for this in the appendix. The final results for harmonics can be written in terms of those obtained in the commutative case (we denote the corresponding terms for the fuzzy sphere as $`Y^IY^J_N`$ and $`Y^IY^JY^K_N`$):
$$\begin{array}{cc}& Y^IY^J_N=\frac{(N+l_1)!}{(2\rho )^{2l_1}N(Nl_11)!}Y^IY^J,\hfill \\ & Y^IY^JY^K_N=\frac{(N+\frac{l_1+l_2+l_3}{2})!}{(2\rho )^{l_1+l_2+l_3}N(N1\frac{l_1+l_2+l_3}{2})!}Y^IY^JY^K,\hfill \end{array}$$
where $`\rho `$ is given $`(4.1)`$. In the $`N\mathrm{}`$ limit, the factors coming from the fuzzy sphere go to $`1`$ and the results for the commutative sphere are obtained.
We have now prepared all the necessary ingredients to proceed to our main calculations and evaluate correlation functions associated with gravity on $`AdS_2\times S_{fuzzy}^2`$.
Consider now the following action:
$$S=_{AdS_2}d^2x\sqrt{g}\frac{1}{N}Tr(\varphi \varphi +J_i\varphi J_i\varphi +\lambda \varphi ^3)$$
where the integration over a (commutative) sphere is replaced by the $`Tr`$ over symmetric product of functions defined on the non-commutative (fuzzy) sphere. Expanding in terms of associated spherical harmonics, we obtain the same action on $`AdS_2`$ as in $`(4.1)`$ but with $`A^{IJK}`$ replaced by:
$$A_N^{IJK}=\frac{\left((N1l_1)!(N1l_2)!(N1l_3)!\right)^{\frac{1}{2}}}{(N1\frac{l_1+l_2+l_3}{2})!}\frac{N^{\frac{1}{2}}(N+\frac{l_1+l_2+l_3}{2})!}{\left((N+l_1)!(N+l_2)!(N+l_3)!\right)^{\frac{1}{2}}}A^{IJK}$$
After we reduce the theory on $`AdS_2`$ in the same way as before we obtain the result. The correlation functions in the case of non-commutative sphere are equal with those in the case of commutative sphere multiplied with the same factor as in $`(4.1)`$. We give below the result for the chiral primaries correlation functions in this case (see $`(3.1)`$):
$$\begin{array}{cc}& 𝙾_I𝙾_J𝙾_K_N=\frac{\lambda C^{IJK}}{(2\pi )^{\frac{1}{2}}}\frac{\left((N1l_1)!(N1l_2)!(N1l_3)!\right)^{\frac{1}{2}}}{(N1\frac{l_1+l_2+l_3}{2})!}\times \hfill \\ & \times \frac{N^{\frac{1}{2}}(N+\frac{l_1+l_2+l_3}{2})!}{\left((N+l_1)!(N+l_2)!(N+l_3)!\right)^{\frac{1}{2}}}\frac{4\left((l_1^2\frac{1}{4})(l_2^2\frac{1}{4})(l_3^2\frac{1}{4})\right)^{\frac{1}{2}}}{\alpha _1\alpha _2\alpha _3(\mathrm{\Sigma }^21)}.\hfill \end{array}$$
This expression contains the essential features relevant for comparison with the CFT. One has the $`SU(2)`$ symmetry, that was represented as R-symmetry in CFT. The result exhibits a cutoff at $`l_{1,2,3}=N`$ and also at $`\frac{l_1+l_2+l_3}{2}=N`$. Most significantly the overall factor that we see in the noncommutative (fuzzy) sphere result is of identical form to the corresponding factor given by the $`S_N`$ orbifold . This is clear evidence for non-commutativity in AdS/CFT. The characteristic features of the factorial terms describing this non-commutativity are seen to be captured by the sphere. It is relevant to stress that the $`AdS`$ contribution is not of the same form, which is understandable as we argued at the beginning of section 3 in terms of the commutative nature of the boundary in deformed ADS. It should be stated that the above expressions are not exactly equal in form to those of the finite N CFT. First the computation of this section is done for a simplified model. In addition, the gravitational coupling being given by $`1/N`$ means that loop effects will also contribute to the final $`N`$ dependence. Finally deformations of AdS will also imply corresponding $`N`$ dependence. But we expect that these two effects, while providing $`N`$ dependence, do not lead to a contribution of the form that we have identified and associated with a fuzzy sphere.
Acknowledgments: We are grateful to Sumit Das, Pei-Ming Ho, Samir Mathur and Jacek Pawelczyk for discussions. This work was supported by the Department of Energy under contract DE-FG02-91ER40688-Task A.
5. Appendix 1
In this appendix we give the formulas and some derivations used in section 3. For the computations for harmonics we use the following expressions:
$$\begin{array}{cc}& \frac{1}{4\pi }_{S^2}\frac{x^{i_1}\mathrm{}x^{i_l}}{\rho ^l}=\frac{_{J_{i_1}}\mathrm{}_{J_{i_l}}d^3xe^{\frac{1}{2}x^2+Jx}|_{J=0}}{4\pi _0^{\mathrm{}}𝑑\rho \rho ^{l+2}e^{\frac{1}{2}\rho ^2}}=\hfill \\ & =\frac{\pi ^{\frac{1}{2}}}{2^{\frac{l+2}{2}}\mathrm{\Gamma }(\frac{l+3}{2})}(\delta ^{i_1i_2}\mathrm{}+\mathrm{}),\hfill \end{array}$$
where $`\mathrm{}`$ mean all possible contractions between $`i`$’s. Then we obtain $`(3.1)`$ from:
$$Y^IY^J=\mathrm{\Omega }_{i_1\mathrm{}i_{l_1}}^I\mathrm{\Omega }_{j_1\mathrm{}j_{l_1}}^J\frac{1}{4\pi }_S^2\frac{x^{i_1}\mathrm{}x^{i_l}x^{j_1}\mathrm{}x^{j_{l_2}}}{\rho ^{l_1+l_2}}$$
and in a similar fashion $`(3.1)`$. The extra combinatorial factors as $`l!`$ for $`(3.1)`$ and $`\alpha _1!\alpha _2!\alpha _3!`$ and $`\mathrm{\Sigma }!`$ come from different combinatorial way to match the indices for $`\mathrm{\Omega }`$’s.
6. Appendix 2
In this appendix, we derive the formulas used in section 4. In order to derive the expression $`(4.1)`$ we use the spin $`\frac{1}{2}`$ representation for $`SO(3)`$ rotations $`e^{iJ^{1,2,3}x}`$, replacing $`x\frac{\sigma }{2}`$ with $`\sigma `$ being the Pauli matrices, and we obtain:
$$cos(\frac{J}{2})+i\frac{sin(\frac{J}{2})}{J}J\sigma =\underset{k=1}{\overset{3}{}}(cos(\frac{J^k}{2})+i\frac{sin(\frac{J^k}{2})}{J^k}J^k\sigma )$$
¿From this, after straightforward algebraic manipulations we obtain $`(4.1)`$.
For $`(4.1)`$, we define:
$$K_{N+1}^l(\frac{d}{d(cosJ)})^l(\frac{sin(J(N+1))}{sinJ})|_{J=0}$$
and using $`sin(J(N+1))=cos(JN)sinJ+sin(JN)cosJ`$ we obtain the following recurrence relations and conditions:
$$\begin{array}{cc}& K_{N+1}^l=K_N^l+(N+l)K_N^{l1},l1,\hfill \\ & K_N^0=N,\hfill \\ & K_1^l=0,l1,\hfill \end{array}$$
It is now easy to see that the expression in the RHS of $`(4.1)`$ satisfies $`(6.1)`$.
References
relax J. Maldacena, The large N limit of superconformal field theories and supergravity, Adv.Theor.Math.Phys.2: 231-252, 1998, hepth/9711200 relax S.S.Gubser,I.R.Klebanov and A.M.Polyakov, Gauge Theory Corellators from Non-Critical String Theory, Phys.Lett.B428 :105-114, 1998, hep-th/9802109 relax E. Witten, Anti-de-Sitter space and holography, Adv.Theor.Math.Phys.2:253-291, 1998, hep-th/9802150 relax A.Jevicki, S.Ramgoolam, Non commutative geometry from the ADS/CFT correspondence, JHEP 9904: 032, 1999, hep-th/9902059 relax A.Jevicki, M.Mihailescu, S.Ramgoolam, Gravity from CFT on $`S^N(X)`$ CFT: Symmetries and Interactions,Nucl.Phys.B577: 47-72, 2000, hep-th/9907144 relax A. Jevicki, A. van Tonder, Finite \[Q oscillator\] representation of 2-D string theory, Mod.Phys.Lett.A11:1397-1410, 1996, hep-th/9601058 relax Pei-Ming Ho, Sanjaye Ramgoolam, Radu Tatar, Quantum space-times and finite N effects in 4-D SuperYang-Mills theories, Nucl.Phys.B573:364-376,2000, hep-th/9907145 relax J. Maldacena, A. Strominger, AdS3 Black Holes and a Stringy Exclusion Principle, JHEP 9812 005, 1998, hep-th/9804085 relax T.Yoneya , String Theory and Space-Time Uncertainty Principle,hep-th/0004074 relax A. Jevicki, T. Yoneya, Space-time uncertainty principle and conformal symmetry in D particle dynamics, Nucl.Phys.B535:335-348,1998, hep-th/980506 relax J. McGreevy, L. Susskind, N. Toumbas, Invasion of the giant gravitons from anti-de Sitter space, JHEP 0006:008,2000, hep-th/0003075 relax R.C.Myers,Dielectric Branes,hep-th/9910053 relax D.Kabat and W.Taylor,Spherical membranes in Matrix theory, hep-th/ 9711078 relax Pei-Ming Ho and M.Li,Fuzzy spheres in AdS/CFT and holography from noncommutativity, hep-th/0004072 relax M.Berkooz and H.Verlinde, Matrix Theory,ADS/CFT and the Higgs-Coulomb Equivalence , hep-th/9907100 relax R. Schiappa, Matrix strings in weakly curved background fields, 2000, hep-th/0005145 relax D.Berenstein,V.Jejjala and R.G.Leigh,hep-th/0005087 relax H. Steinacker, Quantum Anti-de Sitter space and sphere at roots of unity, 1999, hep-th/9910037 relax S. Ferrara, M.A. Lledo, Some aspects of deformations of supersymmetric field theories, JHEP 0005:008, 2000, hep-th/0002084 relax O. Lunin, S. D. Mathur, Correlation functions for $`M^N/S(N)`$ orbifolds, hep-th/0006196 relax D.Freedman, S.Mathur, A.Mathusis and L.Rastelli, Correlation functions in the CFT(d)/AdS(d+1) correspondence, Nucl.Phys.B546(1999): 96-118, hep-th/9804058 relax J. Lee, Three point functions and the effective lagrangian for the chiral primary fields in D = 4 supergravity on $`AdS(2)\times S^2`$, hep-th/0005081 relax T. Masuda, K. Mimachi, Y. Nakagami, M. Noumi, K. Ueno, Representations of quantum groups and a q-analogue of orthogonal polynomials, C.R. Acad. Sci. Paris 307, 559-564 relax P. Podles, Quantum spheres, Lett. Math. Phys. 14, 193-202 (1987) relax T. L. Curtright and C. Zachos, Deforming maps for quantum algebras, Phys. Lett B243, 237-244, 1990 relax H. Suzuki, Rational scalar field theories on quantum spheres, Prog.Theor.Phys.91:379-391, 1994 relax J. Madore, The fuzzy sphere, Class.Quant.Grav. 9: 69-88, 1992 relax H. Grosse, C. Klimcik, P. Presnajder, Towards finite quantum field theory in noncommutative geometry, Int.J.Theor.Phys. 35: 231-244, 1996, hep-th/9505175 relax J. Gratus, An introduction to the noncommutative sphere and some extensions, 1997, q-alg/9710014
|
warning/0006/cond-mat0006313.html
|
ar5iv
|
text
|
# Resonances in the dynamics of ϕ⁴ kinks perturbed by ac forces
## I Introduction
One century and a half after their discovery, solitons and solitary waves have proven themselves ubiquitous in nature, arising in very many physical applications and leading to very important advances in applied mathematics . Generally speaking, it is often the case that the properties of solitary waves are known for certain equations, perhaps integrable, that relate to an oversimplified description of different physical systems; subsequently, one is interested to learn how these properties are modified if terms initially neglected are included as perturbations of the original equation. In most cases, this is a very complicated problem, and shedding light on it usually requires the use of approximate analytical approaches. One of the most succesful and widely applicable of these approaches is the collective coordinate technique or, rather, the family of collective coordinate techniques . The main merit of these procedures is the drastic reduction of the number of degrees of freedom involved in the problem, from the infinity of them in the original partial differential equation to the dynamics of a few degrees of freedom, governed by ordinary differential equations. Quite commonly, the reduction is done to a single degree of freedom, which, in general, can be identified with the center of the wave $`X(t)`$ or its velocity $`V(t)`$, as was first proposed in the mid-seventies . This amounts to mapping the motion of solitons or solitary waves to the motion of a pointlike (perhaps relativistic) particle with an effective mass . Surprisingly, this dramatic simplification leads to excellent results for very many systems . However, there are perturbations which, when acting on solitary waves, change not only the position of the center of mass of the kink, but also its width, as suggested in , or even some radiation can appear in the waves, a phenomenon that can yield the whole collective coordinate idea useless as more and more degrees of freedom are excited .
As we have mentioned above, in 1983 Rice developed a new perturbative method, which he applied to two well-known nonlinear Klein-Gordon problems, the $`\varphi ^4`$ and the sine-Gordon equations , in order to account for variations of the width of their kink solutions under perturbations. In Rice’s approach, the collective coordinates are the kink center $`X(t)`$ and its width $`l(t)`$. His results point out that when, in those systems, the kink is subject to “some perturbations” (sic) the simple translational motion of the kink center can be coupled to an oscillatory motion of the width of the wave, whose frequency he obtained by means of a variational approach. Interestingly, for the $`\varphi ^4`$ equation this so-called Rice frequency, $`\mathrm{\Omega }_R`$, practically coincides with the frequency $`\mathrm{\Omega }_i`$ of the kink internal mode (one of the modes of linear excitations around the kink, corresponding to a nonzero eigenvalue in the discrete spectrum, see, e.g., ). On the contrary, for the sine-Gordon (sG) case, $`\mathrm{\Omega }_R`$ turns out to be within the phonon spectrum and, furthermore, the sG kink does not have an internal mode (its only eigenvalue in the discrete spectrum is zero, corresponding to a Goldstone mode ). For this reason, the deformation of the kink width due to the internal mode has been studied extensively in the $`\varphi ^4`$ equation and, among other interesting results, we now know that the internal mode is able to store and transfer the energy in $`\varphi ^4`$ kink-antikink collisions , in the interaction of $`\varphi ^4`$ kinks with impurities , or in the case when the $`\varphi ^4`$ kink is subject to a periodic spatially modulated potential . This exchange of energy between the internal and the translational modes or among the internal mode and the modes of the impurities explains the resonances that take place in the above perturbed $`\varphi ^4`$ systems.
In view of these results, a question that naturally arises is the possible existence of resonances of the $`\varphi ^4`$ kink when perturbed by external ac forces, a problem that has not been considered so far and that relates to a number of physical contexts. Very recently, we have shown that a strong resonance arises when the $`\varphi ^4`$ system is subject to an external ac force of the form $`f(t)=ϵ\mathrm{sin}(\delta t+\delta _0)`$ with $`\delta `$ close to $`\mathrm{\Omega }_i/2`$, while in the case that $`\delta =\mathrm{\Omega }_i`$ this resonance is weak and even disappears for an appropriate choice of the initial kink velocity and the parameters of the driving force. However, the work reported in contained mostly numerical results and only an intuitive explanation of this striking phenomenon in terms of the collective coordinate equations, and therefore our aim in this paper is to provide a full analytical treatment of those equations (for the undamped case) in order to better understand the anomalous resonance phenomenon. In addition, in doing so we will be already carrying out the same analysis for the sG equation, which may be of interest in order to clarify the question of the existence of an internal quasi-mode for this equation . We deal with these issues along the paper according to the following scheme: In the next Section we use the so called generalized travelling wave Ansatz (GTWA) to obtain the equations governing the dynamics of the kink center and width (which is associated to the excitation of the internal mode). In Sec. III we thoroughly analyze those equations in the absence of damping, identifying all the possible resonances and their locations as a function of the equation parameters. We then focus on the most interesting resonances, namely those at $`\mathrm{\Omega }_i/2`$ and $`\mathrm{\Omega }_i`$. For the damped case we have not been able to solve analytically the equation for $`l(t)`$, but in Sec. IV we present numerical simulations of the full partial differential equation confirming that the resonance at $`\mathrm{\Omega }_i/2`$ remains for weak damping, while the resonance at $`\delta \mathrm{\Omega }_i`$ disappears for any damping value. Additionally, in that section we have compared all our analytical results to the numerical simulations, finding an excellent agreement. Finally, in the last section we summarize our results and discuss their implications for other systems, focusing especially on the sG kink dynamics. In an appendix, we prove the equivalence of the GTWA to a different procedure to obtain collective coordinate equations, which uses the momentum and the energy of the system similarly to the classical approach in .
## II Collective coordinate approach
The dynamics of $`\varphi ^4`$ kinks subject to a periodic force $`f(t)=ϵ\mathrm{sin}(\delta t+\delta _0)`$ is governed by the equation
$$\begin{array}{c}\varphi _{tt}\varphi _{xx}\varphi +\varphi ^3=\beta \varphi _t+f(t),\end{array}$$
(1)
where $`\beta `$ is the damping coefficient, and $`ϵ`$, $`\delta `$ and $`\delta _0`$ represent the amplitude, the frequency and the phase of the periodic force, respectively.
In order to apply the GTWA, first proposed in , we rewrite Eq. (1) in the following way:
$`\dot{\varphi }`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \psi }},`$ (2)
$`\dot{\psi }`$ $`=`$ $`{\displaystyle \frac{\delta H}{\delta \varphi }}+F(x,t,\varphi ,\varphi _t,\mathrm{}),`$ (3)
where $`\psi =\dot{\varphi }`$ and the dot represents derivative with respect to time, $`F(x,t,\varphi ,\varphi _t,\mathrm{})=\beta \dot{\varphi }+f(t)`$, and $`H`$ is the Hamiltonian of the system with the corresponding potential, $`U(\varphi )=(\varphi ^21)^2/4`$, when $`ϵ`$ and $`\beta `$ are zero,
$$\begin{array}{c}H=_{\mathrm{}}^+\mathrm{}𝑑x\left\{\frac{1}{2}\psi ^2+\frac{1}{2}\varphi _x^2+U(\varphi )\right\}.\hfill \end{array}$$
(4)
We now assume that the solution of (2)-(3) has the form
$$\begin{array}{c}\varphi (x,t)=\varphi [xX(t),l(t)],\hfill \end{array}$$
(5)
and, hence, from the definition of $`\psi `$ we have that
$$\begin{array}{c}\psi (x,t)=\psi [xX(t),l(t),\dot{X},\dot{l}].\hfill \end{array}$$
(6)
Here $`\varphi `$ describes the soliton shape, whose center will be given by $`X(t)`$, and where we introduced a second collective variable, $`l(t)`$, which will represent the kink width as we will see below. This Ansatz takes into account that $`l`$ can be different from the Lorentz-contracted width due to the action of the external force. Since the internal mode is related with the kink width, we can expect that by using this approach we will be able to explain the kink motion when the internal mode is excited. However, the linearized problem around the initial kink solution tells us that perturbations of the kink can not only shift its position or change its width, but also make it emit radiation . We have not considered this effect and therefore the Ansatz is only valid when the external force is so small that practically no radiation is emitted. Since the frequencies of the discrete internal mode $`\mathrm{\Omega }_i=\sqrt{3/2}`$ and the continuum phonon band $`\omega _p=\sqrt{2+k^2/2}`$ are apart from each other one can expect that the two kinds of modes are excited at different values of the parameters of the external force, and then when resonances appear as a consequence of the excitation of the internal mode one can in principle neglect the effect of radiation. In any event all our analytical results will have to be confirmed later by numerical simulations, which will show whether or not this assumption is correct.
To obtain the equations for our collective coordinates $`X(t)`$ and $`l(t)`$ we will follow (see the appendix for another possible derivation). First, we insert (5)-(6) into (2) and (3), and then we multiply the first obtained equation by $`\psi /X`$ and the second one by $`\varphi /X`$; subtracting both expressions and integrating we arrive at the following system of equations:
$$\begin{array}{c}_{\mathrm{}}^+\mathrm{}𝑑x\frac{\varphi }{X}\frac{\psi }{\dot{X}}\ddot{X}+_{\mathrm{}}^+\mathrm{}𝑑x[\varphi ,\psi ]\dot{l}+_{\mathrm{}}^+\mathrm{}𝑑x\frac{\varphi }{X}\frac{\psi }{\dot{l}}\ddot{l}F^{stat}(X)=\hfill \\ =_{\mathrm{}}^+\mathrm{}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{})\frac{\varphi }{X},\hfill \end{array}$$
(7)
with
$$\begin{array}{c}[\varphi ,\psi ]=\frac{\varphi }{X}\frac{\psi }{l}\frac{\varphi }{l}\frac{\psi }{X},\hfill \end{array}$$
(8)
$$\begin{array}{c}F^{stat}=_{\mathrm{}}^+\mathrm{}𝑑x\left\{\frac{\delta H}{\delta \varphi }\frac{\varphi }{X}+\frac{\delta H}{\delta \psi }\frac{\psi }{X}\right\}=_{\mathrm{}}^+\mathrm{}𝑑x\frac{}{X}=\frac{E}{X},\hfill \end{array}$$
(9)
where $`E`$ represents the energy of the system, $``$ is the Hamiltonian density and $`F^{stat}`$ is the static force due an the external field or other solitons, equal to zero for the above Hamiltonian. In order to obtain the second equation of motion we proceed as in the previous equation and begin by inserting (5) and (6) into (2) and (3), then multiplying the first and second equation by $`\psi /l`$ and $`\varphi /l`$ respectively, and finally taking the difference and integrating, thus finding
$$\begin{array}{c}_{\mathrm{}}^+\mathrm{}𝑑x[\psi ,\varphi ]\dot{X}+_{\mathrm{}}^+\mathrm{}𝑑x\frac{\varphi }{l}\frac{\psi }{\dot{X}}\ddot{X}+_{\mathrm{}}^+\mathrm{}𝑑x\frac{\varphi }{l}\frac{\psi }{\dot{l}}\ddot{l}K^{int}(X)=\hfill \\ =_{\mathrm{}}^+\mathrm{}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{})\frac{\varphi }{l},\hfill \end{array}$$
(10)
where
$$\begin{array}{c}K^{int}(l,\dot{l},\dot{X})=_{\mathrm{}}^+\mathrm{}𝑑x\frac{}{l}=\frac{E}{l}.\hfill \end{array}$$
(11)
By this procedure, we have obtained two coupled second-order ordinary differential equations for $`X(t)`$ and $`l(t)`$, where up to now we have not imposed any condition on the soliton shape; however, to solve Eqs. (7)–(11) we need the explicit functional dependence of $`\varphi (x,t)`$. Following Rice we assume that
$$\begin{array}{c}\varphi (x,t)=\varphi _0[xX(t),l(t)]=\mathrm{tanh}\left[\frac{xX(t)}{l(t)}\right],\hfill \end{array}$$
(12)
where $`\varphi _0==\mathrm{tanh}[xX_0/l_0]`$ is the static kink solution of the $`\varphi ^4`$ system, centered at $`X_0`$ and of width $`l_0=\sqrt{2}`$. Substituting (12) into (7)–(11) and integrating over $`x`$ we obtain
$`M_0l_0{\displaystyle \frac{\ddot{X}}{l}}M_0l_0{\displaystyle \frac{\dot{X}\dot{l}}{l^2}}`$ $`=`$ $`F^{stat}(X)\beta M_0l_0{\displaystyle \frac{\dot{X}}{l}}+F_{ex},`$ (13)
$`\alpha M_0l_0{\displaystyle \frac{\ddot{l}}{l}}+M_0l_0{\displaystyle \frac{\dot{X}^2}{l^2}}`$ $`=`$ $`K^{int}(l,\dot{l},\dot{X})\beta \alpha M_0l_0{\displaystyle \frac{\dot{l}}{l}}+K,`$ (14)
where
$$\begin{array}{c}F_{ex}=_{\mathrm{}}^+\mathrm{}𝑑xf(t)\frac{\varphi }{X}=qf(t),F^{stat}=0,\hfill \end{array}$$
(15)
$$\begin{array}{c}K=_{\mathrm{}}^+\mathrm{}𝑑xf(t)\frac{\varphi }{l}=0,K^{int}=\frac{E}{l},\hfill \end{array}$$
(16)
$$\begin{array}{c}E=\frac{1}{2}\frac{l_0}{l}M_0\dot{X}^2+\frac{1}{2}\frac{l_0}{l}\alpha M_0\dot{l}^2+\frac{1}{2}M_0\left(\frac{l_0}{l}+\frac{l}{l_0}\right),\hfill \end{array}$$
(17)
with $`\alpha =(\pi ^26)/12`$, $`q=2`$ and $`M_0=4/(3l_0)`$. Denoting $`P(t)M_0l_0\dot{X}/l`$, the Eq. (13) can be written as
$$\begin{array}{c}\frac{dP}{dt}=\beta Pqf(t).\hfill \end{array}$$
(18)
It is interesting to note that this equation may be obtained as well by applying the McLaughlin and Scott procedure with one collective variable only corresponding to the center of the kink. As was shown in , for the sG kink dynamics this is an excellent description of the kink motion, and we have verified that it also describes the $`\varphi ^4`$ kink motion under ac forces away from the resonances we will find and discuss below. Its solution is given by
$$\begin{array}{c}P(t)=\frac{qϵ}{(\beta ^2+\delta ^2)}\left[\delta \mathrm{cos}(\delta t+\delta _0)\beta \mathrm{sin}(\delta t+\delta _0)\right]+\\ +\mathrm{exp}(\beta t)\left\{P(0)+\frac{qϵ}{(\beta ^2+\delta ^2)}\left[\beta \mathrm{sin}(\delta _0)\delta \mathrm{cos}(\delta _0)\right]\right\}.\end{array}$$
(19)
From Eq. (14) the equation that holds for the width $`l(t)`$ is
$$\begin{array}{c}\alpha \left[\dot{l}^22l\ddot{l}2\beta l\dot{l}\right]=\frac{l^2}{l_0^2}\left[1+\frac{P^2}{M_0^2}\right]1.\hfill \end{array}$$
(20)
Note that the term $`P(t)^2`$ in Eq. (20) involves two frequencies $`\delta `$ and $`2\delta `$ when $`\beta =0`$; whereas for $`\beta 0`$ and after some transient time the only frequency that remains is $`2\delta `$. Furthermore, the term with frequency $`\delta `$ vanishes for $`\beta =0`$ and an appropriate choice of initial parameters. This equation represents a nonlinear, damped and parametrically excited oscillator. To solve Eq. (20) we provide the following change of variables, proposed in , $`l(t)=g^2(t)`$, which transforms the above equation into an Ermakov-type equation (or Pinney-type) \[see and references therein\]. Then, the equation for $`g(t)`$ reads
$$\begin{array}{c}\ddot{g}+\beta \dot{g}+\left[\left(\frac{\mathrm{\Omega }}{2}\right)^2+\left(\frac{\mathrm{\Omega }}{2M_0}\right)^2P^2\right]g=\frac{1}{4\alpha g^3},\hfill \\ g(0)=\sqrt{l_s}0,\dot{g}(0)=\frac{\dot{l}(0)}{2\sqrt{l_s}},\hfill \end{array}$$
(21)
where $`\mathrm{\Omega }=1/\sqrt{\alpha }l_0=1.2452`$ is equal to the Rice frequency $`\mathrm{\Omega }_R=1/\sqrt{\alpha }l_s`$ in the case when the kink initially is at rest; this agrees within $`1.7\%`$ with $`\mathrm{\Omega }_i=\sqrt{3/2}=1.2247`$. We have not been able to solve analytically Eq. (21), except when $`\beta =0`$. The next section is devoted to a detailed analysis of that case; we will come back to the nonzero $`\beta `$ problem when discussing our numerical results in Sec. IV.
## III Undamped kink: $`\beta =0`$
When $`\beta =0`$, Eq. (21) reads
$$\begin{array}{c}\ddot{g}+\left[\left(\frac{\mathrm{\Omega }}{2}\right)^2+\left(\frac{\mathrm{\Omega }}{2M_0}\right)^2P^2\right]g=\frac{1}{4\alpha g^3},\end{array}$$
(22)
where $`P(t)`$ is given by $`P(t)=\lambda +qϵ\mathrm{cos}(\delta t+\delta _0)/\delta `$, with $`\lambda M_0\gamma _0u(0)/l_sqϵ\mathrm{cos}(\delta _0)/\delta `$. We thus see that the function $`P(t)^2`$ in Eq. (22) involves trigonometric functions with frequencies $`\delta `$ and $`2\delta `$ if and only if $`\lambda 0`$. On the contrary, when $`\lambda =0`$ the only frequency that remains in the function $`P(t)^2`$ is $`2\delta `$. Interestingly, we note that the relation $`\lambda =0`$ coincides with the condition for the oscillatory motion of the center of the kink, obtained by using the McLaughlin and Scott approach in the absence of dissipation . The solution of Eq. (22) is
$$\begin{array}{c}g(t)=\sqrt{v_1^2+\frac{1}{4\alpha W^2}v_2^2},\end{array}$$
(23)
where $`v_1(t)`$ and $`v_2(t)`$ are two independent solutions of the linear part of Eq. (22), $`W=\dot{v}_1v_2\dot{v}_2v_1`$ is the Wronskian, which in this case is a constant and can be calculated by the initial conditions for $`v_i`$ ($`i=1,2`$) which are $`v_1(0)=\sqrt{l_s}`$, $`\dot{v}_1(0)=\dot{l}(0)/(2\sqrt{l_s})`$, $`v_2(0)=0`$ and $`\dot{v}_2(0)=const0`$.
### A $`\lambda =0`$: Resonance at $`\delta \mathrm{\Omega }_R/2`$
If one denotes $`\tau =\delta t+\delta _0`$ and sets $`\lambda =0`$ in the linear part of Eq. (22), after some manipulations we arrive at the following Mathieu equation for the $`v_i`$ functions:
$$\begin{array}{c}v_i^{\prime \prime }+[a+2\theta \mathrm{cos}(2\tau )]v_i=0,\\ a=\left(\frac{\mathrm{\Omega }}{2\delta }\right)^2\left[1+\frac{q^2ϵ^2}{2\delta ^2M_0^2}\right],\theta =\left(\frac{\mathrm{\Omega }}{2\delta }\right)^2\frac{q^2ϵ^2}{4\delta ^2M_0^2},\end{array}$$
(24)
where prime denotes the derivative with respect to $`\tau `$. Notice that the initial conditions for $`v_i(\tau )`$ become $`v_1(\delta _0)=\sqrt{l_s}`$, $`v_1^{}(\delta _0)=\dot{l}(0)/(2\delta \sqrt{l_s})`$, $`v_2(\delta _0)=0`$ and $`v_2^{}(\delta _0)=\dot{v}(0)/\delta `$. The solution of Eq. (24) (see ) for $`v_1(\tau )`$ and $`v_2(\tau )`$ can be expressed as a linear superposition of the two Mathieu functions $`ce_\nu `$ and $`se_\nu `$ with a non-integer index $`\nu `$, i.e.,
$$\begin{array}{c}v_i(\tau )=A_i\mathrm{ce}_\nu (\tau ,\theta )+B_i\mathrm{se}_\nu (\tau ,\theta ),i=1,2,\hfill \end{array}$$
(25)
where
$$\begin{array}{c}A_i\frac{\mathrm{\Delta }_{A_i}}{\mathrm{\Delta }},B_i\frac{\mathrm{\Delta }_{B_i}}{\mathrm{\Delta }},\hfill \end{array}$$
(26)
and
$`\begin{array}{c}\mathrm{\Delta }=\mathrm{ce}_\nu (\delta _0,\theta )\mathrm{se}_\nu ^{}(\delta _0,\theta )\mathrm{ce}_\nu ^{}(\delta _0,\theta )\mathrm{se}_\nu (\delta _0,\theta ),\hfill \\ \mathrm{\Delta }_{A_i}=v_i(\delta _0)\mathrm{se}_\nu ^{}(\delta _0,\theta )v_i^{}(\delta _0)\mathrm{se}_\nu (\delta _0,\theta ),\hfill \\ \mathrm{\Delta }_{B_i}=v_i^{}(\delta _0)\mathrm{ce}_\nu (\delta _0,\theta )v_i(\delta _0)\mathrm{ce}_\nu ^{}(\delta _0,\theta ),\hfill \end{array}`$with the constraint (characteristic curve for Mathieu functions)
$$\begin{array}{c}a=\nu ^2+\frac{1}{2(\nu ^21)}\theta ^2+O(\theta ^4).\hfill \end{array}$$
(27)
From Eqs. (23), (25) and (26), and taking into account that $`\tau =\delta t+\delta _0`$ we obtain that the kink width $`l(t)`$ is given by
$$\begin{array}{c}l(t)=g^2=v_1^2(t)+\frac{1}{4\alpha W^2}v_2^2(t),\hfill \end{array}$$
(28)
$$\begin{array}{c}v_i(t)=A_i\mathrm{ce}_\nu (\delta t+\delta _0,\theta )+B_i\mathrm{se}_\nu (\delta t+\delta _0,\theta ),i=1,2;\hfill \end{array}$$
(29)
where
$$\begin{array}{c}A_i\frac{v_i(0)\dot{\mathrm{se}}_\nu (\delta _0,\theta )\dot{v}_i(0)\mathrm{se}_\nu (\delta _0,\theta )}{\left[\mathrm{ce}_\nu (\delta _0,\theta )\dot{\mathrm{se}}_\nu (\delta _0,\theta )\dot{\mathrm{ce}}_\nu (\delta _0,\theta )\mathrm{se}_\nu (\delta _0,\theta )\right]},\hfill \\ B_i\frac{v_i(0)\dot{\mathrm{ce}}_\nu (\delta _0,\theta )\dot{v}_i(0)\mathrm{ce}_\nu (\delta _0,\theta )}{\left[\mathrm{ce}_\nu (\delta _0,\theta )\dot{\mathrm{se}}_\nu (\delta _0,\theta )\dot{\mathrm{ce}}_\nu (\delta _0,\theta )\mathrm{se}_\nu (\delta _0,\theta )\right]},\hfill \end{array}$$
(30)
$$\begin{array}{c}W=\sqrt{l_s}\dot{v}_2(0),\hfill \end{array}$$
(31)
and the characteristic curve, Eq. (27) , for our initial parameters can be written up to order $`ϵ^2`$ as
$$\begin{array}{c}\delta =\frac{\mathrm{\Omega }_R}{2\nu }\frac{q^2\nu \mathrm{cos}(2\delta _0)}{2M_0^2\gamma _0^2\mathrm{\Omega }_R}ϵ^2+O(ϵ^4).\end{array}$$
(32)
To obtain a better approximation, we need to take into account more terms of the above series, so in Fig. 1 we have plotted, as solid lines, the characteristic curves obtained numerically with Mathematica 3.0 , for $`\nu =1/2`$ and $`\nu =3/2`$ when $`u(0)=0`$ and $`\delta _0=\pi /2`$. Notice that when $`\nu =m+p/s`$ is rational, with $`m`$ an integer number and $`p/s`$ a rational fraction ($`0<p/s<1`$), $`v_1(t)`$ or $`v_2(t)`$ are $`2\pi s`$ periodic functions, if $`p`$ is odd, and $`\pi s`$-periodic functions, if $`p`$ is even; whereas for irrational $`\nu `$ both functions will be non-periodic, but bounded solutions .
If we know $`l(t)`$, then from the solution $`P(t)`$ \[Eq. (19)\], we can calculate the velocity $`\dot{X}=P(t)l(t)/(M_0l_0)`$ for the kink center. Since the momentum $`P(t)`$ is a periodic function and the kink width $`l(t)`$ at least is a bounded function, $`\dot{X}`$ is bounded as well. For instance, if we take $`\nu =1/2`$ in (32), the frequency of the external force
$$\delta =\mathrm{\Omega }_R\frac{q^2\mathrm{cos}(2\delta _0)}{4M_0^2\gamma _0^2\mathrm{\Omega }_R}ϵ^2+O(ϵ^4),$$
is very close to $`\mathrm{\Omega }_R\mathrm{\Omega }_i`$ if $`ϵ1`$ (see Fig. 1), and $`l(t)`$ is a function of the square of the $`4\pi /\delta `$-periodic solutions $`\mathrm{se}_{1/2}`$ and $`\mathrm{ce}_{1/2}`$, respectively, so the velocity of the kink center, $`\dot{X}(t)`$, and the energy $`E(t)`$ will also be bounded functions for $`\delta =\delta (ϵ)`$.
If we try to find, e.g., $`2\pi `$-periodic solutions of (24) we obtain that in this case the two independent solutions are related to the integer Mathieu functions $`\mathrm{se}_1`$ and $`\mathrm{ce}_1`$. However, these two solutions appear when the characteristic curve $`a(\theta )`$ is
$$a=1+\theta \frac{\theta ^2}{8}\frac{\theta ^3}{64}\frac{\theta ^4}{1536}+\mathrm{},$$
for the even Mathieu function $`\mathrm{ce}_1`$ and
$$a=1\theta \frac{\theta ^2}{8}+\frac{\theta ^3}{64}\frac{\theta ^4}{1536}+\mathrm{},$$
for the odd Mathieu function $`\mathrm{se}_1`$. Since the values of $`a`$ are different for each of the function $`\mathrm{se}_1`$ and $`\mathrm{ce}_1`$, these functions are not solutions of the same equation except when $`\theta =0`$. Although $`\mathrm{se}_1`$ and $`\mathrm{ce}_1`$ are not solutions of Eq. (24) \[recall that in order to find $`l(t)`$ we need to calculate two independent solutions of Eq. (24)\], the characteristic curves of these solutions separate the unstable and stable solutions of the Mathieu equation (24), and can be rewritten as
$$\begin{array}{c}\delta _{}=\frac{\mathrm{\Omega }_R}{2}\left[\frac{q^2\mathrm{cos}(2\delta _0)}{2M_0^2\gamma _0^2\mathrm{\Omega }_R}+\frac{q^2}{4M_0^2\gamma _0^2\mathrm{\Omega }_R}\right]ϵ^2,\hfill \end{array}$$
(33)
and
$$\begin{array}{c}\delta _+=\frac{\mathrm{\Omega }_R}{2}\left[\frac{q^2\mathrm{cos}(2\delta _0)}{2M_0^2\gamma _0^2\mathrm{\Omega }_R}\frac{q^2}{4M_0^2\gamma _0^2\mathrm{\Omega }_R}\right]ϵ^2,\hfill \end{array}$$
(34)
respectively. Interestingly, we note that the width of the unbounded region,
$$\mathrm{\Delta }\delta \delta _+\delta _{}=\frac{q^2ϵ^2}{2M_0^2\gamma _0^2\mathrm{\Omega }_R},$$
decreases when the initial velocity increases, and consequently we will focus on the case $`u(0)=0`$ in our numerical simulations below, as the resonance is then easier to observe. The curves $`\delta _+`$ and $`\delta _{}`$ \[Eqs. (33) and (34)\] are plotted in Fig. 1 for zero initial velocity and $`\delta _0=\pi /2`$, where the shadowed region represents the region where $`l(t)`$ is unbounded. Analogously, we can obtain other characteristic curves, related either with integer Mathieu functions $`\mathrm{se}_n`$ and $`\mathrm{ce}_n`$ ($`n𝖭`$), or with non-integer Mathieu functions $`\mathrm{se}_\nu `$ and $`\mathrm{ce}_\nu `$. In view of this, the above analytical results lead us to expect that if $`\delta `$ is close to $`\mathrm{\Omega }_R\mathrm{\Omega }_i`$; $`l(t)`$, $`u(t)`$ and the energy should be oscillatory (or, at least, bounded) functions, whereas if $`\delta \mathrm{\Omega }_R/2\mathrm{\Omega }_i/2`$ the kink width $`l(t)`$ increases indefinitely and since the energy and the velocity are proportional to $`l(t)`$ \[see Eqs. (19) and (17)\] we should observe that these functions increase with time as well.
### B $`\lambda 0`$ : Resonances at $`\delta \mathrm{\Omega }_R/2`$, $`\mathrm{\Omega }_R`$
When $`\lambda 0`$, the linear part of Eq. (22) becomes, in a manner similar to the procedure leading to Eq. (24), a more general Mathieu equation , namely
$$\begin{array}{c}g^{\prime \prime }+\left[b+2\theta _1ϵ^2\mathrm{cos}(2\tau )+2\theta _2ϵ^2\mathrm{cos}(4\tau )\right]g=0,\hfill \end{array}$$
(35)
where
$$b=\left(\frac{\mathrm{\Omega }}{\delta }\right)^2\left[1+\frac{q^2ϵ^2}{2\delta ^2M_0^2}+\frac{\lambda ^2}{M_0^2}\right],$$
$$\theta _1=\left(\frac{\mathrm{\Omega }}{\delta }\right)^2\frac{\lambda q^2}{\delta ^2M_0^2},\theta _2=\left(\frac{\mathrm{\Omega }}{\delta }\right)^2\frac{q^2}{4\delta ^2M_0^2},$$
and the prime denotes the derivative with respect to $`\tau =(\delta t+\delta _0)/2`$. According to Floquet theory , Eq. (35) has normal solutions of the form $`g=\mathrm{exp}(\sigma \tau )\mathrm{\Phi }(\tau )`$, where $`\mathrm{\Phi }(\tau )`$ is a $`\pi `$-periodic function and $`\sigma `$ is the characteristic exponent. Expanding $`\mathrm{\Phi }(\tau )`$ in a Fourier series, the function $`g`$ can be rewritten as $`g=_{n=\mathrm{}}^+\mathrm{}\mathrm{\Phi }_n\mathrm{exp}[(\sigma +2ni)\tau ]`$, where $`\mathrm{\Phi }_n`$ are the coefficients of the above series. Therefore, the transition curves separating stability from instability correspond to $`\sigma =0`$ ($`\pi `$-periodic solutions) and $`\sigma =i`$ ($`2\pi `$-periodic solutions) . Now we will apply the method of the strained parameters to determine the values of $`b`$, $`\theta _1`$ and $`\theta _2`$ corresponding to these values of the characteristic exponent. First we assume the solutions of (35), which have period $`\pi `$ or $`2\pi `$ and the transition curves $`b=b(ϵ)`$, can be written in the form of perturbation expansions as
$$\begin{array}{c}g(\tau )=g_0+ϵ^2g_1+ϵ^4g_2+O(ϵ^6),\hfill \\ b=b_0+ϵ^2b_1+ϵ^4b_2+O(ϵ^6).\hfill \end{array}$$
(36)
Second, substituting (36) into (35) and equating to zero the coefficients of the same order of $`ϵ`$ we obtain
$$\begin{array}{c}g_0=0,\hfill \end{array}$$
(37)
$$\begin{array}{c}g_1=b_1g_02\theta _1\mathrm{cos}(2\tau )g_02\theta _2\mathrm{cos}(4\tau )g_0,\hfill \end{array}$$
(38)
$$\begin{array}{c}g_2=b_1g_1b_2g_02\theta _1\mathrm{cos}(2\tau )g_12\theta _2\mathrm{cos}(4\tau )g_1,\hfill \end{array}$$
(39)
where $``$ represent the second-order linear differential operator $`={\displaystyle \frac{d^2}{d\tau ^2}}+b_0`$. The solution of Eq. (37) is
$$\begin{array}{c}g_0=A\mathrm{cos}\left(\sqrt{b_0}\tau \right)+B\mathrm{sin}\left(\sqrt{b_0}\tau \right),\hfill \end{array}$$
(40)
where $`b_0=4n^2`$ for the $`\pi `$-periodic solutions and $`b_0=(2n1)^2`$ for the $`2\pi `$-periodic solutions, $`n`$ is an integer number and $`A`$ and $`B`$ are constants.
For the $`\pi `$-periodic solutions we will analyze the case $`n=1`$, corresponding to $`b_0=4`$; we note that the case $`n=0`$ is not possible due to the definition of our parameter b. When $`b_0=4`$, $`g_0=A\mathrm{cos}(2\tau )+B\mathrm{sin}(2\tau )`$. For $`g_1`$ to be a $`\pi `$-periodic function is necessary to eliminate the secular terms in Eq. (38), imposing either $`b_1=\theta _2`$ and $`A=0`$ or $`b_1=\theta _2`$ and $`B=0`$. Then, the transition curves corresponding to the solutions
$$\begin{array}{c}g_+=B\mathrm{sin}(2\tau )+Bϵ^2\left[\frac{\theta _1}{20}\mathrm{sin}(4\tau )+\frac{\theta _2}{32}\mathrm{sin}(6\tau )\right],\hfill \end{array}$$
(41)
and
$$\begin{array}{c}g_{}=A\mathrm{cos}(2\tau )+Aϵ^2\left[\frac{\theta _1}{4}+\frac{\theta _1}{12}\mathrm{cos}(4\tau )+\frac{\theta _2}{32}\mathrm{cos}(6\tau )\right],\hfill \end{array}$$
(42)
can be shown to be
$$b=4\pm ϵ^2\theta _2,$$
or, alternatively,
$`\delta _+`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Omega }_R}{2}}{\displaystyle \frac{qu(0)\mathrm{cos}(\delta _0)}{M_0\gamma _0}}ϵ+\left[{\displaystyle \frac{3q^2}{4M_0^2\gamma _0^2\mathrm{\Omega }_R}}+{\displaystyle \frac{q^2\mathrm{cos}(2\delta _0)}{2M_0^2\gamma _0^2\mathrm{\Omega }_R}}\right]ϵ^2,`$ (43)
$`\delta _{}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Omega }_R}{2}}{\displaystyle \frac{qu(0)\mathrm{cos}(\delta _0)}{M_0\gamma _0}}ϵ+\left[{\displaystyle \frac{5q^2}{4M_0^2\gamma _0^2\mathrm{\Omega }_R}}+{\displaystyle \frac{q^2\mathrm{cos}(2\delta _0)}{2M_0^2\gamma _0^2\mathrm{\Omega }_R}}\right]ϵ^2.`$ (44)
respectively. Since these curves start from $`ϵ=0`$ and $`b=4`$, we are analyzing the transition curves when the frequency of the ac force is close to half of $`\mathrm{\Omega }_R`$. Hence, the resonance at $`\delta \mathrm{\Omega }_i/2\mathrm{\Omega }_R/2`$ appears much as in the case $`\lambda =0`$ even when $`\lambda 0`$. Let us remark that, for given $`ϵ`$, the distance between these two curves $`\delta =\delta (ϵ)`$ is
$$\frac{q^2ϵ^2}{2M_0^2\gamma _0^2\mathrm{\Omega }_R},$$
and when the velocity increases, the unstable region is more narrow, as we found above in the case when $`\lambda =0`$.
The other transition curves are related to the $`2\pi `$-periodic solution of Eq. (37), $`g_0=A\mathrm{cos}(\tau )+B\mathrm{sin}(\tau )`$, which appear when $`b_0=1`$ ($`n=1`$). In this case $`g_1`$ will be a periodic function if $`b_1=\theta _1`$ and $`A=0`$ or $`b_1=\theta _1`$ and $`B=0`$. Hence, the transition curves starting from $`b=1`$ are
$$b=1\pm ϵ^2\theta _1,$$
or, equivalently,
$$\begin{array}{c}\delta _{}=\mathrm{\Omega }_R\frac{qu(0)\mathrm{cos}(\delta _0)}{M_0\gamma _0}ϵ+\left[\frac{q^2}{2M_0^2\gamma _0^2\mathrm{\Omega }_R}+\frac{q^2\mathrm{cos}(2\delta _0)}{4M_0^2\gamma _0^2\mathrm{\Omega }_R}\right]ϵ^2\frac{q^2\lambda }{2M_0^2\gamma _0^2\mathrm{\Omega }_R}ϵ^2,\hfill \end{array}$$
(45)
and correspond to the solutions
$$\begin{array}{c}g_{}=B\mathrm{sin}(\tau )+Bϵ^2\left[\frac{\theta _1\theta _2}{8}\mathrm{sin}(3\tau )+\frac{\theta _2}{24}\mathrm{sin}(5\tau )\right],\hfill \end{array}$$
(46)
$$\begin{array}{c}g_+=A\mathrm{cos}(\tau )+Aϵ^2\left[\frac{\theta _1+\theta _2}{8}\mathrm{cos}(3\tau )+\frac{\theta _2}{24}\mathrm{cos}(5\tau )\right],\hfill \end{array}$$
(47)
respectively. These solutions and its corresponding characteristic curves are related with the driving frequency $`\delta `$ close to $`\mathrm{\Omega }_R`$. Finally, in this case we have
$$\mathrm{\Delta }\delta \delta _+\delta _{}=\frac{q^2\lambda ϵ^2}{M_0^2\gamma _0^2\mathrm{\Omega }_R},$$
which means as above that for fixed $`\lambda `$, $`\mathrm{\Delta }\delta `$ decreases when $`u(0)`$ is increased.
In summary, we have found here that when $`\lambda 0`$ and we drive the system with a frequency close to $`\mathrm{\Omega }_R`$ the solutions will be unstable in the region between the curves $`\delta _+`$ and $`\delta _{}`$. Notice that for $`\lambda =0`$ we recover the previous results, i.e., there are not any resonances at $`\delta \mathrm{\Omega }_R`$.
## IV Numerical verification
The results we have obtained in the previous section have been derived within the collective coordinate assumption that we can describe all the kink dynamics by the two variables $`X(t)`$ and $`l(t)`$, all other degrees of freedom being negligible. As there is no way to know a priori that this is indeed the case, we have to verify this hypothesis by numerical simulations of the full partial differential equation. Furthermore, we have not been able to solve the collective coordinate equations for the damped ($`\beta 0`$) case; it is reasonable to expect that in this situation phenomena similar to those found for the undamped case will arise, but in so far this is not checked this assertion remains a conjecture.
In view of the above considerations, we have computed the numerical solution of the partial differential equations (2) and (3) by using the Strauss-Vázquez scheme and choosing a total length for our numerical system of $`L=400`$, with steps $`\mathrm{\Delta }t=0.01`$, $`\mathrm{\Delta }x=0.1`$, free boundary conditions, and simulating up to a final time equal to $`\mathrm{25\hspace{0.17em}000}`$ with a kink at rest as initial condition. In addition, we have fixed the amplitude $`ϵ=0.01`$ and the phase $`\delta _0=\pi /2`$ of the ac force and then we have changed the value of the driving frequency $`\delta `$ in order to see the resonances. We note that there are many other parameters we could change, such as the initial velocity, the driving amplitude, or the driving phase, but as our main goal in this section is to assert the validity of the general analytical results obtained above we prefer to concentrate on a few cases as mentioned. In all our simulations we have chosen the initial parameters in such a way that $`\lambda =0`$, because when $`\lambda 0`$ the kink moves to the right or to the left and then for large times the kink will leave our finite system. For the aforesaid values of initial parameters the width of the resonance region $`\mathrm{\Delta }\delta =\delta _+\delta _{}`$ predicted at $`\delta \mathrm{\Omega }_R/2\mathrm{\Omega }_i/2`$ in the section III (A) \[see Eqs. (33) and (34)\], is of the order of $`10^4`$. For this reason we have explored the regions around $`\mathrm{\Omega }_i`$ and $`\mathrm{\Omega }_i/2`$ in an interval of that order. Finally, we want to mention that with the Strauss-Vázquez method one can compute very accurately the position and the velocity of the kink center using the integrals of energy and momentum , so it is a good numerical method in order to compare to our analytical predictions.
First of all, in order to verify the results obtained by means of the GTWA with two independent collective coordinates, the position and the width of the kink, we have studied the region around $`\delta \mathrm{\Omega }_i=1.2247`$, i.e., the driving frequency for which we do not expect any resonances when $`\lambda =0`$; hence, the width and the velocity should be, at least, bounded functions. We have verified that this prediction is in excellent agreement with the numerical simulations, an example of which is shown in Fig. 2. In this plot, we can see that the velocity is indeed an oscillatory function and even for large times our theoretical approach describes correctly the evolution of the velocity of the center of mass of the kink. Other values close to $`\delta =1.21`$ behave very much like the presented example.
The next test of our analytical results has to be, of course, the existence of a resonance near $`\mathrm{\Omega }_i/2`$. For frequencies $`\delta `$ around $`\mathrm{\Omega }_R/2`$, our collective variable approach has predicted that the width $`l(t)`$ increases unboundedly and hence the velocity and the energy should increase as well. Again, the prediction is fulfilled, but for $`\mathrm{\Omega }_i/2`$ instead of $`\mathrm{\Omega }_R/2`$: In the numerical simulations we have observed that in this case a resonance takes place when $`\delta \mathrm{\Omega }_i/2=0.6124`$, As a specific example, the results for the kink energy for $`\delta =0.6104`$ and $`\delta =0.608`$ are plotted in Fig. 3, clearly showing a resonant increase of the energy in the former case and oscillatory behavior in the latter. Interestingly, at the resonance the velocity also does not have an oscillatory behaviour as depicted in Fig. 4, in contrast to the behavior off-resonance we found in Fig. 2 for $`\delta =1.21`$. Such a non-periodic evolution of the kink velocity implies that the kink motion is chaotic as we first found in and, eventually, the kink begins emitting radiation. As we discussed in , we believe that this phenomenon has its origin in the energy transfer from the kink internal mode to the rest of modes in the system, a mechanism already demonstrated in ; intuitively, the reason for that is that the collective coordinate prediction that the kink width grows without limit can not be physically true due to the own nature of the kink, and hence, when the internal mode excitation reaches large values the energy ends up being transferred to the rest of the available modes (translation and radiation).
After checking our analytical results for the undamped case, we now have to turn to the damped ($`\beta 0`$) dynamics, in order to find out whether the same phenomena arise there. Simulations for the damped case show that the energy of the system also increases when the resonance takes place (see Fig. 5), but in this case the energy is bounded due to the dissipation, reaches an asymptotic value, and thus does not increase indefinitely as in the absence of damping. For comparison of both situations we have computed the mean energy in the time interval $`\mathrm{10\hspace{0.17em}000}<t<\mathrm{25\hspace{0.17em}000}`$ with and without damping, and our results are summarized in Figs. 6 and 7. For $`\beta =0.001`$ (see points on the solid line in Fig. 6) we have found that the energy increases at $`\delta =0.6103`$, whereas for $`\beta =0`$ (see the points on the dashed line in Fig. 6), the resonance frequency is $`\delta 0.6102`$.
As mentioned above, the numerical solutions of the full partial differential equation show that the resonance at $`\delta \mathrm{\Omega }_i/2`$ also takes place for $`\beta 0`$, whereas when $`\delta \mathrm{\Omega }_i`$ (see Fig. 7) the resonance disappears. Let us recall that in all our simulations we have chosen $`u(0)=0`$ (and hence the resonant region should be the widest possible one) and $`\delta _0=\pi /2`$, in such a way that $`\lambda =0`$. In the case when $`\lambda `$ vanishes we can not expect any resonances at $`\delta \mathrm{\Omega }_i`$. Nevertheless, in Fig. 7 it is clear that the energy increases weakly in this region. We believe that this maximum of the energy is probably due to the fact that the condition for the suppression of the resonance at $`\delta \mathrm{\Omega }_i`$ is extremely difficult to fulfill numerically. On the other hand, it is also possible that this condition, which has been obtained within the collective coordinate framework, is only approximately true, and therefore when we choose $`\lambda =0`$ we are beginning with an initial condition close to the one needed for complete suppression of the resonance but not exactly there, and hence we would still see the small bump in Fig. 7. Another possible reason for this peak is that it is possible that when $`\lambda =0`$ at $`\delta \mathrm{\Omega }_i\mathrm{\Omega }_R`$ a resonance appears in the next order corrections (see the work of Segur , in which the frequency $`2\mathrm{\Omega }_i`$ arises, and could then be parametrically excited by $`\delta =\mathrm{\Omega }_i`$). Were this the case, this “weaker resonance”, that completely disappears for $`\beta 0`$ \[see the solid line in Fig. 7\], is the resonance that we can expect either when $`\lambda 0`$ in the partial differential equation or for $`\lambda =0`$ and large enough times.
## V Conclusions
We have shown analytically and numerically that the internal mode of the $`\varphi ^4`$ model can be excited if we drive the system with an ac force of a frequency close to $`1/2`$ of the internal mode frequency. This is a very surprising result as the driving we are applying to the system is not parametric. At resonance, as a consequence of the increment of the energy of the system, the kink initially at rest begins to move chaotically and also begins to radiate, i.e., the energy is transferred from the internal mode to the translational and radiational ones. The chaotic motion is confirmed by noting that, when the internal mode is excited at $`\delta \mathrm{\Omega }_i/2`$, in the discrete Fourier transform (DFT) of the kink’s velocity \[see Fig. 4\] we see some frequencies in the low frequency part of the spectrum aside from the frequency of the ac force, $`\delta `$. What is more important, we have presented a full analysis of the collective coordinate theory for this problem, whose validity has been undoubtedly confirmed by numerical simulations of the full partial differential equation. In particular, in the absence of damping we have been able to find analytically all the resonances as well as their dependence on the kink initial conditions. On the other hand, we have also shown that the resonance at $`\delta \mathrm{\Omega }_i`$ can be suppressed if we add a small damping to the system or by an appropriate choice of initial parameters of the ac force. We have verified in our simulations that for $`\delta \mathrm{\Omega }_i`$ the lower radiational modes are excited and for this reason the energy in Fig. 7 increases when $`\delta `$ increases. Of course, this excitation of the lowest phonons cannot be explained within the present collective coordinate theory; a much more involved approach including phonon effects (as in for the sG model) would probably account for that resonance.
Beyond the application of our results to the $`\varphi ^4`$ kink dynamics, we believe that the same phenomenology will arise in other systems for which internal modes are present as, for example, in the double sG equation . The analysis we have presented can be straightforwardly extended to this system, and it is quite likely that similar resonance effects will arise, as well as in other models with the same feature of internal modes. A related interesting question concerns the applicability of these results to the sG equation. As we mentioned in the introduction, our calculation applies directly to the sG equation, but in this system the kinks do not posses internal modes. Therefore, the identification $`\mathrm{\Omega }_i\mathrm{\Omega }_R`$ we have done for the $`\varphi ^4`$ kink is not available any more, and, what is worse, the putative $`\mathrm{\Omega }_R`$ lies within the phonon band. However, it has been reported that close to $`\mathrm{\Omega }_R`$ sG kinks might exhibit a quasi-mode with a very long life. If such a mode indeed exists, which is still in doubt since no other reports of its existence have been published, it could be found by looking at resonances, as we have done in this paper, at half its frequency, which would be outside the phonon band. Work along these lines is in progress .
## Acknowledgments
We thank Yuri Gaididei, Francisco Domínguez-Adame, and José Cuesta for discussions. Work at GISC (Leganés) has been supported by DGESIC (Spain) grant PB96-0119. Travel between Bayreuth and Madrid has been supported by “Acciones Integradas Hispano-Alemanas”, a joint program of DAAD (Az. 314-AI) and DGESIC.
## A Equivalence of collective coordinate methods
In this appendix we demostrate that the GTWA, the method we used in section II to obtain the collective coordinate equations, is equivalent to using the variation of the momentum and the energy of the system (1). First of all, from Eqs. (7) and (10) we can obtain the variation of the momentum and the energy of the perturbed $`\varphi ^4`$ equation. Combining the first three terms of Eq. (7) and doing some straightforward transformations the equation for the momentum can be recast in the form
$$\begin{array}{c}\frac{dP}{dt}=_{\mathrm{}}^+\mathrm{}𝑑x\frac{}{X}+_{\mathrm{}}^+\mathrm{}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{})\frac{\varphi }{X},\hfill \\ \\ P(t)=_{\mathrm{}}^+\mathrm{}𝑑x\varphi _x\varphi _t.\hfill \end{array}$$
(A1)
Furthermore, if instead of Eq. (10) we add Eq. (7) rewritten as
$`{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x{\displaystyle \frac{\varphi }{X}}{\displaystyle \frac{\psi }{\dot{X}}}\ddot{X}+{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x[\varphi ,\psi ]\dot{l}+{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x{\displaystyle \frac{\varphi }{X}}{\displaystyle \frac{\psi }{\dot{l}}}\ddot{l}`$ (A2)
$`F^{stat}(X){\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{}){\displaystyle \frac{\varphi }{X}}=0,`$ (A3)
multiply this equation by $`\dot{X}`$, and the expression obtained from (10) by $`\dot{l}`$, we conclude that Eq. (10) is equivalent to
$`{\displaystyle \frac{dH}{dt}}`$ $`=`$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{}){\displaystyle \frac{d\varphi }{dt}},`$ (A4)
$`H(t)`$ $`=`$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left\{{\displaystyle \frac{\varphi _x^2}{2}}+{\displaystyle \frac{\varphi _t^2}{2}}+U(\varphi )\right\}.`$ (A6)
This implies that when we apply the GTWA in those systems we are, in fact, varying the momentum
$`{\displaystyle \frac{dP}{dt}}`$ $`=`$ $`{\displaystyle \frac{d}{dt}}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\varphi _x\varphi _t={\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left[\varphi _{xt}\varphi _t\varphi _x\varphi _{tt}\right],`$ (A7)
and the energy
$`{\displaystyle \frac{dH}{dt}}`$ $`=`$ $`{\displaystyle \frac{d}{dt}}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left\{{\displaystyle \frac{1}{2}}\varphi _x^2+{\displaystyle \frac{1}{2}}\varphi _t^2+U(\varphi )\right\}=`$ (A8)
$`=`$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left\{\varphi _x\varphi _{xx}+\varphi _t\varphi _{tt}+U^{}(\varphi )\varphi _t\right\},`$ (A9)
where
$`\varphi (x,t)=\varphi [xX(t),l(t)],\varphi _t\psi (x,t)=\psi [xX(t),l(t),\dot{X}(t),\dot{l}(t)].`$ (A10)
Moreover, we can also start from Eqs. (A7), (A8) and (A10) and obtain (7) and (10). Doing that, we only need to take into account that there are at least two different ways of transforming the above integrals: in one of them one substitutes (A10) in (A7) and (A8), so that
$`{\displaystyle \frac{dP}{dt}}`$ $`=`$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left\{{\displaystyle \frac{\varphi }{X}}{\displaystyle \frac{\psi }{\dot{X}}}\ddot{X}+[\varphi ,\psi ]\dot{l}+{\displaystyle \frac{\varphi }{X}}{\displaystyle \frac{\psi }{\dot{l}}}\ddot{l}\right\},`$ (A11)
$`{\displaystyle \frac{dH}{dt}}`$ $`=`$ $`\dot{X}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left\{{\displaystyle \frac{}{X}}+{\displaystyle \frac{\varphi }{X}}{\displaystyle \frac{\psi }{\dot{X}}}\ddot{X}+[\varphi ,\psi ]\dot{l}+{\displaystyle \frac{\varphi }{X}}{\displaystyle \frac{\psi }{\dot{l}}}\ddot{l}\right\}+`$ (A13)
$`+`$ $`\dot{l}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x\left\{[\psi ,\varphi ]\dot{X}+{\displaystyle \frac{\varphi }{l}}{\displaystyle \frac{\psi }{\dot{X}}}\ddot{X}+{\displaystyle \frac{\varphi }{l}}{\displaystyle \frac{\psi }{\dot{l}}}\ddot{l}\right\},`$ (A14)
whereas, in the other one, we first substitute the systems of Eqs. (2)-(3) in (A7) and (A8) and then assume that $`\varphi (x,t)`$ and $`\psi (x,t)`$ satisfy (A10). Thus,
$`{\displaystyle \frac{dP}{dt}}={\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x{\displaystyle \frac{}{X}}+{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{}){\displaystyle \frac{\varphi }{X}},`$ (A15)
(A16)
$`{\displaystyle \frac{dH}{dt}}=\dot{X}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{}){\displaystyle \frac{\varphi }{X}}+\dot{l}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xF(x,t,\varphi ,\varphi _t,\mathrm{}){\displaystyle \frac{\varphi }{l}}.`$ (A17)
Equating (A11) and (A15), and (A14) and (A17), respectively, we find that (A7) and (A8) become (7) and (10), respectively.
Let us note that the connection between the GTWA and the the variation of the momentum and energy gives a physical interpretation of such a technique, and furthermore, this second method leads more directly to formulas for $`X(t)`$ and $`l(t)`$. Furthermore, if the solution of Eqs. (2)-(3) for $`F(x,t,\varphi ,\varphi _t,\mathrm{})=\beta \varphi _t+f(x,t)`$ is a kink centered at $`X(t)`$ whose width, $`l(t)`$, depends on time, (5)-(6), then Eqs. (A7) and (A8) become
$`{\displaystyle \frac{dP}{dt}}`$ $`=`$ $`\beta P{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xf(x,t)\varphi _x,P(t)={\displaystyle \frac{M_0l_0\dot{X}(t)}{l(t)}};`$ (A18)
and
$`{\displaystyle \frac{P(t)l(t)}{M_0l_0}}\left[{\displaystyle \frac{dP}{dt}}+\beta P(t)+{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑xf(x,t)\varphi _x\right]+`$ (A19)
$`+\dot{l}[{\displaystyle \frac{P(t)^2}{2M_0l_0}}+{\displaystyle \frac{1}{2}}\alpha M_0l_0({\displaystyle \frac{2\ddot{l}}{l}}{\displaystyle \frac{\dot{l}^2}{l^2}})+{\displaystyle \frac{1}{2}}M_0({\displaystyle \frac{1}{l_0}}{\displaystyle \frac{l_0}{l^2}})+`$ (A20)
$`+\beta \alpha {\displaystyle \frac{l_0}{l}}\dot{l}+{\displaystyle _{\mathrm{}}^+\mathrm{}}d\theta f(l\theta +X(t),t)\theta \varphi _\theta ]=0,`$ (A21)
respectively. The first bracket on the right hand of (A19) vanishes because of (A18), so, the solution of (A19) is $`\dot{l}=0`$ or
$`\alpha [\dot{l}^22l\ddot{l}2\beta l\dot{l}]`$ $`=`$ $`{\displaystyle \frac{l^2}{l_0^2}}\left[1+{\displaystyle \frac{P^2}{M_0^2}}\right]1+`$ (A22)
$`+`$ $`{\displaystyle \frac{2l(t)^2}{M_0l_0}}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑\theta f[\theta l+X(t),t]\theta \varphi _\theta .`$ (A23)
In such a way, (A18) and (A22) can be interpreted as the equations which the collective coordinates, $`X(t)`$ \[$`P(t)`$\] and $`l(t)`$, satisfy, and they are obtained from $`dP/dt`$ and $`dH/dt`$, respectively. For example, if $`f(x,t)=ϵ\mathrm{sin}(\delta t+\delta _0)`$ in (A18) and (A22), we recover the Eqs. (18) and (20), obtained by applying the GTWA in section II.
As a final remark, we point out that, since we have not used in any moment the explicit form for the potential function $`U(\varphi )`$, the equivalence between these two aforesaid methods remains true for any other perturbed nonlinear Klein-Gordon equation of the form (2)-(3), when the Hamiltonian of the system is of the form (4).
|
warning/0006/quant-ph0006052.html
|
ar5iv
|
text
|
# Dynamical description of the buildup process in resonant tunneling: evidence of exponential and non-exponential contributions
\[
## Abstract
The buildup process of the probability density inside the quantum well of a double-barrier resonant structure is studied by considering the analytic solution of the time dependent Schrödinger equation with the initial condition of a cutoff plane wave. For one level systems at resonance condition we show that the buildup of the probability density obeys a simple charging up law, $`\left|\mathrm{\Psi }\left(\tau \right)/\varphi \right|=1e^{\tau /\tau _0},`$ where $`\varphi `$ is the stationary wave function and the transient time constant $`\tau _0`$ is exactly two lifetimes. We illustrate that the above formula holds both for symmetrical and asymmetrical potential profiles with typical parameters, and even for incidence at different resonance energies. Theoretical evidence of a crossover to non-exponential buildup is also discussed.
\]
Since the pioneering work of Esaki and Tsu, the tunneling in one-dimensional semiconductor heterostructures has been the subject of intense investigations . Resonant tunneling in double barrier (DB) systems has received special attention both by its technological applications and by the motivation to clarify the new interesting transport phenomena. Among the fundamental problems that have appeared on the scene, the charge buildup in the quantum well region is considered as one of the most important processes since it governs the ultimate speed of resonant tunneling devices . The need of a direct and comprehensive dynamical study of this phenomenon has been widely recognized . However, up to now we are lacking of an exact description of the buildup process itself.
In this paper we provide an exact description of the buildup process at resonance condition, within the framework of the shutter model. This is based on a full quantum dynamical approach, recently developed by García-Calderón and Rubio , that deals with the solution of the time dependent Schrödinger equation for an arbitrary potential $`V(x)`$ $`(0<x<L)`$, with an initial condition of a cutoff plane wave confined in the half-space $`x<0`$ to the left of an absorbing shutter at $`x=0`$. The sudden opening of the shutter at $`t=0`$ allows the wave function to interact with the potential. As a consequence, they found that the transient solution for the internal region may be written as the stationary solution modulated by a time varying Moshinsky function plus an infinite sum of transient resonance terms associated with the $`S`$matrix poles of the problem.
For the case of a reflecting shutter
$$\mathrm{\Psi }\left(x,k;t=0\right)=\{\begin{array}{cc}e^{ikx}e^{ikx}& \mathrm{}<x0\\ 0& x>0,\end{array}$$
(1)
one can proceed along the lines similar to that discussed in Ref. and obtain the solution for the internal region,
$`\mathrm{\Psi }(x,k;t)`$ $`=`$ $`\varphi (x,k)M(0,k;t)\varphi ^{}(x,k)M(0,k;t)`$ (3)
$`i{\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}T_nM(0,k_n;t),`$
where $`\varphi (x,k)`$ is the stationary wave function and the factors $`T_n=2ku_n(0)u_n(x)/(k^2k_n^2)`$ are given in terms of the resonant eigenfunctions $`u_n(x)`$. The index $`n`$ runs over the complex poles $`k_n`$ distributed in the third and fourth quadrants in the complex $`k`$-plane. The Moshinsky functions , as it is well known, are defined in terms of the complex error function $`w(z)`$ :$`M(y_q)M(0,q;t)=w(iy_q)/2`$, where the argument $`y_q`$ is given by
$$y_q=e^{i\pi /4}\left(\frac{m}{2\mathrm{}}\right)^{1/2}\left[\frac{\mathrm{}q}{m}t^{1/2}\right],$$
(4)
and $`q`$ stands either for $`\pm k`$ or $`k_{\pm n}`$.
The time evolution of wavefunction $`\mathrm{\Psi }(x,k;t)`$ in the internal region may be described by the expression given by Eq. (3), which involves the contribution of the full resonant spectrum of the system. For structures with typical parameters, and incidence energies $`E=\mathrm{}^2k^2/2m`$ near a resonance energy $`\epsilon _n`$, García-Calderón and Rubio showed that the single resonance approximation for the wavefunction $`\mathrm{\Psi }(x,k;t)`$ is valid from a few tenths of the corresponding lifetime onwards. This is the case for the present study, since we are not considering the regime of very short times ($`t\tau _n=\mathrm{}/\mathrm{\Gamma }_n`$), which may require the contribution of far away resonances. The single resonance approximation to Eq. (3) is
$`\mathrm{\Psi }(x,k;t)`$ $`=`$ $`\varphi (x,k)M(0,k;t)\varphi ^{}(x,k)M(0,k;t)`$ (6)
$`iT_nM(0,k_n;t)iT_nM(0,k_n^{};t),`$
where we have used the fact that the poles in the third quadrant $`k_n`$ are related to those of the fourth, $`k_n`$, by $`k_n=k_n^{}`$. This is the one-level expression of the time dependent wave function for the description of the dynamics in the internal region.
In order to exemplify the building up of the probability density in the quantum well for incidence at different resonance energies and for different potential profiles, let us consider the following two numerical examples. The first case corresponds to the symmetrical DB structure with parameters: barrier heights $`V_0=0.5`$ $`eV,`$ barrier widths $`b_0=30`$ Å and well width $`\omega _0=100`$ Å. The resonance parameters for the first three resonant states are: $`\epsilon _1=37.8`$ $`meV`$, $`\mathrm{\Gamma }_1=0.12`$ $`meV`$; $`\epsilon _2=149.2`$ $`meV`$, $`\mathrm{\Gamma }_2=1.40`$ $`meV`$; $`\epsilon _3=325.7`$ $`meV`$, $`\mathrm{\Gamma }_3=8.60`$ $`meV`$. We show in Fig. 1 the time evolution of the probability density calculated by Eq. (6) for incidence at resonance, $`E=\epsilon _n`$ and fixed position $`x`$ (we have considered values of $`x`$ near the maxima of $`|\varphi (x,k)|^2`$ as the most natural choice) for the cases: $`n=1`$, $`x=80`$ Å (solid line); $`n=2`$, $`x=48`$ Å (dashed line); and $`n=3`$, $`x=80`$ Å (dotted line). The second example consists of an asymmetrical DB structure with parameters: barrier heights $`V_1=V_2=0.3`$ $`eV,`$ barrier widths $`b_1=30`$ Å and $`b_2=100`$ Å, and well width $`\omega _0=50`$ Å. The resonance parameters for the first resonant state are $`\epsilon _1=89.1`$ $`meV`$ and $`\mathrm{\Gamma }_1=2.4`$ $`meV`$. The time evolution of $`|\mathrm{\Psi }(x,k;t)|^2`$ for incidence at $`E=\epsilon _1`$ and $`x=55`$ Å is also depicted in Fig. 1 (dashed-dotted line). For all cases, the probability density $`|\mathrm{\Psi }(x,k;t)|^2`$ grows up monotonically towards its asymptotic value. We can see that both the level off and the rate of increase of the curves are quite different. However, a common feature, not evident in Fig. 1, can be appreciated if we replot the normalized probability density $`|\mathrm{\Psi }(x,k;\tau )/\varphi (x,k)|^2`$ as a function of the new variable $`\tau `$, which is now the time given in lifetime units ($`t`$ replaced by $`\tau \mathrm{}/\mathrm{\Gamma }_n`$ for each curve). As a result, all four curves become indistinguishable among them as depicted in Fig. 2. We can see that the full establishment of the stationary situation is preceded by a transient in which the probability density is built up inside the quantum well with a unique characteristic curve. Thus, there must also exist a characteristic transient time constant $`\tau _0`$ that governs the buildup process, with the same value (in lifetimes) for all cases. The observed regularity and the fact that it holds for both symmetrical and asymmetrical cases and at different resonances, is a manifestation that the buildup process in one-level systems is governed by a simple law. In what follows we shall be concerned to find the analytic expression of the actual buildup law and the exact value of the characteristic transient time $`\tau _0`$.
For sharp ($`R_n\epsilon _n/\mathrm{\Gamma }_n1`$) and isolated resonances we know that the stationary wavefunction $`\varphi (x,k)`$ can be written as the one-term expression $`\varphi (x,k)=2iku_n\left(0\right)u_n\left(x\right)/(k^2k_n^2)`$, thus the factors $`iT_n`$ and $`iT_n`$ appearing in Eq. (6) can be identified as $`\varphi (x,k)`$ and $`\varphi ^{}(x,k)`$ respectively. By expressing formula (4) in lifetime units, it can be seen that $`y_q`$ depends only on the ratio $`R_n=\epsilon _n/\mathrm{\Gamma }_n`$, and not on the particular values of the resonance parameters $`\epsilon _n`$ and $`\mathrm{\Gamma }_n`$. For $`q=\pm k_n`$ and $`q=\pm k_n^{}`$, $`y_q`$ reads: $`y_{\pm k_n}=e^{i\pi /4}\left[\left(R_ni/2\right)\tau \right]^{1/2}`$, and $`y_{\pm k_n^{}}=e^{i\pi /4}\left[\left(R_n+i/2\right)\tau \right]^{1/2}`$, respectively. For the cases $`q=\pm k`$ we have $`y_{\pm k}=e^{i\pi /4}\left[R_n\tau \right]^{1/2}`$ (since $`E=\epsilon _n`$). From the above considerations and the well known symmetry relation , $`M(y_q)=e^{y_q^2}M(y_q),`$ applied to the Moshinsky functions $`M(y_k)`$ and $`M(y_{k_n})`$, we obtain a convenient representation for the probability density,
$$\left|\mathrm{\Psi }(x,k;\tau )\right|^2=\left|\varphi (x,k)\right|^2(1e^{\tau /2})^2+\mathrm{\Delta }(\tau ),$$
(7)
where $`\mathrm{\Delta }(\tau )`$ stands for the remaining terms, which involve the square modulus of the Moshinsky functions and several interference terms. It is not difficult to convince oneself that, for very large times, expression (7) possesses the correct asymptotic behavior, i. e. as $`\tau \mathrm{},`$ $`|\mathrm{\Psi }(x,k;\tau )|^2`$ goes into the stationary probability density $`|\varphi (x,k)|^2`$. This follows directly from the presence of the decreasing exponential $`e^{\tau /2},`$ and from the fact that each of the Moshinsky functions involved in $`\mathrm{\Delta }(\tau )`$ can be represented by a series expansion consisting of inverse powers of $`\tau `$. In Ref. it was shown that $`M(y_q)`$ has the asymptotic expansion $`M(y_q)=a_1/y_q+a_2/y_q^2+a_3/y_q^3+\mathrm{}`$ for large values of the variable $`y_q`$ provided that $`\pi /2<arg(y_q)<\pi /2`$. By inspection of our expressions for $`y_q`$, we can see that the above inequality holds for the three cases involved in $`\mathrm{\Delta }(\tau )`$ $`(q=k,`$ $`k_n`$, and $`k_n^{})`$, and as a consequence $`\mathrm{\Delta }(\tau )0`$ when $`\tau \mathrm{}`$, as expected. It is important not only that both $`e^{\tau /2}`$ and $`\mathrm{\Delta }(\tau )`$ go to zero, but also the fact that their rates of decrease are quite different, leading to important consequences on the nature of the buildup. In particular it reveals that there exist exponential and non-exponential contributions to the buildup mechanism, as we shall see later. In view of the series expansions considered above, $`\mathrm{\Delta }(\tau )`$ also contains inverse powers of $`y_q,`$ that is, inverse powers of the product $`(R_n\tau )^{1/2}`$. This means that $`\mathrm{\Delta }(\tau )`$ can be vanishingly small even for $`\tau `$ equal to a few lifetimes, provided that $`R_n1`$. In fact, there exists a finite time interval in which $`\mathrm{\Delta }(\tau )/|\varphi |^2`$ is negligible compared to $`e^{\tau /2}`$, leading to the following exponential buildup law:
$$|\mathrm{\Psi }\left(\tau \right)/\varphi |=1e^{\tau /2}.$$
(8)
This simple formula reproduces successfully the predicted values of expression (6). For comparison, we have included in Fig. 2 a plot of the normalized probability density calculated from (8). The corresponding curve is indistinguishable from all the other curves, showing in particular that $`\mathrm{\Delta }(\tau )`$ has an exceedingly small contribution in the relevant time interval for all of our numerical examples. We see that formula (8) does not depend explicitly on the potential profile parameters nor the resonant state, this explains why all the numerical examples illustrated in Fig. 2 share the same curve, despite the fact that they correspond to different situations. Note that in the exponential regime the buildup law becomes identical to the charging up law of a capacitor in an $`RC`$circuit: $`Q(\tau )/Q_0=1e^{\tau /\tau _C}`$, where $`Q_0`$ is the asymptotic charge, and $`\tau _C=RC`$ is the capacitive time constant. This is relevant, because we find in the literature capacitor-like models used to describe quantum tunneling properties in DB structures, such as the charge buildup and its implications on the speed limit on resonant tunneling devices . According to Eq. (8), the transient time constant $`\tau _0`$ of our “quantum capacitor” is always two lifetimes, and is a characteristic feature of one-level systems. On the experimental side, it is worth to mention that measured values of escape times of the order of $`2\mathrm{}/\mathrm{\Gamma }_n`$ has been reported by Sakaki et al. , arguing that, at coherence conditions, “the buildup time and the tunneling escape time are roughly the same” . An important remark is that the condition $`R_n1`$ is not so restrictive since it is satisfied for most of the resonant structures with typical parameters. In fact we carried out a systematical study (not shown here) and found that for values of $`R_n`$ from $`10`$ onwards this condition is satisfied.
In order to show the existence of deviations from the exponential regime, we shall examine the contributions arising from $`\mathrm{\Delta }(\tau )`$. The explicit calculation of $`\mathrm{\Delta }(\tau )`$ in terms of $`y_k`$, $`y_{k_n}`$, and $`y_{k_n^{}}`$ may result a too involved task, however, for the purpose of our discussion, it is sufficient to realize that the dominant term is proportional to an oscillatory function of $`\tau `$ modulated by the factor $`\tau ^{1/2}`$. We know that at very long times the exponential term goes to zero faster than $`\tau ^{1/2}`$, i. e. $`e^{\tau /2}\tau ^{1/2}`$. Therefore, there must exist a critical time $`\tau _{onset}`$ at which $`e^{\tau /2}`$ and $`\mathrm{\Delta }(\tau )/|\varphi |^2`$ are comparable. Such a critical time defines a crossover from the exponential to a non-exponential regime of the buildup process. In the examples depicted in Fig. 2 the non-exponential contributions to $`|\mathrm{\Psi }\left(\tau \right)/\varphi |^2`$ are overwhelmed and cannot be appreciated due to the scale of the graph. However, if we plot the logarithm of the difference $`\delta (\tau )=|1|\mathrm{\Psi }\left(\tau \right)/\varphi ||`$ versus $`\tau `$, \[using Eq. (6)\] the transition from the exponential to the non-exponential regime is clearly appreciated, see Fig. 3. In this figure, the exponential regime can be identified by the straight line with slope $`1/2`$, extending over a few lifetimes until it reaches the onset of the nonexponential buildup, $`\tau _{onset}`$, which depends on $`R_n`$.
It is interesting to note the similarity of the results depicted in Fig. 3 to the behavior of the survival probability found in studies of the phenomenon of quantum decay , which also exhibits this transition with an oscillatory structure at such crossover. We believe that this striking resemblance is not a simple coincidence but rather a manifestation of the existence of a more profound link between both phenomena. In fact, the survival probability may also be expressed in terms of the Moshinsky functions , which are, in our expressions, the key ingredients for the time evolution. These findings open up new questions about the common features in both processes, for example those about the existence of deviations from the exponential buildup also at early times, as it occurs in the decay process . Such analysis requires the contribution of far away resonances to the transient solution, and is deferred to future work .
Summarizing: (i) We have accomplished the first analytic derivation of the actual buildup law in resonant tunneling structures. It was based on general properties of the solution of the Schrödinger equation, without any assumptions on the potential profile, except that it is finite and support well defined resonances. (ii) We have shown the existence of both exponential and non-exponential contributions to the buildup process. (iii) The exponential regime is characterized by a transient time constant whose value is exactly two lifetimes. (iv) We have illustrated that formula (8) describes very accurately the exponential buildup for a great variety of situations: it works very well for different potential profiles, and is valid not only for the “ground state” $`\left(n=1\right)`$, but also for “excited states” $`\left(n>1\right)`$.
We thank G. García-Calderón for useful discussions.
|
warning/0006/hep-ph0006157.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
In the Standard Model of particle physics, neutrinos are strictly massless. Any evidence for neutrino masses would, therefore, imply physics beyond the Standard Model. Even though the direct experimental determination of a neutrino mass is (probably) far beyond the current experimental reach, experiments have been able to obtain indirect, and recently very strong, evidence for neutrino masses, via neutrino oscillations.
The key evidence for neutrino oscillations comes from the angular dependent flux of atmospheric muon-type neutrinos measured at SuperKamiokande , combined with a large deviation of the muon-type to electron-type neutrino flux ratio from theoretical predictions. This “atmospheric neutrino puzzle” is best solved by assuming that $`\nu _\mu `$ oscillates into $`\nu _\tau `$ and that the $`\nu _e`$ does not oscillate. For a recent analysis of all the atmospheric neutrino data see .
On the other hand, measurements of the solar neutrino flux have always been plagued by a large suppression of the measured solar $`\nu _e`$ flux with respect to theoretical predictions . Again, this “solar neutrino puzzle” is best resolved by assuming that $`\nu _e`$ oscillates into a linear combination of the other flavour eigenstates (for a more conservative analysis of the event rates and the inclusion of the “dark side” of the parameter space, see ). The most recent analysis of the solar neutrino data which includes the mixing of three active neutrino species can be found in .
Neutrino oscillations were first hypothesised by Bruno Pontecorvo in the 1950’s . The hypothesis of three flavour mixing was first raised by Maki, Nakagawa and Sakata . In light of the solar neutrino puzzle, Wolfenstein and Mikheyev and Smirnov realized that neutrino–matter interactions could affect in very radical ways the survival probability of electron-type neutrinos which are produced in the solar core and detected at the Earth (MSW effect).
Since then, significant effort has been devoted to understanding the oscillation probabilities of electron-type neutrinos produced in the Sun. For example, in the survival probability of solar electron-type neutrinos was discussed in the context of three-neutrino mixing including matter effects, and solutions to the solar neutrino puzzle in this context were studied (for example, in ).
In this paper, the understanding of solar neutrino oscillations is extend to the case of other active neutrino species ($`\nu _\mu `$, $`\nu _\tau `$, and antineutrinos) produced in the solar core. Even though only electron-type neutrinos are produced by the nuclear reactions which take place in the Sun’s innards, it is well know that, in a number of dark matter models, dark matter particles can be trapped gravitationally inside the Sun, and that the annihilation of these should yield a flux of high energy neutrinos ($`E_\nu \begin{array}{c}>\hfill \\ \hfill \end{array}1`$ GeV) of all species which may be detectable at the Earth . Indeed, this is one of the goals of very large “neutrino telescopes,” such as AMANDA or BAIKAL . It is important to understand how neutrino oscillations will affect the expected event rates at these experiments.<sup>*</sup><sup>*</sup>*Some effects have already been studied, in the two-neutrino case, in .
The oscillation probability of all neutrino species has, of course, been studied in different contexts, such as in the case of neutrinos produced in the core of supernovae or in the case of neutrinos propagating in constant electron number densities . The latter case has been receiving a considerable amount of attention from neutrino factory studies . The case at hand (GeV solar neutrinos) differs significantly from these mentioned above, in at least a few of the following: source-detector distance, electron number density average value and position dependency, energy average value and spectrum. Neutrino factory studies, for example, are interested in $`O`$(1000) km base-lines, $`O`$(10) GeV electron-type and muon-type neutrinos produced via muon decay propagating in roughly constant, Earth-like (matter densities around 3 g/cm<sup>3</sup>) electron number densities.
The paper is organised as follows. In Sec. 2, the well known case of two-flavour oscillations is reviewed in some detail, while special attention will be paid to neutrinos produced inside the Sun. In Sec. 3 the same discussion is extended to the less familiar case of three-flavour oscillations. Again, special attention is paid to neutrinos produced in the Sun’s core. In Sec. 4 the results presented in Sec. 3 will be analysed numerically, and the three-neutrino multi-dimensional parameter space will be explored. Sec. 5 contains a summary of the results and the conclusions.
It is important to comment at this point that one of the big challenges of studying three-flavour oscillations is the multi-dimensional parameter space, composed of three mixing angles, two mass-squared differences, and one complex phase, plus the neutrino energy. For this reason, the discussions presented here will take advantage of the current experimental situation to constrain the parameter space, and of the possibility of producing neutrinos of all species via dark matter annihilations to constrain the neutrino energies to the range from a few to tens of GeV.
## 2 Two-Flavour Oscillations
In this section, the well studied case of two-flavour oscillations will be reviewed . This is done in order to present the formalism which will be later extended to the case of three-flavour oscillations and describe general properties of neutrino oscillations and of neutrinos produced in the Sun’s core.
### 2.1 Generalities
Neutrino oscillations take place because, similar to what happens in the quark sector, neutrino weak eigenstates are different from neutrino mass eigenstates. The two sets are related by a unitary matrix, which is, in the case of two-flavour mixing, parametrised by one mixing angle $`\vartheta `$.If the neutrinos are Majorana particles, there is also a Majorana phase, which will be ignored throughout since it plays no role in the physics of neutrino oscillations.
$$\left(\begin{array}{c}\nu _e\\ \nu _x\end{array}\right)=\left(\begin{array}{cc}U_{e1}& U_{e2}\\ U_{x1}& U_{x2}\end{array}\right)\left(\begin{array}{c}\nu _1\\ \nu _2\end{array}\right)=\left(\begin{array}{cc}\mathrm{cos}\vartheta & \mathrm{sin}\vartheta \\ \mathrm{sin}\vartheta & \mathrm{cos}\vartheta \end{array}\right)\left(\begin{array}{c}\nu _1\\ \nu _2\end{array}\right),$$
(2.1)
where $`\nu _1`$ and $`\nu _2`$ are neutrino mass eigenstates with masses $`m_1`$ and $`m_2`$, respectively, and $`\nu _x`$ is the flavour eigenstate orthogonal to $`\nu _e`$. All physically distinguishable situations can be obtained if $`0\vartheta \pi /2`$ and $`m_1^2m_2^2`$ or $`0\vartheta \pi /4`$ and no constraint is imposed on the masses-squared.
In the case of oscillations in vacuum, it is trivial to compute the probability that a neutrino produced in a flavour state $`\alpha `$ is detected as a neutrino of flavour $`\beta `$, assuming that the neutrinos are ultrarelativistic and propagate with energy $`E_\nu `$:
$$P_{\alpha \beta }=|U_{\beta 1}|^2|U_{\alpha 1}|^2+|U_{\beta 2}|^2|U_{\alpha 2}|^2+2Re\left(U_{\beta 1}^{}U_{\beta 2}U_{\alpha 1}U_{\alpha 2}^{}e^{i\frac{\mathrm{\Delta }m^2x}{2E_\nu }}\right).$$
(2.2)
Here $`\mathrm{\Delta }m^2m_2^2m_1^2`$ is the mass-squared difference between the two mass eigenstates and $`x`$ is the distance from the detector to the source. It is trivial to note that $`P_{\alpha \beta }=P_{\beta \alpha }`$ since all $`U_{\alpha i}`$ are real and the theory is $`T`$-conserving. Furthermore, note that $`\vartheta `$ is indistinguishable from $`\pi /2\vartheta `$ (or, equivalently, the sign of $`\mathrm{\Delta }m^2`$ is not physical), and all physically distinguishable situations are obtained by allowing $`0\vartheta \pi /4`$ and choosing a fixed sign for $`\mathrm{\Delta }m^2`$.
In the case of nontrivial neutrino–medium interactions, the computation of $`P_{\alpha \beta }`$ can be rather involved. Assuming that the neutrino–medium interactions can be expressed in terms of an effective potential for the neutrino propagation, one has to solve
$$\frac{\mathrm{d}}{\mathrm{d}t}\left(\begin{array}{c}\nu _1(t)\\ \nu _2(t)\end{array}\right)=i\left[\left(\begin{array}{cc}E_1& 0\\ 0& E_2\end{array}\right)+\left(\begin{array}{cc}V_{11}(t)& V_{12}(t)\\ V_{12}(t)^{}& V_{22}(t)\end{array}\right)\right]\left(\begin{array}{c}\nu _1(t)\\ \nu _2(t)\end{array}\right),$$
(2.3)
with the appropriate boundary conditions (either a $`\nu _e`$ or a $`\nu _x`$ as the initial state, for example). In the ultrarelativistic limit one may approximate $`E_2E_1\mathrm{\Delta }m^2/2E_\nu `$, $`\mathrm{d}/\mathrm{d}t=\mathrm{d}/\mathrm{d}x`$, and $`V_{ij}(t)V_{ij}(x)`$. A very crucial assumption is that there is no kind of neutrino absorption due to the neutrino–medium interaction, i.e., the $`2\times 2`$ Hamiltonian for the neutrino system is Hermitian.
It is interesting to argue what can be said about $`P_{\alpha \beta }`$ in very general terms. First, the conservation of probability requires that
$`P_{ee}+P_{ex}`$ $`=`$ $`1,`$ (2.4)
$`P_{xe}+P_{xx}`$ $`=`$ $`1.`$ (2.5)
Second, given that the Hamiltonian evolution is unitary,
$$P_{ee}+P_{xe}=1.$$
(2.6)
It is easy to show that the extra constraint $`P_{ex}+P_{xx}=1`$ is redundant. Eq. (2.6) can be understood by the following “intuitive” argument: if the same amount of $`\nu _e`$ and $`\nu _x`$ is produced, independent of what happens to $`\nu _e`$ and $`\nu _x`$ during flight, the number of $`\nu _e`$ and $`\nu _x`$ detected in the end has to be the same. In light of the constraints above, one can show that there is only one independent $`P_{\alpha \beta }`$, which is normally chosen to be $`P_{ee}`$. The others are given by $`P_{ex}=P_{xe}=1P_{ee}`$ and $`P_{xx}=P_{ee}`$. Note that the equality $`P_{ex}=P_{xe}`$ is not a consequence of $`T`$-invariance, but a consequence of the unitarity of the Hamiltonian evolution and particular only to the two-flavour oscillation case, as will be shown later.
### 2.2 Oscillation of Neutrinos Produced in the Sun’s Core
It is well known that neutrino–Sun interactions affect the oscillation probabilities of neutrinos produced in the Sun’s core in very nontrivial ways. Indeed, all but one solution to the solar neutrino puzzle rely heavily on neutrino–Sun interactions . The survival probability of electron-type solar neutrinos has been computed in many different approximations by a number of people over the years, and can be understood in very simple terms .
In the presence of electrons, the differential equation satisfied by the two neutrino system is, in the flavour basis,
$$\frac{\mathrm{d}}{\mathrm{d}x}\left(\begin{array}{c}\nu _e(x)\\ \nu _x(x)\end{array}\right)=i\left[\frac{\mathrm{\Delta }m^2}{2E_\nu }\left(\begin{array}{cc}|U_{e2}|^2& U_{e2}^{}U_{\mu 2}\\ U_{e2}U_{\mu 2}^{}& |U_{\mu 2}|^2\end{array}\right)+\left(\begin{array}{cc}A(x)& 0\\ 0& 0\end{array}\right)\right]\left(\begin{array}{c}\nu _e(x)\\ \nu _x(x)\end{array}\right),$$
(2.7)
where terms proportional to the $`2\times 2`$ identity matrix were neglected, since they play no role in the physics of neutrino oscillations.
$$A(x)=\sqrt{2}G_FN_e(x)$$
(2.8)
is the charged current contribution to the $`\nu _e`$-$`e`$ forward scattering amplitude, $`G_F`$ is Fermi’s constant, and $`N_e(x)`$ is the position dependent electron number density. In the case of the Sun (see also ), $`AA(0)6\times 10^3`$ eV<sup>2</sup>/GeV, assuming an average core density of 79 g/cm<sup>3</sup>, and $`A(x)`$ falls roughly exponentially until close to the Sun’s edge. It is safe to say that significantly far away from the Sun’s edge $`A(x)`$ is zero.
A particularly simple way of understanding the propagation of electron-type neutrinos produced in the Sun’s core to the Earth is to start with a $`\nu _e`$ state in the basis of the eigenstates of the Hamiltonian evaluated at the production point, $`|\nu _e=cos\vartheta _M(0)|\nu _L+\mathrm{sin}\vartheta _M(0)|\nu _H`$, where $`|\nu _H`$ ($`|\nu _L`$) correspond to the highest (lowest) instantaneous Hamiltonian eigenstate. The matter mixing angle $`\vartheta _M\vartheta _M(0)`$ is given by
$$\mathrm{cos}2\vartheta _M=\frac{\mathrm{\Delta }m^2\mathrm{cos}2\vartheta 2E_\nu A}{\sqrt{(\mathrm{\Delta }m^2)^2+A^24E_\nu A\mathrm{\Delta }m^2\mathrm{cos}2\vartheta }}.$$
(2.9)
The evolution of this initial state from the Sun’s core is described by an arbitrary unitary matrix until the neutrino reaches the Sun’s edge. From this point on, one can rotate the state to the mass basis and follow the vacuum evolution of the state. Therefore, $`P_{ee}(x)`$, where $`x`$ is is the distance from the Sun’s edge to some point far away from the Sun (for example, the Earth), is
$$P_{ee}(x)=\left|\left(\begin{array}{cc}U_{e1}^{}& U_{e2}^{}\end{array}\right)\left(\begin{array}{cc}1& 0\\ 0& e^{i\frac{\mathrm{\Delta }m^2x}{2E_\nu }}\end{array}\right)\left(\begin{array}{cc}A& B\\ B^{}& A^{}\end{array}\right)\left(\begin{array}{c}\mathrm{cos}\vartheta _M\\ \mathrm{sin}\vartheta _M\end{array}\right)\right|^2,$$
(2.10)
where overall phases in the amplitude have been neglected. The matrix parametrised by $`A,B`$ represents the evolution of the system from the Sun’s core to vacuum, and also rotates the state into the mass basis.The most general form of a $`2\times 2`$ unitary matrix is $`\left(\begin{array}{cc}A& B\\ B^{}& A^{}\end{array}\right)\left(\begin{array}{cc}1& 0\\ 0& e^{i\zeta }\end{array}\right)`$, where $`|A|^2+|B|^2=1`$ and $`0\zeta 2\pi `$. In the case of neutrino oscillations, however, the physical quantities are $`|A|^2`$ and the phase of $`AB^{}`$, and therefore $`\zeta `$ can be ignored. Expanding Eq. (2.10), and assuming that there is no coherence in the Sun’s core between $`\nu _L`$ and $`\nu _H`$,<sup>§</sup><sup>§</sup>§This is in general the case, because one has to consider that neutrinos are produced at different points in space and time. one arrives at the well known expression (these have been first derived using a different language in and )
$$P_{ee}(x)=P_1\mathrm{cos}^2\vartheta +P_2\mathrm{sin}^2\vartheta \mathrm{cos}2\vartheta _M\sqrt{P_c(1P_c)}\mathrm{sin}2\vartheta \mathrm{cos}\left(\frac{\mathrm{\Delta }m^2x}{2E_\nu }+\delta \right),$$
(2.11)
where $`\delta `$ is the phase of $`AB^{}`$, $`P_c|B|^2=1|A|^2`$ is the “level crossing probability”, and $`P_1=1P_2=\frac{1}{2}+\frac{1}{2}\left(12P_c\right)\mathrm{cos}2\vartheta _M`$ is interpreted as the probability that the neutrino exits the Sun as a $`\nu _1`$.
Eq. (2.11) should be valid in all cases of interest, and contains a large amount of features. In the case of the solar neutrino puzzle, the neutrino energies of interest range between hundreds of keV to ten MeV, and matter effects start to play a role for values of $`\mathrm{\Delta }m^2`$ as high as $`10^4`$ eV<sup>2</sup>. In the adiabatic limit ($`P_c0`$) very small values of $`P_{ee}`$ are attainable when $`\mathrm{cos}2\vartheta _M1`$ and $`\mathrm{sin}^2\vartheta `$ is small. More generally, in this limit $`P_{ee}=\mathrm{sin}^2\vartheta `$. This is what happens for all solar neutrino energies in the case of the LOW solution,See for the labelling of the regions of the parameter space that solve the solar neutrino puzzle for solar neutrino energies above a few MeV in the case of the LMA solution, and for 400 keV $`\begin{array}{c}<\hfill \\ \hfill \end{array}E_\nu \begin{array}{c}<\hfill \\ \hfill \end{array}1`$ MeV energies in the case of the SMA solution. In the extremely nonadiabatic limit, which is reached when $`\mathrm{\Delta }m^2/2E_\nu A`$, $`P_c\mathrm{cos}^2\vartheta `$ and $`\mathrm{cos}2\vartheta _M1`$, the original vacuum oscillation expression is obtained, up to the “matter phase” $`\delta `$. This is generically what happens in the VAC solution to the solar neutrino puzzle.
If the electron number density is in fact exponential, one can solve Eq. (2.7) exactly . For $`N_e(x)=N_e(0)e^{x/r_0}`$, where $`x=0`$ is the centre of the Sun,
$$P_c=\frac{e^{\gamma \mathrm{sin}^2\vartheta }e^\gamma }{1e^\gamma },$$
(2.12)
where
$$\gamma =2\pi r_0\frac{\mathrm{\Delta }m^2}{2E_\nu }=1.05\left(\frac{\mathrm{\Delta }m^2}{10^6\mathrm{eV}^2}\right)\left(\frac{1\mathrm{GeV}}{E_\nu }\right),$$
(2.13)
for $`r_0=R_{}/10.54=6.60\times 10^4`$ km . In the case of the Sun, the exponential profile approximation has been examined , and was shown to be very accurate, especially if one allows $`r_0`$ to vary as a function of $`\mathrm{\Delta }m^2/2E_\nu `$.
The exact expression for $`\delta `$ has also been obtained , and the readers are referred to for details concerning physical implications of the matter phase. Its effects will not be discussed here any further.
### 2.3 The Case of Antineutrinos
Antineutrinos that are produced in the Sun’s core obey a differential equation similar to Eq. (2.7), except that the sign of the matter potential changes, i.e. $`A(x)A(x)`$, and $`U_{\alpha i}U_{\alpha i}^{}`$ (this is immaterial since, in the two-flavour mixing case, all $`U_{\alpha i}`$ are real).
Instead of working out the probability of an electron-type antineutrino being detected as an electron-type antineutrino $`P_{\overline{e}\overline{e}}`$ from scratch, there is a very simple way of relating it to $`P_{ee}`$. One only has to note that, if the following transformation is applied to Eq. (2.7): $`\vartheta \pi /2\vartheta `$, subtract the matrix $`A(1_{2\times 2})`$, where $`1_{2\times 2}`$ is the $`2\times 2`$ identity matrix and relabel $`\nu _e(x)\nu _x(x)`$, the equation of motion for antineutrinos is obtained.If one decides to limit $`0\vartheta \pi /4`$, a similar result can be obtained if $`\mathrm{\Delta }m^2\mathrm{\Delta }m^2`$, explicitly $`P_{\overline{e}\overline{e}}(\mathrm{\Delta }m^2)=P_{ee}(\mathrm{\Delta }m^2)`$. Therefore, $`P_{\overline{e}\overline{e}}(\vartheta )=P_{xx}(\pi /2\vartheta )=P_{ee}(\pi /2\vartheta )`$ (this was pointed out in ). Remember that, in the case of vacuum oscillations, $`\vartheta `$ is physically equivalent to $`\pi /2\vartheta `$, so $`P_{\overline{e}\overline{e}}=P_{ee}`$. In the more general case of nontrivial matter effects, this is clearly not the case, since the presence of matter (or antimatter) explicitly breaks $`CP`$-invariance.
It is curious to note that, in the case of two-flavour oscillations, there is no $`T`$-noninvariance, i.e., $`P_{\alpha \beta }=P_{\beta \alpha }`$, while there is potentially large $`CP`$ violation, i.e., $`P_{\alpha \beta }P_{\overline{\alpha }\overline{\beta }}`$, even if the Hamiltonian for the system is explicitly $`T`$-noninvariant and $`CP`$-noninvariant, as is the case of the propagation of neutrinos produced in the Sun (namely $`A(t)`$ is a generic function of time and $`A(t)`$ for neutrinos is $`A(t)`$ for antineutrinos).
## 3 Three Flavour Oscillations
Currently, aside from the solar neutrino puzzle, there is an even more convincing evidence for neutrino oscillations, namely the suppression of the muon-type neutrino flux in atmospheric neutrino experiments . This atmospheric neutrino puzzle is best solved by $`\nu _\mu \nu _\tau `$ oscillations with a large mixing angle . Furthermore, the values of $`\mathrm{\Delta }m^2`$ required to solve the atmospheric neutrino puzzle are at least one order of magnitude higher than the values required to solve the solar neutrino puzzle. For this reason, in order to solve both neutrino puzzles in terms of neutrino oscillations, three neutrino families are required.
In this section, the oscillations of three neutrino flavours will considered. In order to simplify the discussion, I will concentrate on neutrinos with energies ranging from a few to tens of GeV, which is the energy range expected for neutrinos produced by the annihilation of dark matter particles which are possibly trapped inside the Sun. Furthermore, a number of experimentally inspired constraints on the neutrino oscillation parameter space will be imposed, as will become clear later.
### 3.1 Generalities
Similar to the two-flavour case, the “mapping” between the flavour eigenstates, $`\nu _e`$, $`\nu _\mu `$ and $`\nu _\tau `$ and the mass eigenstates $`\nu _i`$, $`i=1,2,3`$ with masses $`m_i`$ can be performed with a general $`3\times 3`$ unitary matrix, which is parametrised by three mixing angles ($`\theta `$, $`\omega `$, and $`\xi `$) and a complex phase $`\varphi `$. In short hand notation $`\nu _\alpha =U_{\alpha i}\nu _i`$ where $`\alpha =e,\mu ,\tau `$ and $`i=1,2,3`$. The MNS mixing matrix will be written, similar to the standard CKM quark mixing matrix , as
$$\left(\begin{array}{ccc}U_{e1}& U_{e2}& U_{e3}\\ U_{\mu 1}& U_{\mu 2}& U_{\mu 3}\\ U_{\tau 1}& U_{\tau 2}& U_{\tau 3}\end{array}\right)=\left(\begin{array}{ccc}c\omega c\xi & s\omega c\xi & s\xi e^{i\varphi }\\ s\omega c\theta c\omega s\theta s\xi e^{i\varphi }& c\omega c\theta s\omega s\theta s\xi e^{i\varphi }& s\theta c\xi \\ s\omega s\theta c\omega c\theta s\xi e^{i\varphi }& c\omega s\theta s\omega c\theta s\xi e^{i\varphi }& c\theta c\xi \end{array}\right),$$
(3.1)
where $`c\zeta \mathrm{cos}\zeta `$ and $`s\zeta \mathrm{sin}\zeta `$ for $`\zeta =\omega ,\theta ,\xi `$. If the neutrinos are Majorana particles, two extra phases should be added to the MNS matrix, but, since they play no role in the physics of neutrino oscillations, they can be safely ignored. All physically distinguishable situations can be obtained if one allows $`0\varphi \pi `$, all angles to vary between $`0`$ and $`\pi /2`$ and no restriction is imposed on the sign of the mass-squared differences, $`\mathrm{\Delta }m_{ij}^2m_i^2m_j^2`$. Note that there are only two independent mass-squared differences, which are chosen here to be $`\mathrm{\Delta }m_{21}^2`$ and $`\mathrm{\Delta }m_{31}^2`$.
All experimental evidence from solar, atmospheric, and reactor neutrino experiments can be satisfied,<sup>*</sup><sup>*</sup>*There is evidence for neutrino oscillations coming from the LSND experiment . Such evidence has not yet been confirmed by another experiment, and will not be considered in this paper. If, however, it is indeed confirmed, it is quite likely that a fourth, sterile, neutrino will have to be introduced into the picture. somewhat conservatively, by assuming : $`10^4`$ eV$`{}_{}{}^{2}\begin{array}{c}<\hfill \\ \hfill \end{array}|\mathrm{\Delta }m_{31}^2||\mathrm{\Delta }m_{32}^2|\begin{array}{c}<\hfill \\ \hfill \end{array}10^2`$ eV<sup>2</sup>, $`0.3\begin{array}{c}<\hfill \\ \hfill \end{array}\mathrm{sin}^2\theta \begin{array}{c}<\hfill \\ \hfill \end{array}0.7`$, $`10^{11}`$ eV$`{}_{}{}^{2}\begin{array}{c}<\hfill \\ \hfill \end{array}|\mathrm{\Delta }m_{21}^2|\begin{array}{c}<\hfill \\ \hfill \end{array}10^4`$ eV<sup>2</sup>, $`\mathrm{sin}^2\xi \begin{array}{c}<\hfill \\ \hfill \end{array}0.1`$, while $`\omega `$ is mostly unconstrained. There is presently no information on $`\varphi `$. In determining these bounds, it was explicitly assumed that only three active neutrinos exist.
A few comments about the constraints imposed above are in order. First, one may complain that $`\omega `$ is more constrained than mentioned above by the solar neutrino data. The situation is far from definitive, however. As pointed out recently in if the uncertainty on the <sup>8</sup>B neutrino flux is inflated or if some of the experimental data is not considered (especially the Homestake data ) in the fit, a much larger range of $`\mathrm{\Delta }m_{21}^2`$ and $`\omega `$ is allowed. Furthermore, if three-flavour mixing is considered , different regions in the parameter space $`\mathrm{\Delta }m_{21}^2`$-$`\mathrm{sin}^2\omega `$ are allowed for different values of $`\mathrm{sin}^2\xi `$, even if $`\mathrm{sin}^2\xi `$ is constrained to be small.
Second, the limit from the Chooz and Palo Verde reactor experiments do not constrain $`\mathrm{sin}^2\xi `$ for $`|\mathrm{\Delta }m_{31}^2|\begin{array}{c}<\hfill \\ \hfill \end{array}10^3`$ eV<sup>2</sup>. Furthermore, their constraints are to $`\mathrm{sin}^22\xi `$, so values of $`\mathrm{sin}^2\xi `$ close to one should also be allowed. However, the constraints from the atmospheric neutrino data require $`\mathrm{cos}^2\xi `$ to be close to one. This is easy to understand. Assuming that $`L_{21}^{\mathrm{osc}}`$ is much larger than the Earth’s diameter and that $`\mathrm{\Delta }m_{31}^2=\mathrm{\Delta }m_{32}^2`$,
$$P_{\mu \mu }^{\mathrm{atm}}=14\mathrm{cos}^2\xi \mathrm{sin}^2\theta (1\mathrm{cos}^2\xi \mathrm{sin}^2\theta )\mathrm{sin}^2\left(\frac{\mathrm{\Delta }m_{31}^2x}{4E_\nu }\right),$$
(3.2)
according to upcoming Eq. (3.3). Almost maximal mixing implies that $`\mathrm{cos}^2\xi \mathrm{sin}^2\theta 1/2`$. With the further constraint from $`P_{ee}^{\mathrm{atm}}`$, namely $`\mathrm{sin}^22\xi 0`$, one concludes that $`\mathrm{cos}^2\xi 1`$ and $`\mathrm{sin}^2\theta 1/2`$.
In the case of oscillations in vacuum, it is straight forward to compute the oscillation probabilities $`P_{\alpha \beta }`$ of detecting a flavour $`\beta `$ given that a flavour $`\alpha `$ was produced.
$$P_{\alpha \beta }=\underset{i,j}{}U_{\alpha i}^{}U_{\alpha j}U_{\beta i}U_{\beta j}^{}e^{i\frac{\mathrm{\Delta }m_{ij}^2x}{2E_\nu }}$$
(3.3)
The three different oscillation lengths, $`L_{\mathrm{osc}}^{ij}`$, are numerically given by
$$L_{\mathrm{osc}}^{ij}=\frac{4\pi E_\nu }{\mathrm{\Delta }m_{ij}^2}=2.47\times 10^8\mathrm{km}\left(\frac{E}{1\mathrm{GeV}}\right)\left(\frac{10^8\mathrm{eV}^2}{\mathrm{\Delta }m_{ij}^2}\right),$$
(3.4)
which are to be compared to the Earth-Sun distance (1 a.u.$`=1.496\times 10^8`$ km). In the energy range of interest, 1 Gev$`\begin{array}{c}<\hfill \\ \hfill \end{array}E_\nu \begin{array}{c}<\hfill \\ \hfill \end{array}100`$ GeV and given the experimental constraints on the parameter space described above, it is easy to see that $`L_{\mathrm{osc}}^{31}`$ and $`L_{\mathrm{osc}}^{32}`$ are much smaller than 1 a.u., and that its effects should “wash out” due to any realistic neutrino energy spectrum, detector energy resolution, or other “physical” effects. Such terms will therefore be neglected henceforth. In contrast, $`L_{\mathrm{osc}}^{21}`$ maybe as large as (and maybe even much larger than!) the Earth-Sun distance. Note that a nonzero phase $`\varphi `$ implies $`T`$-violation, i.e., $`P_{\alpha \beta }P_{\beta \alpha }`$, unless $`L_{\mathrm{osc}}^{21}1`$ a.u.. This will be discussed in more detail later.
In the presence of neutrino–medium interactions, the situation is, in general, more complicated (indeed, much more!). Similar to the two-neutrino case, it is important to discuss what is known about the oscillation probabilities. From the conservation of probability one has
$`P_{ee}+P_{e\mu }+P_{e\tau }`$ $`=`$ $`1,`$
$`P_{\mu e}+P_{\mu \mu }+P_{\mu \tau }`$ $`=`$ $`1,`$ (3.5)
$`P_{\tau e}+P_{\tau \mu }+P_{\tau \tau }`$ $`=`$ $`1,`$
and, similar to the two-neutrino case, unitarity of the Hamiltonian evolution implies
$`P_{ee}+P_{\mu e}+P_{\tau e}`$ $`=`$ $`1,`$
$`P_{e\mu }+P_{\mu \mu }+P_{\tau \mu }`$ $`=`$ $`1,`$ (3.6)
A third equation of this kind, $`P_{e\tau }+P_{\mu \tau }+P_{\tau \tau }=1`$, is redundant. As before, Eqs. (3.6) can be understood by arguing that, if equal numbers of all neutrino species are produced, the number of $`\nu _\beta `$’s to be detected should be the same, regardless of $`\beta `$, simply because the neutrino propagation is governed by a unitary operator.
One may therefore express all $`P_{\alpha \beta }`$ in terms of only four quantities. Here, these are chosen to be $`P_{ee}`$, $`P_{e\mu }`$, $`P_{\mu \mu }`$, and $`P_{\tau \tau }`$. The others are given by
$`P_{e\tau }`$ $`=`$ $`1P_{ee}P_{e\mu },`$
$`P_{\mu e}`$ $`=`$ $`1+P_{\tau \tau }P_{ee}P_{\mu \mu }P_{e\mu },`$
$`P_{\mu \tau }`$ $`=`$ $`P_{ee}+P_{e\mu }P_{\tau \tau },`$ (3.7)
$`P_{\tau e}`$ $`=`$ $`P_{\mu \mu }+P_{e\mu }P_{\tau \tau },`$
$`P_{\tau \mu }`$ $`=`$ $`1P_{\mu \mu }P_{e\mu }.`$
Note that, in general, $`P_{\alpha \beta }P_{\beta \alpha }`$.
### 3.2 Oscillation of Neutrinos Produced in the Sun’s Core
The propagation of neutrinos in the Sun’s core can, similar to the two-neutrino case, be described by the differential equation
$$\frac{\mathrm{d}}{\mathrm{d}x}\nu _\alpha (r)=i\left(\underset{i=2}{\overset{3}{}}\left(\frac{\mathrm{\Delta }m_{i1}^2}{2E_\nu }\right)U_{\alpha i}^{}U_{\beta i}+A(x)\delta _{\alpha e}\delta _{\beta e}\right)\nu _\beta (r),$$
(3.8)
where $`\delta _{\eta \zeta }`$ is the Kronecker delta symbol. Terms proportional to the identity $`\delta _{\alpha \beta }`$ are neglected because they play no role in the physics of neutrino oscillations. The matter induced potential $`A(x)`$ is given by Eq.(2.8).
As in the two-neutrino case, it is useful to first discuss the initial states $`\nu _\alpha `$ in the Sun’s core, and to express them in the basis of instantaneous Hamiltonian eigenstates, which will be referred to as $`|\nu _H`$, $`|\nu _M`$, and $`|\nu _L`$ ($`H=`$ high, $`M`$= medium, and $`L`$= low). Therefore
$$|\nu _\alpha =H_\alpha |\nu _H+M_\alpha |\nu _M+L_\alpha |\nu _L,$$
(3.9)
where $`\nu _\alpha |\nu _\alpha ^{}=\delta _{\alpha \alpha ^{}}`$. As before (see Eq. 2.10), the probability of detecting this initial state as a $`\beta `$-type neutrino far away from the Sun (e.g., at the Earth) is given by
$$P_{\alpha \beta }=\left|\left(\begin{array}{ccc}U_{\beta 1}^{}& U_{\beta 2}^{}& U_{\beta 3}^{}\end{array}\right)\left(\begin{array}{ccc}1& 0& 0\\ 0& e^{i\frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }}& 0\\ 0& 0& e^{i\frac{\mathrm{\Delta }m_{31}^2x}{2E_\nu }}\end{array}\right)\left(V_{3\times 3}\right)\left(\begin{array}{c}L_\alpha \\ M_\alpha \\ H_\alpha \end{array}\right)\right|^2,$$
(3.10)
where $`V_{3\times 3}`$ is an arbitrary $`3\times 3`$ unitary matrix which takes care of propagating the initial state until the edge of the Sun and rotating the state into the mass basis.
In order to proceed, it is useful take advantage of the constraints on the neutrino parameter space and the energy range of interest. Note that $`A\begin{array}{c}>\hfill \\ \hfill \end{array}\frac{|\mathrm{\Delta }m_{31}^2|}{2E_\nu }\frac{|\mathrm{\Delta }m_{21}^2|}{2E_\nu }`$ (remember that the energy range of interest is 1 GeV$`\begin{array}{c}<\hfill \\ \hfill \end{array}E_\nu \begin{array}{c}<\hfill \\ \hfill \end{array}100`$ GeV and that $`A6\times 10^3`$ eV<sup>2</sup>/GeV). It has been shown explicitly , assuming the neutrino mass-squared hierarchy to be $`m_3^2>m_2^2>m_1^2`$, I will work under this assumption for the time being. that, if the mass-squared differences are very hierarchical ($`|\mathrm{\Delta }m_{31}^2||\mathrm{\Delta }m_{21}^2|`$), the three-level system “decouples” into two two-level systems, i.e., one can first deal with matter effects in the “$`HM`$” system and then with the matter effects in the “$`ML`$” system. One way of understanding why this is the case is to realize that the “resonance point” corresponding to the $`\mathrm{\Delta }m_{31}^2`$ is very far away from the resonance point corresponding to $`\mathrm{\Delta }m_{21}^2`$. With this in mind, it is fair to approximate (this is similar to what is done, for example, in )
$$V_{3\times 3}=\left(\begin{array}{ccc}A^L& B^L& 0\\ B^L& A^L& 0\\ 0& 0& 1\end{array}\right)\left(\begin{array}{ccc}1& 0& 0\\ 0& A^H& B^H\\ 0& B^H& A^H\end{array}\right),$$
(3.11)
where $`|B^H|^2=1|A^H|^2P_c^H`$, $`|B^L|^2=1|A^L|^2P_c^L`$. The superscripts $`H`$, $`L`$ correspond to the “high” and the “low” resonances, respectively.
It also possible to obtain an approximate expression for the initial states in the Sun’s core. Following the result outline above, this state should be described by two matter angles, $`\xi _M`$ and $`\omega _M`$, corresponding to each of the two-level systems. Both should be given by Eq. (2.9), where, in the case of $`\mathrm{cos}2\xi _M`$, $`\vartheta `$ is to be replaced by $`\xi `$ and $`\mathrm{\Delta }m^2`$ by $`\mathrm{\Delta }m_{31}^2`$, while in the case of $`\mathrm{cos}2\omega _M`$, $`\vartheta `$ is to be replaced by $`\omega `$, $`\mathrm{\Delta }m^2`$ by $`\mathrm{\Delta }m_{21}^2`$ and $`A`$ is to be replaced by $`A\mathrm{cos}\xi `$ . Furthermore, because $`A\mathrm{cos}\xi \frac{|\mathrm{\Delta }m_{21}^2|}{2E_\nu }`$, $`\mathrm{cos}2\omega _M`$ can be safely replaced by -1 (remember that $`\mathrm{cos}^2\xi \begin{array}{c}>\hfill \\ \hfill \end{array}0.9`$). Within these approximations, in the Sun’s core,
$`|\nu _e`$ $`=`$ $`\mathrm{sin}\xi _M|\nu _H+\mathrm{cos}\xi _M|\nu _M,`$
$`|\nu _\mu `$ $`=`$ $`\mathrm{sin}\theta \mathrm{cos}\xi _M|\nu _H\mathrm{sin}\theta \mathrm{sin}\xi _M|\nu _M\mathrm{cos}\theta |\nu _L,`$ (3.12)
$`|\nu _\tau `$ $`=`$ $`\mathrm{cos}\theta \mathrm{cos}\xi _M|\nu _H\mathrm{cos}\theta \mathrm{sin}\xi _M|\nu _M+\mathrm{sin}\theta |\nu _L.`$
The accuracy of this approximation has been tested numerically in the range of parameters of interest, and the difference between the “exact” result and the approximate result presented in Eq. (3.2) is negligible.
Keeping all this in mind, it is straight forward to compute all oscillation probabilities, starting from Eq. (3.10). From here on, $`\varphi =0`$ (no $`T`$-violating phase in the mixing matrix, such that all $`U_{\alpha i}`$ are real) will be assumed, in order to simplify expressions and render the results cleaner. In the end of the day one obtains
$`P_{\alpha \beta }`$ $`=`$ $`a_\alpha ^2(U_{\beta 1})^2+b_\alpha ^2(U_{\beta 2})^2+c_\alpha ^2(U_{\beta 3})^2+2a_\alpha b_\alpha (U_{\beta 1}U_{\beta 2})\mathrm{cos}\left({\displaystyle \frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }}+\delta ^L\right)`$ (3.13)
$`\mathrm{or}`$
$`P_{\alpha \beta }`$ $`=`$ $`\left(a_\alpha U_{\beta 1}+b_\alpha U_{\beta 2}\right)^2+c_\alpha ^2(U_{\beta 3})^24a_\alpha b_\alpha (U_{\beta 1}U_{\beta 2})\mathrm{sin}^2\left({\displaystyle \frac{\mathrm{\Delta }m_{21}^2x}{4E_\nu }}+\delta ^L\right),`$
where $`\delta ^L`$ is the matter phase, induced in the low resonance, and
$`a_e`$ $`=`$ $`\sqrt{P_2^HP_c^L},`$
$`b_e`$ $`=`$ $`\sqrt{P_2^H(1P_c^L)},`$
$`c_e`$ $`=`$ $`\sqrt{P_3^H},`$
$`a_\mu `$ $`=`$ $`\sqrt{(1P_c^L)}\mathrm{cos}\theta \sqrt{P_3^HP_c^L}\mathrm{sin}\theta ,`$
$`b_\mu `$ $`=`$ $`\sqrt{P_c^L}\mathrm{cos}\theta \sqrt{P_3^H(1P_c^L)}\mathrm{sin}\theta ,`$ (3.14)
$`c_\mu `$ $`=`$ $`\sqrt{P_2^H\mathrm{sin}^2\theta },`$
$`a_\tau `$ $`=`$ $`\sqrt{(1P_c^L)}\mathrm{sin}\theta \sqrt{P_3^HP_c^L}\mathrm{cos}\theta ,`$
$`b_\tau `$ $`=`$ $`\sqrt{P_c^L}\mathrm{sin}\theta \sqrt{P_3^H(1P_c^L)}\mathrm{cos}\theta ,`$
$`c_\tau `$ $`=`$ $`\sqrt{P_2^H\mathrm{cos}^2\theta },`$
and $`P_2^H=1P_3^H=(|A^H|^2\mathrm{cos}^2\xi _M+|B^H|^2\mathrm{sin}^2\xi _M)`$, which can also be written as $`P_2^H=\frac{1}{2}+\frac{1}{2}\left(12P_c^H\right)\mathrm{cos}2\xi _M`$. This is to be compared with the expression for $`P_1`$ obtained in the two-flavour case. Note that $`a_\alpha ^2+b_\alpha ^2+c_\alpha ^2=1`$. The effect of $`\delta ^L`$ will not be discussed here and from here on $`\delta ^L`$ will be set to zero. For details about the significance of $`\delta ^L`$ for solar neutrinos in the two-flavour case, readers are referred to .
Many comments are in order. First, in the nonadiabatic limit which can be obtained for very large energies, $`P_c^H\mathrm{cos}^2\xi `$, $`P_c^L\mathrm{cos}^2\omega `$ and $`\mathrm{cos}2\xi _M1`$. It is trivial to check that in this limit $`a_\alpha U_{\alpha 1}`$, $`b_\alpha U_{\alpha 2}`$, $`c_\alpha U_{\alpha 3}`$, and the vacuum oscillation result is reproduced, up to the matter induced phase $`\delta ^L`$.
Second, $`P_{ee}`$ can be written as
$$P_{ee}=P_2^H\mathrm{cos}^2\xi (P_{ee}^{2\nu })+P_3^H\mathrm{sin}^2\xi ,$$
(3.15)
where $`P_{ee}^{2\nu }`$ is the two-neutrino result obtained in the previous section (see Eq. (2.11)) in the limit $`\mathrm{cos}2\vartheta _M1`$. It is easy to check that Eq. (2.11) would be exactly reproduced (with $`\vartheta _M`$ replaced by $`\omega _M`$, of course) if the $`\mathrm{cos}2\omega _M=1`$ approximation were dropped.
For solar neutrino energies (100 keV$`\begin{array}{c}<\hfill \\ \hfill \end{array}E_\nu \begin{array}{c}<\hfill \\ \hfill \end{array}10`$ MeV), $`\xi _M\xi `$, $`P_c^H0`$ and therefore $`P_2^H(P_3^H)\mathrm{cos}^2\xi (\mathrm{sin}^2\xi )`$, reproducing correctly the result of the survival probability of electron-type solar neutrinos in a three-flavour oscillation scenario (see and references therein). In this scenario there is no “$`HL`$” resonance inside the Sun, because $`\frac{|\mathrm{\Delta }m_{31}^2|}{2E_\nu }A`$ for solar neutrino energies.
On the other hand, in the case $`P_c^H0`$ and $`\mathrm{cos}2\xi _M1`$, $`P_3^H1`$ and electron-type neutrinos exit the Sun as a pure $`\nu _3`$ mass eigenstate, and do not undergo vacuum oscillations even if $`\mathrm{\Delta }m_{21}^2`$ is very small. In contrast, $`\nu _\mu `$ and $`\nu _\tau `$ always undergo vacuum oscillations if $`\mathrm{\Delta }m_{21}^2`$ is small enough. The reason for this is simple. The generic feature of matter effects is to “push” $`\nu _e`$ into the heavy mass eigenstate, while $`\nu _\mu `$ and $`\nu _\tau `$ are “pushed” into the light mass eigenstates. This situation is changed by nonadiabatic effects, as argued above.
Finally, it is important to note that all equations obtained are also valid in the case of inverted hierarchies ($`m_3^2<m_{1,2}^2`$ or $`m_2^2<m_1^2`$). This has been discussed in detail in the two-neutrino oscillation case , and is also applicable here. It is worthwhile to point out that, in the approximation $`\mathrm{\Delta }m_{31}^2\mathrm{\Delta }m_{32}^2`$ the transformation $`\mathrm{\Delta }m_{21}^2\mathrm{\Delta }m_{21}^2`$ can be reproduced by transforming $`\omega \pi /2\omega `$, $`\theta \pi \theta `$ and redefining the sign of $`\nu _\tau `$. Therefore, one is in principle allowed to fix the sign of $`\mathrm{\Delta }m_{21}^2`$ as long as $`\theta `$ is allowed to vary between $`0`$ and $`\pi `$.
In the case of inverted hierarchies (especially when $`\mathrm{\Delta }m_{31}^2<0`$) one expects to see no “level crossing” (indeed, matter effects tend to increase the distance between the “energy” levels in this case), but matter effects are still present, because the initial state in the Sun’s core can be nontrivial ( i.e., $`\vartheta _M\vartheta `$). Note that $`\nu _e`$ is still “pushed” towards $`\nu _H`$, even in the case of inverted hierarchies, and the expressions for the matter mixing angles Eq. (2.9), and the initial states inside the Sun Eq. (3.2) are still valid. The consequence of no “level crossing” is that the adiabatic limit does not connect, for example, $`\nu _H\nu _3`$ but $`\nu _H\nu _2`$ (or $`\nu _1`$, depending on the sign of $`\mathrm{\Delta }m_{21}^2`$). This information is in fact contained in the equations above. The crucial feature is that, for example, when $`\mathrm{\Delta }m_{31}^2<0`$, $`P_c^H1`$ in the “adiabatic limit,” and the matrix $`V_{3\times 3}`$ correctly “connects” $`\nu _H\nu _2`$ (or $`\nu _1`$)! Another curious feature is that, in the limit $`|\mathrm{\Delta }m_{31}^2|/2E_\nu A`$, $`\mathrm{cos}2\xi _M\mathrm{cos}2\xi `$, $`P_c^H1`$ and Eq. (3.15) correctly reproduces the survival probability of electron-type solar neutrinos in the three-flavour oscillation case. Note that on this case the sign of $`\mathrm{\Delta }m_{31}^2`$ does not play any role, as expected. On the other hand, it is still true that $`P_c^{(H,L)}\mathrm{cos}^2(\xi ,\omega )`$ in the extreme nonadiabatic limit, and vacuum oscillation results are reproduced, as expected. Again, in this limit, one is not sensitive to the sign of $`\mathrm{\Delta }m_{31}^2`$, as expected.
### 3.3 The Case of Antineutrinos
As in the two-neutrino case, the difference between neutrinos and antineutrinos is that the equivalent of Eq. (3.8) for antineutrinos can be obtained by changing $`A(x)A(x)`$ and $`U_{\alpha i}U_{\alpha i}^{}`$. Unlike the two-flavour case, however, there is no set of variable transformations that allows one to exactly relate the differential equation for the neutrino and antineutrino systems. One should, however, note that if the signs of both $`\mathrm{\Delta }m^2`$ are changed and $`U_{\alpha i}U_{\alpha i}^{}`$, the neutrino equation turns into the antineutrino equation, up to an overall sign. This means, for example, that the instantaneous eigenvalues of the antineutrino Hamiltonian can be read from the eigenvalues of the neutrino Hamiltonian with $`\mathrm{\Delta }m_{ij}^2\mathrm{\Delta }m_{ij}^2`$, $`U_{\alpha i}U_{\alpha i}^{}`$ plus an overall sign.
When it comes to computing $`P_{\overline{\alpha }\overline{\beta }}`$ this global sign difference is not relevant, and therefore $`P_{\overline{\alpha }\overline{\beta }}(\mathrm{\Delta }m_{ij}^2,U_{\alpha i})=P_{\alpha \beta }(\mathrm{\Delta }m_{ij}^2,U_{\alpha i}^{})`$.
## 4 Results and Discussions
This section contains the compilation and discussion of a number of results concerning the oscillation of GeV neutrinos of all species produced in the Sun’s core. The goal here is to explore the multidimensional parameter space spanned by $`\mathrm{\Delta }m_{21}^2`$, $`\mathrm{\Delta }m_{31}^2`$, $`\mathrm{sin}^2\omega `$, $`\mathrm{sin}^2\theta `$, and $`\mathrm{sin}^2\xi `$ (and $`E_\nu `$).
It will be assumed throughout that the electron number density profile of the Sun is exponential, so that Eq. (2.12) can be used. As mentioned before, the numerical accuracy of this approximation is quite good, and certainly good enough for the purposes of this paper. Therefore, both $`P_c^H`$ and $`P_c^L`$ which appear in Eq. (3.13) will be given by Eqs. (2.12, 2.13), with $`\vartheta \xi `$, $`\mathrm{\Delta }m^2\mathrm{\Delta }m_{31}^2`$ in the former, and $`\vartheta \omega `$, $`\mathrm{\Delta }m^2\mathrm{\Delta }m_{21}^2`$ in the latter.
When computing $`P_{\alpha \beta }`$, an averaging over “seasons” is performed, which “washes out” the effect of very small oscillation wavelengths. Furthermore, integration over neutrino energy distributions is performed. Finally, all $`P_{\alpha \beta }`$ to be computed should be understood as the value of $`P_{\alpha \beta }`$ in the Earth’s surface, i.e., Earth matter effects are not included. This is done in order to make the Sun matter effects in the evaluation of $`P_{\alpha \beta }`$ more clear. It should be stressed that Earth matter effects may play a significant role for particular regions of the parameter space, but the discussion of such effects will be left for another opportunity.
Because the parameter space to be explored is multidimensional, it is necessary to make two-dimensional projections of it, such that “illustrative” points are required. The following points in the parameter space are chosen, all inspired by the current experimental situation:
* ATM: $`\mathrm{\Delta }m_{31}^2=3\times 10^3`$ eV<sup>2</sup>, $`\mathrm{sin}^2\theta =0.5`$, and $`\mathrm{sin}^2\xi =0.01`$,
* LMA: $`\mathrm{\Delta }m_{21}^2=2\times 10^5`$ eV<sup>2</sup>, $`\mathrm{sin}^2\omega =0.2`$,
* SMA: $`\mathrm{\Delta }m_{21}^2=6\times 10^6`$ eV<sup>2</sup>, $`\mathrm{sin}^2\omega =0.001`$,
* LOW: $`\mathrm{\Delta }m_{21}^2=1\times 10^7`$ eV<sup>2</sup>, $`\mathrm{sin}^2\omega =0.4`$,
* VAC: $`\mathrm{\Delta }m_{21}^2=1\times 10^{10}`$ eV<sup>2</sup>, $`\mathrm{sin}^2\omega =0.55`$.
ATM corresponds to the best fit point of the solution to the atmospheric neutrino puzzle , and a value of $`\mathrm{sin}^2\xi =0.01`$ which is consistent with all the experimental bounds. Note that some “subset” of ATM will always be assumed (for example, $`\mathrm{sin}^2\theta `$ is fixed while exploring the ($`\mathrm{\Delta }m_{31}^2\times \mathrm{sin}^2\xi `$)- plane). For each analysis, it will be clear what are the “variables” and what quantities are held fixed at their “preferred point” values. All other points refer to sample points in the regions which best solve the solar neutrino puzzle , and the notation should be obvious. Initially a flat neutrino energy distribution with $`E_\nu ^{min}=1`$ GeV and $`E_\nu ^{max}=5`$ GeV is considered (for concreteness), and the case of higher average energies is briefly discussed later.
### 4.1 The Case of Vacuum Oscillations
If neutrinos were produced and propagated exclusively in vacuum, the oscillation probabilities would be given by Eq. (3.3). This would be the case of neutrinos produced in the Sun’s core if either the electron number density were much smaller than its real value or if very low energy neutrinos were being considered. Nonetheless, it is still useful to digress some on the “would be” vacuum oscillation probabilities in order to understand better the matter effects.
In the case of pure vacuum oscillations, it is trivial to check that $`P_{\alpha \beta }=P_{\beta \alpha }`$ (remember that the MNS matrix phase $`\varphi `$ has been set to zero), and therefore all $`P_{\alpha \beta }`$ can be parametrised by three quantities, namely $`P_{\alpha \alpha }`$, $`\alpha =e,\mu ,\tau `$. It is easy to show that
$$P_{\alpha \beta }=P_{\beta \alpha }P_{e\mu }=\frac{1}{2}(1+P_{\tau \tau }P_{\mu \mu }P_{ee}).$$
(4.1)
From Eq. (3.3)
$`P_{\alpha \alpha }`$ $`=`$ $`U_{\alpha 1}^4+U_{\alpha 2}^4+U_{\alpha 3}^4+2U_{\alpha 1}^2U_{\alpha 2}^2\mathrm{cos}\left({\displaystyle \frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }}\right)`$ (4.2)
$`\mathrm{or}`$
$`P_{\alpha \alpha }`$ $`=`$ $`(1U_{\alpha 3}^2)^2+U_{\alpha 3}^44U_{\alpha 1}^2U_{\alpha 2}^2\mathrm{sin}^2\left({\displaystyle \frac{\mathrm{\Delta }m_{21}^2x}{4E_\nu }}\right).`$
Note that there is no dependency on $`\mathrm{\Delta }m_{31}^2`$. Particularly simple limits can be reached when $`L_{\mathrm{osc}}^{21}`$ is either very small or very large compared with the Earth-Sun distance. In both limits $`P_{\alpha \alpha }`$ is independent of $`\mathrm{\Delta }m_{21}^2`$ and, in the latter case, $`P_{\alpha \alpha }`$ depends only on $`U_{\alpha 3}^2`$. Fig. 1 depicts constant $`P_{\alpha \beta }`$ contours in the ($`\mathrm{\Delta }m_{21}^2\times \mathrm{sin}^2\omega `$)-plane, at ATM. Remember that, here, $`P_{e\mu }`$ is not an independent quantity but is a linear combination of all $`P_{\alpha \alpha }`$. Note that $`P_{ee}`$ is symmetric for $`\omega \pi /2\omega `$, and that $`P_{\mu \mu }P_{\tau \tau }`$ when $`\omega \pi /2\omega `$. The latter property is a consequence of $`\theta =\pi /4`$. Also, in the case of $`P_{ee}`$, the $`L_{\mathrm{osc}}^{21}\mathrm{}`$ coincides with the $`\omega 0,\pi /2`$ limit for any $`L_{\mathrm{osc}}^{21}`$ (this is because either $`U_{e1}`$ or $`U_{e2}`$ go to zero). This is not true of $`P_{\mu \mu }`$ or $`P_{\tau \tau }`$ unless $`\mathrm{sin}^2\xi =0`$. Another important consequence of $`L_{\mathrm{osc}}^{21}1`$ a.u. is that $`T`$-violating effects are absent, even if $`\varphi `$ is nonzero. This can be seen by looking at the second expression in Eq. (4.2), which is a function only of $`|U_{\alpha 3}|^2`$ in the limit $`L_{\mathrm{osc}}^{21}\mathrm{}`$.
Finally, one should note that oscillatory effects are maximal for $`\mathrm{\Delta }m_{21}^22\times 10^8`$ eV<sup>2</sup>. In this region $`\mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }\right)1`$, and the largest suppression to all $`P_{\alpha \alpha }`$ is obtained when $`U_{\alpha 1}^2U_{\alpha 2}^2`$ is maximum. For example, $`P_{ee}`$ is smallest when $`\omega =\pi /4`$, since $`U_{e1}^2U_{e2}^2\mathrm{sin}^22\omega `$. There are no “localised” maxima for $`P_{\alpha \alpha }`$ because $`U_{\alpha 1}^2U_{\alpha 2}^2`$ is positive definite.
### 4.2 “Normal” Neutrino Hierarchy
When matter effects are “turned on,” the situation can be dramatically different. This is especially true in the case of normal neutrino mass hierarchies ($`m_1^2<m_2^2<m_3^2`$), which will be discussed first.
The first effect one should observe is that, even though $`L_{\mathrm{osc}}^{31}1`$ a.u., $`P_{\alpha \beta }`$ depend rather nontrivially on $`\mathrm{\Delta }m_{31}^2`$. This dependency comes from the terms $`P_3^H`$ and $`P_2^H=1P_3^H`$ in Eq. (3.13). Remember that $`P_3^H`$ is interpreted as the probability that a $`\nu _e`$ produced in the Sun’s core exits the Sun as a $`\nu _3`$ mass eigenstate. When matter effects are negligible (such as in the limit of small neutrino energies) $`P_3^H\mathrm{sin}^2\xi `$, its “vacuum limit.” Fig. 2 depicts constant $`P_3^H`$ contours in the ($`\mathrm{\Delta }m_{31}^2/E_\nu \times \mathrm{sin}^2\xi `$)-plane.
Note that, for $`\mathrm{\Delta }m_{31}^2/E_\nu 10^2`$ eV<sup>2</sup>/GeV, $`P_3^H1`$, even for small values of $`\mathrm{sin}^2\xi `$. In this region, $`\nu _e`$’s produced in the Sun’s core exit the Sun as pure $`\nu _3`$’s. Therefore, $`P_{e\alpha }U_{\alpha 3}^2`$. Because of unitarity in the propagation, $`\nu _\mu `$’s and $`\nu _\tau `$’s exit the Sun as linear combinations of the light mass eigenstates, and may not only undergo vacuum oscillations but are also susceptible to further matter effects (dictated by the “$`ML`$” system, as described in Sec. 3). For future reference, at ATM, $`P_3^H0.87`$ when averaged over the energy range mentioned in the beginning of this section.
As $`\mathrm{\Delta }m_{31}^2/E_\nu `$ decreases (as is the case for higher energy neutrinos) the nonadiabaticity of the “$`HM`$” system starts to become relevant, and $`P_3^H\mathrm{sin}^2\xi `$, as argued in Sec. 3.2. A hint of this behaviour can already be seen in Fig. 2, for small values of $`\mathrm{\Delta }m_{31}^2/E_\nu `$.
The information due to the “$`ML`$” matter effect is encoded in $`P_c^L`$, present in Eq. (3.13). Fig. 3 depicts contours of constant $`1P_c^L`$ in the ($`\mathrm{\Delta }m_{21}^2/E_\nu \times \mathrm{sin}^2\omega `$)-plane. One should note that $`1P_c^L`$ reaches its extreme nonadiabatic limit, $`\mathrm{sin}^2\omega `$, when $`\mathrm{\Delta }m_{21}^2/E_\nu \begin{array}{c}<\hfill \\ \hfill \end{array}10^7`$ eV<sup>2</sup>/GeV. For $`\mathrm{\Delta }m_{21}^2/E_\nu \begin{array}{c}>\hfill \\ \hfill \end{array}10^7`$ eV<sup>2</sup>/GeV, matter effects increase the value of $`1P_c^L`$.
One can use the intuition from the two-flavour solution to the solar neutrino puzzle to better appreciate the results presented here. In the case of the solutions to the solar neutrino puzzle, the energies of interest range from 100 keV to 10 MeV, and large matter effects happen around $`\mathrm{\Delta }m^210^5`$ eV<sup>2</sup>. Furthermore, at $`\mathrm{\Delta }m^210^{10}`$ eV<sup>2</sup> one encounters the “just-so” solution, which is characterised by very long wave-length vacuum oscillations. Rescaling to $`O`$(GeV) energies, the equivalent of the “just-so” solution happens for $`\mathrm{\Delta }m_{21}^2(10^810^7)`$ eV<sup>2</sup>, while large matter effects would be present at $`\mathrm{\Delta }m^2(10^310^2)`$ eV<sup>2</sup>. Indeed, one observes large matter effects for $`\mathrm{\Delta }m_{31}^2(10^310^2)`$ eV<sup>2</sup>. $`\mathrm{\Delta }m_{21}^2(10^510^6)`$ eV<sup>2</sup> corresponds to the region between the LOW and VAC solutions, where matter effects distort $`P_{\alpha \beta }`$ from its pure vacuum value, but no dramatic suppression or enhancement takes place. Incidently, this behaviour has physical consequences in the solution to the solar neutrino problem, as was first pointed out in .
Figs. 4 and 5 depict contours of constant $`P_{\alpha \alpha }`$ and $`P_{e\mu }`$ in the ($`\mathrm{\Delta }m_{31}^2\times \mathrm{sin}^2\xi `$)-plane. As expected, in the region where $`P_3^H1`$, $`P_{ee}`$ and $`P_{e\mu }`$ do not depend on $`\mathrm{\Delta }m_{21}^2`$ or $`\mathrm{sin}^2\omega `$, namely $`P_{ee}\mathrm{sin}^2\xi `$ and $`P_{e\mu }0.5\mathrm{cos}^2\xi `$. Remember that the results depicted in Figs. 4 and 5 (and all other plots from here on) are for an energy band from 1 to 5 GeV. On the other hand, $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ do depend on the point (LMA, SMA, etc), even for $`P_3^H1`$, as foreseen. This dependence will be discussed in what follows.
In the limit $`P_3^H=1`$, $`\mathrm{sin}^2\theta =0.5`$
$`c_\mu ^2=`$ $`c_\tau ^2=`$ $`0,`$ (4.3)
$`a_\mu ^2=`$ $`b_\tau ^2=`$ $`0.5(1+2\sqrt{P_c^L(1P_c^L))},`$ (4.4)
$`b_\mu ^2=`$ $`a_\tau ^2=`$ $`0.5(12\sqrt{P_c^L(1P_c^L))},`$ (4.5)
$`a_\mu b_\mu =`$ $`a_\tau b_\tau =`$ $`0.5(12P_c^L),`$ (4.6)
and
$`P_{(\mu \mu ,\tau \tau )}`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1U_{(\mu ,\tau )3}^2)\pm \sqrt{P_c^L(1P_c^L)}(U_{(\mu ,\tau )1}^2U_{(\mu ,\tau )2}^2)`$
$`\pm `$ $`(12P_c^L)U_{(\mu ,\tau )1}U_{(\mu ,\tau )2}\mathrm{cos}\left({\displaystyle \frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }}\right).`$
At both LMA and SMA, the oscillatory term averages out to zero, while at VAC $`\mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }\right)=1`$. It is only at LOW that the oscillatory term is nontrivial, as was mentioned in the analogy between the situation at hand and the solutions to the solar neutrino puzzle.
Furthermore, at SMA, $`1P_c^L`$ is tiny (see Fig. 3), so it is fair to approximate $`P_{\mu \mu }P_{\tau \tau }0.5(10.5\times 0.99)0.25`$, in agreement with Fig. 4(bottom). At LMA, it is fair to approximate $`\mathrm{sin}^2\xi =0`$. In this limit, $`U_{\mu 1}^2U_{\mu 2}^20.5\mathrm{cos}2\omega +\mathrm{sin}2\omega \mathrm{sin}\xi `$, while $`U_{\tau 1}^2U_{\tau 2}^20.5\mathrm{cos}2\omega \mathrm{sin}2\omega \mathrm{sin}\xi `$. Therefore, because $`\mathrm{cos}2\omega =0.6>0`$, $`P_{\mu \mu }`$ is significantly less than $`P_{\tau \tau }`$, since $`\sqrt{P_c^L(1P_c^L)}`$ is nonnegligible. Roughly, $`P_{\mu \mu }0.15`$ and $`P_{\mu \mu }0.4`$, using the approximations above. Again, there is agreement with Fig. 4(top).
In order to understand the behaviour at LOW and VAC, one should take advantage of the fact that $`1P_c^L\mathrm{sin}^2\omega `$. In this case, it proves more advantageous to use the second form of Eq. (3.13) in order to express all $`P_{\alpha \beta }`$
$`P_{ee}`$ $`=`$ $`P_2^H\mathrm{cos}^2\xi +P_3^H\mathrm{sin}^2\xi (\mathrm{Osc})_{ee},`$
$`P_{e\mu }`$ $`=`$ $`\left(P_2^H\mathrm{sin}^2\xi +P_3^H\mathrm{cos}^2\xi \right)\mathrm{sin}^2\theta (\mathrm{Osc})_{e\mu },`$ (4.8)
$`P_{\mu \mu }`$ $`=`$ $`\left(\mathrm{cos}^2\theta +\sqrt{P_3^H}\mathrm{sin}\xi \mathrm{sin}^2\theta \right)^2+P_2^H\mathrm{cos}^2\xi \mathrm{sin}^4\theta (\mathrm{Osc})_{\mu \mu },`$
$`P_{\tau \tau }`$ $`=`$ $`\left(\mathrm{sin}^2\theta +\sqrt{P_3^H}\mathrm{sin}\xi \mathrm{cos}^2\theta \right)^2+P_2^H\mathrm{cos}^2\xi \mathrm{cos}^4\theta (\mathrm{Osc})_{\tau \tau },`$
where
$$(\mathrm{Osc})_{\alpha \beta }=4a_\alpha b_\alpha U_{\beta 1}U_{\beta 2}\mathrm{sin}^2\left(\frac{\mathrm{\Delta }m_{12}^2x}{4E_\nu }\right)$$
(4.9)
are the oscillatory terms. When $`L_{\mathrm{osc}}^{21}1`$ a.u., the oscillatory terms are zero, and $`P_{\alpha \beta }`$ are particularly simple. Note that on this limit many simplifications happen: $`P_{\alpha \beta }`$ is independent of $`\omega `$ and $`\mathrm{\Delta }m_{21}^2`$, and $`P_{\mu \mu }=P_{\tau \tau }`$ if $`\mathrm{sin}^2\theta =\mathrm{cos}^2\theta `$, as can be observed in Fig. 5(bottom). A very important fact is that, when the oscillatory terms are neglected, $`2P_{e\mu }=1+P_{\tau \tau }P_{\mu \mu }P_{ee}`$, as one may easily verify directly. As argued before, when this condition is satisfied, $`P_{\alpha \beta }=P_{\beta \alpha }`$. This is not the case in the presence of nonnegligible oscillation effects or when $`P_c\mathrm{cos}^2\omega `$. Both statements are trivial to verify directly. For example,
$$4a_eb_eU_{\mu 1}U_{\mu 2}=P_2^H\mathrm{sin}2\omega \left[\mathrm{sin}2\omega \left(\mathrm{sin}^2\xi \mathrm{sin}^2\theta \mathrm{cos}^2\theta \right)\mathrm{sin}\xi \mathrm{sin}2\theta \mathrm{cos}2\omega \right]$$
(4.10)
while
$$4a_\mu b_\mu U_{e1}U_{e2}=\mathrm{cos}^2\xi \mathrm{sin}2\omega \left[\mathrm{sin}2\omega \left(P_3^H\mathrm{sin}^2\theta \mathrm{cos}^2\theta \right)\sqrt{P_3^H}\mathrm{sin}2\theta \mathrm{cos}2\omega \right],$$
(4.11)
so $`\mathrm{Osc}_{e\mu }\mathrm{Osc}_{\mu e}`$.
Figs. 6 and 7 depict $`P_{\alpha \beta }`$ as a function of $`\mathrm{sin}^2\theta `$ at LMA and SMA, and at LOW and VAC, respectively. In these figures, all $`P_{\alpha \beta }`$ are plotted, in order to illustrate that $`P_{\alpha \beta }P_{\beta \alpha }`$ at LMA, SMA and VAC. Note that at LMA and SMA, the difference comes from the fact that $`P_3^H\mathrm{sin}^2\xi `$ and $`P_c^L\mathrm{cos}^2\omega `$. At LOW, $`P_c^L\mathrm{cos}^2\omega `$, but $`P_3^H\mathrm{sin}^2\xi `$ and nontrivial oscillatory terms render $`P_{\alpha \beta }P_{\beta \alpha }`$. At VAC, even though $`P_3^H\mathrm{sin}^2\xi `$, $`P_{\alpha \beta }=P_{\beta \alpha }`$ because $`P_c^L=\mathrm{cos}^2\omega `$ and because “1-2” oscillations don’t have “time” to happen.
From Eqs. (4.2) one can roughly understand the dependency of $`P_{\alpha \beta }`$ on $`\mathrm{sin}^2\theta `$. Obviously $`P_{ee}`$ does not depend on $`\theta `$ (by the very form of the MNS matrix, Eq. (3.1)), while $`P_{e\mu }`$ ($`P_{e\tau }`$) depends almost exclusively on $`\mathrm{sin}^2\theta `$ ($`\mathrm{cos}^2\theta `$). This is guaranteed by the fact that $`P_3^HP_2^H`$ even at LMA and LOW, when one expects the interference terms to play a significant role. It is also worthwhile to note that, as expected, at VAC and SMA the curves are very similar, a behaviour that can be understood from earlier discussions.
Finally, Fig. 8 depicts constant $`P_{\alpha \beta }`$ contours in the ($`\mathrm{\Delta }m_{21}^2\times \mathrm{sin}^2\omega `$)-plane, at ATM. In light of the previous discussions, the shapes and forms can be readily understood. First note that the shapes of the constant $`P_{ee}`$ and $`P_{e\mu }`$ regions resemble those of the pure vacuum oscillations depicted in Fig. 1, with two important differences. First, the constant values of the contours are quite different. For example, $`P_{ee}`$ varies from a few percent to less then 15%, while in the case of pure vacuum oscillations, $`P_{ee}`$ varies from 30% to 100%. This can be roughly understood numerically by noting that $`P_{(ee,e\mu )}P_2^HP_{(ee,e\mu )}^{\mathrm{vac}}`$ (remember that $`P_2^H=1P_3^H0.13`$ when averaged over the energy range of interest). Second, at high $`\mathrm{\Delta }m_{21}^2`$, the regions are distorted. This is due to nontrivial matter effects in the “$`ML`$” system. Note that the contours follow the constant $`1P_c^L`$ curves depicted in Fig. 3.
The $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ contours are a lot less familiar, and require some more discussion. Many features are rather prominent. For example, the plane is roughly divided into a $`\mathrm{sin}^2\omega >0.5`$ and $`\mathrm{sin}^2\omega <0.5`$ structure, and large (small) values of $`P_{\mu \mu }`$ ($`P_{\tau \tau }`$) are constrained to the $`\mathrm{sin}^2\omega >0.5`$ half, and vice-versa. Also, there is a rough $`P_{\mu \mu }(\omega )P_{\tau \tau }(\pi /2\omega )`$ symmetry in the picture, which was present in the pure vacuum case (see Fig. 1). This symmetry is absent for large values of $`\mathrm{\Delta }m_{21}^2`$, similar to what happens in the case of $`P_{ee}`$, and is due, as mentioned in the previous paragraph, to the fact that $`P_c^L`$ is significantly different from $`\mathrm{cos}^2\omega `$ in this region.
The other features are also fairly simple to understand, and are all due to fact that $`P_3^H\mathrm{sin}^2\xi `$. It is convenient to start the discussion in the limit when $`L_{21}^{\mathrm{osc}}1`$ a.u. (the very small $`\mathrm{\Delta }m_{21}^2`$ region). As was noted before, $`P_{\alpha \beta }`$ are given by Eq. (4.2) where the Osc<sub>αβ</sub> terms vanish. It is therefore easy to see that $`P_{\alpha \beta }`$ do not depend on $`\omega `$ or $`\mathrm{\Delta }m_{21}^2`$ (as mentioned before), and furthermore it is trivial to compute the value of $`P_{\alpha \beta }`$ given that we are at ATM and that $`P_3^H0.87`$. The next curious feature is that there is a “band” around $`\mathrm{sin}^2\omega =1/2`$ where $`P_{\mu \mu }P_{\mu \mu }(L_{21}^{\mathrm{osc}}\mathrm{})`$. The same is true of $`P_{\tau \tau }`$. This is due to the fact that, around $`\mathrm{sin}^2\omega 1/2`$, $`a_\mu b_\mu `$ and $`a_\tau b_\tau `$ vanish when $`P_3^H`$ is large. In the limit $`P_3^H=1`$ one can use Eq.(4.6) and note that indeed both $`a_\mu b_\mu `$ and $`a_\tau b_\tau `$ vanish at $`P_c^L=1/2`$. However, for values of $`\mathrm{\Delta }m_{21}^2\begin{array}{c}<\hfill \\ \hfill \end{array}10^7`$ eV<sup>2</sup> $`P_c^L\mathrm{cos}^2\omega `$, which explains the band around $`\omega \pi /4`$. Slight distortions are due to the fact that $`P_3^H1`$, and are easily computed from the exact expressions.
Again in the limit $`P_3^H=1`$, $`P_c^L=\mathrm{cos}^2\omega `$, the coefficient of the $`\mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }\right)`$ term Eq. (4.2) is
$$\pm \frac{1}{2}(12\mathrm{cos}^2\omega )\left(\frac{1}{2}\mathrm{sin}2\omega 0.1\mathrm{cos}2\omega \right),$$
(4.12)
if $`\mathrm{sin}^2\xi `$ terms are neglected. The $`+,`$ signs are for $`P_{\mu \mu }`$ while the $`,+`$ signs for $`P_{\tau \tau }`$. It is trivial to verify numerically (if a little tedious) that the $`P_{\mu \mu }`$ term has a maximum at $`\mathrm{sin}^2\omega 0.1`$ and a minimum at $`\mathrm{sin}^2\omega 0.8`$. For $`P_{\tau \tau }`$ the maximum (minimum) is at $`\mathrm{sin}^2\omega 0.9(0.2)`$. It is important to comment that the minima are negative numbers. On the other hand, from Fig. 1 (as mentioned before) it is easy to see that $`\mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }\right)`$ is minimum for $`\mathrm{\Delta }m_{21}^22\times 10^8`$ eV<sup>2</sup> (this is where all $`P_{\alpha \alpha }`$ are maximally suppressed in Fig. 1). Combining both informations, it is simple to understand the maxima/minima of $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ at $`\mathrm{\Delta }m_{21}^22\times 10^8`$ eV<sup>2</sup>: Minima occur when the coefficient is maximum (e.g., at $`\mathrm{sin}^2\omega 0.1`$ for $`P_{\mu \mu }`$) while maxima occur when the coefficient is minimum (e.g., at $`\mathrm{sin}^2\omega 0.8`$ for $`P_{\mu \mu }`$). A description of what has happened is the following: The matter effects “compress” the constant $`P_{\mu \mu }`$ ($`P_{\tau \tau }`$) contours from the pure vacuum oscillation case (presented in Fig. 1) to the $`\mathrm{sin}^2\omega <1/2`$ ($`>1/2`$) half of the plane, and a new region “appears” on the other half. This other region is characterised by negative values to the coefficients of the oscillatory terms, which are not attainable in the case of pure vacuum oscillations (see Eq. (4.2)).
At last, the contours in the region where the oscillatory effects average out, $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ are also best understood from Eq. (4.2) and the paragraphs which follow it, in the limit that $`\mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2x}{2E_\nu }\right)0`$. It is simple to see, for example, that $`P_{\mu \mu }<P_{\tau \tau }`$ if $`\mathrm{cos}2\omega >0`$ ($`\mathrm{sin}^2\omega >1/2`$), while the situation is reversed if $`\mathrm{cos}2\omega <0`$. This is indeed what one observes in Fig. 8.
### 4.3 “Inverted” Neutrino Hierarchy
Here I turn to the case of an “inverted” neutrino hierarchy, namely $`\mathrm{\Delta }m_{31}^2<0`$. Currently, there is no experimental hint as to what the sign of $`\mathrm{\Delta }m_{31}^2`$ should be, so there is no reason to believe that the “normal” hierarchy is to be preferred over the “inverted” hierarchy. Indeed, even from a theoretical/ model building point of view, there are no strong reasons for or against a particular neutrino mass hierarchy .
The discussion will be restricted to $`\mathrm{\Delta }m_{21}^2>0`$ for two reasons. First, the $`\mathrm{\Delta }m_{21}^2<0`$ can be approximately read off from the $`\mathrm{\Delta }m_{21}^2>0`$ case by changing $`\omega \pi /2\omega `$, as mentioned before. Second, and most important, there is some experimental hints as to what is the sign of $`\mathrm{\Delta }m_{21}^2`$ . For example, the SMA solution only exists for one sign of $`\mathrm{\Delta }m_{21}^2`$, while the LMA and LOW solutions prefer one particular sign. Even in the case of VAC there is the possibility of obtaining information concerning the sign of $`\mathrm{\Delta }m_{21}^2`$ from solar neutrino data . Therefore, the notation introduced in the beginning of this section (ATM, SMA, LMA, LOW, VAC) still applies, and one should simply remember that here $`\mathrm{\Delta }m_{31}^2<0`$.
As advertised, the largest effect of $`\mathrm{\Delta }m_{31}^2<0`$ is the typical values of $`P_c^H`$. From Eq. (2.12), keeping in mind that here $`\gamma `$ is negative,
$$P_c^H=\frac{1e^{|\gamma |\mathrm{cos}^2\xi }}{1e^{|\gamma |}},$$
(4.13)
where $`\gamma `$ is given by Eq. (2.13) with $`\mathrm{\Delta }m^2\mathrm{\Delta }m_{31}^2`$. Since $`|\gamma |1`$ (see Eq. (2.13)), $`P_c^H=1`$ for all values of $`\mathrm{\Delta }m_{31}^2`$ and $`\mathrm{sin}^2\xi `$ of interest. Indeed, this is true for any value of $`\mathrm{sin}^2\xi `$ as long as $`\mathrm{\Delta }m_{31}^210^6`$ eV<sup>2</sup>. This is to be contrasted to the normal hierarchy case, when there is always some value of $`\mathrm{sin}^2\xi `$ (which is a function of $`\mathrm{\Delta }m_{31}^2`$) below which $`P_c`$ deviates significantly from its adiabatic limit.
All of the $`\mathrm{\Delta }m_{31}^2`$ dependency of $`P_{\alpha \beta }`$ is therefore encoded in $`\xi _M`$. However, in the case $`\mathrm{\Delta }m_{31}^2<0`$ it is trivial to show that $`1<\mathrm{cos}2\xi _M<\mathrm{cos}2\xi `$, where the upper bound is reached in the limit $`|\mathrm{\Delta }m_{31}^2/2E_\nu |A`$ (this has been mentioned before. The minus sign takes care of the “unorthodox” $`P_c1`$ adiabatic limit). Since one is interested in $`\mathrm{sin}^2\xi <0.1`$ ($`\mathrm{cos}2\xi <0.8`$), the range for $`\xi _M`$ is rather limited, and therefore any $`\mathrm{\Delta }m_{31}^2`$ effects are bound to be very small. Larger $`\mathrm{\Delta }m_{31}^2`$ effects are expected for larger $`\mathrm{sin}^2\xi `$.
In light of this, Fig. 9 depicts $`P_{\alpha \beta }`$ and $`P_{e\mu }`$ as a function of $`\mathrm{sin}^2\xi `$ at the various points (LMA, SMA, LOW, VAC), for $`\mathrm{\Delta }m_{31}^2=3\times 10^3`$ eV<sup>2</sup> and $`\mathrm{sin}^2\theta =0.5`$.
It is interesting to compare the results presented here with the pure vacuum case. In the limit $`P_2^H=1`$
$$P_{ee}=\mathrm{cos}^2\xi \left[P_c^L\mathrm{cos}^2\omega +(1P_c^L)\mathrm{sin}^2\omega +\sqrt{P_c(1P_c^l)}\mathrm{sin}2\omega \mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2L}{2E_\nu }\right)\right],$$
(4.14)
while the pure vacuum result in the same region of the parameter space is
$$P_{ee}^{\mathrm{vac}}=\mathrm{cos}^4\xi \left[\mathrm{cos}^4\omega +\mathrm{sin}^4\omega +2\mathrm{sin}^2\omega \mathrm{cos}^2\omega \mathrm{cos}\left(\frac{\mathrm{\Delta }m_{21}^2L}{2E_\nu }\right)\right]+\mathrm{sin}^4\xi .$$
(4.15)
In the limit $`P_c^L=\mathrm{cos}^2\omega `$, the difference
$$P_{ee}P_{ee}^{\mathrm{vac}}=\left(P_{ee}^{2\nu ,\mathrm{vac}}\right)\mathrm{cos}^2\xi \mathrm{sin}^2\xi \mathrm{sin}^4\xi ,$$
(4.16)
where $`P_{ee}^{2\nu ,\mathrm{vac}}`$ is the electron neutrino survival probability in the two-flavour case with $`\mathrm{\Delta }m^2=\mathrm{\Delta }m_{21}^2`$ and vacuum mixing angle $`\omega `$. This difference vanishes at $`\mathrm{sin}^2\xi =0`$, and $`\mathrm{sin}^2\xi =\frac{P_{ee}^{2\nu ,\mathrm{vac}}}{1+P_{ee}^{2\nu ,\mathrm{vac}}}`$, (which is between 0 and 0.5). Furthermore, it is a convex function of $`\mathrm{sin}^2\xi `$, which means that $`P_{ee}`$ is larger than the pure vacuum case for values of $`\mathrm{sin}^2\xi <\frac{P_{ee}^{2\nu ,\mathrm{vac}}}{1+P_{ee}^{2\nu ,\mathrm{vac}}}`$. Away from the limit $`P_c^L=\mathrm{cos}^2\omega `$, keeping in mind that the oscillatory terms average out, $`P_{ee}`$ is still larger than the pure vacuum case if $`\mathrm{cos}^2\omega >\mathrm{sin}^2\omega `$ since $`P_c^L\mathrm{cos}^2\omega `$, as one can easily verify.
Also, in the limit $`P_2^H=1`$, $`\mathrm{sin}^2\theta =1/2`$,
$$P_{\mu \mu }=\frac{1}{2}\left[(1P_c^L)U_{\mu 1}^2+P_c^LU_{\mu 2}^2+U_{\mu 3}^2+2\sqrt{P_c^L(1P_c^L)}U_{\mu 1}U_{\mu 2}\mathrm{cos}\left(\frac{\mathrm{\Delta }m^221L}{2E_\nu }\right)\right].$$
(4.17)
The same expression applies for $`P_{\tau \tau }`$ with $`U_{\mu i}U_{\tau i}`$. This is a consequence of $`\mathrm{sin}^2\theta =\mathrm{cos}^2\theta `$. Furthermore, in the limit $`\mathrm{sin}^2\xi 0`$ (and for $`\mathrm{sin}^2\theta =\mathrm{cos}^2\theta `$), $`U_{\mu i}=U_{\tau i}`$, which explains why $`P_{\mu \mu }=P_{\tau \tau }`$ for $`\mathrm{sin}^2\xi \begin{array}{c}<\hfill \\ \hfill \end{array}10^2`$. At VAC this equality remains for all values of $`\mathrm{sin}^2\xi `$. The reason for this is that, at VAC, the expression simplifies tremendously and $`P_{\mu \mu }=P_{\tau \tau }=\frac{1}{4}\left(1+\mathrm{cos}^2\xi \right)`$. In the same region of the parameter space, the pure vacuum oscillation case yields $`P_{\mu \mu }^{\mathrm{vac}}=P_{\tau \tau }^{\mathrm{vac}}=\frac{1}{2}\mathrm{cos}^4\xi \mathrm{cos}^2\xi +1`$. Note that, in this region of the parameter space $`P_{\mu \mu }^{\mathrm{vac}}P_{\mu \mu }`$, the inequality being saturated at $`\mathrm{cos}^2\xi =1`$.
The same result also applies (approximately) at SMA, since the oscillatory terms are proportional to $`\sqrt{P_c^L(1P_c^L)}`$ and $`1P_c^L`$ is very small at SMA (see Fig 3). The equality $`P_{\mu \mu }=P_{\tau \tau }`$ is broken at larger values of $`\mathrm{sin}^2\xi `$ because $`P_c^L\mathrm{cos}^2\omega `$ at SMA.
It remains to discuss how $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ diverge from the pure vacuum case at LMA and LOW. In the limit $`P_c^L=\mathrm{cos}^2\omega `$, and averaging out the oscillatory terms,
$$P_{\mu \mu }P_{\mu \mu }^{\mathrm{vac}}=\frac{\mathrm{sin}\xi }{2}\left[\mathrm{sin}\xi \left(U_{\mu 3}^2(\mathrm{cos}^2\omega U_{\mu 1}^2+\mathrm{sin}^2\omega U_{\mu 2}^2)\right)\mathrm{sin}2\omega (U_{\mu 1}^2U_{\mu 2}^2)\right].$$
(4.18)
This difference goes to zero as $`\mathrm{sin}^2\xi 0`$. This is to be expected, since in this limit the difference of $`P_2^H`$ and $`\mathrm{cos}^2\xi `$ disappears. For small values of $`\mathrm{sin}^2\xi `$, the last term in Eq. 4.18 dominates, and, as discussed before, $`U_{\mu 1}^2U_{\mu 2}^2=0.5\mathrm{cos}2\omega +O(\mathrm{sin}\xi )`$. Therefore, $`P_{\mu \mu }P_{\mu \mu }^{\mathrm{vac}}>0`$ ($`<0`$) for $`\mathrm{cos}2\omega >0`$ ($`<0`$). The expression for $`P_{\tau \tau }`$ can be obtained from Eq.(4.18) by replacing $`U_{\mu i}U_{\tau i}`$ and changing the sign of the last term. Therefore, since $`U_{\tau 1}^2U_{\tau 2}^2=0.5\mathrm{cos}2\omega +O(\mathrm{sin}\xi )`$, $`P_{\tau \tau }P_{\tau \tau }^{\mathrm{vac}}>0`$ ($`<0`$) for $`\mathrm{cos}2\omega <0`$ ($`>0`$). When the oscillatory terms do not average out, it is easy to verify explicitly that the behaviour of the oscillatory terms follows the behaviour of the average terms, discussed above, and the inequalities obtained above still apply.
The situation, however, changes, when $`P_c^L\mathrm{cos}^2\omega `$, i.e., when matter effects due to the “M-L” system are relevant. In this region, a behaviour similar to the one observed in the “normal” hierarchy case is expected, since $`\mathrm{\Delta }m_{21}^2>0`$. Fig. 10 depicts constant $`P_{\alpha \beta }`$ contours in the ($`\mathrm{\Delta }m_{21}^2\times \mathrm{sin}^2\omega `$)-plane. One should be able to see upon close inspection that the region $`P_{ee}<30\%`$ is smaller in Fig. 10 than the same region in the pure vacuum oscillation case, Fig 1. Also, the constant $`P_{\mu \mu }`$ ($`P_{\tau \tau }`$) contours are shifted to larger (smaller) values of $`\mathrm{sin}^2\omega `$. The other prominent (and expected, as mentioned above) feature is the distortion of the contours at large values of $`\mathrm{\Delta }m_{21}^2`$. This behaviour is similar to the one observed in Fig. 8.
I conclude this subsection with a comment on antineutrinos. As discussed previously, $`P_{\overline{\alpha }\overline{\beta }}(\mathrm{\Delta }m_{21}^2,\mathrm{\Delta }m_{31}^2)=P_{\alpha \beta }(\mathrm{\Delta }m_{21}^2,\mathrm{\Delta }m_{31}^2)`$, such that the “normal” hierarchies yield “inverted” hierarchy results for antineutrinos, and vice-verse. One cannot, however, apply Fig. 8 and Fig. 10 for the antineutrinos because both $`\mathrm{\Delta }m_{ij}^2`$ have to change sign, not just $`\mathrm{\Delta }m_{31}^2`$. Qualitatively, however, it is possible to understand the constant $`P_{\overline{\alpha }\overline{\beta }}`$ contours by examining figures Fig. 8 and Fig. 10 reflected in a mirror positioned at $`\mathrm{sin}^2\omega =0.5`$, meaning that $`P_{\overline{\alpha }\overline{\beta }}(\mathrm{sin}^\omega ,\mathrm{\Delta }m_{31}^2)P_{\alpha \beta }(\mathrm{cos}^2\omega ,\mathrm{\Delta }m_{31}^2)`$. The equality is not complete because one is also required to exchange $`\theta \pi \theta `$, as mentioned earlier.
### 4.4 Higher Neutrino Energies
As the average neutrino energy increases, the values of $`P_{\alpha \beta }`$ start to resemble more the pure vacuum case. This is easy to see from Figs. 2 and 3. Any deviation of $`1P_c^L`$ from $`\mathrm{sin}^2\omega `$ goes away even at LMA for $`E_\nu 50`$ GeV, while “H-M” effects remain important up to $`E_\nu 1`$ TeV, even though quantitatively the effect decreases noticeably. This can be illustrated by the value of $`P_3^H`$ at ATM, for example, which drops from 0.87 for energies which range from 1 to 5 GeV (see the previous subsections) to 0.058, for energies which range from 100 to 110 GeV.
Furthermore, all $`L_{ij}^{\mathrm{osc}}`$ increase as the energy increases, for fixed values of $`\mathrm{\Delta }m_{ij}^2`$. Therefore, LOW becomes indistinguishable from VAC at $`E_\nu 100`$ GeV. For $`O`$(TeV) neutrinos the sensitivity to $`\mathrm{\Delta }m_{21}^2`$ remains only for its highest allowed values, while one should start worrying about nontrivial oscillatory effects due to $`L_{31}^{\mathrm{osc}}`$.
The case of higher energy neutrinos contains a more serious complications: neutrino absorption inside the Sun. As the neutrino energy increases, one has to start worrying about the fact that absorptive neutrino interactions can take place. According to , for neutrinos produced in the Sun’s core, absorption becomes important for $`E_\nu \begin{array}{c}>\hfill \\ \hfill \end{array}200`$ GeV. In this case, $`\nu _e`$ and $`\nu _\mu `$ interact with nuclear matter and produce electrons and muons, respectively. The former are capture and “lost” inside the Sun, while the latter stop before decaying into low energy neutrinos. The case of $`\nu _\tau `$-Sun interactions is more interesting, because the $`\tau `$-leptons produced via charged current interactions decay before “stopping”, yielding $`\nu _\tau `$’s with slightly reduced energies. Therefore, it is possible to get a flux of very high energy initial state $`\tau `$-neutrinos but not muon or electron-type neutrinos. Such effects have been studied for high energy galactic neutrinos traversing the Earth .
The effect of neutrino oscillations inside the Sun in the presence of nonnegligible neutrino absorption is certainly of great interest but is beyond the scope of this paper.
## 5 Conclusions
The oscillation probability of $`O`$(GeV) neutrinos of all flavours produced in the Sun’s core has been computed, including matter effects, which are, in general, nontrivial.
In particular, it was shown that, unlike the two-flavour oscillation case, in the three-flavour case the probability of a neutrino produced in the flavour eigenstate $`\alpha `$ to be detected as a flavour eigenstate $`\beta `$ ($`P_{\alpha \beta }`$) is (in general) different from $`P_{\beta \alpha }`$, even if the $`CP`$-violating phase of the MNS matrix vanishes. This is, of course, expected since Sun–neutrino interactions explicitly break $`T`$-invariance. Indeed, it is the case of two-flavour oscillations which is special, in the sense that the number of independent oscillation probabilities is too small because of unitarity.
The results of a particular scan of the parameter space are presented in Sec. 4. In this case, special attention was paid to the regions of the parameter space which are preferred by the current experimental situation.
It turns out that, in the case of a “normal” neutrino mass hierarchy, it is possible to suppress $`P_{ee}`$ tremendously with respect to its pure vacuum oscillation values, by a mechanism that is similar to the well known MSW effect in the case of two-flavour oscillations: the parameters are such that electron-type neutrinos produced in the Sun’s core exit the Sun (almost) as pure mass eigenstates, and the $`\nu _e`$ component of this eigenstate is small. Both $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ can be significantly suppressed, and the constant $`P_{\mu \mu }`$ and $`P_{\tau \tau }`$ contours as a function of the “solar” angle and the smaller mass-squared differences are nontrivial. One important feature is that when $`P_{\mu \mu }`$ is significantly suppressed, $`P_{\tau \tau }`$ is not, and vice-versa. One consequence of this is that, for some regions of the parameter space, it is possible to have an enhancement of $`\nu _\tau `$’s detected in the Earth with respect to the number of $`\nu _\mu `$’s (or vice-versa). This may have important implications for solar WIMP annihilation searches at neutrino telescopes, and will be studied in another oportunity. It is important to note that the effect of neutrino oscillations on the expected event rate at neutrino telescopes will depend on the expected production rate of individual neutrino species inside the Sun, which is, of course, model dependent.
In the case of an “inverted” mass hierarchy, the situation is very similar to the pure vacuum case, and no particular suppression of any $`P_{\alpha \alpha }`$ is possible. Indeed, for a large region of the parameter space $`P_{ee}`$ is in fact enhanced, a feature which is also observed in the two-flavour case .
The case of higher energy neutrinos was very briefly discussed, and the crucial point is to note that, for neutrino energies above a few hundred GeV, the absorption of neutrinos by the Sun becomes important. The study of absorption effects is beyond the scope of this paper.
Finally, it is important to reemphasise that the values of $`P_{\alpha \beta }`$ computed here are to be understood as if they were evaluated at the Earth’s surface. No Earth-matter effects have been included. It is possible that Earth-matter effects are important, especially the ones related to $`\mathrm{\Delta }m_{31}^2`$, in the advent that $`U_{e3}^2\mathrm{sin}^2\xi `$ turns out to be “large.”
## Acknowledgements
I would like to thank John Ellis for suggesting the study of GeV solar neutrinos, and for many useful discussions and comments on the manuscript. I also thank Amol Dighe and Hitoshi Murayama for enlightening discussions and for carefully reading this manuscript and providing useful comments.
|
warning/0006/hep-ph0006291.html
|
ar5iv
|
text
|
# On classical description of radiation from neutral fermion with anomalous magnetic moment
## Abstract
Electromagnetic radiation from an uncharged spin 1/2 particle with an anomalous magnetic moment moving in the classical electromagnetic external field originates from quantum spin-flip transitions. Although this process has a purely quantum nature, it was observed for certain particular external field configurations that, when quantum recoil is neglected, the radiation power corresponds to the classical radiation from an evolving magnetic dipole. We argue that this correspondence has a more general validity in the case of an unpolarized particle and derive a general formula for radiation in terms of the external field strength and its derivative. A classical dynamics of the spin is described by the Bargmann-Michel-Telegdi equation.
Moscow State University,
Department of Theoretical physics.
$`117234`$, Moscow, Russia
PACs: 03. 65. Sq
Electromagnetic radiation from an uncharged spin 1/2 particle possessing an anomalous magnetic moment moving in an external classical electromagnetic field was studied in a number of papers \[1-8\]. The usual approach consists in a computation of the radiative transitions within the Furry’s picture of quantum electrodynamics (QED). This appeals to the knowledge of exact solutions of the Dirac equation with an anomalous magnetic moment in the external electromagnetic field. Such solutions are known for various particular configurations of the electromagnetic field: homogeneous fields , plane-waves \[3-6\], and some other , for them an explicit computation of radiative transitions was performed. Radiative transitions in the case of the uncharged particles correspond to the spin-flip amplitudes and thus have an essentially quantum nature. Meanwhile it was observed that under certain conditions the radiation process may be described in purely classical terms using the Bargmann-Michel-Telegdi (BMT) spin evolution equation . In such a pseudoclassical treatment the radiation power is given by the well-known formula for radiation from a magnetic moment $`\mu ^\nu `$ :
$$\frac{dI}{d𝒪}=\frac{1}{4\pi (lu)^5u^0}\{\ddot{\mu }^\nu \ddot{\mu }_\nu (ul)^2+(l^\nu \ddot{\mu }_\nu )^2\}$$
(1)
where $`(ul)=u^\nu l_\nu `$, $`u^\nu `$ is the 4-velocity of a particle, $`l^\nu =\{1,𝐥\},`$ $`𝐥`$ is the unit vector in the direction of an emitted wave; a dot denotes the differentiation with respect to the proper time $`\tau `$. We use the units $`\mathrm{}=c=1`$.
The radiation power calculated within the framework of QED was found to correspond to the result obtained from the equation (1) under the following conditions. The particle motion has to be quasiclassical, in a sense that:
i) the binding energy due the magnetic moment in the rest frame is much smaller than the particle mass (here we use gaussian units):
$$H_0H_{cr},H_{cr}=m^2c^3/e\mathrm{},$$
(2)
($`H_0`$ is the magnetic field strength in the rest frame),
ii) the external field varies slowly at the distances of the order of a Compton’s length, which amounts to conditions:
$$\mathrm{}\dot{H}_0/mc^2H_01,(H_0/H_{cr})(\omega /\omega _c)1,$$
(3)
(here $`\omega `$ is the characteristic frequency change of the external field, and $`\omega _c=eH/mc`$ is the cyclotron frequency).
If one is interested in the radiation from an unpolarized particle, an additional requirement to be imposed consists in averaging over the initial spin states and summing over the final polarizations. It can be expected that the averaging of the quantum transition amplitudes should correspond to the averaging over the initial orientation of the magnetic dipole moment within the classical picture.
Although no general derivation of such a pseudoclassical description from QED is available for the case of an arbitrary external field, it seems very plausible that under the above assumptions the result should follow indeed from the classical formula (1). Our conjecture is that one has to replace the magnetic moment by the quantity
$$\mu ^\nu =\mu _0S^\nu ,$$
(4)
where $`S^\nu `$ is expectation value of the spin vector, and perform an averaging over the polarizations at $`\tau =\tau _0`$:
$$\frac{dI}{d𝒪}=\frac{\mu _0^2}{4\pi (lu)^5u^0}<\{\ddot{S}^\mu \ddot{S}_\mu (ul)^2+(l^\mu \ddot{S}_\mu )^2\}>=$$
$$=\frac{\mu _0^2}{4\pi (lu)^5u^0}<e^{\mu \nu \rho \lambda }\ddot{S}_\nu u_\rho l_\lambda e_{\mu \nu ^{}\rho ^{}\lambda ^{}}\ddot{S}^\nu ^{}u^\rho ^{}l^\lambda ^{}>,$$
(5)
An evolution of the classical spin is described by the BMT equation :
$$\dot{S}^\nu =2\mu _0\{F^{\nu \alpha }S_\alpha u^\nu (u_\alpha F^{\alpha \beta }S_\beta )\},$$
(6)
whose solution has to be constrained by the conditions $`S^\nu S_\nu =1`$, $`S^\nu u_\nu =0`$. The validity of this equation is ensured by the conditions (2), (3) . Our main goal is to show that when the averaging over polarization states is performed the resulting expression for the radiation power will depend only on the external field intensity and thus will be valid in the case of arbitrary external field subject to conditions (2),(3). We check that it is true indeed for the particular fields studied before.
An important point in the derivation is that the neutral particle moves in the external field with a constant velocity. Of course, the true quantum description of radiation demands that the quantum recoil in the photon emission process should be taken into account. But when conditions (2), (3) are satisfied, energy of emitted photon is small, therefore we can neglect the recoil, i.e. the change of particle velocity. This assumption was made in the formula (5). In the BMT equation one can pass to the rest frame of the particle where the required averaging over polarizations can be easily performed.
It is more convenient to rewrite the BMT equation in term of the $`SL(2,𝐂)`$ spin-tensors. Instead of the 4-vector $`S^\nu `$ we will use matrices $`\underset{}{𝑆}=S^0\sigma _0+𝑺𝝈`$ and $`\stackrel{~}{S}=S^0\sigma _0𝑺𝝈`$, where $`\sigma _\nu `$ are the Pauli matrices (we follow a notation of ).
Then the evolution of spin is given by the following matrix operator :
$$\underset{}{𝑆}(\tau )=\underset{}{𝐿}\underset{}{𝑅}\underset{}{𝐿}^1\underset{}{S_0}(\underset{}{𝐿}^1)^+\underset{}{𝑅}^+\underset{}{𝐿}^+.$$
(7)
Here $`S_0S(\tau _0)`$ is the polarization at the initial moment of time $`\tau _0`$, $`\underset{}{𝐿}^1`$ is a transformation to the particle rest frame
$$\underset{}{𝐿}^1=\frac{1+\stackrel{~}{u}}{\sqrt{2(1+u^0)}},\underset{}{𝐿}=\frac{1+\underset{}{𝑢}}{\sqrt{2(1+u^0)}};$$
and $`\underset{}{𝑅}`$ is a rotation operator satisfying the equation
$$\dot{\underset{}{𝑅}}=i\mu _0\underset{}{H_0}\underset{}{𝑅},$$
(8)
where
$$𝐇_0(x)𝐇_0=u^0𝐇𝐮\times 𝐄\frac{𝐮(\mathrm{𝐮𝐇})}{1+u^0}$$
(9)
is the magnetic field in the rest frame at the point of the particle location. In this notation the formula (5) can be presented in the form:
$$\frac{dI}{d𝒪}=\frac{\mu _0^2}{8\pi (lu)^5u^0}<Sp\{\ddot{\underset{}{𝑆}}\ddot{\stackrel{~}{S}}(ul)^2+\frac{1}{4}[\underset{}{𝑙}\ddot{\stackrel{~}{S}}\underset{}{𝑙}\ddot{\stackrel{~}{S}}+\ddot{\underset{}{𝑆}}\stackrel{~}{l}\ddot{\underset{}{𝑆}}\stackrel{~}{l}]\}>.$$
(10)
After a simple algebra with account for the Eq.(7) we obtain for the radiation power of unpolarized particle:
$$\frac{dI}{d𝒪}=\frac{\mu _0^2}{16\pi (lu)^3u^0}Sp\underset{i}{}\{\ddot{(\sigma _R)_i}\ddot{(\sigma _R)_i}\ddot{(\sigma _R)_i}(𝝈𝒏)\ddot{(\sigma _R)_i}(𝝈𝒏)\}.$$
(11)
where
$$(\sigma _R)_i=\underset{}{𝑅}\sigma _i\underset{}{𝑅}^+,𝐧=\frac{1}{(ul)}\{𝐥\frac{1+(ul)}{1+u^0}𝐮\}$$
is the unit vector in the direction of radiation in the rest frame of the particle. In (11) the summation over $`i`$ corresponds to an averaging over the initial spin states and summation over the final spin states.
Using Eqs. (8) and (11) we arrive at the final result
$$\frac{dI}{d𝒪}=\frac{\mu _0^4}{\pi u^0(ul)^3}\{4\mu _0\{\mu _0𝐇_0^4+\mu _0𝐇_0^2(𝐇_0𝐧)^\mathrm{𝟐}+(𝐇_\mathrm{𝟎}𝐧)[\dot{𝐇}_\mathrm{𝟎}𝐇_\mathrm{𝟎}𝐧]\}+\dot{𝐇}_{\mathrm{𝟎}}^{}{}_{}{}^{\mathrm{𝟐}}+(\dot{𝐇}_\mathrm{𝟎}𝐧)^\mathrm{𝟐}\}.$$
(12)
Therefore the angle distribution of radiated power is expressed entirely in terms of the magnetic field strength in the rest frame of the particle and its derivative with respect to the proper time.
Formula (12) can be rewritten in a covariant form:
$$\begin{array}{cc}& \\ \hfill \frac{dI}{d𝒪}=& \frac{\mu _0^2}{\pi u^0(ul)^5}\{[4(\mu _0^2u_\rho H^{\rho \lambda }H_{\lambda \sigma }u^\sigma )^2+(\mu _0^2u_\rho \dot{H}^{\rho \lambda }\dot{H}_{\lambda \sigma }u^\sigma )](ul)^2\hfill \\ & \\ & +4(\mu _0^2u_\rho H^{\rho \lambda }H_{\lambda \sigma }u^\sigma )(\mu _0u_\rho H^{\rho \lambda }l_\lambda )^2+(\mu _0u_\rho \dot{H}^{\rho \lambda }l_\lambda )^2\hfill \\ & \\ & +4\mu _0^2(\mu _0u_\rho H^{\rho \lambda }l_\lambda )e^{\mu \nu \rho \lambda }u_\mu \dot{H}_{\nu \sigma }u^\sigma H_{\rho \delta }u^\delta l_\lambda \},\hfill \end{array}$$
(13)
where $`H^{\nu \alpha }=\frac{1}{2}e^{\nu \alpha \beta \gamma }F_{\beta \gamma }`$ is the tensor dual to electromagnetic field tensor.
Integrating over angles, we obtain the formula for the total radiation power:
$$I=\frac{16}{3}\mu _0^2\{4(\mu _0𝐇_0)^4+(\mu _0\dot{𝐇}_0)^2\},$$
(14)
which has the following covariant counterpart:
$$I=\frac{16}{3}\mu _0^2\{4(\mu _0^2u_\rho H^{\rho \lambda }H_{\lambda \sigma }u^\sigma )^2+\mu _0^2u_\rho \dot{H}^{\rho \lambda }\dot{H}_{\lambda \sigma }u^\sigma \}.$$
(15)
We verified that, substituting into formulas (12), (14), the fields which were earlier used for calculations both by quantum and classical methods, we obtain the full agreement with the results obtained when we neglect the recoil, i.e., terms of the higher order in the Planck constant.
Now let us discuss some possible application for a neutron. An important features of the Eqs. (12) and (14) is the presence of the second term, whose magnitude is defined by the field strength and its derivative. Therefore the radiation power is increased substantially when the boundary of field region which is crossed by a particle is very sharp. This phenomenon similar to transition radiation can be important in astrophysical conditions. Possibly this effect can be observed under the laboratory conditions when fast neutrons pass through the ferromagnetic.
It is interesting to compare in the order of magnitude the radiation power $`I`$ from a neutral particle (neutron) and the classical radiation power $`I_0`$ from a charged particle (see, for example, ) with the same mass and the energy (proton). We find
$$\frac{I}{I_0}max\{\left(\frac{H_0}{H_{cr}}\right)^2,\left(\frac{u^0\mathrm{}\omega }{mc^2}\right)^2\}$$
Thus, the charge radiation power and the radiation power from a magnetic moment become comparable in very strong fields, or in the fields of high frequency.
Acknowledgments.
Authors are very grateful to Prof. D. V. Gal’tsov for continuous attention to their work, Prof. A. V. Borisov and Prof. V. Ch. Zhukovsky for fruitful discussions.
|
warning/0006/cond-mat0006080.html
|
ar5iv
|
text
|
# A Bosonic Model of Hole Pairs
## I Introduction
The study of strongly correlated electrons in ladder systems has been an active field in recent years. Ladders consisting of a number of coupled legs (or chains) show a rich variety of phases depending on the number of legs and the electron density. One property of particular interest is the binding of holes into pairs when weakly doped into a half-filled two leg ladder, to form a Luther-Emery liquid. In this liquid the spin degrees of freedom are gapped and the low energy sector can be mapped to a boson liquid of hole pairs. This can be studied analytically in the weak coupling limit described, for example, by a one-band Hubbard model with weak interactions. In the strongly interacting limit described by a $`t`$-$`J`$ model, one must use numerical methods such as Lanczos diagonalization and the recently developed Density Matrix Renormalization Group (DMRG) method. Large simulations can be carried out using the DMRG yielding information on the equal time correlations.
In this paper our goal is to use the hole density distributions determined by DMRG to determine the form of the effective boson model that we know on general grounds describes the hole pairs. We use these DMRG results to parameterize their dispersion and their effective mutual interactions and also the interaction with hard wall boundary conditions that occur in DMRG. In this first paper we limit ourselves to the simplest case of the two leg ladder. The same approach can be made for wider ladders with an even number of legs and is currently in progress.
We start with the $`t`$-$`J`$ ladder Hamiltonian
$`H`$ $`=`$ $`t{\displaystyle \underset{i,j,\sigma }{}}𝒫(c_{i,j,\sigma }^{}c_{i+1,j,\sigma }+c_{i+1,j,\sigma }^{}c_{i,j,\sigma })𝒫`$ (4)
$`t^{}{\displaystyle \underset{i,\sigma }{}}𝒫(c_{i,1,\sigma }^{}c_{i,2,\sigma }+c_{i,2,\sigma }^{}c_{i,1,\sigma })𝒫`$
$`+J{\displaystyle \underset{i,j}{}}(𝐒_{i,j}𝐒_{i+1,j}{\displaystyle \frac{1}{4}}n_{i,j}n_{i+1,j})`$
$`+J^{}{\displaystyle \underset{i}{}}(𝐒_{i,1}𝐒_{i,2}{\displaystyle \frac{1}{4}}n_{i,1}n_{i,2})`$
where $`i`$ runs over $`L`$ rungs, $`j(=1,2)`$ and $`\sigma `$ $`(=,)`$ are leg and spin indices. The projection operator $`𝒫_{i,j}(1n_{i,j,}n_{i,j,})`$ prohibits double occupancy of a site. Unless noted otherwise we set $`t=t^{}`$ and $`J=0.35t`$.
## II The structure of a hole pair
In this section we examine a hole pair on a $`t`$-$`J`$ ladder, studying the internal hole-hole correlation (hhc) function. The hhc function gives us information about the inner structure of a hole pair and depends on the parameters $`t`$, $`t^{}`$, $`J`$ and $`J^{}`$ in the $`t`$-$`J`$ Hamiltonian given above. Since the DMRG computations are performed with open boundaries, special care must be taken in the measurement of the hhc-function, which we define as
$$g_j(r_1,r_2)=\frac{n_{r_1,1}^hn_{r_2,j}^h}{n_{r_1,1}^hn_{r_2,j}^h}n^h.$$
(5)
The index $`j=1,2`$ denotes intra- and inter-leg-correlations respectively, $`n_{i,j}^h=1n_{i,j}`$ denotes the hole density at site $`(i,j)`$, and $`n^h`$ the average hole density. Introducing relative and center of mass coordinates, $`r`$ and $`R_0`$, we measure $`g_j(r,R_0)`$ on ($`40\times 2`$)-ladders with open boundaries for $`R_0L/2`$. In a range $`R_0=L/2\pm 5`$, $`g_j(r,R_0)`$ depends only weakly on $`R_0`$. Furthermore, we have tested that better accuracy for the hhc-function cannot be achieved on larger ladders up to ($`60\times 2`$) sites. We conclude that these measurements give good values for $`g_j(r)`$ on an infinite ladder.
We denote by
$$g(r)=\underset{j}{}g_j(r)$$
(6)
the sum of intra- and inter-leg correlations. We have examined $`g(r)`$ as a function of the rung exchange coupling $`J^{}`$. As can be seen in Fig. 1, a significant change in the hhc-function occurs when the isotropic case ($`J^{}J`$) is approached from the strong rung coupling regime ($`J^{}J`$). In the latter case, the holes sit mainly on the same rung whereas in the former case, they are found predominantly on adjacent rungs. We study the long distance behavior of the hhc-function by fitting the tail to an exponentially decaying function, $`g(r)\mathrm{exp}(r/\xi _{hhc})`$. The results are shown in Fig. 2, where $`\xi _{hhc}`$ is plotted as a function of $`J^{}`$. Starting from strong rung coupling, the correlation length first increases as expected, but decreases significantly as $`J^{}`$ is reduced and bound holes sit mainly on adjacent rungs. As the isotropic case is approached, the correlation length increases again.
This behavior can be explained by looking at the different kinds of bonds in the region between two holes of a pair as the holes move apart. In Fig. 3, we show the exchange fields around a pair of holes. These are obtained by measuring $`𝐒_i𝐒_j`$ after projecting out a particular configuration of two holes from the ground state $`|\psi `$ given for two holes on a ($`16\times 2)`$-ladder. Denoting the corresponding projection operator by $`P_h`$, $`P_h𝐒_i𝐒_jP_h`$ has been measured and normalized by $`\psi |P_h|\psi `$. The procedure is well described by White and Scalapino. This kind of measurement gives us a “snapshot” of the spin configuration around a dynamic hole. In the following we use the term “bond” simply to indicate that $`𝐒_i𝐒_j<0`$. If this expectation value is close to $`0.75`$ for two sites $`i`$ and $`j`$, we say that there is a “singlet bond” connecting $`i`$ and $`j`$.
For $`J^{}>1.2t`$ the rung exchange dominates. When a hole pair virtually splits up, the region between the two holes reverts to rung singlets, as shown in Fig. 3. A high energy is needed to split up the holes, but after they are separated there is no further increase in energy in moving them apart, since the diagonal exchange bonds are weak. For $`J^{}<0.7t`$ the region between the holes does something more complicated, which facilitates hole hopping at the expense of higher exchange energy on the intervening rungs. As shown in Fig. 3, various diagonal exchange bonds are formed, which turn into the rung singlets when the holes hop back together. This allows for easy hopping and low kinetic energy, but the exchange energy cost is high. For $`J^{}=0.7t`$, more and more distorted exchange bonds are created as the holes move apart. This leads to a stronger interaction for larger separation than in the case for $`J^{}=1.2t`$. This weakening of the long range interaction between the two holes explains the increase of $`\xi _{hhc}`$ in Fig. 2 in the range between $`J^{}=0.7t`$ and $`J^{}1.2t`$.
Somewhat similar analysis can be found in Ref. where the authors have argued that the exchange bonds seen in Fig. 3 are responsible for the pairing seen in the isotropic case, as well as for stripe formation.
## III The effective Model for Hole Pairs
In the strong coupling limit $`J^{}J`$, in the ground state, two bound holes are on the same rung and a description of tightly bound hole pairs moving in a background of singlet rungs is appropriate. Considering these hole pairs as hard core bosons (hcb) one can map to an effective boson model as discussed in Ref. and obtain the hole density directly from the corresponding hcb density as well as the correlations between hole pairs.
Here we propose an effective model that applies for any value of $`J^{}`$ between strong coupling and the isotropic case. Since holes on a ladder can pair with more weight on adjacent rungs, our effective model should incorporate the possibility that the “center of mass” of a hole pair can lie on a rung or between two rungs as shown in Fig. 4. Note that for even (odd) distance $`r`$ along the legs between two holes, the center of mass lies on a rung (between two rungs).
To motivate our effective model we study first the case of pairing between two holes on a two-leg ladder with length L where the Hamiltonian is given by
$`H`$ $`=`$ $`t_h{\displaystyle \underset{j=1}{\overset{2}{}}}{\displaystyle \underset{i=1}{\overset{L}{}}}(b_{i,j}^{}b_{i+1,j}+b_{i+1,j}^{}b_{i,j})`$ (8)
$`t_h{\displaystyle \underset{i=1}{\overset{L}{}}}(b_{i,1}^{}b_{i,2}+b_{i,2}^{}b_{i,1})+V_h.`$
Here $`b_i^{}`$ and $`b_i`$ denote hole creation and annihilation operators respectively. Only single occupation of a site is allowed and we use periodic boundary conditions. We choose the attractive interaction $`V_h`$ between the two holes strong enough to give an even parity bound state. We want to find an effective Hamiltonian for the paired holes, expressed in terms of a new effective boson operator $`B_i^{}`$ which describes the pair as a whole, and then answer the question, how do we obtain the hole density $`n_{i,j}^b=b_{i,j}^{}b_{i,j}`$ from the density $`N_i=B_i^{}B_i`$ of the effective bosons?
For the Hamiltonian above, with the help of the operators
$`p_{R,r}^{}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(b_{R\frac{r}{2},1}^{}b_{R+\frac{r}{2},1}^{}+b_{R\frac{r}{2},2}^{}b_{R+\frac{r}{2},2}^{})(1\delta _{r,0})`$ (9)
$`\stackrel{~}{p}_{R,r}^{}`$ $`=`$ $`\{\begin{array}{cc}\frac{1}{\sqrt{2}}(b_{R\frac{r}{2},1}^{}b_{R+\frac{r}{2},2}^{}+b_{R\frac{r}{2},2}^{}b_{R+\frac{r}{2},1}^{})\hfill & r0\hfill \\ b_{R\frac{r}{2},1}^{}b_{R+\frac{r}{2},2}^{}\hfill & r=0\hfill \end{array}`$ (12)
the wavefunction $`\mathrm{\Psi }^{}|0`$ for the ground state can be written as
$`\mathrm{\Psi }^{}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{L}}}{\displaystyle \underset{R=\mathrm{int}.}{}}{\displaystyle \underset{r_{\mathrm{even}}}{}}\left(\phi _1(r)p_{R,r}^{}+\phi _2(r)\stackrel{~}{p}_{R,r}^{}\right)`$ (14)
$`+{\displaystyle \frac{1}{\sqrt{L}}}{\displaystyle \underset{R=\mathrm{half}}{}}{\displaystyle \underset{r_{\mathrm{odd}}}{}}\left(\phi _1(r)p_{R,r}^{}+\phi _2(r)\stackrel{~}{p}_{R,r}^{}\right).`$
Here $`R`$ and $`r(=0,1,2,\mathrm{})`$ denote the center of mass and relative coordinates along the legs as sketched in Fig. 4. $`\phi _j(r)`$ depends only on $`r`$ and the interaction $`V_h`$ in Eq. (8). The pair correlation function $`g_j(r)`$ for the holes $`b_i^{}`$ in this model is simply given by $`g_j(r)=|\phi _j(r)|^2`$, where $`j=1,2`$ denote intra- and inter-leg correlations respectively.
The probability of finding the center of mass of a pair centered on a rung, $`w_{\mathrm{int}}`$, or between two rungs, $`w_{\mathrm{half}}`$, is given by
$`w_{\mathrm{int}}`$ $`=`$ $`{\displaystyle \underset{r_{\mathrm{even}}}{}}|\phi _1(r)|^2+|\phi _2(r)|^2={\displaystyle \underset{r_{\mathrm{even}}}{}}g(r)`$ (15)
$`w_{\mathrm{half}}`$ $`=`$ $`{\displaystyle \underset{r_{\mathrm{odd}}}{}}|\phi _1(r)|^2+|\phi _2(r)|^2={\displaystyle \underset{r_{\mathrm{odd}}}{}}g(r)`$ (16)
where $`g(r)`$ denotes the sum of the inter- and intra-leg correlations as defined in Eq. (6). Note that the probability to find the pair on or between rungs depends only on whether $`R`$ is integer or half integer.
The same occupation probabilities can be obtained with the Hamiltonian
$`H`$ $`=`$ $`t^{}{\displaystyle \underset{R=\frac{1}{2},1,\mathrm{}}{\overset{L}{}}}(B_R^{}B_{R+\frac{1}{2}}+B_{R+\frac{1}{2}}^{}B_R)+ϵ{\displaystyle \underset{R=\frac{1}{2},\frac{3}{2},\mathrm{}}{\overset{L+\frac{1}{2}}{}}}N_R`$ (17)
for one boson $`B_R^{}`$ which moves on a closed chain of length $`L`$ under the action of a periodically varying onsite potential $`ϵ`$. Here $`N_R=B_R^{}B_R`$.
Figure 4 shows the mapping of hole pairs from the ladder to effective bosons on a single chain. Note that the center of mass coordinate $`R`$ of the pairs determines the position of the effective boson. Here the ratio of the probabilities $`w_{\mathrm{half}}/w_{\mathrm{int}}`$ to find the boson on a site with half integer or integer $`R`$ depends only on $`ϵ/t^{}`$ and can easily be obtained as
$$\frac{w_{\mathrm{half}}}{w_{\mathrm{int}}}=\frac{ϵ^2+(4t^{})^2ϵ\sqrt{ϵ^2+(4t^{})^2}}{ϵ^2+(4t^{})^2+ϵ\sqrt{ϵ^2+(4t^{})^2}}.$$
(18)
From this equation and Eq. (15) we obtain for $`ϵ`$
$$ϵ=4t^{}\mathrm{sinh}(\mathrm{ln}\left|\sqrt{\frac{_{r_{\mathrm{even}}}g(r)}{_{r_{\mathrm{odd}}}g(r)}}\right|).$$
(19)
The boson operator $`B_R^{}`$ can be expressed in terms of the operators $`p_{R,r}^{}`$ and $`\stackrel{~}{p}_{R,r}^{}`$ as
$`B_R^{}`$ $`=`$ $`\{\begin{array}{cc}\frac{_{r_{\mathrm{even}}}\left(\phi _1(r)p_{R,r}^{}+\phi _2(r)\stackrel{~}{p}_{R,r}^{}\right)}{\left[_{r_{\mathrm{even}}}g(r)\right]^{\frac{1}{2}}}& \text{for int. }R\hfill \\ \frac{_{r_{\mathrm{odd}}}\left(\phi _1(r)p_{R,r}^{}+\phi _2(r)\stackrel{~}{p}_{R,r}^{}\right)}{\left[_{r_{\mathrm{odd}}}g(r)\right]^{\frac{1}{2}}}& \text{for half int. }R\text{.}\hfill \end{array}`$ (22)
Once the density $`N_R`$ for the model (17) has been computed, we obtain the density $`n_{i,j}^b=b_{i,j}^{}b_{i,j}`$ for the model (8), taking into account the inner structure of the effective boson (22) by the convolution
$`n_{i,j}^b`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{_{r_{\mathrm{even}}}g(r)(N_{i\frac{r}{2}}+N_{i+\frac{r}{2}})}{_{r_{\mathrm{even}}}g(r)}}`$ (24)
$`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{_{r_{\mathrm{odd}}}g(r)(N_{i\frac{r}{2}}+N_{i+\frac{r}{2}})}{_{r_{\mathrm{odd}}}g(r)}}.`$
We are interested in the ground state properties of the $`t`$-$`J`$ model. Since the spin part of the ground state wavefunction is antisymmetric and the spin excitations are gapped, the charge degrees of freedom can be described considering hole pairs as effective (hard core) bosons moving in a spin liquid.
To obtain the effective Hamiltonian describing hole pairs in the $`t`$-$`J`$ model (4), we model the holes as in Eq. (8) and define in analogy to the preceding
$`H`$ $`=`$ $`t^{}{\displaystyle \underset{R=1,\frac{1}{2},\mathrm{}}{\overset{L\frac{1}{2}}{}}}(B_R^{}B_{R+\frac{1}{2}}+B_{R+\frac{1}{2}}^{}B_R)`$ (26)
$`+ϵ{\displaystyle \underset{R=\frac{3}{2},\frac{5}{2},\mathrm{}}{\overset{L\frac{1}{2}}{}}}N_R+V_{\mathrm{int}}+V_b`$
where $`N_R=B_R^{}B_R`$ and the $`B_R^{}`$ and $`B_R`$ denote hcb creation and destruction operators respectively. Since we are using open boundaries, we have to take into account the interaction of the hole-pairs with the boundaries. The potential $`V_b`$ has been introduced to describe this effect. The potential $`V_{\mathrm{int}}`$ gives the interaction between hcb, i.e. hole pairs in the $`t`$-$`J`$ model. The onsite potential $`ϵ`$ on half-integer sites has the same meaning as explained above.
## IV Computing the model parameters
In this section we determine the effective model parameters $`t^{}`$, $`ϵ`$ and the interactions $`V_b`$ and $`V_{\mathrm{int}}`$.
### A Onsite potential $`ϵ`$
The onsite potential $`ϵ`$ has been calculated with the hhc-function $`g(r)`$ obtained numerically by the DMRG for two holes on ($`40\times 2`$) $`t`$-$`J`$ ladders for various $`J^{}`$. The results are given in Fig. 5. They clearly show that for the isotropic case, i.e. $`J=J^{}`$, where $`ϵ<0`$, the hole pair is mainly centered between two rungs, whereas for strong coupling, i.e. $`J^{}J`$, where $`ϵ>0`$, both holes of a pair sit on the same rung and can be well described by the effective model given in Ref. .
### B Interaction with the boundaries
As $`ϵ`$ is determined by the hhc-function alone, we can compute the hcb density $`N_i`$, convolute it with $`g(r)`$ according to Eq. (24) and compare it with the hole density $`n_i^h`$ of the corresponding $`t`$-$`J`$ system. The hole density $`n_{i,j}^h`$ can be obtained using Eq. (24) by identifying $`n_{i,j}^b`$ with $`n_{i,j}^h`$. In this way we obtain $`V_b`$ and $`V_{\mathrm{int}}`$ by fitting the density profile obtained with the effective model to the one obtained with the $`t`$-$`J`$ model. Since $`V_{\mathrm{int}}`$ in Eq. (17) gives no contribution for one hcb, we can obtain $`V_b`$ by considering one hole pair in the $`t`$-$`J`$ model. We choose an exponentially decreasing form for $`V_b`$
$`V_b`$ $`=`$ $`v_b{\displaystyle \underset{R}{}}N_R\left(e^{\frac{R1}{\xi _b}}+e^{\frac{LR}{\xi _b}}\right)`$ (27)
with increasing distance from the boundary. Figure 6 shows fits for $`J^{}=0.35t`$ and $`J^{}=10t`$. and the parameters $`v_b`$ and $`\xi _b`$ are shown in Fig. 7. Up to $`J^{}=3.0t`$, $`v_b`$ is positive and $`\xi _b`$ is finite. Above this value $`v_b`$ becomes negative and $`\xi _b`$ is zero. In this case only the outermost rungs become attractive for hole pairs as can be expected for large $`J^{}`$, due to the charge part $`\frac{1}{4}n_{i,j}^hn_{i+1,j}^h`$ of the $`J^{}`$ term.
### C The interaction term $`V_{\mathrm{int}}`$
To obtain the interaction potential $`V_{\mathrm{int}}`$ we proceed in the same way as for $`V_b`$. We choose a hard core analytical form
$`V_{\mathrm{int}}`$ $`=`$ $`{\displaystyle \underset{R}{}}{\displaystyle \underset{R^{}>R}{}}v_{\mathrm{int}}(|RR^{}|)N_RN_R^{}`$ (28)
$`v_{\mathrm{int}}(r)`$ $`=`$ $`\{\begin{array}{cc}\mathrm{}\hfill & \text{if }r<r_{min}\hfill \\ v_1\hfill & r=r_{min}\hfill \\ ve^{\frac{rr_{min}\frac{1}{2}}{\xi }}\hfill & \text{if }r>r_{min}\hfill \end{array}`$ (32)
with $`r_{min}1`$ and considered ($`30\times 2`$) $`t`$-$`J`$ ladders with four holes and the corresponding effective model with two hcb. The infinite repulsion for $`r=\frac{1}{2}`$ reflects the hard core condition for hole pairs, so that the minimum distance of two hole pairs in the $`t`$-$`J`$ model is 1.
Since the hole pair becomes broader with decreasing $`J^{}`$, it was useful to let $`r_{min}`$ vary. Again, we have made fits between the density profiles as shown in Fig. 8 and obtained the parameters represented in Table I. Figure 9 shows $`v_{\mathrm{int}}`$ for different $`J^{}`$ values. Starting from the isotropic case, the interaction is repulsive and becomes less long ranged as $`J^{}`$ increases. For $`J^{}=5t`$, $`v_{\mathrm{int}}`$ becomes negative for $`r=2`$. The same holds for $`J^{}=10t`$, where $`v_{\mathrm{int}}`$ vanishes for $`r>2`$. This can be expected from a perturbative approach as can be found in Ref. , where the hole pairs on a $`t`$-$`J`$ ladder have been mapped to hcb on a chain with next-nearest neighbor interaction.
We have tested the results obtained in this way by comparing the density profiles for various numbers of hole pairs and system lengths. From strong coupling down to the isotropic case we find good agreement between the density profiles obtained from the $`t`$-$`J`$ model and the effective model, as can be seen in Fig. 10.
We should note here that the fits for $`v_{\mathrm{int}}`$ were not as stable as the fits for $`v_b`$. Varying simultaneously the parameters $`v_1`$, $`v`$ and $`\xi `$ in Eq. (27), one can find another set of parameters which also gives a good fit. However, we found that the Luttinger liquid parameter $`K_\rho `$, which we calculate in the next section, only weakly depends on this ambiguity. Using only two parameters to determine $`v_{\mathrm{int}}`$ suppressing $`v_1`$ in Eq. (27), one can also fit the density profiles with a different set of parameters, but in this case $`K_\rho `$ is strongly affected.
### D The hopping matrix element $`t^{}`$
Up to now the hopping matrix element $`t^{}`$ did not play any role. We were only interested in the hcb density profile, which is not affected by the energy scale fixed by $`t^{}`$.
We compute $`t^{}`$ by finite size scaling with the ansatz that the difference between the ground state energy for two holes $`E_{2\mathrm{h}}`$ and for zero holes $`E_{0\mathrm{h}}`$ in the $`t`$-$`J`$ model is given by
$`E_{2\mathrm{h}}(L)E_{0\mathrm{h}}(L)`$ $``$ $`\mathrm{const}+t^{}E_{\mathrm{eff}}(L).`$ (33)
Here $`E_{\mathrm{eff}}`$ denotes the ground state energy of the corresponding effective model with one boson and with $`t^{}=1`$. We obtained $`t^{}`$ by fitting Eq. (33) to a straight line. The result is shown in Fig. 11.
According to the effective model given in Ref. , one can obtain the corresponding effective hopping matrix element $`t^o`$ by a finite size scaling with the ansatz
$`E_{2\mathrm{h}}(L)E_{0\mathrm{h}}(L)`$ $``$ $`\mathrm{const}+t^o(\pi /(L+1))^2.`$ (34)
Here, $`E_{2\mathrm{h}}(L)`$ and $`E_{0\mathrm{h}}(L)`$ are for the $`t`$-$`J`$ model. A perturbative estimate for $`t^o`$ is given in Ref. . For large $`J^{}`$ one obtains
$`t^o`$ $`=`$ $`2t^2/(J^{}4t^2/J^{}).`$ (35)
Figure 11 shows $`t^o`$ versus $`J^{}`$. It can be seen that for $`J^{}\mathrm{}`$ the $`t^o`$ obtained by perturbation theory tends to the one obtained by finite size scaling. Note that $`t^o`$ stands for the hopping of a hole pair from one rung to another, whereas in our effective model $`t^{}`$ mediates between $`i`$ and $`i+\frac{1}{2}`$. We can obtain $`t^o`$ from $`t^{}`$ by expanding the single boson dispersion of the effective model for small momenta $`k`$. In this way we obtain
$`t^o`$ $`=`$ $`{\displaystyle \frac{t_{}^{}{}_{}{}^{2}}{\sqrt{ϵ^2+(4t^{})^2}}}.`$ (36)
Figure 11 shows that $`t^o`$ obtained in this way from our effective model fits very well to the one obtained by finite size scaling.
An independent test for $`t^{}`$ comes from the comparison of the inverse compressibility $`\kappa ^1`$ obtained from the $`t`$-$`J`$ model and the effective model. The inverse compressibility could be obtained from
$`\kappa ^1(\rho )`$ $`=`$ $`{\displaystyle \frac{\rho ^2}{\mathrm{\Omega }}}{\displaystyle \frac{^2E_0(\rho )}{\rho ^2}}.`$ (37)
Here the volume $`\mathrm{\Omega }=2L`$ for the $`t`$-$`J`$ model and $`\mathrm{\Omega }=2L1`$ for the effective model. With $`N`$ we denote the hole or hcb number and with $`\rho =N/\mathrm{\Omega }`$ the corresponding density. The second derivative of the ground state energy with respect to the particle number $`N`$ was calculated by the discretization
$`{\displaystyle \frac{^2E_0}{N^2}}`$ $``$ $`{\displaystyle \frac{E_0(N+\mathrm{\Delta })2E_0(N)+E_0(N\mathrm{\Delta })}{\mathrm{\Delta }^2}}.`$ (38)
Here $`\mathrm{\Delta }=2`$ and $`\mathrm{\Delta }=1`$ has been used for the $`t`$-$`J`$ model and the effective model respectively. Figure 12 shows the results for $`\kappa ^1`$ computed for a ($`40\times 2`$) $`t`$-$`J`$ ladder with 2 holes and 4 holes. As can be expected, the results agree better for large $`J^{}`$, where the hole pairs are narrower and the interaction between pairs is less long-ranged, but the agreement is satisfactory for all values of $`J^{}/t`$.
## V Luttinger liquid parameter $`K_\rho `$
In the previous section we obtained the interaction between two hcb in the effective model (i.e. two hole pairs in the $`t`$-$`J`$ model). In order to obtain information about the long range correlations we have calculated the Luttinger liquid parameter $`K_\rho `$ in the effective model.
Our effective model belongs to the same universality class as the hcb model with next-nearest neighbor interaction, which in turn is equivalent to the XXZ model and which has been solved exactly by a bosonization approach and conformal field theory. From Ref. we obtain a power law decay at large distances for the charge density wave correlations and the superconducting correlations given as
$`N_rN_0`$ $``$ $`\mathrm{const}\times r^2+\mathrm{const}\times \mathrm{cos}(2\pi \rho r)r^{2K_\rho }`$ (39)
$`B_r^{}B_0`$ $``$ $`\mathrm{const}\times r^{\frac{1}{2K_\rho }}.`$ (40)
Here $`\rho `$ denotes the electron density on the ladder and half filling corresponds to $`\rho =1`$. These relations show that the superconducting correlations $`B_r^{}B_0`$ are dominant if $`K_\rho >\frac{1}{2}`$.
For hcb in one dimension, $`K_\rho `$ can also be obtained from the relations
$`K_\rho `$ $`=`$ $`\pi v_cL\left({\displaystyle \frac{^2E_0}{N_b^2}}\right)^1`$ (41)
$`v_cK_\rho `$ $`=`$ $`{\displaystyle \frac{\pi }{L}}{\displaystyle \frac{^2E_0(\mathrm{\Phi })}{\mathrm{\Phi }^2}}|_{\mathrm{\Phi }=0}.`$ (42)
Here $`E_0`$ denotes the ground state energy for a closed ring of length $`L`$ with $`N_b`$ hcb and $`E_0(\mathrm{\Phi })`$ is the ground state energy of the system penetrated by a magnetic flux $`\mathrm{\Phi }`$ which modifies the hopping by the usual Peierls phase factor, $`tt\mathrm{exp}(\pm i\mathrm{\Phi }/L)`$. From these two equations the charge velocity $`v_c`$ can be eliminated. The second derivative of the ground state energy with respect to the particle number was calculated by the discretization given in Eq. (38) with $`\mathrm{\Delta }=1`$. We used exact diagonalization for system lengths between $`32`$ and $`220`$ and with $`N_b`$ between 2 and 4. Analogously, the second derivative with respect to the magnetic flux was calculated for the same configurations.
We calculated the Luttinger liquid parameter $`K_\rho `$ for the interaction potentials given by the parameters in Table I and found the universal value $`K_\rho =1`$ as half filling is approached, for any value of $`J^{}`$ between strong coupling and the isotropic case (Figs. 13 and 14). For $`N_b/L0`$, corresponding to a very dilute hcb gas, we have $`K_\rho =1+O(N_b/L)`$, independent of the value of the interaction and consistent with Refs. . Down to $`\rho 0.875`$ the superconducting correlations are dominant, since $`K_\rho >\frac{1}{2}`$ and since we are far away from phase separation. Below $`\rho 0.875`$, $`K_\rho `$ becomes less than one half and at one quarter doping ($`\rho =0.75`$) $`K_\rho <\frac{1}{2}`$ for $`J^{}t`$. Figure 13 shows also earlier results on small clusters from Ref. and Ref. , obtained using exact diagonalization. The deviations from our results are most probably due to finite size effects close to half filling in Ref. and Ref. .
We briefly want to discuss the possibility of commensurability effects. At a filling of $`\rho =0.75`$, one might expect that commensurability effects would stabilize long ranged charge density wave ordering, since $`K_\rho `$ is quite small ($`K_\rho =0.232`$ for the isotropic case). In the finite, open systems studied here, a static CDW is pinned by the boundaries, as can be seen in Fig. 16 for the $`t`$-$`J`$ model as well as in Fig. 15 for the effective model. A careful analysis using large systems and finite size scaling is necessary to determine if the CDW order is long ranged. The decay of CDW oscillations away from the edges of the system is quite slow and is consistent with long ranged CDW order, but a very slow algebraic decay cannot be ruled out. A more detailed analysis of the effective model (using 200, 240, 280, 320 and 400 sites) shows a decreasing charge gap with increasing system size. Finite size scaling is consistent with a vanishing or possibly a small charge gap in the infinite system.
## VI conclusions
The effective model derived in this paper works well for the hole density of a two-leg $`t`$-$`J`$ ladder for various fillings. In other words, the low energy physics of hole pairs on a ladder can be well described by a model of hard core bosons on a chain with each boson representing a pair of holes. The interaction between the hard core bosons was determined by fitting the density profile obtained with the effective model to that of the $`t`$-$`J`$ model, taking into account the inner structure of the hole pair given by the hole-hole correlation function. Starting from the isotropic case, with equal exchange couplings on the rungs and legs the interaction between two hole pairs is long ranged and repulsive but becomes attractive and of nearest neighbor type when the strong coupling regime is approached. The same holds for the interaction of a hole pair with the boundaries. We choose a simple form for the interaction between the bosons in order to use only a few parameters. The results obtained from the effective model are insensitive to the specific ansatz used for this interaction. The Luttinger liquid parameter $`K_\rho `$ has been calculated for electron densities from $`\rho =0.982`$ down to $`\rho =0.75`$ (half filling corresponds to $`\rho =1`$). Down to $`\rho 0.875`$ the superconducting correlations are found to be dominant and $`K_\rho 1`$ for $`\rho 1`$. For commensurate filling, $`\rho =0.75`$, there might be true charge density wave ordering and a charge gap. Further investigations are necessary to clarify this question. The hopping matrix element for the bosons in the effective model, which allows one to calculate the inverse compressibility, could also be determined. Comparing to the $`t`$-$`J`$ model we find good agreement for strong rung couplings and only small deviation (less than $`\pm 10\%`$) near the isotropic case.
An interesting feature appeared in the hole-hole correlation function. The correlation length, $`\xi _{hhc}`$, does not monotonically increase as one approaches the isotropic case from strong rung coupling. Instead, it decreases in the interval $`0.7tJ^{}1.2t`$. This unexpected behavior can be traced to the fact that the interaction between holes is dominated by the simple rung exchange bonds in the strong rung coupling regime and by the diagonal rung exchange bonds near the isotropic case, $`J^{}J`$. Reflections of the structure in $`\xi _{hhc}`$ as a function of $`J^{}`$, shown in Fig. 2, can also be found in other properties, e.g. the interaction of the effective bosons with the boundaries and the compressibility.
###### Acknowledgements.
We wish to thank Karyn Le Hur for helpful discussions. SRW acknowledges support from the NSF under grant #DMR98-70930. The DMRG calculations have been performed on the SGI Cray SV1 of ETH Zürich. MT was supported by the Swiss National Science Foundation.
|
warning/0006/physics0006046.html
|
ar5iv
|
text
|
# 1 Problem
## 1 Problem
Deduce the form of a cylindrically symmetric plane electromagnetic wave that propagates in vacuum.
A scalar, azimuthally symmetric wave of frequency $`\omega `$ that propagates in the positive $`z`$ direction could be written as
$$\psi (𝐫,t)=f(\rho )e^{i(k_zz\omega t)},$$
(1)
where $`\rho =\sqrt{x^2+y^2}`$. Then, the problem is to deduce the form of the radial function $`f(\rho )`$ and any relevant condition on the wave number $`k_z`$, and to relate that scalar wave function to a complete solution of Maxwell’s equations.
The waveform (1) has both wave velocity and group velocity equal to $`\omega /k_z`$. Comment on the apparent superluminal character of the wave in case that $`k_z<k=\omega /c`$, where $`c`$ is the speed of light.
## 2 Solution
As the desired solution for the radial wave function proves to be a Bessel function, the cylindrical plane waves have come to be called Bessel beams, following their introduction by Durnin et al. . The question of superluminal behavior of Bessel beams has recently been raised by Mugnai et al. .
Bessel beams are a realization of super-gain antennas in the optical domain. A simple experiment to generate Bessel beams is described in .
Sections 2.1 and 2.2 present two methods of solution for Bessel beams that satisfy the Helmholtz wave equation. The issue of group and signal velocity for these waves is discussed in sec. 2.3. Forms of Bessel beams that satisfy Maxwell’s equations are given in sec. 2.4.
### 2.1 Solution via the Wave Equation
On substituting the form (1) into the wave equation,
$$^2\psi =\frac{1}{c^2}\frac{^2\psi }{t^2},$$
(2)
we obtain
$$\frac{d^2f}{d\rho ^2}+\frac{1}{\rho }\frac{df}{d\rho }+(k^2k_z^2)f=0.$$
(3)
This is the differential equation for Bessel functions of order 0, so that
$$f(\rho )=J_0(k_r\rho ),$$
(4)
where
$$k_\rho ^2+k_z^2=k^2.$$
(5)
The form of eq. (5) suggests that we introduce a (real) parameter $`\alpha `$ such that
$$k_\rho =k\mathrm{sin}\alpha ,\text{and}k_z=k\mathrm{cos}\alpha .$$
(6)
Then, the desired cylindrical plane wave has the form
$$\psi (𝐫,t)=J_0(k\mathrm{sin}\alpha \rho )e^{i(k\mathrm{cos}\alpha z\omega t)},$$
(7)
which is commonly called a Bessel beam. The physical significance of parameter $`\alpha `$, and that of the group velocity
$$v_g=\frac{d\omega }{dk_z}=\frac{\omega }{k_z}=v_p=\frac{c}{\mathrm{cos}\alpha }$$
(8)
will be discussed in sec. 2.3.
While eq. (7) is a solution of the Helmholtz wave equation (2), assigning $`\psi (𝐫,t)`$ to be a single component of an electric field, say $`E_x`$, does not provide a full solution to Maxwell’s equations. For example, if $`𝐄=\psi \widehat{𝐱}`$, then $`𝐄=\psi /x0`$. Bessel beams that satisfy Maxwell’s equations are given in sec. 2.4.
### 2.2 Solution via Scalar Diffraction Theory
The Bessel beam (7) has large amplitude only for |ρ|
<
1/ksinα
<
𝜌1𝑘𝛼\left|\rho\right|\mathrel{\vbox{\kern 0.0pt\hbox{$<$} \kern 0.0pt\hbox{$\sim$} }}1/k\sin\alpha, and maintains the same radial profile over arbitrarily large propagation distance $`z`$. This behavior appears to contradict the usual lore that a beam of minimum transverse extent $`a`$ diffracts to fill a cone of angle $`1/a`$. Therefore, the Bessel beam (7) has been called “diffraction free” .
Here, we show that the Bessel beam does obey the formal laws of diffraction, and can be deduced from scalar diffraction theory.
According to that theory , a cylindrically symmetric wave $`f(\rho )`$ of frequency $`\omega `$ at the plane $`z=0`$ propagates to point r with amplitude
$$\psi (𝐫,t)=\frac{k}{2\pi i}\rho ^{}𝑑\rho ^{}𝑑\varphi f(\rho ^{})\frac{e^{i(kR\omega t)}}{R},$$
(9)
where $`R`$ is the distance between the source and observation point. Defining the observation point to be $`(\rho ,0,z)`$, we have
$$R^2=z^2+\rho ^2+\rho ^{}_{}{}^{}22\rho \rho ^{}\mathrm{cos}\varphi ,$$
(10)
so that for large $`z`$,
$$Rz+\frac{\rho ^2+\rho ^{}_{}{}^{}22\rho \rho ^{}\mathrm{cos}\varphi }{2z}.$$
(11)
In the present case, we desire the amplitude to have form (1). As usual, we approximate $`R`$ by $`z`$ in the denominator of eq. (9), while using approximation (11) in the exponential factor. This leads to the integral equation
$`f(\rho )e^{ik_zz}`$ $`=`$ $`{\displaystyle \frac{k}{2\pi i}}{\displaystyle \frac{e^{ikz}e^{ik\rho ^2/2z}}{z}}{\displaystyle _0^{\mathrm{}}}\rho ^{}𝑑\rho ^{}f(\rho ^{})e^{ik\rho ^{}_{}{}^{}2/2z}{\displaystyle _0^{2\pi }}𝑑\varphi e^{ik\rho \rho ^{}\mathrm{cos}\varphi /z}`$ (12)
$`=`$ $`{\displaystyle \frac{k}{i}}{\displaystyle \frac{e^{ikz}e^{ik\rho ^2/2z}}{z}}{\displaystyle _0^{\mathrm{}}}\rho ^{}𝑑\rho ^{}f(\rho ^{})J_0(k\rho \rho ^{}/z)e^{ik\rho ^{}_{}{}^{}2/2z},`$
using a well-known integral representation of the Bessel function $`J_0`$.
It is now plausible that the desired eigenfunction $`f(\rho )`$ is a Bessel function, say $`J_0(k_\rho \rho )`$, and on consulting a table of integrals of Bessel functions we find an appropriate relation ,
$$_0^{\mathrm{}}\rho ^{}𝑑\rho ^{}J_0(k_\rho \rho ^{})J_0(k\rho \rho ^{}/z)e^{ik\rho ^{}_{}{}^{}2/2z}=\frac{iz}{k}e^{ik\rho ^2/2z}e^{ik_\rho ^2z/2k}J_0(k_\rho \rho ).$$
(13)
Comparing this with eq. (12), we see that $`f(\rho )=J_0(k_\rho \rho )`$ is indeed an eigenfunction provided that
$$k_z=k\frac{k_\rho ^2}{2k}.$$
(14)
Thus, if we write $`k_\rho =k\mathrm{sin}\alpha `$, then for small $`\alpha `$,
$$k_zk(1\alpha ^2/2)k\mathrm{cos}\alpha ,$$
(15)
and the desired cylindrical wave again has form (7).
Strictly speaking, the scalar diffraction theory reproduces the “exact” result (7) only for small $`\alpha `$. But the scalar diffraction theory is only an approximation, and we predict with confidence that an “exact” diffraction theory would lead to the form (7) for all values of parameter $`\alpha `$. That is, “diffraction-free” beams are predicted within diffraction theory.
It remains that the theory of diffraction predicts that an infinite aperture is needed to produce a beam whose transverse profile is invariant with longitudinal distance. That a Bessel beam is no exception to this rule is reviewed in sec. 2.3.
The results of this section were inspired by . One of the first solutions for Gaussian laser beams was based on scalar diffraction theory cast as an eigenfunction problem .
### 2.3 Superluminal Behavior
In general, the group velocity (8) of a Bessel beam exceeds the speed of light. However, this apparently superluminal behavior cannot be used to transmit signals faster than lightspeed.
An important step towards understanding this comes from the interpretation of parameter $`\alpha `$ as the angle with respect to the $`z`$ axis of the wave vectors of an infinite set of ordinary plane waves whose superposition yields the Bessel beam . To see this, we invoke the integral representation of the Bessel function to write eq. (7) as
$`\psi (𝐫,t)`$ $`=`$ $`J_0(k\mathrm{sin}\alpha \rho )e^{i(k\mathrm{cos}\alpha z\omega t)}`$
$`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}𝑑\varphi e^{i(k\mathrm{sin}\alpha x\mathrm{cos}\varphi +k\mathrm{sin}\alpha y\mathrm{sin}\varphi +k\mathrm{cos}\alpha z\omega t)}`$
$`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}𝑑\varphi e^{i(𝐪𝐫\omega t)},`$
where the wave vector q, given by
$$𝐪=k(\mathrm{sin}\alpha \mathrm{cos}\varphi ,\mathrm{sin}\alpha \mathrm{sin}\varphi ,\mathrm{cos}\alpha ),$$
(17)
makes angle $`\alpha `$ to the $`z`$ axis as claimed.
We now see that a Bessel beam is rather simple to produce in principle . Just superpose all possible plane waves with equal amplitude and a common phase that make angle $`\alpha `$ to the $`z`$ axis,
According to this prescription, we expect the $`z`$ axis to be uniformly illuminated by the Bessel beam. If that beam is created at the plane $`z=0`$, then any annulus of equal radial extent in that plane must project equal power into the beam. For large $`\rho `$ this is readily confirmed by noting that $`J_0^2(k\mathrm{sin}\alpha \rho )\mathrm{cos}^2(k\mathrm{sin}\alpha \rho +\delta )/(k\mathrm{sin}\alpha \rho )`$, so the integral of the power over an annulus of one radial period, $`\mathrm{\Delta }\rho =\pi /(k\mathrm{sin}\alpha )`$, is independent of radius.
Thus, from an energy perspective a Bessel beam is not confined to a finite region about the $`z`$ axis. If the beam is to propagate a distance $`z`$ from the plane $`z=0`$, it must have radial extent of at least $`\rho =z\mathrm{tan}\alpha `$ at $`z=0`$. An arbitrarily large initial aperture, and arbitrarily large power, is required to generate a Bessel beam that retains its “diffraction-free” character over an arbitrarily large distance.
Each of the plane waves that makes up the Bessel beam propagates with velocity $`c`$ along a ray that makes angle $`\alpha `$ to the $`z`$ axis. The intersection of the $`z`$ axis and a plane of constant phase of any of these wave moves forward with superluminal speed $`c/\mathrm{cos}\alpha `$, which is equal to the phase and group velocities (8).
This superluminal behavior does not represent any violation of special relativity, but is an example of the “scissors paradox” that the point of contact of a pair of scissors could move faster than the speed of light while the tips of the blades are moving together at sublightspeed. A ray of sunlight that makes angle $`\alpha `$ to the surface of the Earth similarly leads to a superluminal velocity $`c/\mathrm{cos}\alpha `$ of the point of contact of a wave front with the Earth.
However, we immediately see that a Bessel beam could not be used to send a signal from, say, the origin, $`(0,0,0)`$, to a point $`(0,0,z)`$ at a speed faster than light. A Bessel beam at $`(0,0,z)`$ is made of rays of plane waves that intersect the plane $`z=0`$ at radius $`\rho =z\mathrm{tan}\alpha `$. Hence, to deliver a message from $`(0,0,0)`$ to $`(0,0,z)`$ via a Bessel beam, the information must first propagate from the origin out to at least radius $`\rho =z\mathrm{tan}\alpha `$ at $`z=0`$ to set up the beam. Then, the rays must propagate distance $`z/\mathrm{cos}\alpha `$ to reach point $`z`$ with the message. The total distance traveled by the information is thus $`z(1+\mathrm{sin}\alpha )/\mathrm{cos}\alpha `$, and the signal velocity $`v_s`$ is given by
$$v_sc\frac{\mathrm{cos}\alpha }{1+\mathrm{sin}\alpha },$$
(18)
which is always less than $`c`$. The group velocity and signal velocity for a Bessel beam are very different. Rather than being a superluminal carrier of information at its group velocity $`c/\mathrm{cos}\alpha `$, a modulated Bessel beam could be used to deliver messages only at speeds well below that of light.
### 2.4 Solution via the Vector Potential
To deduce all components of the electric and magnetic fields of a Bessel beam that satisfies Maxwell’s equation starting from a single scalar wave function, we follow the suggestion of Davis and seek solutions for a vector potential A that has only a single component. We work in the Lorentz gauge (and Gaussian units), so that the scalar potential $`\mathrm{\Phi }`$ is related by
$$𝐀+\frac{1}{c}\frac{\mathrm{\Phi }}{t}=0.$$
(19)
The vector potential can therefore have a nonzero divergence, which permits solutions having only a single component. Of course, the electric and magnetic fields can be deduced from the potentials via
$$𝐄=\mathrm{\Phi }\frac{1}{c}\frac{𝐀}{t},$$
(20)
and
$$𝐁=\times 𝐀.$$
(21)
For this, the scalar potential must first be deduced from the vector potential using the Lorentz condition (19). We consider waves of frequency $`\omega `$ and time dependence of the form $`e^{i\omega t}`$, so that $`\mathrm{\Phi }/t=ik\mathrm{\Phi }`$. Then, the Lorentz condition yields
$$\mathrm{\Phi }=\frac{i}{k}𝐀,$$
(22)
and the electric field is given by
$$𝐄=ik\left[𝐀+\frac{1}{k^2}(𝐀)\right].$$
(23)
Then, $`𝐄=0`$ since $`^2(𝐀)+k^2(𝐀)=0`$ for a vector potential A of frequency $`\omega `$ that satifies the wave equation (2), etc.
We already have a scalar solution (7) to the wave equation, which we now interpret as the only nonzero component, $`A_j`$, of the vector potential for a Bessel beam that propagates in the $`+z`$ direction,
$$A_j(𝐫,t)=\psi (𝐫,t)J_0(k\mathrm{sin}\alpha \rho )e^{i(k\mathrm{cos}\alpha z\omega t)}.$$
(24)
We consider five choices for the meaning of index $`j`$, namely $`x`$, $`y`$, $`z`$, $`\rho `$, and $`\varphi `$, which lead to five types of Bessel beams. Of these, only the case of $`j=z`$ corresponds to physical, azimuthally symmetric fields, and so perhaps should be called the Bessel beam.
#### 2.4.1 $`j=x`$
In this case,
$$𝐀=\frac{\psi }{x}=\frac{k\mathrm{sin}\alpha x}{\rho }J_1(k\mathrm{sin}\alpha \rho )e^{i(k\mathrm{cos}\alpha z\omega t)}.$$
(25)
In calculating $`(𝐀)`$ we use the identity $`J_1^{}=(J_0J_2)/2`$. Also, we divide E and B by the factor $`ik`$ to present the results in a simpler form. We find,
$`E_x`$ $`=`$ $`\left\{J_0(\varrho ){\displaystyle \frac{\mathrm{sin}^2\alpha }{\rho ^2}}\left[{\displaystyle \frac{y^2J_1(\varrho )}{\varrho }}{\displaystyle \frac{x^2}{2}}\left(J_0(\varrho )J_2(\varrho )\right)\right]\right\}e^{i(k\mathrm{cos}\alpha z\omega t)},`$
$`E_y`$ $`=`$ $`{\displaystyle \frac{\mathrm{sin}^2\alpha xy}{\rho ^2}}\left[{\displaystyle \frac{J_1(\varrho )}{\varrho }}{\displaystyle \frac{1}{2}}\left(J_0(\varrho )J_2(\varrho )\right)\right]e^{i(k\mathrm{cos}\alpha z\omega t)},`$ (26)
$`E_z`$ $`=`$ $`i\mathrm{sin}2\alpha {\displaystyle \frac{x}{2\rho }}J_1(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$
where
$$\varrho k\mathrm{sin}\alpha \rho ,$$
(27)
and
$`B_x`$ $`=`$ $`0,`$
$`B_y`$ $`=`$ $`\mathrm{cos}\alpha J_0(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$ (28)
$`B_z`$ $`=`$ $`i\mathrm{sin}\alpha {\displaystyle \frac{x}{\rho }}J_1(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)}.`$
A Bessel beam that obeys Maxwell’s equations and has purely $`x`$ polarization of its electric field on the $`z`$ axis includes nonzero $`y`$ and $`z`$ polarization at points off that axis, and does not exhibit the azimuthal symmetry of the underlying vector potential.
#### 2.4.2 $`j=y`$
This case is very similar to that of $`j=x`$.
#### 2.4.3 $`j=z`$
In this case the electric and magnet fields retain azimuthal symmetry, so that it is convenient to display the $`\rho `$, $`\varphi `$ and $`z`$ components of the fields. First,
$$𝐀=\frac{\psi }{z}=ik\mathrm{cos}\alpha J_0(k\mathrm{sin}\alpha \rho )e^{i(k\mathrm{cos}\alpha z\omega t)}.$$
(29)
Then, we divide the electric and magnetic fields by $`k\mathrm{sin}\alpha `$ to find the relatively simple forms:
$`E_\rho `$ $`=`$ $`\mathrm{cos}\alpha J_1(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$
$`E_\varphi `$ $`=`$ $`0,`$ (30)
$`E_z`$ $`=`$ $`i\mathrm{sin}\alpha J_0(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$
and
$`B_\rho `$ $`=`$ $`0,`$
$`B_\varphi `$ $`=`$ $`J_1(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$ (31)
$`B_z`$ $`=`$ $`0.`$
This Bessel beam is a transverse magnetic (TM) wave. The radial electric field $`E_\rho `$ vanishes on the $`z`$ axis (as it must if that axis is charge free), while the longitudinal electric field $`E_z`$ is maximal there. Cylindrically symmetric waves with radial electric polarization are often called axicon beams .
#### 2.4.4 $`j=\rho `$
In this case,
$$𝐀=\frac{1}{\rho }\frac{\rho \psi }{\rho }=\left[\frac{J_0(k\mathrm{sin}\alpha \rho )}{\rho }k\mathrm{sin}\alpha J_1(k\mathrm{sin}\alpha \rho )\right]e^{i(k\mathrm{cos}\alpha z\omega t)}.$$
(32)
After dividing by $`ik`$, the electric and magnetic fields are
$`E_\rho `$ $`=`$ $`\{J_0(\varrho )\mathrm{sin}^2\alpha [{\displaystyle \frac{J_0(\varrho )}{\varrho ^2}}+{\displaystyle \frac{J_1(\varrho )}{\varrho }}+{\displaystyle \frac{1}{2}}(J_0(\varrho J_2(\varrho ))]\}e^{i(k\mathrm{cos}\alpha z\omega t)},`$
$`E_\varphi `$ $`=`$ $`0,`$ (33)
$`E_z`$ $`=`$ $`i\mathrm{cos}\alpha \mathrm{sin}\alpha \left[{\displaystyle \frac{J_0(\varrho )}{\varrho }}J_1(\varrho )\right]e^{i(k\mathrm{cos}\alpha z\omega t)},`$
and
$`B_\rho `$ $`=`$ $`0,`$
$`B_\varphi `$ $`=`$ $`\mathrm{cos}\alpha J_0(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$ (34)
$`B_z`$ $`=`$ $`0.`$
The radial electric field diverges as $`1/\rho ^2`$ for small $`\rho `$, so this case is unphysical.
#### 2.4.5 $`j=\varphi `$
Here,
$$𝐀=\frac{1}{\rho }\frac{\psi }{\varphi }=0.$$
(35)
After dividing by $`ik`$, the electric and magnetic fields are
$`E_\rho `$ $`=`$ $`0,`$
$`E_\varphi `$ $`=`$ $`J_0(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$ (36)
$`E_z`$ $`=`$ $`0,`$
and
$`B_\rho `$ $`=`$ $`\mathrm{cos}\alpha J_0(\varrho )e^{i(k\mathrm{cos}\alpha z\omega t)},`$
$`B_\varphi `$ $`=`$ $`0,`$ (37)
$`B_z`$ $`=`$ $`i\mathrm{sin}\alpha \left[{\displaystyle \frac{J_0(\varrho )}{\varrho }}J_1(\varrho )\right]e^{i(k\mathrm{cos}\alpha z\omega t)}.`$
These fields are unphysical due to the finite value of $`E_\varphi `$ at $`\rho =0`$, and the divergence of $`B_z`$ as $`\rho 0`$.
|
warning/0006/cond-mat0006269.html
|
ar5iv
|
text
|
# Local off-cubic distortion the cause for the low- and high-spin states of the Co3+ ion*
## Abstract
It is pointed out that it is the local symmetry that determines the realization of the low- and high-spin state of the Co<sup>3+</sup> ion. We can rigorously prove it treating the Co<sup>3+</sup> ion as the highly-correlated electron system 3d<sup>6</sup>, that fulfiles the Hund’s rules with taking into account the spin-orbit coupling. As the function of the sign of the off-cubic crystal-field distortion the magnetic or non-magnetic state is realized.
Keywords: highly-correlated electron system, crystal field, low-spin state, Co<sup>3+</sup> ion, spin-orbit coupling
PACS: 71.70.E, 75.10.D, 75.30.Gw
Receipt date by Phys.Rev.Lett. 29.03.1999
(29.03.1999)
The origin of low- and high-spin states of the Co<sup>3+</sup> ion formed in different ionic compounds is under the long debate in the 3d magnetism.<sup>1-6</sup> A non-magnetic state observed in LaCoO<sub>3</sub>, for instance, is very intriguing having in mind the strong magnetic Co state realized in CoO that is antiferromagnet with T<sub>N</sub> of 289 K. An argument that the cobalt ion in the high oxidation state (Co<sup>3+</sup>) likes to be in the low-spin state does not hold as SrCoO<sub>2.5</sub>, where also the trivalent cobalt ion exists, shows extremely strong antiferromagnetic state (T<sub>N</sub> of 570 K). Following Van Vleck the realization of the low- or high-spin state is often discussed in the one-electron model as resulting from the delicate interplay of the crystal-field and intra-atomic exchange (Hund coupling) energies (see e.g. Ref. 5). Despite of the numerous very different versions of the one-electron description (LDA, LSDA, …. +GGA, … ) this approach still has serious difficulty in the satisfactory systematic description of e.g. the LaMO<sub>3</sub> (M=Ti-Cu) series. It produces e.g. often wrongly the metallic state instead of the insulating/semiconducting state.<sup>2</sup>
The aim of this Letter is to point out that it is the local symmetry that determines the realization of the low- or high-spin state of the Co<sup>3+</sup> ion. We can prove it for the Co<sup>3+</sup> ion placed at the slightly distorted octahedral site. It turns out that in the CoO<sub>6</sub> octahedra the Co<sup>3+</sup> ion at the absolute zero temperature can have the magnetic moment as large as 3.6 $`\mu _B`$ or null as the function of the sign of the local off-cubic distortion. In case of the rhombohedral distortion the change of the sign is associated with the compression or the elongation of the octaedron along the main cube diagonal. The same strong dependence of the atomic magnetic moment holds for the tetragonal distortion.
We treat the Co<sup>3+</sup> ion in a solid as the highly-correlated electron system 3d<sup>6</sup> with 6 electrons in the unfilled 3d shell. These high correlations assure the realization of Hund’s rules. We have calculated the energy spectrum of such the system for the ground state described by the Hund’s rules quantum numbers S=2 and L=2 taking into account the spin-orbit (s-o) coupling. We have taken the spin-orbit coupling rigorously into account, not by approximate perturbation methods as is usually made in the current literature, if the s-o coupling is considered at all. This 25-level discrete energy spectrum depends on the detailed shape of the electric-field potential formed by local charge surroundings. This detailed shape of the electric potential can be represented by means of the multipolar expansion of spherical harmonics. These multipolar charge interactions in the CoO<sub>6</sub> octahedra can be accounted for by consideration of the crystal-field Hamiltonian<sup>7</sup>
$`H_{CF}=`$B$`{}_{4}{}^{d}(O_4^0`$20$`\sqrt{\text{2}}O_4^3)+`$B$`{}_{2}{}^{0}O_{2}^{0}.`$
The first term accounts for the multipolar charge interactions of the cubic symmetry with z axis taken along the main diagonal. The second term accounts for the simplest off-cubic distortion and $`O_2^0`$=3$`L_z^2`$-L(L+1). The cubic term in combination with the s-o coupling, $`\lambda _{so}`$ $`LS,`$ yields<sup>8</sup> 3-fold degenerated ground state in the $`|\text{LSL}_\text{z}\text{S}_\text{z}`$ space with the corresponding magnetic moments of 0 (singlet) and $`\pm `$3.5 $`\mu _B`$ (doublet). The off-cubic distortion splits these states making the ground state magnetic (doublet) or non-magnetic (the singlet) as the function of the sign of the B$`{}_{}{}^{0}{}_{2}{}^{}`$ parameter. The change of the sign of B$`{}_{}{}^{0}{}_{2}{}^{}`$ can be realized by the compression or the elongation of the octaedron along the main cube diagonal in the case of the rhombohedral distortion. The magnetic doublet forms the long-range magnetic state in contrast to the singlet state that yields the diamagnetic behaviour like it is in LaCoO<sub>3</sub>, for instance. These calculations have been performed with the realistic values: B$`{}_{}{}^{d}{}_{4}{}^{}`$= -11.5 meV and the spin-orbit coupling $`\lambda _{so}`$ of -18 meV. B$`{}_{}{}^{0}{}_{2}{}^{}`$$`>`$ 0 yields the non-magnetic ground state . B$`{}_{}{}^{0}{}_{2}{}^{}`$ of e.g. 2 meV produces the spin-like gap of 1.9 meV with the highly-magnetic excited doublet with the moment of 3.0 $`\mu _B`$. Such the electronic structure has been recently suggested to exist in LaCoO<sub>3</sub> by Zhuang et al.<sup>6</sup> (m=2.9 $`\mu _B`$).
In conclusion, we argue that the formation of the non-magnetic or the magnetic state of the Co<sup>3+</sup>-ion containing compounds, discussed in the current literature as the Co<sup>3+</sup> low- and high-spin states, results from the sign of the local off-cubic crystalline-electric-field distortion. It can be exactly calculated treating the Co<sup>3+</sup> ion as the highly-correlated 3d<sup>6</sup> system that experiences the cubic crystal-field interactions provided the spin-orbit coupling is correctly taken into account. The proposed mechanism, the strong correlation of the local magnetic moment and the detailed shape of the crystal-field potential experienced by the paramagnetic ion, is known to work well for rare-earth ion compounds \[10-12\] but has not been used so far for LaCoO$`_3.`$ We would like to notice that the obtained correlation of the local magnetic moment and the local symmetry is very general. In particular, it does not depend on the used parameters provided the sign of three parameters B$`{}_{}{}^{d}{}_{4}{}^{}`$, B$`{}_{}{}^{0}{}_{2}{}^{}`$ and the spin-orbit coupling $`\lambda _{so}`$ is preserved. Moreover, the proposed single-ion-like crystal-field-based model yields in the natural way the insulating state experimentally-observed for most of 3d-ion compounds like in LaCoO<sub>3</sub>.
\*This paper has been submitted 29.03.1999 to Phys.Rev.Lett. \[LC7763\] but has been rejected by the Managing Editor (Dr G.Wells) and the Editor in Chief (Dr M.Blume) in the course of a special discriminating policy with respect to our papers. Such the policy is the manipulation of Science by the Editors of Phys.Rev.Lett. and violates the fundamental scientific rules. Our request to publish our paper with negative referee reports have been ignored.
|
warning/0006/hep-ph0006321.html
|
ar5iv
|
text
|
# Introduction
## Introduction
Hard exclusive processes, such as deeply virtual Compton scattering (DVCS), owing to the QCD factorization theorems allow to probe so-called skewed parton distributions .
In this note we shall study DVCS amplitude in the parton model, our prime interest will be the question what kind of skewed parton distributions enter DVCS amplitude up to the order $`O(1/Q)`$. Our calculations follow closely ideas of Ref. . We compute the DVCS amplitude in the parton model, where the scattering of virtual photon occurs on a single parton on the mass-shell. In this approach the scattering amplitude is explicitly gauge invariant (transverse). We show that in the parton model in the case of pion target we reproduce results of Ref. . Also we give an expression for the DVCS amplitude for an arbitrary target.
We shall see that even in the simplest parton model the handbag contribution to the DVCS amplitude depend not only on the matrix elements of the light-ray operators with indices projected on the light cone direction but also on the operators projected on the transverse direction. The corresponding additional contributions make the DVCS amplitude explicitly transverse. The additional terms in the DVCS amplitude are proportional either to the transverse component of the momentum transfer or the transverse component of the target polarization, as it follows from the recent operator analysis of Ref. .
Let us emphasize that calculations in the parton model we perform here correspond to the QCD calculation of the handbag diagram in time ordered perturbation theory in the infinite momentum frame. Therefore our calculations are model independent despite the use of the approximation of the parton model. The advantage of such formalism is that at each step of the calculations the amplitude is transverse (e.m. gauge invariant). Of course, the calculations in the parton model do not allow us to determine the contributions of the operators of the type $`\overline{\psi }G\psi `$ to the amplitude. In order to determine contributions of such operators one has to perform the QCD operator analysis.
## DVCS amplitude
Here we give an expression for the handbag contribution to the DVCS amplitude computed in the parton model. The corresponding diagrams are shown in Fig. 1.
It is convenient to introduce the light-cone decomposition of the momenta:
$`p^\mu `$ $`=`$ $`(1+\xi )\stackrel{~}{n}^\mu +(1\xi ){\displaystyle \frac{\overline{M}^2}{2}}n^\mu {\displaystyle \frac{1}{2}}\mathrm{\Delta }_{}^\mu `$
$`p^\mu `$ $`=`$ $`(1\xi )\stackrel{~}{n}^\mu +(1+\xi ){\displaystyle \frac{\overline{M}^2}{2}}n^\mu +{\displaystyle \frac{1}{2}}\mathrm{\Delta }_{}^\mu `$
$`q^\mu `$ $`=`$ $`2\xi \stackrel{~}{n}^\mu +{\displaystyle \frac{Q^2}{4\xi }}n^\mu `$
$`\mathrm{\Delta }^\mu `$ $`=`$ $`(p^{}p)^\mu `$
$`\overline{M}^2`$ $`=`$ $`M_N^2{\displaystyle \frac{\mathrm{\Delta }^2}{4}}.`$ (1)
Here $`\stackrel{~}{n}^\mu `$ and $`n^\mu `$ are arbitrary light-cone vectors such that $`\stackrel{~}{n}n=1`$ and $`\stackrel{~}{n}\mathrm{\Delta }_{}=n\mathrm{\Delta }_{}=0`$. The momentum $`k`$ can be decomposed as follow:
$`k^\mu =x\stackrel{~}{n}^\mu +\beta n^\mu +k_{}^\mu ,`$ (2)
where $`\beta `$ is fixed by the mass-shell conditions for the partons. Let us note that we do not need here the explicit expression for $`\beta `$ because it contributes to the DVCS amplitude at the order $`O(\mathrm{\Delta }_{}^2/Q^2)`$. We are interested here in the order $`O(\mathrm{\Delta }_{}/Q)`$.
Computing scattering of the virtual photon on the mass-shell parton and neglecting terms of the order $`O(\mathrm{\Delta }_{}^2/Q^2)`$ and $`M^2/Q^2`$ one gets the following expression for the DVCS amplitude in the parton model:
$`M^{\mu \nu }{\displaystyle _1^1}𝑑x{\displaystyle \frac{d^2k_{}}{(2\pi )^2}d^4z\delta (nz)\frac{1}{(x\xi )(x+\xi )}\mathrm{exp}[i(kz)]}`$
$`\{\text{Tr}[(k/+{\displaystyle \frac{1}{2}}\mathrm{\Delta }/)\mathrm{\Gamma }^{\mu \nu }(k/{\displaystyle \frac{1}{2}}\mathrm{\Delta }/)n/]p^{}|\overline{\psi }({\displaystyle \frac{z}{2}})n/\psi ({\displaystyle \frac{z}{2}})|p`$ (3)
$`\text{Tr}[(k/+{\displaystyle \frac{1}{2}}\mathrm{\Delta }/)\mathrm{\Gamma }^{\mu \nu }(k/{\displaystyle \frac{1}{2}}\mathrm{\Delta }/)n/\gamma _5]p^{}|\overline{\psi }({\displaystyle \frac{z}{2}})n/\gamma _5\psi ({\displaystyle \frac{z}{2}})|p\}+O({\displaystyle \frac{\mathrm{\Delta }_{}^2}{Q^2}}),`$
where
$`\mathrm{\Gamma }^{\mu \nu }={\displaystyle \frac{\gamma ^\mu (k/\mathrm{\Delta }//2+q/)\gamma ^\nu }{(k\mathrm{\Delta }/2+q)^2+i0}}+{\displaystyle \frac{\gamma ^\nu (k/+\mathrm{\Delta }//2q/)\gamma ^\mu }{(k+\mathrm{\Delta }/2q)^2+i0}}.`$ (4)
Obviously, the amplitude (3) is transverse, $`i.e.`$ $`(q\mathrm{\Delta })^\mu M^{\mu \nu }=q^\nu M^{\mu \nu }=0`$. In the limit $`Q^2\mathrm{}`$ and $`\mathrm{\Delta }_{}^2Q^2`$ the expression (3) can be rewritten as:
$`M^{\mu \nu }`$ $`={\displaystyle \frac{1}{2}}`$ $`{\displaystyle _1^1}dx\{\alpha ^+(x,\xi )[(\stackrel{~}{n}^\mu n^\nu +\stackrel{~}{n}^\nu n^\mu g^{\mu \nu }){\displaystyle \frac{4\xi }{Q^2}}\mathrm{\Delta }_{}^\nu \stackrel{~}{n}^\mu ]n^\alpha F_\alpha `$
$`+`$ $`i\alpha ^{}(x,\xi )\left[\epsilon _{}^{\mu \nu }+{\displaystyle \frac{4\xi }{Q^2}}\stackrel{~}{n}^\mu \epsilon _{}^{\mathrm{\Delta }\nu }\right]n^\alpha F_\alpha ^{(5)}`$
$`+`$ $`[n^\nu +{\displaystyle \frac{8\xi ^2}{Q^2}}\stackrel{~}{n}^\nu ](\alpha ^+(x,\xi )(F^\mu _{}+{\displaystyle \frac{4\xi }{Q^2}}\stackrel{~}{n}^\mu (\mathrm{\Delta }_{}F))`$
$`+`$ $`i\alpha ^{}(x,\xi )(\epsilon _{}^{\mu \rho }F_\rho ^{(5)}+{\displaystyle \frac{4\xi }{Q^2}}\stackrel{~}{n}^\mu \epsilon _{}^{\mathrm{\Delta }\rho }F_\rho ^{(5)}))`$
$`+`$ $`n^\mu (\alpha ^+(x,\xi )F^\nu _{}i\alpha ^{}(x,\xi )\epsilon _{}^{\nu \rho }F_\rho ^{(5)}\left)\right\}+O\left({\displaystyle \frac{\mathrm{\Delta }_{}^2}{Q^2}}\right).`$
Here $`\epsilon _{}^{\mu \nu }=\epsilon ^{\mu \nu \alpha \beta }n_\alpha \stackrel{~}{n}_\beta `$. We also defined:
$`\alpha ^\pm (x,\xi )`$ $`=`$ $`{\displaystyle \frac{1}{x\xi +i0}}\pm {\displaystyle \frac{1}{x+\xi i0}}.`$
Skewed distributions $`F_\mu `$ and $`F_\mu ^{(5)}`$ are defined in terms of bilocal quark operators on the light-cone<sup>1</sup><sup>1</sup>1 The path ordered gauge link is assumed here. Although in the parton model we do not have interactions with the gluon fields the corresponding P-exponential can be obtained using eikonal approximation for the quark propagator in the external gluon field.:
$`F_\mu `$ $`=`$ $`{\displaystyle \frac{d\lambda }{2\pi }e^{ix\lambda }p^{}|\overline{\psi }(\frac{\lambda }{2}n)\gamma _\mu \psi (\frac{\lambda }{2}n)|p},`$
$`F_\mu ^{(5)}`$ $`=`$ $`{\displaystyle \frac{d\lambda }{2\pi }e^{ix\lambda }p^{}|\overline{\psi }(\frac{\lambda }{2}n)\gamma _\mu \gamma _5\psi (\frac{\lambda }{2}n)|p}.`$ (6)
Generically, the expression (3) contains operators off the light-cone, but all of them can be reduced to the operators (6) using obvious operator identities like:
$`\overline{\psi }(y)\left[\gamma _\alpha _\beta \gamma _\beta _\alpha \right]\psi (x)=i\epsilon _{\alpha \beta \rho \sigma }\overline{\psi }(y)\gamma _\rho \gamma _5_\sigma \psi (x),`$ (7)
which are satisfied owing to the QCD equation of motion $`/\psi (x)`$.
Note that two first terms in the expression (DVCS amplitude) coincide exactly with the improved DVCS amplitude used in . This part of the amplitude is characterized by leading-twist skewed parton distributions $`nF`$ and $`nF^{(5)}`$, other contributions are characterized by new additional skewed parton distributions $`F_\mu _{}`$ and $`F_\mu _{}^{(5)}`$. The latter functions are suppressed in differential cross section of the reaction $`e+Ne^{}+\gamma +N`$ by only one power of the hard scale $`1/Q`$ , therefore their estimates are important for the analysis of DVCS observables. In principle, considering azimithal angle and $`Q`$ dependencies of the various spin and charge asymmetries one should be able to disentangle the contributions of the additional SPD’s from the leading twist ones . Note also that in certain spin and azimuthal asymmetries the new functions enter at the same order in $`1/Q`$ as the leading twist one, as examples of such quantities are $`\mathrm{sin}(2\varphi )`$ term in the lepton spin asymmetry, and $`const`$ and $`\mathrm{cos}(2\varphi )`$ term in the lepton charge asymmetry. The measurements of such observables would give us an information on the size of new SPD’s.
It is easy to check that the amplitude (DVCS amplitude) explicitly satisfies the transversality conditions $`(q\mathrm{\Delta })^\mu M^{\mu \nu }=q^\nu M^{\mu \nu }=0`$. In the case of the pion target our expression (DVCS amplitude) coincides with the result obtained by Anikin et al. . For the nucleon target we can write for skewed parton distributions $`F^\mu `$ and $`F^{\mu (5)}`$, for instance, the following decomposition:
$`F^\mu `$ $`=`$ $`\overline{u}(p^{})\{\gamma ^\mu H(x,\xi ,\mathrm{\Delta }^2)+{\displaystyle \frac{i\sigma ^{\mu \nu }\mathrm{\Delta }_\nu }{2M_N}}E(x,\xi ,\mathrm{\Delta }^2)`$ (8)
$`+`$ $`{\displaystyle \frac{\mathrm{\Delta }_{}^\mu }{M_N}}[G_1(x,\xi ,\mathrm{\Delta }^2)+M_Nn/G_2(x,\xi ,\mathrm{\Delta }^2)]+\gamma _{}^\mu G_3(x,\xi ,\mathrm{\Delta }^2)+\mathrm{}\}u(p),`$
$`F^{\mu (5)}`$ $`=`$ $`\overline{u}(p^{})\{\gamma ^\mu \gamma _5\stackrel{~}{H}(x,\xi ,\mathrm{\Delta }^2)+{\displaystyle \frac{\mathrm{\Delta }^\mu }{2M_N}}\gamma _5\stackrel{~}{E}(x,\xi ,\mathrm{\Delta }^2)`$ (9)
$`+`$ $`{\displaystyle \frac{\mathrm{\Delta }_{}^\mu }{M_N}}\gamma _5\stackrel{~}{G}_1(x,\xi ,\mathrm{\Delta }^2)+\gamma _{}^\mu \gamma _5\stackrel{~}{G}_2(x,\xi ,\mathrm{\Delta }^2)+\mathrm{\Delta }_{}^\mu n/\gamma _5\stackrel{~}{G}_3(x,\xi ,\mathrm{\Delta }^2)+\mathrm{}\}u(p),`$
where ellipses stand for terms which do not contribute to the order $`1/Q`$. The two first functions in above equations coincide with the distributions introduced in Ref. . The functions $`G_i(x,\xi ,\mathrm{\Delta }^2)`$ $`\stackrel{~}{G}_i(x,\xi ,\mathrm{\Delta }^2)`$ are additional functions parameterizing the nucleon matrix element of quark light-cone operator. Obviously, these new functions satisfy the following sum rule:
$`{\displaystyle _1^1}𝑑x\stackrel{~}{G}_i(x,\xi ,\mathrm{\Delta }^2)=0,{\displaystyle _1^1}𝑑xG_i(x,\xi ,\mathrm{\Delta }^2)=0.`$ (10)
The additional function $`G_1(x,\xi ,\mathrm{\Delta }^2)`$ receive a contribution from the so-called D-term in the double distributions parametrization of light-ray operators . This contribution to the function $`G_1(x,\xi ,\mathrm{\Delta }^2)`$ has the form:
$`G_1(x,\xi ,\mathrm{\Delta }^2)={\displaystyle \frac{1}{2\xi }}D({\displaystyle \frac{x}{\xi }},\mathrm{\Delta }^2).`$ (11)
In the forward limit the SPD’s $`F_\mu `$ and $`F_\mu ^{(5)}`$ are:
$`F_\mu `$ $``$ $`2\left\{f_1(x)\stackrel{~}{n}_\mu +M_N^2f_4(x)n_\mu \right\}`$
$`F_\mu ^{(5)}`$ $``$ $`2\left\{g_1(x)\stackrel{~}{n}_\mu (nS)+g_T(x)S_\mu +g_3(x)M_N^2n_\mu (nS)\right\}.`$
Hence, we observe that in the forward limit the SPD’s $`F_\mu `$ and $`F_\mu ^{(5)}`$ are reduced to the combinations of twist-2 parton distributions and higher twist distributions $`f_4(x),g_3(x)`$ and $`g_T^{tw3}(x)`$. The functions $`F_\mu `$ and $`F_\mu ^{(5)}`$ can be computed in the chiral quark-soliton model using methods of Refs. .
Very interesting sum rules can be derived if we consider the second Mellin moments of the distributions $`G_i`$ and $`\stackrel{~}{G}_i`$. With help of identities like (7) one can derive relations between the second Mellin moments. Here we restrict ourselves to the following relations:
$`{\displaystyle _1^1}𝑑xxG_3(x,\xi )`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _1^1}𝑑xx\left[H(x,\xi )+E(x,\xi )\right]+{\displaystyle \frac{1}{2}}{\displaystyle _1^1}𝑑x\stackrel{~}{H}(x,\xi ),`$ (12)
$`{\displaystyle _1^1}𝑑xx\stackrel{~}{G}_2(x,\xi )`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _1^1}𝑑xx\stackrel{~}{H}(x,\xi )+{\displaystyle \frac{\xi ^2}{2}}{\displaystyle _1^1}𝑑x\left[H(x,\xi )+E(x,\xi )\right].`$ (13)
The first relation in the forward limit can be related to the quark orbital momentum because:
$`\underset{\mathrm{\Delta }^\mu 0}{lim}{\displaystyle _1^1}𝑑xxG_3(x,\xi )=J_q+{\displaystyle \frac{1}{2}}\mathrm{\Delta }q=L_q.`$ (14)
Here $`L_q`$ is the quark orbital momentum contribution to the proton spin. In derivation we used Ji’s sum rule . We see that the distribution $`G_3`$ is deeply related to the spin structure of the nucleon.
The second relation (13) is the non-forward generalization of the Efremov-Leader-Teryaev sum rule , it reduces to the ELT sum rule in the forward limit, because in the forward limit the SPD’s $`\stackrel{~}{H}`$ and $`\stackrel{~}{G}_2`$ are reduced to the spin distributions $`g_1`$ and $`g_2`$ correspondingly.
As the final remark we note that the new functions $`G_i`$ and $`\stackrel{~}{G}_i`$ can be related to the leading twist functions $`H,\stackrel{~}{H}`$ and $`E,\stackrel{~}{E}`$ if one neglects the contributions of operators of type $`\overline{\psi }G\psi `$. The method of derivation is analogous to derivation of the Wandzura–Wilczek type of relations between twist-2 and twist-3 rho-meson distributions amplitudes . Recently such relations were derived in ref. .
## Discussion
We have demonstrated that the handbag contribution to the DVCS amplitude up to the order $`1/Q`$ depends not only on the leading twist skewed parton distributions $`(nF)`$ and $`(nF^{(5)})`$ but also on the additional functions $`F_\rho _{}`$ and $`F_\rho _{}^{(5)}`$. These additional contributions are suppressed by only one power of the hard scale–$`1/Q`$ in the differential cross section. The estimates of these contributions are important for the extraction of the leading twist skewed parton densities $`H,\stackrel{~}{H}`$ and $`E,\stackrel{~}{E}`$ from observables and measuring the Ji’s sum rule . Also it is encouraging that the new functions can be also related to the spin structure of the nucleon, see eqs. (12,13). This can give additional possibility to extract quark orbital momentum from DVCS observables.
Note also that in our analysis we used the approximation $`\mathrm{\Delta }_{}^2Q^2`$, natural for the parton model. An account of the QCD evolution may lead effectively to a breakdown of this approximation and further complicate description of DVCS. Indeed, in the QCD ladder virtualities of the partons in the rungs of the ladder close to the nucleon are much smaller than $`Q^2`$. They are rather close to $`Q_0^2`$. Hence the accuracy of neglecting $`\mathrm{\Delta }_{}^2`$ in these propagators is $`\mathrm{\Delta }_{}^2/Q_0^2`$, not $`\mathrm{\Delta }_{}^2/Q^2`$. The effects due to evolution of distributions $`F_\mu _{}`$ and $`F_\mu _{}^{(5)}`$ will be studied elsewhere.
## Acknowledgements
We are grateful to V.Yu. Petrov for sharing with us his ideas and valuable contributions to calculations. Discussions with L. Frankfurt, K. Goeke, N. Kivel, D. Müller, P.V. Pobylitsa, A. Radyushkin, A. Schäfer, J. Soffer, O. Teryaev, M. Vanderhaeghen are greatly acknowledged. M.V.P. is thankful to C. Weiss for many interesting discussions. Critical remarks of M. Diehl and N. Kivel were of big help for us.
|
warning/0006/hep-th0006145.html
|
ar5iv
|
text
|
# A note on D-branes in group manifolds: flux quantisation and D0-charge
## 1. Introduction
Recently the issues of D0 charge and $`\mathrm{U}(1)`$ flux quantisation for a class of D-branes in the $`\mathrm{SU}(2)`$ group manifold have attracted a great deal of attention . Our aim here is to show, using a somewhat different line of argument, that the same basic results can be obtained without any reference to a hypothetical $`\mathrm{U}(1)`$ gauge field or, indeed, of its flux.
In this letter we show, using the formalism developed in , that a D-brane in a group manifold sitting on a (twisted) conjugacy class $`𝒞`$, and described, in the framework of the boundary state approach, by the matrix $`R`$ of gluing conditions is characterised by a gauge invariant two-form field $`\omega `$ defined on the worldvolume of the D-brane whose components are determined by $`R`$. By comparing the boundary conditions coming from the gluing conditions with the ones deduced from the classical sigma model action, we are able to identify this two-form field $`\omega `$ with the gauge-invariant combination $`B+2\pi \alpha ^{}F`$.
In order to write the boundary WZW action in terms of the three-form $`H`$ and the two-form $`\omega `$ one is forced to introduce, much as in the case of the standard WZW model, a certain field extension $`\stackrel{~}{g}`$. The requirement that the quantum theory be independent of the choice of field extension imposes two quantisation conditions : the first one, imposed on $`H`$ alone and similar to the closed string case, is an integrality condition on the cohomology class of $`H`$ in $`H^{}(𝐆)`$; whereas the second is that the periods of $`(H,\omega )`$ over cycles in the relative homology $`H_3(𝐆,𝒞)`$ take integer values. In the case of $`\mathrm{SU}(2)`$ at level $`k`$, these conditions yield $`k+1`$ D-branes: two point-like D-branes situated at the two elements in the centre of $`\mathrm{SU}(2)`$, and $`k1`$ spherical D2-branes .
A closer look at this quantisation condition suggests a natural definition for the D0-brane charge of a given D-brane, which reads
$$\frac{1}{4\pi ^2T}Q_0=\frac{1}{2\pi }(_{}\stackrel{~}{g}^{}\omega _{}\stackrel{~}{g}^{}H)(modk),$$
(1)
where $``$ is a three-manifold such that $`g()=𝒞`$. This quantity is naturally gauge invariant and quantised (with integer values), independently of any assumption regarding the existence of a $`\mathrm{U}(1)`$ gauge field on the brane. In the particular case of the D2-branes in $`\mathrm{SU}(2)`$, the $`H`$ field contribution is similar to the Poynting-type bulk contribution advocated in , and is valid also when $`H`$ belongs to a nontrivial cohomology class. This quantity can be thought of as a generalisation of the $`\mathrm{U}(1)`$ flux in the case where the three-form field $`H`$ belongs to a nontrivial cohomology class.
## 2. Semi-classical analysis
### 2.1. Boundary conditions from the boundary state approach
We consider D-branes in a group $`𝐆`$ which preserve conformal invariance and the infinite-dimensional symmetry of the current algebra of the bulk theory. They are described in terms of the following gluing conditions:
$$J=R\overline{J},$$
(2)
where the matrix of gluing conditions $`R:𝔤𝔤`$ is a metric-preserving automorphism of the Lie algebra $`𝔤`$; that is,
$$[R(T_a),R(T_b)]=R([T_a,T_b]),$$
(3)
and
$$R^T\mathrm{\Omega }T=\mathrm{\Omega },$$
(4)
in the obvious notation. This type of gluing conditions describe D-branes whose worldvolumes lie on twisted conjugacy classes. More precisely, $`𝖣`$-branes in a WZW model with group $`𝐆`$ come in several types, classified by the group $`\mathrm{Out}_o(𝐆)`$ of metric-preserving outer automorphisms of $`𝐆`$, which is defined as the quotient $`\mathrm{Aut}_o(𝐆)/\mathrm{Inn}_o(𝐆)`$ of the group of metric-preserving automorphisms by the invariant subgroup of inner automorphisms.
In it was shown that the above gluing conditions give rise, at a given point $`g`$ in $`𝐆`$, to the following boundary conditions
$$g=𝐑(g)\overline{}g,$$
where the map $`𝐑(g):T_g𝐆T_g𝐆`$ is defined as
$$𝐑(g)=(\rho _g)_{}R(\lambda _g)_{}^1.$$
For the purposes of this paper it will be convenient to write the above boundary conditions in a different form. We therefore parametrise $`𝐆`$ by introducing the coordinates $`X^\mu `$, with $`\mu =1,\mathrm{},dim𝐆`$; we also introduce the left- and right-invariant vielbeins
$$g^1dg=e_{}^{a}{}_{\mu }{}^{}dX^\mu T_a\text{and}dgg^1=\overline{e}^a{}_{\mu }{}^{}dX^\mu T_a.$$
These vielbeins are related by $`\overline{e}^a{}_{\mu }{}^{}=e_{}^{b}{}_{\mu }{}^{}A_{}^{a}{}_{b}{}^{}`$, where $`A`$ denotes the adjoint action of the group: $`gT_ag^1=A_{}^{b}{}_{a}{}^{}T_b`$. Using this set-up, one can easily see that the gluing conditions (2) give rise to the following boundary conditions for the component fields $`X^\mu `$:
$$X^\mu =\stackrel{~}{R}(g)^\mu {}_{\nu }{}^{}\overline{}X^\nu ,$$
where the matrix of boundary conditions $`\stackrel{~}{R}(g)`$ is given by
$$\stackrel{~}{R}(g)=\overline{e}^1Re.$$
(5)
A Dirichlet direction is determined by an eigenvector of $`\stackrel{~}{R}(g)`$ with eigenvalue $`1`$, whereas all the other eigenvectors correspond to Neumann directions, that is, directions tangent to the worldvolume of the D-brane.
If we parametrise the worldsheet of the string by $`(\sigma ,\tau )`$ we can rewrite the above boundary conditions in the following form
$$i(\mathbb{𝟙}+\stackrel{~}{R})_\sigma X=(\mathbb{𝟙}\stackrel{~}{R})_\tau X,$$
(6)
where $`,\overline{}=_\tau i_\sigma `$. We know that in this case the worldvolume of a D-brane passing through $`g`$ and being described by (2) is given by the twisted conjugacy class $`𝒞_R(g)`$. We therefore consider the following split
$$T_g𝐆=T_g𝐆^{||}T_g𝐆^{},$$
(7)
of the tangent space of $`𝐆`$ at the point $`g`$, where $`T_g𝐆^{||}`$ is the tangent space to the twisted conjugacy class, and $`T_g𝐆^{}`$ is its orthogonal complement. Using this, one can split the boundary conditions (6) into two sets of conditions:
$`i(\mathbb{𝟙}+\stackrel{~}{R})_\sigma X^{||}`$ $`=(\mathbb{𝟙}\stackrel{~}{R})_\tau X^{||},`$
$`i(\mathbb{𝟙}+\stackrel{~}{R})_\sigma X^{}`$ $`=(\mathbb{𝟙}\stackrel{~}{R})_\tau X^{},`$
in the obvious notation. Since $`\stackrel{~}{R}|_{T_g𝐆^{}}=\mathbb{𝟙}`$, from the second equation above we obtain the Dirichlet boundary conditions
$$_\tau X^{}=0.$$
On the other hand, by using the fact that $`(\mathbb{𝟙}+\stackrel{~}{R})|_{T_g𝐆^{||}}`$ is invertible, we obtain the Neumann boundary conditions
$$_\sigma X^{||}+i\frac{\mathbb{𝟙}\stackrel{~}{R}(g)}{\mathbb{𝟙}+\stackrel{~}{R}(g)}_\tau X^{||}=0,$$
We will now show that the matrix which defines the above Neumann boundary conditions coincides with the one defining the two-form $`\omega `$ on the worldvolume of the D-brane.
### 2.2. Boundary conditions from the sigma model
In the next section we will briefly review the definition of the boundary WZW model. In particular we will see that the action (12) of a generic WZW model on a 2-space with a disc topology is specified in terms of the three-form field $`H`$, familiar from the standard case (when the worldsheet has no boundary), and a two-form field $`\omega `$ defined on the worldvolume $`𝒞`$ of the D-brane, and satisfying $`d\omega =H|_𝒞`$.
The infinitesimal variation of the boundary WZW action contains a bulk term (yielding the same equations of motion as in the closed string case) and a boundary term which reads
$$_\mathrm{\Sigma }𝑑\tau (g^1\delta g)^a\left[G_{ab}(g^1_\sigma g)^bi(g^{}\omega )_{ab}(g^1_\tau g)^b\right]|_{\sigma =0}^{\sigma =\pi }$$
where we have denoted by $`G`$ is the bi-invariant metric on the group manifold.
Here we are interested in D-branes described by (twisted) conjugacy classes. Thus, in order to separate the Neumann and Dirichlet boundary conditions encoded in the boundary term above, we must make use of the specific form of a conjugacy class. We recall (for details see, e.g., ) that this is defined as
$$𝒞_R(g_0)=\{g=r(h)g_0h^1h𝐆\},$$
where the map $`r:𝐆𝐆`$ is defined by
$$r\left(e^{tT}\right)=e^{tR(T)},$$
for $`t`$ small enough and $`T`$ any element in the Lie algebra.
Hence in this case $`g`$ maps the boundary of the worldsheet $`\mathrm{\Sigma }`$ into the conjugacy class $`𝒞_R(g_0)`$ that is, $`g(\mathrm{\Sigma })𝒞_R(g_0)`$, and therefore
$$g^1\delta g|_{𝒞_R(g_0)}=(\mathrm{Ad}_{g^1}R\mathbb{𝟙})\delta hh^1.$$
Assuming that the metric restricts nondegenerately to the worldvolume of the D-brane (this is only a restriction in pseudo-riemannian signature), then the infinitesimal variation $`g^1\delta g`$ can be written as the sum of two terms given by
$`(g^1\delta g)^{||}`$ $`=(\mathrm{Ad}_{g^1}R\mathbb{𝟙})^{||}\delta hh^1,`$
$`(g^1\delta g)^{}`$ $`=(\mathrm{Ad}_{g^1}R\mathbb{𝟙})^{}\delta hh^1.`$
Since $`(\mathrm{Ad}_{g^1}R\mathbb{𝟙})^{}=0`$, the second equation above yields the Dirichlet boundary conditions
$$(g^1\delta g)^{}=0,$$
whereas the boundary term in the infinitesimal variation of the action becomes
$$_\mathrm{\Sigma }(\mathrm{Ad}_{g^1}R\mathbb{𝟙})^{||}(\delta hh^1)\left[G(g^1_\sigma g)i(g^{}\omega )(g^1_\tau g)\right]^{||}.$$
Taking into account that $`(\mathrm{Ad}_{g^1}R\mathbb{𝟙})^{||}`$ is nondegenerate, we obtain the Neumann boundary conditions
$$(g^1_\sigma g)^{||}iG^1(g^{}\omega )(g^1_\tau g)^{||}=0.$$
If we now consider the field $`g`$ to be parametrised as in the previous paragraph, we can rewrite the Dirichlet and Neumann boundary conditions in the following form
$$\delta X^{}=0,$$
$$_\sigma X^{||}i\stackrel{~}{G}^1\stackrel{~}{\omega }_\tau X^{||}=0,$$
where $`\stackrel{~}{G}=e^TGe`$, $`\stackrel{~}{\omega }=e^T(g^{}\omega )e`$.
### 2.3. The two-form field $`\omega `$
By identifying now the Neumann boundary conditions obtained from the boundary state approach with the Neumann conditions obtained from the classical sigma model, we can deduce that the two-form $`\omega `$ is uniquely determined by the matrix of gluing conditions $`R`$. Indeed we first deduce that
$$\stackrel{~}{\omega }=\frac{1}{2}dX,\frac{\mathbb{𝟙}\stackrel{~}{R}(g)}{\mathbb{𝟙}+\stackrel{~}{R}(g)}dX,$$
from where we finally obtain that
$$g^{}\omega =\frac{1}{2}g^1dg,\frac{\mathbb{𝟙}+\mathrm{Ad}_{g^1}R}{\mathbb{𝟙}\mathrm{Ad}_{g^1}R}g^1dg.$$
(8)
Notice that this form is well defined on $`𝒞_R(g_0)`$, as $`\mathbb{𝟙}+\stackrel{~}{R}(g_0)`$ is invertible on $`T_{g_0}𝐆^{||}`$. One can easily check that $`g^{}\omega `$ is antisymmetric, hence it does indeed define a differential two-form. We know that the basic property that this field must satisfy is
$$d(g^{}\omega )=g^{}H|_{𝒞_{R(g_0)}},$$
(9)
where $`H`$ is the WZW three-form. In order to verify that the two-form field defined in (8) does indeed satisfy this property, we use the fact that the left-invariant Maurer–Cartan form evaluated on $`𝒞_R(g_0)`$ reads
$$g^1dg|_{𝒞_R(g_0)}=(\mathrm{Ad}_{g^1}R\mathbb{𝟙})dhh^1.$$
This allows us to evaluate $`H`$ on the conjugacy class
$$g^{}H|_{𝒞_R(g_0)}=ddhh^1,\mathrm{Ad}_{g^1}R(dhh^1).$$
As expected, we obtain that the three-form field $`g^{}H`$ is trivial in de Rham cohomology when restricted to the (twisted) conjugacy class. Furthermore, for $`\omega `$ itself we obtain
$$g^{}\omega =dhh^1,\mathrm{Ad}_{g^1}R(dhh^1).$$
We thus conclude that a D-brane configuration which is given by (2) and described geometrically by a (twisted) conjugacy class $`𝒞_R`$ in a group manifold $`𝐆`$ is endowed with a two-form field $`\omega `$ which is uniquely determined in terms of the matrix of gluing conditions $`R`$. This implies, in particular, that if one makes a certain gauge choice for the $`B`$ field in the bulk, then the field $`F`$ on a given D-brane is uniquely determined in terms of $`\omega `$ and the pull-back on the worldvolume of the D-brane of that $`B`$ field.
## 3. Quantum considerations
We recall that the WZW model is defined by the action
$$I[g]=_\mathrm{\Sigma }g^1g,g^1\overline{}g+_{}\stackrel{~}{g}^{}H,$$
where $`G_{ab}=T_a,T_b`$ defines a bi-invariant metric on the group manifold, $`\mathrm{\Sigma }`$ is Riemann surface without boundary, and $``$ is a three-manifold with boundary $`=\mathrm{\Sigma }`$. As is well known, the WZ term in this case is a nonlocal term, defined in terms of an extension $`\stackrel{~}{g}:𝐆`$ of the map $`g`$, such that $`\stackrel{~}{g}|_{=\mathrm{\Sigma }}=g`$, and given by
$$\stackrel{~}{g}^{}H=\frac{1}{3}\stackrel{~}{g}^1d\stackrel{~}{g},d(\stackrel{~}{g}^1d\stackrel{~}{g}).$$
(10)
Thus the WZ term depends on the choice of extension $`\stackrel{~}{g}`$, which introduces in the action $`I[g]`$ an ambiguity proportional to the periods of $`H`$ over the integer homology $`H_3(𝐆)`$. At the classical level these discrete contributions are not relevant, as they do not affect the equations of motion. However at the quantum level, the requirement that the path integral be independent of the choice of extension $`\stackrel{~}{g}`$ will in general fix the metric. In the case of $`𝐆`$ a compact simple Lie group, the metric can be fixed uniquely by the requirement that
$$\frac{1}{2\pi }_ZH=k,$$
where $`Z`$ is a $`3`$-cycle in $`𝐆`$ representing the generator of $`H_3(𝐆)\mathrm{}`$, and $`k`$ is a positive integer (the level).
The boundary WZW model was analysed in some detail in ; here we review a few aspects of particular interest for our discussion. The classical theory is usually defined by an action
$$S[g]=_\mathrm{\Sigma }g^1g,g^1\overline{}g+_\mathrm{\Sigma }g^{}B+_\mathrm{\Sigma }g^{}A.$$
(11)
In this case the worldsheet $`\mathrm{\Sigma }`$ is a two-dimensional manifold with boundary $`\mathrm{\Sigma }`$, and $`B`$ represents a particular choice for the antisymmetric tensor field, such that $`dB=H`$. A D-brane configuration is characterised in this setting by a two-form $`\omega `$ defined on its worldvolume $`𝒞`$, and satisfying $`d\omega =dB|_𝒞=H|_𝒞`$. Since $`d(B\omega )|_𝒞=0`$, one can define locally the one-form potential $`A`$ such that $`dA=B\omega `$.
One can write the boundary WZW action a manifestly gauge invariant form, by using the three-form $`H`$ and the two-form $`\omega `$. Let us assume, for simplicity, that we have one D-brane sitting on a (twisted) conjugacy class $`𝒞`$ in $`𝐆`$. In this case the worldsheet $`\mathrm{\Sigma }`$ can be represented in terms of a closed surface $`\mathrm{\Sigma }^{}`$, where $`\mathrm{\Sigma }=\mathrm{\Sigma }^{}\backslash D`$, and $`D`$ is a (unit) disk embedded in $`\mathrm{\Sigma }^{}`$. Provided that some topological obstructions can be overcome (which amount to the vanishing of the relative homology group $`H_2(𝐆,𝒞)`$), one can then extend $`g`$ to a map $`g^{}:\mathrm{\Sigma }^{}𝐆`$ such that $`g^{}(D)𝒞`$, and $`g^{}`$ can be further extended to a map $`\stackrel{~}{g}^{}:B^{}𝐆`$, with $`B^{}`$ a three-dimensional manifold such that $`B^{}=\mathrm{\Sigma }^{}`$. This allows us to write the WZ term in a more familiar form
$$S[g]=_\mathrm{\Sigma }g^1g,g^1\overline{}g+_B^{}\stackrel{~}{g}^{{}_{}{}^{}}H_Dg^{{}_{}{}^{}}\omega .$$
(12)
Thus, in this case, the WZ term has a bulk component and a boundary component (defined on the worldvolume of the D-brane). Similarly to the standard case, the WZ term depends on the extension $`\stackrel{~}{g}^{}`$, which introduces an ambiguity in the action
$$\left(_\stackrel{~}{}\stackrel{~}{g}^{}H_{S^2}\stackrel{~}{g}^{}\omega \right),$$
(13)
where $`\stackrel{~}{}`$ is a three-dimensional manifold with $`\stackrel{~}{}=S^2`$ and $`\stackrel{~}{g}:\stackrel{~}{}𝐆`$ such that $`\stackrel{~}{g}(S^2)𝒞`$. As shown in these are proportional to the periods of $`(H,\omega )`$ over the cycles of the relative homology $`H_3(𝐆,𝒞)`$.
In order to evaluate this ambiguity and compare it to the standard case without boundary it is convenient to “fill” $`\stackrel{~}{}`$ with the unit ball $``$ (whose boundary is $`S^2`$), ending up with a three-dimensional manifold $`\widehat{}`$ without boundary; if we also extend $`\stackrel{~}{g}`$ to a map $`\widehat{g}:\widehat{}𝐆`$, we can rewrite (13) as the sum of two terms, where the first one
$$_\widehat{}\widehat{g}^{}H,$$
(14)
has the same form as the ambiguity appearing in the standard WZW action. This fixes the metric just as in the previous case, to ensure that (14) induces no dependence on our field extensions at the level of the path integral. This leaves us with the second term
$$\left(_{}\stackrel{~}{g}^{}\omega _{}\stackrel{~}{g}^{}H\right).$$
(15)
which is characteristic to the boundary WZW model. Hence if we want that the path integral be independent of $`\stackrel{~}{g}`$, this term must take values in $`2\pi \mathrm{}`$. This can be thought of as a generalisation of the Dirac quantisation condition.
## 4. The $`\mathrm{SU}(2)`$ case
### 4.1. (Semi-)classical analysis
Let us now apply the above discussion to the particular case of D-brane configurations given by conjugacy classes in $`\mathrm{SU}(2)`$. We use the following parametrisation :
$$g=e^{i(\psi _1\sigma _1+\psi _2\sigma _2+\psi _3\sigma _3)},$$
where $`(\psi _1,\psi _2,\psi _3)`$ forms a vector of length $`\psi `$ pointing in the direction $`(\theta ,\varphi )`$, and $`(\sigma _1,\sigma _2,\sigma _3)`$ are the Pauli matrices. This parametrisation, whose spacetime fields are $`(\psi ,\theta ,\varphi )`$, has the advantage that one of the coordinates, namely $`\psi `$, corresponds to the Dirichlet direction, as we will see explicitly in a moment. We can compute, as usual, the invariant vielbeins $`e`$ and $`\overline{e}`$, the sigma model metric $`G`$, and the Wess–Zumino three-form $`H`$ thus obtaining
$$G=\frac{k}{2\pi }\left(d\psi ^2+\mathrm{sin}^2\psi (d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2)\right),$$
$$H=\frac{k}{\pi }\mathrm{sin}\theta \mathrm{sin}^2\psi d\psi d\theta d\varphi ,$$
where the level $`k`$ is a positive integer. Furthermore, by using (5), we obtain the matrix of boundary conditions
$$\stackrel{~}{R}(g)=\left(\begin{array}{ccc}1& 0& 0\\ 0& \mathrm{cos}2\psi & \mathrm{sin}\theta \mathrm{sin}2\psi \\ 0& \mathrm{sin}2\psi \mathrm{csc}\theta & \mathrm{cos}2\psi \end{array}\right).$$
From the form of this matrix we can immediately read off that we always have a Dirichlet boundary condition along the “radial” coordinate $`\psi `$. In other words, the D-branes described by these gluing conditions are normal to the $`\psi `$ direction at any given point. On the other hand, we know that these D-branes are conjugacy classes in $`\mathrm{SU}(2)`$—hence every such conjugacy class $`𝒞=𝒞(\psi )`$ is a two-sphere centred around the identity. In particular, for $`\psi =0,\pi `$ we get the zero-dimensional D-branes since $`\stackrel{~}{R}=\mathbb{𝟙}`$.
According to the discussion in Section 2, we can now calculate the two-form field $`\omega `$ associated to a given D-brane $`𝒞(\psi )`$ obtaining
$$\stackrel{~}{\omega }=\frac{k}{4\pi }\mathrm{sin}2\psi \mathrm{sin}\theta d\theta d\varphi .$$
Using the explicit knowledge of this field, we can evaluate the energy of such a configuration from the Born–Infeld action
$`E(\psi )`$ $`=2\pi T{\displaystyle _{𝒞(\psi )}}\sqrt{det(\stackrel{~}{g}+\stackrel{~}{\omega })}`$
$`=4\pi kT\mathrm{sin}\psi ,`$ (16)
where we denoted by $`T`$ the D-brane tension, and by $`\stackrel{~}{g}`$ the metric induced on the worldvolume of the D-brane. From this expression we can immediately see that the energy of such a D-brane configuration reaches its minimum for $`\psi =0,\pi `$, i.e., for the zero-dimensional D-branes. Hence, from a classical point of view, it is only the two D0-branes that give rise to stable configurations.
Let us compare our result with the one obtained in . The main difference between the two approaches lies in the way one determines the $`B+2\pi \alpha ^{}F`$ field. In some gauge choices were involved. Here, this field was determined uniquely, by identifying the boundary conditions coming from the gluing conditions with the ones of the classical sigma model, and thus the expression for the energy (4.1) holds independently of any particular gauge choice. One could argue that the only necessary conditions are that $`dB=H`$ and $`dF=0`$. However, as we showed in Section 2, consistency between the gluing conditions and the classical sigma model boundary conditions constrains $`B+2\pi \alpha ^{}F`$ to be equal to the two-form field $`\omega `$. One might also expect that the energy minimisation procedure itself selects the right combination, but the different results obtained in the two approaches indicate that this is not the case. It is perhaps useful to remark that the gluing conditions fix the shape of the D-brane (spherical in this case) whereas minimising the energy basically fixes its size. Moreover, there is an infinite number of D2-branes which, despite the fact that they satisfy the gluing conditions, do not minimise the Born–Infeld action.
### 4.2. Quantum analysis
Let us now apply the quantum considerations of the previous section to our D-branes $`𝒞(\psi )`$. To this end, let us compute the period of $`(H,\omega )`$ over a cycle in $`H_3(𝐆,𝒞)`$. In this case $`=(\psi )`$ is a three-ball bounded by $`𝒞(\psi )`$ and we calculate
$$\left(_{C(\psi )}\stackrel{~}{g}^{}\omega _{(\psi )}\stackrel{~}{g}^{}H\right)=2k\psi .$$
Hence, in order for the path integral to be independent of our field extensions, $`\psi `$ must be quantised as follows
$$\psi _n=\frac{n\pi }{k},n=0,1,\mathrm{},k.$$
This result, which agrees with the analysis of (for a detailed exposition see also ), allows us to conclude that the $`k+1`$ D-branes singled out in this fashion are stable. Moreover if we evaluate their masses by using the Born–Infeld action, we obtain
$$M_n=4\pi kT\mathrm{sin}\left(\frac{n\pi }{k}\right),n=0,1,\mathrm{},k,$$
which, as pointed out in , agrees with the CFT calculations.
Such a quantisation condition appears to be a non-local condition imposed on $`\psi `$, and concerns about its physical meaningfulness are well founded<sup>1</sup><sup>1</sup>1I thank M Douglas for raising this point.. It seems clear that the reason why $`\psi `$ is determined to a particular value is due to the fact that we are analysing a very symmetric type of D-brane configurations, namely those given by conjugacy classes. Presumably, it is possible to deform the D-brane while preserving the charge and the mass and while respecting conformal invariance of the boundary conditions in such a way that the condition on $`\psi `$ gets smeared.
### 4.3. D0 charge
These results prompt us to propose the following definition for the D0-charge of such D-branes:
$$\frac{1}{4\pi ^2T}Q_0=\frac{1}{2\pi }(_{}\stackrel{~}{g}^{}\omega _{}\stackrel{~}{g}^{}H)(modk).$$
(17)
In the particular case of a D-brane given by $`𝒞(\psi )`$ one obtains
$$Q_0(\psi )=4\pi kT\psi (modk),$$
which takes integer values for the $`k+1`$ D-branes obtained before.
We remark here that these values of the D0 charge appear to be different from the results one obtains from the CFT calculations. On the other hand, if we compute the flux of $`\omega `$ alone, one obtains
$$Q^{\text{BDS}}(\psi )=2\pi T_{𝒞(\psi )}\stackrel{~}{g}^{}\omega =2\pi kT\mathrm{sin}2\psi ,$$
(18)
which agrees, in the case of the $`k+1`$ stable configurations, with and with the CFT results. This seems to indicate that the path integral and the boundary state approach compute two distinct quantities, as suggested recently in (we will come back to this point). Notice that in the regime where $`nk`$ one obtains $`Q_0(n)Q^{\text{BDS}}(n)M_n`$, which is nothing but the mass and charge of $`n`$ D-particles.
This definition for the D0-brane charge has the following virtues: it is manifestly gauge invariant, as the one introduced in , but unlike the one based on the flux of the $`\mathrm{U}(1)`$ field . It is naturally quantised with integer values, as is natural to expect of a RR charge. Moreover, it includes a contribution coming from the bulk field $`H`$, similar to the one advocated in . Notice however that this bulk term does not cancel the $`B`$ field contribution included in the flux of $`\omega `$. It is perhaps useful to discuss this point also in the framework used in , where the Poynting-type contribution to the D0-brane charge reads
$$\frac{1}{6}G_{0ijk}^{(4)}H^{ijk},$$
with obvious notation. In evaluating this contribution we must take into account that, in this case, $`H`$ belongs to a nontrivial cohomology class and hence there is no globally defined gauge invariant $`B`$ field<sup>2</sup><sup>2</sup>2A similar observation was made independently by A Tseytlin.. Therefore this results in a bulk contribution which agrees with the bulk term in our definition of $`Q_0`$ in (17).
Last but not least, notice that, although the relative cohomology class of $`(H,\omega )`$, on which the definition of $`Q_0`$ is based, is not the same as the flux of the gauge field $`F`$ defined on the conjugacy class, it can nevertheless be thought of as a natural generalisation of the $`\mathrm{U}(1)`$ flux in the case of a D-brane in a WZW background, since it reduces to this in the particular case where $`H`$ is an exact form, as one can verify by using the Stokes theorem in the second term in (17).
## 5. Discussion
In this letter we have analysed a class of D-brane configurations in group manifolds, which is characterised by gluing conditions that preserve the maximum possible amount of symmetry of the bulk theory, namely, the current algebra of the WZW theory. We know that the gluing conditions generally fix the “shape” of a D-brane; in particular, this type of gluing conditions give rise to D-branes described by (twisted) conjugacy classes. Here we have shown that consistency between the gluing conditions and the sigma model boundary conditions also fixes the gauge invariant field $`B+2\pi \alpha ^{}F`$. Using this fact, one can estimate the energy of such a D-brane configuration, from the Born–Infeld action, independently of any particular gauge choice for either $`B`$ or $`F`$. One thus obtains that in the $`\mathrm{SU}(2)`$ case, at the classical level, it is only the two D0-branes that are stable. We have then used a quantisation argument based on the path integral which requires that the periods of $`(H,\omega )`$ over the cycles of the relative homology $`H_3(𝐆,𝒞)`$ take integer values; in the $`\mathrm{SU}(2)`$ case this produces a discrete set of allowed D-brane configurations, whose mass spectrum agrees with the CFT calculations.
This quantisation argument also prompted us to make an alternative proposal for the D0-brane charge of such D-branes, which differs from the one introduced in by a bulk term, similar, yet not identical, to the one advocated in . We believe this to be a natural definition for a number of reasons. First of all it is manifestly gauge invariant, and we consider this to be an important feature, as both the boundary WZW model and the Born–Infeld action of a D-brane in such a background are manifestly gauge invariant. This D0 charge is also naturally quantised, taking integer values, and this is clearly a desirable feature for a RR charge. Finally, this definition constitutes a natural generalisation of the $`\mathrm{U}(1)`$ flux.
It is important to remark that this definition also raises a number of interesting questions. One of them is the apparent discrepancy between $`Q_0`$ and the CFT results for the D0 charge. More precisely, it seems that what the boundary state calculates is the flux of $`\omega `$ through the D-brane, without the $`H`$ bulk contribution. According to a recent analysis , both these quantities can be given a natural physical interpretation: the flux (18) of the two-form field $`\omega `$ (gauge invariant but not quantised) can be understood as the brane source charge, whereas $`Q_0`$ can be thought of as the Page charge<sup>3</sup><sup>3</sup>3I thank D Marolf for this observation., which is gauge invariant and quantised. Using this analogy one can understand better the relation between these two charges. Indeed, the brane source charge is not conserved in general and the non-conservation rule is encoded, in our case, in the relation between $`\omega `$ and $`H`$ on the D-brane which is given by (9). Moreover, this very relation also suggests us the way to modify the brane source charge in order to obtain a conserved quantity, which is nothing but our $`Q_0`$. This procedure is reminiscent of the way one constructs the Page charge in supergravity (see, e.g., ). Is is quite remarkable that the D0 charge obtained in this way turns out to agree with the one obtained by imposing the well-definedness of the boundary WZW path integral, and it is this agreement that renders it quantised.
One problematic aspect of this approach is the fact that the quantisation argument seems to impose a non-local condition on the spacetime field $`\psi `$. Notice however that the evaluation of the D0 charge was made for a specific class of D-branes, described by a very special type of gluing conditions; it is possible that this non-local condition will get smeared once we allow for more general D-brane configurations.
Finally, one of the most important and intriguing conclusions of this analysis is that there appears to be no evidence for the existence of a quantised $`\mathrm{U}(1)`$ gauge field flux on this particular class of D-branes. By this we do not mean that we have a $`\mathrm{U}(1)`$ gauge field whose flux is not quantised. Rather, we have shown that one can define and analyse the boundary WZW model, both at the classical and at the quantum level, without ever having to introduce a $`\mathrm{U}(1)`$ gauge field on the brane. Using this approach we obtained a mass spectrum for the allowed D-brane configurations which agrees with the one obtained in the boundary state formalism, we recovered the D0 charge of and its spectrum, and we were able to define a new D0 charge which is manifestly gauge invariant and quantised and modular, with the periodicity given by the level of the WZW model. These results suggest that the role of the $`\mathrm{U}(1)`$ gauge field $`F`$ in the case of a background with $`B=0`$ is taken by over by the globally defined gauge invariant two-form field $`\omega `$ in the case of a background characterised by a nontrivial $`B`$ field.
## Acknowledgements
I would like to thank C Bachas, M Douglas, K Gawȩdzki, M Kreuzer, D Marolf and C Schweigert for useful remarks and criticism, A Recknagel, V Schomerus and S Theisen for their hospitality at AEI Potsdam and ESI Vienna, where part of this work was done, and especially JM Figueroa-O’Farrill and A Tseytlin for many insightful discussions and correspondence, and for helping me fix a stubborn factor of $`2`$ or two.
|
warning/0006/gr-qc0006088.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
A surface theoretic view of non-perturbative quantum gravity was studied some time ago based on 3 dimensional general relativity as a reformulation of the Ponzano-Regge model in Riemannian spacetime . Since 3 dimensional general relativity has no local degrees of freedom, the Ponzano-Regge model is known to be a topological field theory called lattice BF theory in 3 dimensions. A 4 dimensional generalization was made by Ooguri and is known as lattice BF theory in 4 dimensions, also a topological field theory . This Ooguri model also allows the surface theoretic view. However, since 4 dimensional general relativity contains local degrees of freedom, the Ooguri model is not a quantization of general relativity.
In order to pursue 4 dimensional generalizations of the surface theoretic view of the Ponzano-Regge model with local degrees of freedom, some investigations have been being made. They can be divided roughly to four strategies. One is the use of canonical theory called loop quantum gravity based on Ashtekar formulation of general relativity . It uses a Hamiltonian of the canonical theory to compute quantum amplitudes. In this strategy, the definition of “correct” Hamiltonian is still uncertain although the canonical theory is able to provide physical interpretations. Another is the use of the Plebanski action of general relativity in Reimannian spacetime. This action has the form of BF theory with additional constraint terms . It computes a path integral of the action. In this strategy, a well regulated computation of the constraint terms is still missing and, in addition, the action functional is available only in Reimannian spacetime. Another is the use of “quantized” lattice BF theory and additional “quantum level” constraint conditions. In this strategy, the formulation is strictly mathematical and it necessitates help from other theories for physical interpretations. The other is based on physical reasoning such as causality and statistical criticality. This strategy wants a systematic formulation.
In the present work, a fifth strategy is applied to construct a surface theoretic model for 4 dimensional quantum gravity. We introduce a version of the action of general relativity in continuum. This action is derived from the action proposed some time ago by Samuel and by Jacobson and Smolin . The Ashtekar formulation can be constructed from the latter. We define a lattice version of this action and its path integral. The variables integrated are closely related to the SU(2) (real valued) variables of the Ashtekar formulation. However, the model avoids the use of Hamiltonian of canonical theory, which is a main difficulty so far for quantization. The spacetime the model assumes has signature $`(+1,+1,+1,+1)`$, Riemannian spacetime. In order to make the lattice action finite per lattice point, a dimensionless “(inverse) coupling” constant is defined in analogy to Wilson’s formulation of lattice gauge theory . The finiteness of the action with finite lattice allows the path integral without gauge fixing finite. The path integral of the model is computed as a character expansion and shown to be written as a sum over surface-like excitations in spacetime.
We show that a 3 dimensional version of the model exists and can be reduced to lattice BF theory. In this sense, the model is a 4 dimensional generalization of the Ponzano-Regge model with local degrees of freedom. The expectation values of four quantities are computed in 3 dimensions. Two of them are basic variables of the theory and are SU(2) gauge dependent quantities. Another is the Wilson loop. It is SU(2) gauge independent but diffeomorphism dependent. The other is the action itself. It is invariant under not only SU(2) gauge transformations but also diffeomorphisms. The first three are shown to be always zero unless an appropriate gauge is fixed. The other is zero if the (inverse) coupling constant is brought to an appropriate limit. These are very simple results but important checks for consistency. We discuss the meaning of these results. Note that this kind of computations have not been performed in the other strategies while the computation of the path integral (or sometimes called transition amplitude, projector or partition function) is always focussed on in all the strategies.
The model is defined on a fixed hyper cubic lattice for simplicity. However, the construction can be generalized to a randomly triangulated lattice with small modifications. Further more, the sum over different triangulations can be done, if necessary, by applying a technology developed in a recent work .
This paper is organized as follows. In Sec. 2, we restructure the action functional of general relativity so that it can be utilized to define the model. In Sec. 3, the model is defined and its path integral is studied. In Sec. 4, three dimensional version of the model is studied. In Sec. 5, we conclude the work. In Appendix A, mathematical formulae utilized in this work are listed.
## 2 The action of general relativity
The action we make use of is the one proposed by Samuel and by Jacobson and Smolin . It is the Palatini action but an additional term. The additional term is a topological term and not responsible to the local degrees of freedom of general relativity. From the Palatini action, one can construct a canonical theory by splitting space and time. However, it contains constraints of the second class in addition to those of the first class in the sense of Dirac. Therefore, the constrained phase space is not well defined unless the second class constraints are explicitly solved.
On the other hand, from the present action, one can construct Ashtekar’s canonical theory, which contains only constrains of the first class. It means that the constrained phase space is well defined in the sense that it is invariant under transformations produced by Hamiltonian and diffeomorphism generators. The difference between the two canonical theories is due to the canonical transformation produced by the topological term in the Samuel-Jacobson-Smolin action. We restructure this action for the use in the following section.
Let us introduce variables we need. We use tetrad field $`e_\mu ^I`$ rather than metric $`g_{\mu \nu }`$. They are related with each other by $`g_{\mu \nu }=e_\mu ^Ie_\nu ^J\eta _{IJ}`$. Here $`\eta _{IJ}`$ is the Euclidean or Minkowskian metric of “internal” spacetime depending on the value of $`\sigma `$ with signature $`(\sigma ,+1,+1,+1)`$. The capital alphabets $`I,J\mathrm{}`$ are used for internal spacetime indices $`\{0,1,2,3\}`$ and the Greek letters are for spacetime indices $`\{0,1,2,3\}`$. Below, we also use the lower case alphabets $`i,j\mathrm{}`$ for the space componets of the internal spacetime indices $`\{1,2,3\}`$ and for the adjoint indices of SU(2) group.
The so-called spin-connection $`\omega _\mu ^{IJ}`$ is written in terms of the tetrad field as follows.
$`\omega _\mu ^{IJ}`$ $`:=2e^{\sigma [I}_{[\mu }e_{\sigma ]}^{J]}e^{\sigma I}e^{\nu J}e_{\mu K}_{[\sigma }e_{\nu ]}^K`$ (1)
$`=\sigma ϵ^{\sigma \alpha \beta \gamma }ϵ_{MN}^{IJ}e_\alpha ^Me_\beta ^N\left(e_\gamma ^K_{[\mu }e_{\sigma ]K}+{\displaystyle \frac{1}{2}}e_\mu ^K_\gamma e_{\sigma K}\right).`$
Here $`e^{\mu I}`$ is the inverse tetrad. The Riemann tensor $`R_{\mu \nu }^{IJ}`$ is written in terms of the spin-connection as follows.
$`R_{\mu \nu }^{IJ}:=2_{[\mu }\omega _{\nu ]}^{IJ}+2\omega _{[\mu }^{IK}\omega _{\nu ]K}^J.`$ (2)
In terms of these traditional variables, we define connection and curvature variables $`A_\mu ^{IJ}`$ and $`F_{\mu \nu }^{IJ}`$ as follows.
$`A_\mu ^{IJ}:=\omega _\mu ^{IJ}+{\displaystyle \frac{1}{2}}\gamma ϵ_{KL}^{IJ}\omega _\mu ^{KL}={\displaystyle \frac{1}{2}}\gamma ϵ_{KL}^{IJ}A_\mu ^{KL}+(1\gamma ^2\sigma )\omega _\mu ^{IJ},`$ (3)
$`F_{\mu \nu }^{IJ}:=R_{\mu \nu }^{IJ}+{\displaystyle \frac{1}{2}}\gamma ϵ_{KL}^{IJ}R_{\mu \nu }^{KL}={\displaystyle \frac{1}{2}}\gamma ϵ_{KL}^{IJ}F_{\mu \nu }^{KL}+(1\gamma ^2\sigma )R_{\mu \nu }^{IJ},`$ (4)
$`F_{\mu \nu }^{IJ}=2_{[\mu }A_{\nu ]}^{IJ}+A_{[\mu }^{IK}A_{\nu ]K}^J+(1\gamma ^2\sigma )\omega _{[\mu }^{IK}\omega _{\nu ]K}^J.`$ (5)
Here $`\gamma `$ is a parameter and chosen to be $`\gamma =1`$ to relate the model with SU(2) (real valued) Ashtekar formulation. If one chooses the value such that $`\gamma ^2\sigma =1`$, then the variables $`A_\mu ^{IJ}`$ and $`F_{\mu \nu }^{IJ}`$ are self-dual with respect to the internal spacetime indices. We keep $`\gamma `$ unspecified in the course of development within this section.
Among the components of $`A_\mu ^{IJ}`$ and $`F_{\mu \nu }^{IJ}`$, we use specifically $`A_\mu ^{0i}`$ and $`F_{\mu \nu }^{0i}`$. $`F_{\mu \nu }^{0i}`$ is written in terms of $`A_\mu ^{0i}`$ and $`\omega _\mu ^{IJ}`$ as follows.
$`F_{\mu \nu }^{0i}=2_{[\mu }A_{\nu ]}^{0i}\sigma \gamma ϵ_{jk}^{0i}A_\mu ^{0j}A_\nu ^{0k}+(1\gamma ^2\sigma )\left[2\omega _{[\mu }^{0k}\omega _{\nu ]k}^i+{\displaystyle \frac{1}{2}}\gamma ϵ_{mn}^{0k}\omega _{[\mu }^{mn}\omega _{\nu ]k}^i\right].`$ (6)
The pull-back of $`A_\mu ^{0i}`$ to foliated space of spacetime is precisely the SU(2) (real valued) connection variables of the Ashtekar canonical formulation.
Let us explicitly write the action. First, the Palatini action derived from the Hilbert-Einstein action is
$`S_0[e,\omega ]`$ $`:={\displaystyle \frac{1}{l_p^2}}{\displaystyle d^4x\sqrt{\sigma g}R}={\displaystyle \frac{1}{l_p^2}}{\displaystyle d^4xee_I^\mu e_J^\nu R_{\mu \nu }^{IJ}}`$ (7)
$`={\displaystyle \frac{1}{2l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }ϵ_{IJKL}e_\mu ^Ie_\nu ^JR_{\lambda \sigma }^{KL}}.`$
Here, $`g`$ is the determinant of metric $`g_{\mu \nu }`$, $`R`$ is the scalar curvature, $`l_p`$ is the Planck length constant, $`e`$ is the determinant of $`e_\mu ^I`$ and $`ϵ^{\mu \nu \lambda \sigma }`$ is the alternating density. By adding a topological term as mentioned above, the action we make use of is
$`S[e,A]`$ $`:={\displaystyle \frac{1}{\gamma l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }e_{\mu I}e_{\nu J}\left[\frac{1}{2}\gamma ϵ_{KL}^{IJ}R_{\lambda \sigma }^{KL}+R_{\lambda \sigma }^{IJ}\right]}`$ (8)
$`={\displaystyle \frac{1}{\gamma l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }e_{\mu I}e_{\nu J}F_{\lambda \sigma }^{IJ}}.`$
This is the Samuel-Jacobson-Smolin action. By dividing space and time indices for internal spacetime and writing in terms of $`e_\mu ^0`$, $`e_\mu ^i`$ and $`A_\mu ^{0i}`$, the action becomes
$`S[e,A]={\displaystyle \frac{1}{\gamma l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }\left[2e_{\mu 0}e_{\nu i}F_{\lambda \sigma }^{0i}+e_{\mu i}e_{\nu j}F_{\lambda \sigma }^{ij}\right]}`$
$`={\displaystyle \frac{1}{\gamma l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }\left[2e_{\mu 0}e_{\nu i}F_{\lambda \sigma }^{0i}+e_{\mu i}e_{\nu j}\left[\gamma ϵ_{0k}^{ij}F_{\lambda \sigma }^{0k}+(1\gamma ^2\sigma )R_{\lambda \sigma }^{ij}\right]\right]}`$
$`={\displaystyle \frac{1}{\gamma l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }\left[\sigma \left(2e_{[\mu }^0e_{\nu ]i}+\gamma ϵ_{ijk}^0e_\mu ^je_\nu ^k\right)F_{\lambda \sigma }^{0i}+(1\gamma ^2\sigma )e_{\mu i}e_{\nu j}R_{\lambda \sigma }^{ij}\right]}`$
$`={\displaystyle \frac{1}{\gamma l_p^2}}{\displaystyle }d^4xϵ^{\mu \nu \lambda \sigma }\{\sigma (2e_{[\mu }^0e_{\nu ]i}+\gamma ϵ_{ijk}^0e_\mu ^je_\nu ^k)(2_{[\lambda }A_{\sigma ]}^{0i}\sigma \gamma ϵ_{jk}^{0i}A_\lambda ^{0j}A_\sigma ^{0k})`$
$`+(1\gamma ^2\sigma )[\sigma (2e_{[\mu }^0e_{\nu ]i}+\gamma ϵ_{ijk}^0e_\mu ^je_\nu ^k)(2\omega _{[\lambda }^{0k}\omega _{\sigma ]k}^i+{\displaystyle \frac{1}{2}}\gamma ϵ_{mn}^{0k}\omega _{[\lambda }^{mn}\omega _{\sigma ]k}^i)`$
$`+e_{\mu i}e_{\nu j}R_{\lambda \sigma }^{ij}]\}.`$ (9)
Here, we consider $`\omega _\mu ^{IJ}`$ and $`R_{\mu \nu }^{ij}`$ as functions of $`e_\mu ^0`$ and $`e_\mu ^i`$. Note that if $`\gamma ^2\sigma =1`$, then the terms proportional to $`(1\gamma ^2\sigma )`$ go away and the action gets simplified. We will utilize the case that $`\gamma =1`$ and $`\sigma =1`$, Riemannian spacetime.
## 3 The model
### 3.1 The action
The action discussed in the previous section for Riemannian spacetime $`(\sigma =+1)`$ with $`\gamma =1`$ is
$`S_R[e,A]={\displaystyle \frac{1}{l_p^2}}{\displaystyle d^4xϵ^{\mu \nu \lambda \sigma }\left(2e_{[\mu }^0e_{\nu ]i}+ϵ_{ijk}^0e_\mu ^je_\nu ^k\right)\left(2_{[\lambda }A_{\sigma ]}^iϵ_{jk}^{0i}A_\lambda ^jA_\sigma ^k\right)}.`$ (10)
Hereafter, we write $`A_\mu ^i`$ for $`A_\mu ^{0i}`$. We note that if $`2e_{[\mu }^0e_{\nu ]}^i+ϵ_{jk}^{0i}e_\mu ^je_\nu ^k`$ is replaced by a 2-form variable denoted by $`B_{\mu \nu }^i`$ with the condition that $`ϵ^{\mu \nu \lambda \sigma }B_{\mu \nu }^iB_{\lambda \sigma }^j`$ is positionwisely proportional to $`\delta ^{ij}`$, then the resulting action is precisely the Plebanski action. This additional condition is supposed to restore $`B_{\mu \nu }^i=2e_{[\mu }^0e_{\nu ]}^i+ϵ_{jk}^{0i}e_\mu ^je_\nu ^k`$. Ways of constructing a surface theoretic model starting from the Plebanski action have been explored. The condition is usually imposed with Lagrange multipliers. However, it has turned out that the inclusion of the condition in terms of Lagrange multilpiers is still a difficult task. In the present work, we do not follow any of those ways. Instead, we define a corresponding action directly on the lattice without introducing $`B_{\mu \nu }^i`$ and utilizing Lagrange multipliers.
In order to define the model, we rewrite the action in a slightly compact fashion so that the construction of the lattice version is straightforward. An idea is the following. The tetrad variables (or their modifications) are often considered su(2) algebra valued. For example, in the Ashtekar formulation, the pull-back of the (inverse) tetrads to space have su(2) algebra indices and , in the Plebanski formulation, the 2-form variable $`B_{\mu \nu }^i`$ are su(2) algebra valued. However, we do not consider here the tetrad variables su(2) algebra valued but do consider them SU(2) group valued. If one defines $`e_\mu :=e_\mu ^0+i\sigma _ie_\mu ^i`$ with the Pauli matrices $`\sigma _i`$ $`(i=1,2,3)`$, then the complicated expression $`2e_{[\mu }^0e_{\nu ]}^i+ϵ_{jk}^{0i}e_\mu ^je_\nu ^k`$ becomes just the non-trace part of $`e_\mu ^{}e_\nu `$. If $`e_\mu `$ is normalized so that $`e_\mu ^0e_\mu ^0+e_\mu ^1e_\mu ^1+e_\mu ^2e_\mu ^2+e_\mu ^3e_\mu ^3=1`$, then it is an SU(2) group element. This small change of view dramatically simplifies the construction of the model and makes the extension to other dimensional spacetimes straightforward.
Before introducing the lattice, let us count the number of degrees of freedom of the variables. $`e_\mu ^I`$ and $`A_\mu ^i`$ have respectively 16 and 12 degrees of freedom per spacetime point. From thses variables, if one constructs canonical pairs of variables, one finds 2 physical degrees of freedom from $`e_\mu ^I`$ and 2 from $`A_\mu ^i`$. The other degrees of freedoms are for the constraints and gauge trajectries in addition to the freedom of specifying a spacetime foliation. Let us be slightly more concrete. Out of 16 degrees of freedom of $`e_\mu ^I`$, 3 fix a spacetime foliation and 4 are used for Lagrange multipliers imposing three diffeomorphism and a Hamiltonian constraints. Out of 12 degrees of freedom of $`A_\mu ^i`$, 3 are used for Lagrange multipliers imposing three Gauss gauge constraints. Hence, 9 degrees of freedom in each are left. These are exactly the number of degrees of freedom of the Ashtekar canonical theory. Then by subtracting 7 for the constraints and 7 for the gauge degrees of freedom, one finds the correct number of physical degrees of freedom.
### 3.2 Lattice
We fix a lattice. We do not take sum over different lattices. In general, the model can be defined on a randomly triangulated lattice. However, technically it is difficult, if not impossible, to compute the quantum amplitudes analytically or even numerically if the structure of lattice varies from place to place. For this reason, we use a hyper cubic lattice. First we introduce a pair of hyper cubic lattices. One is dual to the other. Denote them $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$ respectively. Then we further restrict ourselves to the case that $`\mathrm{\Delta }^{}`$ is $`\mathrm{\Delta }`$ itself. This is possible because the dual lattice of a hyper cubic lattice is a hyper cubic lattice. This restriction dramatically simplifies the construction of the model and helps computations in practice.
The lattices $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$ are a pair of 4-dimensional hyper cubic lattices dual to each other. They consist of 0-cells (vertices), 1-cells (edges), 2-cells (square faces), 3-cells (cubes) and 4-cells (hyper cubes). Every k-cell of one lattice intersects with a (4-k)-cell of the other lattice at a point inside the cells. Every cell of the lattices has one and only one intersection. The correspondence of k-cells of $`\mathrm{\Delta }`$ and (4-k)-cells of $`\mathrm{\Delta }^{}`$ is one-to-one. The corresponding k-cell of $`\mathrm{\Delta }`$ and (4-k)-cell of $`\mathrm{\Delta }^{}`$ are called dual to each other. There is a limit at which every intersection coincides with one of the vertices so that the two lattices are identical $`\mathrm{\Delta }=\mathrm{\Delta }^{}`$.
Let us be more specific. Fix $`\mathrm{\Delta }`$ and $`\mathrm{\Delta }^{}`$. Both have a lattice spacing $`\epsilon `$. One is the displacement of the other by the distance $`\epsilon /2`$ in all the four directions. Let $`x`$ denote (the position of) a vertex of $`\mathrm{\Delta }`$. The (positions of) adjacent vertices are $`x\pm \epsilon ^\mu `$ with $`\mu =0,1,2,3`$, where $`\epsilon ^\mu `$ is the increment in the $`\mu `$-th direction with the magnitude $`\epsilon `$. Note that $`x\pm \epsilon ^0`$, for example, is a short hand notation for $`(x^0\pm \epsilon ,x^1,x^2,x^3)`$. Let $`x^{}`$ denote (the position) of a vertex of $`\mathrm{\Delta }^{}`$ at $`x+\epsilon ^0/2+\epsilon ^1/2+\epsilon ^2/2+\epsilon ^3/2`$. The face of $`\mathrm{\Delta }`$ specified by (the positions of) four vertices $`x`$, $`x+\epsilon ^\mu `$, $`x+\epsilon ^\nu `$ and $`x+\epsilon ^\mu +\epsilon ^\nu `$ is dual to the face of $`\mathrm{\Delta }^{}`$ specified by (the positions of) four vertices $`x^{}`$, $`x^{}\epsilon ^\lambda `$, $`x^{}\epsilon ^\sigma `$ and $`x^{}\epsilon ^\lambda \epsilon ^\sigma `$, where the indices are chosen such that $`ϵ^{\mu \nu \lambda \sigma }=1`$. There are six pairs of faces for given $`x`$ up to the orientation of the face. These dual faces are important in our construction of the model. After taking the limit $`\mathrm{\Delta }^{}\mathrm{\Delta }`$ such that $`x^{}`$ coincides with $`x`$, the dual faces are still clearly defined on the single lattice $`\mathrm{\Delta }`$. The intersection of a dual pair of faces, one basing at $`x`$ and the other at $`x^{}`$, coincides with $`x=x^{}`$ at the limit. Notice that the face basing at $`x^{}`$ and in the $`\mu \nu `$ plane does not coincide with the face basing at $`x`$ and in the same plane at the limit. One is in the $`\mu `$ and $`\nu `$ directions while the other is in the $`\mu `$ and $`\nu `$ directions from the coincident base point $`x=x^{}`$. We will define our model on the lattice $`\mathrm{\Delta }(=\mathrm{\Delta }^{})`$.
### 3.3 The lattice action
We define an action on the lattice $`\mathrm{\Delta }`$ so that it converges to the action in continuum as the lattice spacing goes to zero. First, let us define variables on the lattice as follows. For an edge basing at $`x`$ and pointing at the $`\mu `$-th direction, define
$`\zeta _\mu (x):=\zeta _\mu ^0(x)+i\sigma _i\zeta _\mu ^i(x):=\mathrm{exp}[i\epsilon A_\mu ^i(x)\sigma _i/2],`$ (11)
$`\eta _\mu (x):=\eta _\mu ^0(x)+i\sigma _i\eta _\mu ^i(x):=(\beta ^{1/2}l_p)^1\epsilon [e_\mu ^0(x)+ie_\mu ^i(x)\sigma _i]/\rho _\mu (x),`$ (12)
$`\rho _\mu (x):=(\beta ^{1/2}l_p)^1\epsilon \sqrt{|e_\mu ^0(x)|^2+|e_\mu ^1(x)|^2+|e_\mu ^2(x)|^2+|e_\mu ^3(x)|^2}.`$ (13)
Here, $`\epsilon `$ is the lattice spacing and $`\sigma _i`$’s are the Pauli matrices, that is,
$`\sigma _1:=\left[\begin{array}{cc}0& 1\\ 1& 0\end{array}\right],\sigma _2:=\left[\begin{array}{cc}0& i\\ i& 0\end{array}\right],\sigma _3:=\left[\begin{array}{cc}1& 0\\ 0& 1\end{array}\right].`$ (14)
Note that the inside the square root in $`\rho _\mu `$ is equal to the metric element $`g_{\mu \mu }`$. $`\beta `$ is the upper bound of $`l_p^2\epsilon ^2g_{\mu \mu }(\mu =0,1,2,3)`$ and introduced to regularize the path integral defined below. $`\rho _\mu `$ is non-negative and less than $`1`$, and $`\zeta _\mu `$ and $`\eta _\mu `$ are SU(2) matrices normalized such that $`|\zeta _\mu ^0|^2+|\zeta _\mu ^1|^2+|\zeta _\mu ^2|^2+|\zeta _\mu ^3|^2=1`$ and $`|\eta _\mu ^0|^2+|\eta _\mu ^1|^2+|\eta _\mu ^2|^2+|\eta _\mu ^3|^2=1`$. For a face basing at $`x`$ and enclosed by four edges in the $`\mu `$ and $`\nu `$-th direstions, define
$`U_{\mu \nu }(x):=\zeta _\mu (x)\zeta _\nu (x+\epsilon ^\mu )\zeta _\mu ^{}(x+\epsilon ^\nu )\zeta _\nu ^{}(x),`$ (15)
without summing over $`\mu `$ and $`\nu `$. Here, $`\epsilon ^\mu `$ is the increment in the $`\mu `$-th direction with the magnitude $`\epsilon `$.
The action on the lattice $`\mathrm{\Delta }`$ is then defined as follows.
$`S_\mathrm{\Delta }[\zeta ,\eta ,\rho ]:=\beta {\displaystyle \underset{x\mathrm{\Delta }}{}}ϵ^{\mu \nu \lambda \sigma }\rho _\mu (x)\rho _\nu (x)\mathrm{Tr}[\eta _\mu ^{}(x)\eta _\nu (x)U_{\lambda \sigma }(x)].`$ (16)
Here, the sums over $`\mu `$, $`\nu `$, $`\lambda `$ and $`\sigma `$ have been performed. This lattice action is finite on the lattice $`\mathrm{\Delta }`$ with $`\epsilon `$ and $`\beta `$ fixed if the lattice size is finite. This is because $`\zeta _\mu `$ and $`\eta _\mu `$ are compact SU(2) variables and $`\rho _\mu `$ has lower and upper bounds in addition to the fact that the number of degrees of freedom is finite on the lattice if the lattice size is finite. The finiteness of the action lets the path integral without gauge fixing non-divergent. In other words, the path integral along a gauge trajectory produces just a finite overall multiplicative constant. We call $`\beta `$ “inverse coupling” in analogy to Wilson’s formulation of lattice gauge theory.
Let us examine if the lattice action really converges to the continuum action as the lattice spacing $`\epsilon `$ goes to zero. Take the limit $`\epsilon 0`$. In the limiting process, it can be shown that
$`\beta \rho _\mu (x)\rho _\nu (x)\eta _{[\mu }^{}(x)\eta _{\nu ]}(x)=l_p^2\epsilon ^2i\sigma _i\left(2e_{[\mu }^0(x)e_{\nu ]}^i(x)+ϵ_{jk}^{0i}e_\mu ^j(x)e_\nu ^k(x)\right),`$ (17)
$`U_{[\lambda \sigma ]}(x){\displaystyle \frac{1}{2}}\epsilon ^2i\sigma _i\left(2_{[\lambda }A_{\sigma ]}^i(x)ϵ_{jk}^{0i}A_\lambda ^j(x)A_\sigma ^k(x)\right)+𝒪(\epsilon ^3).`$ (18)
Therefore, the lattice action converges to
$`S_\mathrm{\Delta }[\zeta ,\eta ,\rho ]`$ $`l_p^2{\displaystyle \underset{x\mathrm{\Delta }}{}}\epsilon ^4ϵ^{\mu \nu \lambda \sigma }(2e_\mu ^0(x)e_{\nu i}(x)+ϵ_{ijk}^0e_\mu ^j(x)e_\nu ^k(x))\times `$ (19)
$`\left(2_\lambda A_\sigma ^i(x)ϵ_{jk}^{0i}A_\lambda ^j(x)A_\sigma ^k(x)\right)+𝒪(\epsilon ^5).`$
### 3.4 The lattice path integral
The exponential of the action can be considered as (an extension of) the graph-cylindrical function discussed in . The integral measure we use for the variable $`A_\mu ^i`$ is the Ashtekar-Lewandowski measure . The measure for SU(2)-compactified part of $`e_\mu ^0`$ and $`e_\mu ^i`$ is analogously defined and that for the other part of them is the one discussed in . The path integral has the form of
$`{\displaystyle }d\mu (A)d\mu (e)\mathrm{\Psi }_{\mathrm{\Delta },\psi }(A,e):={\displaystyle }{\displaystyle \underset{x\mathrm{\Delta }}{}}{\displaystyle \underset{\mu }{}}d\zeta _\mu (x)d\eta _\mu (x)d\rho _\mu ^4(x)\times `$
$`\psi (\zeta (x_1,A),\mathrm{}\zeta (x_l,A);\eta (x_1,e),\mathrm{}\eta (x_m,e);\rho (x_1,e),\mathrm{}\rho (x_n,e)),`$ (20)
where $`\mathrm{\Psi }_{\mathrm{\Delta },\psi }`$ is a cylindrical function defined on the lattice $`\mathrm{\Delta }`$ in terms of a complex valued integrable function $`\psi `$ on $`[SU(2)]^l\times [SU(2)]^m\times [0,1]^n`$. $`d\zeta _\mu `$ and $`d\eta _\mu `$ are the Haar measures for SU(2) elements $`\zeta _\mu `$ and $`\eta _\mu `$ respectively. $`d\rho _\mu ^4=4d\rho _\mu \rho _\mu ^3`$ is the radial integral of the 4-dimensional unit sphere. $`\rho _\mu `$ is integrated from $`0`$ to $`1`$.
The path integral we will compute in the following section is defined as follows.
$`Z_\mathrm{\Delta }:={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4e^{iS_\mathrm{\Delta }[\zeta ,\eta ,\rho ]}}`$
$`={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4\underset{x\mathrm{\Delta }}{}e^{i\beta ϵ^{\mu \nu \lambda \sigma }\rho _\mu (x)\rho _\nu (x)\mathrm{Tr}[\eta _\mu ^{}(x)\eta _\nu (x)U_{\lambda \sigma }(x)]}}.`$ (21)
Note that we integrate the exponential oscillation form $`e^{iS}`$ rather than the exponential decay form $`e^S`$. The latter is commonly and successfully used for Euclidean quantum field theories whose action has quadratic form. However, the action of general relativity is not quadratic but linear in each of the variables. Because of this fact, the action is not bound from the below and the use of the exponential decay form runs into serious technical problems. In addition, it unlikely contains the classical limit since the value of the action for classical solutions is zero. The former is used in other strategies in quantum gravity even for Riemannian spacetime. In the exponential oscillation form, large magnitudes of the action less contribute to the path integral. In order to examine possible consequences of the use of the exponential decay form, we simply replace $`\beta `$ by $`i\beta `$ in the results of the use of the exponential oscillation form.
In terms of the path integral, define the expectation value of an observable $`X`$ as follows.
$`X_\mathrm{\Delta }:=Z_\mathrm{\Delta }^1{\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4X[\zeta ,\eta ,\rho ]e^{iS_\mathrm{\Delta }[\zeta ,\eta ,\rho ]}}.`$ (22)
In particular, we are interested in four quantities: $`\eta _\mu `$, $`\zeta _\mu `$, $`\mathrm{Tr}U_{\mu \nu }`$ and $`S_\mathrm{\Delta }`$. The first two are basic variables of the theory and are SU(2) dependent quantities. They should vanish. This is because there is a symmetric structure along the SU(2) gauge trajectories. It is known that the expactation value of the basic variables of lattice QCD vanishes unless an appropriate gauge is fixed. This fact is often utilized to study gauge fixing ambiguities on lattice. The same thing should happen here since the Gauss gauge constraint structure of the present theory is common with the SU(2) version of QCD.
The third is (a smallest of) the Wilson loop. It is SU(2) gauge independent and a good physical observable for Gauss gauge invariant theories such as lattice QCD. Here, it is possible that the expectation value of the Wilson loop vanishs unless an appropriate gauge is fixed. This would mean that the Gauss gauge invariance is not enough to be a physical observable for quantum gravity and a possible symmetric structure along diffeomorphism gauge trajectories annihilates the expectation value of the Wilson loop. Quantum gravity is invariant not only under Gauss gauge transformations but also under spacetime diffeomorphisms. The Wilson loops are not diffeomorphism invariant and hence cannot be physical observables of quantum gravity.
The last one is the action of the theory. It is invariant under SU(2) gauge transformations and diffeomorphisms and the only known local physical observable in general relativity. It should vanish if the inverse coupling constant is brought to infinity but diverge in the same limit if the path integral has the exponential decay form. The former would be an indication that the path integral successfully eliminates the degrees of freedom constrained since the value of the action on the constraint surface must be identically zero. The latter would be an indication that the path integral of exponential decay form cannot capture the classical solutions. These are non-trivial checks for consistency. These facts are examined in 3 dimensions.
### 3.5 Character expansion
We compute the path integral as a character expansion and show that it can be written as a sum over surface-like excitations in spacetime. First we rewrite the path integral as follows.
$`Z_\mathrm{\Delta }:={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4e^{iS_\mathrm{\Delta }[\zeta ,\eta ,\rho ]}}`$
$`={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4\underset{x\mathrm{\Delta }}{}\underset{(\mu \nu \lambda \sigma )}{}e^{i2\beta \rho _\mu (x)\rho _\nu (x)\mathrm{Tr}[\eta _\mu ^{}(x)\eta _\nu (x)U_{\lambda \sigma }(x)]}e^{i2\beta \rho _\mu (x)\rho _\nu (x)\mathrm{Tr}[\eta _\nu ^{}(x)\eta _\mu (x)U_{\lambda \sigma }(x)]}}.`$
Here, $`(\mu \nu \lambda \sigma )`$ denotes the even permutations of $`(0123)`$; there are six terms, that is, $`(0123)`$, $`(0231)`$, $`(0312)`$, $`(1203)`$, $`(3102)`$, and $`(2301)`$. Then use the formula (36) to expand to the characters as follows.
$`Z_\mathrm{\Delta }={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4\underset{x\mathrm{\Delta }}{}\underset{(\mu \nu \lambda \sigma )}{}\underset{l_{\mu \nu },l_{\nu \mu }}{}\mathrm{\Gamma }_{\mu \nu }(\beta ;\{l,\rho \})\chi _{l_{\mu \nu }}(\eta _\mu ^{}\eta _\nu U_{\lambda \sigma })\chi _{l_{\nu \mu }}(\eta _\nu ^{}\eta _\mu U_{\lambda \sigma })},`$ (24)
with
$`\mathrm{\Gamma }_{\mu \nu }(\beta ;\{l,\rho \}):={\displaystyle \frac{(2l_{\mu \nu }+1)(2l_{\nu \mu }+1)(1)^{l_{\mu \nu }l_{\nu \mu }}}{4\beta ^2\rho _\mu ^2\rho _\nu ^2}}J_{2l_{\mu \nu }+1}(4\beta \rho _\mu \rho _\nu )J_{2l_{\nu \mu }+1}(4\beta \rho _\mu \rho _\nu ).`$
Here, $`\chi _j(U)`$ is the character of SU(2) element $`U`$ in the spin-j representation and $`l_{\mu \nu }`$ and $`l_{\nu \mu }`$ are spins taking values 0, $`\frac{1}{2}`$, 1, $`\frac{3}{2}\mathrm{}`$. $`J_m(x)`$ is the Bessel function of the first kind. One of its definitions is given by (37).
From (24), one can understand how surface-like excitations emerge. $`U_{\lambda \sigma }(x)`$ is a parallel transport along the four edges enclosing a face of the lattice. The face is in the $`\lambda \sigma `$-plane and bases at $`x`$. Let $`f_{\lambda \sigma }(x)`$ denote this face. (We do not distinguish the orientations; $`f_{\lambda \sigma }(x)f_{\sigma \lambda }(x)`$.) Associate the spin $`l_{\mu \nu }`$ of $`\chi _{l_{\mu \nu }}(\eta _\mu ^{}(x)\eta _\nu (x)U_{\lambda \sigma }(x))`$ to the face $`f_{\lambda \sigma }(x)`$. In the same way, $`l_{\nu \mu }`$ is also associated to the same face. By integrating $`\eta _\mu `$ at $`x`$, one introduces an intertwiner to combine six spins $`l_{\mu \nu }`$ and $`l_{\nu \mu }`$ ($`\nu \mu `$ for given $`\mu `$) associated with $`f_{\lambda \sigma }(x)`$ ($`\lambda `$ and $`\sigma `$ are such that $`ϵ^{\mu \nu \lambda \sigma }=1`$). By integrating $`\zeta _\mu `$ shared by six faces $`f_{\mu \nu }(x)`$ and $`f_{\mu \nu }(x\epsilon ^\nu )`$ ($`\nu \mu `$ for given $`\mu `$), one introduces an intertwiner to combine the twelve spins associated with the six faces. Therefore, after integrating all $`\eta `$’s and $`\zeta `$’s, the spins and the intertwiners combining the spins are left. Two spins are associated with a single face and two intertwiners are associated with a single edge. These facts are more clearly understood in the 3 dimensional version of the model as examined in Sec. 4. The path integral is now written as a sum over spins and intertwiners associated with faces and edges respectively. A set of faces jointed by edges represent a surface-like object, mathematically called 2-dimensional piece-wise linear cell complex . A surface-like excitation is a 2-dimensional piece-wise linear cell complex with faces labeled by non-zero spins and with edges labeled by intertwiners. A face with spin-0 means the absence of the face in a surface-like excitation.
The integrations of $`\eta `$’s and $`\zeta `$’s can be explicitly done by using the formulae (38), (39) and (40). The resulting coefficiens contain the spins associated with faces and other spins representing intertwiners associated with edges. The coefficients are expressed in terms of the so-called 3j-coefficients (41). The integrations of $`\rho `$’s are difficult because of the presence of two $`\rho `$’s in the argument of the Bessel functions. This difficulty is absent in the 3 dimensional version of the model.
## 4 The model in 3-dimensions
### 4.1 The action and path integral
From the action in 4 dimensions, it is straightforward to write its 3 dimensional version. We simply replace $`l_p^2d^4xϵ^{\sigma \lambda \mu \nu }e_\sigma ^{}e_\lambda `$ in the 4 dimensional action by $`l_p^1d^3xϵ^{\lambda \mu \nu }e_\lambda `$. We compute the corresponding path integral on the lattice and show that it is reduced to lattice BF theory in 3 dimensions in order to claim that 3 dimensional version of the model exists and is a correct quantum gravity model.
The action of general relativity in 3 dimensional Riemannian spacetime is defined as follows.
$`S_{R_3}:={\displaystyle \frac{1}{l_p}}{\displaystyle d^3xϵ^{\lambda \mu \nu }e_{\lambda i}\left(2_{[\mu }A_{\nu ]}^iϵ_{jk}^iA_\mu ^jA_\nu ^k\right)}.`$ (26)
Here, the Greek letter indices are $`\{0,1,2\}`$ for spacetime and lower alphabet indices are $`\{1,2,3\}`$ for internal spacetime.
Let us count the number of degrees of freedom. Out of 9 degrees of freedom of $`e_\mu ^i`$ at each spacetime point, 3 are for Lagrange multipliers imposing two diffeomorphism and a Hamiltonian constraints. Out of 9 degrees of freedom of $`A_\mu ^i`$ at each spacetime point, 3 are for Lagrange multipliers imposing 3 Gauss constraints. Hence, 6 degrees of freedom in each are left. They are cancelled by the 6 constraints and 6 gauge degrees of freedom and one finds no local degrees of freedom consistently with general relativity. Note that in canonical theory one sees the vectors specifying a spacetime foliation multiplicative $`e_\mu ^i`$. Therefore, the integral over all foliations is taken into account by the integrals of $`e_\mu ^i`$.
The lattice action is defined as follows.
$`S_{\mathrm{\Delta }_3}[\zeta ,\eta ,\rho ]:=\beta {\displaystyle \underset{x\mathrm{\Delta }_3}{}}ϵ^{\lambda \mu \nu }\rho _\lambda (x)\mathrm{Tr}[\eta _\lambda (x)U_{\mu \nu }(x)].`$ (27)
Here, $`\mathrm{\Delta }_3`$ denotes the 3 dimensional cubic lattice. Note that the definitions of the variables are the same as for the 4 dimensional model, (11),(12) and (13), and the additional degrees of freedom $`\eta _\mu ^0`$ cancel in the lattice action and do not play any role. Then the lattice path integral is defined as follows.
$`Z_{\mathrm{\Delta }_3}:={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4\underset{x\mathrm{\Delta }_3}{}e^{i\beta ϵ^{\lambda \mu \nu }\rho _\lambda (x)\mathrm{Tr}[\eta _\lambda (x)U_{\mu \nu }(x)]}}`$
$`={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4\underset{x\mathrm{\Delta }_3}{}\underset{\genfrac{}{}{0pt}{}{(\lambda \mu \nu )=(012),}{(120),(201)}}{}e^{i\beta \rho _\lambda \mathrm{Tr}[\eta _\lambda U_{\mu \nu }]}e^{i\beta \rho _\lambda \mathrm{Tr}[\eta _\lambda U_{\nu \mu }]}}.`$ (28)
Then use the formula (36) to expand to the characters as follows.
$`Z_{\mathrm{\Delta }_3}={\displaystyle 𝑑\zeta 𝑑\eta 𝑑\rho ^4\underset{x\mathrm{\Delta }_3}{}\underset{\genfrac{}{}{0pt}{}{(\lambda \mu \nu )=(012),}{(120),(201)}}{}\underset{l_{\mu \nu },l_{\nu \mu }}{}\mathrm{\Gamma }_\lambda (\beta ,\{l,\rho \})\chi _{l_{\mu \nu }}(\eta _\lambda U_{\mu \nu })\chi _{l_{\nu \mu }}(\eta _\lambda ^{}U_{\mu \nu })},`$ (29)
with
$`\mathrm{\Gamma }_\lambda (\beta ,\{l,\rho \}):={\displaystyle \frac{(2l_{\mu \nu }+1)(2l_{\nu \mu }+1)(1)^{l_{\mu \nu }l_{\nu \mu }}}{\beta ^2\rho _\lambda ^2}}J_{2l_{\mu \nu }+1}(2\beta \rho _\lambda )J_{2l_{\nu \mu }+1}(2\beta \rho _\lambda ).`$ (30)
Here, one can understand the emergence of surface-like excitations, in other words, 2 dimensional piece-wise linear cell complexs with faces labeled by non-zero spins and with edges labeled by intertwiners. $`U_{\mu \nu }`$ is a parallel transport along the four edges enclosing a face of the lattice. The face is in the $`\mu \nu `$-plane and bases at $`x`$. Let $`f_{\mu \nu }(x)`$ denote this face. Associate the spin $`l_{\mu \nu }`$ of $`\chi _{l_{\mu \nu }}(\eta _\lambda U_{\mu \nu })`$ to the face $`f_{\mu \nu }(x)`$. In the same way, $`l_{\nu \mu }`$ is also associated to the same face. By integrating $`\eta _\lambda `$ at $`x`$, one introduces an intertwiner to combine two spins $`l_{\mu \nu }`$ and $`l_{\nu \mu }`$ associated to $`f_{\mu \nu }(x)`$ ($`\mu `$ and $`\nu `$ are such that $`ϵ^{\lambda \mu \nu }=1`$). By integrating $`\zeta _\mu `$ shared by four faces $`f_{\mu \nu }(x)`$ and $`f_{\mu \nu }(x\epsilon ^\nu )`$ ($`\nu \mu `$ for given $`\mu `$), one introduces an intertwiner to combine the eight spins associated with the four faces. Therefore, after integrating all $`\eta `$’s and $`\zeta `$’s, the spins and the intertwiners combining the spins are left. The path integral is now written as a sum over spins and intertwiners representing surface-like excitations.
The integrations of $`\eta `$’s, $`\zeta `$’s and $`\rho `$’s can be explicitly done by using the formulae (38), (39) and (40). Integrate $`\eta `$’s and $`\rho `$’s as follows.
$`Z_{\mathrm{\Delta }_3}={\displaystyle 𝑑\zeta \underset{x\mathrm{\Delta }_3}{}\underset{\genfrac{}{}{0pt}{}{(\mu \nu )=(12),}{(20),(01)}}{}\underset{l_{\mu \nu }}{}\mathrm{\Delta }_{\mu \nu }(\beta )(2l_{\mu \nu }+1)\chi _{l_{\mu \nu }}(U_{\mu \nu }U_{\mu \nu })},`$ (31)
with
$`\mathrm{\Delta }_{\mu \nu }(\beta ):={\displaystyle _0^1}𝑑\rho _\lambda ^4\left[{\displaystyle \frac{J_{2l_{\mu \nu }+1}(2\beta \rho _\lambda )}{\beta \rho _\lambda }}\right]^2=\beta ^4{\displaystyle _0^{2\beta }}𝑑kkJ_{2l_{\mu \nu }+1}(k)J_{2l_{\mu \nu }+1}(k)`$
$`\beta ^4\delta (0)\mathrm{as}\beta \mathrm{}.`$ (32)
In the limit $`\beta \mathrm{}`$, $`\mathrm{\Delta }_{\mu \nu }`$ loses the dependence on $`l_{\mu \nu }`$ and the path integral can be rewritten as follows.
$`\underset{\beta \mathrm{}}{lim}Z_{\mathrm{\Delta }_3}={\displaystyle 𝑑\zeta \underset{x\mathrm{\Delta }_3}{}\underset{\genfrac{}{}{0pt}{}{(\mu \nu )=(12),}{(20),(01)}}{}\underset{l_{\mu \nu }}{}(2l_{\mu \nu }+1)[\chi _{l_{\mu \nu }}(U_{\mu \nu })+\chi _{l_{\mu \nu }}(U_{\mu \nu })]},`$ (33)
up to an overall multiplicative constant. The first half is well known form for 3-dimensional lattice BF theory and is meant that spacetime is flat $`U_{\mu \nu }=1`$. In addition, we have another term also meaning that spacetime is flat but $`U_{\mu \nu }=1`$. The presence of the additional term is due to the definition of the lattice action. Because of the definition of the lattice action, the path integral contains a form analogous to $`𝑑xe^{ix\mathrm{sin}\theta }=\delta (\theta )+\delta (\theta \pi )`$ instead of the well known form analogous to $`𝑑xe^{ix\theta }=\delta (\theta )`$. We have understood that the path integral of the model in 3 dimensions can be reduced to (a generalization of) that of 3 dimensional lattice BF theory.
### 4.2 Expectation values
We compute the expectation values of four quantities, $`\eta _\mu `$, $`\zeta _\mu `$, $`\mathrm{Tr}U_{\mu \nu }`$ and $`S_{\mathrm{\Delta }_3}`$, and find all of them vanish. However, their meanings are different as discussed below. Remember that the path integral contains the integrations of 18 degrees of freedom per lattice point (9 for $`\zeta _\mu `$ and 9 for $`\eta _\mu `$ and $`\rho _\mu `$). If one translates them into the canonical theory, 3 of $`\zeta _\mu `$ and 3 of $`\eta _\mu `$ and $`\rho _\mu `$ are Lagrange multipliers imposing the 6 constraints per lattice point. The other degrees of freedom are 6 degrees of freedom constrained and 6 on the constraint surface. The degrees of freedom constrained do not contribute to the path integral. When we compute the expectation values, the integrations over the degrees of freedom on the constraint surface make non trivial contributions. All of the local degrees of freedom on the constraint surface are gauge degrees of freedom in the 3 dimensional case.
$`\eta _\mu `$ and $`\zeta _\mu `$ are basic variables of the theory. They are SU(2) gauge dependent quantities. It is known that the expectation value of basic variable of lattice QCD vanishes unless an appropriate gauge is fixed because there is a symmetric structure along the SU(2) gauge trajectories. The same Gauss gauge constraint structure is also present in general relativity in the present formulation. Therefore, $`\eta _\mu `$ and $`\zeta _\mu `$ should vanish if the model is consistent.
Let us compute the expectation values of $`\eta _\mu `$ and $`\zeta _\mu `$. Insert $`\eta _0(x)`$ into (29) and integrate $`\eta _0(x)`$. This results in $`l_{21}=l_{12}\pm \frac{1}{2}`$ associated with $`f_{12}(x)`$. On the other hand, the integrations of $`\eta _2(x)`$, $`\eta _0(x\epsilon ^2)`$ and $`\eta _2(x\epsilon ^0)`$ result in $`l_{10}=l_{01}`$ associated with $`f_{01}(x)`$, $`l_{21}=l_{12}`$ associated with $`f_{12}(x\epsilon ^2)`$ and $`l_{10}=l_{01}`$ associated with $`f_{01}(x\epsilon ^0)`$ respectively. $`f_{12}(x)`$, $`f_{01}(x)`$, $`f_{12}(x\epsilon ^2)`$ and $`f_{01}(x\epsilon ^0)`$ share the edge where $`\zeta _1(x)`$ is defined. Hence, the integration of $`\zeta _1(x)`$ introduces an intertwiner combining the spins associated with the four faces. However, no intertwiner consistent with these spins exists because the sum of the 8 spins is a half integer. Therefore, we conclude $`\eta _0=0`$. In the same way, we find $`\eta _1=\eta _2=0`$.
In order to compute $`\zeta _0`$, insert $`\zeta _0(x)`$ into (31) and integrate $`\zeta _0(x)`$. The edge where $`\zeta _0(x)`$ is defined is shared by the four faces $`f_{01}(x)`$, $`f_{02}(x)`$, $`f_{01}(x\epsilon ^1)`$ and $`f_{02}(x\epsilon ^2)`$. Since each of the four faces is associated with two equal spins after the integrations of $`\eta `$’s, the sum of the 8 spins associated with the four faces is an integer. There is no intertwiner combining an integer and $`\frac{1}{2}`$ to produce the spin-0. Therefore, we conclude $`\zeta _0=0`$. In the same way, we find $`\zeta _1=\zeta _2=0`$.
$`\mathrm{Tr}U_{\mu \nu }`$ is the expectation value of the Wilson loop and is an SU(2) gauge independent quantity. It is a good physical observable for Gauss gauge invariant theories such as lattice QCD. However, the Gauss gauge invariance is not enough to be a physical observable for quantum gravity. Physical observables in quantum gravity must be invariant not only under SU(2) gauge transformations but also under spacetime diffeomorphisms. The Wilson loop is not invariant under diffeomorphisms. Therefore, it is possible that $`\mathrm{Tr}U_{\mu \nu }`$ vanishes because of a symmetric structure along diffeomorphism gauge trajectories unless an appropriate gauge is fixed. If it is the case, the vanishing result can be understood as a consequence of the fact that quantum gravity is a spacetime diffeomorphism invariant theory.
The reasoning for $`\zeta _\mu `$ to vanish also holds for $`\mathrm{Tr}U_{\mu \nu }`$ to vanish. Since every face is associated with two equal spins after the integrations of $`\eta `$’s, there is no way of combining the 8 spins associated with the four faces and spin $`\frac{1}{2}`$ of $`\mathrm{Tr}U_{\mu \nu }`$ to produce the spin-0. Therefore, we find $`\mathrm{Tr}U_{\mu \nu }=0`$. The fact that every face is associated with two spins is a notable difference of the model from lattice QCD. One of the reasons for this fact is the presence of the variable $`e_\mu ^i`$ in addition to the connection variable $`A_\mu ^i`$ in the action. Note that $`\mathrm{Tr}U_{\mu \nu }=0`$ does not contradict to the fact that the same quantity evaluated with flat connection is not zero but one since this loop can be contracted to a point. This fact is restored by the computation of the quantity with a gauge fixing corresponding to either $`U_{\mu \nu }=1`$ or $`1`$. The vanishing result is due to the symmetric structure consisting of $`U_{\mu \nu }=1`$ and $`1`$ on the constraint surface.
$`S_{\mathrm{\Delta }_3}`$ is the expectation value of the action. It is invariant under SU(2) gauge transformations and spacetime diffeomorphisms. The value of the action on the constraint surface is identically zero. Therefore, $`S_{\mathrm{\Delta }_3}`$ must vanish if the path integral successfully eliminates the contributions from the degrees of freedom constrained. This must be the case if the path integral adopts the exponential oscillation form with $`\beta \mathrm{}`$. On the other hand, if the path integral adopts the exponential decay form, then the dominant contribution comes from the value of the action far away from zero and hence $`S_{\mathrm{\Delta }_3}`$ unlikely vanishes. This fact can be checked by replacing $`\beta `$ in the path integral by $`i\beta `$.
In order to compute $`S_{\mathrm{\Delta }_3}`$, take a derivative of (31) with respect to $`\beta `$ (with $`i`$ multiplied) instead of inserting $`S_{\mathrm{\Delta }_3}`$ into (29). This procedure results in the following quantity inserted in the path integral.
$`{\displaystyle \underset{x\mathrm{\Delta }}{}}{\displaystyle \underset{\genfrac{}{}{0pt}{}{(\mu \nu )=(12),}{(20),(01)}}{}}{\displaystyle \frac{i\frac{d}{d\beta }\mathrm{\Delta }_{\mu \nu }(\beta )}{\mathrm{\Delta }_{\mu \nu }(\beta )}}`$ (34)
Since there is no particular place or direction in purely empty spacetime, the every term must have equal contrubution and we compute one of the terms as follows.
$`{\displaystyle \frac{i\frac{d}{d\beta }\mathrm{\Delta }_{\mu \nu }(\beta )}{\mathrm{\Delta }_{\mu \nu }(\beta )}}={\displaystyle \frac{4i}{\beta }}+{\displaystyle \frac{4i[J_{2l_{\mu \nu }+1}(2\beta )]^2}{\beta ^3\mathrm{\Delta }_{\mu \nu }}}{\displaystyle \frac{4i}{\beta }}\left(1{\displaystyle \frac{1}{c\pi }}\mathrm{cos}^2[2\beta {\displaystyle \frac{\pi }{2}}(2l_{\mu \nu }+1){\displaystyle \frac{\pi }{4}}]\right),`$
as $`\beta `$ goes to infinity. Here, we have used the asymptotic formula $`J_m(x)\sqrt{\frac{2}{\pi x}}\mathrm{cos}(x\frac{m\pi }{2}\frac{\pi }{4})`$ for large $`x`$ and the fact that $`\mathrm{\Delta }_{\mu \nu }`$ goes to $`\beta ^4\delta (0)`$ or more precisely $`c\beta ^3`$ as $`\beta \mathrm{}`$, where $`c`$ is a constant. We do not need the detailed value of $`c`$. This fact can be understood by rescaling $`\beta `$ in (4.1). From (4.2), we conclude $`S_{\mathrm{\Delta }_3}0`$ as $`\beta \mathrm{}`$. Note that if one replaces $`\beta `$ by $`i\beta `$ in the path integral, then one finds (4.2) with the cosine replaced by the hyperbolic cosine. In this case, $`S_{\mathrm{\Delta }_3}`$ diverges instead of converging to zero as $`\beta \mathrm{}`$.
## 5 Conclusion
In the present work, we constructed a surface-theoretic lattice quantum gravity model in 4 dimensional Riemannian spacetime based on the SU(2) Ashtekar formulation of general relativity. We introduced a version of the action of general relativity and defined its lattice version on a fixed hyper cubic lattice. We introduced a dimensionless “(inverse) coupling” constant so that the magnitude of the action is finite per lattice point. We defined a path integral whose integrand has the exponential oscillation form so that it eliminates the degrees of freedom constrained. The finiteness of the action with finite lattice allowed the path integral without gauge fixing finite. We expanded the path integral in the SU(2) characters and showed that the path integral can be written as a sum over surface-like excitations in spacetime. We showed that the 3 dimensional version of the model exists and its path integral is reduced to that of 3-dimensional lattice BF theory. Therefore, we considered the model as a 4 dimensional generalization of the Ponzano-Regge model with local degrees of freedom. We examined the expectation values of two basic variables of the theory, the Wilson loop and the action of the model in 3 dimensions and showed all of them vanish. We discussed the meaning of each of them and understood that the model has a chance of capturing the physical degrees of freedome of general relativity.
## Appendix A Useful formulae
In the present work, we have used the following mathematical formulae.
$`e^{ix\frac{1}{2}\mathrm{Tr}U}={\displaystyle \underset{j}{}}2{\displaystyle \frac{2j+1}{x}}i^{2j}J_{2j+1}(x)\chi _j(U),`$ (36)
$`J_m(x)={\displaystyle \frac{1}{2\pi }}{\displaystyle _\pi ^\pi }𝑑\theta e^{ix\mathrm{sin}\theta im\theta }.`$ (37)
The first formula can be easily proved and the second is a definition of the Bessel function of the first kind. $`\chi _j(U)`$ is the character of SU(2) element $`U`$ in the spin-j representation and $`j`$ is a spin taking values 0, $`\frac{1}{2}`$, 1, $`\frac{3}{2}\mathrm{}`$.
$`(1)^{nm}D_{nm}^{(j)}(U)=D_{n,m}^{(j)}(U),`$ (38)
$`D_{n_1m_1}^{(j_1)}(U)D_{n_2m_2}^{(j_2)}(U)={\displaystyle \underset{j,n,m}{}}(2j+1)\left(\begin{array}{ccc}j_1& j_2& j\\ n_1& n_2& n\end{array}\right)\left(\begin{array}{ccc}j_1& j_2& j\\ m_1& m_2& m\end{array}\right)D_{nm}^{(j)}(U),`$ (39)
$`{\displaystyle 𝑑UD_{mn}^{(i)}(U)D_{m^{}n^{}}^{(j)}(U)}={\displaystyle \frac{1}{2j+1}}\delta _{ij}\delta _{mm^{}}\delta _{nn^{}},`$ (40)
$`D_{mn}^{(j)}(U)`$ is the spin-j representation matrix of SU(2) element $`U`$ and $`m`$ and $`n`$ run from $`j`$ through $`j`$ with the increment 1. $`\left(\begin{array}{ccc}j_1& j_2& j\\ m_1& m_2& m\end{array}\right)`$ is the so-called 3j-coefficient and defined by
$`\left(\begin{array}{ccc}j_1& j_2& j\\ m_1& m_2& m\end{array}\right):={\displaystyle \frac{(1)^{j_1j_2m}}{\sqrt{2j+1}}}j_1m_1;j_2m_2|j,m,`$ (41)
with the Clebsch-Gordan coefficient $`j_1m_1;j_2m_2|jm`$. The asterisks mean the complex conjugate and the sum of $`j`$ is taken over $`|j_1j_2|`$ through $`j_1+j_2`$ and the sums of $`n`$ and $`m`$ over $`j`$ through $`j`$.
|
warning/0006/hep-lat0006015.html
|
ar5iv
|
text
|
# 1 The phase diagram for the action (), showing the first order bulk phase transition lines. Taken from Ref. [] but with the endpoint D as obtained in Ref. [].
TIFR/TH/00-28
hep-lat/0006015
Suppressing monopoles and vortices : A possibly smoother approach to scaling ?
Rajiv V. Gavai<sup>1</sup><sup>1</sup>1gavai@tifr.res.in
Department of Theoretical Physics
Tata Institute of Fundamental Research
Homi Bhabha Road
Mumbai 400005, India
ABSTRACT
Suppressing monopoles and vortices by introducing large chemical potentials for them in the Wilson action for the $`SU(2)`$ lattice gauge theory, we study the nature of the deconfinement phase transition on $`N_\sigma ^3\times N_\tau `$ lattices for $`N_\tau =`$ 4, 5, 6 and 8 and $`N_\sigma =`$ 8–16. Using finite size scaling theory, we obtain $`\omega \gamma /\nu =1.93\pm 0.03`$ for $`N_\tau =`$ 4, in excellent agreement with universality. Corresponding determinations for the $`N_\tau =`$ 5 and 6 lattices are also found to be in very good agreement with this estimate. The critical couplings for $`N_\tau `$ = 4, 5, 6 and 8 lattices exhibit large shifts towards the strong coupling region when compared with the usual Wilson action, and suggest a lot smoother approach to scaling.
1. INTRODUCTION
Quantum field theories need regularization schemes to control divergences. The regularization schemes, many different types of which have been used in performing calculations, do not themselves affect physics in any manner, as they are eliminated at the end of all calculations. Long distance physics, such as confinement of quarks in quantum chromodynamics or determination of the hadronic spectrum, is conveniently studied using the lattice regularization. There is a lot of freedom in defining a lattice field theory. In particular, a variety of different choices of the lattice action correspond to the same quantum field theory in the continuum. While many numerical simulations have been performed for the Wilson action for the gauge theories, other choices, some motivated by the desire to find a smoother continuum limit, have also been used. These actions differ merely by irrelevant terms in the parlance of the renormalization group: in the naive classical continuum limit of $`a0`$, they all reduce to the same Yang-Mills action and the differing terms are of higher order in $`a`$.
Investigations of the deconfinement phase transition for mixed actions, obtained by extending the Wilson action by addition of an adjoint coupling term, showed surprising challenges to the above notion of universality. These actions are
$`S_{BC}={\displaystyle \underset{P}{}}\left(\beta \left(1{\displaystyle \frac{1}{2}}\mathrm{Tr}_FU_P\right)+\beta _A\left(1{\displaystyle \frac{1}{3}}\mathrm{Tr}_AU_P\right)\right),`$ (1)
and
$`S_V={\displaystyle \underset{P}{}}\left({\displaystyle \frac{1}{2}}\left(\beta +\beta _V\sigma _P\right)\mathrm{Tr}_FU_P\right).`$ (2)
Here $`U_P`$ denotes the directed product of the basic link variables which describe the gauge fields, $`U_\mu (x)`$, around an elementary plaquette $`P`$. $`F`$ and $`A`$ denote that the traces are evaluated in the fundamental and adjoint representations respectively and the formula $`\mathrm{Tr}_AU=|\mathrm{Tr}_FU|^21`$ is used. $`\sigma _P`$ are $`Z_2`$ plaquette fields associated with the plaquette P and the partition function in the Villain case of eq.(2) has a sum over all possible values for each of them. The first term in both the equations describes the standard SU(2) Wilson action whereas the second term adds an adjoint SO(3) action. For zero adjoint coupling, i.e, for the Wilson action, several finite temperature investigations have shown the presence of a second order deconfinement phase transition. Its critical exponents have been shown to be in very good agreement with those of the three dimensional Ising model, as conjectured by Svetitsky and Yaffe. The verification of the universality conjecture strengthened our analytical understanding of the deconfinement phase transition which, however, came under a shadow of doubt by the results for the mixed actions. Following the deconfinement phase transition into the extended coupling plane by simulating these actions at finite temperature, it was found on a range of temporal lattice sizes for both actions (1) and (2) that:
1. The deconfinement transition was of second order, and in agreement with the conjectured universality exponents, for small values of the adjoint coupling. It became definitely of first order for large values $`\beta _A`$ or $`\beta _V`$.
2. The deconfinement order parameter acquired a nonzero value discontinuously at the transition point for large adjoint couplings.
3. There was no evidence of any other separate bulk transition at those large adjoint couplings, as expected from the results of Refs. . In fact, simulations on larger symmetric lattices even suggested a lack of a bulk phase transition at that adjoint coupling where a first order deconfinement transition for a lattice of temporal size four was clearly seen.
Recently it was shown that suppression of some lattice artifacts such as $`Z_2`$ monopoles and vortices do restore the universality for the action (2): no first order deconfinement transition was found in the entire coupling plane in that case. In this paper, we address this question for the action (1) in the same manner and find that unlike the Villain action, one gains an additional bonus. The approach to scaling seems to become smoother than that for the original Wilson action. The organization of the paper is as follows: In section 2 we define the action we investigate and briefly recapitulate the definitions of various observables used and their scaling laws. We present the detailed results of our simulations in the next section and the last section contains a brief summary of our results and their discussion.
2. THE MODEL AND THE OBSERVABLES
Bhanot and Creutz found that the lattice theory defined by the extended action of eq.(1) has a rich phase structure, shown in Fig. 1. Similar results were obtained for the Villain action (2) by Halliday and Schwimmer. In either case, the $`\beta =0`$ axis describes an $`SO(3)`$ model which has a first order phase transition, denoted by point A in Fig. 1. At $`\beta _A(\mathrm{or}\beta _V)=\mathrm{}`$, the theory reduces to a $`Z_2`$ lattice gauge theory with again a first order phase transition at $`\beta ^{\mathrm{crit}}=\frac{1}{2}\mathrm{ln}(1+\sqrt{2})`$ $``$ 0.44 . For both the mixed actions, these first order transitions extend into the coupling plane, as shown in Fig. 1 by continuous lines. These lines meet at a triple point C and continue as a single line of first order transitions which ends at an apparently critical point D. The proximity of D to the $`\beta _A=0`$ line, which defines the Wilson action, has commonly been held responsible for the abrupt change from the strong coupling region to the scaling region for the Wilson action. It has also been attributed as a possible reason for the dip in the non-perturbative $`\beta `$-function obtained by Monte Carlo Renormalization Group methods. Indeed, its relative closeness to the $`\beta _A=0`$ line for the $`SU(2)`$ theory compared to the $`SU(3)`$ theory has been thought of as a possible reason for the shallower dip in the former case.
The bulk transition in the Villain form of the SO(3) gauge theory is known to be caused by a condensation of $`Z_2`$ monopoles in the strong coupling phase. Defining $`\sigma _c=_{pc}\sigma _P`$ for an elementary cube $`c`$, one finds that $`\sigma _c=1`$ characterizes a monopole located in $`c`$. These monopoles are absent in the weak coupling region, and can be suppressed at stronger couplings by adding a term, $`\lambda _M_c\left(1\sigma _c\right)`$ to the action (2) and setting $`\lambda _M`$ large. Using this extra term, Ref. demonstrated a clear merging of the second order deconfinement line with a first order bulk transition line for $`\lambda _M=`$ 1 for the Villain action (2). Moreover, two separate transitions were shown to exist on the same lattice near the merging point, thereby shedding some light on the paradoxes a\]-c\] above. While pointing to the peculiar role the bulk transitions play in affecting the deconfinement transitions, these results also suggested that $`Z_2`$ electric current loops or vortices have to be suppressed in restoring the universality for the mixed action fully. Defining $`\sigma _l`$ for a link $`l(x,\widehat{\mu })`$ at a point $`x`$ on the lattice in the $`\mu `$th direction as a product of $`\sigma _P`$ of all those plaquettes which share the link $`l`$, i.e, $`\sigma _l=_{p\widehat{}l}\sigma _P`$, one finds that $`\sigma _l=1`$ signals the link to be a part of an $`Z_2`$ electric current loop. Adding further a term $`\lambda _E_l\left(1\sigma _l\right)`$ to the action (2) in addition to the monopole suppression term above, Refs. showed that universality is restored in the entire coupling plane for large $`\lambda _E`$.
While a similar mechanism is expected to work for the Bhanot-Creutz mixed action (1) as well, it is clear that both the monopole and vortex suppression terms added above do not depend on the gauge variables $`U_\mu (x)`$ directly and have to be generalized suitably for addition to $`S_{BC}`$. One possible way is to define $`\sigma _c`$ and $`\sigma _l`$ by replacing $`\sigma _P`$ in them by sign($`\mathrm{Tr}_FU_P`$). Thus the mixed action with chemical potentials for these monopoles and vortices in that case is given by
$`S_{BC,S}=`$ $`{\displaystyle \underset{P}{}}\left(\beta \left(1{\displaystyle \frac{1}{2}}\mathrm{Tr}_FU_P\right)+\beta _A\left(1{\displaystyle \frac{1}{3}}\mathrm{Tr}_AU_P\right)\right)`$ (3)
$`+\lambda _M{\displaystyle \underset{c}{}}\left(1\sigma _c\right)+\lambda _E{\displaystyle \underset{l}{}}\left(1\sigma _l\right),`$
where $`\sigma _c=_{pc}\mathrm{sign}(\mathrm{Tr}_FU_P)`$ and $`\sigma _l=_{p\widehat{}l}\mathrm{sign}(\mathrm{Tr}_FU_P)`$. Comparing the naive classical continuum limit of eq. (3) with the standard $`SU(2)`$ Yang-Mills action, one obtains
$`{\displaystyle \frac{1}{g_u^2}}={\displaystyle \frac{\beta }{4}}+{\displaystyle \frac{2\beta _A}{3}}.`$ (4)
Here $`g_u`$ is the bare coupling constant of the continuum theory. Since the asymptotic scaling equation for the above mixed action with additional (irrelevant) couplings $`\lambda _M`$ and $`\lambda _E`$ can be easily written down in terms of $`g_u`$, it is clear that the introduction of a non-zero $`\beta _A`$, $`\lambda _M`$, or $`\lambda _E`$ does not affect the continuum limit: the theory for each $`\beta _A`$, including the usual Wilson theory for $`\beta _A=0.0`$, flows to the same critical fixed point, $`g_u^c=0`$, in the continuum limit for all $`\lambda _M`$ and $`\lambda _E`$ and has the same scaling behavior near the critical point.
One sees that even for $`\beta _A=0`$ eq. (3) corresponds to a modified Wilson action with different densities of the monopoles and vortices depending on the values of $`\lambda _M`$ and $`\lambda _E`$ respectively. By analogy with the works of Refs. for the Villain action, one expects to eliminate all bulk transition lines in Fig. 1 by setting $`\lambda _M`$ and $`\lambda _E`$ large. In particular, the critical point D is expected to be absent in that case, leading perhaps to a lot smoother transition from the strong coupling region to the scaling region for those values. It has to be stressed though that universality has to be tested afresh for eq.(3) even for $`\beta _A=0`$ to be sure that the above naive argument about the $`\lambda _M`$ and $`\lambda _E`$ terms being irrelevant is correct. This is what we do in the following by determining a critical index of the deconfinement phase transition. We then check whether the passage to scaling is affected by studying the deconfinement transition as a function of the temporal lattice size.
It is interesting to note that no similar change occurs on the fundamental axis ($`\beta _V=0`$) in the case of the Villain action since the variables $`\sigma _P`$ are decoupled from the gauge variables in that case and can be integrated out exactly for any observable depending solely on $`U_\mu (x)`$. However, even in that case the $`\sigma _c`$ and $`\sigma _l`$ could have been defined in terms of sign($`\mathrm{Tr}_FU_P`$) and similar results as we obtain below would be obtained.
We studied the deconfinement phase transition on $`N_\sigma ^3\times N_\tau `$ lattices by monitoring its order parameter and the corresponding susceptibility for $`N_\tau =`$ 4, 5, 6 and 8 and $`N_\sigma =`$ 8, 10, 12, 14, 15, and 16. We used the simple Metropolis algorithm and tuned it to have an acceptance rate $`40`$ %. The expectation values of the observables were recorded every 20 iterations to reduce the autocorrelations. Errors were determined by correcting for the autocorrelations and also by the jack knife method. The observables monitored were the average plaquette, P, defined as the average of $`\mathrm{Tr}_FU_P`$/2 over all independent plaquettes, and the absolute value of the average of the deconfinement order parameter , $`L(\stackrel{}{n})`$, over the three dimensional lattice spanned by $`\stackrel{}{n}`$, where $`L`$ is defined by
$`L(\stackrel{}{n})={\displaystyle \frac{1}{2}}\mathrm{Tr}_F{\displaystyle \underset{\tau =1}{\overset{N_\tau }{}}}U_0(\stackrel{}{n},\tau ).`$ (5)
Here $`U_0(\stackrel{}{n},\tau )`$ is the timelike link at the lattice site $`(\stackrel{}{n},\tau )`$. In order to monitor the nature of deconfinement and bulk phase transitions, we also define the susceptibilities for both $`|L|`$ and P:
$`\chi _{|L|}=`$ $`N_\sigma ^3(L^2|L|^2)`$ (6)
$`\chi _P=`$ $`6N_\sigma ^3N_\tau (P^2P^2).`$ (7)
According to the finite size scaling theory, the peak of the $`|L|`$ (or plaquette) susceptibility at the location of the deconfinement (or bulk) transition should grow on $`N_\sigma ^3\times N_\tau `$ lattices like
$$\chi _{|L|\mathrm{or}P}^{\mathrm{max}}N_\sigma ^\omega ,$$
(8)
for fixed $`N_\tau `$. For a second order transition, $`\omega \gamma /\nu `$, where $`\gamma `$ and $`\nu `$ characterize the growth of the $`|L|`$ (plaquette)-susceptibility and the correlation length near the critical temperature (coupling)on an infinite spatial lattice. If the phase transition were to be of first order instead, then one expects the exponent $`\omega =3`$, corresponding to the dimensionality of the space. In addition, of course, the average $`|L|`$ or plaquette is expected to exhibit a sharp, or even discontinuous, jump and the corresponding probability distribution should show a double peak structure in case of a first order phase transition. Such an analysis of the $`|L|`$-susceptibility for the Wilson action, where only $`\beta `$ is nonzero in eq. (3), yielded an exponent $`\omega =1.93\pm 0.03`$, in good agreement with the corresponding value ($`1.965\pm 0.005`$) for the three dimensional Ising model, and the universality conjecture. Universality of the continuum limit of lattice gauge theories predicts a similar deconfinement transition belonging to the same universality class as the three dimensional Ising model for all values of $`\beta _A`$, $`\lambda _M`$ and $`\lambda _E`$.
3. RESULTS OF THE SIMULATIONS
In view of the fact that even for $`\beta _A=0`$, the action (3) differs from the Wilson action in a non-trivial manner for nonzero $`\lambda _M`$ and $`\lambda _E`$, we concentrate here on $`\beta _A=0`$ and large $`\lambda _M`$ and $`\lambda _E`$. Our aim is to check the impact of the suppression of monopoles and vortices, defined by the sign($`\mathrm{Tr}_FU_P`$) as above, on the critical exponent $`\omega `$ on lattices with fixed $`N_\tau `$. We then wish to study the scaling behaviour of the deconfinement transition by varying $`N_\tau `$. We chose $`\lambda _M=1`$ and $`\lambda _E=5`$ throughout this work. Variations with respect to these as well as $`\beta _A`$ should in principle be investigated although one would expect the results to display universality for sufficiently large $`\lambda _M`$ and $`\lambda _E`$ if the $`\beta _A=0`$ results do so.
3.1 $`N_\tau =4`$
The deconfinement phase transition on $`N_\sigma ^3\times 4`$ lattices for $`N_\sigma =`$ 8, 10, 12 and 14 lattices was studied by first making short hysteresis runs on the smallest lattice to look for abrupt or sharp changes in both the average plaquette $`P`$ and the order parameter $`|L|`$. In the region of strong variation of $`|L|`$, longer runs were made to check whether the $`|L|`$-susceptibility exhibits a peak. Histogramming technique was used to extrapolate to nearby couplings for doing this. A fresh run was made at the $`\chi _{|L|}`$ peak position and the process repeated until the input coupling for the run was fairly close to the output peak position of the susceptibility. The same procedure was used for the bigger lattices also but by starting from the $`\beta _c`$ of the smaller lattice. No peak was found in the vanishingly small plaquette susceptibility throughout, suggesting a lack of any nearby bulk transition. This should be contrasted with the results for the $`\lambda _M=\lambda _E=0`$, which is known to exhibit a peak. Typically 100-200 thousand measurements (2-4 million Monte Carlo iterations) were used to estimate the magnitude of the peak height and the peak location for each $`N_\sigma `$. Table 1 lists our final results for all the $`N_\sigma `$ used. The errors on $`\beta _c`$ were estimated by varying the bin size while those for $`\chi ^{\mathrm{max}}`$ were taken to be the maximum of the errors for all bin sizes. Fitting the peak heights in Table 1 to eq.(8), we obtained
$$\omega =1.93\pm 0.03.$$
(9)
Fig. 2 displays the very good quality of the fit. The critical exponent $`\omega `$ is in excellent agreement with the values for both the standard Wilson action and the 3-dimensional Ising model quoted in previous section. Fitting the peak locations $`\beta _{c,N_\sigma }`$ by the usual finite size scaling expression,
$$\beta _{c,N_\sigma }=\beta _{c,\mathrm{}}+B/N_\sigma ^{1/\nu },$$
(10)
where $`\nu `$ = 0.63 is the correlation length exponent for the 3-dimensional Ising model and $`B`$ is a constant, we obtained $`\beta _{c,\mathrm{}}=1.326\pm 0.006`$, which is shifted by about one from the corresponding value for the $`\lambda _M=\lambda _E=0`$ case which is 2.2986 $`\pm `$ 0.0006 . Inspired by the agreement of $`\omega `$ above, we assumed universality to be true for $`\nu `$ as well in eq. (10). However, any reasonable variation of $`\nu `$ between 0.33 and 1 changes the infinite volume extrapolation for $`\beta _c`$ by a few
($`2`$-3) per cent only. Thus the shift $`\delta \beta =\beta _c^{\mathrm{Wilson}}\beta _c`$ = 0.97 appears to be dominantly due to nonzero $`\lambda _M`$ and $`\lambda _E`$, i.e, suppression of monopoles and vortices.
3.2 $`N_\tau =5`$, 6 and 8
In order to minimize finite spatial volume effects, we chose to work with $`N_\sigma 2N_\tau `$ always, as seen in sec. 3.1. Consequently, the full 4-volume $`N_\sigma ^3N_\tau `$ increased rapidly as we increased $`N_\tau `$. This resulted in progressive shrinking of the coupling interval in which the histogramming technique was reliable. We therefore used many longer runs in the region of strong variation of $`|L|`$ to obtain the susceptibility directly and used the histogramming only for the finer determination of the critical coupling. Fig. 3 exhibits our results for $`|L|`$ and $`P`$ (shown in the inset) as a function of $`\beta `$ for $`N_\tau `$ = 5, 6 and 8. A deconfinement phase transition is clearly visible for all of them. The behaviour of the order parameter for two spatial volumes $`12^3`$ and $`15^3`$ for $`N_\tau `$ = 6 also supports the existence of a transition. This is more clearly seen in the corresponding $`\chi _{|L|}`$ determinations, shown in Fig. 4. The plaquette, on the other hand, describes
a smooth and unique curve for all $`N_\tau `$ and $`N_\sigma `$ values. As the inset in Fig. 3 shows, plaquette values for all these lattices fall on the same curve, indicating an absence of any bulk transition. Tables 2 and 3 list the estimated maxima of $`\chi _{|L|}`$ for $`N_\tau `$ = 5 and 6 for two different spatial volumes along with the corresponding peak locations. As seen in Fig. 4 they are rather close to the input $`\beta `$ at which the long runs were made. Using our value for $`\omega `$ from eq.(9), determined for $`N_\tau `$ = 4, and the peak height for the smaller spatial volume, the $`\chi ^{\mathrm{max}}`$ on the bigger lattice can be predicted. These predictions are listed in the respective tables in the last column and can be seen to be in very good agreement with the direct Monte Carlo determinations. Alternatively, one can fit eq.(8) to the peak heights in Tables 2 and 3 and determine $`\omega `$ again :
$`\omega =`$ $`1.99\pm 0.17\mathrm{for}N_\tau =5,\mathrm{and}`$
$`\omega =`$ $`1.80\pm 0.14\mathrm{for}N_\tau =6.`$ (11)
Both these determinations agree with the canonical values as well as our own value in eq.(9). Not only is the universality of the deconfinement phase transition thus verified on three different temporal lattice sizes, but it also confirms that the same physical phase transition is being simulated on them, thus approaching the continuum limit of $`a0`$ in a progressive manner by keeping the transition temperature $`T_c`$ constant in physical units.
3.3 Scaling of $`T_c`$
Table 4 lists the $`\beta _c`$ (estimated by extrapolating to infinite volume whenever possible) for all the $`N_\tau `$ values we used. The corresponding values for the usual Wilson action, i.e., $`\lambda _M`$ = 0 and $`\lambda _E`$= 0 case, are also given in Table 4 along with the shifts caused by switching on these two couplings. The shifts decrease with increasing $`N_\tau `$ but nevertheless remains sizeable even for the largest lattice we used. Their decrease smoothens the approach to the scaling limit, as we shall see below. Fig. 5 shows $`aT_c=N_\tau ^1`$ as a function of the corresponding critical $`\beta `$ for both our simulations with suppression of monopoles and vortices and the standard Wilson action (without any such suppression). The latter are taken from the compilation of Ref. . The full lines in the figure show the 2-loop asymptotic scaling relation
$$aT_c=\frac{1}{N_\tau }\left(\frac{4b_0}{\beta }\right)^{b_1/b_0^2}\mathrm{exp}\left(\frac{\beta }{8b_0}\right),$$
(12)
where
$$b_0=\frac{11}{24\pi ^2},\mathrm{and}b_1=\frac{17}{96\pi ^4},$$
(13)
are the first two coefficients of the perturbative $`\beta `$-function for the $`SU(2)`$ Yang-Mills theory. The curve in each case was normalized to pass through the $`N_\tau `$= 8 data point. The dashed line describes a ‘phenomenological’ scaling equation which is similar to the eq.(12) but with the exponent increased by a factor of two:
$$aT_c=\frac{1}{N_\tau }\left(\frac{4b_0}{\beta }\right)^{b_1/b_0^2}\mathrm{exp}\left(\frac{\beta }{4b_0}\right).$$
(14)
One sees deviations from asymptotic scaling for both the Wilson action and our action with suppression of monopoles and vortices. The deviations for the same range of $`N_\tau `$ seem larger for our action but then one is also considerably deeper in the strong coupling region of the Wilson action where one a priori would not have even expected any scaling behaviour. As the agreement of our results with the dashed line of eq.(14) in Fig. 5 shows, scaling may hold in this region of $`\beta `$ for the suppressed action, since the relation between $`a`$ and $`\beta `$ in this region (or $`g^2`$) is similar to the asymptotic scaling relation, differing only in the exponent which will cancel in dimensionless ratios of physical quantities. It is clear that as $`\beta \mathrm{}`$, the difference between the two actions must vanish. The shifts in Table 4 do show such a trend although the limiting point is not reached by $`N_\tau `$ = 8 definitely. It seems likely though that the trend of evenly spaced transition points for our action will continue and the dashed line traced by its transition points will merge with the Wilson action by $`N_\tau 25`$ or so, as suggested by its approach to the data for the
Wilson action. If this were to be so, a much smoother approach to continuum limit is to be expected after the suppression of monopoles and vortices. In particular, one expects that dimensionless ratios of physical quantities at the deconfinement phase transition couplings should be constant, already from $`\beta 1.33`$, which is the transition point for the $`N_\tau `$ = 4.
One possible interpretation of the results in Fig. 5 is that the proximity of the point D in Fig. 1 for the usual Wilson action causes the nontrivial curvature visible in the data for the Wilson action, and consequently its approach to the continuum limit is not so smooth. A strong suppression of monopoles and vortices, as performed here, eliminates D, resulting in a smoother approach to scaling. Of course, for very small lattice spacings ( or large $`\beta `$), no significant difference between the two will be seen but for sizeable values of the cut-off one may expect the action with suppression to exhibit a better and smoother approach to the continuum limit. We intend to check this by measuring the glueball spectrum at the critical couplings for $`N_\tau `$ = 4–8. In the meantime, one can try to check this hypothesis by using eq. (14) to convert our $`\chi (\beta )`$ results to $`\chi (T/T_c)`$, i.e, as a function of a dimensionless ratio for various $`N_\tau `$. Fig. 6 depicts the susceptibility $`\chi `$ as a function of $`T/T_c`$ on lattices with $`N_\tau `$ = 5, 6 and 8. Ideally one would have expected all susceptibility data for the same physical volume to fall on the same curve for different $`N_\tau `$. These are the three lowest curves with physical volume 8$`T^3`$ but with $`N_\tau `$= 5, 6 and 8. Unfortunately, the order parameter $`|L|`$ is not ultra-violet safe; it contains divergent contributions in the continuum, making it $`N_\tau `$ (or $`a`$)-dependent even as a function of $`T/T_c`$. Consequently, the corresponding $`\chi `$’s are close but not on any universal curve. On the other hand, an increase in physical volume to 15.6 $`T^3`$ (the $`15^3\times 6`$ data) and 27$`T^3`$ (the $`15^3\times 5`$ data) does seem to sharpen the susceptibility peak progressively, as expected. Although no quantitative analysis can be done meaningfully due to the cut-off dependence of the order parameter itself, the results do show the right trend and thus support a possible scaling in the coupling region of these data points.
4. SUMMARY AND DISCUSSION
The phase diagram of the mixed action of eq. (1) in the fundamental and adjoint couplings, $`\beta `$ and $`\beta _A`$, has been a crucial input in understanding many properties of the $`SU(2)`$ and $`SU(3)`$ lattice theories and their continuum limits. The cross-over to the scaling region from the strong coupling region, as well as the dip in the non-perturbative $`\beta `$-function have been attributed to the location of the end point D of the line of bulk first order phase transition. In fact, even the relative shallowness of the dip for the $`SU(2)`$ case compared to the $`SU(3)`$ case is thought to be due to the closeness of the corresponding end point to the $`\beta _A=0`$ Wilson axis. Adding extra irrelevant terms to the action one obtains the modified action of eq.(3) in which monopoles and vortices can be suppressed by setting the additional couplings to large values. Based on the works for Villain action, one expects the phase diagram to change completely in that case. In particular, no phase transition lines or their critical end point D will be there, causing a smoother transition from the strong coupling region to the scaling region.
In this paper we studied the deconfinement phase transition on the fundamental axis in the $`(\beta ,\beta _A)`$ coupling plane but with $`\lambda _M`$ = 1 and $`\lambda _E`$ = 5, i.e., with strong suppression of monopoles and vortices. Our finite size scaling analysis yielded 1.93 $`\pm `$ 0.03 for the critical exponent $`\omega \gamma /\nu `$ for lattices with $`N_\tau `$ = 4. This value is in excellent agreement with that for the Wilson action and the three dimensional Ising model, thus verifying the naive universality of the modified action. However, as a result of the suppression, the critical coupling is shifted by about unity compared to the Wilson case. Our results on $`N_\tau `$ = 5 and 6 also yielded similar values for $`\omega `$ albeit with larger errors, confirming that the same physical transition was being studied this way as a function of the lattice cut-off, $`a`$. While the $`aT_c=N_\tau ^1`$ was found to vary slower than expected from the asymptotic scaling relation (12) for $`N_\tau `$ = 4–8, the data did obey a similar relation with a factor of two larger exponent. A straightforward extrapolation suggests the results from the modified action will merge with those of Wilson action for large $`N_\tau `$ ( of about $``$ 25), as expected in the limit of vanishing lattice spacing $`a`$. This suggests that the suppression makes the approach from the strong coupling side to the scaling side much smoother than that for the unsuppressed Wilson action, allowing us to simulate the theory at smaller $`\beta `$. It will be interesting to see if dimensionless ratios of physical quantities such as glueball masses or string tension with $`T_c`$ are constant in the range of critical couplings explored here. Since the phase diagram for $`SU(3)`$, and indeed for $`SU(N)`$ lattice gauge theories, is similar and the same mechanism is expected to work for them, it will also be interesting to study such suppression in those theories as well. However, additional possibilities for topological objects may add further complications and may make it necessary to suppress them as well.
6. ACKNOWLEDGMENTS
It is a pleasure to acknowledge interesting discussions with Sourendu Gupta.
|
warning/0006/cond-mat0006048.html
|
ar5iv
|
text
|
# Interface fluctuations in disordered systems: Universality and non-Gaussian statistics
## I Introduction
Quenched disorder plays a crucial role in a huge variety of physical systems. One of the most prominent examples for such systems is a domain wall in Ising-like systems. Such interfaces can couple to various types of disorder such as a random pinning potential which can be provided by impurities in the system. For interfaces with internal dimensions $`d4`$, such disorder is known to radically modify the structure of the interface and also its dynamics. The structure of such systems typically is self-affine, resembling pure systems at criticality. In contrast to the statics, the dynamics is very different from a critical one, since it lacks the characteristic scaling relations. It is dominated by high energy barriers which lead to an exponentially slow, glassy dynamics. This behavior was observed experimentally by position-space imaging techniques and it should also be possible to study such systems by means of scattering techniques.
The theoretical understanding of such systems has made substantial progress during the last 15 years. Although the self-affinity of the structure suggests the use of renormalization-group (RG) techniques to calculate the characteristic exponents (such as the roughness exponent), the analysis intricate since the flow of an infinite number of relevant parameters has to be studied. This can be achieved in the framework of a functional renormalization group which was employed by D.S. Fisher on the basis of the hard-cutoff renormalization group (HCRG) scheme of Wegner and Houghton. This study was the prototype for subsequent generalizations to a variety of different physical systems with a higher number of components of the displacement field or to periodic systems. The latter class includes systems such as charge-density waves, Wigner crystals, and vortex lines in type-II superconductors. In all these systems disorder plays a qualitatively similar role.
However, this HCRG is known to suffer from pathologies related to the sharp cutoff as noted already in Ref. . This cutoff procedure leads to a long-ranged and oscillatory behavior of the field correlations in real space, which leads to the generation of highly nonlocal interactions on a coarse-grained level. It requires particular care to interpret these interactions as the renormalization of local quantities such as the interface stiffness. These pathologies emerge not only in the context of disordered systems but also in pure systems, for example in the sine-Gordon model which describes the roughening transition of crystal surfaces. In the latter context, it was pointed out by Nozières and Gallet that these pathologies may affect the renormalization of the stiffness constant and ultimately the scaling behavior at criticality.
For this reason it is of principal interest to reinvestigate the prototype model for disordered systems in the framework of a soft-cutoff renormalization group (SCRG). In this scheme fluctuations are regularized on small length scales by a smooth cutoff function. There exist several variants of SCRGs (for recent review articles on RG schemes, see e.g. Refs. ). The scheme of Wilson and Kogut is a SCRG. Unfortunately, it is very clumsy to work with since already the free interface is described by a rather complicated Gaussian fixed point. We use the scheme of Polchinski, which we find convenient for our purposes.
Starting from this SCRG method we confirm the value of the roughness exponent found in the HCRG scheme. In addition, we explicitly demonstrate its universality, i.e. its independence from the cutoff function. In this respect our calculation parallels the demonstration of universality for the bulk exponent $`\eta `$ of the pure Ising model.
In most theoretical analyses disorder is assumed to be Gaussian distributed, i.e. that higher cumulants of the disorder distribution are negligible. Within the present scheme higher cumulants play a central role. Although in general these cumulants are non-universal, we determine their functional form at the RG fixed-point point since they are physically meaningful for energy fluctuations on large scales and since they will be needed for the analysis of the model to subleading order in $`ϵ`$.
In the next section II we specify the model. In order to keep this paper self-contained we include a brief scaling analysis in Sec. III. The application of the Polchinski SCRG scheme to the present model is discussed in Sec. IV. The renormalization flow equations are derived and evaluated in Sec. V and our results are discussed in Sec. VI.
## II Model
Our analysis is based on the model Hamiltonian
$$=d^dx\left\{\frac{\gamma }{2}(\mathbf{}\varphi )^2+V(\varphi (𝐱),𝐱)\right\},$$
(1)
which is composed of an elastic energy and a pinning energy. $`\gamma `$ is the elastic stiffness constant of the interface. $`V`$ represents a quenched random potential acting on the interface. In the simplest case the disorder potential $`V`$ is Gaussian distributed with zero average and short-ranged correlations in its dependence on $`𝐱`$ and $`\varphi `$,
$$\overline{V(\varphi _1,𝐱_1)V(\varphi _2,𝐱_2)}=R(\varphi _1\varphi _2)\delta ^{(d)}(𝐱_1𝐱_2).$$
(2)
In principle one could allow for a more general form of the correlator (2), where the dependence on $`𝐱`$ and $`\varphi `$ does not factorize and which has a finite correlation length also in $`𝐱`$ directions. However, in a coarse-grained description this correlation length shrinks to zero and one ends up with a dependence of the form (2). Therefore this correlation length can be taken as zero right away without a modification of the large-scale properties of the system. On the contrary, the $`\varphi `$ dependence of the correlator has to be described by the function $`R(\varphi )`$ with a finite width. Otherwise the typical force density $`\frac{}{\varphi }V(\varphi ,𝐱)`$ would be infinite and the problem would be ill defined.
We subsequently focus on disorder of the random-potential type, where $`R(\varphi )`$ decays for large values of $`\varphi `$. Thus, we exclude systems with random-field disorder, for which many large-scale features can be derived already from Imry-Ma type scaling arguments. We further specialize to short ranged random potentials where $`R(\varphi )`$ decays faster than any power of $`\varphi `$. In addition, we assume a statistical reflection symmetry $`\varphi \varphi `$ such that the correlation function is even, $`R(\varphi )=R(\varphi )`$.
For the analysis of the model it is convenient to apply the standard replica trick in order to anticipate the disorder average and to restore translation symmetry. After replicating the system $`n`$ times and averaging over disorder, one obtains the Hamiltonian
$``$ $`=`$ $`_{\mathrm{el}}+_{\mathrm{pin}},`$ (4)
$`_{\mathrm{el}}`$ $`=`$ $`{\displaystyle \frac{\gamma }{2}}{\displaystyle \underset{\alpha =1}{\overset{n}{}}}{\displaystyle d^dx(\mathbf{}\varphi ^\alpha )^2},`$ (5)
$`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{\alpha ,\beta =1}{\overset{n}{}}}{\displaystyle d^dxR(\varphi ^\alpha (𝐱)\varphi ^\beta (𝐱))},`$ (6)
which we have decomposed into an “elastic” part and a “pinning” part. In this formulation, the quenched disorder is represented by an effective interaction between different replicas. $`T`$ denotes the temperature of the system.
Since we have taken the correlation length in $`𝐱`$ direction as zero, the interaction is local in $`𝐱`$ and the interaction energy is invariant under an arbitrary replica-independent tilt $`\varphi ^\alpha (𝐱)\varphi ^\alpha (𝐱)+\delta \varphi (𝐱)`$ of the interface. This symmetry will play an important role below.
As usual, the partition function can be written as a functional integral
$$𝒵=\mathrm{Tr}_\varphi e^{\frac{1}{T}[\varphi ]}$$
(7)
with $`\mathrm{Tr}_\varphi =_𝐤_\alpha 𝑑\varphi ^\alpha (𝐤)`$ as integral over the Fourier modes $`\varphi (𝐤)=d^dxe^{i𝐤𝐱}\varphi (𝐱)`$. This model has to be regularized at large momenta by a momentum cutoff $`\mathrm{\Lambda }`$ related to a microscopic length $`a=2\pi /\mathrm{\Lambda }`$. Our particular choice for the regularization will be discussed in Sec. IV B.
## III Scaling analysis
Before we study this model by means of the renormalization group technique it is instructive to perform a scaling analysis. Although a lucid presentation of this analysis can be found in Ref. , we give a brief summary thereof in order to make this article self-contained.
To describe the shape fluctuations of the interface we are primarily interested in the pair correlation
$`\overline{\varphi (𝐤)\varphi (𝐤^{})}=\underset{n0}{lim}\varphi ^\alpha (𝐤)\varphi ^\alpha (𝐤^{})=:C(𝐤)\delta (𝐤+𝐤^{})`$ (8)
(here $`\alpha `$ is not summed over) after averaging over thermal fluctuations (denoted by $`\mathrm{}`$) and over the disorder distribution (denoted by $`\overline{\mathrm{}}`$). Since the latter average has been anticipated in the replicated system (II) it no longer appears in the central term in Eq. (8).
On large length scales $`k\mathrm{\Lambda }`$ the displacement correlation is found to be self-affine,
$`C(𝐤){\displaystyle \frac{T}{\gamma }}k^{d2\zeta }`$ (9)
with a roughness exponent $`\zeta `$. For $`\zeta >0`$, the interface is rough and the relative displacement
$`w(𝐱𝐱^{})`$ $`:=`$ $`\overline{[\varphi (𝐱)\varphi (𝐱^{})]^2}`$ (10)
$`=`$ $`\underset{n0}{lim}[\varphi ^\alpha (𝐱)\varphi ^\alpha (𝐱^{})]^2`$ (11)
increases like
$`w(𝐱)x^{2\zeta },`$ (12)
on large scales $`xa`$. We always assume $`\zeta <1`$ since otherwise the model breaks down since higher powers of $`\mathbf{}\varphi `$ become relevant in the elastic energy. For $`\zeta <0`$, the interface is flat and $`w(𝐱)`$ converges to a finite value $`w_{\mathrm{}}`$ for $`𝐱\mathrm{}`$ and
$`w_{\mathrm{}}w(𝐱)x^{2\zeta }.`$ (13)
In order to examine the relevance of disorder we analyze the properties of the model under a rescaling of lengths, field, and temperature:
$`𝐱`$ $`=`$ $`e^{\mathrm{}}𝐱_{\mathrm{}},`$ (15)
$`\varphi ^\alpha (𝐱)`$ $`=`$ $`e^\zeta \mathrm{}\varphi _{\mathrm{}}^\alpha (𝐱_{\mathrm{}}),`$ (16)
$`T`$ $`=`$ $`e^\theta \mathrm{}T_{\mathrm{}}.`$ (17)
Here we have introduced the scaling parameter $`\mathrm{}`$ – which can be viewed as logarithmic length scale – and the energy scaling exponent $`\theta `$. The role of $`\theta `$, which is absent in usual critical phenomena, will become clear soon. The scaling hypothesis requires the statistical weights and therefore $`/T`$ to be invariant under rescaling. The elastic energy remains invariant only if the stiffness is rescaled according to $`\gamma _{\mathrm{}}=e^{(d2+2\zeta \theta )\mathrm{}}\gamma `$, which reads in differential form
$`{\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{scal}}\gamma _{\mathrm{}}=(d2+2\zeta \theta )\gamma _{\mathrm{}}.`$ (18)
Thus the stiffness is scale invariant for
$`\theta =2\zeta +d2.`$ (19)
In the absence of disorder one can achieve the scale invariance of both $`\gamma `$ and $`T`$ with the choice $`\theta =0`$ and $`\zeta =\zeta _{\mathrm{th}}:=\frac{2d}{2}`$.
Since in the absence of disorder the interface is flat for $`d>2`$ one may analyze the relevance of disorder (i.e. of $`_{\mathrm{pin}}`$) performing a Taylor expansion of $`R(\varphi )=_{m0}\frac{R_{2m}}{(2m)!}\varphi ^{2m}`$, assuming analyticity of the (unrenormalized) correlator $`R`$ (compare Ref. ). A scale invariance of the statistical weights would require
$`{\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{scal}}R_{2m,\mathrm{}}=(d+2m\zeta 2\theta )R_{2m,\mathrm{}}.`$ (20)
If we insert $`\zeta =\zeta _{\mathrm{th}}`$ and $`\theta =\theta _{\mathrm{th}}=0`$ into this equation, we find $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}R_{2m,\mathrm{}}=(d+m(2d))R_{2m,\mathrm{}}`$. In particular, $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}R_{2,\mathrm{}}=2R_{2,\mathrm{}}`$ and we conclude that disorder is relevant. The couplings $`R_{2m,\mathrm{}}`$ with $`m>1`$ are less relevant for $`d>2`$ under rescaling with the thermal exponents.
The scale invariance of both $`_{\mathrm{el}}/T`$ and $`_{\mathrm{pin}}/T`$ can be achieved only for $`\theta 0`$. Requiring $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}\gamma _{\mathrm{}}=0`$ and $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}R_{2,\mathrm{}}=0`$ one finds $`\theta =\theta _{\mathrm{rf}}=:2`$ and $`\zeta =\zeta _{\mathrm{rf}}:=\frac{4d}{2}`$, which implies the roughness of the interface in $`d<4`$. This “random-force” value of the roughness exponent is the value one finds in a perturbative treatment of disorder, where the dependence of the pinning force on $`\varphi `$ is neglected an which is represented by the Hamiltonian
$`_{\mathrm{rf}}={\displaystyle \underset{\alpha \beta }{}}{\displaystyle d^dx\left\{\frac{\gamma }{2}\delta _{\alpha \beta }(\mathbf{}\varphi ^\alpha )^2\frac{R_2}{4T}(\varphi ^\alpha \varphi ^\beta )^2\right\}}`$ (21)
which is the contribution to (II) bilinear in the displacement.
If we reexamine the relevance of the couplings $`R_{2m}`$ with the “random-force” exponents, we find $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}R_{2m,\mathrm{}}=(4d)(m1)R_{2m,\mathrm{}}`$. Thus the couplings $`R_{2m}`$ with $`m>1`$ are irrelevant only in $`d>4`$. Their relevance in $`d<4`$ indicates that the roughness cannot be obtained from perturbation theory and that the correct value of the roughness exponent most likely is not given by $`\zeta _{\mathrm{rf}}`$.
In order to obtain the roughness exponent in $`d4`$, a renormalization group (RG) calculation is required. This RG has to be a functional RG since all terms in the Taylor series of $`R(\varphi )`$ are relevant, as $`\zeta >0`$ and the relevance of coefficients $`R_{2m}`$ increases with increasing $`m`$ for any value of $`\zeta >0`$.
Before we start such a RG calculation, it is worthwhile to mention that a good estimate of the scaling exponent in $`d4`$ can be obtained from the Flory argument. In this argument one assumes that for a rough interface the short-ranged disorder correlator can be approximated by $`R(\varphi )R^{}\delta (\varphi )`$ with a weight $`R^{}:=𝑑\varphi R(\varphi )`$. Then $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}R_{\mathrm{}}^{}=(d\zeta 2\theta )R_{\mathrm{}}^{}`$. The requirement that $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}\gamma _{\mathrm{}}=0`$ and $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}R_{\mathrm{}}^{}=0`$ then leads to the Flory value of the roughness exponent $`\zeta =\zeta _\mathrm{F}:=\frac{4d}{5}`$.
## IV Renormalization Group
In $`d>4`$ the couplings $`R_{2m}`$ are irrelevant for $`m>1`$ and the large-scale properties of the model are governed by the Gaussian random-force model (21). In analogy to usual critical phenomena (see, e.g. Refs. ) we assume that in $`d=4ϵ<4`$ dimensions the large-scale fluctuations are described by a fixed point which is close to this Gaussian fixed point for small $`ϵ`$. However, unlike for usual critical phenomena, this will be a zero-temperature fixed point: According to Eq. (17), which is equivalent to
$`{\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{scal}}T_{\mathrm{}}=\theta T_{\mathrm{}},`$ (22)
the effective scale-dependent temperature $`T_{\mathrm{}}`$ flows to zero on large length scales $`xe^{\mathrm{}}`$ provided disorder increases the roughness of the interface, i.e. $`\zeta >\zeta _{\mathrm{th}}`$ and $`\theta >0`$ according to Eq. (19).
### A Rescaling
Since we must perform a functional RG analysis, we first reformulate the flow of the couplings (such as $`\gamma `$ and $`R_{2m}`$ in section III) resulting from the rescaling (III) in a closed form:
$`{\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{scal}}`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle d^dx\{𝐱\mathbf{}\varphi ^\alpha (𝐱)+\zeta \varphi ^\alpha (𝐱)\}\frac{\delta }{\delta \varphi ^\alpha (𝐱)}}`$ (24)
$`+\theta T^2{\displaystyle \frac{}{T}}{\displaystyle \frac{}{T}}.`$
Here and henceforth we drop the subscript $`\mathrm{}`$ for simplicity of notation. In writing $`\frac{d}{d\mathrm{}}|_{\mathrm{scal}}`$ we stress that this is only the scaling contribution to the flow. To obtain the full RG flow, this contribution has to be combined with a contribution $`\frac{d}{d\mathrm{}}|_{\mathrm{int}}`$ which arises from integrating out modes of the field $`\varphi `$.
### B Regularization
To calculate this second contribution, Fisher and Balents and Fisher have used the “hard cutoff” scheme of Wegner and Houghton in their analysis of the present problem. We choose the “soft cutoff” scheme of Polchinski for the reasons outlined in the introduction.
We regularize the theory by modifying the propagator using Schwinger’s proper time method (see, e.g., Ref. ). To this end, we rewrite $`_{\mathrm{el}}`$ in Fourier space,
$$_{\mathrm{el}}=\frac{1}{2}(\varphi ,G_\mathrm{\Lambda }^1\varphi ):=\frac{1}{2}\underset{\alpha }{}\frac{d^dk}{(2\pi )^d}G_\mathrm{\Lambda }^1(𝐤)|\varphi ^\alpha (𝐤)|^2.$$
(25)
Herein the propagator
$$G_\mathrm{\Lambda }(𝐤):=G_{\mathrm{}}(𝐤)f(k/\mathrm{\Lambda })\text{ with }G_{\mathrm{}}(𝐤)=1/(\gamma k^2)$$
(26)
is regularized by the cutoff function $`f`$. In order to suppress fluctuations on short length scales, $`f(z)`$ has to vanish for large $`z`$, whereas it has to satisfy $`f1`$ for $`z0`$ in order not to modify the properties of the model on large length scales. $`f(k/\mathrm{\Lambda })`$ can roughly be interpreted as “weight of modes”: In the calculation of the pair correlation (11) one may consider the density of modes in Fourier space as being reduced to a fraction $`f`$ with respect to the unregularized model.
We will keep the function $`f`$ general as long as possible, which is desirable to verify the independence of the results on the regularization procedure. However, we implicitly assume $`f`$ to be monotonous in order to have smooth functions and to have no further intrinsic length scales. For illustrative purposes we occasionally choose the specific form
$`f(z)=e^{z^2/2}.`$ (27)
As long as $`f^1(k/\mathrm{\Lambda })`$ is analytic, a Taylor expansion of $`f^1(k/\mathrm{\Lambda })`$ shows that the regularization is achieved by contributions to the Hamiltonian which are irrelevant on large length scales since they involve higher orders of $`k`$. The HCRG method of Wegner and Houghton can be considered as non-analytic choice $`f(z)=\mathrm{\Theta }(1z)`$.
### C Mode integration
In Polchinski’s scheme, the mode integration works as follows (see also Ref. ). One introduces an additional field $`\varphi ^<`$ in the partition sum and defines $`\varphi ^>:=\varphi \varphi ^<`$. This can be achieved in such a way that
$`𝒵`$ $``$ $`\mathrm{Tr}_{\varphi ^<}e^{\frac{1}{2T}(\varphi ^<,G_<^1\varphi ^<)}`$ (29)
$`\times \mathrm{Tr}_{\varphi ^>}e^{\frac{1}{2T}(\varphi ^>,G_>^1\varphi ^>)\frac{1}{T}_{\mathrm{pin}}[\varphi ^<+\varphi ^>]}.`$
The “slow modes” $`\varphi ^<`$ are regularized by the propagator $`G_<:=G_{\mathrm{\Lambda }_<}`$ with an infinitesimally reduced cutoff $`\mathrm{\Lambda }_<:=e^d\mathrm{}\mathrm{\Lambda }`$, whereas the “fast modes” $`\varphi ^>`$ have a propagator $`G_>:=G_\mathrm{\Lambda }G_<`$. In Eq. (29) — which is derived from Eq. (7) in Appendix A — proportionality factors independent of $`\varphi `$ have been dropped.
In the representation (29) the modes $`\varphi ^>`$ can be integrated out. This can be done exactly in the limit, where $`G_>`$ is small (it is of order $`d\mathrm{}`$) since then also $`\varphi ^>`$ is small (of order $`(d\mathrm{})^{1/2}`$). Then an expansion of $`_{\mathrm{pin}}[\varphi ^<+\varphi ^>]`$ to second order in $`\varphi ^>`$ is sufficient to establish the differential RG. For infinitesimal $`d\mathrm{}`$ the propagator of the fast modes is
$`G_>(𝐤)`$ $`=`$ $`g(𝐤)d\mathrm{}+𝒪(d\mathrm{}^2),`$ (31)
$`g(𝐤)`$ $`=`$ $`\mathrm{\Lambda }{\displaystyle \frac{d}{d\mathrm{\Lambda }}}G_\mathrm{\Lambda }(𝐤).`$ (32)
Integration over the field $`\varphi ^>`$ then yields a flow (see Appendix A):
$`{\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{int}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{\alpha }{}}{\displaystyle _{12}}g_{12}\left[T{\displaystyle \frac{\delta ^2_{\mathrm{pin}}}{\delta \varphi _1^\alpha \delta \varphi _2^\alpha }}{\displaystyle \frac{\delta _{\mathrm{pin}}}{\delta \varphi _1^\alpha }}{\displaystyle \frac{\delta _{\mathrm{pin}}}{\delta \varphi _2^\alpha }}\right].`$ (33)
We introduce the abbreviations $`_{ijk\mathrm{}}:=_i_j_k\mathrm{}`$, $`_i:=d^dx_i`$, $`\varphi _i:=\varphi (𝐱_i)`$ and $`g_{ij}:=g(𝐱_i𝐱_j)`$ to keep expressions compact. We call the first term in Eq. (33) the “contraction” part of the generator (since legs of a vertex are contracted in a diagrammatic representation) and the second term the “composition” part (since new vertices are composed by linking two vertices).
The flow contribution (33) has a simpler form than the corresponding contribution in the HCRG scheme, since (33) contains terms only up to first order in $`g`$ (and second order in the Hamiltonian), whereas the HCRG generator contains terms of arbitrarily high order. This difference is due to the fact that $`G_>`$ is a bounded function of order $`d\mathrm{}`$, in contrast to the HCRG scheme, where $`G_>`$ is singular for momenta at the cutoff, for which reason the power counting of orders in $`d\mathrm{}`$ breaks down and also higher powers in $`G_>`$ contribute to $`𝒪(d\mathrm{})`$. The simplicity of the generator is one important argument in favor of the SCRG scheme.
The complete functional RG consists of the combination of integration over modes (33) with rescaling (24),
$`{\displaystyle \frac{d}{d\mathrm{}}}={\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{scal}}+{\displaystyle \frac{d}{d\mathrm{}}}|_{\mathrm{int}}.`$ (34)
Although we have defined the regularization in Fourier space, we ultimately find it more convenient to evaluate the RG flow in position space because of the functional structure of $`\frac{d}{d\mathrm{}}|_{\mathrm{int}}`$.
To close this exposition of the RG method we emphasize that the RG (34) is exact, provided the dependence of the functional $`_{\mathrm{pin}}[\varphi ^<+\varphi ^>]`$ on small $`\varphi ^>`$ (which is of order $`(d\mathrm{})^{1/2}`$because of Eq. (31)) is captured in order $`d\mathrm{}`$ by a functional Taylor expansion to second order in $`\varphi ^>`$. This analyticity assumption is common to HCRG and SCRG approaches and will be reexamined later on after we have found the fixed point for the disordered interface and the generation of nonanalytic features.
### D General features
We now turn to the evaluation of the RG flow for our model (II). Due to the locality of the disorder correlator in $`𝐱`$ space, the model owns a stochastic symmetry under a tilt $`\varphi ^\alpha (𝐱)\varphi ^\alpha (𝐱)+𝐯𝐱`$ with a constant vector $`𝐯`$. As a consequence of this symmetry there is no renormalization of the stiffness $`\gamma `$ arising from the mode integration. Hence the flow of $`\gamma `$ arises only from rescaling,
$`{\displaystyle \frac{d}{d\mathrm{}}}\gamma =(d2+2\zeta \theta )\gamma .`$ (35)
This implies the validity of (19) not only in a scaling analysis but also at the RG fixed point.
In contrast to $`\gamma `$, the couplings of the disorder part of the Hamiltonian, $`_{\mathrm{pin}}`$, will not only be rescaled but also renormalized by the mode integration. This renormalization results not only in a modification of the second-order cumulant $`R`$ of the disorder distribution but also in the generation of contributions to $`_{\mathrm{pin}}`$ of an extended functional form. The functional form of the emerging terms can be recognized from the action of the generator (33) on $`_{\mathrm{pin}}`$. The unrenormalized functional (6) evaluates the field in two different replicas but only at a single point $`𝐱`$. Therefore we call functionals of this type “2-replica” functionals and “1-point” functionals. The action of the first “contraction” term in (33) results again in a 1-point functional. However, the “composition” term evaluates the field at two different positions, it is a 2-point and 3-replica term. Fortunately, at the expense of the more complicated functional form we get “smaller” terms (i.e. of higher order in $`R`$). This 2-point term is of second order in $`R`$. Successive iterations generate $`p`$-point and $`(p+1)`$-replica terms in order $`R^p`$.
The extended functional form of the Hamiltonian can be captured in the form
$`{\displaystyle \frac{1}{T}}_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle \frac{1}{T}}𝒬{\displaystyle \frac{1}{2T^2}}+{\displaystyle \frac{1}{3!T^3}}𝒮\mathrm{}`$ (36)
where $`𝒬`$, $``$, and $`𝒮`$ are 1-, 2-, and 3-replica terms, which represent cumulants of the disorder distribution. To become specific, let us denote the pinning energy of the replicated system before disorder averaging by $`_V=_\alpha _𝐱V(\varphi ^\alpha (𝐱),𝐱)`$. Then the contributions to the extended Hamiltonian can be identified as cumulants of $`_V`$:
$`𝒬=_V,=_V^2_\mathrm{c},𝒮=_V^3_\mathrm{c},\mathrm{}`$ (37)
We will not keep track of $`𝒬`$ since this is a field-independent constant because of the stochastic symmetry $`\varphi (𝐱)\varphi (𝐱)+\delta \varphi `$ for an arbitrary constant $`\delta \varphi `$.
Anticipating that $`R=𝒪(ϵ)`$ at the fixed point, it is sufficient to retain $`p`$-point and $`(p+1)`$-replica terms to study the RG flow in $`𝒪(ϵ^p)`$. As we will see in Sec. V, it is possible to retain the full functional form of the Hamiltonian without need for truncations. To keep track of this functional form it is more convenient to perform the RG analysis in position space than in momentum space.
## V RG flow and fixed point
In this section we analyze the RG flow in order $`ϵ^2`$, from which we obtain the roughness exponent $`\zeta `$ in order $`ϵ`$. We find agreement with previous HCRG calculations in this exponent. In addition, we are able to demonstrate the universality of this exponent in the sense of its independence on the cutoff function $`f`$. We also explicitly obtain the third-order cumulant $`𝒮`$ of the disorder distribution.
In the HCRG analysis it was not necessary to keep track of 3-replica terms for the determination of $`\zeta `$ in order $`ϵ`$. Although 3-replica terms are generated, they do not feed back into the equation that determines $`\zeta `$ (in order $`ϵ`$) and there is no need to keep track of $`𝒮`$. The situation is different in the SCRG scheme, since $``$ is renormalized only via $`𝒮`$. However, 3-replica terms are generated also in the HCRG scheme. For the analysis of the consistency of the $`ϵ`$-expansion one has to include also this term into consideration.
In order to determine $`\zeta `$ in order $`ϵ`$, the it is sufficient to sufficient to consider the Hamiltonian in a functional form which can be parameterized by functions $`R_+`$, $`R_{++}`$ and $`S_{}`$ according to
$``$ $`=`$ $`{\displaystyle \underset{\alpha \beta }{}}{\displaystyle _1}R_+(\varphi _1^{\alpha \beta })`$ (40)
$`+{\displaystyle \underset{\alpha \beta }{}}{\displaystyle _{12}}R_{++}(\varphi _1^{\alpha \beta },\varphi _2^{\alpha \beta };𝐱_{12})+\mathrm{},`$
$`𝒮`$ $`=`$ $`{\displaystyle \underset{\alpha \beta \gamma }{}}{\displaystyle _{12}}S_{}(\varphi _1^{\alpha \beta },\varphi _2^{\alpha \gamma };𝐱_{12})+\mathrm{},`$ (41)
where we rewrite $`R_+(\varphi )R(\varphi )`$ to display that $`R`$ is an even function of $`\varphi `$. To compress notation we further introduce $`\varphi _i^{\alpha \beta }:=\varphi ^\alpha (𝐱_i)\varphi ^\beta (𝐱_i)`$. We thus keep functional contributions up to 3-replica and 2-point type. It is not necessary to keep these functionals in the most general form. Instead, the functions obey certain symmetry properties. $`R_{++}`$ is an even function of both field arguments, $`S_{}`$ is an odd function of both field arguments (the detailed symmetry properties are specified below). Both functions evaluate the fields at two points and can therefore depend also on the distance $`𝐱_{ij}:=𝐱_i𝐱_j`$ between these points.
A priori there is an ambiguity in representing a functional (such as $``$) in terms of $`p`$-point kernels (such as the 1-point and 2-point functions $`R_+`$ and $`R_{++}`$). For example, the simultaneous replacement $`R_+(\varphi _1^{\alpha \beta })R_+(\varphi _1^{\alpha \beta })+\mathrm{\Delta }R_+(\varphi _1^{\alpha \beta })`$ and $`R_{++}(\varphi _1^{\alpha \beta },\varphi _2^{\alpha \beta };𝐱_{12})R_{++}(\varphi _1^{\alpha \beta },\varphi _2^{\alpha \beta };𝐱_{12})\mathrm{\Delta }R_+(\varphi _1^{\alpha \beta })\delta (𝐱_{12})`$ leaves the functional $``$ unchanged. In this way one could absorb the 1-point kernel into the 2-point kernel (in general, low-point kernels into higher-point kernels). In the case of $`𝒮`$ we avoid introducing a 1-point kernel. In the case of $``$ we extract the 1-point kernel in order to ensure that the 2-point kernel contributes to $``$ only in order $`ϵ^2`$ (see below). A unique distinction between the 1-point and higher-point kernels in a functional is achieved by requirement that the integral over higher-point kernels must vanish for a spatially constant field $`\varphi ^{\alpha \beta }(𝐱)=\varphi _0^{\alpha \beta }`$, in particular $`d^dxR_{++}(\varphi _0^{\alpha \beta },\varphi _0^{\alpha \beta };𝐱)=0`$.
The flow equation (34) for the Hamiltonian can be represented as flow equation of the parameter functions (for a diagrammatic representation, see Fig.1)
$`{\displaystyle \frac{d}{d\mathrm{}}}R_+(\varphi )`$ $`=`$ $`\{d2\theta +\zeta \varphi \}R_+(\varphi )+{\displaystyle \frac{2}{3}}{\displaystyle _2}_1_2S_{}(\varphi _1,0;𝐱_{12})g_{12}`$ (44)
$`{\displaystyle \frac{1}{3}}{\displaystyle _2}_1_2S_{}(\varphi _1,\varphi _1;𝐱_{12})g_{12}+\mathrm{},`$
$`{\displaystyle \frac{d}{d\mathrm{}}}R_{++}(\varphi _1,\varphi _2;𝐱_{12})`$ $`=`$ $`\{2d2\theta +\zeta \varphi _i_i+𝐱_i\mathbf{}_i\}R_{++}(\varphi _1,\varphi _2;𝐱_{12}){\displaystyle \frac{1}{3}}_1_2S_{}(\varphi _1,\varphi _2;𝐱_{12})g_{12}`$ (46)
$`+{\displaystyle \frac{1}{3}}\delta _{12}{\displaystyle _3}_1_2S_{}(\varphi _1,\varphi _1;𝐱_{13})g_{13}+\mathrm{},`$
$`{\displaystyle \frac{d}{d\mathrm{}}}S_{}(\varphi _1,\varphi _2;𝐱_{12})`$ $`=`$ $`\{2d3\theta +\zeta \varphi _i_i+𝐱_i\mathbf{}_i\}S_{}(\varphi _1,\varphi _2;𝐱_{12})3R_+(\varphi _1)R_+(\varphi _2)g_{12}+\mathrm{}.`$ (47)
$`\mathbf{}_i:=\frac{}{𝐱_i}`$ is a partial derivative that acts only on the explicit position arguments $`𝐱_{ij}`$ but not on the arguments of the fields. In $`𝐱_i\mathbf{}_i`$ summation over $`i`$ is assumed implicitly ($`1ip`$ in $`p`$-point kernels). Further on, $`_i`$ denotes the partial derivative acting on the the $`i`$th field argument of the kernel function (which is not necessarily $`\varphi _i`$; if there is only one argument we drop the subscript, $`:=_1`$).
In the flow equations (V) we neglect all terms proportional to temperature since $`T`$ flows to zero according to Eq. (22). One immediately recognizes that even if there is only a function $`R_+`$ in the unrenormalized model, the kernel $`S_{}`$ is generated from the composition of two $`R_+`$. The kernel $`S_{}`$ then feeds back into the flow equations for $`R_+`$ and generates $`R_{++}`$.
Since $`R_+(\varphi )`$ and $`g(𝐱)`$ are even functions, $`S_{}`$ has the symmetry properties
$`S_{}(\varphi _1,\varphi _2;𝐱)=S_{}(\varphi _1,\varphi _2;𝐱)=S_{}(\varphi _2,\varphi _1;𝐱)`$ (48)
$`=S_{}(\varphi _1,\varphi _2;𝐱)=S_{}(\varphi _1,\varphi _2;𝐱).`$ (49)
These symmetries immediately imply
$`R_{++}(\varphi _1,\varphi _2;𝐱)=R_{++}(\varphi _1,\varphi _2;𝐱)=R_{++}(\varphi _2,\varphi _1;𝐱)`$ (50)
$`=R_{++}(\varphi _1,\varphi _2;𝐱)=R_{++}(\varphi _1,\varphi _2;𝐱).`$ (51)
The last terms in Eq. (44) and Eq. (46) are an “insertion of zero” since their contributions to $``$ cancel exactly. They represent the shift of a 1-point term from the flow of $`R_{++}(\varphi _1,\varphi _2;𝐱_{12})`$ to the flow of $`R_+(\varphi _1)`$ in order to achieve $`_𝐱R_{++}(\varphi _0,\varphi _0;𝐱)=0`$ according to the requirement specified above. This requirement has to be imposed on $`R_{++}`$ (unlike $`S_{}`$) in order to ensure that an $`ϵ`$ expansion can be performed consistently.
In order to obtain the roughness exponent to lowest order in $`ϵ`$,
$$\zeta =ϵ\zeta ^{(1)}+\text{h.o.}$$
(52)
(“h.o.” stands for higher orders in $`ϵ`$) we must determine the fixed-point Hamiltonian up to order $`ϵ^2`$,
$`_{\mathrm{pin}}=ϵ_{\mathrm{pin}}^{(1)}+ϵ^2_{\mathrm{pin}}^{(2)}+\text{h.o.}.`$ (53)
To lowest order there is only a 2-replica 1-point contribution
$`_{\mathrm{pin}}^{(1)}={\displaystyle \frac{1}{2T}}{\displaystyle \underset{\alpha \beta }{}}{\displaystyle _1}R_+^{(1)}(\varphi _1^{\alpha \beta }).`$ (54)
In order $`ϵ^2`$ the RG flow generates 2-point terms which are of 2-replica or 3-replica type:
$`_{\mathrm{pin}}^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{\alpha \beta }{}}{\displaystyle _1}R_+^{(2)}(\varphi _1^{\alpha \beta })`$ (57)
$`{\displaystyle \frac{1}{2!T}}{\displaystyle \underset{\alpha \beta }{}}{\displaystyle _{12}}R_{++}^{(2)}(\varphi _1^{\alpha \beta },\varphi _2^{\alpha \beta };𝐱_{12})`$
$`+{\displaystyle \frac{1}{3!T^2}}{\displaystyle \underset{\alpha \beta \gamma }{}}{\displaystyle _{12}}S_{}^{(2)}(\varphi _1^{\alpha \beta },\varphi _2^{\alpha \gamma };𝐱_{12}).`$
Inserting the Hamiltonian (54) and (57) into equation (34) we obtain the fixed-point equations for the parameter functions. We establish these equations and find their solutions in the order as they are generated by the RG.
The fixed-point condition $`\frac{d}{d\mathrm{}}S_{}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})=0`$ reads
$`0`$ $`=`$ $`\{2+𝐱_i\mathbf{}_i\}S_{}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})`$ (59)
$`3R_+^{(1)}(\varphi _1)R_+^{(1)}(\varphi _2)g_{12}.`$
It is solved by a function $`S_{}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})`$ in which the dependences on all arguments factorize:
$`S_{}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})=3R_+^{(1)}(\varphi _1)R_+^{(1)}(\varphi _2)\sigma (𝐱_{12}).`$ (60)
The explicit spatial dependence is carried by the function $`\sigma `$ which is determined by the differential equation
$`2\sigma (𝐱)+𝐱\mathbf{}\sigma (𝐱)`$ $`=`$ $`g(𝐱).`$ (61)
The solution of this differential equation (which we require to preserve the rotation symmetry in $`𝐱`$ of the unrenormalized model) involves one constant of integration, which can be fixed by the requirement that the Fourier transform $`\sigma (𝐤)`$ be analytic at $`𝐤=\mathrm{𝟎}`$. This is a common requirement, which is crucial also for the solution of the $`\varphi ^4`$ model, see Refs. . The Fourier transform of Eq. (61) is $`(2ϵ+𝐤\mathbf{}_𝐤)\sigma (𝐤)=g(𝐤)`$ which is solved by
$`\sigma (𝐤)={\displaystyle _0^1}𝑑zz^{1ϵ}g(z𝐤).`$ (62)
This solution is analytic at $`𝐤=\mathrm{𝟎}`$ for $`d>2`$ since $`g(𝐤)`$ is analytic. The explicit form of various appearing kernel functions is given in appendix B for the special cutoff function (27). Since $`R^{(1)}`$ is of order $`ϵ`$, $`\sigma `$ is needed only in order $`ϵ^0`$ to determine $`S_{}^{(2)}`$ from Eq. (60). In general, $`g(𝐤=\mathrm{𝟎})0`$ and then also $`\sigma (𝐤=\mathrm{𝟎})0`$. The $`𝐤=\mathrm{𝟎}`$ contribution to the functional $`S_{}^{(2)}`$ could in principle be split off as a 1-point term. We refrain from doing so because both the 1-point term and the 2-point term are of order $`ϵ^2`$ and this would unnecessarily increase the number of terms.
In order $`ϵ^2`$ the fixed-point conditions for $`R_+^{(2)}`$ and $`R_{++}^{(2)}`$ read
$`0`$ $`=`$ $`\{14\zeta ^{(1)}+\zeta ^{(1)}\varphi _1\}R_+^{(1)}(\varphi _1)+{\displaystyle \frac{2}{3}}{\displaystyle _2}_1_2S_{}^{(2)}(\varphi _1,0;𝐱_{12})g_{12}{\displaystyle \frac{1}{3}}{\displaystyle _2}_1_2S_{}^{(2)}(\varphi _1,\varphi _1;𝐱_{12})g_{12},`$ (64)
$`0`$ $`=`$ $`\{4+𝐱_i\mathbf{}_i\}R_{++}^{(2)}(\varphi _1,\varphi _2;𝐱_{12}){\displaystyle \frac{1}{3}}_1_2S_{}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})g_{12}+{\displaystyle \frac{1}{3}}\delta _{12}{\displaystyle _3}_1_2S_{}^{(2)}(\varphi _1,\varphi _1;𝐱_{13})g_{13}.`$ (65)
Plugging the solution (60) into the fixed-point equation (65), $`R_{++}^{(2)}`$ is found in the form
$`R_{++}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})=^2R_+^{(1)}(\varphi _1)^2R_+^{(1)}(\varphi _2)\rho (𝐱_{12})`$ (66)
with a nonlocal kernel $`\rho `$ satisfying the equation
$`4\rho (𝐱)+𝐱\mathbf{}\rho (𝐱)`$ $`=`$ $`c(𝐱)c_0\delta (𝐱),`$ (67)
where we defined
$`c(𝐱)`$ $`:=`$ $`\sigma (𝐱)g(𝐱),`$ (69)
$`c_0`$ $`:=`$ $`{\displaystyle _𝐱}c(𝐱)=c(𝐤=\mathrm{𝟎}).`$ (70)
The differential equation for $`\rho `$ is solved in Fourier space in analogy to Eq. (61) imposing the analyticity requirement at small wave vectors,
$`\rho (𝐤)={\displaystyle _0^1}𝑑zz^{1ϵ}[c_0c(z𝐤)].`$ (71)
Since $`g(𝐤)`$ is analytic, $`\sigma (𝐤)`$ and $`c(𝐤)`$ are analytic and so is $`\rho (𝐤)`$. It is important to notice that it is possible to find a solution which is analytic and finite for $`ϵ0`$ only due to the presence of the last term in equation (65). In the absence of this term, $`c_0`$ would be absent in Eq. (71) and $`\rho (𝐤=\mathrm{𝟎})1/ϵ`$ would become singular for $`d4^+`$. This would contradict our assumption that the 2-point kernel $`R_{++}`$ contributes to $``$ only in order $`ϵ^2`$; instead it would contribute to $``$ in order $`ϵ`$ and generate even more complicated contributions to $``$ in order $`ϵ^2`$.
Now we turn to the determination of the fixed-point function $`R_+^{(1)}`$ from equation (64). Inserting the given solution for $`S_{}^{(2)}`$ into this equation we find
$`0`$ $`=`$ $`\{14\zeta ^{(1)}+\zeta ^{(1)}\varphi \}R_+^{(1)}(\varphi )`$ (73)
$`+c_0[^2R_+^{(1)}(\varphi )^2R_+^{(1)}(\varphi )2^2R_+^{(1)}(\varphi )^2R_+^{(1)}(0)].`$
Remarkably, the entire information about the cutoff function $`f`$ (as well as the surface stiffness $`\gamma `$) is contained in the constant $`c_0`$. This constant can be eliminated by a rescaling $`RR/c_0`$, after which our flow equation coincides with the one derived with the HCRG scheme.
Since the fixed-point equation (73) is homogeneous, i.e. invariant under a simultaneous rescaling
$`\varphi \stackrel{~}{\varphi }:=b\varphi \text{ and }R_+(\varphi )\stackrel{~}{R}_+(\stackrel{~}{\varphi }):=b^4R_+(\varphi )`$ (74)
for any real $`b`$, it has a continuous set of solutions. Therefore one can search a solution of Eq. (73) (which is a nonlinear differential equation of second order for $`R_+^{(1)}`$) without loss of generality with the initial condition $`R_+^{(1)}(0)=1`$. The second condition $`R_+^{(1)}(0)=0`$ follows from the symmetry of $`R_+`$. The fixed point has to be determined numerically. For a trial value of $`\zeta ^{(1)}`$ one integrates Eq. (73) from $`\varphi =0`$ to $`\varphi =\mathrm{}`$. The fixed point describing elastic interfaces in disorder with short-ranged correlations is determined from the condition that $`R_+^{(1)}(\varphi )`$ decays monotonously and faster than any power. We find
$$\zeta ^{(1)}=0.208298(1)$$
(75)
and the fixed-point function and its derivatives are illustrated in Fig. 2. As discussed in Ref. , from this flow equation (73) one actually can determine not only the fixed point and scaling behavior for short-ranged correlations but also for long-ranged correlations which we will not further analyze here.
## VI Results and discussion
The RG result (52) with (75) is close to the Flory value $`\zeta _\mathrm{F}=\frac{ϵ}{5}`$. For $`d=1`$ the value $`\zeta (d=1)=\frac{2}{3}`$ of the roughness exponent is known exactly. In $`d=2`$ and $`d=3`$ numerical calculations resulted in the values $`\zeta (d=2)=0.41(1)`$ and $`\zeta (d=3)=0.22(1)`$. Although it is questionable whether the corresponding values of $`ϵ=3`$ are small enough to justify the neglection of higher order terms, the lowest order estimate (52) yields values $`\zeta (d=1)0.6249`$, $`\zeta (d=2)0.4166`$, and $`\zeta (d=3)0.2083`$, which are surprisingly close to the exact or numerical values.
Since $`f`$ enters the fixed-point equation only via $`c_0`$ which can be eliminated by rescaling, we have explicitly shown universality, i.e. that $`\zeta ^{(1)}`$ is independent of the cutoff function $`f`$. However, the fixed-point Hamiltonian does depend on this function. In particular, also the disorder cumulants are nonuniversal. We wish to stress again that in our analysis the nonlocality has been treated exactly, whereas Fisher and Balents and Fisher have truncated the flow to spatially constant fields which is sufficient to determine $`\zeta ^{(1)}`$. The nonlocality is important for the evaluation of the RG flow in higher orders of $`ϵ`$. However, only $`S_{}^{(2)}`$ feeds back into the flow equation for $`R_+^{(2)}`$ from which $`\zeta ^{(1)}`$ is determined. In contrast, $`R_{++}^{(2)}`$ with the proper definition of $`\rho `$ \[where $`\rho (𝐤=\mathrm{𝟎})=0`$ ensures that $`R_{++}`$ is of order $`ϵ^2`$\] does not feed back. Nevertheless, it is important to know $`R_{++}^{(2)}`$ in order to determine $`\zeta `$ in higher orders of $`ϵ`$.
The most distinctive feature of the fixed-point solution is the nonanalyticity of $`R_+^{(1)}`$, as pointed out by Fisher and Balents and Fisher and which is implied by the fixed-point equation (73) as follows. According to Eq. (2) a physically meaningful solution must satisfy $`R_+^{(1)}(0)>0`$ and $`R_+^{(1)}(\varphi )`$ can be expected to decay monotonously decaying for increasing $`\varphi `$. Searching a solution which is as smooth as possible, we may assume $`R_+^{(1)}(\varphi )`$ to be continuous and to vanish for $`\varphi =0`$. Evaluating Eq. (73) at $`\varphi =0`$ one then finds $`^2R_+^{(1)}(0)=[(14\zeta ^{(1)})R_+^{(1)}(0)/c_0]^{1/2}`$ which can have a solution only for $`\zeta ^{(1)}\frac{1}{4}`$. Then the second derivative of Eq. (73) implies
$`^3R_+^{(1)}(0^\pm )=\pm \sqrt{(12\zeta ^{(1)})[^2R_+^{(1)}(0)/2c_0]}.`$ (76)
The sign in front of the root is determined by the requirement that the amplitude of $`^2R_+^{(1)}(\varphi )`$ should have a maximum at $`\varphi =0`$. Thus $`^3R_+^{(1)}(0^\pm )0`$, i.e. $`R_+^{(1)}(\varphi )`$ is discontinuous at $`\varphi =0`$. Consequently, $`^4R_+^{(1)}(\varphi )`$ contains a singular contribution proportional to $`\delta (\varphi )`$.
Now we look back to verify that this nonanalyticity does not invalidate the RG analysis performed so far. Since at the fixed point $`S_{}^{(2)}`$ is proportional to $`R_+^{(1)}(\varphi )`$, discontinuities appear if $`S_{}^{(2)}`$ is derived twice with respect to one field argument. In the derivation of Eq. (44) some terms \[for example $`_𝐱_2^2S_{}(\varphi _1,0;𝐱)`$\] have been dropped because of the symmetry properties (49) of $`S_{}`$. This amounts to the implicit assumption $`^3R_+(0)=0`$, which seems arbitrary because of the discontinuity of $`^3R_+`$. However, in general there are additional odd factors since the energy functional has to be even in $`\varphi `$. In the aforementioned example $`_𝐱_2^2S_{}^{(2)}(\varphi _1,0;𝐱)R_+^{(1)}(\varphi _1)^3R_+^{(1)}(0)`$, i.e. there is an additional odd factor $`R_+^{(1)}(\varphi _1)`$. The corresponding contribution to the functional (57) then vanishes after summing over the replica indices irrespective of the value of $`^3R_+(0)`$. Therefore there is no need to retain this term.
The situation becomes more severe as soon as a fourth derivative $`^4R_+(\varphi )`$ appears. Since the fixed point function $`R_{++}^{(2)}`$ involves second derivatives of $`R_+^{(1)}`$, this can happen where second derivatives of $`R_{++}^{(2)}`$ with respect to one field argument appear. In the above flow equations (V) and the resulting fixed-point equations in order $`ϵ^2`$ this is not the case. However, in these equations we have neglected terms proportional to temperature $`T`$. In particular, we have neglected contributions to the flow of $`R_{++}^{(2)}`$, which are proportional to $`T_1^2R_{++}^{(2)}(\varphi _1,\varphi _2;𝐱_{12})T^4R_+^{(1)}(\varphi )`$. From a similar contribution to the flow equation of $`^2R_+^{(2)}`$ one expects $`^4R_+^{(1)}(0)T^1`$, where $`T`$ is the effective temperature that decreases on large scales. Thus these particular terms are proportional to $`T^0`$. This means that temperature is a dangerously irrelevant variable. (For a discussion of the role of temperature in $`d=1`$ see Refs. .) However, since $`R_{++}^{(2)}`$ does not feed back into the fixed-point equation for $`R_+^{(1)}`$, the roughness exponent is not affected in leading order. But these terms give contributions to the fluctuations of free energy (that scale for large system sizes $`L`$ proportional to $`L^\theta `$) which do not vanish in the limit $`T0^+`$.
To sum up, we have shown universality of the roughness exponent to leading order in $`ϵ`$ with the SCRG method elaborated here. Avoiding locality truncations, we have obtained the nonlocal functional form of the fixed-point Hamiltonian. In particular, we have determined higher cumulants of the effective pinning energy distribution on large scales the implications of which will be examined elsewhere.
Although in $`d=2`$ and $`d=3`$ the agreement between the RG result for $`\zeta `$ and numerical values is quite good, it is of fundamental interest to examine the possibility to extend the analytic theory beyond this leading order. It was argued by Fisher and Balents and Fisher that because of the nonanalyticity of the fixed point function $`R_+(\varphi )`$ in $`𝒪(ϵ)`$ the next higher order after $`𝒪(ϵ^2)`$ to the fixed-point equations should be $`𝒪(ϵ^{5/2})`$. Consequently, the subleading contribution to $`\zeta `$ would be of order $`ϵ^{3/2}`$. However, their reasoning is based on a $`T=0`$ argument. As shown above, temperature is not a truly irrelevant variable and the validity of a $`T=0`$ argument for $`T>0`$ is questionable. From preliminary studies we expect that the subleading contribution to $`\zeta `$ is of order $`ϵ^2`$. The SCRG scheme presented here lays the foundation for an extended analysis of the functional RG beyond leading order in a systematic way. In addition, this method is free of the pathologies hampering the HCRG scheme and better tractable than other SCRG schemes.
## Acknowledgments
The authors thank S. Bogner, T. Emig, T. Nattermann and H. Rieger for helpful discussions. S.S. is indebted to H. Wagner for stimulating suggestions already quite a long time ago. This work was supported financially by Deutsche Forschungsgemeinschaft through SFB 341.
## A Mode integration
In this appendix we sketch the derivation of Eq. (29) from Eq. (7) and deduce the RG generator (33) following Polchinski.
Starting from $`G_\mathrm{\Lambda }=G_<+G_>`$, elementary manipulations lead to the identity
$$G_\mathrm{\Lambda }^1=G_>^1G_>^1(G_<^1+G_>^1)^1G_>^1.$$
(A1)
We consider the partition sum of the pure interface as functional of the propagator $`G`$,
$`𝒵_{\mathrm{el}}[G]:=\mathrm{Tr}_\varphi e^{\frac{1}{2T}(\varphi ,G_\mathrm{\Lambda }^1\varphi )}.`$ (A2)
From $`G_\mathrm{\Lambda }=G_<+G_>`$ one can immediately derive
$`𝒵_{\mathrm{el}}[G_\mathrm{\Lambda }]={\displaystyle \frac{𝒵_{\mathrm{el}}[G_<]𝒵_{\mathrm{el}}[G_>]}{𝒵_{\mathrm{el}}[(G_<^1+G_>^1)^1]}}.`$ (A3)
Now the partition sum (7) is transformed by the following steps: we use Eq. (A1), introduce the additional field $`\varphi ^<`$, regroup fields in the bilinear Hamiltonian, use (A3), and substitute $`\varphi =\varphi ^<+\varphi ^>`$:
$`𝒵`$ $`=`$ $`\mathrm{Tr}_\varphi e^{\frac{1}{2T}(\varphi ,G_\mathrm{\Lambda }^1\varphi )\frac{1}{T}_{\mathrm{pin}}[\varphi ]}`$ (A4)
$`=`$ $`\mathrm{Tr}_\varphi e^{\frac{1}{2T}(\varphi ,G_>^1\varphi )+\frac{1}{2T}(G_>^1\varphi ,(G_<^1+G_>^1){}_{}{}^{1}G_{>}^{1}\varphi )\frac{1}{T}_{\mathrm{pin}}[\varphi ]}`$ (A5)
$`=`$ $`{\displaystyle \frac{1}{𝒵_{\mathrm{el}}[(G_<^1+G_>^1)^1]}}\mathrm{Tr}_{\varphi ^<}\mathrm{Tr}_\varphi e^{\frac{1}{2T}(\varphi ^<,(G_<^1+G_>{}_{}{}^{1})\varphi ^<)+\frac{1}{T}(\varphi ^<,G_>^1\varphi )\frac{1}{2T}(\varphi ,G_>{}_{}{}^{1}\varphi )\frac{1}{T}_{\mathrm{pin}}[\varphi ]}`$ (A6)
$`=`$ $`{\displaystyle \frac{𝒵_{\mathrm{el}}[G_\mathrm{\Lambda }]}{𝒵_{\mathrm{el}}[G_<]𝒵_{\mathrm{el}}[G_>]}}\mathrm{Tr}_{\varphi ^<}\mathrm{Tr}_\varphi e^{\frac{1}{2T}(\varphi ^<,G_<^1\varphi ^<)\frac{1}{2T}((\varphi \varphi ^<),G_>^1(\varphi \varphi ^<))\frac{1}{T}_{\mathrm{pin}}[\varphi ]}`$ (A7)
$`=`$ $`{\displaystyle \frac{𝒵_{\mathrm{el}}[G_\mathrm{\Lambda }]}{𝒵_{\mathrm{el}}[G_<]𝒵_{\mathrm{el}}[G_>]}}\mathrm{Tr}_{\varphi ^<}\mathrm{Tr}_{\varphi ^>}e^{\frac{1}{2T}(\varphi ^<,G_<^1\varphi ^<)\frac{1}{2T}(\varphi ^>,G_>^1\varphi ^>)\frac{1}{T}_{\mathrm{pin}}[\varphi ^<+\varphi ^>]}`$ (A8)
$`=:`$ $`{\displaystyle \frac{𝒵_{\mathrm{el}}[G_\mathrm{\Lambda }]}{𝒵_{\mathrm{el}}[G_<]}}\mathrm{Tr}_{\varphi ^<}e^{\frac{1}{2T}(\varphi ^<,G_<^1\varphi ^<)}e^{\frac{1}{T}_{\mathrm{pin}}[\varphi ^<+\varphi ^>]}_>`$ (A9)
This is Eq. (29) apart from the field independent factors $`𝒵_{\mathrm{el}}`$.
The RG generator (33) follows from evaluating the expectation value in the last expression of Eq. (A9). The coarse-grained pinning Hamiltonian $`_{\mathrm{pin}}^{}[\varphi ^<]`$ for the slow modes is defined by
$`e^{\frac{1}{T}_{\mathrm{pin}}^{}[\varphi ^<]}:=e^{\frac{1}{T}_{\mathrm{pin}}[\varphi ^<+\varphi ^>]}_>`$ (A10)
and is found from a cumulant expansion:
$`_{\mathrm{pin}}^{}[\varphi ^<]`$ $`=`$ $`_{\mathrm{pin}}[\varphi ^<+\varphi ^>]_>`$ (A12)
$`{\displaystyle \frac{1}{2T}}_{\mathrm{pin}}^2[\varphi ^<+\varphi ^>]_>^\mathrm{c}+\mathrm{}`$
A Taylor expansion of $`_{\mathrm{pin}}[\varphi ^<+\varphi ^>]`$ in $`\varphi ^>`$ yields ($`\varphi _i^>:=\varphi ^>(𝐱_i)`$ etc.)
$`_{\mathrm{pin}}[\varphi ^<+\varphi ^>]_>`$ $`=`$ $`_{\mathrm{pin}}[\varphi ^<]`$
$`+{\displaystyle \frac{T}{2}}{\displaystyle _{12}}{\displaystyle \frac{\delta ^2_{\mathrm{pin}}}{\delta \varphi _1^>\delta \varphi _2^>}}G_{12}^>+𝒪(G_>^2),`$
$`_{\mathrm{pin}}^2[\varphi ^<+\varphi ^>]_>^\mathrm{c}`$ $`=`$ $`T{\displaystyle _{12}}{\displaystyle \frac{\delta _{\mathrm{pin}}}{\delta \varphi _1^>}}{\displaystyle \frac{\delta _{\mathrm{pin}}}{\delta \varphi _2^>}}G_{12}^>+𝒪(G_>^2).`$
All neglected terms arising from higher cumulants of higher orders of Taylor expansion involve at least two propagators $`G_>`$. Since $`G_>`$ is bounded, terms of order $`G_>^2`$ result in functionals of order $`d\mathrm{}^2`$ for small $`d\mathrm{}`$ and do not contribute to the differential RG. Plugging the contributions of the last equations into (A12) result in the generator (33). For further details see also Ref. .
## B Kernels, coefficients
Here we give explicit expressions for the kernel functions and the coefficients entering the flow equations. This is interesting for illustrative purposes but also important to demonstrate, that they are well defined. We start from the Gaussian cutoff function (27) and perform calculations in $`d=4`$. We find from Eqs. (IV C), (62), (70), and (71)
$`g(𝐤)`$ $`=`$ $`{\displaystyle \frac{1}{\gamma \mathrm{\Lambda }^2}}e^{k^2/2\mathrm{\Lambda }^2}`$
$`\sigma (𝐤)`$ $`=`$ $`{\displaystyle \frac{1}{\gamma k^2}}\left(1e^{k^2/2\mathrm{\Lambda }^2}\right)`$
$`c(𝐤)`$ $`=`$ $`{\displaystyle \frac{g_0}{2}}\sigma (𝐤/\sqrt{2})`$
$`\rho (𝐤)`$ $`=`$ $`c_0{\displaystyle _0^1}𝑑t{\displaystyle \frac{1}{t}}\left\{1{\displaystyle \frac{4\mathrm{\Lambda }^2}{t^2k^2}}\left[1e^{t^2k^2/4\mathrm{\Lambda }^2}\right]\right\}`$
$``$ $`\{\begin{array}{cc}\frac{c_0k^2}{16\mathrm{\Lambda }^2}\hfill & \text{ for }k0,\hfill \\ c_0\mathrm{ln}\frac{k}{2\mathrm{\Lambda }}\hfill & \text{ for }k\mathrm{}.\hfill \end{array}`$
with $`g_0:=g(𝐱=\mathrm{𝟎})=\frac{\mathrm{\Lambda }^2}{4\pi ^2\gamma }`$ and $`c_0=\frac{1}{(4\pi \gamma )^2}`$. These kernels are apparently well-defined in $`d=4`$ and analytic for $`𝐤\mathrm{𝟎}`$.
|
warning/0006/hep-ph0006329.html
|
ar5iv
|
text
|
# Enhanced Electric Dipole Moment of the Muon in the Presence of Large Neutrino Mixing
## I Introduction
It has long been recognized that electric dipole moments (edm) of fermions can provide a unique window to probe into the nature of the forces that are responsible for CP violation. Experimental limits on the edm of neutron have reached the impressive level of $`6\times 10^{26}`$ ecm and have already helped constrain and sometimes exclude theoretical models of CP violation. Currently efforts are under way to improve this limit by at least two orders of magnitude, which will no doubt have very important implications for physics beyond the standard model. Electric dipole moment of the electron has severely been constrained by atomic measurements in $`Cs`$ ($`d_e^e10^{26}`$) and $`T\mathrm{}`$ ($`d_e^e4.3\times 10^{27}`$ ecm) . The limits on the muon edm on the other hand are much weaker, the present limit derived from the CERN $`(g2)`$ experiment is $`d_\mu ^e1.1\times 10^{18}`$ ecm. There has been a recent proposal to improve this limit on $`d_\mu ^e`$ to the level of $`10^{24}`$ ecm . In this paper we will argue that there is a strong motivation for this proposed improvement, related to the observation of neutrino masses and oscillations. We will show that a natural understanding of small neutrino masses with large oscillation angles in the framework of the seesaw mechanism will lead to an enhancement of $`d_\mu ^e`$, to values as large as $`5\times 10^{23}`$ ecm, which is well within the reach of the proposed experiment.
As for the theory of leptonic edm, in a large class of models a generic scaling law holds, given by $`d_\mu ^e/d_e^em_\mu /m_e`$. If such a relation is valid, even prior to any detailed calculation, one can infer that the present upper limit on electron edm will constrain the muon edm to be less than about $`10^{24}`$ ecm. This scaling law arises due to the chiral structure of the edm operator, which is very similar to the operator corresponding to the fermion mass. To the lowest order in the light fermion Yukawa couplings, the edm becomes proportional linearly to the fermion mass. In specific models, it may so happen that other constraints put the electron edm itself at a much lower value; e.g., the standard model prediction for the electron edm is $`10^{41}`$ ecm. The scaling law then suggests that the corresponding value for the muon edm would be at the level of $`10^{39}`$ ecm, which is beyond the reach of any conceivable experiment. In multi–Higgs doublet extensions of the standard model, the dominant contribution to the leptonic edm arises from a two–loop diagram involving $`\gamma V`$–Higgs vertex, where $`V=Z,W`$ . Since such a vertex is flavor universal, when converted to the fermion edm, the above–mentioned scaling law will hold. Recently an extended Higgs model has been analyzed, where it has been shown that for large values of the parameter $`\mathrm{tan}\beta `$ (ratio of the two Higgs vacuum expectation values), the one–loop diagram that scales as $`m_\mu \lambda _\mu ^2`$, where $`\lambda _\mu `$ is the muon Yukawa coupling, can compete with the two–loop diagram , leading to order one violation of the scaling law.
In the supersymmetric extension of the standard model (MSSM), under the usual assumptions about supersymmetry breaking terms, i.e., universality of scalar mass terms and proportionality of the trilinear $`A`$ terms with the corresponding Yukawa couplings, a similar scaling law would hold. A leading contribution to leptonic edm in such models is the one–loop diagram involving the bino virtual state and a complex $`A_{\mathrm{}\mathrm{}}`$ term. The assumption of proportionality of $`A`$ terms then implies that the above mentioned scaling relation remains. A similar remark holds when the chargino diagram is considered, with a complex $`\mu `$ term, again due to the universality of the CP violating parameter. (For a discussion of edm of electrons in MSSM and SUGRA models, see ref..) Evaluation of these bino and chargino diagrams leads to a value for the muon edm of about $`8\times 10^{25}`$ ecm, once the upper limit on electron and neutron edm are satisfied. The expected reach of a proposed BNL experiment for the muon edm is $`10^{24}`$ ecm, which is somewhat above the largest value allowed within the MSSM.
Recent experimental evidence for neutrino masses, especially from the SuperKamiokande atmospheric neutrino data , suggests that the MSSM must be extended to account for it. A natural place for small neutrino masses is the left–right symmetric extension of the standard model . We have recently advocated a simple supersymmetric realization of left–right symmetry (SUSYLR) which accommodates the neutrino masses via the seesaw mechanism. Our proposal is simply to embed the MSSM into a left–right symmetric gauge structure at a high scale $`v_R10^{11}10^{15}`$ GeV. The effective MSSM that emerges from this model at scales below the left-right symmetry breaking scale, $`v_R`$, is a constrained MSSM with far fewer number of phases. In particular, it has a built–in solution to the SUSY CP problem. Owing to the constraints of parity symmetry, the Yukawa coupling matrices and the trilinear $`A`$ matrices become hermitian in this model. Similarly, the $`\mu `$ term, the soft $`B\mu `$ parameter, and the gluino mass parameters all become real, eliminating potentially excessive CP violation from the MSSM. Furthermore, $`R`$–Parity arises automatically in this model as part of the gauge symmetry, since the gauge structure involves $`BL`$ symmetry.
In this paper we wish to investigate the CP violating muon edm $`d_\mu ^e`$ and $`(g2)_\mu `$ in this class of models. We will show that the interactions responsible for the Majorana masses of the right–handed neutrinos will lead to an enhancement of $`d_\mu ^e`$. We find $`d_\mu ^e`$ as large as $`5\times 10^{23}`$ ecm and $`(g2)_\mu `$ as much as $`13\times 10^{10}`$. These values arise even for small $`\mathrm{tan}\beta 3`$. Our main effect arises through the renormalization group extrapolation from the Planck scale to the left–right scale $`v_R`$ . In this interval the Yukawa couplings of the $`\nu _R`$ fields which induce their Majorana masses, as well as the associated trilinear $`A`$ terms, will affect the soft supersymmetry breaking parameters of the effective MSSM, leading to the enhancement of $`d_\mu ^e`$. Sine the Majorana Yukawa couplings do not obey $`e\mu `$ universality, the scaling law $`d_\mu ^e/d_e^e=m_\mu /m_e`$ is not obeyed by these new diagrams.
For concreteness, we will work within the framework of a minimal version of the high scale SUSYLR (or $`SO(10)`$) model. It is minimal in the sense that we have only one multiplet of Higgs field that gives rise to the usual Dirac fermion masses, i.e., one left-right bidoublet $`\mathrm{\Phi }`$ (10 in the case of SO(10)). With one such multiplet, only one Yukawa coupling matrix is allowed in the quark sector, leading to the proportionality of the up and the down Yukawa coupling matrices. We call this up–down unification. It has the consequence that all the flavor mixings vanish at the tree level. We have shown that acceptable values of the mixing angles can arise from the one–loop diagrams involving the gluino (and the chargino), proportional to the flavor structure of the trilinear $`A`$ terms. This considerably restricts the flavor and CP violating interactions in the model and makes it very predictive. The model has been shown to lead to a consistent picture of Kaon CP violation including $`ϵ`$ and $`ϵ^{}`$ and it predicts neutron edm at the level of $`10^{27}`$ ecm. The leptonic sector of the model was investigated in Ref., we shall work within that framework to calculate the edm and $`(g2)`$ of the muon. We have verified that going to non–minimal models, e.g., by employing more than one bidoublet Higgs field, does not affect our results by much in the leptonic sector.
## II Brief overview of the model
Let us briefly review the salient features of the minimal SUSYLR model. The electroweak gauge group of the model is $`SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ with the standard assignment of quarks and leptons – left–handed quarks and leptons ($`Q,L`$) transform as doublets of $`SU(2)_L`$, while the right–handed ones ($`Q^c,L^c`$) are doublets of $`SU(2)_R`$. The Dirac masses of fermions arise through their Yukawa couplings to a Higgs bidoublet $`\mathrm{\Phi }(2,2,0)`$. The $`SU(2)_R\times U(1)_{BL}`$ symmetry is broken to $`U(1)_Y`$ by $`BL=2`$ triplet scalar fields, the left triplet $`\mathrm{\Delta }`$ and right triplet $`\mathrm{\Delta }^c`$ (accompanied by $`\overline{\mathrm{\Delta }}`$ and $`\overline{\mathrm{\Delta }^c}`$ fields, their conjugates to cancel anomalies). These fields also couple to the leptons and are responsible for inducing large Majorana masses for the $`\nu _R`$. An alternative is to use $`BL=1`$ doublets $`\chi `$ (left) and $`\chi ^c`$ (right) along with $`\overline{\chi }`$ and $`\overline{\chi ^c}`$ instead of the $`\mathrm{\Delta }`$ fields. Here we shall adopt the $`BL=2`$ triplet option, which allows direct couplings to the leptons and which conserve $`R`$–Parity automatically. Let us write down the gauge invariant matter part of the superpotential involving these fields:
$`W`$ $`=`$ $`𝐘_qQ^T\tau _2\mathrm{\Phi }\tau _2Q^c+𝐘_lL^T\tau _2\mathrm{\Phi }\tau _2L^c`$ (1)
$`+`$ $`(𝐟L^Ti\tau _2\mathrm{\Delta }L+𝐟_cL_{}^{c}{}_{}{}^{T}i\tau _2\mathrm{\Delta }^cL^c).`$ (2)
Under left–right parity, $`QQ^c,LL^c,\mathrm{\Phi }\mathrm{\Phi }^{}`$, $`\mathrm{\Delta }\mathrm{\Delta }^c`$, along with $`W_{SU(2)_L}W_{SU(2)_R}^{}`$, $`W_{BL}W_{BL}^{}`$ and $`\theta \overline{\theta }`$. Here the transformations apply to the respective superfileds. As a consequence, $`𝐘_q=𝐘_q^{}`$, $`𝐘_l=𝐘_l^{}`$, and $`𝐟=𝐟_c^{}`$ in Eq. (1). Furthermore, the trilinear $`A_q`$ and $`A_l`$ terms will be hermitian, gluino mass term will be real, and the supersymmetric mass term for $`\mathrm{\Phi }`$ (the $`\mu `$–term) as well as the supersymmetry breaking $`B\mu `$ term will be real. Departures from these boundary conditions below $`v_R`$ due to the renormalization group extrapolation is small. The model thus provides a natural resolution to the supersymmetric CP problem.
Below $`v_R`$, the effective theory is the MSSM with the $`H_u`$ and $`H_d`$ Higgs multiplets. These are contained in the bidoublet $`\mathrm{\Phi }`$ of the SUSYLR model, but in general they can also reside partially in other multiplets having identical quantum numbers under the MSSM symmetry. Allowing for such a possibility, the single coupling matrix $`𝐘_q`$ of Eq. (1) describes the flavor mixing in the MSSM in both the up and the down sectors leading to the relations
$`𝐘_u=\gamma 𝐘_d,𝐘_{\mathrm{}}=\gamma 𝐘_{\nu ^D},`$ (3)
which we call up–down unification. Here $`\gamma `$ is a parameter characterizing how much of $`H_u`$ and $`H_d`$ of MSSM are in the bidoublet $`\mathrm{\Phi }`$. The case of $`H_{u,d}`$ entirely in $`\mathrm{\Phi }`$ will correspond to $`\gamma =1`$ and $`\mathrm{tan}\beta =m_t/m_b`$. At first sight the first of the relations in Eq.(2) might appear phenomenologically disastrous since it leads to vanishing quark mixings and unacceptable quark mass ratios. We showed in Ref. that including the one–loop diagrams involving the gluino and the chargino and allowing for a flavor structure for the $`A`$ terms, there exists a large range of parameters (though not the entire range possible in the usual MSSM) where correct quark mixings as well as masses can be obtained consistent with flavor changing constraints. In Ref., we explored the parameter space that allowed for arbitrary squark masses and mixings as well as arbitrary form for the supersymmetry breaking $`A`$ matrix. We found a class of solutions for large $`\mathrm{tan}\beta `$ $`3540`$ ($`\gamma `$ =1), and for small $`\mathrm{tan}\beta 4`$ where all quark masses mixings and CP violating phenomena could be explained. The smaller value $`\mathrm{tan}\beta `$ requires larger values of $`\gamma `$, since $`\gamma \mathrm{tan}\beta =m_t/m_b`$ is fixed. In this paper, we use small $`\mathrm{tan}\beta `$ scenarios which is less constrained.
Since the parameter $`\gamma `$ plays a crucial role in determining the value of $`\mathrm{tan}\beta `$, let us explain its origin in an explicit high scale model. We will also show how the solution to the SUSY CP problem can be maintained even for the case of small $`\mathrm{tan}\beta `$. $`\gamma `$ arises from the mixing of the bidoublet $`\mathrm{\Phi }`$ with other weak doublets in the high scale theory. We assume that only one pair of doublets, $`H_u`$ and $`H_d`$ of MSSM, remain light below $`v_R`$. A concrete example which also maintains automatic $`R`$–Parity of the left–right model involves the addition of the following new fields: $`\rho (2,2,2)+\overline{\rho }(2,2,2)`$ and $`\mathrm{\Omega }_L(3,1,0)+\mathrm{\Omega }_R(1,3,0)`$. They lead to the following new terms in the superpotential:
$`W_{\mathrm{new}}`$ $`=`$ $`\mu _\mathrm{\Delta }(\mathrm{\Delta }\overline{\mathrm{\Delta }}+\mathrm{\Delta }^c\overline{\mathrm{\Delta }^c})+\mu _\mathrm{\Phi }\mathrm{\Phi }^2+\mu _\rho \overline{\rho }\rho +\mu _\mathrm{\Omega }(\mathrm{\Omega }_L^2+\mathrm{\Omega }_R^2)`$ (4)
$`+`$ $`\lambda _1\left[\mathrm{Tr}(\rho \mathrm{\Delta }^c\mathrm{\Phi })+\mathrm{Tr}(\overline{\rho }\mathrm{\Delta }\mathrm{\Phi })\right]+\lambda _2\left[\mathrm{Tr}(\overline{\rho }\overline{\mathrm{\Delta }}^c\mathrm{\Phi })+\mathrm{Tr}(\rho \overline{\mathrm{\Delta }}\mathrm{\Phi })\right]`$ (5)
$`+`$ $`\lambda _3\mathrm{Tr}(\overline{\mathrm{\Delta }}^c\mathrm{\Omega }_R\mathrm{\Delta }^c+\overline{\mathrm{\Delta }}\mathrm{\Omega }_L\mathrm{\Delta })+\lambda _4\mathrm{Tr}(\rho \mathrm{\Omega }_R\overline{\rho }+\overline{\rho }\mathrm{\Omega }_L\rho ).`$ (6)
The coupling and the mass parameters in Eq. (3) are guaranteed to be real by parity symmetry, $`P`$, defined earlier in combination with the charge conjugation symmetry $`C`$ under which all superfields (except $`\rho `$ and $`\mathrm{\Phi }`$) transform as $`\mathrm{\Psi }\mathrm{\Psi }^c`$, where $`\mathrm{\Psi }`$ stands for a relevant superfield in the theory; the $`W_LW_R`$ and $`BB`$. The fields $`\rho `$ and $`\mathrm{\Phi }`$ transform as follows: $`\rho \tau _2\overline{\rho }^T\tau _2`$ and $`\mathrm{\Phi }\tau _2\mathrm{\Phi }^T\tau _2`$. We will assume that the supersymmetry breaking terms respect only $`P`$ and not $`C`$.
It can be shown (see e.g., Ref. ) that this model has a ground state where $`\mathrm{\Omega }_R\mathrm{\Delta }^c=\overline{\mathrm{\Delta }^c}v_R`$ and $`\mathrm{\Omega }_L=\mathrm{\Delta }=\overline{\mathrm{\Delta }}=0`$. The $`\rho `$ superfield contains an $`H_u`$–like MSSM doublet and $`\overline{\rho }`$ contains an $`H_d`$–like one. Once the right handed gauge symmetry is broken by $`\mathrm{\Delta }^c`$ vev, the doublets in $`\mathrm{\Phi }`$ and those in $`\rho `$ and $`\overline{\rho }`$ mix via a matrix, which is given by $`W_{\mathrm{mass}}=\left(\begin{array}{cc}\rho _u& \mathrm{\Phi }_u\end{array}\right)M_{\mathrm{doublet}}\left(\begin{array}{c}\rho _d\\ \mathrm{\Phi }_d\end{array}\right)`$, where
$`M_{\mathrm{doublet}}`$ $`=`$ $`\left(\begin{array}{cc}\mu _\rho & \lambda _1v_R\\ \lambda _2v_R& \mu _\mathrm{\Phi }\end{array}\right).`$ (7)
$`M_{\mathrm{doublet}}`$ being an asymmetric matrix leads to light eigenstates given by $`H_u=\mathrm{cos}\theta _1\mathrm{\Phi }_u+\mathrm{sin}\theta _1\rho _u`$ and $`H_d=\mathrm{cos}\theta _2\mathrm{\Phi }_d+\mathrm{sin}\theta _2\rho _d`$. Here $`\theta _1`$ is the $`\rho _u\mathrm{\Phi }_u`$ mixing angle, which is unrelated (due to the asymmetry of the matrix) to $`\theta _2`$, the $`\rho _d\mathrm{\Phi }_d`$ mixing angle. This gives $`\gamma =\frac{\mathrm{cos}\theta _1}{\mathrm{cos}\theta _2}`$, which can take any arbitrary value.
We note that due to the combination of $`P`$ and softly broken $`C`$ symmetry, all dimension four couplings are real, leading to a solution to the SUSY CP problem. To see this, note that due to these symmetries, all entries in the mass matrix of Eq. (4) are real, so that the effective $`\mu `$ term of MSSM stays real. (With parity symmetry alone, the $`\lambda _{1,2}`$ couplings in Eq. (4) could be complex, which would make the effective $`\mu `$ term of the MSSM complex.) Furthermore, since only the dimension 3 and 2 terms of the SUSY breaking Lagrangian are assumed to respect P, but not $`C`$, such a scenario is completely stable under renormalization. (This scheme is distinct from scenarios where CP symmetry is imposed on the MSSM Lagrangian at a high scale to solve the SUSY CP problem . Since the gauge structure of MSSM does not have parity symmetry, the phases of the soft SUSY breaking terms will have to be small in that case.)
Unlike the large $`\mathrm{tan}\beta `$ case (corresponding to $`\gamma =1`$), we are finding that CP violation in the quark sector has to arise from soft terms. We have analyzed this possibility in Ref. and shown its consistency. We are pursuing this possibility further . An immediate outcome of this scenario for hadronic CP violation is that although there is KM type CP violation, generically it tends to be sub–leading to SUSY CP violation.
In the absence of the $`\mathrm{\Omega }_R`$ field in Eq. (3), the doubly charged field $`\mathrm{\Delta }^{c++}`$ in $`\mathrm{\Delta }^c`$ (as well as $`\mathrm{\Delta }^c`$ in $`\overline{\mathrm{\Delta }^c}`$) will remain massless – it will pick up mass only of order the weak scale, or of order $`v_R^2/M_{\mathrm{string}}`$, if non–renormalizable operators are included. Inclusion of $`\mathrm{\Omega }_R`$ (and its left–handed partner $`\mathrm{\Omega }_L`$) lifts the mass of $`\mathrm{\Delta }^{c++}`$ to the scale $`v_R`$ . We will analyze two cases, one with the inclusion of $`\mathrm{\Omega }_{L,R}`$ fields, and one without. In the latter case, we will take the mass of $`\mathrm{\Delta }^{c++}`$ to be $`v_R^2/M_{\mathrm{string}}`$.
## III Leptonic CP violation and muon EDM
To discuss CP violation in the lepton sector, we need to specify the leptonic superpotential $`W_{\mathrm{}}`$ and the most general soft breaking Lagrangian, $`_{\mathrm{soft}}^{\mathrm{}}`$, in the lepton superpartners. The leptonic $`W_{\mathrm{}}`$ is given in Eq. (1), $`_{\mathrm{soft}}^{\mathrm{}}`$ is given by:
$$_{\mathrm{soft}}^{\mathrm{}}=𝐦_{LL}^2\stackrel{~}{L}^{}\stackrel{~}{L}+𝐦_{RR}^2\stackrel{~}{L^c}^{}\stackrel{~}{L^c}+[A_l\stackrel{~}{L}\mathrm{\Phi }\stackrel{~}{L^c}+A_f(\stackrel{~}{L}\stackrel{~}{L}\mathrm{\Delta }+\stackrel{~}{L^c}\stackrel{~}{L^c}\mathrm{\Delta }^c)+H.c.]$$
(8)
To generate a nonvanishing muon edm, one needs a complex valued $`(A_l)_{22}`$ and/or complex soft mass-squared terms. But above the scale where the parity symmetry is valid, $`A_l`$ is hermitian and therefore its diagonal elements are all real. This element can however be complex due to radiative corrections below the parity breaking scale. There are two ways this can happen: (i) if only parity symmetry is broken but gauge symmetry $`SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ is unbroken at the string scale by introduction of parity odd singlets; (ii) if both parity and the left-right gauge symmetry are broken, but some remnant of the $`𝐟`$ and $`A_f`$ couplings remain below the $`v_R`$ scale. This has been shown to happen in supersymmetric left-right models with minimal field content. In the explicit version described in Sec. II, if the $`\mathrm{\Omega }_{L,R}`$ fields are absent, the $`\mathrm{\Delta }^{c++}`$ field from $`\mathrm{\Delta }^c`$ will have a mass of order $`v_R^2/M_{\mathrm{string}}10^{12}`$ GeV. So between $`M_{\mathrm{string}}`$ and $`M_{\mathrm{\Delta }^{c++}}`$, the effects of $`𝐟`$ and $`A_f`$ couplings will be felt, and $`(A_l)_{22}`$ can become complex. This will also induce flavor violating complex soft mass–squared terms proportional to $`A_fA_f^{}`$, even if we start with diagonal soft masses at $`M_{\mathrm{string}}`$.
In case (i), the way $`(A_l)_{ij}`$ become complex is as follows. Below the D-parity (discrete parity) breaking scale, $`M_{\mathrm{string}}`$, only $`\mathrm{\Delta }^c`$’s (and not $`\mathrm{\Delta }`$’s) contribute to renormalization group equations (RGE) describing the evolution of $`(A_l)_{ij}`$ since the $`\mathrm{\Delta }`$’s acquire masses of order $`M_{\mathrm{string}}`$. The RGE are given in the Appendix for this case. We have, from Eq. (25) of Appendix,
$`{\displaystyle \frac{dA_l}{dt}}`$ $``$ $`{\displaystyle \frac{1}{16\pi ^2}}(4A_f𝐟^{}𝐘_l+2\mathrm{𝐟𝐟}^{}A_l).`$ (9)
The first term on the RHS of Eq. (6) will introduce phase in $`A_l`$. Note that $`A_f`$ is not constrained to be hermitian at the string scale by parity symmetry (unlike $`A_l`$, which must be hermitian at $`M_{\mathrm{string}}`$). We will allow for complex entries in the $`23`$ block of $`A_f`$ in our analysis.
Below the D–parity breaking scale, the soft mass parameters $`𝐦_{LL}^2`$ and $`𝐦_{RR}^2`$ will evolve differently. In particular, $`𝐦_{RR}^2`$ will feel the effects of $`𝐟`$ and $`A_f`$ couplings. In order to explain the large oscillation angle needed for the atmospheric neutrino data, we will find that $`f_{23}`$ is not much smaller than $`f_{33}`$. Thus $`(A_f)_{23}`$ is not much smaller than $`(A_f)_{33}`$. Consequently, $`(𝐦_{RR}^2)_{23}`$ will become large and complex. This is the main source of the enhanced edm of the muon in the model. This qualitative feature becomes more transparent if we examine the RGE for $`𝐦_{RR}^2`$ (see Eq. (27) of Appendix). It has the form:
$`{\displaystyle \frac{d𝐦_{RR}^2}{dt}}`$ $``$ $`{\displaystyle \frac{1}{16\pi ^2}}(2A_fA_f^{}).`$ (10)
It is clear from Eq.(7) how $`(𝐦_{RR}^2)_{23}`$ becomes large and complex.
The dominant contribution to the edm of muon arises from a diagram which has right and left–handed muon in the external legs and a lighter stau inside the loop. It utilizes the above–mentioned $`23`$ mixing which is large and complex. For example, the diagram can have $`\mu _L\stackrel{~}{\tau }_R`$ and $`\stackrel{~}{\tau }_L\mu _R`$ vertices along with the stau mass flip inside the loop or it can involve just the $`\mu _L\stackrel{~}{\tau }_R`$ and $`\stackrel{~}{\tau }_R\mu _R`$ vertices. It might be suspected that similar diagrams will also induce large edm for the electron. However, in this model, since $`f_{13}`$ and $`f_{12}`$ are much smaller, such contributions are negligible. Essentially, we have a scenario where $`e\mu `$ flavor symmetry is broken by a large amount by the $`𝐟`$ and $`A_f`$ terms. As a result the scaling law alluded to in the introduction does not hold. If we assume, as we do in our analysis, the existence of phases only in the $`23`$ block of the $`A_f`$ matrix, or if $`A_f`$ has negligible entries in its first row and column, no appreciable edm for the electron gets induced due to mixing effect. Below $`v_R`$, we have only the MSSM field content. Due to the new $`𝐟`$ couplings above the $`v_R`$ scale, the $`\stackrel{~}{\tau }_1`$ mass is lower than usual SUGRA model for the same values of the parameter space (i.e., $`m_0`$, $`m_{1/2}`$, $`A_0`$, $`\mathrm{tan}\beta `$). This is why the diagram involving the $`\stackrel{~}{\tau }`$ tends to dominate in $`d_\mu ^e`$.
In case (ii), we use the fact that in the minimal SUSYLR model (without $`\mathrm{\Omega }_{L,R}`$), $`\mathrm{\Delta }^{c++}`$ and $`\mathrm{\Delta }^c`$ remain below the $`v_R`$ scale; therefore their couplings to the charged fermions via RGE’s lead to imaginary parts in $`(A_l)_{22}`$ by an amount $`\frac{(𝐟A_f^{}𝐘_l)_{22}}{16\pi ^2}`$. Again the soft masses become complex in the same fashion as in case (i). These fields get decoupled at somewhat lower scale $`10^{11}`$ GeV, below which the spectrum is that of MSSM.
## IV Results
Let us first discuss the neutrino mass fits in this model. We start with a basis where the charged leptons masses are diagonal and Dirac neutrino masses are given by
$$M_{\nu ^D}=\gamma \mathrm{tan}\beta M_l,$$
(11)
where $`M_l=Diag(m_e,m_\mu ,m_\tau )`$. The light Majorana neutrino mass matrix is then given by:
$$M_\nu =\frac{\gamma ^2\mathrm{tan}^2\beta }{v_R}M_l𝐟^\mathrm{𝟏}M_l,$$
(12)
where $`𝐟`$ is the right–handed Majorana Yukawa coupling matrix.
In our fit, we first use the small angle MSW oscillations solution for the solar neutrino deficit with $`\mathrm{\Delta }m_{e\mu }^2(0.31)\times 10^5`$ eV<sup>2</sup> and $`2\times 10^3\mathrm{sin}^22\theta _{e\mu }2\times 10^3`$. We also use the $`\nu _\mu \nu _\tau `$ oscillation scenario to explain the observed deficit in the flux of muon neutrinos from the atmosphere . The mass splitting is taken to be $`\mathrm{\Delta }m_{\mu \tau }^2(0.11)\times 10^2`$ eV<sup>2</sup> and the oscillation angle to be $`\mathrm{sin}^22\theta _{\mu \tau }0.81`$.
For tan$`\beta `$ = 3, we find a good fit to the solar and atmospheric neutrino data by choosing $`𝐟`$ at $`M_{\mathrm{string}}`$ to be
$`𝐟`$ $`=`$ $`\left(\begin{array}{ccc}1.00\times 10^4& 8.8\times 10^4& 2.2\times 10^5\\ 8.8\times 10^4& 1.3\times 10^2& 1.03\times 10^1\\ 2.4\times 10^5& 1.03\times 10^1& 1.59\end{array}\right).`$ (13)
The resulting neutrino masses at $`v_R=10^{15.3}`$ GeV are: $`(6.27\times 10^6,\mathrm{\hspace{0.17em}2.5}\times 10^3,\mathrm{\hspace{0.17em}5.2}\times 10^2)`$ eV. The leptonic mixing matrix is given by:
$`U`$ $`=`$ $`\left(\begin{array}{ccc}0.99& 4.2\times 10^2& 8.4\times 10^5\\ 3.1\times 10^2& 0.74& 0.67\\ 2.9\times 10^2& 0.71& 0.71\end{array}\right).`$ (14)
$`U_{21}`$ is the mixing angle relevant for solar neutrino oscillations. (Our notation is such that $`UM_\nu U^T=M_\nu ^{\mathrm{diagonal}}`$.) This choice leads to a simultaneous explanation of the solar and atmospheric neutrino anomalies. Note that we have taken all Yukawa couplings to be real, consistent with our assumption that $`C`$ and $`P`$ symmetry are respected by $`d=4`$ terms.
It is possible to fit the large angle oscillations solution to satisfy the solar neutrino deficit. In that case we take $`𝐟`$ matrix is at $`v_R10^{15.6}`$ GeV to be
$`𝐟`$ $`=`$ $`\left(\begin{array}{ccc}1.77\times 10^7& 1.42\times 10^6& 0\\ 1.42\times 10^6& 3.9\times 10^3& 6.4\times 10^2\\ 0& 6.4\times 10^2& 1.28\end{array}\right).`$ (15)
With these values the neutrino masses are $`(1.7\times 10^3,\mathrm{\hspace{0.17em}2.0}\times 10^3,\mathrm{\hspace{0.17em}3.4}\times 10^2)`$ eV and the corresponding leptonic mixing matrix is:
$`U`$ $`=`$ $`\left(\begin{array}{ccc}0.89& 3.3\times 10^1& 4.0\times 10^1\\ 4.5\times 10^1& 0.63& 0.63\\ 2.8\times 10^2& 0.72& 0.69\end{array}\right).`$ (16)
We use the one–loop Yukawa and two–loop gauge RGE to extrapolate all parameters between the string scale and the $`v_R`$ scale. Since the new couplings $`𝐟`$ affect the RGE for the leptonic Yukawas, one needs to make sure that the charged lepton masses come out to be correct at the weak scale. For simplicity we choose a universal scenario, i.e., all the scalar masses are given by a common mass parameter $`m_0`$ at the string scale. We also assume a common trilinear mass $`A_0`$($`\times 𝐘_l`$) for all generations. For $`A_f`$ we use a structure similar to $`𝐟`$. But we do not impose $`A_f𝐟`$. We demand electroweak symmetry to be broken radiatively. In case (i), where parity is broken at $`M_{\mathrm{string}}`$, $`\mathrm{\Delta }`$ fields get decoupled and only the $`\mathrm{\Delta }^c`$ fields contribute to the RGE for soft masses. Consequently the renormalized right handed slepton masses get lowered due to the presence of the new couplings $`𝐟`$. Furthermore, $`A_l`$ will pick up off–diagonal elements and will lose its hermitian structure through renormalization. The $`\mathrm{\Delta }^c`$ fields get decoupled at the left-right breaking scale $`v_R`$, below which we use the RGE corresponding to the MSSM degrees.
The EDM for a spin $`1/2`$ fermion is given by the following effective Lagrangian:
$$L_f=\frac{1}{2}d_f\overline{\psi }\sigma _{\mu \nu }\gamma _5\psi F^{\mu \nu }.$$
(17)
In this model, we have only the neutralino-slepton loop contribution to the edm of muon. This contribution is given as :
$$d_\mu ^e/e=\frac{\alpha _{em}}{4\pi sin^2\theta _w}\underset{i=1}{\overset{6}{}}\underset{i=1}{\overset{4}{}}Im(\eta _{\mu ik})\frac{\stackrel{~}{m}_{\chi _i^0}}{\stackrel{~}{m}_k^2}Q_\mu B(\frac{\stackrel{~}{m}_{\chi _i^0}^{}{}_{}{}^{2}}{\stackrel{~}{m}_k^2})$$
(18)
where $`\eta _{\mu ik}=[\sqrt{2}(\mathrm{tan}\theta _W(Q_\mu T_{3\mu })X_{1i}+T_{3\mu }X_{2i})\mathrm{\Gamma }_{L2k}^{}+x_\mu X_{3i}\mathrm{\Gamma }_{R2k}^{}](\sqrt{2}\mathrm{tan}\theta _WQ_\mu X_{1i}\mathrm{\Gamma }_{R2k}x_\mu X_{3i}\mathrm{\Gamma }_{L2k}^{})`$ and $`x_\mu =\frac{m_\mu }{\sqrt{2}m_W\mathrm{cos}\beta }`$. $`X`$ diagonalizes the neutralino mass matrix, $`X^TM_{\chi ^0}X=diag(m_{\chi _1^0},m_{\chi _2^0},m_{\chi _3^0},m_{\chi _1^04})`$. Here $`\mathrm{\Gamma }_{L,R}`$ are $`6\times 3`$ matrices given by $`\stackrel{~}{q}_L=\mathrm{\Gamma }_{L,R}\stackrel{~}{q}`$ and $`B(r)=\frac{1}{2(r1)^2}(1+r+2r\mathrm{ln}r/(1+r))`$.
We first analyzed the case where $`A_f`$ and $`𝐟`$ are proportional. It still allows for an overall phase in $`A_f`$, consistent with $`P`$ invariance. In this case $`d_\mu ^e`$ is highly suppressed, $`d_\mu ^e10^{26}`$ ecm. The reason is that with only one matrix structure $`𝐟`$, when the effective $`(A_l)_{22}`$ is computed in the original gauge bases, it will remain real. Small contribution will arise in the mixed $`\mu \tau `$ EDM operator, which can lead to a small value of $`d_\mu ^e`$ since the physical $`\mu `$ is a linear combination of the two states. However, this $`\mu \tau `$ mixing turns out to be small. As soon as the proportionality $`A_f𝐟`$ is relaxed, $`d_\mu ^e`$ becomes much larger. We have analyzed the case where $`A_f`$ and $`𝐟`$ are non–proportional, but the magnitudes $`|A_f|_{ij}`$ are proportional to $`|𝐟|_{ij}`$. We allow phases of order 1 in the (23), (32) and (33) elements of $`A_f`$ matrix, while keeping $`f_{ij}`$ real. In this case we find the maximum muon edm to be $`7\times 10^{25}`$ ecm. When this assumption of proportionality of the magnitudes is relaxed, even larger value of $`d_\mu ^e`$ results. We give an explicit example for this case below. It should be mentioned that large values of $`A_f`$ reduces stau mass while it increases $`d_\mu ^e`$. So in exploring regions of large $`d_\mu ^e`$, we need to consider the experimental limits on stau. In our calculation we take the lightest stau mass ($`\stackrel{~}{\tau }_1`$) to be $`80`$ GeV (which is above the current experimental limit of 70 GeV at $`\sqrt{s}=202`$ GeV). In Fig. 1, we exhibit the case which has small angle oscillation solution. The large angle solution, however, does not show any difference. In Fig. 1 we plot the muon edm parameter $`k_\mu Log_{10}[\frac{d_\mu ^e}{1\times 10^{23}\mathrm{ecm}}]`$ for case (i) for tan$`\beta =3`$. This corresponds to $`D`$–parity broken at the string scale, but left–right gauge symmetry broken at $`v_R10^{15.3}`$ GeV. At the string scale (taken to be $`10^{17}`$ GeV), we have assumed (in GeV units throughout)
$`A_f`$ $`=`$ $`\left(\begin{array}{ccc}2\times 10^3& 1\times 10^2& 0\\ 1.0\times 10^2& 1\times 10^2e^{i\pi /2}& 4.7\times 10^2e^{i\pi /2}\\ 0& 4.7\times 10^2e^{i\pi /2}& 3.3\times 10^2e^{i\pi /2}\end{array}\right).`$ (19)
We put $`A_0=120`$ GeV (with $`A_l=A_0𝐘_l`$). The solid line in Fig. 1 is drawn for $`m_0`$=160 GeV. The extreme left corner of the curve corresponds to lighter stau mass ($`\stackrel{~}{\tau }_1`$)=82 GeV. At the same spot in the parameter space, the lightest chargino ($`\chi _1^\pm `$) and the lightest neutralino masses ($`\chi _1^0`$) are 106 GeV and 52 GeV respectively. We can see that the muon edm can be as large $`3\times 10^{23}`$ ecm in this case. The dotted line is drawn for $`m_0`$=170 GeV for the same set of input values.
In Fig. 2 we plot the muon edm parameter $`k_\mu `$, for case (ii) with tan$`\beta =3`$ and $`m_0=160`$ GeV. This case corresponds to $`\mathrm{\Delta }^{c++}`$ surviving below $`v_R`$. We assume the scale at which it decouples to be $`10^{12}`$ GeV. We have used the universal scenario for the slepton masses and have used the same $`𝐟`$ matrix as before. At the string scale, we take (in GeV)
$`A_f`$ $`=`$ $`\left(\begin{array}{ccc}2\times 10^3& 1\times 10^2& 0\\ 1.0\times 10^2& 1\times 10^1e^{i\pi /2}& 3.0\times 10^2e^{i\pi /2}\\ 0& 3.0\times 10^2e^{i\pi /2}& 1.1\times 10^2e^{i\pi /2}\end{array}\right).`$ (20)
We take $`A_0=0`$ GeV. The extreme left corner of the curve in Fig. 2 corresponds to lighter stau mass ($`\stackrel{~}{\tau }_1`$) mass of 80 GeV. At the same spot, as before, the $`\chi _1^\pm `$ and the $`\chi _1^0`$ masses ($`\chi _1^0`$) are 106 GeV and 52 GeV respectively. As can be seen from the figure, large values of $`d_\mu ^e`$ are possible, as large as $`5\times 10^{23}`$ ecm.
We have assumed non–proportionality of $`A_f`$ and $`𝐟`$ in the preceding two examples. We will argue that this is not unnatural. First of all, there are no strong experimental hints that suggest proportionality of the two (unlike the case of $`A_l`$ and $`𝐘_l`$). Second, we have proposed recently a model based on horizontal gauge symmetry which allows for all parameters of the soft breaking sector to be arbitrary, subject only to the constraints of the horizontal symmetry $`H`$ . The symmetry $`H`$ was taken to be $`SU(2)_H\times U(1)_H`$, with the first two generations of fermions falling into $`SU(2)_H`$ doublets and the thrid generation into singlets. The first two generations have $`U(1)_H`$ charges of $`1`$, while the third generation is neutral. $`H`$ is spontaneously broken by a pair of doublet \[$`\varphi (+1),\overline{\varphi }(1)`$\] and singlet \[$`\chi (+1),\overline{\chi }(1)`$\] scalars fields whose vev’s are below the string scale. We denote $`ϵ_\varphi \varphi /M_{\mathrm{string}}`$, $`ϵ_\chi \chi /M_{\mathrm{string}}`$ with $`ϵ_\varphi 1/7,ϵ_\chi 1/25`$. The effective Yukawa couplings involving the light two generations will be proportional to powers of $`ϵ_\varphi `$ and $`ϵ_\chi `$. The $`U(1)_H`$ also alleviates potential problems with $`D`$–terms associated with horizontal symmetries.
Within the $`SU(2)_H\times (U1)_H`$ model, it is not necessary to assume universality of scalar masses or proportionality of $`A`$ terms and the Yukawa couplings. For the first two generations, the scalar masses will be approximately equal, owing to the non–Abelian sector of the horizontal symmetry. With the horizontal charge assignment given above, we can write down the most general $`H`$–symmetric Yukawa couplings, soft mass terms and $`A`$ terms. Since the $`A`$ terms become hierarchical, all FCNC constraints can be satisfied, even without proportionality assumption .
We will now present an example for the muon edm within this horizontal symmetric framework. We will embed the model of Ref. into left–right symmetry at a high scale. Unlike in Ref. , all the CKM mixing will vanish at tree–level now. In a basis where the Yukawa couplings are diagonalized, the Majorana neutrino coupling can be written in the following hierarchical form:
$`𝐟`$ $`=`$ $`\left(\begin{array}{ccc}f_{11}ϵ_\chi ^4/ϵ_\varphi ^2& f_{12}ϵ_\chi & f_{13}ϵ_\chi ^2/ϵ_\varphi \\ f_{12}ϵ_\chi & f_{22}ϵ_\varphi ^2& f_{23}ϵ_\varphi \\ f_{13}ϵ_\chi ^2/ϵ_\varphi & f_{23}ϵ_\varphi & f_{33}\end{array}\right).`$ (21)
The bilinear soft mass matrix and the A matrix are given as:
$`m_{RR}^2`$ $`=`$ $`\left(\begin{array}{ccc}m_0^2& m_0^2(x_{12})ϵ_\chi & m_0^2(x_{13})ϵ_\chi ^2/ϵ_\varphi \\ m_0^2(x_{12})ϵ_\chi & m_0^2& m_0^2(x_{23})ϵ_\varphi \\ m_0^2(x_{13})ϵ_\chi ^2/ϵ_\varphi & m_0^2(x_{23})ϵ_\varphi & m_{33}^2\end{array}\right);`$ (22)
$`A`$ $`=`$ $`A_0\left(\begin{array}{ccc}(y_{11})ϵ_\chi ^4/ϵ_\varphi ^2& (y_{12})ϵ_\chi & (y_{13})ϵ_\chi ^2/ϵ_\varphi \\ (y_{21})ϵ_\chi & (y_{22})ϵ_\varphi ^2& (y_{23})ϵ_\varphi \\ (y_{31})ϵ_\chi ^2/ϵ_\varphi & (y_{32})ϵ_\varphi & y_{33}\end{array}\right).`$ (24)
We also have $`m_{LL}^2=m_{RR}^2`$. This structure for $`A`$ hold for both $`A_l`$ and $`A_f`$ (as well as for $`A_q`$). At $`M_{\mathrm{string}}`$ we will take $`A_l`$ to be hermitian. In order to fit the experimental values of quark and lepton masses we choose $`ϵ_\varphi =1/7`$ and $`ϵ_\chi =1/25`$. In this new scenario, the muon edm can be enhanced to $`5\times 10^{23}`$. We have taken soft masses for all the Higgs fields to be 85 GeV. In Fig. 3, we exhibit the results for $`d_\mu ^e`$ for one such example. To generate this plot, the input values we have used at the string scale are as follows:
$`m_{RR}^2`$ $`=`$ $`85^2\left(\begin{array}{ccc}1& \frac{1}{2}ϵ_\chi & \frac{1}{7}ϵ_\chi ^2/ϵ_\varphi \\ \frac{1}{2}ϵ_\chi & 1& ϵ_\varphi \\ \frac{1}{7}ϵ_\chi ^2/ϵ_\varphi & ϵ_\varphi & 1.8\end{array}\right);A_l=30\left(\begin{array}{ccc}ϵ_\chi ^4/ϵ_\varphi ^2& \frac{1}{2}ϵ_\chi & \frac{1}{3}ϵ_\chi ^2/ϵ_\varphi \\ \frac{1}{2}ϵ_\chi & \frac{1}{3}ϵ_\varphi ^2& \frac{1}{3}ϵ^{i\pi /3}ϵ_\varphi \\ \frac{1}{3}ϵ_\chi ^2/ϵ_\varphi & \frac{1}{3}ϵ^{i\pi /3}ϵ_\varphi & 4\end{array}\right).`$ (25)
$`𝐟`$ $`=`$ $`\left(\begin{array}{ccc}1.01\times 10^4& 9.0\times 10^4& 1.4\times 10^3\\ 9.0\times 10^4& 1.2\times 10^2& 1.04\times 10^1\\ 1.4\times 10^3& 1.04\times 10^1& 1.59\end{array}\right);`$ (26)
$`A_f`$ $`=`$ $`500\left(\begin{array}{ccc}ϵ_\chi ^4/ϵ_\varphi ^2& \frac{1}{3}ϵ_\chi & \frac{1}{7}ϵ_\chi ^2/ϵ_\varphi \\ \frac{1}{3}ϵ_\chi & 3ϵ^{i\pi /2}ϵ_\varphi ^2& 4ϵ^{i\pi /2}ϵ_\varphi \\ \frac{1}{7}ϵ_\chi ^2/ϵ_\varphi & 4ϵ^{i\pi /2}ϵ_\varphi & 0.6ϵ^{i\pi /2}\end{array}\right).`$ (28)
Note that we have allowed for all coefficients to be order one, consistent with the horizontal symmetry. (This is also true for the $`𝐟`$ matrix elements.) The $`(13),(31)`$ elements of the $`𝐟`$ are no longer very small like our previous example because of the symmetry requirement.
The choice of $`𝐟`$ matrix in this case corresponds to the following light neutrino masses: $`(6.27\times 10^6,\mathrm{\hspace{0.17em}2.9}\times 10^3,\mathrm{\hspace{0.17em}4.4}\times 10^2)`$ eV. The corresponding leptonic mixing matrix is:
$`U`$ $`=`$ $`\left(\begin{array}{ccc}0.99& 4.2\times 10^2& 3.2\times 10^3\\ 3.2\times 10^2& 0.70& 0.71\\ 2.9\times 10^2& 0.74& 0.66\end{array}\right).`$ (29)
Note that in this example $`\nu _e\nu _\mu `$ oscillation explains the solar neutrino data via small angle MSW oscillation. $`\nu _\mu \nu _\tau `$ oscillation explains the atmospheric neutrino data. We have found that by varying the order one couplings slightly, it is also possible to obtain a different scenario whre $`\nu _e\nu _\tau `$ oscillation is relevant for solar neutrinos, while $`\nu _\mu \nu _\tau `$ oscillations with $`m_{\nu _\mu }m_{\nu _\tau }`$ explains the atmospheric neutrino data . The predictions for $`d_\mu ^e`$ is not much altetered in such a scenario.
Now we turn to the evaluation of $`(g2)`$ of the muon. In MSSM, the $`(g2)_\mu `$ gets contribution from the chargino and neutralino diagrams . The relevant expressions can be found in Ref. . In this model we have contributions from both these loops. The chargino contribution is somewhat bigger than the neutralino loop. We find the magnitude of $`(g2)_\mu `$ to be $`(610)\times 10^{10}`$ for the curves in Figs. 1 and 2 and $`(813)\times 10^{10}`$ for the model with horizontal symmetry given in Fig. 3.
As for other rare processes, the branching ratio of $`\tau \mu \gamma `$ is one to two orders of magnitude below the present experimental limit. Since this process cnnot be made much smaller, it will be of great interest to improve the present limit by two orders of magnitude, which does not appear to be out of question. In all cases that we studied, the edm for electron is of order $`10^{28}`$ ecm. As for $`\mu e\gamma `$, it is three to four orders of magnitude smaller than current limits for cases (i) and (ii), and one order of magnitude smaller than current limits in the case of horizontal symmetry.
In conclusion, we have shown that in supersymmetric extensions of the stanadard model that accommodates small neutrino masses via the seesaw mechanism, there is an enhancement of the muon electric dipole moment. Interactions responsible for the generation of Majoran masses for the right–handed neutrinos are responsible for this enhancemnt through renormalization group effects. We have found values of $`d_\mu ^e`$ as large as $`5\times 10^{23}`$ ecm. Our finding should provide a strong motivation to improve the limit of $`d_\mu ^e`$ to the level of $`10^{24}`$ ecm, as has recently been proposed. Probing $`d_\mu ^e`$ at this level could reveal the underlying structure responsible for CP violation as well as for the generation of neutrino masses.
## V Acknowledgments
One of the authors (R.N.M) would like to thank K. Jungman for discussions on the present ideas for the searches for muon edm. K.S.B is thankful to the Theory Group at Fermilab where part of this work was done for its warm hospitality. The work of K.S.B is supported by Department of Energy Grant No. DE-FG03-98ER41076 and by a grant from the Research Corporation. B.D is supported by National Science Foundation Grant No. PHY-9722090. R.N.M is supported by NSF Grant No. PHY-9802551.
## VI Appendix
In this Appendix we give the renormalization group equations appropriate for the momentum range between $`M_{\mathrm{string}}`$ and $`v_R`$ for the case where parity is broken at $`M_{\mathrm{string}}`$.
$`{\displaystyle \frac{d𝐟}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}[4\pi (7\alpha _R+{\displaystyle \frac{9}{2}}\alpha _{BL})\mathrm{𝟏}+2𝐟𝐟^{}+4𝐘_l𝐘_l^{}+\mathrm{Tr}(𝐟𝐟^{})]𝐟,`$ (30)
$`{\displaystyle \frac{dY_l}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}[4\pi (3\alpha _R+3\alpha _L+{\displaystyle \frac{3}{2}}\alpha _{BL})\mathrm{𝟏}+2𝐟𝐟^{}+4𝐘_l𝐘_l^{}+\mathrm{Tr}(3𝐘_q𝐘_{q}^{}{}_{}{}^{}+𝐘_l𝐘_{l}^{}{}_{}{}^{})]𝐘_l,`$ (31)
$`{\displaystyle \frac{dA_l}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}[4\pi (3\alpha _R+3\alpha _L+{\displaystyle \frac{3}{2}}\alpha _{BL})A_l`$ (32)
$`+`$ $`8\pi (3\alpha _RM_R+3\alpha _LM_L+{\displaystyle \frac{3}{2}}\alpha _{BL}M_{BL})𝐘_l`$ (33)
$`+`$ $`4A_l𝐘_l^{}𝐘_l+8𝐘_l𝐘_l^{}A_l+2𝐟𝐟^{}A_l+4A_f𝐟^{}𝐘_l`$ (34)
$`+`$ $`2\mathrm{T}\mathrm{r}(A_l𝐘_l^{})𝐘_l+\mathrm{Tr}(𝐘_l𝐘_l^{})A_l+6\mathrm{T}\mathrm{r}(A_q𝐘_q^{})𝐘_l+3\mathrm{T}\mathrm{r}(𝐘_q𝐘_q^{})𝐘_l],`$ (35)
$`{\displaystyle \frac{dA_f}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}[4\pi (7\alpha _R+{\displaystyle \frac{9}{2}}\alpha _{BL})A_f+8\pi (7\alpha _RM_R+{\displaystyle \frac{9}{2}}\alpha _{BL}M_{BL})𝐟`$ (36)
$`+`$ $`8A_l𝐘_l^{}𝐟+4𝐘_l𝐘_l^{}A_f+2A_f𝐟^{}𝐟+4𝐟𝐟^{}A_f`$ (37)
$`+`$ $`2\mathrm{T}\mathrm{r}(A_f𝐟^{})𝐟+\mathrm{Tr}(𝐟𝐟^{})A_f],`$ (38)
$`{\displaystyle \frac{d𝐦_{RR}^2}{dt}}`$ $`=`$ $`{\displaystyle \frac{2}{16\pi ^2}}[4\pi (3/2\alpha _{BL}M_{BL}^2+3\alpha _LM_L^2)`$ (39)
$`+`$ $`{\displaystyle \frac{1}{2}}((𝐘_l𝐘_l^{}+𝐟𝐟^{})𝐦_{RR}^2+𝐦_{RR}^2(𝐘_l𝐘_l^{}+𝐟𝐟^{})+2(𝐘_l𝐦_{RR}^2𝐘_l^{})`$ (40)
$`+`$ $`2(\mathrm{𝐟𝐦}_{RR}^2𝐟^{}+m_\mathrm{\Phi }^2𝐘_l𝐘_l^{}+m_{\mathrm{\Delta }^c}^2𝐟𝐟^{}+A_lA_l^{}+A_fA_f^{}))].`$ (41)
|
warning/0006/cond-mat0006358.html
|
ar5iv
|
text
|
# Correlation gap in the optical spectra of the two-dimensional organic metal (BEDT-TTF)4[Ni(dto)2]
## Abstract
Optical reflection measurements within the highly conducting ($`a,b`$)-plane of the organic metal (BEDT-TTF)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] reveal the gradual development of a sharp feature at around 200 cm<sup>-1</sup> as the temperature is reduced below 150 K. Below this frequency a narrow Drude-like response is observed which accounts for the metallic behavior. Since de Haas-von Alphen oscillations at low temperatures confirm band structure calculations of bands crossing the Fermi energy, we assign the observed behavior to a two-dimensional metallic state in the proximity of a correlation induced metal-insulator transition.
Low-dimensional metals have attracted considerable interest recently due to observed deviations from the simple Fermi-liquid behavior. By now the electronic properties of quasi one-dimensional Bechgaard salts are fairly well understood in terms of a Tomonaga-Luttinger liquid ; but there exists only little agreement in the understanding of the quasi two-dimensional systems like organic or high-temperature superconductors .
The two-dimensional organic BEDT-TTF salts, where BEDT-TTF, abbreviated by ET, stands for bisethylenedithio-tetrathiafulvalene, are widely studied for their superconducting properties (which are still heavily debated ); only a few investigations deal with the metallic state, which seems to be highly interesting because its vicinity to unconventional states like a Mott-Hubbard insulator or antiferromagnetism. Most of the systems consist of monovalent anions with each ET molecule donating half an electron. The new charge transfer complex based on the electron donor ET and the acceptor nickelbis(dithiooxalate) \[Ni(dto)<sub>2</sub>\]<sup>2-</sup> was proven to be metallic down to low temperatures . The aim of our investigation was twofold: first to understand why the material does not become superconducting although it is highly ordered; second to study the effect of electronic correlations in this prototypical two-dimensional metal. Here we report on the optical properties of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] which strongly indicates that a correlation gap opens at low temperatures.
Single crystals of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] grown by electrochemical methods were as large as $`2\times 3\times 0.5`$ mm<sup>3</sup>. The crystal structure consists of ET stacks along the $`a`$-axis where the ethylene groups are highly ordered below 200 K as known from X-ray analysis. This resembles the $`\beta `$-modification of the ET salts with one-dimensional stacks of ET molecules, in contrast to the $`\kappa `$-phase where pairs of organic molecules are rotated by 90 with respect to each other . Due to S-S contacts of the ET molecules in neighboring stacks, (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] forms a conducting ($`a,b`$)-plane. These layers are separated along the $`c`$-direction by sheets of the \[Ni(dto)<sub>2</sub>\]<sup>2-</sup>. Although the absolute value of the conductivity perpendicular to the planes is about one to two orders of magnitude smaller , (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] exhibits a metal-like temperature dependence in all three directions (lower panel of Fig. 1), i.e. the conductivity increases like $`T^2`$ below 100 K.
Our samples were also characterized by electron spin resonance (ESR) experiments in the temperature range from 300 K down to 4 K (Fig. 1) which yield an absolute value of the spin susceptibility of $`6.4\times 10^4`$ emu/mole at room temperature. It decreases continuously by a factor of three when the temperature is lowered with the strongest reduction in the temperature range $`150\mathrm{K}>T>100`$ K indicating that the density of states might be reduced . The ESR linewidth $`\mathrm{\Delta }H`$ exhibits a maximum just below 100 K. This behavior is seen in all three directions. For $`T<80`$ K the drop of $`\mathrm{\Delta }H`$ and the constant spin susceptibility are in qualitative agreement with Elliott’s prediction for an isotropic metal. Internal fields lead to a pronounced anisotropy as reflected in the $`g`$-values (not shown); they are basically independent of temperature: $`g_a=2.0037`$, $`g_b=2.0048`$, and $`g_c=2.0110`$.
The Fermi surface of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] as obtained by band structure calculations (inset of Fig. 1) is similar to that of $`\alpha `$-(ET)<sub>2</sub>(KHg(SCN)$`{}_{4}{}^{})`$ and to the typical Fermi surface observed in the $`\kappa `$-phase of ET salts. It consists of a two-dimensional hole pocket ($`\alpha `$-orbit) which covers about 14% of the first Brillouin zone; as well as one-dimensional open trajectories with a flat dispersion which are less than half filled. These calculations are fully confirmed by Shubnikov-de Haas and de Haas-van Alphen (dHvA) experiments at low temperatures .
The polarized optical reflectivity $`R(\omega )`$ of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] was measured at $`5\mathrm{K}<T<300`$ K for the electric field parallel to the $`a`$ and to the $`b`$ axes. Using two Fourier-transform spectrometers and a quasi-optical submillimeter spectrometer we covered the spectral range from 20 cm<sup>-1</sup> up to 8000 cm<sup>-1</sup>. We combined all results on $`R(\omega )`$, and after using appropriate low-frequency (Hagen-Rubens) and high-frequency ($`\omega ^2`$ decay) extrapolations finally performed a Kramers-Kronig analysis to obtain the optical conductivity $`\sigma (\omega )`$ and dielectric constant spectra.
The frequency dependence of the reflectivity and the corresponding conductivity are displayed in Fig. 2 for both polarizations of the highly conducting plane. For the $`b`$-direction the plasma edge is located at somewhat higher frequencies as compared to the $`a`$-direction. In both directions no simple Drude response is observed. As the temperature is reduced below 150 K, pronounced new features develop in the low-frequency domain: a dip in $`R(\omega )`$ with a corresponding mode in $`\sigma (\omega )`$ around 200 cm<sup>-1</sup>, which we ascribe to excitations across a gap. As discussed below, we assign the origin of this gap to correlation effects. Below this gap, a narrow Drude-like peak appears at low temperatures. The position of the maximum is the same for both axes and does not change with temperature. The intensity (oscillator strength) increases linearly by about a factor of 10 when going from $`150\mathrm{K}`$ to 50 K and stays constant below.
We now discuss possible scenarios for the understanding of our experimental results. The major challenge is to reconcile two aspects of the experimental data. On one hand, the observed quantum oscillations are in complete agreement with calculations of the electronic band structure which do not take into account correlation effects. On the other hand, the optical conductivity shows signs of a gap. This situation stands in contrast to other materials such as (ET)<sub>2</sub>AuBr<sub>2</sub> which also exhibits a low temperature gap in the conductivity spectrum at 130 cm<sup>-1</sup>, the origin being assigned to a spin-density-wave ground state . This magnetic instability, originating from nesting properties of the one-dimensional parts of the Fermi surface, leads to a correlation-induced reconstruction of the Fermi surface topology which does not match band structure calculations. This has been confirmed by dHvA experiments . In the case of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] no indication of magnetic ordering was found at low temperatures, hence the formation of a spin density wave can definitely be ruled out. In particular the $`g`$-value does not show any shift in this range of temperature which would indicate the development of an internal magnetic field . No X-ray study of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] has been performed at $`T<200`$ K which would give a clear answer to whether a charge density wave ground state is present. However, we do not observe an abrupt change in any physical quantity and therefore we do not expect an increase in size of the cell as reported in typical charge density wave systems . Furthermore, such transitions should show up in the Fermi surface topology.
Given the above, the most natural way of understanding the experimental data is in terms of a metallic state in the proximity of a quantum phase transition which partially or totally gaps the Fermi surface. The feature in $`\sigma (\omega )`$ at 200 cm<sup>-1</sup> is interpreted as a precursor effect of this transition. As one approaches the transition (by changing the relevant microscopic parameters) spectral weight will be transfered from the Drude response to frequencies above the charge gap. The question is then: what is the nature of this transition and which microscopic parameter controls it?
To illustrate this idea, we consider one-dimensional organics such as (TMTSF)<sub>2</sub>PF<sub>6</sub>. Along the chains the optical conductivity has very similar features to those of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\]: a gap feature at 200 cm<sup>-1</sup> with a small Drude peak containing 1 % of the spectral weight. The Bechgaard salts consist of quarter-filled chains which are insulating at low temperature due to umklapp processes and/or the combination of dimerization and strong Coulomb interaction . The metallic behavior is attributed to interchain hopping which effectively warps the one-dimensional Fermi surface thus leading to doping away form the ideal quarter-filled case. In this picture, the conductivity spectrum corresponds to that of a metallic state in the proximity of a metal-insulator transition, the control parameter being the (dimensionality driven) doping away from quarter-filling . The Fermi-surface of the (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] material has one-dimensional features; however, interchain hopping is a sizable fraction of the intrachain hopping, which explains the nearly isotropic dc conductivity within the ($`a,b`$)-plane. Although tempting, an interpretation of the data in terms of strongly coupled chains seems out of reach.
Since the conductivity of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] is roughly isotropic in the ($`a,b`$)-plane it is appropriate to start with a two dimensional model. The unit-cell contains four ET molecules each donating half an electron to the acceptor. Band structure calculations reveal two conduction bands. In comparison to the energy scale of the gap, the other two bands lie well below the Fermi energy and we will omit them. In real space this leads to a model with two $`s`$-orbitals per unit-cell each orbital accommodating one hole, i.e. half-filling. The reduction to a two-band model may be explicitly carried out by noting that there is one hopping matrix element between the ET molecules which dominates, thus defining a dimer. Each unit cell contains a pair of dimers, and the two conduction bands are well reproduced by considering only the antibonding combination of orbitals on the dimer. A similar mapping was carried out for the $`\kappa `$-phase of the ET salts .
Correlation effects are taken into account with a Hubbard $`U`$-term which sets an energy cost $`U`$ for doubly occupied $`s`$-orbitals. Since we have precisely one hole per orbital large values of $`U`$ lead to the localization of holes and hence a Mott insulating state . The critical value of $`U_c`$ at which the metal-insulator transition occurs depends sensitively upon details of the band structure. In particular, if nesting is present, as in the prototype single band two-dimensional half-filled Hubbard model with nearest neighbor hopping, $`U_c=0`$ and the gap in a mean-field approach scales as $`e^{t/U}`$ where $`t`$ is the hopping matrix element. On the other hand, when frustration is present thus prohibiting magnetic ordering, one expects $`U_c`$ to be a sizable fraction of the bandwidth.
The antiferromagnetic insulating phase of the $`\kappa `$-phase salts is naturally described within the above approach if one chooses $`U>U_c`$. For the present material it is tempting to argue that $`U<U_c`$. This qualitatively explains the experimental data: the system remains metallic at arbitrary temperatures and no phase transition is present. Since the Mott transition is a mass divergent transition, one expects the Fermi surface to remain essentially unaltered as a function of $`U<U_c`$. As mentioned above this is consistent with the dHvA experiments . An indication of the mass enhancement due to proximity of the Mott transition stems from the fact that in $`a`$ and $`b`$ directions respectively only 2% and 4% of the total spectral weight of the optical conductivity is contained in the narrow Drude-like mode. Comparing those values to the large Fermi surface observed by quantum oscillations infers an enhanced mass.
The major problem within the above approach is the small energy scale of the gap which is more than an order of magnitude smaller than the bandwidth. Approximate nesting properties of the Fermi surface will certainly reduce the value of the gap. A calculation of the spin and charge suceptibilities within the Hükel tight-binding model shows only slight enhancements at wavevectors corresponding to nesting of the one-dimensional portions of the Fermi surface.
Finally some remarks on the superconducting state which was not detected in (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] down to 0.06 K; the effect of the dimensionality is not clear yet, but it should be noted that the system is the least two-dimensional ET conductor. As argued above, in (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] the Coulomb repulsion is not strong enough to drive the system to a Mott insulator. Given the phase diagram of the $`\kappa `$-phase as a function of pressure , we can only speculate that reducing the pressure by chemical substitution will drive the system to a superconductor and ultimately to a Mott insulator. This would agree with superconductivity observed at quantum phase transitions in numerous systems . We also predict a similar scenario in $`\alpha `$-(ET)<sub>2</sub>(KHg(SCN)$`{}_{4}{}^{})`$; optical investigations are in progress.
Thus we can conclude that the energy gap which develops in the optical conductivity at around 200 cm<sup>-1</sup> for temperatures $`T150`$ K in both directions of the highly conducting ($`a,b`$)-plane of (ET)<sub>4</sub>\[Ni(dto)<sub>2</sub>\] is due to electronic correlations. To reconcile the facts that the Fermi surface shows no correlation-induced gap and that there is no phase transition as the temperature is lowered, we have argued that the system is in the proximity of two-dimensional Mott metal-insulator transition. The narrow Drude-like contribution to the optical conductivity which is responsible for the metallic conductivity and quantum oscillations at low temperatures contains only of few percent of the spectral weight. The effective mass of these carriers has to be enhanced.
We thank A. Darjushkin for assistance and G. Grüner, H. J. Koo, and M.H. Whangbo for instructive discussions.
|
warning/0006/astro-ph0006060.html
|
ar5iv
|
text
|
# The stars near the centre of supergiant shell LMC 4: Further constraints on triggering scenariosBased on observations collected at the European Southern Observatory (ESO), La Silla, Chile.
## 1 Introduction
The creation mechanism for interstellar shells of diameters larger than 300 pc continues to be debated. Such sizes rule out that a single (central) stellar association can provide the energy (the combined effect of SN explosions, stellar winds and radiation pressure) to create such a supershell (see Tenorio-Tagle & Bodenheimer 1988 for a review). Apparently much more energy is needed so that different formation mechanisms have been considered.
In the Large Magellanic Cloud (LMC) Meaburn and collaborators (see e.g. Meaburn 1980) have identified nine supergiant shells (SGSs) based on their filamentary ring-like structure on H$`\alpha `$ plates. These SGSs have diameters in the range of $`0.61.4`$kpc. The largest and best studied is LMC 4, which appears in the light of H$`\alpha `$ as an ellipse of 1 kpc (in E-W) and 1.4 kpc diameter (in N-S direction). In H i it is a thick shell with a gas depleted central area (see, e.g., Luks & Rohlfs 1992; Kim et al. 1999).
The inner part of LMC 4 is dominated by the stellar superassociation LH 77. In literature one finds the name Shapley Constellation III, which is also used for the entire inner part of this SGS. Recently Efremov & Elmegreen (1998) pointed out that this term was used for the wrong object. Originally McKibben Nail & Shapley (1953) ‘defined’ it to be a triple aggregation of $`26^{}\times 26^{}`$ area and to contain NGC 1974. So this name would not be appropriate for the LH 77 region (see Fig. 7), but for the N 51 region, containing LH 63, LH 60, and LH 54. Thus we consider it best to avoid the ambiguous term Sh III and to only use appropriate object names based on a clear naming convention, like LH$`nn`$ (Lucke & Hodge 1970; Lucke 1972), N$`nn`$ (Henize 1956), or LMC 4 (Goudis & Meaburn 1978).
Since the identification of LMC 4 as SGS, several photometric investigations have been carried out. These aimed, in part, at finding the age of the stars inside LMC 4 and at its rim with the goal of tracing the history of the SGS. These studies include those of Isserstedt (1984) and Reid et al. (1987), which indicated propagating star fomation, and the recent CCD photometries of Braun et al. (1997) and Dolphin & Hunter (1998), showing a nearly coeval population without any age gradient (see Sect. 3). By now, the CCD studies, in combination with new data presented in Sect. 2, cover about half of the area inside the H$`\alpha `$ rim. All these studies of the interior agree on the young age of the stars, being of the order of 11-13 Myr, with very little spread. Of course, a much older background population is present, too. But it must be the young stars which are related with the formation of the SGS LMC 4.
Additional studies of star groups near the outer edge include those about NGC 1948 (Vallenari et al. 1993; Will et al. 1996), LH 63, LH 60, LH 54 (Petr et al. 1994), LH 76 (Wilcots et al. 1996), LH 95, LH 91 (Gouliermis et al. 2000), NGC 2030 (Laval et al. 1986), and the association LH 72 (Olsen et al. 1997) in the (2D projected) central region. The surprising result of these studies is, that the stars in the SGS (those near the H$`\alpha `$ emission) have an age only marginally younger than those in the interior.
The small difference in age between the interior and the edge of LMC 4 makes it likely that the stars near the H$`\alpha `$ emission were formed as a result of the events taking place inside the SGS. However, the sheer volume of space interior to LMC 4 containing stars of the same age makes LMC 4 a truly exceptional area.
In the present study we have compiled the data and the results on ages and population structure for the LMC 4 area. We also add some new photometry, consisting of three fields near the centre covering a total of $`167.3\mathit{}^{}`$.
The overall data are analysed to find further clues to the origin of the SGS LMC 4. Various mechanisms have been proposed for the formation of LMC 4. These include the collision with a high-velocity H i cloud (HVC), the stochastic self-propagating star formation (SSPSF), the triggering of star cluster arcs, and the LMC bow-shock as star formation trigger. These will be discussed, including remarks about their viability and verifiability.
## 2 Observations and data reduction
The new data<sup>1</sup><sup>1</sup>1 The entire Table 4 of this publication is only available electronically, at the CDS (see Editorial in A&A 280, E1, 1993) or at the Astronomical Institutes of Bonn University (’ftp ftp.astro.uni-bonn.de’; further information can be obtained at the URL ’http://www.astro.uni-bonn.de/$`\stackrel{~}{}`$jbraun/phdt\_lmc4c.html’). were taken in January, 1999 with the $`1.54`$m Danish Telescope at ESO observatory on Cerro La Silla. The telescope was equipped with DFOSC and the 2k$`^2`$pix<sup>2</sup> LORAL CCD (W7 Chip) resulting in a scale factor of $`0.39^{\prime \prime }`$ pix<sup>-1</sup>. The chip possesses one 6-fold, six double and six single bad columns plus a 5-fold bad line produced by the electronics on two object frames. Additionally, there were some flatfield problems causing a residual large-scale gradient of up to 4% in $`V`$ sky flats, so flatfields had to be derived by combining object frames and twilight flats.
The data reduction was performed on GNU/Linux (i386) workstations with MIDAS (see e.g. Banse et al. 1983), IRAF (Tody 1986) and DAOPHOT II (Stetson 1987).
The calibration was achieved using the standard fields SA$`\mathrm{\hspace{0.17em}95}42`$, Rubin$`\mathrm{\hspace{0.17em}149}`$, PG$`\mathrm{\hspace{0.17em}1047}+003`$, and PG$`\mathrm{\hspace{0.17em}0918}+029`$ (Landolt 1992), observed in $`B,V`$ passbands at different airmasses on 14th January 1999, yielding an atmospheric extinction of $`k_V=0.163(32)`$mag and $`k_B=0.297(42)`$mag and the calibration constants of Eqs. (1) and (2). After normalization (indicated by the index ’n’) to airmass 0 and 1 s exposure time including the PSF to aperture shift ($`\delta _V=0.066(15)`$mag and $`\delta _B=0.005(20)`$mag), the following relations were applied:
$$(BV)=\begin{array}{c}\left[(BV)_\mathrm{n}0.261(46)\text{mag}\right]/\mathrm{\hspace{0.17em}0.931}(18)\hfill \end{array}$$
(1)
$$V=\begin{array}{c}V_\mathrm{n}0.683(15)\text{mag}+0.021(6)(BV)\hfill \end{array}$$
(2)
The dataset contains the three fields C<sub>1-3</sub> of $`167.3\mathit{}^{}`$ area (see Fig. 1 for their location and Fig. 3a for an image composit). Each was observed in $`B,V`$ passbands with short (30 and $`15`$s) and long (480 and $`240`$s, respectively) exposures (for field C<sub>3</sub> the exposure times were 240 and $`480`$s) during a typical seeing of $`1.5^{\prime \prime }`$. The entire CCD mosaic shown in Fig. 3a covers $`328.5\mathit{}^{}`$.
The brightest star located near the centre of the analyzed area is the galactic F8 dwarf HD 37195 ($``$ P 1313 $``$ MACS J0532$``$666#013, see Table 3, #7). The compact cluster HS 343 = KMHK 1030 (Hodge & Sexton 1966; Kontizas et al. 1990) of $`15`$pc diameter can be found 2.9 arcmin to the east of star #7 while 2.5 arcmin to the west of star #10 (Table 3 and Fig. 3a) the loose cluster KMHK 1000 of $`13`$pc diameter is located.
As the western part of the centre field C<sub>3</sub> overlaps with the N-S strip of the ’J’-shaped region (see Fig. 1), Fig. 3 shows both CCD mosaics at the same scale, so e.g. stars marked #9, 13, 15, and 16 on the left panel can easily be found in the right panel.
Table 1 lists photometric errors derived by DAOPHOT and their standard deviation for three magnitude intervals (comparable to the statistical errors given in Table 2 of Braun et al. 1997). The values for the single stars can be found in the resulting data table (Table 4). The quality of all three large photometric studies inside LMC 4 can be checked by plotting the magnitude and colour differences of stars in common (see Fig. 2). The mean value, standard deviation, minimum and maximum value of the differences between the photometries of Braun et al. (1997, ’J’), Dolphin & Hunter (1998, ’DH’) and the present one (’C’) are given in Table 2. Because the data were paired by software, in particular at the faint end mismatches occured, which affect these values. To determine better statistics, we applied selection criteria given in the footnote of Table 2.
Even though the new dataset covers a larger area ($`A_{\mathrm{C}_{13}}1.1A_\mathrm{J}`$), the colour-magnitude diagram of the new data (Fig. 4a) shows less upper main sequence stars than the CMD of the earlier data (Fig. 4b). This is caused by the larger stellar density in the E-W portion, i.e. fields 0-9 of the ’J’-shaped region containing a part of LH 77. The combined datasets cover $`574\mathit{}^{}`$ on the sky, i.e. about 11% of the LMC 4 area as defined by the H$`\alpha `$ filaments.
To calculate absolute magnitudes we used the distance modulus of $`(mM)_0=18.5`$mag, which is still the best applicable value. Gibson (2000) gives a range of current determinations of $`[18.07,18.74]`$mag (see also Groenewegen & Oudmaijer 2000).
The isochrones of the Geneva group (Schaerer et al. 1993) for LMC metallicity ($`Z=0.008`$ or \[Fe/H\]$`=0.34`$dex) had been fitted by eye, yielding age and interstellar extinction as given in the next section.
## 3 Results of photometric analyses
To determine the age of stellar populations inside LMC 4 we had already obtained CCD photometry in $`B,V`$ passbands (Braun et al. 1997). The two strips with a total of 25 CCD positions (see Fig. 1) cover 298 arcmin<sup>2</sup> equivalent to 6% of the area inside the H$`\alpha `$ filaments. This ’J’-shaped area constitutes an E-W strip of $`400`$pc in the region of the OB superassociation LH 77 and a S-N strip of $`850`$pc.
The colour-magnitude diagrams (CMDs) and luminosity functions (LFs) derived from the data yielded ages of $`916`$Myr, a colour excess of mostly only 0.11 mag (in some places even less), and LF slopes of 0.22-0.41 (Braun et al. 1997; Braun 1996). The age range given is maximized but nevertheless by a factor 2 smaller than the necessary age spread required by the global SSPSF scenario (see Sect. 5.2). Additionally, no correlation with the distance to the centre of LMC 4 was found. In all CMDs a young population of about 11 Myr is present, even though the stellar density of this population gets low towards the north leading to the higher LF slopes (compared to an expected slope $`\gamma =0.27`$, see Will 1996).
Dolphin & Hunter (1998) analyzed $`UBV`$ photometry of four fields inside LMC 4 and one field outside, near its rim. This study yields ages of 12-16 Myr and $`BV`$ excess reaching from 0.04 to 0.07 mag inside LMC 4 and $`7`$Myr and 0.09 mag for the NGC 1955 field (i.e. the N 51 region). The lack of an age gradient inside the SGS is supported.
These studies can now be extended by the present $`B,V`$ CCD photometry for three positions of the central region inside SGS LMC 4 (see Fig. 1), which are centered near the clusters HS 343 and KMHK 1000. The observed area of 69 500 pc<sup>2</sup> would contain the stellar population predicted to be 30 Myr of age according to the model of the triggered star cluster arcs (see Sect. 5.3), even if that population were spatially dispersed.
The isochrone fit to the new dataset yields an age of $`11(2)`$ Myr and a colour excess $`E_{BV}=0.10(3)`$mag. This value was also checked by deriving intrinsic colours and thus reddening by the Wesenheit function (see e.g. Hill et al. 1994): $`W:=V3.1(BV)`$ and an appropriate fit to the stellar evolution model. We used the main sequence (MS) stars with $`V[14.7,17.5]`$mag and an isochrone of $`\mathrm{log}(t/[\mathrm{yr}])=7.05`$ yielding $`(BV)_00.693\text{mag}+0.025W`$. The $`E_{BV}`$-distribution of these 629 MS stars has its maximum at 0.1 mag with a mean (and deviation) of $`0.12(4)`$mag.
The similarity of the morphology visible in the CMDs of the two datasets, the one (Fig. 4a, fields C<sub>1-3</sub>) containing the populations at the LMC 4 centre and the other (Fig. 4b, fields $`010`$ and $`1724`$) being dominated by the superassociation LH 77, is striking. No population of age 30 Myr is present in the central part. This follows directly from the comparison of the region in both CMDs to the right of the main sequence, near the $`30`$Myr isochrone.
The members of the group of six A-type supergiants near HS 343 (see Fig. 6; these stars selected by Efremov & Elmegreen 1998 from the data of Rousseau et al. 1978 are No. 5, 8, 9, 10, 12, and 13 of Table 3), quoted as relicts of a $`30`$Myr population, are marked by crosses in Fig. 4a. The CMDs for the two clusters near the geometric centre of the filamentary H$`\alpha `$ boundary show their ages to be $`0.1`$Gyr for HS 343 and $`0.3`$Gyr for KMHK 1000 (see Fig. 5).
## 4 Age structure inside LMC 4
The similarity of the morphology visible in the CMDs of all datasets is striking. Therefore the stars inside LMC 4 must have been formed in a short period of time, so the H$`\alpha `$ and H i feature called LMC 4 must have been created in a turbulent manner rather than in a ’bubble blowing’ action. This fits also to the small (i.e. $`500`$pc) disk thickness of the LMC (e.g. Kim 1998 derived in Sect. 5.2.1 an exponential scale height of the gas of $`170`$pc, i.e. a disk thickness of $`340`$pc). This implies that this SGS resembles a truncated cylinder rather than a bubble. The onset of this star burst can be dated by isochrone fitting giving as average age of the interior population $`12(2)`$ Myr. No other population of age less than 100 Myr is present in the central part.
The star groups near the edge are younger indeed than the stars in the interior of LMC 4. The age at the edge (age of stars associated with the H$`\alpha `$ emission outside the H$`\alpha `$ rim of the SGS) is 4-7 Myr (see Table 5 and Sect. 6.1). This implies that $`5`$Myr after the giant central volume formed stars, the edge was pushed into star formation mode, too.
One is thus faced with several questions pertaining to the origin of the structure called LMC 4. What mechanism can stimulate simultaneous star formation over an area (in a volume) with projected size of about $`800\text{ pc}\times 1200`$pc (the interior of LMC 4)? By what mechanism was the gas of the birth cloud dispersed so efficiently to create the presently low column densities in the H i over such an area? What stimulated the star formation at the outskirts of this area (creating the H$`\alpha `$ structure called LMC$`\mathrm{\hspace{0.17em}4}`$) in a relatively homogeneous manner?
## 5 Discussion of scenarios proposed for LMC 4’s creation
Several star formation scenarios possibly leading to the formation of supergiant shells are mentioned in the literature. Here we will comment on each in relation with our knowledge of LMC 4.
### 5.1 HVC infall
An infalling high-velocity cloud might compress the gas in the LMC disk and so trigger star formation.
Infall would affect star formation in a large region indeed and we see that the ages derived for the large inner part of LMC 4 are equal (see Sect. 3). However, the data about the density and radial velocity of LMC 4 gas contradict this scenario (Domgörgen et al. 1995), and infall was not the actual trigger for the formation of LMC 4.
In addition, given the motion of the LMC through the halo of the Milky Way (Kroupa & Bastian 1997), the velocity of such a HVC should have been rather orthogonal to that of the LMC itself, thus having a velocity vector pointing more or less toward the Sun. A HVC moving away from the Milky Way is very unlikely. HVCs moving toward the Milky Way at $`+60`$ and $`+120\mathrm{km}\mathrm{s}^1`$ are known in this direction (Savage & de Boer 1979). Since this high velocity gas is seen over the entire face of the LMC (see Savage & de Boer 1981), with the LMC at $`v_{\mathrm{rad}}250\mathrm{km}\mathrm{s}^1`$, it is rather nearby.
### 5.2 SSPSF
In the stochastic self-propagating star formation scenario the stars formed initially trigger further star formation in their vicinity.
SSPSF leads to an age gradient from the starting point, e.g. the centre of a SGS, outward to the rim (Feitzinger et al. 1981). It has been a popular model and was often cited in papers about supershells in distant galaxies where no photometry of individual stars is available.
The predictions by SSPSF for the morphology of SGSs agree with the observed density and radial velocity of gas in the interior of LMC 4. However, the large radius of the SGS would lead to a continuous age spread of the stars of at least 15 Myr range. Even if one takes a possible overlap of stellar populations having different ages into account, the photometric data are in contradiction with such a population structure and thus indicate incompatibility of the SSPSF scenario with the interior of LMC 4. This has now been sufficiently noted (see Braun et al. 1997; Dolphin & Hunter 1998).
### 5.3 Triggering of stellar arcs
Star clusters or $`\gamma `$-ray bursts may trigger star formation to form arc like structures.
Efremov & Elmegreen (1998) suggested as trigger for star formation in LMC 4 the self-gravitational collapse of parts of rims from shells initially driven by star clusters, leading to the building of star cluster arcs. They identify three of these structures (see Fig. 6), circular (in the 2d projection) in shape, and this may indicate that they have been formed by swept-up material. Efremov et al. (1998) subsequently suggested that gamma-ray bursts (GRBs) may produce the pressurizing wave.
These suggestions are rather ad hoc. Moreover, they would create localized structures and nothing of the kind like LMC 4. The collapse into a cluster or the explosion of a GRB, formation of a shell of swept up material which itself forms stars, which then leads to a next shell or the acceleration of the old shell (the scenario for the ‘Quadrant’ and subsequently for LMC 4 after Efremov & Elmegreen 1998) may explain some features, but not the entire SGS with the documented age structure (see Sect. 4). Furthermore it is hard to explain why ‘Quadrant’ is coeval (Braun et al. 1997) while ‘Sextant’ shows a clear age gradient (Petr et al. 1994) if the same mechanism would be valid. The geometry of the (after Efremov & Fargion 2000) four cluster arcs are rather due to cloud structure long before star formation than being directly connected to the formation process.
### 5.4 Star formation triggered by the LMC bow-shock
The motion of the LMC through the Milky Way halo leads to a bow-shock triggering star formation.
Since a very large area created stars at the same time, and since (due of the rotation of the LMC) the present day LMC 4 was near the leading edge about 15 Myr ago, the bow-shock of the LMC may have triggered the conditions for large-scale star formation. Searching for a possible creation mechanism of supergiant shell LMC 4 and maybe of all the supershells of the Large Magellanic Cloud, one finds that the superstructures of the LMC (see Fig. 7) are all located at the outskirts. Looking at these features clockwise from south-east to north-west, the expected correlation of age with traversed distance is visible (see de Boer et al. 1998).
The scenario has as essential part the motion of the LMC through the galactic halo and the LMCs clockwise rotation. The gas of the LMC rotates through the compression zone at the south-east, to eventually find itself in the north-east (LMC 4) with all the consequences of the star formation triggered inside that gas.
## 6 Summary and conclusions
In the review of the scenarios above we have indicated that the first three are not compatible with the observational facts. The most important fact is that over the entire interior of LMC 4 star formation burst out on a large scale about $`12`$Myr ago (see Table 5 and Sects. 3 and 4). Small stellar associations cannot be used as trigger for large-scale structures and there is no evidence for a recent collision with a HVC. The bow-shock formation scenario is the only viable one. Note that the bow-shock scenario explains only the formation of large structures around the edge of the LMC. At any given location in the LMC star formation continues as is readily visible from the ubiquitous H$`\alpha `$ structures everywhere in the LMC.
### 6.1 Evolution after the star burst took place
After formation of the original cloud of stars these stars emitted their energy into the surroundings, driving the gas out of this region rather more in a turbulent than a pressurized manner. The total energy produced by the massive stars, (stellar winds, ionizing photons, and supernova explosions) went into ionizing and expansion of the original birth cloud. These aspects have been discussed by Braun et al. (1997) and by Dolphin & Hunter (1998) and are not repeated here.
After the formation of this 1st generation of stars, star formation was initiated at the edge of the central region, creating populations of younger ages visible by the H ii regions surrounding this SGS. Examples for these stellar populations are LH 63, LH 60 and LH 54 (Petr et al. 1994), situated in the N 51 region, and LH 76 (Wilcots et al. 1996) more to the east (see Fig. 7 and Table 5). Also part of LH 72 (Olsen et al. 1997) is one of these, assuming it lies somewhat outside the plane of LMC 4, being at a different depth in the LMC. This phase with the formation of stars at the outskirts of the original cloud affects only regions of small size (i.e. $`<300`$pc) as expected from the limits due to inhomogenity (density, temperature, magnetic fields etc.) of the medium.
### 6.2 Details of the bow-shock formation and consequences for the overall structure of the LMC
The convergence of gas near the bow-shock, due to the LMC space motion combined with the LMCs rotation (see de Boer et al. 1998) creates a large and condensing gas cloud. In it star formation sets in, and the rotation of the LMC moves this region around the edge in clock-wise fashion.
The role of ram-pressure, especially the comparison of strength of hydrodynamic and gravitational effects, is still uncertain. Observational findings like X-ray shadowing (Blondiau et al. 1997), the steep edge in H i near the location of the largest velocity with respect to the halo gas (see e.g. the map by Kim et al. 1999), and H$`\alpha `$ emission of some main components of the Magellanic Stream (Weiner & Williams 1996) are all indications for bow-shocks due to motion of Magellanic material through halo gas. To fit the H i maps of the Stream and other structures, recent models for the tidal interactions of the Magellanic Clouds (MCs) are extended by a drag term (Gardiner 1999), accounting for the motion of the MCs through the Milky Way halo. However, problems remain and not all inconsistencies are solved (see e.g. Moore & Davis 1994 and Murali 2000).
The shape of the dark cloud presently seen south of 30 Dor is elongated, oriented perpendicular to the direction of the rotation centre of the LMC. The most dense region of stars inside LMC 4, being the hatched region in Fig. 7 with the embedded smaller associations LH 65 and LH 84, has the same orientation. We note that at their radial distance from the centre (both at approximately 2 kpc) the LMC rotation curve is rather level, and rotation will hardly change the original shape of structures.
Since the density of the LMC drops radially outward, it is to be expected that older SGSs will have opened up to the LMC edge. The morphology of the H$`\alpha `$ distribution may therefore give further evidence for the expected age sequence of the SGSs. Continuing in clockwise direction, the supergiant shell LMC 5 to the east of LMC 4 is still complete, while LMC 6 is a non-contiguous ring. Note also, that the star density in LMC 4 drops outward and that LMC 4 is elongated toward the lower density on the outside.
Apart from the young stars in LMC 4 older populations are present. These may date from a previous burst, of one or more revolutions ago. Such earlier events will have left numerous earlier generations. However, a bow-shock triggered burst need not take place continuously, since apparently gas accumulates intermittantly and the power of the trigger is sensitive to the current halo density at the front of the LMC. If it were continous, one would see a continuous ring of young stars around the LMC’s edge.
The age of small-scale structures may or may not be related to the age of the large-scale structures such as LMC 4. One full rotation of the LMC edge takes $`100`$Myr so that larger older structures may be part and/or consequence of a previous large-scale event.
### 6.3 Comparison with other dwarf galaxies
Looking at other dwarf galaxies in H$`\alpha `$ or H i, one recognizes that supershells are quite common features in galaxies and especially in dwarf galaxies (e.g. Brinks & Walter 1998), where the symmetry is not disturbed by differential rotation. However the different environment of these galaxies, such as tidal forces from neighbouring galaxies or being completely isolated, makes it difficult to explain these supershells by the same creation mechanism as the supergiant shells in the Magellanic Clouds. Still, in such galaxies larger volumes of space must attract sufficient amounts of gas in dense form that a starburst takes place, setting off star formation nearly simultaneous in huge volumes.
Clearly, photometric studies of resolved stellar populations in more distant (dwarf) galaxies are needed to achieve a general picture on these large-scale star forming processes.
### 6.4 Conclusions
The photometry of large areas inside LMC 4 shows that the stars are of the same age, being $`12`$Myr. The rotation speed of the LMC makes clear that the LMC 4 region was at time 0 near the bow-shock. At the bow-shock one currently finds a huge dark cloud being in the process of forming stars. Combining these two observational facts led to the bow-shock scenario and further evidence has been brought forward to substantiate this scenario.
Observing the interior of further supergiant shells will provide clearer insight in the formation history of the star groups inside such structures. Except for the complicated 30 Dor region (with the neighbouring supergiant shells LMC 2 and LMC 3) all other SGSs should show a coeval large-scale stellar population fitting to the age sequence proposed by the bow-shock scenario. Were this found to be not so, then the bow-shock trigger model would be falsified. One then would be in need of finding another star formation trigger. Given the current observations the only other possible (however unlikely) scenario would then be a special HVC collision. As yet, the bow-shock induced star formation fits the available observational data best.
## Acknowledgments
JMB was partly supported through the Deutsche Forschungsgemeinschaft (DFG) in the Graduiertenkolleg ‘The Magellanic Clouds and other Dwarf Galaxies’, MA by the DFG through grant Bo 779/21.
We wish to thank Dr. Deidre A. Hunter for providing us a CCD image of their LH 77 field for star identification.
|
warning/0006/hep-ph0006230.html
|
ar5iv
|
text
|
# Parton densities and structure functions beyond the next-to-leading order
## 1 Introduction
Structure functions in deep-inelastic scattering (DIS) are among the quantities best suited for measuring the strong coupling constant $`\alpha _s`$. They also form the backbone of our knowledge of the proton’s parton densities, indispensable for analyses of hard scattering processes at proton–(anti-)proton colliders like Tevatron and the LHC. During the past two decades DIS experiments have proceeded towards high accuracy and a greatly extended kinematic coverage .
To make full use of these results requires transcending the standard next-to-leading order (NLO) formalism . Indeed besides the QCD $`\beta `$-functions to even NNNLO , the NNLO (2-loop) coefficient functions for DIS have been calculated some time ago . However, only partial results have been obtained so far for the corresponding 3-loop splitting functions . The derivation of the full results is under way .
In we have derived approximate expressions for the 3-loop $`\overline{\text{MS}}`$ splitting functions which are sufficient for reliable NNLO analyses down to $`x10^3`$. These functions turn out to be much less important than the 2-loop coefficient functions at $`x10^2`$. Thus it is possible, based on partial results on the 3-loop coefficient functions, to proceed to NNNLO at large $`x`$ , especially for the non-singlet case most important for extractions of $`\alpha _s`$ from DIS.
## 2 Parton densities: formalism
It is convenient to work with the flavour non-singlet (NS) and singlet (S) combinations of the (anti-)quark and gluon densities, $`q_i`$, $`\overline{q}_i`$ and $`g`$:
$`q_{\mathrm{NS}}^\pm `$ $`=`$ $`q_i\pm \overline{q}_i(q_k\pm \overline{q}_k)`$
$`q_{\mathrm{NS}}^V`$ $`=`$ $`{\displaystyle \underset{r=1}{\overset{N_f}{}}}(q_r\overline{q}_r)`$ (1)
$`q_S^{}`$ $`=`$ $`\left(\begin{array}{c}\mathrm{\Sigma }\\ g\end{array}\right),\mathrm{\Sigma }={\displaystyle \underset{r=1}{\overset{N_f}{}}}(q_r+\overline{q}_r).`$ (4)
Here $`N_f`$ is the number of effectively massless flavours. As in (2) we often suppress the dependence on the momentum fraction $`x`$ and the renormalization and factorization scales, $`\mu _r`$ and $`\mu _f`$.
Using (2) the evolution equations are decomposed into $`2N_f1`$ scalar (NS) equations and the $`2\times 2`$ singlet system, all schematically written as
$$\frac{d}{d\mathrm{ln}\mu _f^2}q=𝒫q_x^1\frac{dy}{y}𝒫(y)q\left(\frac{x}{y}\right).$$
(5)
At $`\mu _r=\mu _f`$ the expansion of the splitting functions $`𝒫`$ up to NNLO is given by
$$𝒫a_sP^{(0)}(x)+a_s^2P^{(1)}(x)+a_s^3P^{(2)}(x).$$
(6)
Our choice of the expansion parameter reads
$$a_s\alpha _s/4\pi .$$
(7)
The expression for $`\mu _r\mu _f`$ is obtained from (6) by inserting the expansion of $`a_s(\mu _f^2)`$ in terms of $`a_s(\mu _r^2)`$ and $`L_R=\mathrm{ln}(\mu _f^2/\mu _r^2)`$. Large logarithms in $`𝒫`$ are avoided by choosing $`\mu _r=O(\mu _f)`$.
## 3 Splitting functions
The functions $`P^{(0)}`$and $`P^{(1)}`$ in (6) are known . The current information on $`P_{\mathrm{NS}}^{(2)\pm }(x)`$ comprises
* the first five even-integer moments of $`P_{\mathrm{NS}}^{(2)+}`$ given by $`P_{\mathrm{NS}}^{(2)+}(N)=_0^1𝑑xx^{N1}P_{\mathrm{NS}}^{(2)+}(x)`$ , and the first moment ($`N`$=1) of $`P_{\mathrm{NS}}^{(2)}`$,
* the complete $`N_f^2`$ contribution ,
* the leading small-$`x`$ terms $`\mathrm{ln}^4x`$ .
The difference of $`P_{\mathrm{NS}}^{(2)}`$ and $`P_{\mathrm{NS}}^{(2)+}`$ is expected to be negligible at large $`x`$. It has been conjectured that the leading large-$`x`$ terms are $`1/[1x]_+`$.
The following partial results have been derived so far for the singlet splitting functions $`P_{ij}^{(2)}(x)`$:
* $`P_{ij}^{(2)}(N)`$ for $`N=2,\mathrm{\hspace{0.17em}4},\mathrm{\hspace{0.17em}6}`$ and 8 ,
* the $`C_AN_f^2`$ contribution to $`P_{gg}^{(2)}(x)`$ ,
* the leading small-$`x`$ terms $`(1/x)\mathrm{ln}x`$ of $`P_{qq}^{(2)}`$, $`P_{qg}^{(2)}`$ and $`P_{gg}^{(2)}`$ , see also .
The $`1/[1x]_+`$ terms of $`P_{gg}^{(2)}`$ and $`P_{qq}^{(2)}`$ are expected to be related by a factor $`C_A/C_F=9/4`$.
We have derived approximate expressions for $`P_{\mathrm{NS}}^{(2)\pm }(x)`$ and $`P_{ij}^{(2)}(x)`$ from these constraints. After decomposing the functions into
$$P^{(2)}=P_0^{(2)}+N_fP_1^{(2)}+N_f^2P_2^{(2)},$$
(8)
we employ the ansatz (cf. )
$$P_m^{(2)}(x)=\underset{n=1}{\overset{n_m}{}}A_nf_n(x)+f_e(x).$$
(9)
The basis functions $`f_n`$ are build up of $`1/[1x]_+`$, $`\delta (1x)`$ and of powers of $`\mathrm{ln}(1x)`$, $`x`$, and $`\mathrm{ln}x`$. The coefficients $`A_n`$ are determined from the $`n_m=5`$ ($`n_m=4`$) linear equations provided by the non-singlet (singlet) moments of after taking into account the other constraints collected in $`f_e`$ in (9). The remaining uncertainties are estimated by ‘reasonably’ varying the choice of the basis functions $`f_n`$, typically considering some 20 to 40 trial functions. Finally two approximations spanning the error band are selected, except for the highest unknown $`N_f`$-contributions in (8) for which one central representative is sufficient.
This procedure is briefly illustrated in Fig. 1 for the $`N_f=0`$ part of $`P_{\mathrm{NS}}^{(2)+}(x)`$. The upper plot shows 24 trial functions. The approximations A and B emphasized in the plot have been selected, after considering also the convolution with a typical input shape shown for these two functions in the lower plot. As can be inferred from Fig. 1, the presence of the convolution in (5) considerably increases the effective accuracy of our approximations illustrated in Figs. 1 and 2: The convolutions smoothen out the oscillating large-$`x`$ differences between different approximations to a large extent. They also partly compensate the large small-$`x`$ uncertainties of $`P^{(2)}`$ present despite the $`x0`$ constraints of .
## 4 Parton densities: results
We illustrate the impact of the NNLO terms on the parton evolution by the derivatives $`\dot{q}d\mathrm{ln}q/d\mathrm{ln}\mu _f^2`$, $`q=q_{\mathrm{NS}}^+,\mathrm{\Sigma },g`$, at a reference scale
$$\mu _f^2=\mu _{f,0}^2\mathrm{\hspace{0.25em}30}\text{ GeV}^2$$
(10)
corresponding to $`\alpha _s(\mu _r^2=\mu _{f,0}^2)=0.2`$. The input densities adopted for the non-singlet case read
$$xq_{\mathrm{NS}}^+=x^{0.5}(1x)^3.$$
(11)
For the singlet distributions we employ
$`x\mathrm{\Sigma }(x,\mu _{f,0}^2)`$ $`=`$ $`0.6x^{0.3}(1x)^{3.5}(1+5x^{0.8})`$
$`xg(x,\mu _{f,0}^2)`$ $`=`$ $`1.0x^{0.37}(1x)^5.`$ (12)
The dependence of the results on the renormalization scale is presented via
$$\mathrm{\Delta }\dot{q}\frac{\mathrm{max}\dot{q}\mathrm{min}\dot{q}}{2|\mathrm{average}\dot{q}|},$$
(13)
where $`\mu _r`$ is varied over the conventional interval
$$1/4\mu _f^2\mu _r^2\mathrm{\hspace{0.25em}4}\mu _f^2.$$
(14)
The NNLO effects on the derivatives $`\dot{q}`$ and the NLO and NNLO scale uncertainties $`\mathrm{\Delta }\dot{q}`$ are shown in Figs. 3 and 4. The present inaccuracies of the NNLO results caused by the uncertainties remaining for the functions $`P^{(2)}`$ are represented by the bands spanned by the NNLO<sub>A</sub> and NNLO<sub>B</sub> curves. The central results $`\frac{1}{2}(\text{NNLO}_A+\text{NNLO}_B`$) are not shown separately.
The uncertainties of the NNLO derivatives $`\dot{q}`$ due to the approximations for $`P^{(2)}`$ are entirely negligible for $`x\stackrel{>}{}0.1`$. They increase towards very small values of $`x`$, but do not exceed $`\pm 2\%`$ above $`x10^3`$ (or a few times this number for scales $`\mu _f`$ much smaller than (10) ). Given the small size of the NNLO corrections and the weak $`\mu _r`$-dependence remaining at NNLO, one can safely estimate that contributions beyond NNLO affect the parton evolution, for $`\alpha _s\stackrel{<}{}0.2`$, by less than 1% at large $`x`$ and 2% down to $`x10^3`$.
## 5 Structure functions: formalism
The unpolarized non-singlet ($`a=1,\mathrm{\hspace{0.17em}2},\mathrm{\hspace{0.17em}3}`$) and singlet ($`a=1,\mathrm{\hspace{0.17em}2}`$) structure functions $`F_a`$ are obtained by convoluting the solutions $`q(\mu _f^2,\mu _r^2)`$ of (5) with the corresponding coefficient functions,
$`\eta _aF_a(x,Q^2)=\left[𝒞_a(a_s,L_M,L_R)q\right](x).`$ (15)
For $`L_M=\mathrm{ln}(Q^2/\mu _f^2)0`$, but $`\mu _r=\mu _f`$ ($`L_R=0)`$:
$`𝒞_a`$ $`=`$ $`c_a^{(0)}(x)+{\displaystyle \underset{l=1}{}}a_s^l\left\{c_a^{(l)}(x)+{\displaystyle \underset{m=1}{\overset{l}{}}}c_a^{(l,m)}(x)L_M^m\right\}`$
with $`𝒞_a=(𝒞_{a,q},𝒞_{a,g})`$ in the singlet case. The coefficients $`\eta _a`$ in (LABEL:ave13) include the charge factors so that $`c_{a,\mathrm{NS}}^{(0)}=c_{a,q}^{(0)}=\delta (1x)`$, whereas, of course,
$`c_{a,g}^{(0)}=0`$. The contributions $`c_a^{(l,m)}`$ fixed by renormalization-group constraints are build up of the $`c_a^{(k)}`$ and the splitting functions $`P^{(k)}`$ up to $`k=l1`$. The generalization of (LABEL:ave13) to $`\mu _r\mu _f`$ proceeds as indicated below (7).
The scaling violations of the non-singlet structure functions can be conveniently expressed in terms of these structure functions themselves (thereby removing any dependence on $`\mu _f`$), viz
$$\frac{d}{d\mathrm{ln}Q^2}F_{a,\mathrm{NS}}=𝒦_{a,\mathrm{NS}}F_{a,\mathrm{NS}}.$$
(17)
The kernels $`𝒦_{a,\mathrm{NS}}`$ are derived by differentiating (15) with respect to $`\mathrm{ln}Q^2`$ and then eliminating the quark densities using the same equation.
## 6 Coefficient functions
Besides the functions $`c_a^{(1)}`$ in (LABEL:ave13), see , also the NNLO contribution $`c_a^{(2)}`$ are known . Those expressions are rather lengthy and involve higher transcendental functions. We have thus provided compact approximations which are sufficiently accurate for any foreseeable application.
As illustrated below (Fig. 6), the impact of the functions $`c_a^{(2)}`$ (especially of the quark coefficient functions which contain large soft-gluon emission terms) is much larger than that of the splitting functions $`P^{(2)}`$ at $`x>10^2`$. The same situation is expected for the NNNLO quantities $`c_a^{(3)}`$ and $`P^{(3)}`$. Hence a good approximation to the NNNLO at large $`x`$ can be obtained by just retaining the $`c_a^{(3)}`$.
The current information on the $`c_a^{(3)}`$ comprises
* the first five even-integer moments of $`c_{2,\mathrm{NS}}^{(3)+}`$, and the first four of $`c_{2,q}^{(3)}`$ and $`c_{2,g}^{(3)}`$ ,
* the four leading large-$`x`$ terms $`\mathrm{ln}^k(1x)/`$ $`[1x]_+`$, $`k=2,\mathrm{},\mathrm{\hspace{0.17em}5}`$ of $`c_{a,\mathrm{NS}}^{(3)}`$ and $`c_{a,q}^{(3)}`$ , fixed by the results of , together with those of for $`k=2`$.
For $`c_{2,\mathrm{NS}}^{(3)}`$ only the first moment (= 0) is known from the Adler sum rule. However, the results on $`c_{2,\mathrm{NS}}^{(2)\pm }`$ indicate that the difference $`c_{2,\mathrm{NS}}^{(3)}c_{2,\mathrm{NS}}^{(3)+}`$ has a negligibly small effect. Results for the lowest moments of $`c_{\mathrm{\hspace{0.17em}3}}^{(3)}`$ will become available soon .
Focusing on the non-singlet case most relevant for $`\alpha _s`$-determinations from structure functions, we have employed the above information to derive approximations of $`c_{2,\mathrm{NS}}^{(3)}(x)`$ . The impact of their residual uncertainties is small for $`x0.2`$.
## 7 Structure functions: results
We illustrate the effect of the NNLO (and NNNLO) terms on the structure functions at
$$Q^2\mathrm{\hspace{0.25em}30}\text{ GeV}^2.$$
(18)
In (17) we employ the non-singlet input shape
$$F_{2,\mathrm{NS}}^+(x,Q^2)=x^{0.5}(1x)^3.$$
(19)
For the singlet case we fix, besides $`\alpha _s(Q^2)=0.2`$, the input parton densities (4), hence not $`F_{2,S}`$. The $`\mu _r`$-dependence of the results for $`f=F_{2,S}`$, $`\dot{F}_{2,\mathrm{NS}}d\mathrm{ln}F_{2,\mathrm{NS}}/d\mathrm{ln}Q^2`$, $`\dot{F}_{2,S}`$ and $`F_{2,S}^{}dF_{2,S}/d\mathrm{ln}Q^2`$ is presented via
$$\mathrm{\Delta }\dot{f}\frac{\mathrm{max}\dot{f}\mathrm{min}\dot{f}}{2|\mathrm{average}\dot{f}|},\frac{1}{4}Q^2\mu _r^24Q^2.$$
(20)
The singlet results are shown for $`\mu _f=\mu _r\mu `$.
The results for $`\dot{F}_{2,\mathrm{NS}}`$ are shown in Fig. 5. The uncertainty bands for the NNNLO predictions take into account both the remaining inaccuracies of the coefficient functions $`c_{2,\mathrm{NS}}^{(3)}(x)`$ and the possible effects of the splitting functions $`P_{\mathrm{NS}}^{(3)}`$. At $`0.25\stackrel{<}{}x\stackrel{<}{}0.7`$ the $`\mu _r`$-uncertainties $`\mathrm{\Delta }\dot{F}_{2,\mathrm{NS}}`$ are reduced by a factor of two (four) or more at NNLO (NNNLO). These uncertainties lead to the following estimates for the errors of $`\alpha _s(M_Z^2)`$ due to the truncation of the perturbation series:
$`\mathrm{\Delta }\alpha _s(M_Z^2)_{\mathrm{NLO}}`$ $``$ $`0.005`$
$`\mathrm{\Delta }\alpha _s(M_Z^2)_{\mathrm{NNLO}}`$ $``$ $`0.002`$ (21)
$`\mathrm{\Delta }\alpha _s(M_Z^2)_{\mathrm{NNNLO}}`$ $``$ $`0.001.`$
A 1% accuracy is achieved at the NNNLO level. As the scaling violations for the same $`\alpha _s(Q^2)`$ are stronger at NNLO and NNNLO than at NLO, higher-order fits of data on $`F_{2,\mathrm{NS}}`$ will yield somewhat lower central values of $`\alpha _s(M_Z^2)`$,
$$\alpha _s(M_Z^2)_{\mathrm{NNNLO}}\alpha _s(M_Z^2)_{\mathrm{NLO}}\mathrm{\hspace{0.17em}0.002}.$$
(22)
The results for the singlet case are presented in Fig. 6. $`F_{2,S}`$ receives large positive corrections at large $`x`$, caused by the soft-gluon parts of the quark coefficient functions. The sizeable negative NNLO corrections at small $`x`$ are dominated by the gluon contribution. It is worth noting that the positive $`1/x`$ term of $`c_{2,g}^{(2)}`$ does not dominate this correction even at $`x<10^3`$.
The $`Q^2`$-derivative of $`F_{2,S}`$ is dominated by the quark contribution at $`x>0.3`$, and by the gluon contribution at $`x<0.03`$. Thus we present the logarithmic derivative $`\dot{F}_{2,S}`$ in the former $`x`$-range, and the linear derivative $`F_{2,S}^{}`$ in the latter region. Note that the positive NNLO gluon contribution reaches 5% of the total $`|\dot{F}_{2,S}|`$ at $`x=0.5`$, enough to jeopardize purely non-singlet analyses of $`F_2^p`$ data also in the region $`x>0.3`$.
The reduced $`\mu _r`$-dependence of both $`F_2`$ and its derivatives leads to a better theoretical accuracy of determinations of the parton densities from data on $`F_{2,S}`$ and $`dF_{2,S}/d\mathrm{ln}Q^2`$ at $`Q^230\text{ GeV}^2`$: NNLO uncertainties of less than 2% from the truncation of the perturbation series are obtained for the quark density at $`10^3<x<0.5`$ and for the gluon density at $`310^3<x<0.2`$.
## 8 Summary
We have briefly discussed the evolution of unpolarized parton densities and structure functions in the $`\overline{\text{MS}}`$ scheme. Our approximate results for the 3-loop splitting functions $`P^{(2)}(x)`$ pave the way for promoting, even though only at $`x>10^3`$, global analyses of DIS and related processes to NNLO accuracy. We will also provide approximations for the 3-loop non-singlet coefficient functions $`c_{\mathrm{NS}}^{(3)}`$, thus enabling NNNLO determinations of $`\alpha _s`$ from structure functions at least at $`x0.2`$.
At very large $`x`$, $`x\stackrel{>}{}0.8`$, terms even beyond NNNLO are relevant. Here results are available from soft-gluon resummation . Progress towards the important HERA small-$`x`$ region of $`x\stackrel{<}{}10^3`$ at moderate/low $`Q^2`$ requires the full calculation of the three-loop splitting functions .
Fortran subroutines of our parametrizations of the 2-loop coefficient functions and our approximations of the 3-loop splitting functions can be obtained from neerven@lorentz.leidenuniv.nl or avogt@lorentz.leidenuniv.nl.
|
warning/0006/physics0006020.html
|
ar5iv
|
text
|
# Laser generated wake fields as a new diagnostic tool for magnetized plasmas
## I Figure captions
Fig. 1 Illustration of the discretization scheme. (a) Spectrum path discretization (b) Plasma slab discretization
Fig. 2 Example of a reconstruction. (a) Predicted wake field spectrum normalized against the maximum frequency (b) . Normalized plasma profiles. The solid line marks the assumed profile and the crosses marks the reconstructed profile.
|
warning/0006/hep-ph0006207.html
|
ar5iv
|
text
|
# Extraction of skewed parton distributions.
## 1 Definition and properties of Skewed parton distributions.
Skewed parton distributions (SPD’s) were introduced for about one decade as a generalization of both Feynman’s parton densities as well as of exclusive distribution amplitudes $`^\mathrm{?}`$. They are defined as expectation values of light-ray operators sandwiched between hadronic states $`|P_iS_i`$ with different momentum $`P_i`$ and spin $`S_i`$ dependence. At twist-two level they read in the quark sector:
$`\left\{{\displaystyle \genfrac{}{}{0pt}{}{{}_{}{}^{Q}q_{}^{V}}{{}_{}{}^{Q}q_{}^{A}}}\right\}(x,\xi ,\mathrm{\Delta }^2,\mu )={\displaystyle \frac{d\kappa }{2\pi }e^{i\kappa xP_+}P_2S_2|\overline{\psi }(\kappa n)\left\{\genfrac{}{}{0pt}{}{\gamma _+}{\gamma _+\gamma _5}\right\}\psi (\kappa n)|P_1S_1_\mu },`$ (1)
where $`x`$ is the longitudinally momentum fraction with respect to $`P_+=n(P_1+P_2)`$ \[$`n`$ is a light-cone vector which project onto the + component\], $`\xi =\mathrm{\Delta }_+/P_+`$ is the so-called skewedness parameter with $`\mathrm{\Delta }=P_2P_1`$ and $`\mu `$ is the renormalization scale of the operators. They describe the probability amplitude to find a quark with momentum fraction $`(x\xi )P_+/2`$ which goes for the DGLAP–region, i.e. $`|x||\xi |`$, into a final quark with momentum fraction $`(x+\xi )P_+/2`$ or forms together with the second quark a mesonic like state into the ER-BL–region, i.e. $`|x||\xi |`$.
More recently, it has been realized that they contain valuable non-perturbative information, which, for instance, may offer the possibility to measure the angular momentum fraction of quarks and gluons in the nucleon. A number of properties follow from first principals by means of the definition (1):
* Support properties: $`q(x,\xi )=0`$ for $`|x|>1`$.
* Polynomial property of moments: $`_1^1𝑑xx^jq(x,\xi ,\mathrm{\Delta }^2)=_{k=0}^jq_{jk}(\mathrm{\Delta }^2)\xi ^k`$.
* Lowest moment is given by form factors: $`q_{00}=P_2S_2|n^\mu {}_{}{}^{Q}J_{\mu }^{I}|P_1S_1/P_+.`$
* Hermiticity as well as time reversal invariance provide: $`q(x,\xi )=q(x,\xi )`$.
* Evolution equation follows from renormalization group invariance:
$`\mu ^2\frac{d}{d\mu ^2}q(x,\xi ,\mathrm{\Delta }^2,\mu )=_1^1\frac{dy}{|\xi |}V(x/\xi ,y/\xi ;\alpha _s(\mu ))q(y,\xi ,\mathrm{\Delta }^2,\mu )`$.
* Inclusive connection: $`q(x,\mu ^2)=q(x,\xi =0,\mathrm{\Delta }^2=0,\mu ^2)_{|S_2S_1}`$, where $`q(x,\mu ^2)`$ is the usual parton density.
* Exclusive connection: $`\mathrm{\Phi }(x,\mu ^2)=q(x,\xi =1,\mathrm{\Delta }^2=M^2,\mu ^2)_{P_2,S_2|0|}`$, where $`|P_1,S_1`$ is now a meson state and $`\mathrm{\Phi }`$ is the distribution amplitude.
## 2 Models for skewed parton distributions.
As we see from the list given above, the SPD’s give us a link between exclusive quantities like electromagnetic form factors and parton densities measured in inclusive reactions. Based on model assumptions, there exist different proposals for the SPD’s with quite different characteristics. For instance, the bag-model predicts a valence quark distribution which is rather independent on the skewedness parameter $`^\mathrm{?}`$. While the chiral solution model in the large $`N_c`$ limit takes into account the Dirac sea and, thus, predicts a more complex shape of SPD’s containing zeros and a strong skewedness dependence $`^\mathrm{?}`$. Note that the typical sclale for these predictions is very low, i.e. $`Q_00.4`$ GeV and $`Q_00.6`$ GeV, respectively.
Other suggestions are inspired by the inclusive connection and are based on different mappings of the parton densities to the SPD’s, for instance, equating them, mapping the Mellin moments to the conformal ones due to an integral transformation, or expressing the SPD’s in terms of double distributions $`^\mathrm{?}`$. Thereby, it is assumed that the $`\mathrm{\Delta }^2`$ dependence is factorized, which is certainly true up to logarithmic corrections for large $`\mathrm{\Delta }^2`$ of a few GeV. Note that in the forward limit the spin flip part vanishes, and therefore, further assumptions are needed. For very small values of $`\xi `$, the non-spin flip part in the DGLAP region is essentially determined by the forward parton distributions.
These prescriptions have to be supplemented by the scale at which they are applied. It seems to be a good idea to use a low input scale that is typically for non-perturbative model calculations. Afterwards one evolves the SPD’s to the scale that is used in hard scattering experiments, i.e. higher than at least one GeV. The advantage is twice: i) one study the stability under evolution and ii) perturbative QCD information are taken into account. Note that evolution to an asymptotic large scale predicts, independently on the initial conditions, that the SPD’s are concentrated in the ER-BL region and vanish for $`x=\pm \xi `$. However, this fact is not of practical relevance for the scales accessible in experiments.
## 3 Experiments to access skewed parton distributions.
There is a growing interest to access these non-perturbative functions in lepton-hadron scattering experiments, where the virtuality of the intermediate photon has to be larger than of at least one GeV. On the theoretical side there are formal factorization proofs for the following hard processes to leading twist-two accuracy in a perturbative calculable hard-scattering part and universal SPD’s: deeply virtual Compton scattering (DVCS) $`e^\pm pe^\pm p\gamma `$, exclusive meson $`M`$ production $`e^\pm pe^\pm BM`$<sup>a</sup><sup>a</sup>a Since $`B`$ denotes an arbitrary baryon, the SPD’s defined in eq. (1) are generalized to “off-diagonal” ones. Note also that factorization is only proofed for longitudinal polarized photon., and exclusive lepton pair production $`e^\pm pe^\pm pl^+l^{}`$. In the first two cases the amplitude reads in leading order:
$`𝒜(\xi ,\mathrm{\Delta }^2,Q^2){\displaystyle \underset{j=u,d,s}{}}{\displaystyle _1^1}𝑑x\left({\displaystyle \frac{C_j}{x\xi +iϵ}}+{\displaystyle \frac{\overline{C}_j}{x+\xi iϵ}}\right)q_j(x,\xi ,\mathrm{\Delta }^2,Q^2),`$ (2)
where the coefficients $`C_j,\overline{C}_j`$ depend on the considered process. If $`𝒜`$ could be very precisely measured as function of $`\xi `$ and $`Q^2`$, the deconvolution of this formula exists in principal. Practically, one has to compare the model predictions with the experimental data or defines characteristic functions that together with eq. (2) can distinguish between different models:
$`R={\displaystyle \frac{\mathrm{Re}𝒜(\xi ,\mathrm{\Delta }^2,Q^2)}{\mathrm{Im}𝒜(\xi ,\mathrm{\Delta }^2,Q^2)}},S=1{\displaystyle \frac{\mathrm{PV}_1^1𝑑x\frac{1}{x\xi }\mathrm{Im}𝒜(x,\mathrm{\Delta }^2,Q^2)}{(\pi )\mathrm{Re}𝒜(\xi ,\mathrm{\Delta }^2,Q^2)}}.`$ (3)
The latter one can be considered as a measure for the skewedness dependence.
Unfortunately, this perturbative leading order analysis can be spoiled by the size of perturbative as well as higher twist corrections. The first ones have been considered for DVCS in the valence quark region at next-to-leading order. It turns out that the corrections due to evolution are small, i.e. about 10% or less, while the corrections to the hard scattering amplitude depend on the chosen model and can be of the order of 50% or even more $`^\mathrm{?}`$.
|
warning/0006/astro-ph0006272.html
|
ar5iv
|
text
|
# The Effect of Diffusion on Pulsations of Stars on the Upper Main Sequence.
## 1 Introduction
Until very recently the theory for the formation of chemically peculiar (CP) stars and the observations of variable stars agreed that CP stars were not variable and variable stars were chemically normal. A few exceptions were known for many years but their rarity and their mild chemical anomalies did not challenge the standard picture.
In recent years this picture has evolved considerably and the dichotomy between CP stars and variable stars is no longer clear-cut. Indeed, many types of CP stars are now known to exhibit various types of pulsations. Significantly, most of the variable CP stars belong to the magnetic Ap and Bp families. But some non-magnetic CP stars are known to be variable and it has been claimed that in some cases the observed variability cannot be easily reconciled with the theory of diffusion for CP stars.
In this paper we examine how recent progress in modeling Am stars affects our understanding of the problem of variability in A-type CP stars. We first review the current observational and theoretical pictures for variability in CP stars. The improved models of diffusion in A stars are then presented. The linear nonadiabatic oscillation equations are solved in selected models of A-type stars, both with and without diffusion. These results highlight the differential effect of diffusion on the stability of the models, and allow to draw conclusions on how well the models with diffusion can be reconciled with observations.
## 2 Metallicism and pulsations: observations
CP stars are found almost everywhere in the HR diagram. Some of these stars are evolved and their observed abundance peculiarities reflect nuclear processes. Others are compact objects in which diffusion is well established (Michaud & Fontaine MichaudFontaine84 (1984)). Many of these objects are variable and they are very favourable for asteroseismology.
We concentrate mainly on main-sequence stars in which diffusion is thought to be the principal cause of abundance anomalies. These are found from early F-type (Fm) stars to late B-type (HgMn) stars, including many varieties of A stars \[Am, Ap, $`\lambda `$ Bootis, $`\rho `$ Puppis\]. Within these spectral classes many types of variable stars are also found \[$`\gamma `$ Doradus, several classes of $`\delta `$ Scuti, and Slowly Pulsating B (SPB) stars\].
In A-type stars, almost 70% of non-CP stars are $`\delta `$ Scuti variables at current levels of sensitivity. Most non-variable A stars are Am stars.
Amongst the CP stars which happen to be pulsating, the most conspicuous are the “rapidly oscillating Ap stars” (roAp) first discovered by Kurtz (1978a ). These are found amongst the coolest of the magnetic Ap stars which exhibit large abundance anomalies of many heavy elements. They have generated considerable interest because of the relatively large number of observed overstable modes which makes them promising objects for asteroseismology. See, for example, Matthews (Matthews91 (1991)) for a review.
In non-magnetic stars, however, variability and anomalous abundances are found in very few stars simultaneously. Over the years some mildly metallic stars have been found to exhibit some variability. Baglin (Baglin72 (1972)) suggested that if diffusion is the cause of the Am phenomenon, Am stars should not pulsate. Some mild Am stars and evolved CP stars were then already known to be variable (Kurtz Kurtz76 (1976)) but all classical Am stars which were thought to be variable in the early 70’s were subsequently found to be stable. Since then Kurtz and his collaborators have been the principal investigators in the search for variable CP stars. They have assembled a short list of metallic stars, most if not all of them fairly evolved, which are also $`\delta `$ Scuti-type pulsators (Kurtz 1978b , Kurtz84 (1984), Kurtz89 (1989); Kurtz et al. Kurtzetal95 (1995); Kurtz Kurtz00 (2000)).
Amongst these stars, some have been thought to be problematic. In particular, Kurtz (Kurtz89 (1989)) claimed to have discovered a pulsating classical Am star. The problem lies in its apparently large abundance anomalies, not typical of the other known variable non-magnetic CP stars. Also, the variable evolved Am star HD40765 has been considered by Kurtz et al. (Kurtzetal95 (1995)) to be problematic because of the possibly large surface velocities involved. A point that has been made repeatedly is that the large velocities caused by pulsations, estimated in this star to be of the order of $`14\mathrm{km}\mathrm{s}^1`$ at the surface, might generate turbulence in the interior which would in turn hinder the formation of the required surface-abundance anomalies.
In addition to variable CP stars there are some $`\delta `$ Scuti stars that are particularly interesting from our point of view. First there are some otherwise seemingly run-of-the-mill $`\delta `$ Scuti stars<sup>1</sup><sup>1</sup>1Amongst these is 20CVn which was classified as $`\delta `$ Delphini by Kurtz (Kurtz76 (1976)) in which peculiar abundances (Russell Russell95 (1995); Rachkovskaya Rachkovskaya94 (1994)) superficially consistent with what is expected as a result of diffusion seem to have been found.
Second, the high-amplitude $`\delta `$ Scuti stars are interesting because they are characterized by very small $`v\mathrm{sin}i`$ (Solano & Fernley SolanoFernley97 (1997)). As CP stars are mostly slow rotators ($`v_{\mathrm{rot}}<100\mathrm{km}\mathrm{s}^1`$), one should find different velocity distributions in CP stars and in $`\delta `$ Scuti stars. Indeed, observations show that $`\delta `$ Scuti stars are on the average fairly fast-rotating stars, with the exception of the high-amplitude stars. In other respects the high-amplitude $`\delta `$ Scuti stars do not differ from their more common low-amplitude counterparts. They are, however, known to be evolved stars and as such might not feature significant abundance anomalies.
Although observations do not completely rule out variability in Am stars, they do pose rigorous constraints on them. Either it is an extremely rare occurrence or pulsations are of extremely low amplitude. An extensive search of the Hipparcos database by many authors revealed many new variable stars (Aerts et al. Aertsetal98 (1998); Waelkens et al. Waelkensetal98 (1998); Paunzen & Maitzen PaunzenMaitzen98 (1998)). Significantly, all the newly discovered variable CP stars were found to be magnetic (Ap or Bp). All the other variable stars found in this way are of the known families of variable stars, i.e., $`\gamma `$ Doradus, SPB and a few $`\beta `$ Cephei stars. The expected sensitivity of these surveys ranges from as low as 3 to 55 mmag depending on the brightness of the object (Eyer & Grenon EyerGrenon98 (1998)).
The CP stars found most recently to be variable are $`\lambda `$ Bootis stars in which nonradial pulsations have been detected (Paunzen et al. Paunzenetal98 (1998)).
## 3 Metallicism and pulsations: theory
Elements migrate with respect to each other because of differential forces mostly due to inward gravity and outward radiative pressure. This segregation of different atomic species is what is termed here diffusion. Diffusion is a rather fragile process because typical diffusion velocities are of the order of fractions of cm per second. Large-scale motion of matter or turbulence quickly overwhelm diffusion and homogenize the chemical composition outside of nuclear burning regions. Diffusion is efficient only in stars where the competing processes are weak. This explains why CP stars are mainly of spectral types ranging from early F to late B-type stars. In these spectral types the surface convection zone is thin enough and the mass loss rates small enough (a mass loss of around $`10^{14}`$ M$`{}_{\mathrm{}}{}^{}\mathrm{yr}_{}^{1}`$ is large enough to remove all anomalies) to permit the development of significant abundance anomalies. As a result, CP stars are typically relatively unevolved, slowly rotating stars. There are some CP stars, Ap or $`\lambda `$ Bootis stars, for example, in which other factors come into play.
In the standard picture for FmAm, HgMn and Ap stars, the observed abundance anomalies of heavy elements develop as a consequence of the settling of helium. As it disappears from the superficial convection zone it can no longer provide the opacity to sustain convection in the He II zone and the result is a much thinner convection zone due to the ionization of H I. At this depth, the radiative forces on the various heavy elements are compatible with the pattern of surface-abundance anomalies. This model agrees qualitatively with observations but requires additional assumptions to obtain a quantitative agreement as the predicted anomalies are generally too large. No calculation based on this model has ever reproduced the abundances of individual Am stars.
The major consequence of this model for stellar stability is that helium is no longer present to excite pulsations typical of the classical instability strip. Many studies related to this problem have been carried out. The most thorough, by Cox et al. (Coxetal79 (1979)), showed that variability is possible with a low helium content but that the width of the instability strip decreases as the helium abundance decreases. Their conclusion was that classical Am stars should not be variable but that metallicism and variability were not mutually exclusive in the red part of the classical instability strip if the surface helium abundance fell marginally below 0.1 but without dropping to a very low value.
When the diffusion of heavy elements is included in a consistent fashion this picture changes drastically. These models are dubbed here The New Montreal Models (NMM).
Magnetic Ap stars and $`\lambda `$ Bootis stars will be ignored at this time. We note in passing that for roAp stars a few models have been proposed based on the so-called $`\kappa `$ mechanism, where the dominant driving comes from the effects of the opacity $`\kappa `$. As only microscopic diffusion can explain the observed abundances at present, the proposed models include its effects. Amongst the possibilities are: the replenishment of superficial helium by advection from mass loss (Vauclair & Dolez VauclairDolez90 (1990)), the replacement of helium by silicon pushed in the driving region by radiative levitation (Matthews Matthews88 (1988)), and hydrogen overabundances in the H I ionization zone as a result of helium settling (Dziembowski & Goode DziembowskiGoode96 (1996)). As for $`\lambda `$ Bootis stars, the currently preferred but so far unproven accretion models for these stars assume that the helium abundance would remain normal in the driving region (Turcotte & Charbonneau TurcotteCharbonneau93 (1993)), accounting for the necessary opacity for the $`\kappa `$ mechanism to work.
## 4 The New Montreal Models
### 4.1 New results with the NMM
The NMM include the diffusion of all major elements up to nickel consistently by using monochromatic opacity tables of the Livermore group (Iglesias & Roger OPAL96 (1996)) to compute the opacity and the radiative forces accurately at all points in the star and for the local chemical composition; the basic procedures were outlined by Turcotte et al. (1998a ) and Richer et al. (Richeretal98 (1998)).
The evolution of the abundances and of the structure is completely consistent through the opacity. One very important property of these models is the presence of a convectively unstable zone around $`\mathrm{200\hspace{0.17em}000}`$ K where iron-group elements dominate the opacity. This convection zone appears naturally as the consequence of abundance changes if they are large enough. The NMM then assume that this deeper convection zone is connected to the helium and hydrogen convection zones through convective overshoot.
The large depth of the convective mixing relative to standard models for chemically peculiar stars results in much smaller surface-abundance variations, more in line with observations. Richer, Michaud & Turcotte (RMT00 (2000)) showed that the radiative forces at the base of the iron convection zone follow the correct pattern over a relatively narrow region for the Am signature to be recovered without additional assumptions. A quantitative agreement with observed abundances for Am stars does require additional turbulence below the superficial convective envelope. In the NMM, the mixing necessary to reproduce the abundances observed in Am stars prevents the formation of this convection zone.
A significant result in the context of stellar pulsations is that helium is still substantially present in the He II driving region in these new models. Also, as the opacity in the vicinity of the so-called “metal opacity bump”<sup>2</sup><sup>2</sup>2 We shall refrain from using this expression and otherwise use the expression “iron-peak-element opacity bump”, as those elements dominate the opacity there is increased relative to standard models, one can expect that this region will contribute to the excitation of longer-period modes. On the other hand, in some extreme cases this region might be convective in contrast to standard models, which would reduce the radiative flux and by consequence the driving in that region.
We shall examine the consequence of these results on the possible instability of Am stars. As a complement, we shall also estimate the effect of mild abundance anomalies on predicted pulsations of $`\delta `$ Scuti stars.
### 4.2 The basic properties of the NMM
The NMM include the detailed diffusion of 21 major elements from H through Ni plus several light elements and isotopes for a total of 28 species. The opacity data used in the evolution code and in the following analysis are the OPAL monochromatic opacity tables (Iglesias & Rogers OPAL96 (1996)) which allow us to calculate accurate Rosseland mean opacities and radiative forces for any peculiar chemical composition necessary. At low temperatures, for which the OPAL data is lacking, we supplement them with the Kurucz (Kurucz91 (1991)) opacity tables. Although it would be desirable to take into account changes in the composition of individual heavy elements in the atmosphere, the transition occurs at such low temperatures that it might be of little consequence for the stellar interior.
The models incorporate standard procedures for the equation of state and the nuclear reaction rates. The mixing-length formalism for convection is used and is calibrated using the Sun (Turcotte et al. 1998a ). All models have an identical, homogeneous, initial chemical composition as specified by Turcotte et al. (1998b ). All models are one-dimensional, non-rotating, and non-magnetic.
Individual models for a given stellar mass only differ in the parameters adopted for the coefficient of turbulent diffusion. In all calculations presented in this paper, the coefficients are chosen so that the zone mixed by turbulence goes from the surface to somewhat below the iron convection zone. The coefficient of turbulent diffusion is modeled with the following three-parameter expression
$$D_\mathrm{T}=D_0\left(\frac{\rho _0}{\rho }\right)^n,$$
(1)
where the free parameters are $`D_0`$, $`\rho _0`$ and $`n`$. The evolution of the abundances is very sensitive to the depth of the mixing but not so much to the profile of $`D_\mathrm{T}`$. The models are named with reference to the number $`R`$ which specifies the ratio of $`D_\mathrm{T}`$ to the coefficient of atomic diffusion of helium $`D_{\mathrm{He}}`$ at the point where the density $`\rho `$ is equal to the reference value $`\rho _0`$. For example, model 1.90R1000-2 is a 1.90 M star with $`R=1000`$ and $`n=2`$; for simplicity, ‘10K’ is used to refer to models with $`R=\mathrm{10\hspace{0.17em}000}`$. In all the models discussed in this paper $`\rho _0`$ is $`8\times 10^6`$g cm<sup>-3</sup>. The reader is referred to Richer et al. (RMT00 (2000)) and Richard et al. (RMR00 (2000)) for further details.
For every stellar mass examined, one model that does not include any effects of diffusion is also included. For computational efficiency this comparison model uses the mean Rosseland opacity tables of Iglesias & Rogers (OPAL96 (1996)). Assuming that there is no separation of elements in the star implies that some unspecified mixing is necessarily assumed. This mixing is required to be large and deep enough to keep superficial regions at a constant chemical composition without mixing too deeply, to avoid dredging up nuclearly processed matter to the surface. They are named according to the mass and are labeled with the tag “ND” (e.g., 1.90-ND).
## 5 Diffusion and the $`\kappa `$ mechanism
To determine the frequencies of modes of oscillation for a star requires only that we solve the adiabatic equations. Solving the full nonadiabatic equations of stellar oscillation allows us to calculate the growth rates of the modes, and hence to determine which of the modes are overstable; also, by considering the work integral we can investigate the contributions of the different parts of the star to the excitation and damping of the mode.
The nonadiabatic oscillation package used was generously provided to us by W. Dziembowski and follows the procedure first described by Dziembowski (Dziembowski77 (1977)). We are mainly concerned here with excitation via the $`\kappa `$ mechanism on which abundance variations have a direct impact. We note, however, that the present calculations lack a good modeling of the effect of convection; this must be kept in mind in the analysis of the results.
The physics of the $`\kappa `$ mechanism has been reviewed extensively (e.g. Cox Cox80 (1980); Unno et al. Unnoetal79 (1979); Gautschy & Saio GautschySaio95 (1995)). As a quick reminder, generating pulsations in a star requires that the energy gained by an oscillation mode over a complete cycle be larger than the energy lost. We are then looking for a positive net work over the entire star over one cycle. In the case of the $`\kappa `$ mechanism, the energy is transferred from the outward radiation flux to the oscillation mode via the opacity. A mode becomes overstable by this mechanism if the opacity profile and its derivatives have the right features.
Following Unno et al. (Unnoetal79 (1979)), from the definition of the work ($`W`$) as the variation of the kinetic energy ($`E`$) over a cycle,
$$W=\frac{\mathrm{d}E}{\mathrm{d}t}dt,$$
(2)
one can write
$$W=\frac{\pi }{\omega }_0^{M_r}\frac{\delta T}{T}\delta [ϵ_\mathrm{N}\frac{1}{\rho }\dot{(}F_\mathrm{R}+F_\mathrm{c})]\mathrm{d}M_r,$$
(3)
where $`\delta `$ denotes Lagrangian perturbations. Also, $`\omega `$ is the (angular) oscillation frequency, $`T`$ is temperature, $`M_r`$ is the mass interior to the radius $`r`$, $`ϵ_\mathrm{N}`$ is the nuclear energy generation rate, and $`F_\mathrm{R}`$ and $`F_\mathrm{c}`$ are the radiative and convective fluxes. If one neglects the contribution from the nuclear ($`ϵ_\mathrm{N}`$) and convective terms ($`F_\mathrm{c}`$), and only keeps the perturbation of the radiative flux ($`F_\mathrm{R}`$), one can isolate the contribution of the $`\kappa `$ mechanism to the driving of a given mode of oscillation. To obtain a simple estimate of this contribution to the work integral we make the quasi-adiabatic approximation (i.e., evaluate the work integral by means of adiabatic eigenfunctions), and furthermore assume that the adiabatic thermodynamic derivatives $`\mathrm{\Gamma }_1`$, $`\mathrm{\Gamma }_3`$ and $`_{\mathrm{ad}}`$ are constant. Then the work done by the $`\kappa `$ mechanism is proportional to
$$\left(\frac{\delta T}{T}\right)^2\frac{\mathrm{d}}{\mathrm{d}r}\left[\left(\kappa _T+\frac{\kappa _\rho }{\mathrm{\Gamma }_31}\right)L_r\right]dM_r,$$
(4)
where $`L_r`$ is the luminosity at $`r`$, and $`\kappa _T=(\mathrm{ln}\kappa /\mathrm{ln}T)_\rho `$, $`\kappa _\rho =(\mathrm{ln}\kappa /\mathrm{ln}\rho )_T`$; thus regions where
$$\frac{\mathrm{d}}{\mathrm{d}r}\left(\kappa _T+\frac{\kappa _\rho }{\mathrm{\Gamma }_31}\right)>0$$
(5)
contribute to the excitation. Local increases in the logarithmic derivatives of $`\kappa `$ are necessary and a decrease in $`\mathrm{\Gamma }_31`$ in partial ionization zones of a dominant species (H or He) is helpful. It also follows that regions where the gradients of $`\kappa _T`$ and $`\kappa _\rho `$ are negative contribute to damping of the pulsation. The $`\kappa _T`$ term usually dominates over the other term.
The numerical results reported in the text, including the growth parameters and work integrals, are computed using the full nonadiabatic procedure of the Dziembowski code.
In order that the excitation by the $`\kappa `$ mechanism should not be cancelled by damping elsewhere, it is necessary that the driving region lie in the so-called transition zone between the quasi-adiabatic and nonadiabatic regimes; in that case, the oscillations are strongly nonadiabatic outside the driving region, and this part of the star therefore does not contribute to the damping, giving rise to net driving. This leads to an approximate relation between the period $`\mathrm{\Pi }`$ of a given mode of pulsation and the position of the transition region in a star (Cox Cox80 (1980)):
$$\frac{c_vT_{\mathrm{tr}}\mathrm{\Delta }M_{\mathrm{tr}}}{L\mathrm{\Pi }}1,$$
(6)
here $`\mathrm{\Delta }M_{\mathrm{tr}}`$ is the mass outside the transition region, $`\mathrm{}_{\mathrm{tr}}`$ is the average over that part of the star, $`c_v`$ being the specific heat, and $`L`$ is the luminosity.
The normalized growth rate is defined as
$$\eta =\frac{\mathrm{d}W}{\mathrm{d}\mathrm{log}T}\mathrm{d}\mathrm{log}T/\left|\frac{\mathrm{d}W}{\mathrm{d}\mathrm{log}T}\right|\mathrm{d}\mathrm{log}T.$$
(7)
In this formulation, $`\eta `$ varies from $`+1`$, if there is driving in the entire star, to $`1`$, if there is damping in the entire star. The value of zero defines neutral stability.
Diffusion affects the $`\kappa `$ mechanism by decreasing driving from helium in favour of driving from metals. As a consequence of Eq. (6), the pulsation period of the unstable modes depends on the depth of the driving region. During a star’s evolution the helium ionization zone gradually shifts deeper in the star, thereby increasing the period of the observed pulsation modes. Additionally, as the driving in the deeper iron-peak driving region increases while the driving due to helium decreases as a result of diffusion, one might expect the observed pulsation periods to shift to even longer periods. The effect of abundance variations on the opacity profiles is discussed below for selected models (see Fig. 3).
## 6 Opacity and its derivatives
We recomputed many thermodynamic quantities and the opacities in the process of preparing the models for input to the nonadiabatic oscillation package. Every effort has been made to be consistent with the procedures followed in the evolution code. The most tricky operation at this point is the determination of accurate opacity derivatives. The opacities are interpolated linearly and smoothed locally in the evolution code. As first and second derivatives of the opacity are needed in the oscillation package a more refined interpolation procedure had to be adopted to guarantee smoother derivatives. We chose the Houdek (Houdek & Rogl HoudekRogl96 (1996)) routines which use two-dimensional rational splines. For each mesh point of the models we constructed a $`7\times 7`$ opacity grid in the $`(\rho ,T)`$ plane in which the splines were fitted and the opacity derivatives were determined.
## 7 The models of A stars
Amongst the numerous NMM models some were selected to provide a satisfactory sampling of the instability strip and of Am stars. The selected models of 1.9, 2.0 and 2.2 M are summarized in Tables 1, 2, and 3 respectively. (The naming convention for the models was described in the discussion relating to Eq. 1.) Of these, the models with $`R=1000`$ and $`n=2`$ reproduce reasonably well the abundances of Am stars (Richer et al. RMT00 (2000)). They are not optimal, however, and models with $`R=1000`$ and with shallower turbulence, $`n=3`$ or 4, would be more accurate. We feel that using such models would not alter the essence of the interpretation of the results.
## 8 $`\delta `$ Scuti-type pulsations in Am stars
Most pulsating A stars lie in the He II instability strip. Consequently, most attempts to model the pulsations of chemically peculiar A stars have centered on how helium can remain or be replenished in the He II zone, or possibly which element can replace it as the motor of the excitation mechanism. As previously discussed, an important characteristic of the NMM is the relatively high superficial helium abundance remaining throughout the evolution in Am stars (see Sect. 4). This section aims at answering some questions raised by these new models: Is helium sufficiently abundant to generate variability in Am stars? Do the NMM clash with our current observational understanding of metallicism and pulsations? Does the larger driving in the iron-peak opacity bump play a role in exciting g modes in CP stars?
### 8.1 The variability of an Am star across the instability strip
The evolution of a 1.9 M star from the ZAMS to hydrogen exhaustion in the core spans almost the full width of the instability strip as shown in Fig. 1. Studying the evolution of instability in models for such a star as it evolves gives a first idea of the behaviour of Am stars in general. The luminosity dependence of the instability strip implies that the lower part of the instability strip will be populated by young stars and the upper part by generally more evolved stars. This should be taken into account if one attempts to extrapolate the following results to stars of different masses.
Tables 4 and 5 show the general pattern of overstability for radial ($`l=0`$) and nonradial ($`l=1`$) p modes, respectively, in an evolving 1.9 M model with and without diffusion. The model with diffusion shown here is an adequate representation of a classical Am star (see Fig. 22 of Richer et al. RMT00 (2000)).
First, one can see that no overstable modes, radial or otherwise, are found in the young model with diffusion (Am star) whereas high-order modes are predicted to be overstable in the model without diffusion (a standard model for $`\delta `$ Scuti stars, i.e., a homogeneous envelope with solar composition). As the stars age, the lower-order modes are gradually more excited and eventually the fundamental mode becomes overstable in both models. In the model with diffusion, the superficial helium abundance reaches its lowest value at 750 Myr where it is 0.114 by mass. The model is stable at that time. The more evolved models tend to be more unstable. At an age of 1.098 Gyr, for example, the helium mass fraction rises to 0.126. The small difference in the helium abundance for these two models suggests that the pulsations in evolved stars occur without relying on the dredge-up of helium nor on hypothetic mass-loss but only as a result of their evolution. This result is in agreement with Cox et al. (Coxetal79 (1979)) and earlier work which showed that stars would be expected to be variable in the red part of the instability strip if their helium abundance were of the order of 0.1 or higher.
As far as nonradial modes are concerned, the same overall behaviour as for radial modes is repeated, although the labeling of the modes presents some problems. Nonetheless, a presentation such as Table 5 illustrates the general evolution of overstability in nonradial modes; here the mathematical mode orders, defined such that the mode order of a given mode is constant during the evolution of the model, are shown. This is not directly related to the physical nature of the mode, however. For p modes, including radial modes, the dimensionless frequency<sup>3</sup><sup>3</sup>3e.g. $`\sigma `$, defined by $`\sigma ^2=R^3\omega ^2/(GM)`$ where $`M`$ and $`R`$ are total mass and surface radius of the star, and $`G`$ is the gravitational constant., normalized with the dynamical timescale of the star, is approximately constant during the evolution. In contrast, the dimensionless frequency increases for g modes, which are strongly affected by the composition structure in the stellar interior. Where the frequency of a g modes meets the frequency of a p mode, an avoided crossing takes place, at which occurs a shift in mode order for a given frequency, or conversely a shift in frequency for a given mode order (e.g. Osaki 1975). Thus, in contrast to radial modes, there is no firm correspondence between mode order and dimensionless frequency through the life of a star. Even so, a table corresponding to Table 5 but based on a definition of the order more closely related to the physical nature of the modes, would have shown a similar overall pattern. The more evolved models feature overstable nonradial p and g modes. In the most evolved model, the last model of 1.90-ND, the overstable modes all have the physical nature of g modes, spanning periods of 1 to 3.5 hours. Some modes, at avoided crossings, are mixed modes which share p and g mode characters.
Two models with a turbulence depth bracketing that of the model discussed so far are also available at this mass. In the model with shallower mixing (1.90R300-2) only one mode is found to have a positive growth rate, the $`l=0,n=3`$ mode for the model at 1.098 Gyr. No overstable nonradial modes have been found. The minimum helium abundance at 750 Myr is 0.0734 and increases to 0.0871 at 1.098 Gyr for this star. In the model with deeper mixing (1.90R10K-2) the minimum helium abundance reached is higher (0.180) and the youngest age at which overstability is found is 850 Myr. This confirms that the blue edge of the instability strip is sensitive to the helium abundance, which has long been established (Cox et al. Coxetal79 (1979) and references therein).
In an attempt to give a better picture of the evolution of the excitation of various modes in the star, Fig. 2 shows the change in the growth rate as a function of time for 3 different radial modes in all four models of the 1.9 M star. This figure reflects the preceding discussion in the sense that the growth rates of low-order modes increase with passing time and that high-order modes become less excited. It also shows that the effect of diffusion is much stronger for higher-order modes. The different evolution of the growth rates of the $`l=0,n=8`$ mode after 700 Myr in the models with and without diffusion is noteworthy. This will be discussed again shortly.
The detailed effect of diffusion on the opacities of A stars is illustrated in Fig. 3 where the opacity is shown for the same four models at 670 Myr. The other panels show the evolution of the argument of Eq. (5) for ages of roughly 100, 670 and 1000 Myr. The most striking effect is the increase around $`\mathrm{log}T5.3`$ caused by the heightened iron-peak opacity bump. The effect of the settling of helium around $`\mathrm{log}T4.5`$ appears minor in the figure, but this ends up as being the dominant effect.
Considering just Fig. 3 can give a false impression of the importance of diffusion in different regions. A truer impression is obtained from Fig. 4 where the growth rates for radial modes are shown together with the work integrals for two of those modes, the $`l=0,n=1`$ and the $`l=0,n=8`$, again at 100, 670 and 1000 Myr.
Evidently, the work integrals for different models are close to each other in most regions of the star. The differences are larger for the $`n=8`$ modes which is reflected in the growth rates. The driving due to helium at $`\mathrm{log}T4.7`$ is the feature most obviously affected by diffusion, decreasing as helium settles out of the convection zone. Diffusion has some marginal effect in the iron-peak opacity bump for the $`n=1`$ mode where it partially compensates the effect of He settling but it has no effect whatsoever for the higher-frequency $`n=8`$ mode. The differences from model to model are more easily seen in the growth rates which have already been discussed.
Comparing the evolution of driving in the models with and without diffusion in Fig. 4 explains the different evolution of the growth rates of low and high-order modes in the presence of diffusion presented in Fig. 2. In Fig. 4, the height of the peaks due to H I, at $`\mathrm{log}T4.1`$, increases with time at a rate quite similar in all models shown. However, the evolution of the peak due to helium, at $`\mathrm{log}T4.7`$, differs significantly for models with and without diffusion. As discussed previously diffusion causes a decrease in helium driving, but it has little effect on driving from hydrogen in the H I ionization zone. The relative contribution of hydrogen to the net excitation of a mode is larger in higher-order modes, as illustrated by comparing the evolution of the $`n=8`$ and the $`n=1`$ modes in Fig. 4, and is also larger in models with diffusion because of the lower excitation from helium.
As a result, with the passage of time, the net excitation of a high-order mode is less affected by diffusion than a low-order mode. This can explain why, for higher-order modes shown in the bottom panel of Fig. 2, the growth rates are less adversely affected by diffusion than would have been expected from a simple extrapolation of the evolution in the young stars.
There are some apparent numerical effects in the work integrals that are caused by the manipulation of the opacity tables. First, there are obvious spurious oscillations of the work integrals for the $`n=1`$ mode. Although these oscillations change the growth rates slightly, one still gets an accurate assessment of the differential effect of diffusion as the numerical features are the same for all models. Also, the apparent temperature shift towards the surface at 1.0 Gyr between the 1.90-ND and other models is probably an age effect. The model without diffusion is slightly more evolved than the models with diffusion at a given evolution time because the former was computed using the Livermore Rosseland mean opacity tables rather than the monochromatic opacity tables.
Fig. 5 illustrates the dependence of the growth rates on the surface helium abundance for different ages of the 1.9 M models. Three modes were selected, the fundamental ($`l=0,n=1`$), a higher-order p mode ($`l=0,n=8`$) and a g mode ($`l=1,n=14`$) at 100, 670 and 1000 Myr. As one could have predicted, both p modes are very sensitive to the depletion of helium at the surface. The higher the frequency of the mode, the larger the slope of the correlation. For the g mode, there is a slight anti-correlation which shows that it is weakly correlated to the level of driving in the iron-peak opacity bump.
The periods of the modes predicted to be overstable are summarized in Table 6. First one can verify that the periods are within the observed period range for $`\delta `$ Scuti stars. Their periods are known to range from 30 minutes to 6 hours. The young models discussed here feature periods of 20 minutes while older models are characterized by longer periods, roughly between one and four hours. While we find periods well below the currently observed lower limit for $`\delta `$ Scuti stars, we know of no reason to exclude their existence outright.
Diffusion has little effect on the periods of oscillation. There is a systematic trend for diffusion to increase slightly the period of a given mode. The relative difference between the same modes in the models shown in Table 6 is of the order of a few per cent with no clear dependence on the order of the modes. Actually, most of the difference could be consistent with only evolutionary effects related to the use of Rosseland opacity tables for the ND model. If we compare the periods of the radial modes of models 1.90R1000-2 and 1.90R10K-2, which are both computed with the monochromatic opacity tables, we find period differences typically only a fifth of the differences between the models calculated with differing opacities.
### 8.2 The effect of diffusion on the instability strip
The effect of the settling of helium on the width of the instability strip has already been studied by Cox et al. (Coxetal79 (1979)). They found that the main effect was a reduction of its width, the blue edge shifting towards the red edge, eventually leading to the disappearance of instability when helium was sufficiently depleted. As has been touched upon in the preceding section, the present models seem to follow the same pattern.
In order to gain a somewhat more complete vision of the effect of diffusion on the width of the instability strip we need to examine a larger sample of models of A-type stars with and without diffusion. For this purpose, additional models of 2.0 and 2.2 M with various assumed turbulence were selected.
The effect of diffusion on the width of the classical instability strip can be estimated from Fig. 6 where our models which exhibit variability are shown in the HR diagram. In each panel of the figure, all models have the same turbulence parameterization. Clearly, the blue edge of the instability strip is sensitive to the efficiency of the diffusion processes, which means that it is sensitive to the assumed turbulence. It provides additional and independent constraints on the turbulence model. The superficial metal abundances constrain the turbulence (see Richer et al. RMT00 (2000)). However, the superficial He abundance is poorly determined by observations. It is much better determined via the driving of $`\delta `$ Scuti-type pulsations.
In practice, each mode of oscillation has an individual blue edge and each mode is affected differently by diffusion. Tables 4 and 5 and Fig. 5 suggest that the blue edge for the fundamental mode is only slightly shifted to the red compared with normal stars except if turbulence is better represented by the R300-2 model. The results of Richer et al. (RMT00 (2000)) do not exclude this model, at least not for all Am stars. The blue edge for higher-frequency modes is shifted significantly as soon as the He abundance is slightly reduced.
As indicated in Fig. 6 we have considered only models hotter than the observational red edge of the instability strip. It is likely that the return to stability for cooler models would be dominated by convective effects (e.g. Houdek et al. Houdeketal99 (1999)), which are ignored here. On the basis of time-dependent mixing-length calculations (Gough Gough77 (1977); Balmforth Balmforth92 (1992)) one finds that the convective effects grow rapidly as the red edge is approached (e.g. Houdek Houdek00 (2000)), although they have significant influence on the stability properties also in somewhat hotter stars. Even so, we expect that our qualitative conclusions, and the results on differential effects of diffusion, are robust.
### 8.3 Comparison with variable evolved CP stars
The $`\delta `$ Delphini spectral sub-type has been shown to be a very inhomogeneous group of stars consisting of classical Am stars, evolved Am stars and other stars which are essentially normal A stars (Gray & Garrison GrayGarrison89 (1989)). Amongst those, the $`\rho `$ Puppis stars are those which are thought to be evolved Am stars. Some of these evolved Am stars have been shown to be variable. It was shown in Sect. 8.1 that models of evolved Am stars can be variable; thus, it remains to be seen whether our evolved Am-star models are comparable with the observed variable CP stars. To verify if this is the case, the most evolved models of 1.90R1000-2, 2.00R1000-2 and 2.20R1000-2 are compared here with the variable metallic stars observed by Kurtz (Kurtz76 (1976)).
The positions of these stars are first shown in a temperature – gravity diagram (Fig. 7) to see if the models are in an evolutionary state comparable with the stars in our reference sample. The variable 1.90R1000-2 model at 1.094 Myr and the variable 2.00R1000-2 model at 850 Myr are both on the cool side (roughly 500 K cooler than the average) of the observed stars. Surprisingly, the 2.20R1000-2 model at 670 Myr, which is apparently a close match to two variable metallic A stars and which is much more compatible with the sequence of variable metallic stars than the cooler models, is predicted to be stable. This may or may not be a serious difficulty given the uncertainty of effective-temperature determinations for CP stars.
The surface-abundance signatures of the Kurtz (Kurtz76 (1976)) $`\delta `$ Delphini stars and of the three selected models of the evolved Am stars are compared in Fig. 8. In a second panel we repeat the surface abundance of the evolved 2.20 M but then compared with the two stars closest to it in Fig. 7, which are HR1706 and HR6561. These two stars have very nearly normal abundances that are closer to the surface abundances of a 2.20R10K-2 model which would probably pulsate at that $`T_{\mathrm{eff}}`$ and $`\mathrm{log}g`$, if it behaves similarly to the 1.9R10K-2 model.
Fig. 8 shows that there is little variation in the surface-abundance profiles between the three evolved Am-star models. This is due to the fact that the depth of the mixed region is similar in these models since we adopted the same turbulence parameterization in all three. There is a significant scatter in the observed abundances of these stars. Nonetheless, the models lie in general within the envelope defined by the observations. As Richer et al. (RMT00 (2000)) pointed out, care must be taken when comparing with observed abundances due to large variations in the abundances determined by different observers. Although one cannot claim that the global properties of the present models are a perfect match to those of individual variable CP stars, there is a qualitative agreement for the stars taken as a group. Better observational data for these stars would greatly help the interpretation of the results.
### 8.4 The excitation of g modes in A stars with diffusion
As the driving regions shift deeper into the star as it evolves, the g modes become gradually more and more excited. The effect of diffusion on the excitation of p modes and g modes of a 1.9 M star was discussed in Sect. 8.2. It was shown there that p modes are stabilized through diffusion whereas g modes tend to be excited as a result of that process. In the more evolved stars presented here, some g modes are predicted to become overstable. For example, we find all overstable modes in model 1.90-ND at 1.105 Gyr to be of the physical nature of g modes (the mathematical mode orders are very confusing for that model). However, the modes which are found to be overstable in our models are low-order modes for which the driving from helium is significant. Therefore, the effect of diffusion on their excitation is similar to that of low-order p modes, i.e., the settling of helium leads to a reduction of their excitation.
As an example for one of these g modes, we get the following growth parameters for the $`l=1,n=1`$ mode in the different models of 1.9 M at roughly 1 Gyr, in order of decreasing helium content in the driving region: 0.0864 for 1.9-ND, 0.0891 for 1.9R10K-2, 0.0294 for 1.9R1000-2 and $`0.0098`$ for 1.9R300-2.
The detection of g modes in $`\delta `$ Scuti stars has been problematic. Models for evolved A stars predict a very dense spectrum of overstable g modes while observations so far show little evidence of such variations (Guzik Guzik00 (2000)). In addition, Breger & Beichbuchner (BregerBeichbuchner96 (1996)) show that there is no observational evidence for long-period g modes (of the nature of those observed in cooler $`\gamma `$ Doradus stars) in $`\delta `$ Scuti stars. The models suggest that they would not be easier to find in evolved Am stars, especially considering that convective effects should become increasingly important in models leaving the main sequence.
## 9 Is there a seismic signature of diffusion in $`\delta `$ Scuti stars?
In general, $`\delta `$ Scuti stars are fairly fast rotators in which diffusion is not expected to be important. In those stars for which diffusion could be important it is legitimate to ask in what measure it could influence their modeling.
The high-amplitude $`\delta `$ Scuti (HADS) stars are typically slow rotators and as such are candidates to become Am stars. They are characterized by high-amplitude pulsations compared with their normal counterparts. All evidence currently points to HADS being normal stars (Høg & Petersen HogPetersen97 (1997)) following normal stellar evolution (Petersen & Christensen-Dalsgaard PetersenJCD96 (1996)). They are not distributed uniformly in the instability strip but occupy only a strip some 300 K wide. If the high amplitudes were linked to a particular non-linear effect caused by the low rotational velocity this could have interesting repercussions on the stability of other slowly rotating A stars, such as the Am stars. Certainly, nothing in our results would lead us to believe that diffusion, by itself, would generate higher-amplitude modes in A stars. The lack of abundance anomalies in HADS might simply be due to their advanced evolutionary state and the depth of the convection zone at that time.
As has been briefly touched upon in the introduction (Sect. 2) some $`\delta `$ Scuti stars may exhibit abundance anomalies. These anomalies are not small contrary to what one might expect. In fact, one can see a scatter of \[Ca/Fe\] from $`0.6`$ to $`0.7`$ and depletions of carbon of up to \[C/Fe\] $`1.3`$ in Russell’s (Russell95 (1995)) results. Referring to Fig. 8 it is evident that the anomalies reported by Russell are higher than those observed in evolved Am stars. Similar results have been also published by Rachkovskaya (Rachkovskaya94 (1994)) and her other work referenced therein.
Although the anomalies have the overall appearance of diffusion, e.g. depleted C and Ca, normal Si and enhanced Fe and Ti, the scatter of the observed abundances is so large as to make any correlation between the abundances of the different elements very difficult. There is no systematic signature of diffusion as found in Am stars (cf. Richer et al. RMT00 (2000)) or even in the more evolved Am stars shown in Fig. 8. Moreover, some of those stars exhibit a $`v\mathrm{sin}i`$ as large as 100 to 150 km s<sup>-1</sup>.
In other more standard $`\delta `$ Scuti stars the abundance anomalies are expected to be small. In such cases, the effect of diffusion on the seismology of those stars would be subtle. Previous experience in lower-mass stars (Turcotte & Christensen-Dalsgaard TJCD98 (1998)) suggests that any seismic signature of diffusion would be overwhelmed by other effects such as convective-core overshoot, for example. In the event that the iron convection zone would develop, one could expect a clear seismic signature if a sufficient number of oscillation modes were identified. Guzik (Guzik93 (1993)) pointed out that helium settling may have some consequence on observed properties of $`\delta `$ Scuti stars, such as the light curve.
## 10 Pulsation and turbulence
In their discussion of metallicism and pulsations, Kurtz and collaborators have made the point that pulsations involve rapid motion to a substantial depth in the star. They argue that velocity of the displacement generated by the pulsation remains high throughout the region in which the abundance anomalies are thought to be formed (referring for example to the eigenfunctions displayed in Fig. 8.2 of Cox Cox80 (1980)). They speculate that the high velocities might generate turbulent mixing and therefore inhibit the occurrence of metallicism.
Although the velocity of the radial displacement might be large, turbulence is expected to be generated not by the speed of the displacement itself but rather by velocity gradients: a fast but uniform displacement will not become turbulent. The scale on which the displacement occurs is much greater than other relevant scales, e.g. the pressure scale height. The gradient of the displacement velocity is small over those scales and one would not expect turbulence to be necessarily generated as a result.
Our understanding of the extent to which pulsations are laminar or generate turbulence is too incomplete for us to speculate further on the link between the turbulence assumed in Am stars and the observed pulsations.
## 11 Conclusion
We have investigated the impact of diffusion on the stability of A-type main-sequence Population I stars. The models include the consistent abundance evolution of all important elements and its impact on stellar structure. In these models helium remains present in the He II ionization zone and the opacity in the iron opacity bump increases substantially, raising the possibility of a strong relationship between variability and diffusion.
We present evidence that young Am stars are stable against driving from the $`\kappa `$ mechanism and that, as the stars evolve, they become unstable, but only when near the red edge of the instability strip. Hot Am stars need to be more evolved than cool Am stars before variability can occur. The blue edge of the instability strip for metallic A stars is sensitive to the magnitude of the abundance variations and is thus indicative of the depth of mixing by turbulence.
The stability of A stars is more sensitive to the evolution of the abundance of helium than to the accumulation of iron-peak elements. In stars with very little turbulence the iron-peak driving region can become convectively unstable thereby reducing the radiative flux there and negating the driving effect of the enhanced opacity bump. However, in the models which are representative of Am stars, the turbulence is high enough to prevent the formation of that iron convection zone while allowing a significant increase of the opacity bump due to iron-peak elements. Still, only a marginal positive effect can be seen in the long-period g modes. The higher-frequency modes depend mostly on the helium abundance and somewhat on the hydrogen abundance.
There are a number of caveats relevant to the present work.
There is no direct link between the normalized growth rates $`\eta `$ and the actual amplitude of the pulsations as evidenced by the lower number of modes observed in $`\delta `$ Scuti stars relative to their predicted number. Additionally, comparisons to models at solar composition by J. Christensen-Dalsgaard and W. Dziembowski have shown that the growth rates are sensitive to the details of the modeling. In general, in the stars we compared at 1.8 and 1.9 M, the growth rates were lower in our models than in either of their models.
The treatment of convection in the present work is simplistic. Convection and turbulence are known to damp pulsation to a certain extent. The most striking and well known evidence of the effect of convective damping is the red edge of the instability strip \[Gough (Gough77 (1977)); Gonczi & Osaki (1980); Balmforth & Gough (BalmforthGough88 (1988)); see also Buchler et al. (BYKG99 (1999)) for a recent review\]. Turbulence affects pulsations in two ways: 1) through turbulent viscosity, which always dampens pulsations, and 2) through the phase difference between entropy variations and the modulation of the convective flux, which can either excite or dampen pulsations. It is the latter effect which is responsible for the red edge of the classical instability strip. While the interaction between turbulent convection and pulsations has been studied, the effect of turbulence outside of the convection zones is unknown.
In the models including diffusion, the coupling between turbulence and diffusion might be very important. For the present models reproducing Am stars, the turbulence extends to a significant depth below the convection zone. While the energy flux related to the turbulence is expected to be very small, the effect of turbulent viscosity might not be negligible. For stars in which the iron convection zone is allowed to form, three separate convection zones exist with highly turbulent regions between them. The behavior of pulsations in such stars might be heavily affected by a proper treatment of convective effects.
While it is true that our analysis does not take this into account, one should not forget that in the standard models without diffusion an ad hoc mixing mechanism is implied to prevent the formation of abundance gradients. This presumably turbulent mixing (at least in the more slowly rotating $`\delta `$ Scuti stars) is also never taken into account. It is, by hypothesis, larger than that in the models where atomic diffusion is important. Under these circumstances, we may perhaps consider the simple treatment of convection and turbulence applied here as adequate for identifying the differential effect of diffusion on stellar stability.
The NMM used here still contain an element of arbitrariness in that they extend turbulent mixing a little beyond that expected from iron convection zones, without providing a physical mechanism for this extension. This extension is required to fit the observed surface abundances of Am stars. The extension is, however, sufficient to cause the iron-peak abundances to decrease sufficiently for the convection zone to disappear in models representative of these stars. The models do not include the potential effects of mass loss or rotation, although differential rotation is one mechanism that is now being investigated as a source of the instabilities that could provide this mixing zone extension.
When compared against each other the present models do, however, illustrate the effect of diffusion on the stability of main-sequence A stars as well as present modeling allows.
## ACKNOWLEDGEMENTS
This work was supported in part by the Danish National Research Foundation through its establishment of the Theoretical Astrophysics Center. We are deeply indebted to Wojtek Dziembowski for his pulsation stability code and to Günter Houdek for his help in implementing his interpolation routines. We also acknowledge the very valuable comments of W. Dziembowski on the computations, as well as on an earlier version of the manuscript. Thanks are extended to D. O. Gough, D. Kurtz and J.-P. Zahn for insightful discussions, and J. O. Petersen for careful reading of the originally submitted manuscript.
|
warning/0006/hep-th0006050.html
|
ar5iv
|
text
|
# Nahm Transform For Periodic Monopoles And N=2 Super Yang-Mills Theory
## 1 Introduction and Summary
### 1.1 The Bogomolny Equation
Let $`X`$ be a three-dimensional oriented Riemannian manifold, $`E`$ be a vector bundle over $`X`$ with structure group $`G`$, $`A`$ be a connection on $`E`$, and $`\varphi `$ be a section of $`\mathrm{End}(E)`$. Bogomolny equation is the reduction of the self-duality equation to three dimensions which reads
$$F_A=d_A\varphi .$$
(1)
Here $``$ is the Hodge star operator. In what follows we set $`G=SU(2)`$, with $`E`$ being associated with the fundamental representation of $`SU(2)`$. In this case $`\varphi `$ is Hermitian and traceless. We will assume that all functions and connections are infinitely differentiable, unless specified otherwise.
It is well known that for $`X=^3`$ with flat metric the Bogomolny equation admits finite energy solutions, so-called BPS monopoles. The energy and magnetic charge of a pair $`(A,\varphi )`$ are defined as follows:
$`E(A,\varphi )`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle _X}\mathrm{Tr}(F_AF_A+d_A\varphi d_A\varphi ),`$
$`m(A,\varphi )`$ $`=`$ $`\underset{R\mathrm{}}{lim}{\displaystyle _{|x|=R}}{\displaystyle \frac{\mathrm{Tr}\left(F_A\varphi \right)}{4\pi \varphi }}.`$
Here
$$\varphi ^2=\frac{1}{2}\mathrm{Tr}\varphi ^2.$$
For BPS monopoles $`\varphi `$ tends to a constant value $`v`$ at infinity, while $`F_A`$ decreases as $`1/r^2`$. It follows that the energy of a BPS monopole is proportional to its magnetic charge:
$$E(A,\varphi )=2\pi vm(A,\varphi ).$$
BPS monopoles are absolute minima of the energy function $`E(A,\varphi )`$ in a subspace with fixed magnetic charge and fixed asymptotic value of $`\varphi `$.
### 1.2 Periodic Monopoles
Solutions of the Bogomolny equation on $`^2\times \mathrm{SS}^1`$ have not been studied previously. One of the reasons is that any monopole on $`^2\times \mathrm{SS}^1`$ with a nonzero magnetic charge must have infinite energy. This happens because the magnetic field of a magnetically charged object on $`^2\times \mathrm{SS}^1`$ decays only as $`1/r`$, where $`r`$ is the radial distance on $`^2`$. Hence the magnetic energy density decays as $`1/r^2`$, and its integral over $`^2\times \mathrm{SS}^1`$ diverges.
Since the energy of a periodic monopole is infinite, it cannot be regarded as a solitonic particle. Still, periodic monopoles do play a role in certain physical problems. For example, we will see that the centered moduli space of an $`SU(2)`$ periodic monopole with magnetic charge $`k`$ is a hyperkähler manifold of dimension $`4(k1)`$. It turns out that this hyperkähler manifold coincides with the quantum Coulomb branch of the $`𝒩=2`$ super Yang-Mills theory on $`^3\times \mathrm{SS}^1`$ with gauge group $`SU(k)`$ (see below). For $`k=2`$ this manifold is a very interesting deformation of the Atiyah-Hitchin manifold (the reduced moduli space of a $`k=2`$ monopole on $`^3`$) and is an example of a new class of asymptotically locally flat self-dual gravitational instantons. The properties of the moduli spaces of periodic monopoles will be discussed in more detail in a forthcoming paper .
The goal of this paper is two-fold. On one hand, we want to compute the dimension of the moduli spaces of periodic monopoles and to establish a correspondence between periodic monopoles and solutions of Hitchin equations on a cylinder. The latter correspondence is a particular instance of Nahm transform. On the other hand, we want to explain the relation of our results to the four-dimensional $`𝒩=2`$ gauge theories and to the brane configurations of the type first considered by Chalmers and Hanany and further explored by Hanany and Witten , Witten , and many others. These brane configurations were used to find the exact Coulomb branch of the super Yang-Mills theory with eight supercharges on $`^3`$ and $`^4`$. In particular, Witten showed how to obtain the exact solution of the $`𝒩=2`$ super Yang-Mills on $`^4`$ from the physics of the M-theory fivebrane. As explained below, the Nahm transform approach not only reproduces the classical physics of the fivebrane, but goes considerably further by resumming the effects of membrane instantons.
In a companion paper we study solutions of Bogomolny equations on $`^2\times \mathrm{SS}^1`$ with prescribed singularities. Their moduli spaces provide more examples of novel self-dual gravitational instantons and are related to four-dimensional $`𝒩=2`$ gauge theories with matter compactified on a circle.
In the remainder of this section we define periodic monopoles more precisely and formulate our main result, the correspondence between periodic monopoles and solutions of Hitchin equations on a cylinder.
### 1.3 Periodic Dirac Monopoles
Before investigating the nonabelian Bogomolny equation on $`^2\times \mathrm{SS}^1`$, it is instructive to write down solutions of the Bogomolny equation in the case $`G=U(1)`$. In this case there are no nontrivial smooth solutions, so we allow for singularities at some finite number of points on $`^2\times \mathrm{SS}^1`$. A solution with one singularity represents a Dirac monopole on $`^2\times \mathrm{SS}^1`$.
For $`G=U(1)`$ the Bogomolny equation implies that the Higgs field satisfies the Laplace equation
$$^2\varphi =0.$$
Let $`z`$ be a complex affine coordinate on $`^2`$ and $`\chi [0,2\pi ]`$ be the periodic coordinate on $`\mathrm{SS}^1`$. We will denote by $`x`$ the pair $`(z,\chi )`$. The solution corresponding to a Dirac monopole at $`x=0`$ is given by
$$\varphi (x)=v+kV(x)v+k\frac{\mathrm{log}(4\pi )\gamma }{2\pi }\frac{k}{2}\underset{p=\mathrm{}}{\overset{\mathrm{}}{^{}}}\left[\frac{1}{\sqrt{|z|^2+(\chi 2\pi p)^2}}\frac{1}{2\pi |p|}\right],$$
where the prime means that for $`p=0`$ the second term in the square brackets must be omitted, and $`\gamma `$ is the Euler’s constant. $`V(x)`$ satisfies the Laplace equation everywhere except $`z=0,\chi =0mod2\pi `$. Near this point $`V(x)`$ diverges:
$$V(x)\frac{1}{2\sqrt{|z|^2+|\chi |^2}}+O(1).$$
For large $`|z|`$ the function $`V(x)`$ is given by
$$V(x)\frac{\mathrm{log}|z|}{2\pi }+o(1).$$
The connection $`A`$ corresponding to this Higgs field has the following asymptotics for $`|z|\mathrm{}`$ (up to a gauge transformation):
$$A_z\frac{a}{z}+o(1/z),A_\chi =\frac{k}{2\pi }\mathrm{arg}z+b+o(1).$$
Here $`a`$ and $`b`$ are real constants. For these formulas to define a connection on a $`U(1)`$ bundle, the parameter $`k`$ must be an integer.
The magnetic charge of a $`U(1)`$ monopole is defined as the first Chern class of the monopole bundle restricted to the 2-torus $`|z|=R`$ for sufficiently large $`R`$. It is easy to see that the magnetic charge of the above monopole is $`k`$.
The Higgs field of a solution describing several periodic Dirac monopoles has the form
$$\varphi (x)=v+\underset{\alpha }{}k_\alpha V(xx_\alpha ).$$
(2)
It is singular at $`x=x_\alpha `$ and for large $`|z|`$ behaves as
$$\varphi (x)v+\frac{\mathrm{log}|z|}{2\pi }\underset{\alpha }{}k_\alpha +o(1).$$
(3)
### 1.4 Asymptotics Of A Periodic Monopole
It is well known that finite-energy solutions of $`SU(2)`$ Bogomolny equations on $`^3`$ are exponentially close to the Dirac monopole at large distances . Then it is natural to require that periodic $`SU(2)`$ monopoles be close to the periodic Dirac monopole at large $`|z|`$. Accordingly, we will look for solutions of $`SU(2)`$ Bogomolny equations on $`^2\times \mathrm{SS}^1`$ such that outside a compact set $`T^2\times \mathrm{SS}^1`$ one has
$`\varphi (x)`$ $``$ $`g(x)\sigma _3\varphi _D(x)g(x)^1+o(1),`$
$`d_A\varphi (x)`$ $``$ $`g(x)\sigma _3d\varphi _D(x)g(x)^1+o(1/|z|),`$
$`A(x)`$ $``$ $`g(x)\sigma _3A_D(x)g(x)^1+g(x)dg^1(x)+o(1).`$ (4)
Here $`\sigma _3=\mathrm{diag}\{1,1\}`$$`g(x)`$ is an $`SU(2)`$-valued function on $`\left(^2\times \mathrm{SS}^1\right)\backslash T`$, and $`\varphi _D`$ and $`A_D`$ are a 0-form and a $`U(1)`$ connection defined by
$`\varphi _D(x)`$ $`=`$ $`v+k{\displaystyle \frac{\mathrm{log}|z|}{2\pi }},`$ (5)
$`A_D(x)`$ $`=`$ $`b+{\displaystyle \frac{k}{2\pi }}\mathrm{arg}z.`$ (6)
This means that up to terms vanishing at infinity a periodic $`SU(2)`$ monopole is gauge-equivalent to a periodic Dirac monopole with charge $`k`$ embedded in a $`U(1)`$ subgroup of $`SU(2)`$.
The real parameters $`v`$ and $`b`$ will often appear in a combination $`v+ib`$. We will denote this combination $`𝔳.`$
Note that we implicitly set the circumference of circle parameterized by $`\chi `$ to be $`2\pi `$. This does not entail a loss of generality, as the Bogomolny equation is invariant with respect to rescalings of the metric on $`X`$. One has to keep in mind that rescaling the circumference by a factor $`\lambda `$ requires rescaling the Higgs field $`\varphi `$ by the same factor. Thus the “large circumference limit” is equivalent to the “large $`v`$ limit.” We will use these terms interchangeably.
The magnetic charge of a periodic monopole is defined in analogy with the case of monopoles on $`^3`$. It follows from Eq. (1.4) that for large enough $`|z|`$ the eigenvalues of $`\varphi `$ are distinct (and opposite). Hence for large enough $`|z|`$ one has a well-defined line bundle $`L_+E`$, the eigenbundle of $`\varphi `$ associated with the positive eigenvalue. The magnetic charge can be defined as the first Chern class of $`L_+`$ restricted to a 2-torus $`|z|=R`$, where $`R`$ is large enough. Thus the magnetic charge is given by the formula
$$m(A,\varphi )=\underset{R\mathrm{}}{lim}_{|z|=R}\frac{\mathrm{Tr}\left(F_A\varphi \right)}{4\pi \varphi }.$$
(7)
It is easy to see that the magnetic charge of a periodic monopole is nonnegative. Substituting the asymptotics Eqs. (1.4-6) into this formula, one finds that $`m(A,\varphi )=k`$, so $`k`$ must be nonnegative too. Unlike the case of monopoles on $`^3`$, the energy of a monopole is infinite for $`k0`$.
### 1.5 Nahm Transform For Periodic Monopoles
Let $`\mathrm{\Sigma }`$ be a Riemann surface, $`V`$ be a unitary vector bundle on $`\mathrm{\Sigma }`$, $`\widehat{A}`$ be a connection on $`V`$, and $`\mathrm{\Phi }`$ be a section of $`\mathrm{End}(V)\mathrm{\Omega }_\mathrm{\Sigma }^{1,0}.`$ Hitchin equations for $`\widehat{A},\mathrm{\Phi }`$ are the equations
$$\overline{}_{\widehat{A}}\mathrm{\Phi }=0,F_{\widehat{A}}+\frac{i}{4}[\mathrm{\Phi },\mathrm{\Phi }^{}]=0,$$
where the commutator is understood in a graded sense, i.e. $`[\mathrm{\Phi },\mathrm{\Phi }^{}]=\mathrm{\Phi }\mathrm{\Phi }^{}+\mathrm{\Phi }^{}\mathrm{\Phi }.`$ Hitchin equations are the reduction of the self-duality equation to two dimensions.
Our main result is that there is a one-to-one correspondence (modulo gauge transformations) between $`SU(2)`$ periodic monopoles with magnetic charge $`k`$ and solutions of $`U(k)`$ Hitchin equations on a cylinder $`\times \mathrm{SS}^1`$ with the Higgs field growing exponentially at infinity. To describe the asymptotics of the Higgs field more precisely, let us regard $`\times \mathrm{SS}^1`$ as a strip $`0\mathrm{Im}s1`$ on the complex $`s`$–plane with the boundaries identified in an obvious manner. Then the Higgs field behaves as follows for $`\mathrm{Re}s\pm \mathrm{}`$:
$$\mathrm{\Phi }(s)g_\pm (s,\overline{s})e^{\pm \frac{2\pi s}{k}}\mathrm{diag}(1,\omega ,\omega ^2,\mathrm{},\omega ^{k1})g(s,\overline{s})_\pm ^1ds.$$
(8)
Here $`g(s,\overline{s})_\pm `$ are some (multi-valued) functions with values in $`U(k)`$, and $`\omega `$ is a $`k`$–th root of unity. The curvature of the $`U(k)`$ connection, on the other hand, approaches zero as $`1/|\mathrm{Re}\mathrm{s}|^{3/2}`$ for $`\mathrm{Re}\mathrm{s}\pm \mathrm{}.`$
For $`k=1`$ it is easy to write down an explicit solution of Hitchin equations with this asymptotics. Then Nahm transform implies that there exists a periodic $`SU(2)`$ monopole with $`k=1`$. One can also argue that solutions of Hitchin equations exist for all positive $`k`$, and even describe their moduli space. This implies that periodic monopoles exist for all $`k>0`$. It would be interesting to find an explicit formula for periodic monopole, at least in the $`k=1`$ case.
### 1.6 Outline
The paper is organized as follows. In Section 2 we explain the relation between periodic monopoles and $`𝒩=2`$ super Yang-Mills theory compactified on a circle. This section requires familiarity with the physics of branes in Type II string theory. The rest of the paper does not depend on it.
In Section 3 we show that the Nahm transform takes periodic monopoles to solutions of Hitchin equations on a cylinder. In Section 4 we explain how to associate algebro-geometric data to a periodic monopole. These data consist of an algebraic curve and a line bundle over it and are important in the study of Nahm transform. On the other hand, it is well known that to every solution of the Hitchin equations one can associate so-called spectral data also consisting of an algebraic curve and a line bundle. In Section 5 we show that the algebro-geometric data associated to the periodic monopole coincide with the spectral data of its Nahm transform. In Section 6 we use this information to determine the asymptotic behavior of the solutions of Hitchin equations arising from periodic monopoles. In Section 7 we describe the “inverse” Nahm transform which produces a solution of the Bogomolny equation on $`^2\times \mathrm{SS}^1`$ from a solution of Hitchin equations on a cylinder. In Section 8 we study the asymptotic behavior of the resulting solution of the Bogomolny equation and show that it is given by (1.4). In Section 9 we prove that the composition of the “direct” and “inverse” Nahm transform takes a periodic monopole to a gauge-equivalent periodic monopole. The proof is modelled on that of Schenk and requires rather tedious computations. Another approach to the proof which uses the spectral sequence technology is sketched in the Appendix. The results of Sections 3-9 imply that the Nahm transform establishes a one-to-one correspondence between periodic monopoles and solutions of Hitchin equations on a cylinder with a particular asymptotics. In Section 10 we give (nonrigorous) arguments that periodic monopoles exist for all $`k>0`$ and are (almost) completely determined by their spectral data. Assuming that this is true, we show in Section 11 that the centered moduli space of a charge $`k`$ periodic monopole has dimension $`4(k1)`$, and describe a distinguished complex structure on it. We also argue that the centered moduli space carries a natural hyperkähler metric.
## 2 Periodic Monopoles And Brane Configurations
This section assumes familiarity with the Chalmers-Hanany-Witten-type brane configurations and their use in solving quantum gauge theories with eight supercharges.
Consider two parallel flat NS5-branes in Type IIB string theory. For definiteness, let us assume that their worldvolumes are given by the equations
$$x^6=x^7=x^8=x^9=0$$
and
$$x^6=v,x^7=x^8=x^9=0.$$
This brane configuration is BPS (preserves sixteen supercharges), and its low-energy dynamics is described by a $`d=6`$ supersymmetric Yang-Mills theory with gauge group $`U(2)`$.
Consider now a D3-brane with the worldvolume given by
$$x^3=x^4=x^5=x^7=x^8=x^9=0,0x^6v.$$
This is an open D3-brane, in the sense that its worldvolume has boundaries. This is possible because the boundaries lie on the NS5-branes. One can say that such a D3-brane is suspended between the NS5-branes.
From the point of view of the Yang-Mills theory describing the NS5 branes, the suspended D3-brane is a static solution of the Yang-Mills equations of motion with a unit magnetic charge . Moreover, since a suspended D3-brane preserves eight supercharges, it is a BPS soliton, and must solve the Bogomolny equation.
Similarly, $`k`$ suspended D3-branes are described in the Yang-Mills theory by a charge $`k`$ monopole .
Let us now compactify the $`x^3`$ coordinate on a circle of radius $`R`$, i.e. let $`x^3`$ take values in $`/(2\pi R)`$ rather than in $``$. In such a situation we may still consider suspended D3-branes. The same arguments as in the uncompactified case lead one to the conclusion that $`k`$ D3 branes are described in the Yang-Mills theory by a BPS monopole on $`^2\times \mathrm{SS}^1`$ with charge $`k`$. The $`\mathrm{SS}^1`$ has circumference $`2\pi R`$.
Now let us apply T-duality in the $`x^3`$ direction. This has the effect of taking us to Type IIA string theory. We will denote the spatial coordinates in Type IIA by $`y^1,\mathrm{}y^9`$, so that $`y^3`$ can be identified with the Fourier dual of $`x^3`$, while the rest of the $`y`$ coordinates are identified with the corresponding $`x`$ coordinates. If we choose the units in which the Regge slope $`\alpha ^{}`$ is unity, then $`y^3`$ has period $`2\pi /R`$.
The usual T-duality rules tell us that the Type IIB NS5-branes are mapped under T-duality to the Type IIA NS5 branes with the worldvolumes given by
$$y^6=y^7=y^8=y^9=0$$
and
$$y^6=v,y^7=y^8=y^9=0.$$
A suspended D3-brane is mapped to a D4-brane with the worldvolume given by
$$y^4=y^5=y^7=y^8=y^9=0,0y^6v.$$
This D4-brane is suspended between the NS5-branes.
Such a brane configuration in Type IIA string theory has been first studied by E. Witten and subsequently by many other authors. The only difference with is that in our case $`y^3`$ is a periodic variable. Witten argued that the low-energy dynamics of $`k`$ suspended D4-branes is described by the $`d=4`$ $`𝒩=2`$ supersymmetric Yang-Mills theory with gauge group $`SU(k)`$. The classical gauge coupling of this theory depends on $`v`$: $`1/g_{YM}^2v/g_{st}`$. Thus we are dealing with a four-dimensional super Yang-Mills theory on $`^3\times \mathrm{SS}^1`$ where the circumference of $`\mathrm{SS}^1`$ is given by $`2\pi /R`$.
In the quantum theory the gauge coupling depends on the renormalization scale $`\mu `$: $`1/g_{YM}^2(\mu )=1/g_{YM}^2(\mu _0)+k\mathrm{log}(\mu /\mu _0)`$. From the string theory viewpoint, taking into account quantum corrections on the D4-brane worldvolume is equivalent to taking into account the back-reaction of the D4-branes on the NS5-branes. This back-reaction results in the bending of the NS5-branes, as a consequence of which the distance between them in the $`y^6`$ direction starts to depend on $`u=y^4+iy^5`$:
$$\delta x^6(u)=const+k\mathrm{Re}\mathrm{log}u.$$
Since D3-branes suspended between NS5-branes in Type IIB string theory are T-dual to D4-branes suspended between NS5-branes in Type IIA string theory, their moduli spaces must coincide (as Riemann manifolds). The moduli space of the former coincides with the moduli space of $`k`$ periodic monopoles. The moduli space of the latter is the Coulomb branch of the $`d=4`$ $`𝒩=2`$ supersymmetric Yang-Mills with gauge group $`SU(k)`$ compactified on a circle of radius $`1/R`$.
Assuming that this correspondence is true, we may predict the dimension of the moduli space of periodic monopoles. As explained in , the Coulomb branch of a $`d=4`$ $`𝒩=2`$ super-Yang-Mills theory on $`^3\times \mathrm{SS}^1`$ is a hyperkähler manifold of dimension $`4\mathrm{rank}(G)`$, where $`G`$ is the gauge group. Thus the moduli space of a periodic monopole of charge $`k`$ must have real dimension $`4(k1)`$.
A particular case of this correspondence has been known for some time from the work of Chalmers and Hanany . These authors showed that the centered moduli space of $`k`$ periodic monopoles on $`^3`$ coincides with the Coulomb branch of the $`d=3`$ $`𝒩=4`$ supersymmetric Yang-Mills with gauge group $`SU(k)`$ on $`^3`$. This statement follows from ours in the limit $`R\mathrm{}`$. In this limit monopoles on $`^2\times \mathrm{SS}^1`$ reduce to ordinary monopoles on $`^3`$. On the other hand, the radius of the dual circle goes to zero, and therefore the $`d=4`$ $`𝒩=2`$ super-Yang-Mills undergoes Kaluza-Klein reduction to the $`d=3`$ $`𝒩=4`$ super-Yang-Mills.
We pause here to explain one subtlety in the above arguments. The Coulomb branch of the $`d=3`$ super-Yang-Mills theory with gauge group $`SU(k)`$ is related to the centered monopole moduli space , while the Coulomb branch of the $`d=4`$ gauge theory on a circle appears to be related to the uncentered moduli space of periodic monopoles. If this were the case, we would not get an exact agreement between the two statements in the limit $`R\mathrm{}`$. In fact, when considering periodic monopoles, one is forced to fix their center-of-mass if one wants to get a well-defined metric on the moduli space. The reason is that the translational zero modes of a periodic monopole are not normalizable. This is explained in more detail in Section 11. In this way the contradiction is avoided. (The fact that the translational zero modes for suspended D4 branes are not normalizable was explained from the string theory point of view in . This ”freezing out” of the center-of-mass motion is the ultimate reason why the suspended D4-branes are described by an $`SU(k)`$ rather than $`U(k)`$ gauge theory.)
Another interesting limit is $`R0`$. In this limit the circle on which the $`d=4`$ $`𝒩=2`$ super-Yang-Mills theory is compactified becomes arbitrarily large, while the monopole interpretation looses meaning. The Coulomb branch of this theory with all quantum corrections has been determined in . It is a special Kähler manifold of real dimension $`2(k1)`$. (Note that the dimension of the Coulomb branch jumps by a factor two as soon as one compactifies one dimension a circle. The reason for this is explained in .) The simplest way to derive the answer uses the Type IIA brane configuration with suspended D4-branes described above . One notices that the metric on the Coulomb branch does not depend on the string coupling if $`g_{YM}`$ is kept fixed, so one can consider the limit $`g_{st}\mathrm{},v\mathrm{}.`$ In this limit Type IIA string theory reduces to $`d=11`$ supergravity, and the configuration with D4-branes suspended between two NS5 branes turns into a single smooth M5-brane. The metric on the Coulomb branch with all quantum corrections taken into account can be obtained by a classical computation with an M5-brane.
It would certainly be nice if the quantum Coulomb branch of the compactified theory could also be determined by a classical computation in $`d=11`$ supergravity. However, it is easy to see that this is not the case. The reason is that upon compactification on a circle there appear new kinds of instantons in the gauge theory, namely virtual BPS monopoles and dyons whose worldlines wrap the compactified circle. In a strongly coupled string theory such effects are captured by membrane instantons. These instantons are represented by Euclidean open M2-branes whose boundaries lie on the M5-brane. Clearly, directly summing up all such instantons is a hopeless task.
Nevertheless, one can give a “classical” recipe for computing the complete quantum Coulomb branch of the compactified super-Yang-Mills theory by exploiting the correspondence with periodic monopoles. Computing the metric on the moduli space of periodic monopoles is a well-defined problem which appears much simpler than summing up membrane instantons. One could hope to determine this metric using twistor methods, similarly to how it has been done for ordinary monopoles. Alternatively, one could apply Nahm transform to make the problem more manageable. Below we show that the Nahm transform of a periodic monopole is described by Hitchin equations on $`^{}`$. These equations are somewhat simpler than the original Bogomolny equation. The properties of the moduli space of periodic monopoles will be studied in detail in a forthcoming publication .
## 3 From Periodic Monopoles To Solutions Of Hitchin Equations
In this section we show that Nahm transform associates to every periodic $`SU(2)`$ monopole with charge $`k`$ a solution of $`U(k)`$ Hitchin equations on a cylinder. We follow where the Nahm transform for instantons on $`T^4`$ is discussed. In fact, periodic monopoles can be regarded as a limiting case of instantons on $`T^4`$ invariant with respect to a subgroup of translations. Another closely related work is , where Nahm transform for instantons on $`^2\times T^2`$ is studied. We will use many of the techniques of and .
Let the pair $`(A,\varphi )`$ be a periodic $`SU(2)`$ monopole with asymptotics (1.4). Let $`S`$ be the spinor bundle on $`X=^2\times \mathrm{SS}^1`$. This means that $`S`$ is a trivial unitary rank 2 bundle on $`X`$ equipped with an injective bundle morphism $`\sigma :T^{}XSS^{}`$ which is Hermitian and has zero trace. By a change of trivialization, one can always bring $`\sigma `$ to the standard form $`\sigma (dx_j)=\sigma _j,j=1,2,3,`$ where $`\sigma _j`$ are the Pauli matrices. Let $`L`$ be a trivial unitary line bundle on $`X`$ with a flat unitary connection $`a`$ whose monodromy around $`\mathrm{SS}^1`$ is $`\mathrm{exp}(2\pi it),t/`$ (these conditions define a unique connection).
Consider a Dirac–type operator $`D:ESLESL`$ of the form
$$D=\sigma d_{A+a}(\varphi r).$$
(9)
We will be interested in its $`L^2`$ kernel and cokernel. Using the fact that the norm of the Higgs field $`\varphi `$ grows logarithmically at infinity, one can show that $`D`$ is Fredholm for any $`(r,t)\times /`$. Thus its $`L^2`$–index is independent of $`r,t`$. As explained in the end of this section, the index is equal to the negative of the magnetic charge $`k`$.
The Weitzenbock formula for $`D`$ reads:
$$D^{}D=_{A+a}^2+(\varphi r)^2+\sigma (d_A\varphi F_A).$$
(10)
This formula together with the Bogomolny equation imply that $`D^{}D`$ is a positive-definite operator, and therefore $`D`$ has a trivial $`L^2`$ kernel. It follows that $`\mathrm{Ker}D^{}`$ is a rank $`k`$ trivial bundle over the $`(r,t)`$–plane. Actually, since $`t`$ is a periodic variable, we get a rank $`k`$ bundle over a cylinder $`\widehat{X}=\times \mathrm{SS}^1^{}.`$ This trivial bundle will be denoted $`\widehat{E}`$.
From the growth of $`\varphi `$ at infinity it follows that for all $`s=r+it`$ the elements of $`\mathrm{Ker}D^{}`$ decay at least exponentially. Thus for all $`s`$ we have a well-defined Hermitian inner product on $`\widehat{E}_s`$. If we choose a basis $`\psi _1(x,s),\mathrm{},\psi _k(x,s),xX,`$ of $`\mathrm{Ker}D^{}`$ at point $`s`$, then the explicit formula for the inner product is
$$\psi _\alpha ,\psi _\beta =\psi _\alpha (x,s)^{}\psi _\beta (x,s)d^3x.$$
This inner product makes $`\widehat{E}`$ into a unitary bundle. Below it will be assumed that the vectors $`\psi _\alpha ,\alpha =1,\mathrm{},k,`$ are chosen to form an orthonormal basis of $`\mathrm{Ker}D^{}`$ for all $`s`$.
Next we want to define a connection $`\widehat{A}`$ on $`\widehat{E}`$ and a Higgs field $`\widehat{\varphi }\mathrm{\Gamma }(\mathrm{End}(\widehat{E}))`$. The Higgs field at a point $`s\widehat{X}`$ is a linear map from $`\widehat{E}_s`$ to $`\widehat{E}_s`$. We define this map as a composition of two maps: multiplication by $`z`$ and projection to $`\widehat{E}_s.`$ An explicit formula for $`\widehat{\varphi }`$ in an orthonormal basis is
$$\widehat{\varphi }(s)_\beta ^\alpha =\psi _\alpha (x,s)^{}z\psi _\beta (x,s)d^3x.$$
(11)
Since all $`\psi _\alpha `$ decay at infinity faster than any power of $`z,`$ this is well-defined.
The connection $`\widehat{A}`$ on $`\widehat{E}`$ is induced by the zero connection on the trivial infinite-dimensional bundle whose fiber at a point $`s\widehat{X}`$ consists of all smooth $`L^2`$ sections of $`ESL`$. In components:
$$\widehat{A}_s(s)_\beta ^\alpha ds=i\psi _\alpha (x,s)^{}𝑑s\left(\frac{}{s}\psi _\beta (x,s)\right)d^3x.$$
(12)
It is easy to see that $`\widehat{A}`$ is a unitary connection on $`\widehat{E}`$. As for $`\widehat{\varphi }\mathrm{\Gamma }(\mathrm{End}(\widehat{E}))`$, it is not Hermitian, unlike its counterpart $`\varphi \mathrm{\Gamma }(\mathrm{End}(E))`$.
Now we will show that $`\widehat{A}`$ and $`\widehat{\varphi }`$ satisfy Hitchin equations. We will need the following commutation relations:
$$\begin{array}{cccc}[D,z]=\sigma _+,\hfill & [D,\overline{z}]=\sigma _{},\hfill & [D,]=p_+,\hfill & [D,\overline{}]=p_{},\hfill \\ [D^{},z]=\sigma _+,\hfill & [D^{},\overline{z}]=\sigma _{},\hfill & [D^{},]=p_{},\hfill & [D^{},\overline{}]=p_+.\hfill \end{array}$$
(13)
Here $`=/s,\overline{}=/\overline{s},`$ $`\sigma _\pm =\sigma _1\pm i\sigma _2,`$ $`p_\pm =\frac{1}{2}(1\pm \sigma _3)`$. We will denote the projector to $`\mathrm{Ker}D^{}`$ by $`P`$. Its explicit form is
$$P=1D(D^{}D)^1D^{}.$$
The projector to the orthogonal complement of $`\mathrm{Ker}D^{}`$ will be denoted by $`Q`$:
$$Q=D(D^{}D)^1D^{}.$$
First let us compute $`\overline{}_{\widehat{A}}\widehat{\varphi }`$:
$$\begin{array}{ccc}\overline{}_{\widehat{A}}\widehat{\varphi }\hfill & =& [P\overline{},Pz]\hfill \\ & =& (P\overline{}(1Q)zPz(1Q)\overline{})\hfill \\ & =& P(zQ\overline{}\overline{}Qz).\hfill \end{array}$$
Using the identity $`PD=0,`$ and keeping in mind that $`\overline{}_{\widehat{A}}\widehat{\varphi }`$ should be thought of as acting on $`\mathrm{Ker}D^{}`$ from the right, we can rewrite this expression as follows:
$$\begin{array}{ccc}\overline{}_{\widehat{A}}\widehat{\varphi }\hfill & =& P([z,D](D^{}D)^1[D^{},\overline{}][\overline{},D](D^{}D)^1[D^{},z])\hfill \\ & =& P(\sigma _+(D^{}D)^1p_++p_{}(D^{}D)^1\sigma _+).\hfill \end{array}$$
To go from the first line to the second line we used the commutation relations (13). The Weitzenbock formula (10) tells us that $`D^{}D`$ commutes with all $`\sigma _j,j=1,2,3`$, and since $`p_{}\sigma _+=\sigma _+p_+=0`$, we get the “complex” Hitchin equation
$$\overline{}_{\widehat{A}}\widehat{\varphi }=0.$$
(14)
The curvature $`\widehat{F}`$ of the connection $`\widehat{A}`$ is given by
$$\widehat{F}=i[P,P\overline{}]dsd\overline{s}.$$
We can simplify this as follows:
$$\begin{array}{ccc}\hfill \widehat{F}& =& iP(Q\overline{}Q\overline{}QQ)dsd\overline{s}\hfill \\ & =& iP(D(D^{}D)^1\overline{}D^{}\overline{}D(D^{}D)^1D^{})dsd\overline{s}\hfill \\ & =& iP(p_+(D^{}D)^1p_+p_{}(D^{}D)^1p_{})dsd\overline{s}\hfill \\ & =& iP(D^{}D)^1\sigma _3dsd\overline{s}.\hfill \end{array}$$
(15)
Here we again used the commutation relations (13) and the fact that $`D^{}D`$ commutes with all $`\sigma _j`$. On the other hand, let us compute the commutator $`[\widehat{\varphi },\widehat{\varphi }^{}]`$:
$$\begin{array}{ccc}[\widehat{\varphi },\widehat{\varphi }^{}]\hfill & =& [Pz,P\overline{z}]\hfill \\ & =& (P\overline{z}QzPzQ\overline{z})\hfill \\ & =& P([\overline{z},D](D^{}D)^1[D^{},z][z,D](D^{}D)^1[D^{},\overline{z}])\hfill \\ & =& P(\sigma _{}(D^{}D)^1\sigma _+\sigma _+(D^{}D)^1\sigma _{})\hfill \\ & =& 4P(D^{}D)^1\sigma _3.\hfill \end{array}$$
Comparing with the expression for the curvature of $`\widehat{A}`$, we obtain the “real” Hitchin equation:
$$\widehat{F}_{s\overline{s}}+\frac{i}{4}[\widehat{\varphi },\widehat{\varphi }^{}]=0.$$
(16)
To make the last equation covariant with respect to diffeomorphisms of $`\widehat{X}`$ one should think of $`\mathrm{\Phi }=\widehat{\varphi }ds`$ as a section of $`\mathrm{End}(\widehat{E})\mathrm{\Omega }_{\widehat{X}}^{1,0}`$. Then the “real” Hitchin equation takes the form
$$\widehat{F}+\frac{i}{4}[\mathrm{\Phi },\mathrm{\Phi }^{}]=0,$$
(17)
where the commutator is understood in the graded sense.
Following , we can associate to any solution of Hitchin equations an algebraic curve $`𝐂.`$ In the present case the curve is a hypersurface in $`\times ^{}`$ defined by the equation
$$det(z\widehat{\varphi }(s))=0.$$
(18)
Here $`z`$ is an affine parameter on $``$, while $`s`$ parameterizes $`\widehat{X}^{}.`$ The left-hand-side of the above equation is a polynomial in $`z`$ of degree $`k`$, and it follows from the “complex” Hitchin equation that its coefficients are holomorphic functions on $`^{}`$. This shows that the above equation defines an algebraic curve which is noncompact and is a $`k`$–fold cover of $`^{}`$.
The eigenvectors of $`\widehat{\varphi }`$ obviously form a sheaf $`N`$ on $`𝐂`$ whose stalk at a general point is one-dimensional. The direct image of $`N`$ under the projection map $`\pi :𝐂^{}`$ is the bundle $`\widehat{E}`$. We will call the pair $`(𝐂,N)`$ the spectral data of a Hitchin pair $`(\widehat{A},\widehat{\varphi })`$, and refer to $`𝐂`$ as the Hitchin spectral curve.
For a general Hitchin pair the curve $`𝐂`$ is nonsingular. If this is the case, then the sheaf $`N`$ is a line bundle. Indeed, since $`\pi _{}(N)`$ is a vector bundle, $`N`$ is a torsion free sheaf, hence a subsheaf of a locally free sheaf. But any subsheaf of a locally free sheaf on a smooth algebraic curve is locally free (this follows from the fact that a nonsingular curve has cohomological dimension one). Thus $`N`$ must be a line bundle.
Finally, let us justify the assertion that $`\mathrm{Ind}D=k`$. The index can be computed using the heat kernel method. Alternatively one may use the approach of Callias who computed the index of a Dirac-type operator on $`^{2n+1}`$ for all $`n`$. One can check that the proof goes through for $`^2\times \mathrm{SS}^1`$. Either way, we find:
$$\begin{array}{ccc}\mathrm{Ind}D\hfill & =& \underset{R\mathrm{}}{lim}_{|z|=R}\frac{\mathrm{Tr}((_A\varphi )\varphi )}{4\pi \varphi }\hfill \\ & =& \underset{R\mathrm{}}{lim}\frac{1}{2\pi }_{|z|=R}\frac{}{r}\varphi d(\mathrm{arg}z)d\chi .\hfill \end{array}$$
Thus $`\mathrm{Ind}D=m(A,\varphi )=k.`$ Below we will compute the index in another way, which also provides some information on the spatial structure of the zero modes.
## 4 Spectral Data Of A Periodic Monopole
In the previous section we showed that the Nahm transform of a charge $`k`$ periodic monopole is a pair $`(\widehat{A},\mathrm{\Phi })`$, where $`\widehat{A}`$ is a connection on a trivial rank $`k`$ bundle $`\widehat{E}`$ over $`\widehat{X}=\times \mathrm{SS}^1,`$ $`\mathrm{\Phi }`$ is a section of $`\mathrm{End}(\widehat{E})\mathrm{\Omega }_{\widehat{X}}^{1,0}`$, and the pair $`\widehat{A},\mathrm{\Phi }`$ satisfies the Hitchin equations (14,16). Since $`\widehat{X}`$ is noncompact, it is important to determine the behavior of the pair $`(\widehat{A},\mathrm{\Phi })`$ at $`r=\pm \mathrm{}`$. The simplest way to do this uses an algebraic curve associated to the periodic monopole. In this section we explain how to construct this curve and a line bundle over it. These algebro-geometric data associated to a periodic monopole will be called the monopole spectral data.
Let $`B`$ be a (nonunitary) connection on $`E`$ defined by
$$B(x)=A(x)i\varphi (x)d\chi .$$
Let $`\zeta `$. Consider a loop $`\gamma _\zeta :\mathrm{SS}^1X`$ given by
$$\gamma _\zeta :u(z(u),\chi (u))=(\zeta ,u),u/2\pi .$$
We denote the value of $`B(\gamma _\zeta (u))`$ on the vector $`/u`$ by $`B_u`$. Suppose we want to compute the holonomy of $`B`$ along $`\gamma `$. To do this, we must solve the matrix equation
$$\left(\frac{d}{du}iB_u\right)V(\zeta ,u)=0$$
(19)
with the initial condition $`V(\zeta ,0)=1_{2\times 2}.`$ The holonomy is equal to $`V(\zeta ,2\pi ).`$
Note now that the Bogomolny equation implies
$$[_{\overline{z}}iA_{\overline{z}}(\zeta ,u),\frac{d}{du}iB_u]=iF_{\overline{z}\chi }(_{\overline{z}}\varphi i[A_{\overline{z}},\varphi ])|_{z=\zeta }=0.$$
Hence the commutator
$$W(\zeta ,u)=[_{\overline{z}}iA_{\overline{z}},V(z,u)]|_{z=\zeta }$$
also satisfies the differential equation (19). On the other hand, since $`V(\zeta ,0)=1_{2\times 2}`$ for all $`\zeta `$, $`W(\zeta ,0)=0`$ for all $`\zeta `$. The equation (19) being first order, this means that $`W(\zeta ,u)=0`$ for all $`\zeta `$ and $`u`$. Recalling the definition of $`W`$, we see that the characteristic polynomial of $`V(z,u)`$ is a holomorphic function of $`z`$ for any $`u`$. Hence the function
$$F(w,z)=det(wV(z,2\pi ))$$
is a holomorphic function of both $`z`$ and $`w`$. It is also easy to see that $`F(w,z)`$ is gauge-invariant and independent of the choice of origin on the circle parameterized by $`\chi `$.
We define the spectral curve $`𝐒`$ of a periodic monopole to be the zero set of $`F(w,z)`$, i.e. $`𝐒`$ is an algebraic curve in $`^2`$ given by
$$det(wV(z,2\pi ))=0.$$
(20)
Since both $`\varphi `$ and $`A`$ are traceless, $`detV(z,2\pi )=1`$. It follows that $`𝐒`$ does not have common points with the set $`w=0`$ in $`^2`$, and therefore may be regarded as a complete curve in $`\times ^{},`$ where $`^{}`$ is a complex $`w`$–plane with the origin removed.
Let us examine the curve $`𝐒`$ more closely. Since we are dealing with $`SU(2)`$ monopoles, the equation of $`𝐒`$ is really
$$w^2w\mathrm{Tr}V(z,2\pi )+1=0,$$
(21)
i.e. $`𝐒`$ is a double cover of the $`z`$–plane. One can also show that $`\mathrm{Tr}V(z,2\pi )`$ is a degree $`k`$ polynomial in $`z`$. Indeed, we already know that $`\mathrm{Tr}V(z,2\pi )`$ is an entire function of $`z`$. Its behavior for large $`z`$ can be computed from the known behavior of $`A`$ and $`\varphi `$ described by (1.4). This yields
$$\mathrm{Tr}V(z,2\pi )=z^k\mathrm{exp}\left(2\pi 𝔳\right)(1+o(1)),$$
(22)
with $`𝔳=v+ib`$. Since the function $`\mathrm{Tr}V(z,2\pi )`$ is entire and bounded by a multiple of $`z^k`$, it must be a polynomial of degree $`k`$. The leading coefficient of this polynomial is determined by the asymptotic conditions imposed on the monopole (i.e. by $`b`$ and $`v`$), while the remaining $`k`$ coefficients are the moduli of the periodic monopole.
A periodic monopole also provides us with a coherent sheaf $`M`$ on $`𝐒`$, namely the sheaf of eigenvectors of $`V(z,2\pi )`$. The stalk of $`M`$ at a general point is one-dimensional. The direct image of $`M`$ under the projection map $`\pi :𝐒`$ is of course the bundle $`E`$ restricted to $`\chi =0`$. We will call the pair $`(𝐒,M)`$ the spectral data of a periodic monopole.
For a general monopole the curve $`𝐒`$ is nonsingular. If this is the case, then $`M`$ is a line bundle. The reasoning leading to this conclusion is the same as for the Hitchin spectral data.
A periodic monopole with charge $`k`$ can be thought of as consisting of $`k`$ monopoles of charge $`1`$. With the help of the spectral curve one may suggest a precise definition of the location of these constituent monopoles on $``$. These are the points where the holonomy $`V(z,2\pi )`$ has an eigenvalue $`1`$, i.e. the roots of the equation $`\mathrm{Tr}V(z,2\pi )=2.`$ Since $`\mathrm{Tr}V(z,2\pi )`$ is a polynomial of degree $`k`$, for a generic monopole this equation has $`k`$ distinct roots $`\zeta _1,\mathrm{},\zeta _k`$. We expect that when these points are well-separated, the energy density is concentrated in their neighborhood.
If we assume that the curve $`𝐒`$ is nonsingular, then at $`z=\zeta _\alpha `$ the Jordan normal form of $`V(z,2\pi )`$ is
$$\left(\begin{array}{cc}1& 1\\ 0& 1\end{array}\right).$$
This implies that at $`z=\zeta _\alpha `$ the holonomy $`V(z,2\pi )`$ has a single eigenvector with eigenvalue one. In other words, if we consider the restriction of $`E`$ to the $`\mathrm{SS}^1`$ given by $`z=\zeta _\alpha `$, and equip it with a (nonunitary) connection $`B`$, then this bundle has a covariantly constant section unique up to a scalar multiplication. On the other hand, for other values of $`\zeta `$ the holonomy of $`B`$ has both eigenvalues distinct from $`1`$, and the restriction of $`E`$ to the circle does not have sections covariantly constant with respect to $`B`$. This elementary observation plays an important role in the next section.
## 5 Coincidence Of The Spectral Data
The purpose of this section is to demonstrate that the two kinds of spectral data defined in sections 3 and 4 coincide. Recall that starting from a periodic monopole $`(A,\varphi )`$ twisted by $`s=r+it`$ we defined a unitary bundle $`\widehat{E}`$ on $`^{}`$, formed by zero-modes of the twisted Dirac operator $`D^{}`$, as well as a unitary connection on $`\widehat{E}`$, and a Higgs field $`\widehat{\varphi }\mathrm{\Gamma }(\mathrm{End}(\widehat{E})).`$ The Higgs field $`\widehat{\varphi }(s)`$ was defined as a composition of multiplication by the affine coordinate $`z`$ on $`^2`$ and projection to $`\mathrm{Ker}(D^{})`$. The coincidence of the spectral curves $`𝐂`$ and $`𝐒`$ is equivalent to the following statement: if $`\zeta `$ is an eigenvalue of the transformed Higgs field $`\widehat{\varphi }`$ at a point $`s=\sigma `$, then $`e^{2\pi \sigma }`$ is an eigenvalue of the holonomy of $`B=Ai\varphi d\chi `$ around the loop $`\gamma _\zeta `$ which winds around the $`\mathrm{SS}^1`$ at $`z=\zeta `$. This is the statement that will be proved below. We will also show that the zero modes of the Dirac operator $`D^{}`$ are in one-to-one correspondence with the points $`\zeta `$ such that the restriction of $`E`$ to the circle $`z=\zeta `$ has a covariantly constant section (with respect to the connection $`B`$). As explained in the previous section, for a general monopole there are $`k`$ such points, so we see again that $`dim\mathrm{ker}D^{}=k`$.
### 5.1 Cohomological Description Of The Nahm Transform
We proceed to reformulate the Nahm transform of section 3 in cohomological terms. The benefits of such a reformulation will become apparent shortly. In particular the cohomological definition of the transformed Higgs field $`\widehat{\varphi }`$ is extremely simple.
Let us denote by $`\mathrm{\Lambda }^{0,1}(X,E)`$ the bundles on $`X=^2\times \mathrm{SS}^1`$ whose sections have the form $`fd\overline{z}+gd\chi `$, where $`f,g\mathrm{\Gamma }(E).`$ $`\mathrm{\Lambda }^{0,2}(X,E)`$ will denote the bundle whose sections have the form $`fd\overline{z}d\chi `$, where $`f\mathrm{\Gamma }(E)`$. The bundles $`\mathrm{\Lambda }^{0,1}(X,E)`$ and $`\mathrm{\Lambda }^{0,2}(X,E)`$ are subbundles of the bundles of $`E`$-valued differential forms $`\mathrm{\Lambda }^1(X,E)`$ and $`\mathrm{\Lambda }^2(X,E),`$ respectively. Their names betray their origin in the Hodge decomposition of forms on $`^2`$. For uniformity of notation, we also set $`\mathrm{\Lambda }^{0,0}(X,E)=\mathrm{\Gamma }(E)`$.
Pursuing this analogy, we can identify spinor bundles $`S^+(E)`$ and $`S^{}(E)`$ as follows:
$$S^+(E)=\mathrm{\Lambda }^{0,0}(X,E)\mathrm{\Lambda }^{0,2}(X,E),S^{}(E)=\mathrm{\Lambda }^{0,1}(X,E).$$
(23)
To any trivial vector bundle $`E`$ on $`X`$ we can associate a locally free sheaf of vector spaces defined in the following way. Over the whole $`X`$ its space of sections is the space of smooth global sections of $`E`$ which belong to the Schwarz space (i.e. all of their derivatives decay faster than any negative power of $`|z|`$). Over any open set $`𝒪X`$ its space of sections is obtained by restriction from $`X`$. In what follows we will identify a trivial vector bundle on $`X`$ and the corresponding sheaf.
Let us define differentials
$$\overline{𝒟}_p=d\overline{z}2(\frac{}{\overline{z}}iA_{\overline{z}})+d\chi (\frac{}{\chi }iA_\chi \varphi +s),$$
(24)
acting from $`\mathrm{\Lambda }^{0,p}(X,E)`$ to $`\mathrm{\Lambda }^{0,p+1}(X,E)`$, $`p=0,1.`$ Note that $`\overline{𝒟}_1\overline{𝒟}_0=0`$, as a consequence of the Bogomolny equation. Thus we have a differential complex $`K`$:
$$K:0\mathrm{\Lambda }^{0,0}\stackrel{\overline{𝒟}_0}{}\mathrm{\Lambda }^{0,1}\stackrel{\overline{𝒟}_1}{}\mathrm{\Lambda }^{0,2}0.$$
(25)
Since the operators $`\overline{𝒟}_p`$ depend on $`e^{2\pi s}^{}`$, so does the complex $`K`$, and it would be more precise to call it $`K_s`$. We will omit the subscript $`s`$ where this cannot lead to confusion.
Since all the bundles we are dealing with are trivial, we are free to identify $`SES^+(E)S^{}(E)`$. Then the twisted Dirac operator $`D:SESE`$ becomes simply
$$D=\overline{𝒟}_0\overline{𝒟}_1^{},$$
and its adjoint $`D^{}:S^{}S^+`$ is $`\overline{𝒟}_0^{}\overline{𝒟}_1`$.
As explained in section 3, the only $`L^2`$ solution of the equation $`D\psi =0,\psi \mathrm{\Gamma }(SE),`$ is the trivial one. In other words the equation $`\overline{𝒟}_0\mathrm{\Psi }\overline{𝒟}_1^{}\mathrm{\Psi }=0`$, $`\mathrm{\Psi }\mathrm{\Lambda }^{0,0}(X,E)\mathrm{\Lambda }^{0,2}(X,E),`$ has only the trivial $`L^2`$ solution. It follows that the complex (25) is exact in the first and the third terms: $`H^0(K)=H^2(K)=0`$. We want to show that $`H^1(K)`$ is isomorphic to the kernel of the twisted Dirac operator $`D^{}`$. In one direction this is easy: for any $`\psi \mathrm{Ker}D^{}`$ we have $`\overline{𝒟}_1\psi =\overline{𝒟}_0^{}\psi =0`$, and therefore $`\psi `$ is a harmonic representative of a class in $`H^1(K)`$. It is obvious that this map from $`\mathrm{Ker}D^{}`$ to $`H^1(K)`$ is injective. The inverse map is constructed as follows. For any representative $`\theta `$ of a class $`[\theta ]H^1(K)`$ we have to find $`\rho \mathrm{\Lambda }^{0,0}(X,E)`$ such that $`(\overline{𝒟}_0^{}\overline{𝒟}_1)(\theta +\overline{𝒟}_0\rho )=0`$. Since $`H^0(K)`$ is trivial, the kernel of the operator $`\overline{𝒟}_0^{}\overline{𝒟}_0`$ is empty and the operator itself is invertible. Thus we may solve for the function $`\rho `$:
$$\rho =(\overline{𝒟}_0^{}\overline{𝒟}_0)^1\overline{𝒟}_0^{}\theta .$$
(26)
This yields a map from $`H^1(K)`$ to $`\mathrm{Ker}D^{}.`$ It is easy to see that it is the inverse of the map from $`\mathrm{Ker}D^{}`$ to $`H^1(K)`$ constructed above.
Since $`\overline{𝒟}_p`$ and multiplication by $`z`$ commute, the action of $`\widehat{\varphi }`$ on $`H^1(K)`$ is simply multiplication by $`z`$, without a need for a projection. This is the reason the cohomological description of $`\mathrm{Ker}D^{}`$ is useful.
### 5.2 Explicit Argument
Suppose the point $`(\zeta ,e^{2\pi \sigma })\times ^{}`$ belongs to the Hitchin spectral curve $`𝐂`$. In this case there exists a nonzero vector $`\mathrm{\Theta }H^1(K_\sigma )`$ such that
$$\widehat{\varphi }(\sigma )\mathrm{\Theta }=\zeta \mathrm{\Theta }.$$
(27)
As explained above, $`\widehat{\varphi }`$ acts on $`H^1(K_\sigma )`$ as multiplication by $`z`$. Let $`\theta `$ be a one-form representing $`\mathrm{\Theta }H^1(K_\sigma )`$. Then the equation (27) means that there exists $`\psi \mathrm{\Lambda }^{0,0}(X,E)`$ such that
$$(z\zeta )\theta =\overline{𝒟}_1\psi .$$
(28)
It follows that $`\overline{𝒟}_1\psi `$ vanishes at $`z=\zeta `$. In particular we have
$$\left(\frac{}{\chi }iA_\chi \varphi +\sigma \right)\psi |_{z=\zeta }=0,$$
(29)
i.e. the restriction of $`\psi `$ to the circle $`z=\zeta `$ is covariantly constant with respect to the connection $`B+i\sigma d\chi `$. If $`\psi `$ is not identically zero on the circle $`z=\zeta `$, this implies that the holonomy matrix $`V(\zeta ,2\pi )`$ has an eigenvalue equal to $`e^{2\pi \sigma }`$, and consequently the point $`(\zeta ,e^{2\pi \sigma })`$ belongs to the monopole spectral curve $`𝐒`$.
To complete the proof of $`𝐂=𝐒`$ it remains to show that $`\psi `$ does not vanish identically on the circle $`z=\zeta `$. Suppose it does vanish. Then (28) implies that on the circle $`z=\zeta `$ we have $`\frac{^j}{\overline{z}^j}\psi =0`$ for all $`j0`$. It follows that the function $`\psi `$ has the form $`\psi =(z\zeta )a(z,\overline{z},\chi )+b(z,\overline{z},\chi )`$, where both $`a`$ and $`b`$ are smooth, and as $`z\zeta `$ the function $`b`$ approaches zero faster than any power of $`|z\zeta |`$. Hence $`\psi `$ is divisible by $`(z\zeta )`$, i.e. there exists a smooth function $`\phi \mathrm{\Lambda }^{0,0}(X,E)`$ such that $`\psi =(z\zeta )\phi .`$ Then $`\theta =\overline{𝒟}_0\phi `$, which contradicts the assumption that $`\theta `$ represents a nontrivial class in $`H^1(K)`$.
### 5.3 Cohomological Argument
Consider a complex of sheaves of vector spaces:
$$0E\stackrel{z\zeta }{}E\stackrel{rest.}{}E|_{z=\zeta }\stackrel{}{}0,$$
(30)
where the second map is multiplication by $`z\zeta `$, and the map $`rest.`$ is restriction to the circle $`z=\zeta `$. This complex fails to be exact in the second term. Nevertheless, as shown below, there is a long exact sequence in $`\overline{𝒟}`$ cohomology:
$$0H_{\overline{𝒟}}^0(\mathrm{SS}^1,E|_{z=\zeta })H_{\overline{𝒟}}^1(X,E)\stackrel{z\zeta }{}H_{\overline{𝒟}}^1(X,E)\stackrel{rest.}{}H_{\overline{𝒟}}^1(\mathrm{SS}^1,E|_{z=\zeta })0.$$
(31)
Here $`H_{\overline{𝒟}}^1(X,E)`$ is the same as $`H^1(K)`$, while $`H_{\overline{𝒟}}^j(\mathrm{SS}^1,E|_{z=\zeta })`$ is the $`j`$-th cohomology of the restriction of $`K`$ to the circle $`z=\zeta `$. Note that the restriction of $`\overline{𝒟}_p`$ to $`z=\zeta `$ is simply the covariant differential with respect to the connection $`B+isd\chi `$ restricted to $`z=\zeta `$.
To understand where this exact sequence comes from, it is helpful to think about solutions of self-duality equations on $`\times T^2`$. Periodic monopoles are a particular class of such solutions which are invariant with respect to translations in one direction on the torus. Now the variable $`\chi `$ gets promoted to a complex variable parameterizing the universal cover of the torus, and the operator $`\overline{𝒟}`$ becomes simply a $`\overline{}`$–operator on the bundle $`E`$. Thus $`E`$ has a structure of a holomorphic bundle over $`\times T^2`$. The cohomology of $`\overline{𝒟}`$ is the Dolbeault cohomology of $`E`$. If we forget about noncompactness, the Dolbeault cohomology can be identified with the Čech cohomology of the holomorphic bundle $`E`$. On the other hand, if we work in the holomorphic category, the sequence of sheaves (30) is exact and hence induces a long exact sequence of Čech cohomology groups.
In our situation, we cannot use the holomorphic interpretation. Instead, in the next subsection we derive the exact sequence (31) from a spectral sequence of a double complex.
The coincidence of the Hitchin and the monopole spectral data is an immediate consequence of the exactness of the sequence (31). Indeed, if the point $`(\zeta ,\mathrm{exp}(2\pi \sigma ))`$ belongs to the Hitchin spectral curve $`𝐂`$, then the kernel of the map $`(z\zeta )`$ from $`H^1(K_\sigma )`$ to $`H^1(K_\sigma )`$ is nontrivial. But the cohomology exact sequence implies an isomorphism
$$\mathrm{Ker}(z\zeta )H_{\overline{𝒟}}^0(\mathrm{SS}^1,E|_{z=\zeta }).$$
(32)
Therefore $`H_{\overline{𝒟}}^0(\mathrm{SS}^1,E|_{z=\zeta })`$ is nontrivial as well. This means that the holonomy of $`B`$ along the circle $`z=\zeta `$ has $`\mathrm{exp}(2\pi \sigma )`$ as one of its eigenvalues. Thus the point $`(\zeta ,\mathrm{exp}(2\pi \sigma ))\times ^{}`$ belongs to the monopole spectral curve $`𝐒`$. Moreover, the fibers of the spectral line bundles on $`𝐂`$ and $`𝐒`$ are given by $`\mathrm{Ker}(z\zeta )`$ and $`H_{\overline{𝒟}}^0(\mathrm{SS}^1,E|_{z=\zeta }),`$ respectively. Thus we also get an isomorphism of the line bundles.
### 5.4 Exactness Of The Cohomology Sequence
Consider again the complex of sheaves
$$0E\stackrel{(z\zeta )}{}E\stackrel{rest.}{}E|_{z=\zeta }0,$$
(33)
where $`rest.`$ is the restriction to $`z=\zeta `$. Since $`\overline{𝒟}_p,p=0,1,`$ commutes with $`(z\zeta )`$ and $`rest.`$, this complex is included in a double complex $`D^{p,q}`$:
Computing the cohomology of the rows, we obtain the first term of the “vertical” spectral sequence :
On the second level the “vertical” spectral sequence degenerates to zero: $`E_{\mathrm{}}^{p,q}=0`$.
Now let us compute the “horizontal” spectral sequence. Its first term is simply the $`\overline{𝒟}`$ cohomology of $`D^{p,q}`$: $`\stackrel{~}{E}_1^{p,q}=H_{\overline{𝒟}}(D^{p,q})`$. The second term $`\stackrel{~}{E}_2^{p,q}`$ is given by
At the next level the “horizontal” spectral sequence degenerates. Since the “vertical” spectral sequence converges to zero, so should the “horizontal” one. From this we infer the isomorphisms
$`\mathrm{Ker}rest.|_{H^1(K)}`$ $`\mathrm{Im}(z\zeta )|_{H^1(K)},`$
$`H_{\overline{𝒟}}^1(E|_{z=\zeta })`$ $`\mathrm{Im}rest.|_{H^1(K)},`$
$`\mathrm{Ker}(z\zeta )|_{H^1(K)}`$ $`H_{\overline{𝒟}}^0\left(E|_{z=\zeta }\right).`$
These isomorphisms are equivalent to the exactness of the cohomology sequence (31).
### 5.5 Revisiting The Index Computation
Using the spectral sequence technology, we can give another proof that $`dimH^1(K)=k.`$ The advantage of this method of proof is that it makes it clear that the zero modes of the Dirac operator are “localized” near the points $`z=\zeta _1,\mathrm{},\zeta _k`$.
Consider a bundle morphism $`\mathrm{\Delta }`$ from $`E`$ to $`E`$ defined as multiplication by the polynomial
$$det(z\widehat{\varphi }|_{s=0})=(z\zeta _1)\mathrm{}(z\zeta _k).$$
Here $`\zeta _1,\mathrm{},\zeta _k`$ are the roots of the characteristic polynomial of $`\widehat{\varphi }`$, or equivalently the solutions of the equation $`\mathrm{Tr}V(z,2\pi )=2.`$ Suppose all $`\zeta _i`$ are distinct. Let $`Y`$ be the restriction of the sheaf $`E`$ to the union of $`k`$ circles $`z=\zeta _1,\mathrm{},z=\zeta _k`$, and let $`rest.`$ be the restriction map. Obviously, $`rest.\mathrm{\Delta }=0`$, so we get a complex
$$0E\stackrel{\Delta }{}E\stackrel{rest.}{}Y0.$$
As before, this complex is not exact in the middle term, but nevertheless leads to an exact cohomology sequence:
$$0H_{\overline{𝒟}}^0(Y)\stackrel{}{}H^1(K)\stackrel{\Delta }{}H^1(K)\stackrel{rest.}{}H_{\overline{𝒟}}^1(Y)0.$$
The proof of exactness is identical to the one given above. Now note that the map $`\mathrm{\Delta }`$ sends $`H^1(K)`$ to zero by virtue of the Cayley-Hamilton theorem on the characteristic polynomial of a matrix. Hence $`H^1(K)`$ is isomorphic to $`H_{\overline{𝒟}}^0(Y)`$. On the other hand, if all the numbers $`\zeta _\alpha `$ are distinct, we have
$$H_{\overline{𝒟}}^0(Y)=_{\alpha =1}^kH_{\overline{𝒟}}^0\left(E|_{z=\zeta _\alpha }\right)^k.$$
Hence $`dimH^1(K)=k`$. Moreover, we see that each of the circles $`z=\zeta _\alpha `$ gives rise to a vector in $`H^1(K)\mathrm{Ker}D^{}`$, and all these vectors are linearly independent. Thus we may think of the $`k`$ zero modes of the Dirac operator $`D^{}`$ as “localized” in the neighborhood of $`k`$ circles $`z=\zeta _\alpha ,\alpha =1,\mathrm{},k.`$
## 6 Asymptotic Behavior Of The Hitchin Data
The fact that the spectral curves $`𝐂`$ and $`𝐒`$ coincide provides a wealth of information about periodic monopoles. In particular, it allows to determine the behavior of the Higgs field $`\widehat{\varphi }`$ for $`\mathrm{Re}s\pm \mathrm{}.`$
Let us rewrite the equation of the curve $`𝐒`$ in the following form:
$$f(z)\mathrm{exp}(2\pi s)\mathrm{exp}(2\pi s)=0.$$
(34)
Here we used the identification of the eigenvalue $`w`$ of $`V(z,2\pi )`$ with $`\mathrm{exp}(2\pi s)`$. Recalling the definition of the Hitchin spectral curve $`𝐂`$, we infer that all coefficients of the characteristic polynomial of $`\widehat{\varphi }`$ except $`det\widehat{\varphi }`$ are independent of $`s`$. Furthermore, we know from section 4 that $`f(z)`$ is a degree $`k`$ polynomial in $`z`$ with leading coefficient $`\mathrm{exp}(2\pi 𝔳)`$. It follows that the determinant of $`\widehat{\varphi }`$ is given by
$$det\widehat{\varphi }=(1)^{k+1}\left(e^{2\pi s}+e^{2\pi s}\right)\mathrm{exp}(2\pi 𝔳).$$
Another piece of information comes from the equation (15) which says that the curvature of $`\widehat{A}`$ is proportional to the restriction of the operator $`(D^{}D)^1`$ to the subspace $`\mathrm{Ker}D^{}`$. The operator $`(D^{}D)^1`$ is a integral operator on the space of $`L^2`$ sections of $`ES`$, and from (10) it is clear that its norm vanishes in the limit $`\mathrm{Re}s\pm \mathrm{}`$. (In fact, it is easy to see that the norm is bounded from above by a multiple of $`1/|\mathrm{Re}s|^{3/2}.`$) Hence the curvature of $`\widehat{A}`$ also goes to zero in this limit. Bogomolny equations then imply that $`[\widehat{\varphi }^{},\widehat{\varphi }]0`$ asymptotically.
Combining this with the information about the characteristic polynomial of $`\widehat{\varphi }`$, we infer that for $`\mathrm{Re}s\pm \mathrm{}`$ the Higgs field $`\widehat{\varphi }`$ behaves as follows:
$$\widehat{\varphi }(s)\mathrm{exp}\left(\frac{2\pi }{k}\left(𝔳+\frac{i}{2}\right)\right)g_\pm (s)e^{\pm \frac{2\pi s}{k}}(1+o(1))\mathrm{diag}(1,\omega ,\omega ^2,\mathrm{},\omega ^{k1})g_\pm (s)^1.$$
(35)
Here $`\omega `$ is a $`k`$–th root of unity and $`g_\pm (s)`$ are multi-valued functions on $`\widehat{X}`$ with values in $`U(k)`$. In order for $`\widehat{\varphi }`$ to be well-defined, the functions $`g_\pm `$ must satisfy
$$g_\pm (s+i)=g_\pm (s)V^{\pm 1}e^{i\beta }.$$
Here $`\beta `$ is a real number, and $`VSU(k)`$ is the so-called “shift” matrix:
$$V=\left(\begin{array}{cccccc}0& 0& 0& \mathrm{}& 0& 1\\ 1& 0& 0& \mathrm{}& 0& 0\\ 0& 1& 0& \mathrm{}& 0& 0\\ 6\\ 0& 0& 0& \mathrm{}& 0& 0\\ 0& 0& 0& \mathrm{}& 1& 0\end{array}\right).$$
(36)
We can reformulate these results as follows. The Nahm transform of a periodic monopole of charge $`k`$ is a pair $`(\widehat{A},\widehat{\varphi })`$ satisfying the $`U(k)`$ Hitchin equations and the following asymptotic conditions:
(i) The functions $`\mathrm{Tr}\widehat{\varphi }(s)^\alpha ,\alpha =1,\mathrm{},k1`$ are bounded;
(ii) The function $`\mathrm{exp}(2\pi s)det\widehat{\varphi }(s)`$ behaves as
$$(1)^{k+1}\mathrm{exp}(2\pi 𝔳)+O(\mathrm{exp}(2\pi s))$$
for $`\mathrm{Re}s\pm \mathrm{}`$;
(iii) $`F_{z\overline{z}}^2{\displaystyle \frac{C}{|\mathrm{Re}s|^3}}.`$
Since the functions $`\mathrm{Tr}\widehat{\varphi }(s)^\alpha `$ and $`det\widehat{\varphi }(s)`$ are holomorphic functions on $`^{}\times \mathrm{SS}^1`$ by virtue of the Hitchin equations, the first two conditions are equivalent to the statement that the spectral curve of $`(\widehat{A},\widehat{\varphi })`$ has the form (34), with $`f(z)`$ being a polynomial of degree $`k`$ with the leading coefficient $`\mathrm{exp}(2\pi 𝔳)`$, and the rest of the coefficients being arbitrary constants.
In the next three sections we will show that the correspondence between the solutions of Hitchin equations satisfying (i)-(iii) and periodic monopoles is one-to-one.
## 7 The Inverse Nahm Transform
In this section we show how to associate a periodic $`SU(2)`$ monopole of charge $`k`$ to any solution of $`U(k)`$ Hitchin equations on $`\widehat{X}`$ with asymptotic behavior as above. This procedure will be called the inverse Nahm transform. It will take us from Hitchin data associated with a bundle $`\widehat{E}\widehat{X}`$ to monopole data on a bundle $`\stackrel{ˇ}{\widehat{E}}X`$. Later on, in section 9, we will show that the monopole on $`\stackrel{ˇ}{\widehat{E}}`$ coincides with that on $`E`$, so in this section we shall use a simplified notation in which the symbol $`\stackrel{ˇ}{\widehat{}}`$ is omitted. Since the original monopole data on $`E`$ are not used in this section, this should not lead to confusion.
Let $`\widehat{E}`$ be a trivial unitary rank $`k`$ bundle over $`\widehat{X}=^{}`$, $`\widehat{A}`$ be a connection on $`\widehat{E}`$ and $`\widehat{\varphi }`$ be a section of $`\mathrm{End}(\widehat{E})`$. Furthermore, let the pair $`(\widehat{A},\widehat{\varphi })`$ be a solution of $`U(k)`$ Hitchin equations on $`\widehat{X}`$ such that $`\widehat{\varphi }`$ has the asymptotics as in (8). Let $`\widehat{L}`$ be a trivial line bundle over $`\widehat{X}`$ with a flat unitary connection $`\widehat{a}`$ such that the holonomy of $`\widehat{a}`$ around the positively oriented loop encircling the origin of $`^{}`$ is $`\mathrm{exp}(i\chi ).`$ The variable $`\chi `$ is assumed to take values in the interval $`[0,2\pi ].`$ Consider a Dirac–type operator $`\widehat{D}:\widehat{E}\widehat{L}^2\widehat{E}\widehat{L}^2`$ given by
$$\widehat{D}=\left(\begin{array}{cc}\widehat{\varphi }+z& 2_{\widehat{A}+\widehat{a}}\\ 2\overline{}_{\widehat{A}+\widehat{a}}& \widehat{\varphi }^{}+\overline{z}.\end{array}\right)$$
Here $`z`$ is a complex parameter. The operator $`\widehat{D}`$ is Fredholm for any $`z`$ and $`\chi `$ because $`\widehat{\varphi }`$ grows without bound as $`t\pm \mathrm{}`$.
The Weitzenbock formula for $`\widehat{D}`$ reads:
$$\widehat{D}^{}\widehat{D}=\left(\begin{array}{cc}(\widehat{\varphi }^{}\overline{z})(\widehat{\varphi }z)4_{\widehat{A}+\widehat{a}}\overline{}_{\widehat{A}+\widehat{a}}& 2_{\widehat{A}}\widehat{\varphi }^{}\\ 2\overline{}_{\widehat{A}}\widehat{\varphi }& (\widehat{\varphi }z)(\widehat{\varphi }^{}\overline{z})4\overline{}_{\widehat{A}+\widehat{a}}_{\widehat{A}+\widehat{a}}.\end{array}\right)$$
If the Hitchin equations are satisfied, then this formula simplifies:
$$\widehat{D}^{}\widehat{D}=_{\widehat{A}+\widehat{a}}^2+\frac{1}{2}((\widehat{\varphi }^{}\overline{z})(\widehat{\varphi }z)+(\widehat{\varphi }z)(\widehat{\varphi }^{}\overline{z})).$$
This operator is clearly positive definite on the space of smooth rapidly decreasing sections of $`\widehat{E}\widehat{L}^2`$. It is easy to see that any $`L^2`$ eigenvector of $`\widehat{D}`$ with zero eigenvalue must be smooth and decreasing faster than any negative power of $`r=\mathrm{Re}s`$, hence $`\widehat{D}`$ has trivial $`L^2`$ kernel. Thus the dimension of the kernel of $`\widehat{D}^{}`$ is minus the index of $`\widehat{D}`$.
Computing the $`L^2`$ index of $`\widehat{D}`$ turns out to be rather tricky. We will do it in the next section by reinterpreting $`\mathrm{Ker}\widehat{D}^{}`$ as a certain cohomology group and computing it using the spectral sequence of a double complex, similarly to how it was done in Section 5. The result of this computation is that $`\mathrm{Ker}\widehat{D}^{}`$ has dimension $`2`$ for all $`z`$ and $`\chi `$.
We conclude that $`\mathrm{Ker}\widehat{D}^{}`$ forms a trivial rank $`2`$ bundle on the manifold $`X=\times \mathrm{SS}^1`$ parameterized by $`(z,\chi )`$. Since the elements of the kernel are square-integrable, we have a well-defined Hermitian inner product on $`\mathrm{Ker}\widehat{D}^{}`$ for all $`z,\chi `$. In this way we obtain a unitary rank 2 bundle $`E`$ over $`X`$.
Now we need to define a connection $`A`$ on $`E`$ and a traceless Hermitian section $`\varphi `$ of $`\mathrm{End}(E)`$. The connection on $`\mathrm{Ker}\widehat{D}`$ is induced from a trivial connection on a trivial infinite-dimensional bundle on $`X`$ whose fiber consists of all smooth $`L^2`$ sections of $`\widehat{E}\widehat{S}\widehat{L}`$. If we introduce the projectors $`\widehat{P}=1\widehat{D}(\widehat{D}^{}\widehat{D})^1\widehat{D}^{}`$ and $`\widehat{Q}=1\widehat{P}`$, we may write
$$d_A=\widehat{P}d=\widehat{P}\left(dz\frac{}{z}+d\overline{z}\frac{}{\overline{z}}+d\chi \frac{}{\chi }\right).$$
The value of the Higgs field $`\varphi `$ at a point $`xX`$ is a linear map $`EE`$ defined as a composition of multiplication by $`r`$ and projection to $`\mathrm{Ker}D^{}`$, i.e.
$$\varphi =\widehat{P}r.$$
It remains to show that $`\varphi `$ and $`A`$ satisfy the Bogomolny equation (1). In the coordinates $`z,\chi `$ used above this equation is equivalent to a pair of equations
$`F_{\overline{z}\chi }=i\overline{}_A\varphi ,`$ (37)
$`F_{z\overline{z}}={\displaystyle \frac{i}{2}}\left(i_{/\chi }d_A\right)\varphi .`$ (38)
Here $`\overline{}_A`$ means $`i_{/\overline{z}}d_A.`$
To show that these equations are satisfied, we have to use the commutation relations
$$\begin{array}{cccc}[\widehat{D},\frac{}{\chi }]=i\sigma _2,\hfill & [\widehat{D},\frac{}{z}]=p_+,\hfill & [\widehat{D},\frac{}{\overline{z}}]=p_{},\hfill & [\widehat{D},r]=\sigma _1,\hfill \\ [\widehat{D}^{},\frac{}{\chi }]=i\sigma _2,\hfill & [\widehat{D}^{},\frac{}{z}]=p_{},\hfill & [\widehat{D}^{},\frac{}{\overline{z}}]=p_+,\hfill & [\widehat{D}^{},r]=\sigma _1,\hfill \end{array}$$
and the fact that $`\widehat{D}^{}\widehat{D}`$ commutes with all $`\sigma _i`$. We find for the curvature of $`A`$:
$$\begin{array}{ccc}\hfill F_{\overline{z}\chi }& =& i\widehat{P}(\overline{}\widehat{Q}_\chi \widehat{Q}_\chi \widehat{Q}\overline{}\widehat{Q})\hfill \\ & =& i\widehat{P}([\overline{},\widehat{D}](\widehat{D}^{}\widehat{D})^1[_\chi ,\widehat{D}^{}][_\chi ,\widehat{D}](\widehat{D}^{}\widehat{D})^1[\overline{},\widehat{D}^{}])\hfill \\ & =& \widehat{P}(p_{}(\widehat{D}^{}\widehat{D})^1\sigma _2+\sigma _2(\widehat{D}^{}\widehat{D})^1p_+)\hfill \\ & =& 2i\widehat{P}(\widehat{D}^{}\widehat{D})^1\sigma _{},\hfill \\ \hfill F_{z\overline{z}}& =& i\widehat{P}(\widehat{Q}\overline{}\widehat{Q}\overline{}\widehat{Q}\widehat{Q})\hfill \\ & =& i\widehat{P}([,\widehat{D}](\widehat{D}^{}\widehat{D})^1[\overline{},\widehat{D}^{}][\overline{},\widehat{D}](\widehat{D}^{}\widehat{D})^1[,\widehat{D}^{}])\hfill \\ & =& i\widehat{P}(p_+(\widehat{D}^{}\widehat{D})^1p_+p_{}(\widehat{D}^{}\widehat{D})^1p_{})\hfill \\ & =& i\widehat{P}(\widehat{D}^{}\widehat{D})^1\sigma _3.\hfill \end{array}$$
(39)
The covariant derivatives of $`\varphi `$ can also be easily computed:
$$\begin{array}{ccc}\hfill \overline{}_A\varphi & =& [\widehat{P}\overline{},\widehat{P}r]\hfill \\ & =& \widehat{P}r\widehat{Q}\overline{}\widehat{P}\overline{}\widehat{Q}r\hfill \\ & =& \widehat{P}([r,\widehat{D}](\widehat{D}^{}\widehat{D})^1[\widehat{D}^{},\overline{}][\overline{},\widehat{D}](\widehat{D}^{}\widehat{D})^1[\widehat{D}^{},r])\hfill \\ & =& \widehat{P}(\sigma _1(\widehat{D}^{}\widehat{D})^1p_++p_{}(\widehat{D}^{}\widehat{D})^1\sigma _1)\hfill \\ & =& 2\widehat{P}(\widehat{D}^{}\widehat{D})^1\sigma _{},\hfill \\ \hfill (i_{/\chi }d_A)\varphi & =& [\widehat{P}_\chi ,\widehat{P}r]\hfill \\ & =& \widehat{P}r\widehat{Q}_\chi \widehat{P}_\chi \widehat{Q}r\hfill \\ & =& \widehat{P}([r,\widehat{D}](\widehat{D}^{}\widehat{D})^1[\widehat{D}^{},_\chi ][_\chi ,\widehat{D}](\widehat{D}^{}\widehat{D})^1[\widehat{D}^{},r])\hfill \\ & =& i\widehat{P}(\sigma _1(\widehat{D}^{}\widehat{D})^1\sigma _2\sigma _2(\widehat{D}^{}\widehat{D})^1\sigma _1)\hfill \\ & =& 2\widehat{P}(\widehat{D}^{}\widehat{D})^1\sigma _3.\hfill \end{array}$$
(40)
Comparing (39) and (40), we see that $`A`$ and $`\varphi `$ indeed satisfy the Bogomolny equation.
## 8 Spectral Data And The Inverse Nahm Transform
In the previous section we showed that the inverse Nahm transform applied to a solution of Hitchin equations on $`\widehat{X}^{}`$ yields a solution of the Bogomolny equation on $`X^2\times \mathrm{SS}^1`$. In this section we prove that if the solution of the Hitchin equations has the asymptotics (35), then the corresponding solution of the Bogomolny equation has the asymptotics (1.4). The use of the monopole spectral data greatly facilitates this proof. We first give a “pedestrian” proof of the coincidence of the Hitchin and monopole spectral data, and then a more conceptual one using the spectral sequence of a double complex. This spectral sequence will also be used to compute the index of the Dirac operator $`\widehat{D}`$, thereby filling a gap in the derivation of Section 7. In fact, we will show that the zero modes of the twisted Dirac operator $`\widehat{D}_{z,\chi }`$ are in one-to-one correspondence with the points on $`\widehat{X}`$ where $`\widehat{\varphi }`$ has an eigenvalue $`z`$.
### 8.1 Cohomological Interpretation Of The Inverse Nahm Transform
In order to give a cohomological interpretation of the inverse Nahm transform, let us consider the following complex constructed from the sheaves of vector spaces $`\widehat{\mathrm{\Lambda }}^0=\mathrm{\Lambda }^{0,0}(\widehat{X},\widehat{E})`$ and $`\widehat{\mathrm{\Lambda }}^1=\mathrm{\Lambda }^{0,1}(\widehat{X},\widehat{E})`$:
$$0\widehat{\mathrm{\Lambda }}^0\stackrel{\widehat{\delta }_0}{}\widehat{\mathrm{\Lambda }}^0\widehat{\mathrm{\Lambda }}^1\stackrel{\widehat{\delta }_1}{}\widehat{\mathrm{\Lambda }}^10,$$
(41)
with $`\widehat{\delta }_0`$ and $`\widehat{\delta }_0`$ defined by
$$\widehat{\delta }_0:f\left(\begin{array}{c}(\widehat{\varphi }\zeta )f\\ 2\overline{}_{\widehat{A}+\widehat{a}}f\end{array}\right),\widehat{\delta }_1:\left(\begin{array}{c}g_0\\ g_1\end{array}\right)\left(2\overline{}_{\widehat{A}+\widehat{a}}g_0(\widehat{\varphi }\zeta )g_1\right).$$
(42)
Since every element $`g_1\widehat{\mathrm{\Lambda }}^1`$ has the form $`g_1=g(s)d\overline{s}`$, we can identify $`\widehat{\mathrm{\Lambda }}^0g`$ with $`\widehat{\mathrm{\Lambda }}^1g_1`$.
In terms of this complex $`\widehat{K}_{z,\chi }`$ the Dirac operator $`\widehat{D}^{}`$ of section 7 is given by
$$\widehat{D}^{}=\widehat{\delta }_0^{}\widehat{\delta }_1.$$
(43)
One can easily see that the square-integrable zero modes of $`\widehat{D}^{}`$ are in one-to-one correspondence with the first $`L^2`$ cohomology of the above complex. Thus, if we denote the inverse Nahm transform of the bundle $`\widehat{E}`$ by $`\stackrel{ˇ}{\widehat{E}}`$, we have a canonical identification of the fiber of $`\stackrel{ˇ}{\widehat{E}}`$ at a point $`(z,\chi )X`$ with $`H^1(\widehat{K}_{z,\chi })`$. In other words, the spaces $`H^1(\widehat{K}_{z,\chi })`$ form a vector bundle over $`X`$ which is canonically isomorphic to $`\stackrel{ˇ}{\widehat{E}}`$.
At this stage of the discussion it becomes crucial to keep track of the periodicity conditions along the $`\chi `$ and $`t`$ directions. Let us denote the circles parameterized by $`\chi `$ and $`t`$ by $`\mathrm{SS}^1`$ and $`\widehat{\mathrm{SS}}^1`$, respectively. Previously we worked with one circle at a time and could choose a trivialization of any bundle on a circle so that the components of a section be periodic functions. However, if one considers a bundle on $`\mathrm{SS}^1\times \widehat{\mathrm{SS}}^1`$, both circles are in the game. If the bundle on $`\mathrm{SS}^1\times \widehat{\mathrm{SS}}^1`$ is nontrivial, then there is no trivialization in which sections are periodic functions along both periodic directions. This is in fact what happens in our case.
Let us introduce some notation. The circle $`\widehat{\mathrm{SS}}^1`$ parameterizes line bundles with a unitary connection on $`\mathrm{SS}^1`$. Namely, a point $`t\widehat{\mathrm{SS}}^1`$ corresponds to a line bundle with a monodromy $`e^{2\pi it}`$. We will denote this line bundle with a connection by $`L_t`$. If one chooses a “periodic” trivialization alluded to above, then the connection on $`L_t`$ is $`td\chi `$. Alternatively, if one chooses a “quasiperiodic” trivialization in which sections are represented by functions satisfying $`f(\chi +2\pi )=e^{2\pi it}f(\chi )`$, then the connection on $`L_t`$ is trivial. Conversely, $`\mathrm{SS}^1`$ parameterizes unitary line bundles on $`\widehat{\mathrm{SS}}^1`$; the point $`\chi \mathrm{SS}^1`$ corresponds to a line bundle $`\widehat{L}_\chi `$ whose monodromy is $`e^{i\chi }`$.
Recall now the definition of the Poincaré line bundle $`𝒫`$ on $`\mathrm{SS}^1\times \widehat{\mathrm{SS}}^1`$ (see e.g. ). It is a line bundle with a unitary connection whose restriction to any circle $`t=t_0`$ is isomorphic to $`L_{t_0}`$ (as a line bundle with a connection), while its restriction to any circle $`\chi =\chi _0`$ is isomorphic to $`\widehat{L}_{\chi _0}`$. The curvature of the connection on $`𝒫`$ is given by $`d\chi dt`$. The Poincaré line bundle is nontrivial and has the first Chern class equal to $`1`$. If we choose a trivialization of $`𝒫`$ such that the connection on $`𝒫`$ is given by $`𝒜=\chi dt`$, then its sections are represented by functions $`f(\chi ,t)`$ satisfying $`f(\chi +2\pi ,t)=e^{2\pi it}f(\chi ,t)`$ and $`f(\chi ,t+1)=f(\chi ,t)`$. Similarly, the dual line bundle $`𝒫^{}`$ has a connection 1-form $`\chi dt`$, and its sections are represented by functions $`f(\chi ,t)`$ satisfying $`f(\chi +2\pi ,t)=e^{2\pi it}f(\chi ,t)`$ and $`f(\chi ,t+1)=f(\chi ,t)`$.
By making use of $`𝒫^{}`$, the inverse Nahm transform can be rephrased as follows. We pull back the bundle $`\widehat{E}`$ to $`X\times \widehat{X}`$ using the natural projection $`\widehat{\pi }:X\times \widehat{X}\widehat{X}`$. Then we twist $`\widehat{\pi }^{}(\widehat{E})`$ by a line bundle with a unitary connection $`\chi dt`$ and trivial periodicity condition in the $`t`$ direction, i.e. by a line bundle $`𝒫^{}`$. The complex $`\widehat{K}`$ is also twisted by $`𝒫^{}`$, as is clear from its definition, and in addition by the Higgs field $`zds`$. Finally, we form a bundle on $`X`$ whose fiber over $`x`$ is the first cohomology group of the complex $`\widehat{K}_{z,\chi }`$. The direct Nahm transform can be similarly reformulated using the line bundle $`𝒫`$. (This description suggests that it is useful to think about the derived functor of the Nahm transform, which reduces to the ordinary Nahm transform when both the initial and the transformed complexes happen to have only a single nonvanishing cohomology).
The upshot of this discussion is that sections of $`𝒫^{}\times \widehat{E}`$ should be thought of as vector-valued functions on $`X\times \widehat{X}`$ satisfying
$$e^{2\pi it}f(\chi +2\pi ,t)=f(\chi ,t),f(\chi ,t+1)=f(\chi ,t),$$
while the covariant derivatives along $`\frac{}{\chi }`$ and $`\frac{}{t}`$ are given by
$$_\chi =\frac{}{\chi },_t=\frac{}{t}+i\chi i\widehat{A}_t.$$
### 8.2 Coincidence Of The Spectral Curves: An Explicit Argument
Let us pick a point $`(\zeta ,\mathrm{exp}2\pi \sigma )\times ^{}`$ belonging to the monopole spectral curve $`𝐒`$. In other words, $`\mathrm{exp}2\pi \sigma `$ is one of the eigenvalues of the holonomy $`V(\zeta ,2\pi )`$:
$$det(V(\zeta ,2\pi )e^{2\pi \sigma })=0.$$
(44)
This implies that there is a family of sections $`\mathrm{\Psi }`$ of $`\widehat{\mathrm{\Lambda }}^0\widehat{\mathrm{\Lambda }}^1`$ parameterized by $`\chi `$, such that $`[\mathrm{\Psi }(\chi )]`$ is a nonzero element of $`H^1(\widehat{K}_{z,\chi })`$ and
$$\left[\left(\frac{}{\chi }r+\sigma \right)\mathrm{\Psi }\right]=[0].$$
(45)
Here the brackets designate the cohomology class in $`H^1(\widehat{K}_{z,\chi })`$. We also denote $`s=r+it`$, as usual.
Let us unwrap the equation (45). We will write $`\mathrm{\Psi }`$ as follows:
$$\mathrm{\Psi }(s,\chi )=\left(\begin{array}{c}a(\chi ,s)\\ b(\chi ,s)\end{array}\right),$$
For a fixed $`\chi `$ the functions $`a`$ and $`b`$ are sections of $`\widehat{E}`$. We can choose the cohomology representative $`\mathrm{\Psi }(s,\chi )`$ so that $`a`$ and $`b`$ are sections of $`𝒫^{}\widehat{\pi }^{}(\widehat{E})`$, and therefore satisfy
$$e^{2\pi it}\left(\begin{array}{c}a(\chi +2\pi ,s)\\ b(\chi +2\pi ,s)\end{array}\right)=\left(\begin{array}{c}a(\chi ,s)\\ b(\chi ,s)\end{array}\right).$$
The equation (45) means that there exists a section $`h`$ of $`𝒫^{}\widehat{\pi }^{}(\widehat{E})`$ such that
$$\left(\frac{}{\chi }r+\sigma \right)\left(\begin{array}{c}a\\ b\end{array}\right)=\left(\begin{array}{c}(\widehat{\varphi }\zeta )h\\ 2\overline{}_{\widehat{A}+\widehat{a}}h\end{array}\right).$$
Introducing
$$F(s)=_0^{2\pi }e^{i\chi t}h(\chi ,s)𝑑\chi ,$$
we find that
$$\left(\widehat{\varphi }(\sigma )\zeta \right)F(\sigma )=0.$$
If $`F(\sigma )`$ is nonzero, this implies that $`\zeta `$ is an eigenvalue of $`\widehat{\varphi }(\sigma )`$ and therefore the point $`(\zeta ,e^{2\pi \sigma })`$ belongs to the Hitchin spectral curve $`𝐂`$. This shows that the curves coincide.
To prove that $`F(\sigma )`$ is nonzero, let us assume the contrary. Then it follows from the Fredholm Alternative that the equation
$$\left(\frac{}{\chi }r+\sigma \right)g=h$$
has a solution $`g\mathrm{\Gamma }(𝒫^{}\widehat{\pi }^{}(\widehat{E}))`$. One can easily see that $`a`$ and $`b`$ are expressible in terms of $`g`$ as follows:
$$\left(\begin{array}{c}a\\ b\end{array}\right)=\left(\begin{array}{c}(\widehat{\varphi }\zeta )g\\ 2\overline{}_{\widehat{A}+\widehat{a}}g\end{array}\right).$$
But this contradicts the assumption that $`\mathrm{\Psi }`$ represents a nontrivial cohomology class in $`H^1(\widehat{K}_{z,\chi }).`$ Thus if $`\mathrm{exp}(2\pi \sigma )`$ is an eigenvalue of $`V(\zeta ,2\pi )`$, then $`\zeta `$ is an eigenvalue of $`\widehat{\varphi }(\sigma )`$.
### 8.3 Cohomological Argument
The above argument can be conveniently rephrased in cohomological terms. Consider the manifold $`\mathrm{SS}^1\times \widehat{X}=\mathrm{SS}^1\times \widehat{\mathrm{SS}}^1\times `$ and a bundle $`=𝒫^{}\widehat{\pi }^{}(\widehat{E})`$ over it. We have two operators acting on its sections, namely $`\widehat{\delta }_{\zeta ,\chi }`$ of Eq. (42) and $`\mathrm{}`$ given by
$$\mathrm{}=\frac{}{\chi }r+\sigma .$$
(46)
Let us note that these operators, $`\widehat{\delta }_{\zeta ,\chi }`$ and $`\mathrm{}`$, commute.
$$[\widehat{\delta }_{\zeta ,\chi },\mathrm{}]=0.$$
(47)
Consider a complex of sheaves of vector spaces:
$$0\stackrel{}{}\stackrel{int}{}\widehat{\mathrm{E}}_{\mathrm{s}=\sigma }0,$$
(48)
where $`\mathrm{int}`$ acts as
$$\mathrm{int}:f(\chi ,s)_0^{2\pi }f(\chi ,\sigma )e^{i\chi \mathrm{Im}\sigma }𝑑\chi $$
(49)
Even though this short sequence is not exact, by an argument similar to that in subsection 5.4, one can show that there still is a long exact sequence of cohomology groups. To show this, we consider a double complex whose lowest row is the complex (48), and the vertical differential is given by $`\widehat{\delta }_{\zeta ,\chi }`$. The first level $`\stackrel{~}{E}_1^{p,q}`$ of the spectral sequence of this double complex contains the cohomology groups $`H_{\widehat{\delta }_{\zeta ,\chi }}^j(\widehat{X},\widehat{E})`$, as well as maps between them. Comparison with the total cohomology of this double complex provides us with an exact sequence
$$0H_{\widehat{\delta }_{\zeta ,\chi }}^0(\widehat{E}|_{s=\sigma })H_{\widehat{\delta }_{\zeta ,\chi }}^1(\widehat{X},\widehat{E})\stackrel{}{}H_{\widehat{\delta }_{\zeta ,\chi }}^1(\widehat{X},\widehat{E})\mathrm{}$$
(50)
From the definition of $`\widehat{\delta }_{\zeta ,\chi }`$ we have $`H_{\widehat{\delta }_{\zeta ,\chi }}^0(\widehat{E}|_{s=\sigma })=\mathrm{Ker}\left(\widehat{\varphi }(\sigma )\zeta \right)`$. Recalling the identification of $`\mathrm{Ker}\widehat{D}_{\zeta ,\chi }^{}`$ with $`H_{\widehat{\delta }_{\zeta ,\chi }}^1(\widehat{X},\widehat{E})`$, the above exact sequence implies an isomorphism
$$\mathrm{Ker}\left(\widehat{\varphi }(\sigma )\zeta \right)\mathrm{Ker}\mathrm{}|_{\mathrm{Ker}\widehat{D}_{\zeta ,\chi }^{}}.$$
(51)
If the point $`(\zeta ,\mathrm{exp}(2\pi \sigma )`$ belongs to the monopole spectral curve $`𝐒`$, then the right-hand side of this equation is nonempty, and therefore the point belongs to the Hitchin spectral curve as well. The converse statement is also true. Thus the two curves coincide. Moreover, the spectral line bundle $`\mathrm{Ker}\left(\widehat{\varphi }(\sigma )\zeta \right)`$ on $`𝐂`$ is identified with $`\mathrm{Ker}\left(\mathrm{}|_{H_{\widehat{\delta }_{\zeta ,\chi }}^1(\widehat{X},\widehat{E})}\right)`$, which is the line bundle on the monopole spectral curve $`𝐒`$. Thus the line bundles on $`𝐂`$ and $`𝐒`$ are also isomorphic.
The isomorphism (51) also enables one to compute the index of $`\widehat{D}`$. Indeed, since for $`\sigma _1\sigma _2`$ the kernels of $`\mathrm{}_{\sigma _1}`$ and $`\mathrm{}_{\sigma _2}`$ do not intersect, one can easily see that
$$\mathrm{Ker}\widehat{D}_{\zeta ,\chi }^{}_{\sigma \widehat{X}}\mathrm{Ker}\left(\widehat{\varphi }(\sigma )\zeta \right).$$
The dimension of the right-hand side is just the number of points at which the spectral curve $`𝐂`$ intersects the cylinder in $`\times ^{}`$ given by $`z=\zeta `$. From the equation of the curve (34) we see that there are two such points for any $`\zeta `$. Since the kernel of $`\widehat{D}`$ is trivial, the index of $`\widehat{D}`$ equals $`2`$, as promised.
### 8.4 The Asymptotic Behavior Of The Monopole
We are now ready to show that the inverse Nahm transform produces a solution of Bogomolny equations with the asymptotics (1.4). By assumption, we started from a solution of Hitchin equations with the spectral curve
$$f(z)\mathrm{exp}(2\pi s)\mathrm{exp}(2\pi s)=0,$$
where $`f(z)`$ is a degree $`k`$ polynomial with leading coefficient $`\mathrm{exp}(2\pi 𝔳)`$. We proved that the monopole spectral curve is given by the same equation. This implies that $`\mathrm{Tr}V(z,2\pi )`$ grows as
$$\mathrm{Tr}V(z,2\pi )z^ke^{2\pi 𝔳}$$
(52)
for large $`z`$, while $`detV(z,2\pi )=1`$ everywhere.
Another piece of information that we need is that $`F_A^2`$ is bounded by a multiple of $`1/|z|^2`$. This follows from (39) and a simple estimate of the norm of $`(\widehat{D}^{}\widehat{D})^1`$. Together with the Bogomolny equation this fact implies that $`_\chi \varphi `$ goes to zero for large $`z`$, while the components of the connection $`A`$ can be chosen to be bounded.
Let us now investigate the consequence of these two observations. First, one can show that the inverse Nahm transform yields a traceless connection and a traceless Higgs field (this is not obvious from their definition). Indeed, if the trace part of the curvature were nonzero, it would satisfy the Laplace equation (as a consequence of the Bogomolny equation) and grow at infinity, in contradiction with the above estimate. Hence the curvature is traceless and $`\mathrm{Tr}A`$ is a flat connection on $`^2\times \mathrm{SS}^1`$. Furthermore, $`\mathrm{Tr}\varphi `$ must be constant by virtue of the Bogomolny equation. Now, since the monopole spectral curve tells us that $`detV(z,2\pi )=1`$ everywhere, this means that $`\mathrm{Tr}\varphi =0`$ and $`\mathrm{Tr}A`$ has zero monodromy. Since $`\mathrm{Tr}A`$ is also flat, it must be gauge-equivalent to zero.
Second, the fact that $`\mathrm{Tr}V(z,2\pi )`$ grows as (52) at infinity implies that for large $`z`$ the eigenvalues of the Higgs field are
$$\pm \frac{k}{2\pi }\mathrm{log}|z|+v+o(1).$$
This proves that the Higgs field has the asymptotics (1.4). To show that the gauge field has the correct asymptotics, it suffices to prove that the components of the connection orthogonal to $`\varphi `$ go to zero for large $`|z|`$, i.e. that the $`SU(2)`$ monopole approaches at infinity a $`U(1)`$ monopole embedded in $`SU(2)`$. The argument for this is exactly the same as for the monopole on $`^3`$ (it uses only the fact that at large distances $`\varphi `$ is bounded from below by a strictly positive constant). In fact, proves that the “nonabelian” components of the curvature decay exponentially fast. From the physical point of view this can be explained as follows. Since $`\varphi 1`$ for large enough $`|z|`$, the $`SU(2)`$ gauge group is broken down to $`U(1)`$, the Higgs effect makes all the “nonabelian” components of the gauge field massive, and they decay exponentially.
## 9 Closing The Circle
In this section we prove that the composition of the direct and inverse Nahm transform takes a periodic monopole to a gauge-equivalent periodic monopole. Together with the results of Sections 38, this implies that there is a one-to-one correspondence between the gauge-equivalence classes of periodic $`SU(2)`$ monopoles with charge $`k`$ and gauge-equivalence classes of solutions of $`U(k)`$ Hitchin equations on a cylinder with the asymptotic behavior as described in Section 6. Our proof is modelled on the argument given by Schenk for instantons on a four-torus. Another proof, similar to that given by Donaldson and Kronheimer for instantons on $`T^4`$, is sketched in the Appendix.
The direct Nahm transform is defined in terms of square-integrable sections $`\psi _1(x,s),\mathrm{},\psi _k(x,s)`$ of $`ES`$ which form an orthonormal basis of $`\mathrm{Ker}D^{}`$. Here $`S`$ is the spin bundle on $`X`$, and $`D^{}`$ is twisted by $`s`$. The sections $`\psi _1(x,s),\mathrm{},\psi _k(x,s)`$ span a fiber of $`\widehat{E}`$ at a point $`s`$, so by combining them into a matrix $`\mathrm{\Psi }=(\psi _1,\psi _2,\mathrm{},\psi _k)`$ we obtain a section $`\mathrm{\Psi }(x,s)`$ of $`\widehat{\pi }^{}(\widehat{E})\pi ^{}(ES)`$. Here $`\pi :X\times \widehat{X}X`$ and $`\widehat{\pi }:X\times \widehat{X}\widehat{X}`$ are the natural projections. By definition, $`\mathrm{\Psi }`$ satisfies
$$D^{}\mathrm{\Psi }=0.$$
(53)
Since $`S`$ is trivial and two-dimensional, we can view $`\mathrm{\Psi }`$ as a pair of bundle morphisms $`(\mathrm{\Psi }_1,\mathrm{\Psi }_2)`$ from $`\widehat{\pi }^{}(\widehat{E})`$ to $`\pi ^{}(E)`$. (We remind that we have Hermitean inner products on $`E`$ and $`\widehat{E}`$ and thus can identify $`E`$ and $`\widehat{E}`$ with their duals.)
In terms of $`\mathrm{\Psi }`$ the expression for $`(\widehat{A},\widehat{\varphi })`$ reads
$$\frac{}{s}i\widehat{A}_s=_Xd^3x\mathrm{Tr}_{\mathrm{spin}}\mathrm{\Psi }^{}(x,s)\frac{}{s}\mathrm{\Psi }(x,s),\widehat{\varphi }=_Xd^3x\mathrm{Tr}_{\mathrm{spin}}\mathrm{\Psi }^{}(x,s)z\mathrm{\Psi }(x,s).$$
(54)
We denote by $`\frac{}{s}\mathrm{\Psi }`$ the composition of $`\mathrm{\Psi }`$ and $`\frac{}{s}`$, while the derivative of $`\mathrm{\Psi }`$ with respect to $`s`$ will be denoted by $`[\frac{}{s},\mathrm{\Psi }]`$.
In order to perform the inverse Nahm transform, we have to find a pair of sections $`\widehat{\psi }_1(s,x),\widehat{\psi }_2(s,x)`$ of $`\widehat{\pi }^{}(\widehat{E})^2`$ which span $`\mathrm{Ker}\widehat{D}_{z,\chi }^{}`$ for all $`(z,\chi )X`$. In other words, if we combine them into a $`k\times 2\times 2`$ matrix $`\widehat{\mathrm{\Psi }}=(\widehat{\psi }_1,\widehat{\psi }_2)`$, $`\widehat{\mathrm{\Psi }}`$ must satisfy
$$\widehat{D}^{}\widehat{\mathrm{\Psi }}=0,$$
(55)
and, with proper normalization,
$$_{\widehat{X}}d^2s\mathrm{Tr}_{\mathrm{spin}}\widehat{\mathrm{\Psi }}(s,x)\widehat{\mathrm{\Psi }}^{}(s,x)=1_E.$$
(56)
Here $`1_E`$ is the identity endomorphism $`EE`$.
Given $`\widehat{\mathrm{\Psi }}`$, the inverse Nahm transform $`(\stackrel{ˇ}{\widehat{A}},\stackrel{ˇ}{\widehat{\varphi }})`$ is given by
$$d_{\stackrel{ˇ}{\widehat{A}}}=_{\widehat{X}}d^2s\widehat{\mathrm{\Psi }}^{}(s,x)d_x\widehat{\mathrm{\Psi }}(s,x),\stackrel{ˇ}{\widehat{\varphi }}(s)=_{\widehat{X}}d^2s\widehat{\mathrm{\Psi }}^{}(s,x)r\widehat{\mathrm{\Psi }}(s,x).$$
(57)
The difficulty in establishing the equivalence of $`(A,\varphi )`$ and $`(\stackrel{ˇ}{\widehat{A}},\stackrel{ˇ}{\widehat{\varphi }})`$ lies in finding $`\widehat{\mathrm{\Psi }}(s,x)`$ in terms of $`\mathrm{\Psi }(x,s)`$. In the case of the Nahm transform on a four-torus this was accomplished by Schenk , whose results we adapt to the case at hand. Let $`\widehat{\mathrm{\Psi }}^T`$ denote $`\widehat{\mathrm{\Psi }}`$ with the spinor indices transposed. We claim that
$$\widehat{\mathrm{\Psi }}^T(s,x)=2\sqrt{2\pi }_Xd^3y\mathrm{\Psi }^{}(y,s)(D^{}D)^1(y,x;s)e^{i\chi _xt},$$
(58)
where $`t=\mathrm{Im}s`$, as usual. In what follows it will be convenient to regard $`x,yX`$ as continuous labels and think of $`\mathrm{\Psi }`$ as an object with one continuous and three discrete labels. Integration over $`x`$ is then regarded as a summation over a continuous label and is not shown explicitly. The dependence on $`s`$ will not be shown explicitly either. In this shortened notation Eq. (58) takes the form
$$\widehat{\mathrm{\Psi }}^T=2\sqrt{2\pi }\mathrm{\Psi }^{}(D^{}D)^1e^{i\chi t}.$$
The first thing to check is whether $`\widehat{\mathrm{\Psi }}`$ is a well-defined section of $`\widehat{E}`$, that is, whether $`\widehat{\mathrm{\Psi }}(s+i,x)=\widehat{\mathrm{\Psi }}(s,x)`$. Since the twisted derivative along $`\chi `$ is given by $`\left(_\chi iA_\chi +it\right)`$, we see that $`\mathrm{\Psi }(x,s+i)`$ is related to $`e^{i\chi }\mathrm{\Psi }(x,s)`$ by a $`U(k)`$ gauge transformation. By making an $`s`$-dependent change of basis in $`\mathrm{Ker}D^{}`$, we can always ensure that $`\mathrm{\Psi }(x,s+i)=e^{i\chi }\mathrm{\Psi }(x,s)`$. Furthermore, we have
$$(D^{}D)^1(x,y;s+i)=e^{i\chi _x}(D^{}D)^1(x,y;s)e^{i\chi _y}.$$
It follows that $`\widehat{\mathrm{\Psi }}(s+i,x)=\widehat{\mathrm{\Psi }}(s,x)`$.
In terms of $`\sigma _\pm =\sigma _1\pm i\sigma _2`$ and $`p_\pm =(1\pm \sigma _3)/2`$ the untwisted operators $`\widehat{D}`$ and $`\widehat{D}^{}`$ take the following form
$$\widehat{D}=\left(\begin{array}{cccc}_{\widehat{A}},& \overline{}_{\widehat{A}},& \widehat{\varphi },& \widehat{\varphi }^{}\end{array}\right)\left(\begin{array}{c}\sigma _+\\ \sigma _{}\\ p_+\\ p_{}\end{array}\right)$$
and
$$\widehat{D}^{}=(1)\left(\begin{array}{cccc}_{\widehat{A}},& \overline{}_{\widehat{A}},& \widehat{\varphi },& \widehat{\varphi }^{}\end{array}\right)\left(\begin{array}{c}\sigma _+\\ \sigma _{}\\ p_{}\\ p_+\end{array}\right).$$
(59)
The statement that $`\widehat{\mathrm{\Psi }}(s,x)`$ given by (58) satisfies (55) is equivalent to the following identity
$$\left(\mathrm{\Psi }^{}([\frac{}{s},\mathrm{\Psi }],[\frac{}{\overline{s}},\mathrm{\Psi }],z\mathrm{\Psi },\overline{z}\mathrm{\Psi })\mathrm{\Psi }^{}(D^{}D)^1\mathrm{\Psi }^{}(D^{}D)^1(0,0,z,\overline{z})\right)\left(\begin{array}{c}\sigma _+^T\\ \sigma _{}^T\\ p_{}^T\\ p_+^T\end{array}\right)=0.$$
(60)
We remind that $`z`$ stands for an operator of multiplication by $`x_1+ix_2`$, or equivalently for an integral operator with a kernel $`(x_1+ix_2)\delta (x,y)`$.
Making use of the identities
$$\mathrm{\Psi }\mathrm{\Psi }^{}=1D(D^{}D)^1D^{},$$
(61)
$`[{\displaystyle \frac{}{s}},(D^{}D)^1]=(D^{}D)^1(p_{}DD^{}p_+)(D^{}D)^1,`$
$`[{\displaystyle \frac{}{\overline{s}}},(D^{}D)^1]=(D^{}D)^1(p_+DD^{}p_{})(D^{}D)^1,`$ (62)
$`[z,(D^{}D)^1]=(D^{}D)^1(\sigma _+D+D^{}\sigma _+)(D^{}D)^1,`$
$`[\overline{z},(D^{}D)^1]=(D^{}D)^1(\sigma _{}D+D^{}\sigma _{})(D^{}D)^1,`$
and $`\sigma _\pm ^T=\sigma _{}`$ and $`p_\pm ^T=p_\pm `$, one can see that Eq. (60) is a consequence of a matrix identity
$$\left((p_{},p_+,\sigma _+,\sigma _{})D+(p_+,p_{},\sigma _+,\sigma _{})D^{}+D^{}(p_+,p_{},\sigma _+,\sigma _{})\right)\left(\begin{array}{c}\sigma _{}\\ \sigma _+\\ p_{}\\ p_+\end{array}\right)=0,$$
which can be readily verified. Thus we conclude that $`\widehat{\mathrm{\Psi }}`$ defined by Eq. (58) indeed solves $`\widehat{D}^{}\widehat{\mathrm{\Psi }}=0`$.
Next we verify Eq. (56). With the help of Eqs. (62) we find
$`\mathrm{Tr}_{\mathrm{spin}}(D^{}D)^1\left(1D(D^{}D)^1D^{}\right)(D^{}D)^1=`$ (63)
$`={\displaystyle \frac{1}{8}}\mathrm{Tr}_{\mathrm{spin}}\left(4_{\overline{s}}_s(D^{}D)^1[\overline{z},[z,(D^{}D)^1]]\right).`$
Combining this identity with the the formula for $`\widehat{\mathrm{\Psi }}`$, we obtain
$`{\displaystyle _{\widehat{X}}}d^2s\mathrm{Tr}_{\mathrm{spin}}\widehat{\mathrm{\Psi }}^{}(x_1,s)\widehat{\mathrm{\Psi }}(x_2,s)=`$
$`=\pi \mathrm{Tr}_{\mathrm{spin}}{\displaystyle _{\widehat{X}}}d^2se^{it(\chi _1\chi _2)}\left(4_{\overline{s}}_s(\overline{z_1}\overline{z_2})(z_1z_2)\right)(D^{}D)^1(x_1,x_2;s).`$ (64)
Integrating by parts and considering the limit of $`x_2`$ approaching $`x_1`$, we see that in this limit the integral is dominated by the region of large $`r=\mathrm{Re}s`$. Thus to estimate the integral it is sufficient to consider the large $`r`$ limit, where $`(D^{}D)^1`$ reduces to the Green’s function of the operator $`^2+r^2`$. We conclude that in the limit $`x_1x_2`$ the right-hand side of Eq. (9) reduces to
$$2\pi 1_E\underset{x_1x_2}{lim}_{\mathrm{}}^+\mathrm{}𝑑r\left(|x_1x_2|^2\right)\frac{e^{|r||x_1x_2|}}{4\pi |x_1x_2|}.$$
Performing the integral over $`r`$, we get (56).
Now, having constructed $`\widehat{\mathrm{\Psi }}`$, one can find the result of the inverse Nahm transform. Let us start with a few useful identities valid for any smooth section $`\mathrm{\Xi }(s)`$ of $`\widehat{E}`$ which decays rapidly as $`|s|\mathrm{}`$ together with all its derivatives (i.e. belongs to the Schwarz space):
$`{\displaystyle _{\widehat{X}}}d^2s\mathrm{\Psi }(x,s)_{\widehat{A}+\widehat{a}}\mathrm{\Xi }(s)`$ $`=`$ $`{\displaystyle _{\widehat{X}}}d^2sD(D^{}D)^1p_{}\mathrm{\Psi }\mathrm{\Xi }(s),`$
$`{\displaystyle _{\widehat{X}}}d^2s\mathrm{\Psi }(x,s)\overline{}_{\widehat{A}+\widehat{a}}\mathrm{\Xi }(s)`$ $`=`$ $`{\displaystyle _{\widehat{X}}}d^2sD(D^{}D)^1p_+\mathrm{\Psi }\mathrm{\Xi }(s),`$
$`{\displaystyle _{\widehat{X}}}d^2s\left(\mathrm{\Psi }(x,s)\widehat{\varphi }z\mathrm{\Psi }(x,s)\right)\mathrm{\Xi }(s)`$ $`=`$ $`{\displaystyle _{\widehat{X}}}d^2sD(D^{}D)^1\sigma _+\mathrm{\Psi }\mathrm{\Xi }(s),`$
$`{\displaystyle _{\widehat{X}}}d^2s\left(\mathrm{\Psi }(x,s)\widehat{\varphi }^{}\overline{z}\mathrm{\Psi }(x,s)\right)\mathrm{\Xi }(s)`$ $`=`$ $`{\displaystyle _{\widehat{X}}}d^2sD(D^{}D)^1\sigma _{}\mathrm{\Psi }\mathrm{\Xi }(s).`$
Now substitute into these formulas $`\mathrm{\Xi }(s)=e^{it\chi ^{}}\widehat{\mathrm{\Psi }}^T(s,x^{})`$, set $`x^{}=x`$, multiply them from the right by $`\sigma _{},\sigma _+,p_{},`$ and $`p_+`$, respectively, and sum them up. The left-hand side of the resulting identity will be proportional to $`\left(D^{}\widehat{\mathrm{\Psi }}\right)^T`$ and therefore will vanish. Thus we get
$$_{\widehat{X}}d^2sD(D^{}D)^1(p_{},p_+,\sigma _+,\sigma _{})\mathrm{\Psi }e^{i\chi t}\widehat{\mathrm{\Psi }}^T\left(\begin{array}{c}\sigma _{}\\ \sigma _+\\ p_{}\\ p_+\end{array}\right)=0.$$
(65)
Substituting $`2\sqrt{2\pi }e^{i\chi t}(D^{}D)^1\mathrm{\Psi }=\left(\widehat{\mathrm{\Psi }}^T\right)^{}`$ and using an identity
$$(p_{},p_+,\sigma _+,\sigma _{})M\left(\begin{array}{c}\sigma _{}\\ \sigma _+\\ p_{}\\ p_+\end{array}\right)=(\sigma _++\sigma _{})\mathrm{Tr}_{\mathrm{spin}}M$$
(66)
valid for any $`2\times 2`$ matrix $`M`$, we are left with
$$_{\widehat{X}}d^2se^{i\chi t}De^{i\chi t}(\sigma _++\sigma _{})\mathrm{Tr}_{\mathrm{spin}}\left(\widehat{\mathrm{\Psi }}^T\right)^{}\widehat{\mathrm{\Psi }}^T=0.$$
(67)
This operator equation in spin space can be rewritten as four “scalar” equations which express the coincidence of $`(A,\varphi )`$ and $`(\stackrel{ˇ}{\widehat{A}},\stackrel{ˇ}{\widehat{\varphi }})`$. For example, the vanishing of the coefficient of $`p_+`$ in Eq. (67) implies
$$_{\widehat{X}}d^2s\left(\left(\frac{}{z}iA_z\right)\widehat{\mathrm{\Psi }}_1^{}(s,x)\right)\widehat{\mathrm{\Psi }}_1(s,x)+\left(\left(\frac{}{z}iA_z\right)\widehat{\mathrm{\Psi }}_2^{}(s,x)\right)\widehat{\mathrm{\Psi }}_2(s,x)=0,$$
(68)
where $`\widehat{\mathrm{\Psi }}_1`$ and $`\widehat{\mathrm{\Psi }}_2`$ are the two spinor components of $`\widehat{\mathrm{\Psi }}`$. This equation simply says that $`A_z=\stackrel{ˇ}{\widehat{A}}_z`$. Writing out the coefficient of $`\sigma _+`$, we obtain
$$\begin{array}{c}_{\widehat{X}}d^2s\left(\left(\frac{}{\chi }iA_\chi \varphi +r\right)\widehat{\mathrm{\Psi }}_1^{}(s,x)\right)\widehat{\mathrm{\Psi }}_1(s,x)\hfill \\ \hfill +\left(\left(\frac{}{\chi }iA_\chi \varphi +r\right)\widehat{\mathrm{\Psi }}_2^{}(s,x)\right)\widehat{\mathrm{\Psi }}_2(s,x)=0,\end{array}$$
(69)
which implies $`A_\chi =\stackrel{ˇ}{\widehat{A}}_\chi `$ and $`\varphi =\stackrel{ˇ}{\widehat{\varphi }}`$. This completes the proof.
## 10 Remarks On The Existence Of Periodic Monopoles
It is intuitively plausible that periodic monopoles exist for all $`k>0`$.<sup>1</sup><sup>1</sup>1From the string theory point of view, periodic $`SU(2)`$ monopoles can be identified with D4 branes suspended between two parallel NS5 branes, with one direction common to D4-branes and NS5-branes compactified on a circle (see section 2 for details). This brane configuration surely exists, so one is tempted to dismiss the question of the existence of periodic monopoles as trivial. But it is far from obvious that suspended D4-branes are represented by nonsingular field configurations on the NS5-branes after a T-duality along the compact direction. In fact, when the parameter $`v`$ in (5) is large, one can propose a simple way of constructing approximate solutions of Bogomolny equations on $`^2\times S^1`$. One considers a charge $`k`$ $`SU(2)`$ monopole on $`^3`$ located near the origin of $`^3`$. It approaches a charge $`k`$ Dirac monopole exponentially fast at distances larger than $`1/v`$ . Then one can patch it with a periodic Dirac monopole solution of Section 1 at distances larger than $`1/v`$ but smaller than $`1`$, and obtain an accurate approximation to a nonabelian periodic monopole.
To prove the existence of periodic monopoles for all $`v`$ and $`k`$ it is easier to use the correspondence between periodic monopoles and solutions of Hitchin equations on a cylinder established above. For $`k=1`$ one can write down explicitly a family of solutions of Hitchin equations with required asymptotics:
$$\widehat{\varphi }(s)=e^{2\pi (𝔳+s)}+e^{2\pi (𝔳s)}+c,c,A=\beta dt,\beta /(2\pi ).$$
(70)
This proves that a periodic monopole of charge $`1`$ exists. Moreover, it is easy to see that any solution of $`U(1)`$ Hitchin equations with the boundary conditions described in section 6 is gauge-equivalent to (70). Thus a periodic monopole with $`k=1`$ has three real moduli ($`\mathrm{Re}c,\mathrm{Im}c,\beta `$). (Caution: we do not claim that there is a natural metric on this moduli space, and in fact we will see below that this is not true.) They arise from the translational invariance of the Bogomolny equations and parameterize $`^2\times \mathrm{SS}^1=X`$. We may regard them as describing the location of the monopole on $`X`$.
For $`k>1`$ finding solutions of Hitchin equations is harder, and we do not have a satisfactory proof of their existence. Below we merely sketch a possible approach to the proof based on the holomorphic description of solutions of Hitchin equations. The idea of the holomorphic approach is familiar to physicists in the guise of the following principle: the space of solutions of D and F-flatness conditions modulo a compact gauge group is the same as the space of solutions of the F-flatness conditions modulo the complexified gauge group (this principle is often referred to as the Luty-Taylor theorem ).
Let us apply this principle to our problem. The “complex” Hitchin equation is invariant with respect to the complexified gauge transformations, i.e. gauge transformations which are $`GL(k,)`$-valued. The “real” Hitchin equation is invariant only with respect to $`U(k)`$ gauge transformations. Thus from the physical point of view, the “real’ and “complex” Hitchin equations play the role of the D-flatness and F-flatness conditions, respectively, and it is natural to consider the space of solutions of the “complex” Hitchin equation modulo the complexified gauge group. The Hermitian inner product on the bundle $`\widehat{E}`$ then plays no role, and all we have is a holomorphic bundle over $`\widehat{X}^{}`$. The “complex” Hitchin equation says that $`\widehat{\varphi }`$ is a holomorphic section of $`\mathrm{End}(\widehat{E})`$. Such a pair, a holomorphic bundle $`\widehat{E}`$ on $`\widehat{X}`$ and a holomorphic section of $`\mathrm{End}(\widehat{E})\mathrm{\Omega }_{\widehat{X}}`$, is called a Higgs bundle. Obviously, we have a forgetful map from the moduli space of solutions of the full Hitchin equations to the moduli space of Higgs bundles.
It is very easy to construct Higgs bundles on $`\widehat{X}`$, and in the next section we will give a rather explicit description of their moduli space. Thus if the above-mentioned map is surjective, the existence of solutions of Hitchin equations will be established.
Let us explain why it is plausible that every suitable Higgs bundle comes from a solution of Hitchin equations. In the case of Hitchin equations on a compact Riemann surface, one can prove that any stable Higgs bundle is related by a $`GL(k,)`$ gauge transformation to a solution of Hitchin equations. The role of the stability condition is to ensure that the complexified gauge group acts freely on the Higgs bundles. One may also consider solutions of Hitchin equations on a punctured Riemann surface with “tame” singularities at the punctures . (“Tame” means that the eigenvalues of the Higgs field grow at most as $`1/r`$ as one approaches the puncture.) The corresponding holomorphic object is a Higgs bundle on the punctured Riemann surface whose Higgs field has simple poles at the punctures. Again there is a stability condition on the Higgs bundle which ensures that the complexified gauge group acts freely on its orbit, and any stable Higgs bundle on a punctured Riemann surface comes from a solution of Hitchin equations with tame singularities .
In our case the Higgs field is not “tame” at infinity. To see this, let us make a conformal transformation $`w=\mathrm{exp}(2\pi s)`$ which maps the cylinder to $`^{}`$. Keeping in mind that the Higgs field is a section of $`\mathrm{End}(\widehat{E})\mathrm{\Omega }^{1,0}`$, we see that its eigenvalues near $`w=0`$ behave as $`|w|^{11/k}`$, i.e. the singularity is not “tame.” A similar problem occurs at $`w=\mathrm{}`$. This means that we cannot use the results of . Still, the above discussion suggests that the important thing is for complexified gauge transformations to act freely on the Higgs bundles. It appears that this condition is always satisfied if the spectral curve of the Higgs bundle is given by
$$z^k+a_1z^{k1}+\mathrm{}+a_k2e^{2\pi 𝔳}\mathrm{cosh}(2\pi s)=0.$$
(71)
Indeed, if there were a $`GL(k,)`$ transformation which would leave our Higgs bundle $`(\widehat{E},\widehat{\varphi })`$ invariant, this would mean that $`\widehat{E}`$ has a rank one holomorphic subbundle invariant with respect to $`\widehat{\varphi }`$. However, this is clearly not true for large $`|\mathrm{Re}s|`$ because of the asymptotic behavior of the eigenvalues of $`\widehat{\varphi }`$: they are all distinct and cyclically permuted as one goes around the circumference of the cylinder. Thus it seems plausible that any Higgs bundle whose spectral curve has the form (71) is related by a $`GL(k,)`$ transformation to a solution of Hitchin equations with required asymptotics.
One could ask if there could be a one-to-one correspondence between solutions of Hitchin equations on a cylinder modulo $`U(k)`$ gauge transformations and holomorphic Higgs bundles with the spectral curve (71) modulo $`GL(k,)`$ gauge transformations. This would be the analogue of the Luty-Taylor theorem for Hitchin equations on a cylinder. If the question is posed this way, the answer is negative. Indeed, we already saw that rank one solutions of Hitchin equations are parameterized by a complex number $`c`$ which describes the Higgs field, and a real number $`\beta `$ which parameterizes the monodromy of $`\widehat{A}`$ around the circumference of the cylinder. On the other hand, the bundle $`\widehat{E}`$ is holomorphically trivial, so all the information about the holomorphic Higgs bundle is described by $`c`$. In other words, the information about the monodromy of $`\widehat{A}`$ is lost in the holomorphic picture.
In the case of Higgs bundles with “tame” singularities, the situation is similar: the information about the monodromy of $`\widehat{A}`$ around the punctures is lost upon passing to a Higgs bundle. But there is a way to fix this: one has to consider Higgs bundles with “parabolic structure” at the punctures . Parabolic structure essentially encodes the conjugacy class of the monodromy. In our case it is reasonable to conjecture that what is missing in the naive holomorphic description is precisely the information about the monodromy of $`\widehat{A}`$ at infinity. In fact, we have an a priori knowledge (see Section 6) that the monodromy is given by $`\mathrm{exp}(i\beta )V^{\pm 1}`$, where $`V`$ is the “shift” matrix (36), and $`\beta /(2\pi )`$. Thus the holomorphic description misses one real parameter $`\beta /(2\pi )`$.
We are led to the following conjecture. Let $`_{Hi,k}`$ be the moduli space of solutions of $`U(k)`$ Hitchin equations on a cylinder with asymptotics (35). Let $`_{HB,k}`$ be the moduli space of holomorphic Higgs bundles on $`^{}`$ whose spectral curve has the form (71). The forgetful map from $`_{Hi,k}`$ to $`_{HB,k}`$ is surjective, and moreover is a fiber bundle with fiber $`\mathrm{SS}^1`$.
In the next section we will test this conjecture by computing the dimension of $`_{HB,k}`$ and comparing with expectations from $`𝒩=2`$ super-Yang-Mills.
## 11 The Moduli Space Of Periodic Monopoles
In this section we describe the moduli space $`_{HB,k}`$ of solutions of the complex Hitchin equation on $`\times \mathrm{SS}^1`$ with the spectral curve (71). Assuming that the analogue of the Luty-Taylor theorem formulated in the previous section is true, the moduli space of charge $`k`$ periodic monopoles is fibered over $`_{HB,k}`$ with fiber $`\mathrm{SS}^1`$. As explained below, there is an alternative way to view the relation between $`_{HB,k}`$ and periodic monopoles: a certain submanifold in $`_{HB,k}`$ of complex codimension one coincides with the centered moduli space of charge $`k`$ periodic monopoles. We compare our results with the expectations from string theory and discuss the existence of a hyperkähler metric on the centered moduli space of periodic monopoles.
We already know how to associate a spectral curve $`𝐂^{}\times ^{}`$ and a coherent sheaf $`N`$ on it to every solution of the complex Hitchin equation. For a generic solution, the curve $`𝐂`$ is nonsingular, and $`N`$ is a line bundle (see Section (3)). The curve has the form
$$w^2wf(z)+1=0,$$
where $`f(z)`$ is a polynomial of degree $`k`$ whose leading coefficient is a known constant. Thus to specify the polynomial $`f(z)`$ we need to specify its $`k`$ coefficients $`a_1,\mathrm{},a_k`$.
For $`k=1`$ the spectral curve is rational, and its compactification is a $`^1`$. To understand what happens for $`k>1`$, it is convenient to rewrite the equation of the curve $`𝐂`$ in the form
$$\stackrel{~}{w}^2=\frac{1}{4}f(z)^21,$$
(72)
where $`\stackrel{~}{w}=wf(z)/2`$. This equation implies that the compactification of $`𝐂`$ is a hyperelliptic curve in $`^2`$. It is well known that a hyperelliptic curve in $`^2`$ has singularities, in this case over the point $`z=\mathrm{}`$. Its desingularization has genus $`k1`$ and will be denoted $`\stackrel{~}{𝐂}`$. The pull-back of the line bundle $`N`$ to $`\stackrel{~}{𝐂}`$ will be denoted by the same letter $`N`$. Since $`\widehat{E}`$ is a trivial bundle, $`N`$ has zero degree. The moduli space of line bundles over $`\stackrel{~}{𝐂}`$ with fixed degree is simply the Jacobian of $`\stackrel{~}{𝐂}`$, which is an Abelian variety of dimension $`k1`$.
Conversely, starting from a hyperelliptic curve $`\stackrel{~}{𝐂}`$ and a line bundle over it, we can reconstruct the solution of the complex Hitchin equation. The bundle $`\widehat{E}`$ over $`^{}`$ is obtained by pushing forward $`N`$ with respect to the projection $`(w,z)w`$. The Higgs field $`\widehat{\varphi }\mathrm{\Gamma }(\mathrm{End}(\widehat{E}))`$ is defined as follows:
$$\widehat{\varphi }:v_{w,z}zv_{w,z}.$$
Thus we obtain the following description of the moduli space of $`_{HB,k}`$ valid in an open set: it is the space of pairs $`(\stackrel{~}{𝐂},N)`$, where $`\stackrel{~}{𝐂}`$ is a hyperelliptic curve of genus $`k1`$ and $`N`$ is a degree $`0`$ line bundle on it. Hence $`_{HB,k}`$ is a complex manifold of dimension $`2k1`$ fibered over $`^k`$ by Abelian varieties of dimension $`k1`$.
It follows that the moduli space $`_{Hi,k}`$, and therefore the moduli space of periodic monopoles of charge $`k`$, has real dimension $`4k1`$. This coincides with the dimension of the moduli space of $`SU(2)`$ monopoles of charge $`k`$ on $`^3`$. Furthermore, string theory predicts that the centered moduli space of periodic monopoles has dimension $`4k4`$ (see section 2). Centering the monopole amounts to setting $`\beta =0`$, and $`a_1=0`$, where $`a_1`$ is the coefficient of $`z^{k1}`$ in $`f(z)`$. Indeed, we already explained in Section 4 that the positions of the constituent charge $`1`$ monopoles on $`^2`$ are the roots of the equation $`f(z)=2`$, so setting $`a_1=0`$ has the effect of making the center-of-mass of the monopole located at $`z=0`$. It is also easy to check that a translation along $`\mathrm{SS}^1`$ has the effect of shifting $`\beta `$. Thus the centered moduli space of periodic monopoles of charge $`k`$ is a hypersurface in $`_{HB,k}`$ given by the equation $`a_1=0`$. It has complex dimension $`2k2`$, in agreement with string theory predictions. This lends support to the conjectured correspondence between solutions of Hitchin equations on a cylinder and a special class of Higgs bundles.
Moreover, one expects on physical grounds that the moduli space of the $`𝒩=2`$ super Yang-Mills compactified on a circle has a distinguished complex structure in which it is a complex manifold fibered over $`^{k1}`$ by Abelian varieties of dimension $`k1`$ . We saw above that this is indeed true.
Let us now turn to the issue of the hyperkähler metric on the moduli space of periodic monopoles. Supersymmetry implies that the Coulomb branch of the $`𝒩=2`$ $`SU(k)`$ super-Yang-Mills theory compactified on a circle must be a complete hyperkähler manifold , so we expect that the hyperkähler metric exists for the centered moduli space. In contrast to monopoles on $`^3`$, we do not expect to have a well-defined metric on the uncentered moduli space. The reason for this is that the uncentered monopoles would correspond to a $`U(k)`$ gauge theory in $`d=4`$, but the latter does not make sense as a quantum theory because it is not asymptotically free.
We can also explain this in a purely classical way, which does not involve quantum $`𝒩=2`$ super-Yang-Mills theory. The difference between the centered and the uncentered moduli spaces is that in the former case we mod out by translations of $`^2\times \mathrm{SS}^1`$, while in the latter case we don’t. The reason why one needs to divide by the translations group to get a well-defined metric is that the tangent vectors to the moduli space corresponding to the translations on $`^2`$ are not normalizable, i.e. their $`L^2`$ norm diverges. This tangent vector is given by $`(\delta A,\delta \varphi )=(_zA,_z\varphi ).`$ According to (5), $`_z\varphi `$ decays only as $`1/z`$, therefore the $`L^2`$ norm of this tangent vector is logarithmically divergent.
The above arguments demonstrate that there is no well-defined metric on the uncentered moduli space, but they do not prove that there is one on the centered moduli space. This can be argued as follows. As explained in Section 6, for large $`|z|`$ a nonabelian periodic monopole is exponentially close to a periodic Dirac monopole embedded in $`SU(2)`$. Thus to count $`L^2`$ deformations it is sufficient to use the abelian asymptotics (5,6). Then it is easy to see that changing the locations of monopoles while keeping their center-of-mass fixed changes the Higgs field only by terms which decay as $`1/|z|^2`$ (this is essentially multipole expansion). Thus all such deformations have finite $`L^2`$ norm. There are $`3k3`$ such tangent vectors. Using the quaternionic structure of the tangent space (see below), one can show that the remaining $`k1`$ tangent vectors are also normalizable.
From the mathematical point of view, it may be easier to count $`L^2`$ deformations in the Nahm-transformed picture. Note that setting $`\beta =0`$, $`a_1=0`$ amounts to setting $`\mathrm{Tr}\widehat{\varphi }=0`$ and passing from the $`U(k)`$ to $`SU(k)`$ Hitchin equations. $`SU(k)`$ Hitchin equations may be regarded as hyperkähler moment map equations for the action of the $`SU(k)`$ gauge group on the cotangent bundle of the space of $`SU(k)`$ connections on $`\times \mathrm{SS}^1`$ . Formally, the hyperkähler quotient construction implies that the moduli space of $`SU(k)`$ Hitchin equations has a hyperkähler metric. In order to prove the existence of a hyperkähler metric on the centered moduli space, it is sufficient to show that the space of $`L^2`$ deformations of $`SU(k)`$ Hitchin equations on a cylinder has the expected dimension $`4k4`$.
The properties of the hyperkähler metric on the centered moduli space of periodic monopoles will be discussed elsewhere . One thing is clear though: in the limit when the circumference of $`\mathrm{SS}^1`$ goes to infinity, the centered moduli space of periodic monopoles smoothly goes over to the centered moduli space of monopoles on $`^3`$. In particular, the metric on the centered moduli space of a charge $`2`$ periodic monopole is a deformation of the Atiyah-Hitchin metric. It would be very interesting to find the explicit form of this metric. On physical grounds, we expect that it is hyperkähler and asymptotically locally flat. But unlike the Atiyah-Hitchin metric, which has an $`SU(2)`$ isometry, the new metric seems to have no continuous isometries.
## Acknowledgments
We are grateful to Nigel Hitchin for a very helpful conversation concerning the definition of the monopole spectral data, and to Dmitri Orlov and Marcos Jardim for discussions. We also wish to thank the organizers of the workshop “The Geometry and Physics of Monopoles,” Edinburgh, August-September 1999, for creating a very stimulating atmosphere during the meeting and for providing us with an opportunity to present a preliminary version of this work. The work of S.Ch. was supported in part by NSF grant PHY9819686. The work of A.K. was supported in part by a DOE grant DE-FG02-90ER4054442.
## Appendix
In section 9 we proved that the composition of the Nahm transform of Section 3 and the inverse Nahm transform of Section 7 is the identity map on the gauge-equivalence classes of periodic monopole configurations. For the sake of completeness, we present here an outline of a cohomological proof of this fact in the spirit of reference . Both the direct and inverse Nahm transforms, as well as a map identifying the result of the composition of the two transforms with the initial configuration, will emerge from the spectral sequence of a double complex.
Nahm transform, as described in subsection 5.1, is given in terms of the cohomology of the following complex:
$$0\mathrm{\Lambda }^{0,0}(X,E)\stackrel{\overline{𝒟}_0}{}\mathrm{\Lambda }^{0,1}(X,E)\stackrel{\overline{𝒟}_1}{}\mathrm{\Lambda }^{0,2}(X,E)0.$$
(73)
The differentials $`\overline{𝒟}_p,p=0,1,`$ here are twisted by $`\widehat{x}\widehat{X}`$. In the trivialization of the Poincare bundle defined in Section 8 the operators $`\overline{𝒟}_0`$ and $`\overline{𝒟}_1`$ are given by:
$$\overline{𝒟}_p=d\overline{z}2(\frac{}{\overline{z}}iA_{\overline{z}})+d\chi (\frac{}{\chi }iA_\chi \varphi +r).$$
(74)
Consider a trivial bundle $`𝔈X\times \widehat{X}`$, such that its restriction to $`X\times \widehat{x}`$ is $`E`$ and restriction to $`x\times \widehat{X}`$ is a trivial bundle with fiber $`E|_x`$. For each $`\widehat{x}\widehat{X}`$ it has an action of the operator $`\overline{𝒟}`$ twisted by $`\widehat{x}`$. Let $`\mathrm{\Omega }^p`$ denote the sheaf of $`𝔈`$-valued rapidly decaying $`p`$-forms spanned by the differentials $`d\chi `$ and $`d\overline{z}`$ with coefficients depending on $`x`$ and $`\widehat{x}`$. Here by “rapidly decaying” we mean rapidly decaying both for large $`|z|`$ and large $`\mathrm{Re}s`$. Consider a double complex
(75)
where $`\delta _0`$ and $`\delta _1`$ act as follows:
$$\delta _0:f\left(\begin{array}{c}zf\\ 2\left(\frac{}{\overline{s}}+\frac{1}{2}\chi \right)f\end{array}\right),\delta _1:\left(\begin{array}{c}g_0\\ g_1\end{array}\right)\left(2\left(\frac{}{\overline{s}}+\frac{1}{2}\chi \right)g_0zg_1\right).$$
(76)
It is easy to check that $`\overline{𝒟}`$ and $`\delta `$ commute. The zeroth and second cohomology of $`\overline{𝒟}`$ vanish, and the first cohomology yields $`\widehat{E}`$. Thus the cohomology of columns with respect to $`\overline{𝒟}`$ is:
(77)
which contains exactly the sequence (41) defining the inverse Nahm transform. Thus on the second level the sequence degenerates to
(78)
Computation of the other spectral sequence $`E^{p,q}`$ is exactly analogous to that of , the result being
(79)
Comparing the total cohomologies of the double complex (75)
$$_{p+q=n}E_{\mathrm{}}^{p,q}=_{p+q=n}\stackrel{~}{E}_{\mathrm{}}^{p,q},$$
(80)
we conclude that $`E|_{x=0}=\stackrel{ˇ}{\widehat{E}}|_{x=0}`$.
Chasing the spectral sequence we can obtain the isomorphism $`\omega :\stackrel{ˇ}{\widehat{E}}|_{x=0}E|_{x=0}`$ explicitly. Namely, an element of $`\stackrel{ˇ}{\widehat{E}}|_{x=0}`$ can be represented by $`\alpha \mathrm{\Omega }^1\mathrm{\Omega }^1`$ harmonic with respect to $`\overline{𝒟}`$ and such that $`\delta _1\alpha =\overline{𝒟}\beta `$ for some $`\beta \mathrm{\Omega }^0`$. The isomorphism $`\omega `$ takes $`\alpha `$ to the value of $`\beta `$ at $`\widehat{x}=0`$ integrated over $`X\times \{0\}`$. The equation for $`\beta `$ is solved by $`\beta =\left(\overline{𝒟}^{}\overline{𝒟}\right)^1\overline{𝒟}^{}\delta _1\alpha ,`$ thus
$$\omega :\alpha _{X\times \{0\}}𝑑x\left(\overline{𝒟}^{}\overline{𝒟}\right)^1[\overline{𝒟}^{},\delta _1]\alpha .$$
(81)
Note that $`\left(\overline{𝒟}^{}\overline{𝒟}\right)^1`$ is proportional to the Green’s function $`(D^{}D)^1`$, while the commutator $`[\overline{𝒟}^{},\delta _1]`$ has a simple form: it maps $`(g_0,g_1)𝒞^{1,1}=\mathrm{\Omega }^1\mathrm{\Omega }^1`$ to $`2(g_{0,\chi }+g_{1,\overline{z}})`$.
The point $`x=0`$ was not distinguished in any natural way, and twisting the vertical operator $`\delta `$ of the double complex (75) by $`(x_0)X`$ and computing the spectral sequence would lead to an isomorphism $`\omega :\stackrel{ˇ}{\widehat{E}}|_{x=x_0}E|_{x=x_0}`$. Therefore the isomorphism $`\omega `$ is an isomorphism of bundles on $`X`$. Using the above explicit formula for $`\omega `$, it can be checked that it commutes with the differential $`\overline{𝒟}`$. This shows that $`\stackrel{ˇ}{\widehat{A}}_{\overline{z}}=A_z`$ and $`(i\stackrel{ˇ}{\widehat{A}}_\chi +\stackrel{ˇ}{\widehat{\varphi }})=(iA_\chi +\varphi )`$.
In order to conclude that the isomorphism $`\omega `$ takes the original monopole data $`(A,\varphi )`$ to $`(\stackrel{ˇ}{\widehat{A}},\stackrel{ˇ}{\widehat{\varphi }})`$ we need to say a few words regarding the naturalness of the above construction. The way to present the direct (as well as inverse) Nahm transform in cohomological terms is not unique. For example, we could have replaced $`\overline{𝒟}`$ with a differential
$$(dx_1id\chi )\left(\frac{}{x_1}iA_1i\left(\frac{}{\chi }iA_\chi \right)\right)+dx_2\left(\frac{}{x_1}iA_2+(\varphi r)\right),$$
(82)
and modified the cohomological construction accordingly. This amounts to identifying $`X\times ^{}`$ instead of $`X\times \mathrm{SS}^1`$. This arbitrariness is exactly the same as the arbitrariness in the choice of complex structure in the twistor description of monopoles, as well as in the discussion of Higgs bundles and Nonabelian Cohomology in . As in the case of a four-torus , we could have constructed an isomorphism $`\eta :\stackrel{ˇ}{\widehat{E}}E`$ preserving an appropriate differential for each choice of the identification $`X\times ^{}`$. We would have discovered then that the isomorphism $`\eta `$ is always given by the formula (81) and therefore coincides with $`\omega `$. Therefore, $`\omega `$ maps $`(\stackrel{ˇ}{\widehat{A}},\stackrel{ˇ}{\widehat{\varphi }})`$ to $`(A,\varphi )`$.
|
warning/0006/quant-ph0006017.html
|
ar5iv
|
text
|
# Kolmogorov and von Mises viewpoints to the Greenburger-Horne-Zeilinger paradox
## 1 Introduction
It is well known that violations of Bell’s inequality by quantum correlations in the Einstein-Podolsky-Rosen (EPR) framework may be interpreted as the evidence of the impossibility to use the local realism in quantum theory (see, for example, , ). Such a viewpoint was strongly supported by experiments of Aspect which demonstrated violations of generalized Bell’s inequality (see also ). Despite of the general attitude to connect violations of Bell’s inequality with such problems as determinism and locality, there exists sufficiently strong opposition - to such a conclusion. This opposition, despite of the great diversity of approaches, can be called the probability opposition. The general viewpoint of adherents of the probabilistic interpretation of violations of Bell’s inequality is that the derivation of this inequality is based on numerous (hidden) probabilistic assumptions. Unfortunately at the present time there are no experimental facts which can justify these probabilistic assumptions. It seems that theoretical as well as experimental investigations of the EPR paradox (in particular, Bell’s inequality) must be at least partly reoriented to the investigation of probabilistic roots of this paradox.
The viewpoint that the notion of probability plays the large role in Bell’s (and in EPR’s) considerations is not so new, -. The main consequence of all these probabilistic analyses is that the EPR experiment could not be described (as it was assumed by J. Bell, ) by the unique Kolmogorov probability distribution. In fact, these are various forms of contextual interpretation of quantum formalism): (1) De Broglie, Lochak, Nelson, De Muynck, De Baere, Marten, Stekelenborg, , thermodynamical approach to Bell’s problem, difference between hidden and observed probabilities; (2) Beltrametti and Cassinelli , quantum logic; (3) Accardi , quantum probabilities, no Bayes’ formula; (4) Pitowsky and Gudder , probability manifolds; (5) De Baere , fluctuating probabilities; (6) Fine and Rastal , no simultaneous probability distribution; (7) Muckenheim , negative probabilities; (8) Khrennikov , $`p`$-adic probabilities; fluctuating probabilities and modified Bell’s inequality .
However, a new strong argument in the favour of nonlocal (or nonreal) interpretation of the EPR paradox was given by so called Greenberger-Horne-Zeilinger (GHZ) paradox, . The GHZ scheme is based on the probability one arguments. From the first point of view all probabilistic circumstances of the GHZ scheme are so straightforward that there is no more place for probabilistic counter arguments. However, the careful probabilistic analysis demonstrates that the GHZ paradox has even deeper connection to foundations of probability theory than Bell’s inequality. Roughly speaking the root of the GHZ paradox might be in the use of the conventional probability calculus, namely Kolmogorov’s (axiomatic) measure theoretical approach, 1933, .
In this paper we shall consider the GHZ paradox from the viewpoint of so called frequency probability theory, R. von Mises, 1919 (see for the advanced formalism). In the opposite to Kolmogorov’s model of probability theory which is characterized by the highest degree of abstraction, von Mises’ model of probability theory is characterized by its concreteness. By R. von Mises we cannot consider a probability distributions without the relation to the concrete collective (random sequence). Von Mises’ slogan was: ”first collective and then probability distribution”. Analysis of the GHZ paradox based on von Mises’ approach demonstrated that it is rather doubtful that there exists a collective which produces the probability distribution which is formally (via Kolmogorov’s approach) used in the GHZ considerations. Hence if we use a mathematical model of probability theory which is different from Kolmogorov’s model, namely von Mises’ model, we observe no paradox in the GHZ considerations. In particular, there is no contradiction between the quantum formalism with the frequency interpretation of probability and local realism.
Of course, our probabilistic considerations could not be considered as arguments in favour of either locality or determinism. It may be that physical reality is nonlocal or even nonreal. However, Bell’s as well as GHZ’s approaches do not give definite arguments to deny locality or determinism. Both these approaches are strongly based on the use of one particular model of probability theory, Kolmogorov’s model.
Our frequency analysis clarifies the measure-theoretical roots of the GHZ paradox. In fact, this paradox can be escaped even in the measure-theoretical approach if it would not be assumed that probability distributions corresponding to different settings of measurement devices are equivalent measures. We discuss the connection between equivalence/singularity dichotomy in measure theory and the existence of compatible and noncompatible observables. It seems that the splitting of physical reality to classical and quantum realities is just a consequence of the general (mathematical) property of probability measures. So it is just a property of our (mathematical) description of physical reality.
The Kolmogorov definition of a probability space is well known , . This is a triple ($`\mathrm{\Omega },F,𝐏`$), where $`\mathrm{\Omega }`$ is an abstract set, $`F`$ is a $`\sigma `$-field of subsets (events) of $`\mathrm{\Omega },𝐏`$ is the probability (normalized by 1 and $`\sigma `$-additive) measure on $`F.`$ On the other hand, the frequency probability theory of R. von Mises is now days practically forgotten. So we must present an extended introduction to this approach, see section 2.
We must remark that von Mises’ theory was strongly criticized due to rather informal definition of randomness, . In fact, this purely mathematical critique was one of the reasons to eliminate the frequency approach from quantum formalism. We do not relate our use of frequency theory to sophisticated mathematical problems of randomness . There are two main reasons to eliminate the problem of randomness from physical considerations and justify the use of the frequency formalism. The first is von Mises’ observation that the class of place selections must be determined not by some mathematical theory, but by the concrete physical phenomenon. This viewpoint is supported by Wald’s theorem by that if we fix a countable set of place selections, then there exist sufficiently many collectives with respect to this set of place selections. The second is my own observation that it seems to be that the property of randomness is not related (at least directly) to physical measurements (at least for present experiments). We are always interested only in one property of a sequence of observations: the statistical stabilization of relative frequencies $`\nu _N=\frac{n}{N}`$ to some limiting quantities $`𝐏`$ (probabilities).
## 2 Frequency probability theory
2.1. History. The frequency probability theory was developed by R. von Mises in 1919 (see , , for the details). In fact, the basis of the frequency approach was provided in the work of J. Venn, 1866, see . The frequency theory was used as the motivation of Kolmogorov’s axiomatic, 1933, of the conventional probability theory (see remarks in ). The main advantage of the conventional theory is its abstractness. Here we work with abstract probability distributions which are not directly related to the concrete physical model. Thus results of the conventional probability theory can be used without any modification in any physical models. However, this advantage may become in some circumstances a disadvantage, because the abstractness of the formalism does not give the possibility to analyse the origin (and even the existence) of probability distributions. On the other hand, the frequency theory of probability is concrete. Here to introduce a probability distribution, we must be sure that there exists a collective (random sequence) which produces this probability distribution. The collective is more primary object than a probability distribution. The collective has more direct connection with a physical phenomenon. However, in the frequency approach we cannot obtain results which are valid for ‘all probability distributions’. The probability distribution without a collective is nothing. Typically such a concreteness is considered as the large disadvantage of the frequency approach (comparing with the conventional measure theoretical approach). Of course, it is more attractive to prove some probabilistic statement ones and then to apply it to numerous physical models. This was one of the reasons to eliminate the frequency approach from applications in the favour of the measure-theoretical approach. <sup>1</sup><sup>1</sup>1Another reason was the problem of the rigorous mathematical definition of a collective, random sequence, see, for example, .
In the present paper we demonstrate that the frequency analysis of probabilistic assumptions for the derivation of Bell’s inequality can give some new sights to this problem. These sights would be impossible to obtain in the conventional abstract framework. Analysis of collectives can give more than analysis of abstract probability distributions.
2.2. Collective. Let $``$ be an ensemble of physical systems. We take elements of $``$ and form a sequence $`\pi =(\pi _1,\pi _2,\mathrm{},\pi _N,\mathrm{}).`$ Suppose that elements of $``$ have some properties. <sup>2</sup><sup>2</sup>2It is not important in general either these properties are objective (properties of an object) or ‘created’ in the process of observation by an observer, see . Suppose that these properties can be described by natural numbers, $`L=\{1,2,\mathrm{},m\}`$ (the set of ‘labels’). In principle we can consider continuous label sets, see . Thus, for each $`\pi _j\pi ,`$ we have a number $`\alpha _jL.`$ So $`\pi `$ induces a sequence
$$x=(\alpha _1,\alpha _2,\mathrm{},\alpha _N,\mathrm{}),\alpha _jL.$$
(1)
For each fixed $`\alpha L,`$ we have the relative frequency $`\nu _N(\alpha )=n_N(\alpha )/N`$ of the appearance of $`\alpha `$ in $`(\alpha _1,\alpha _2,\mathrm{},\alpha _N).`$
R. von Mises said that $`x`$ satisfies to the principle of the statistical stabilization of relative frequencies, if, for each fixed $`\alpha L,`$ $`|\nu _N(\alpha )\nu _M(\alpha )|0,N,M\mathrm{}.`$ The corresponding limit
$$𝐩(\alpha )=\underset{N\mathrm{}}{lim}\nu _N(\alpha )$$
(2)
is said to be a probability. This probability can be extended to the field of all subsets of $`L:`$
$$𝐩(B)=\underset{N\mathrm{}}{lim}\nu _N(\alpha B)=\underset{N\mathrm{}}{lim}\underset{\alpha B}{}\nu _N(\alpha )=\underset{\alpha B}{}𝐩(\alpha ),BL$$
(3)
(the situation becomes sufficiently complex for an infinite $`L,`$ see Tornier ). We remark that $`𝐩(L)=1.`$
R. von Mises said that $`x`$ satisfies the principle of randomness if limits (2) are invariant with respect to choices of some subsequences in $`x.`$ These choices of subsequences, so called place selections, have some properties, see , or (which are unimportant for our investigation). <sup>3</sup><sup>3</sup>3The class of place selections was not defined precisely by R. von Mises. This induced numerous discussions. However, the problem can be solved (at least partially) by the consideration of countable classes of place selections, Wald theorem, or , p.43. In principle the reader may forget about the principle of randomness and consider only the principle of the statistical stabilization. It seems that only this principle is important (at least at the moment) in physics in that we study behaviour of frequencies.
Sequence (1) which satisfies to two von Mises’ principles is said to be a collective; $`𝐩`$ is said to be a probability distribution of the collective $`x.`$ We will often use the symbols $`𝐩(B;x)`$ (and $`\nu _N(B;x),n_N(B;x)),BL,`$ to indicate the dependence on the concrete collective $`x.`$
The frequency probability formalism is not a calculus of probabilities. It is a calculus of collectives. Thus instead of operations for probabilities (as it is in the conventional probability theory), we define operations for collectives.
2.3. Operation of combining of collectives. This operation will play the crucial role in our analysis of probabilistic foundations of Bell’s arguments. Let $`x=(x_j)`$ and $`y=(y_j)`$ be two collectives with label sets $`L_x`$ and $`L_y`$, respectively. We define a new sequence $`z=(z_j),z_j=\{x_j,y_j\}`$ (in general $`z`$ is not a collective). Let $`aL_x`$ and $`bL_y`$. Among the first $`N`$ elements of $`z`$ there are $`n_N(a;z)`$ elements with the first component equal to $`a`$. As $`n_N(a;z)=n_N(a;x)`$ is a number of $`x_j=a`$ among the first $`N`$ elements of $`x`$, we obtain that $`lim_N\mathrm{}\frac{n_N(a;z)}{N}=𝐩(a;x)`$. Among these $`n_N(a;z)`$ elements, there are a number, say $`n_N(b/a;z)`$ whose second component is equal to $`b`$. The frequency $`\nu _N(a,b;z)`$ of elements of the sequence $`z`$ labeled $`(a,b)`$ will then be
$$\frac{n_N(b/a;z)}{N}=\frac{n_N(b/a;z)}{n_N(a;z)}\frac{n_N(a;z)}{N}.$$
We set $`\nu _N(b/a;z)=\frac{n_N(b/a;z)}{n_N(a;z)}`$. Let us assume that, for each $`aL_x`$, the subsequence $`y(a)`$ of $`y`$ which is obtained by choosing $`y_j`$ such that $`x_j=a`$ is an collective. Then, for each $`aL_x`$, $`bL_y`$, there exists
$$𝐩(b/a;z)=\underset{N\mathrm{}}{lim}\nu _N(b/a;z)=\underset{N\mathrm{}}{lim}\nu _N(b;y(a))=𝐩(b;y(a)).$$
(4)
We have $`_{bL_2}𝐩(b/a;z)=1.`$ The existence of $`𝐩(b/a;z)`$ implies the existence of $`𝐩(a,b;z)=lim_N\mathrm{}\nu _N(a,b;z)`$. Moreover, we have
$$𝐩(a,b;z)=𝐩(a;x)𝐩(b/a;z)$$
(5)
and $`𝐩(b/a;z)=𝐩(a,b;z)/𝐩(a;x),`$ if $`𝐩(a;x)0`$. We have
$$\underset{aL_a}{}\underset{bL_2}{}𝐩(a,b;z)=1.$$
Thus in this case the sequence $`z`$ is an collective and the probability distribution $`𝐩(a,b;z)`$ well defined. The collective $`y`$ is said to be combinable with the collective $`x`$. The relation of combining is a symmetric relation on the set of pairs of collectives with strictly positive probability distributions $`(𝐩>0).`$
2.4. Independent collectives. Let $`x`$ and $`y`$ be collectives. Suppose that they are combinable. The $`y`$ is said to be independent from $`x`$ if all collectives $`y(a)`$, $`aL_x`$, have the same probability distribution which coincides with the probability distribution $`𝐩(b;y)`$ of $`y`$. This implies that
$$𝐩(b/a;z)=\underset{N\mathrm{}}{lim}\nu _N(b/a;z)=\underset{N\mathrm{}}{lim}\nu _N(b;y(a))=𝐩(b;y).$$
Here the conditional probability $`𝐩(b/a;z)`$ does not depend on $`a.`$ Hence
$$𝐩(a,b;z)=𝐩(a;x)𝐩(b;y),aL_x,bL_y.$$
From the physical viewpoint the notion of independent collectives is more natural than the notion of independent events in the conventional probability theory. In latter the relation $`𝐩(a,b)=𝐩(a)𝐩(b)`$ can hold just occasionally (as the result of a game with numbers, see or , p.53).
## 3 Kolmorogov’s viewpoint to the GHZ scheme
From the probabilistic viewpoint the GHZ experiment can be described in the following way ( in the Kolmorogov approach). Let $`(\mathrm{\Omega },\mathrm{F},𝐏)`$ be a Kolmogorov probability space which describes hidden variables. For each setting ($`\varphi _1,\varphi _2,\varphi _3)`$ of phase shifts we define random variables $`\mathrm{A}(\varphi _1,\omega ),\mathrm{B}(\varphi _2,\omega ),\mathrm{C}(\varphi _3,\omega )`$ corresponding to physical observables $`\mathrm{A}(\varphi _1),\mathrm{B}(\varphi _2),\mathrm{C}(\varphi _3)`$ (given by measurements for photons 1,2,3 respectively, in the triple (1,2,3)). Quantum formalism predicts that there exist four settings ($`\varphi _1^i,\varphi _2^i,\varphi _3^i),i=1,2,3,4`$ such that
$$\mathrm{A}(\varphi _1^i,\omega )\mathrm{B}(\varphi _2^i,\omega )\mathrm{C}(\varphi _3^i,\omega )=1,$$
(6)
$$\omega \mathrm{\Omega }_\mathrm{i}^+\mathrm{F},𝐏(\mathrm{\Omega }_\mathrm{i}^+)=1,\mathrm{i}=1,2,3;$$
(7)
$$\mathrm{A}(\varphi _1^4,\omega )\mathrm{B}(\varphi _2^4,\omega )\mathrm{C}(\varphi _3^4,\omega )=1,$$
(8)
$$\omega \mathrm{\Omega }_4^{}\mathrm{F},𝐏(\mathrm{\Omega }_4^{})=1.$$
(9)
By using algebraic properties $`(\mathrm{A},\mathrm{B},\mathrm{C}=\pm 1)`$ we obtain that
$$\mathrm{\Sigma }^+=\mathrm{\Omega }_1^+\mathrm{\Omega }_2^+\mathrm{\Omega }_3^+\mathrm{\Omega }_4^+=\mathrm{\Omega }\mathrm{\Omega }_4^{}.$$
(10)
The trivial mathematical considerations in Kolmorogov’s framework imply that by (7)
$$𝐏(\mathrm{\Sigma }^+)=1.$$
(11)
On the other hand, by (9) and (10) we have
$$𝐏(\mathrm{\Sigma }^+)=0.$$
(12)
This is the GHZ paradox. The typical conclusion is that we could not use the local deterministic description.
From the Kolmorogov viewpoint it seems that all was right in the GHZ derivation.
## 4 Von Mises’ viewpoint to the GHZ paradox
Here we could not start with an abstract probability distribution of hidden parameters. First we have to define a collective which produces this distribution. To introduce a collective, we have to define the label set $`L`$ of this collective. It is convenient to use symbol $`\mathrm{\Omega }`$ instead of $`L`$ (to use formulas of the previous section). However, it is just the same symbol and nothing more. Here $`\mathrm{\Omega }`$ has the following structure: $`\mathrm{\Omega }=\mathrm{\Lambda }\times \mathrm{\Lambda }_1\times \mathrm{\Lambda }_2\times \mathrm{\Lambda }_3,`$ where $`\mathrm{\Lambda }`$ is the set of hidden variables for a quantum system (a triple of photons), $`\mathrm{\Lambda }_\mathrm{j},\mathrm{j}=1,2,3,`$ are sets of hidden variables for measurement devices (for $`A,B`$ and $`C,`$ respectively). <sup>4</sup><sup>4</sup>4To simplify considerations, we assume that all sets of hidden variables are finite.
For each setting $`\varphi _1,\varphi _2,\varphi _3`$ of phase shifts, we may consider (in the hidden variables framework) a sequence $`\mathrm{x}_{\varphi _1\varphi _2\varphi _3}=(\omega _1,\omega _2,\mathrm{},\omega _\mathrm{N}.\mathrm{}),\omega _\mathrm{j}=(\lambda _\mathrm{j},\lambda _\mathrm{j}^1,\lambda _\mathrm{j}^2,\lambda _\mathrm{j}^3)\mathrm{\Omega },`$ where $`\omega _\mathrm{j}`$ is the configuration of hidden variables for $`\mathrm{jth}`$ quantum system $`\pi _\mathrm{j}`$ (a triple of photons) + three measurement devices at the instants of measurements $`j=1,2,\mathrm{}`$
The first question is the following: Is $`x_{\varphi _1\varphi _2\varphi _3}`$ a collective? We have no experimental reasons to suppose that micro parameters have the property of the statistical stabilization (as macro parameters). It may be that the property of the statistical stabilization on the macro level is just a consequence of the average over huge ensembles of hidden parameters. Well, suppose that $`x_{\varphi _1\varphi _2\varphi _3}`$ is a collective. Thus the frequency probability distribution
$$𝐏_{\varphi _1\varphi _2\varphi _3}(\lambda =k,\lambda ^1=s_1,\lambda ^2=s_2,\lambda ^3=s_3)=$$
$$\underset{N\mathrm{}}{lim}\frac{n_N(\lambda =k,\lambda ^1=s_1,\lambda ^2=s_2,\lambda ^3=s_3)}{N}$$
is well defined. <sup>5</sup><sup>5</sup>5The consideration of hidden variables for measurement apparatuses is quite natural from the physical viewpoint. In fact, it is the hidden variable representation of Bohr’s ideas. So, for four different settings $`(\varphi _1^\mathrm{i},\varphi _2^\mathrm{i},\varphi _3^\mathrm{i})`$ of phase shifts we have four collectives $`x^\mathrm{i}=x_{\varphi _1^\mathrm{i},\varphi _2^\mathrm{i},\varphi _3^\mathrm{i}},\mathrm{i}=1,2,3,4,`$ with probability distributions $`𝐏_i𝐏_{x^\mathrm{i}}.`$ By the GHZ scheme we obtain that
$$𝐏_i(\mathrm{\Omega }_\mathrm{i}^+)=1,\mathrm{i}=1,2,3,\text{and}𝐏_4(\mathrm{\Omega }_4^+)=0.$$
(13)
Of course, by (10) we obtain
$$𝐏_4(\mathrm{\Sigma }^+)=0.$$
(14)
However, the first three equations in (13) do not imply that
$$𝐏_4(\mathrm{\Sigma }^+)=1.$$
(15)
Hence there is no paradox. To obtain the paradox, we need to obtain (15). Thus there must be some special restrictions on collectives (and consequently probability distributions) which imply (15). One of such restrictions is that the probability distribution does not depend on the setting $`(\varphi _1,\varphi _2,\varphi _3)`$ of phase shifts:
$$𝐏=𝐏_{\varphi _1,\varphi _2,\varphi _3}.$$
(16)
However, such an assumption has no physical justification (compare with -). First of all we have to assume so called ensemble reproducibility for hidden variable $`\lambda `$ (see and ): the preparation procedure for quantum systems must precisely reproduce the probability distribution of hidden variables in different runs of the experiment (in particular, for different settings $`\varphi _1,\varphi _2,\varphi _3`$). Despite of the common opinion that such a reproducibility is a natural property of quantum systems (preparation procedures), at the present stage of experimental research it is impossible to test this hypothesis. Moreover, the hypothesis of reproducibility is a form of the postulate on the completeness of quantum mechanics. By the hypothesis on reproducibility we have that quantum state $`\psi `$ uniquely determines all statistical properties of the (ideal infinite) ensemble of quantum particles described by $`\psi .`$
So by accepting this hypothesis we turn back (at least indirectly) to the original discussion of Einstein, Podolsky, Rosen and Bohr on the completeness of quantum mechanics. In some sense this is the logical loop, because one of the main aims of J.Bell and his followers was to transform the EPR polemic on the completeness of quantum mechanics into polemic on locality and determinism.
Remark (On the interpretation of a wave function).Of course, all our previous considerations on the hypothesis of reproducibility and the completeness of quantum mechanics strongly depend on the interpretation of a wave function. In fact, we used so called statistical interpretation of quantum mechanics (see, for example, L. Ballentine ): a wave function gives the description of statistical properties of an ensemble of quantum particles. Here the statistical reproducibility of macro properties need not be based on the statistical reproducibility of micro properties. For an adherent of the orthodox Copenhagen interpretation (by that the wave function provides the complete description of an individual quantum system), there are no doubts in the validity of the hypothesis of reproducibility.
However, even if we suppose that there are no ensemble fluctuations, there are still some doubts in the validity of (16). It is more natural to think that different settings of apparatuses produce different distributions of micro states of these apparatuses (compare with -).
## 5 Singularity/equivalence dichotomy and the principle of complementarity
Of course, (16) is only a sufficient condition for obtaining the GHZ paradox. In fact, we need only that
$$𝐏_{\varphi _1\varphi _2\varphi _3}(\mathrm{E})=0𝐏_{\varphi _1^{}\varphi _2^{}\varphi _3^{}}(\mathrm{E})=0$$
(17)
for any two settings $`\varphi _1\varphi _2\varphi _3`$ and $`\varphi _1^{}\varphi _2^{}\varphi _3^{}`$ of measurement devices. This condition is well known in the measure theory, namely this is the condition of equivalence of two measures: they are absolutely continuous with respect to each other. The absolute continuity implies that the transition from one setting of measurement devices to another is sufficiently smooth (in measure-theoretical sense). There exists so called Radon- Nikodim derivative:
$$\frac{\mathrm{d}𝐏_{\varphi _1\varphi _2\varphi _3}}{\mathrm{d}𝐏_{\varphi _1^{}\varphi _2^{}\varphi _3^{}}}(\omega )=\mathrm{f}(\omega ;\varphi _1\varphi _2\varphi _3/\varphi _1^{}\varphi _2^{}\varphi _3^{}).$$
The GHZ paradox (via our frequency analysis) demonstrated that quantum measurement procedures induce probability distributions which transform nonsmoothly (in measure-theoretical sense) from one setting to another.
Measure-theoretical singularity is described by the notion of singularity: $`𝐏^{}𝐏^{\prime \prime }`$ if there is a set $`\mathrm{E}\mathrm{F}`$ such that $`𝐏^{\prime \prime }(\mathrm{E})=1`$ and $`𝐏^{}(\mathrm{E})=0.`$ Suppose that $`𝐏_\mathrm{i}𝐏_\mathrm{j},\mathrm{i},\mathrm{j}=1,2,\mathrm{},4,`$ where $`𝐏_\mathrm{j}`$ are probability distributions in the GHZ scheme. Let $`\mathrm{\Omega }_\mathrm{j}^+,\mathrm{j}=1,2,3,`$ play the role of E in the definition of $`𝐏_\mathrm{j}𝐏_4:𝐏_\mathrm{j}(\mathrm{\Omega }_\mathrm{j}^+)=1`$ and $`𝐏_4(\mathrm{\Omega }_\mathrm{j})=0,\mathrm{j}=1,2,3.`$ Then $`𝐏_4(\mathrm{\Sigma }^+)=𝐏_4(\mathrm{\Omega }_1^+\mathrm{\Omega }_2^+\mathrm{\Omega }_3^+)=0.`$ Thus there is no GHZ paradox.
We remark that if the space of hidden variables has infinite dimension, then, for many classes of probability distributions (in particular, Gaussian), we have equivalence/singularity dichotomy: either equivalent or singular . It may be that the split of reality into classical and quantum is just the exhibition of such a dichotomy.
## 6 ‘Gedanken kollektiven’(counterfactural arguments)
We note that in the frequency approach the GHZ paradox can be obtained via counterfactural arguments (compare with , ). These arguments are represented here via the use of ’gedanken kollektiven’. In fact, the GHZ scheme is applied to four settings $`(\pi /2,0,0),(0,\pi /2,0),(0,0,\pi /2),(\pi /2,\pi /2,\pi /2).`$ Let us consider a ‘gedanken kollektiv’ corresponding to the simultaneous imaginary measurement for all angles involved in the GHZ scheme: $`\varphi _1=0,\pi /2,\varphi _2=0,\pi /2,\varphi _3=0,\pi /2.`$ Such an imaginary measurement would be described by the hidden variable:
$$\stackrel{~}{\omega }=(\lambda ,\lambda _0^1,\lambda _{\pi /2}^1,\lambda _0^2,\lambda _{\pi /2}^2,\lambda _0^3,\lambda _{\pi /2}^3).$$
(18)
Of course, such a measurement is forbidden by the quantum theory. We recognize this. However, we continue our frequency analysis trying to find the origin of the impossibility of such a measurement. We may image that there are two settings $`\varphi _1=0,\pi /2`$ for the first photon, two settings $`\varphi _2=0,\pi /2`$ for the second photon and two settings $`\varphi _3=0,\pi /2`$ for the third photon in the triple. At the moment of interaction (imaginary) with photons measurement devices with these settings have hidden parameters included in (18). If we assume that the sequence of parameters $`\stackrel{~}{\omega }`$ corresponding to the sequence of imaginary measurements, $`x=(\stackrel{~}{\omega }_\mathrm{j},\mathrm{j}=1,\mathrm{},\mathrm{})`$ is a collective, then we obtain the frequency probability distribution $`𝐏=𝐏_\mathrm{x}`$ which can be used in the GHZ scheme (and induce the paradox). The origin of the nonexistence of $`𝐏`$ (statistical stabilization in $`\mathrm{x}`$) is that collectives corresponding to incompatible settings of measurement devices are not combinable: $`\mathrm{x}_1=(\stackrel{~}{\omega }_\mathrm{j}^1),\stackrel{~}{\omega }_\mathrm{j}^1=(\lambda ,\lambda ^1,\lambda _0^2,\lambda _0^3),`$ and $`\mathrm{x}_2=(\stackrel{~}{\omega }_\mathrm{j}^2),\stackrel{~}{\omega }_\mathrm{j}^2=(\lambda ,\lambda _{\pi _2}^1,\lambda _{\pi _2}^2,\lambda _{\pi _2}^3)`$ or $`\mathrm{x}_1=(\stackrel{~}{\omega }_\mathrm{j}^1),\stackrel{~}{\omega }_\mathrm{j}^1=(\lambda ,\lambda _{\pi _2}^1,\lambda _0^2,\lambda _0^3),`$ and $`\mathrm{x}_2=(\stackrel{~}{\omega }_\mathrm{j}^2),\stackrel{~}{\omega }_\mathrm{j}^2=(\lambda ,\lambda _0^1,\lambda _{\pi _2}^2,\lambda _{\pi _2}^3),\mathrm{}`$
Thus our frequency counterfactural analysis demonstrated again that the origin of the GHZ paradox is the existence of incompatible settings of measurement apparatuses (uncombinable collectives).
References
J.S. Bell, Rev. Mod. Phys., 38, 447–452 (1966). J. S. Bell, Speakable and unspeakable in quantum mechanics. Cambridge Univ. Press (1987).
J.F. Clauser , M.A. Horne, A. Shimony, R. A. Holt, Phys. Rev. Letters, 49, 1804-1806 (1969); J.F. Clauser , A. Shimony, Rep. Progr.Phys., 41 1881-1901 (1978). A. Aspect, J. Dalibard, G. Roger, Phys. Rev. Lett., 49, 1804-1807 (1982); D. Home, F. Selleri, Nuovo Cim. Rivista, 14, 2–176 (1991). H. P. Stapp, Phys. Rev., D, 3, 1303-1320 (1971); P.H. Eberhard, Il Nuovo Cimento, B, 38, N.1, 75-80(1977); Phys. Rev. Letters, 49, 1474-1477 (1982); A. Peres, Am. J. of Physics, 46, 745-750 (1978). P. H. Eberhard, Il Nuovo Cimento, B, 46, N.2, 392-419 (1978); J. Jarrett, No s, 18, 569 (1984).
B. d’Espagnat, Veiled Reality. An anlysis of present-day quantum mechanical concepts. Addison-Wesley(1995). A. Shimony, Search for a naturalistic world view. Cambridge Univ. Press (1993); T. Maudlin, Quantum non-locality and relativity. Blackwill.
L. de Broglie, La thermodynamique de la particule isolee. Gauthier-Villars, Paris, 1964; G. Lochak, Found. Physics, 6 , 173-184 (1976); E. Nelson, Quantum fluctuation. Princeton Univ. Press, 1985; W. De Muynck and W. De Baere W., Ann. Israel Phys. Soc., 12, 1-22 (1996); W. De Muynck, W. De Baere, H. Marten, Found. of Physics, 24, 1589–1663 (1994); W. De Muynck, J.T. Stekelenborg, Annalen der Physik, 45, N.7, 222-234 (1988).
E. Beltrametti, G. Cassinelli, The logic of quantum mechanics. Addison-Wesley, Reading (1981).
L. Accardi, Urne e Camaleoni: Dialogo sulla realta, le leggi del caso e la teoria quantistica. Il Saggiatore, Rome (1997); Accardi L., The probabilistic roots of the quantum mechanical paradoxes. The wave–particle dualism. A tribute to Louis de Broglie on his 90th Birthday, Edited by S. Diner, D. Fargue, G. Lochak and F. Selleri. D. Reidel Publ. Company, Dordrecht, 47–55(1984);
I. Pitowsky, Phys. Rev. Lett, 48, N.10, 1299-1302 (1982); Phys. Rev. D, 27, N.10, 2316-2326 (1983); S.P. Gudder, J. Math Phys., 25, 2397- 2401 (1984); A. Fine, Phys. Rev. Letters, 48, 291–295 (1982); P. Rastal, Found. Phys., 13, 555 (1983). W. Muckenheim, Phys. Reports, 133, 338–401 (1986); W. De Baere, Lett. Nuovo Cimento, 39, 234-238 (1984); 25, 2397- 2401 (1984).
A. Yu. Khrennikov, Dokl. Akad. Nauk SSSR, ser. Matem., 322, No. 6, 1075–1079 (1992); J. Math. Phys., 32, No. 4, 932–937 (1991); Physics Letters A, 200, 119–223 (1995); Physica A, 215, 577–587 (1995); Int. J. Theor. Phys., 34, 2423–2434 (1995); J. Math. Phys., 36, No.12, 6625–6632 (1995); A.Yu. Khrennikov, $`p`$-adic valued distributions in mathematical physics. Kluwer Academic Publishers, Dordrecht (1994); A.Yu. Khrennikov, Non-Archimedean analysis: quantum paradoxes, dynamical systems and biological models. Kluwer Acad.Publ., Dordreht, The Netherlands, 1997;
A. Yu. Khrennikov, Bell and Kolmogorov: probability, reality and nonlocality. Reports of Vaxjo Univ., N. 13 (1999); A. Yu. Khrennikov, Interpretations of probability. VSP Int. Sc. Publ., Utrecht, 1999.
A. Einstein, B. Podolsky, N. Rosen, Phys. Rev., 47, 777–780 (1935).
D. Greenberger, M. Horne, A. Zeilinger, Going beyond Bell’s theorem, in Bell’s theorem, quantum theory, and conceptions of the universe. Ed. M.Kafatos, Kluwer Academic, Dordrecht, 73-76 (1989).
A. N. Kolmogoroff, Grundbegriffe der Wahrscheinlichkeitsrech Springer Verlag, Berlin (1933); reprinted: Foundations of the Probability Theory. Chelsea Publ. Comp., New York (1956).
R. von Mises, Grundlagen der Wahrscheinlichkeitsrechnung.Math.Z., 5, 52–99 (1919); R. von Mises, Probability, Statistics and Truth, Macmillan, London (1957).
R. von Mises, The mathematical theory of probability and statistics. Academic, London (1964); E. Tornier , Wahrscheinlichkeitsrechnunug und allgemeine Integrationstheorie., Univ. Press, Leipzing (1936);
A. N. Shiryayev, Probability. Springer, New York-Berlin-Hei (1984).
E. Kamke, Über neuere Begründungen der Wahrscheinlichkeitsrechnung. Jahresbericht der Deutschen Matemati, 42, 14-27 (1932); J. Ville, Etude critique de la notion de collective, Gauthier– Villars, Paris (1939);M. Van Lambalgen, Von Mises’ definition of random sequences reconsidered. J. of Symbolic Logic, 52, N. 3 (1987).
A. Wald, Die Widerspruchsfreiheit des Kollektivbegriffs in der Wahrscheinlichkeitsrechnung. Ergebnisse eines Math. Kolloquiums, 8, 38-72 (1938).
Venn J., The logic of chance. London (1866); reprint: Chelsea, New York (1962).
L. E. Ballentine, Rev. Mod. Phys., 42, 358–381 (1970).
I. Kvart, A theory of Counterfactuals. Indiapolis: Hackett(1986); D. Lewis, Counterfactuals. Cambridge, Mass.: Harvard Univ. Press (1973);
|
warning/0006/hep-th0006207.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The traditional goal of fundamental physics is to infer the rules by which the “present state” of a system’s dynamical variables determines their future state. Since Newton’s time, most attention has been given to models for which the “present state” of a system’s dynamical variables means their values at some instant in time and possibly also the values of their first time derivatives. This restriction corresponds to equations of motion which are local in time and contain no more than second time derivatives. It has not so far proved useful, in describing the physical universe on the most fundamental level, to invoke equations of motion which are either nonlocal in time or which even possess more than two time derivatives.
The deep reason behind this surprising simplification of fundamental theory seems to be the result obtained by the 19th century physicist Ostrogradski . He showed that Lagrangians which possess a finite number of higher time derivatives, and are not degenerate in the highest one, must give rise to Hamiltonians which are linear in essentially half of the canonical variables. This is a nonperturbative result. Further, it cannot be altered by quantization since the instability occurs over a large volume of the canonical phase space. I will review Ostrogradski’s construction in Section 2 of this paper. For now it suffices to note that the instability must apply as well to nonlocal theories which can be represented as the limits of ever higher derivative ones.
Much of the interest in nonlocal quantum field theories has been driven by the close connection between ultraviolet divergences and local interactions . Of course it does no good to avoid divergences by introducing an infinite number of instabilities against which there is not even any barrier to decay! It is therefore of interest to know when a nonlocal Lagrangian possesses a higher derivative representation and, consequently, the Ostrogradskian instability. The higher derivative representation does seem to be valid for cases such as string field theory, where the nonlocality enters through entire functions of the derivative operator and the Lagrangian cannot be made local by a field redefinition . On the other hand, the higher derivative representation is certainly not valid for the inverse differential operators which result from integrating out a local field variable. It also fails for “maximal nonlocality” in which the action is a nonlinear function of local actions . The purpose of this paper is to demonstrate by construction that the higher derivative representation is valid for “nonlocality of finite extent” in which the Lagrangian connects no times differing by more than some constant $`\mathrm{\Delta }t`$.
Although the results of this paper apply as well to field theories I will work in the context of a one dimensional, point particle whose position as a function of time is $`q(t)`$. A nonlocal Lagrangian of finite extent $`\mathrm{\Delta }t`$ is one which definitely depends upon (and mixes) $`q(t)`$ and $`q(t+\mathrm{\Delta }t)`$, and potentially depends as well upon $`q(t^{})`$ for $`t<t^{}<t+\mathrm{\Delta }t`$. An example would be the following nonlocal generalization of the harmonic oscillator:
$$L[q](t)=\frac{1}{2}m\dot{q}^2(t+\mathrm{\Delta }t/2)\frac{1}{2}m\omega ^2q(t)q(t+\mathrm{\Delta }t).$$
(1)
The deterministic way of viewing such theories is that the equations of motion give the dynamical variable at the latest time — $`q(t+\mathrm{\Delta }t)`$ — as a function of earlier times in the range $`t\mathrm{\Delta }tt^{}<t+\mathrm{\Delta }t`$. In our example the equation of motion is
$$_0^{\mathrm{\Delta }t}𝑑r\frac{\delta L[q](tr)}{\delta q(t)}=m\left\{\ddot{q}(t)+\frac{1}{2}\omega ^2q(t+\mathrm{\Delta }t)+\frac{1}{2}\omega ^2q(t\mathrm{\Delta }t)\right\}=0,$$
(2)
and its deterministic interpretation is
$$q(t+\mathrm{\Delta }t)=q(t\mathrm{\Delta }t)\frac{2}{\omega ^2}\ddot{q}(t).$$
(3)
This paper contains six sections of which this Introduction is the first. Section 2 is devoted to a review of Ostrogradski’s result for local Lagrangians depending upon $`N`$ time derivatives. My canonical formalism is presented in Section 3 and shown to correctly realize the dynamics of nonlocal Lagrangians of finite extent. This formalism is applied in Section 4 to the Lagrangian (1) discussed above. The connection with Ostrogradski’s formalism is demonstrated in Section 5. My conclusions comprise Section 6.
## 2 Ostrogradski’s construction
Consider a Lagrangian $`L(q,\dot{q},\mathrm{},q^{(N)})`$ which depends upon the first $`N`$ derivatives of the dynamical variable $`q(t)`$. I shall assume only that the Lagrangian is nondegenerate, i.e., that the equation
$$P_N=\frac{L}{q^{(N)}},$$
(4)
can be inverted to solve for $`q^{(N)}`$ as a function of $`P_N`$, $`q`$ and the first $`N1`$ derivatives of $`q`$. This just means that the action’s dependence upon $`q^{(N)}`$ cannot be eliminated by partial integration, so the equation of motion,
$$\underset{I=0}{\overset{N}{}}\left(\frac{d}{dt}\right)^I\frac{L}{q^{(I)}}=0,$$
(5)
contains $`q^{(2N)}`$.
Since the equation of motion determines $`q^{(2N)}`$ as a function of $`q`$ and its first $`2N1`$ derivatives, one can obviously specify the initial values of these $`2N`$ variables. The canonical phase space must accordingly contain $`N`$ coordinates and $`N`$ conjugate momenta. In Ostrogradski’s construction the $`I`$-th coordinate is just the $`(I1)`$-th derivative of $`q`$,
$$Q_Iq^{(I1)}.$$
(6)
The momentum canonically conjugate to $`Q_I`$ is,
$$P_I=\underset{J=I}{\overset{N}{}}\left(\frac{d}{dt}\right)^{JI}\frac{L}{q^{(J)}}.$$
(7)
A consequence of nondegeneracy is that the derivatives $`q^{(N+I)}`$ can be determined from $`P_{NI},P_{NI+1},\mathrm{},P_N`$ and the $`Q_J`$’s. In particular $`q^{(N)}`$ involves only $`P_N`$ and the $`Q_J`$’s,
$$q^{(N)}=𝒬(\stackrel{}{Q},P_N).$$
(8)
Ostrogradski’s Hamiltonian is,
$`H`$ $`=`$ $`{\displaystyle \underset{I=1}{\overset{N}{}}}P_I\dot{Q}_IL,`$ (9)
$`=`$ $`{\displaystyle \underset{I=1}{\overset{N1}{}}}P_IQ_{I+1}+P_N𝒬(\stackrel{}{Q},P_N)L(\stackrel{}{Q},𝒬(\stackrel{}{Q},P_N)),`$ (10)
and his canonical equations are the ones suggested by the notation,
$$\dot{Q}_I=\frac{H}{P_I},\dot{P}_I=\frac{H}{Q_I}.$$
(11)
It is straightforward to check that the various canonical evolution equations reproduce the equation of motion and the structure of the canonical formalism: $`\dot{Q}_I`$ gives the canonical definition (6) for $`Q_{I+1}`$; $`\dot{Q}_N`$ gives the canonical definition for $`P_N`$ in its inverse form (8); $`\dot{P}_{I+1}`$ gives the canonical definition (7) for $`P_I`$; and $`\dot{P}_1`$ gives the equation of motion (5). So there is no doubt that Ostrogradski’s Hamiltonian generates time evolution. When the Lagrangian is free of explicit time dependence $`H`$ is also the conserved current associated with time translation invariance.
The instability consequent upon $`H`$’s linearity in $`P_1,P_2,\mathrm{},P_{N1}`$ explains why higher derivative theories have not been of use in describing physics on the fundamental level. Note the generality of the problem. It does not depend upon any approximation scheme, nor upon any feature of the Lagrangian except nondegeneracy. Further, it must continue to afflict the theory after quantization because the instability is not confined to a small region of the classical phase space. If a fully nonlocal Lagrangian can be represented as the limit of such higher derivative Lagrangians it must inherit their instability.
The limit of infinite $`N`$ is facilitated by regarding Ostrogradski’s formalism as the result of constraining a larger system with an extra pair of canonical variables,
$$Q_{N+1}q^{(N)},P_{N+1}0.$$
(12)
The Hamiltonian is,
$$H=\underset{I=1}{\overset{N}{}}P_IQ_{I+1}L(\stackrel{}{Q},Q_{N+1}),$$
(13)
and requiring that $`P_{N+1}`$ remains zero imposes the canonical definition of $`P_N`$ as another constraint,
$$\dot{P}_{N+1}=\frac{H}{Q_{N+1}}=P_N+\frac{L}{Q_{N+1}}0.$$
(14)
Since the Poisson bracket with $`P_{N+1}`$ gives the second derivative of the Lagrangian with respect to $`Q_{N+1}`$, nondegeneracy implies that the two constraints are second class. The resulting Dirac brackets are,
$`\{Q_I,Q_J\}_D`$ $`=`$ $`\left(\delta _{IN+1}\delta _{JN}+\delta _{IN}\delta _{JN+1}\right)\left[{\displaystyle \frac{^2L}{Q_{N+1}^2}}\right]^1,`$ (15)
$`\{Q_I,P_J\}_D`$ $`=`$ $`\delta _{IJ}\delta _{IN+1}\left[{\displaystyle \frac{^2L}{Q_{N+1}^2}}\right]^1{\displaystyle \frac{^2L}{Q_JQ_{N+1}}},`$ (16)
$`\{P_I,P_J\}_D`$ $`=`$ $`0.`$ (17)
Note that there is not even any difference between Dirac brackets and Poisson brackets provided one avoids the highest $`Q`$ — that is, $`Q_{N+1}`$.
## 3 My construction for finite nonlocality
I define a nonlocal Lagrangian $`L[q](t)`$ of finite extent $`\mathrm{\Delta }t`$ as one which potentially depends upon the dynamical variable from time $`t`$ to time $`t+\mathrm{\Delta }t`$, with guaranteed mixing between $`q(t)`$ and $`q(t+\mathrm{\Delta }t)`$. The requirement of mixing is the generalization of nondegeneracy and it implies,
$$\frac{\delta ^2L[q](t)}{\delta q(t)\delta q(t+\mathrm{\Delta }t)}0.$$
(18)
I shall also require that the Lagrangian contain no derivatives of either $`q(t)`$ or $`q(t+\mathrm{\Delta }t)`$.
I label the canonical variables by a continuum parameter $`0s\mathrm{\Delta }t`$. They are defined as follows:
$`Q(s,t)`$ $``$ $`q(s+t),`$ (19)
$`P(s,t)`$ $``$ $`{\displaystyle _s^{\mathrm{\Delta }t}}𝑑r{\displaystyle \frac{\delta L[q](s+tr)}{\delta q(s+t)}}.`$ (20)
Note that (20) implies the constraint $`P(\mathrm{\Delta }t,t)0`$. Note also that whereas the $`s`$ and $`t`$ derivatives of $`Q(s,t)`$ are identical those of $`P(s,t)`$ are not,
$$\frac{d}{ds}P(s,t)=\frac{d}{dt}P(s,t)\frac{\delta L[q](t)}{\delta q(s+t)}.$$
(21)
Since $`L[q](t)`$ involves the dynamical variable from $`q(t)`$ up to $`q(t+\mathrm{\Delta }t)`$ we see that $`P(s,t)`$ involves $`q(s+t\mathrm{\Delta }t)`$ up to $`q(t+\mathrm{\Delta }t)`$. So decreasing $`s`$ allows one to reach back further before time $`t`$, all the way to time $`t\mathrm{\Delta }t`$ at $`s=0`$.
Note that the equation of motion is $`P(0,t)=0`$. This emerges as an additional constraint from surface variations of the canonical Hamiltonian,
$$H(t)_0^{\mathrm{\Delta }t}𝑑rP(r,t)\frac{d}{dr}Q(r,t)L[Q](t).$$
(22)
We can find the canonical equations of time evolution from the fact that the only nonzero Poisson bracket is,
$$\{Q(r,t),P(s,t)\}=\delta (rs).$$
(23)
The result for $`Q(s,t)`$ is straightforward,
$`{\displaystyle \frac{d}{dt}}Q(s,t)`$ $`=`$ $`\{Q(s,t),H(t)\},`$ (24)
$`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑r\delta (rs){\displaystyle \frac{d}{dr}}Q(r,t),`$ (25)
$`=`$ $`{\displaystyle \frac{d}{ds}}Q(s,t).`$ (26)
A partial integration is necessary for $`P(s,t)`$ and one must be careful about the resulting surface terms,
$`{\displaystyle \frac{d}{dt}}P(s,t)`$ $`=`$ $`\{P(s,t),H(t)\},`$ (27)
$`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑rP(r,t){\displaystyle \frac{d}{dr}}\delta (rs)+{\displaystyle \frac{\delta L[Q](t)}{\delta Q(s,t)}},`$ (28)
$`=`$ $`{\displaystyle \frac{d}{ds}}P(s,t)+{\displaystyle \frac{\delta L[Q](t)}{\delta Q(s,t)}}\delta (rs)P(r,t)|{\displaystyle \frac{}{}}_0^{\mathrm{\Delta }t}.`$ (29)
For $`0<s<\mathrm{\Delta }t`$ this simply reproduces (21), and hence the canonical definition of $`P(s,t)`$.
Since the two surface terms cannot be canceled by anything else, they must be imposed as constraints,
$$P(s,t)0s=0,\mathrm{\Delta }t.$$
(30)
Requiring that they be preserved under time evolution implies two additional constraints,
$$\frac{d}{ds}P(s,t)+\frac{\delta L[Q](t)}{\delta Q(s,t)}0s=0,\mathrm{\Delta }t.$$
(31)
Nondegeneracy — and the absence in $`L[q](t)`$ of derivatives of $`q(t)`$ and/or $`q(t+\mathrm{\Delta }t)`$ — guarantees that the four constraints are second class.
Note that the $`H(t)`$ is conserved when $`L[q](t)`$ is free of explicit time dependence,
$`{\displaystyle \frac{dH}{dt}}(t)`$ $`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑s\left\{{\displaystyle \frac{dP}{dt}}(s,t){\displaystyle \frac{dQ}{ds}}(s,t)+P(s,t){\displaystyle \frac{d^2Q}{ds^2}}(s,t)\right\}{\displaystyle \frac{d}{dt}}L[Q](t),`$ (33)
$`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑s{\displaystyle \frac{d}{ds}}\left[P(s,t){\displaystyle \frac{d}{ds}}Q(s,t)\right]`$
$`+{\displaystyle _0^{\mathrm{\Delta }t}}𝑑s{\displaystyle \frac{\delta L[Q](t)}{\delta Q(s,t)}}{\displaystyle \frac{d}{ds}}Q(s,t){\displaystyle \frac{d}{dt}}L[Q](t),`$
$`=`$ $`P(s,t){\displaystyle \frac{d}{ds}}Q(s,t)|_0^{\mathrm{\Delta }t},`$ (34)
$``$ $`0.`$ (35)
Note also that the Hamiltonian has inherited the Ostrogradskian instability. After eliminating the constraints it must be linear in all the $`P(s,t)`$ except possibly $`\frac{d}{ds}P(s,t)`$ at $`s=0`$ and at $`s=\mathrm{\Delta }t`$.
## 4 A simple example
It is useful to see how the general construction given in the previous section applies to the Lagrangian (1) presented in Section 1. Of course the canonical coordinates are always $`Q(s,t)=q(s+t)`$ for $`0s\mathrm{\Delta }t`$. To find the canonical momenta note that the functional derivative of the Lagrangian is,
$`{\displaystyle \frac{\delta L[q](t)}{\delta q(s+t)}}`$ $`=`$ $`m\dot{q}\left(t+{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)\delta ^{}\left(s{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)`$ (36)
$`{\displaystyle \frac{1}{2}}m\omega ^2\left[{\displaystyle \frac{}{}}q(t+\mathrm{\Delta }t)\delta (s)+q(t)\delta (s\mathrm{\Delta }t)\right].`$
Substituting in (20) gives,
$`P(s,t)`$ $`=`$ $`m\dot{q}\left(t+{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)\delta \left(s{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)m\ddot{q}(s+t)\theta \left({\displaystyle \frac{\mathrm{\Delta }t}{2}}s\right)`$ (37)
$`{\displaystyle \frac{1}{2}}m\omega ^2\left[{\displaystyle \frac{}{}}q(s+t+\mathrm{\Delta }t)\theta (s)+q(s+t\mathrm{\Delta }t)\theta (\mathrm{\Delta }ts)\right].`$
Note that $`P(\mathrm{\Delta }t,t)=0`$ and that
$$P(0,t)=m\left[\ddot{q}(t)+\frac{1}{2}\omega ^2q(t+\mathrm{\Delta }t)+\frac{1}{2}\omega ^2q(t\mathrm{\Delta }t)\right],$$
(38)
indeed vanishes with the equation of motion.
The canonical Hamiltonian is
$$H(t)=_0^{\mathrm{\Delta }t}𝑑sP(s,t)\frac{dQ}{ds}(s,t)\frac{1}{2}m\left[\frac{dQ}{ds}(\frac{\mathrm{\Delta }t}{2},t)\right]^2+\frac{1}{2}m\omega ^2Q(0,t)Q(\mathrm{\Delta }t,t).$$
(39)
The canonical evolution equations are,
$`{\displaystyle \frac{dQ}{dt}}(s,t)`$ $`=`$ $`{\displaystyle \frac{dQ}{ds}}(s,t),`$ (40)
$`{\displaystyle \frac{dP}{dt}}(s,t)`$ $`=`$ $`{\displaystyle \frac{dP}{ds}}(s,t)m{\displaystyle \frac{dQ}{ds}}({\displaystyle \frac{\mathrm{\Delta }t}{2}},t)\delta ^{}\left(s{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)`$ (41)
$`{\displaystyle \frac{1}{2}}m\omega ^2\left[Q(\mathrm{\Delta }t,t)\delta (s)+Q(0,t)\delta (s\mathrm{\Delta }t)\right].`$
It is simple to check that substituting $`Q(s,t)=q(s+t)`$ and relation (37) for $`P(s,t)`$ indeed verifies these equations.
The constraints are $`P(0,t)0`$, $`P(\mathrm{\Delta }t,t)0`$ and the apparently singular pair,
$`{\displaystyle \frac{dP}{ds}}(0,t){\displaystyle \frac{1}{2}}m\omega ^2Q(\mathrm{\Delta }t,t)\delta (0)`$ $``$ $`0,`$ (42)
$`{\displaystyle \frac{dP}{ds}}(\mathrm{\Delta }t,t){\displaystyle \frac{1}{2}}m\omega ^2Q(0,t)\delta (0)`$ $``$ $`0.`$ (43)
However, the vanishing of $`P(s,t)`$ at the endpoints means that the endpoint derivatives contain delta functions, so the actual constraints are the perfectly regular coefficients of $`\delta (0)`$,
$`P(0^+,t){\displaystyle \frac{1}{2}}m\omega ^2Q(\mathrm{\Delta }t,t)`$ $``$ $`0,`$ (44)
$`P(\mathrm{\Delta }t^{},t){\displaystyle \frac{1}{2}}m\omega ^2Q(0,t)`$ $``$ $`0.`$ (45)
Note that these constraints are implied by (37) and, where necessary, the vanishing of (38). Note also that the contstraints determine both the actual endpoint values of $`P(s,t)`$ and its limit as the endpoints are approached.
Since the Lagrangian (1) has no explicit dependence upon time the Hamiltonian should be conserved. To see that it is, first substitute $`Q(s,t)=q(s,t)`$ and relation (37) for $`P(s,t)`$ to obtain,
$`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑sP(s,t){\displaystyle \frac{dQ}{dq}}(s,t)=m\dot{q}^2\left(t+{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)m{\displaystyle _0^{\mathrm{\Delta }t/2}}𝑑s\dot{q}(s+t)\ddot{q}(s+t)`$ (47)
$`{\displaystyle \frac{1}{2}}m\omega ^2{\displaystyle _0^{\mathrm{\Delta }t}}𝑑sq(s+t\mathrm{\Delta }t)\dot{q}(s+t),`$
$`={\displaystyle \frac{1}{2}}m\left[\dot{q}^2\left(t+{\displaystyle \frac{\mathrm{\Delta }t}{2}}\right)+\dot{q}^2(t)\right]{\displaystyle \frac{1}{2}}m\omega ^2{\displaystyle _0^{\mathrm{\Delta }t}}𝑑sq(s+t\mathrm{\Delta }t)\dot{q}(s+t).`$
Then subtract expression (1) to determine the configuration space Hamiltonian,
$$H(t)=\frac{1}{2}m\dot{q}^2(t)+\frac{1}{2}m\omega ^2q(t)q(t+\mathrm{\Delta }t)\frac{1}{2}m\omega ^2_0^{\mathrm{\Delta }t}𝑑sq(s+t\mathrm{\Delta }t)\dot{q}(s+t).$$
(48)
Now use the fact that the integrand depends upon $`t`$ only through the sum $`s+t`$ to express the derivative of the integral as a surface term,
$`{\displaystyle \frac{dH}{dt}}(t)`$ $`=`$ $`m\dot{q}(t)\ddot{q}(t)+{\displaystyle \frac{1}{2}}m\omega ^2\left[\dot{q}(t)q(t+\mathrm{\Delta }t)+q(t)\dot{q}(t+\mathrm{\Delta }t)\right]`$ (49)
$`{\displaystyle \frac{1}{2}}m\omega ^2q(s+t\mathrm{\Delta }t)\dot{q}(s+t)|_{s=0}^{s=\mathrm{\Delta }t},`$
$`=`$ $`m\dot{q}(t)\left[\ddot{q}(t)+{\displaystyle \frac{1}{2}}\omega ^2q(t\mathrm{\Delta }t)+{\displaystyle \frac{1}{2}}\omega ^2q(t+\mathrm{\Delta }t)\right].`$ (50)
The most straightforward way of demonstrating that the transformation to the constrained phase space is invertible is by exhibiting the inverse. Of course we always have,
$$q(s+t)=Q(s,t),\mathrm{\hspace{0.33em}0}s\mathrm{\Delta }t.$$
(51)
For $`\mathrm{\Delta }t<s<0`$ one recovers $`q(s+t)`$ from relation (37),
$$q(s+t)=\frac{2}{m\omega ^2}P(s+\mathrm{\Delta }t,t)\frac{2}{\omega ^2}\frac{d}{ds}\left[\frac{dQ}{ds}(s+\mathrm{\Delta }t,t)\theta \left(\frac{\mathrm{\Delta }t}{2}s\right)\right].$$
(52)
The endpoint case of $`s=\mathrm{\Delta }t`$ is given by the constraint $`P(0,t)0`$,
$$q(t\mathrm{\Delta }t)=Q(\mathrm{\Delta }t,t)\frac{2}{\omega ^2}\frac{d^2Q}{ds^2}(0,t),$$
(53)
where I am of course defining differentiation in the right-handed sense,
$$\frac{df}{dx}(x)\underset{ϵ0^+}{lim}\frac{f(x+ϵ)f(x)}{ϵ}.$$
(54)
It is amusing to close the section by exhibiting the run-away solutions which are one possible consequence of the Ostrogradskian instability. Since the configuration space equation of motion,
$$\ddot{q}(t)+\frac{1}{2}\omega ^2[q(t+\mathrm{\Delta }t)+q(t\mathrm{\Delta }t)]=0,$$
(55)
is linear and invariant under time translation, the general solution must be a superposition of terms having the form $`e^{ikt}`$. The allowed frequencies are complex numbers $`k`$ which obey,
$$k^2=\omega ^2\mathrm{cos}(k\mathrm{\Delta }t).$$
(56)
The equation is transcendental but graphing both sides shows a single pair of $`\pm `$ real solutions. To find the remaining solutions make the substitution,
$$k=\alpha +i\beta ,$$
(57)
and take the real and imaginary parts of the equation,
$`\alpha ^2\beta ^2`$ $`=`$ $`\omega ^2\mathrm{cos}(\alpha \mathrm{\Delta }t)\mathrm{cosh}(\beta \mathrm{\Delta }t),`$ (58)
$`2\alpha \beta `$ $`=`$ $`\omega ^2\mathrm{sin}(\alpha \mathrm{\Delta }t)\mathrm{sinh}(\beta \mathrm{\Delta }t).`$ (59)
Graphical analysis indicates a conjugate pair of solutions for $`\alpha \mathrm{\Delta }t`$ in each $`2\pi `$ interval of the real line. For large integer $`N`$ these solutions have the form,
$`\alpha \mathrm{\Delta }t`$ $``$ $`2\pi N{\displaystyle \frac{2\mathrm{ln}(N)}{\pi N}},`$ (60)
$`\pm \beta \mathrm{\Delta }t`$ $``$ $`\mathrm{ln}\left({\displaystyle \frac{8\pi ^2N^2}{\omega ^2\mathrm{\Delta }t^2}}\right)+\left({\displaystyle \frac{\mathrm{ln}(N)}{\pi N}}\right)^2.`$ (61)
So this system has the infinite number of solutions predicted by the Ostrogradskian analysis, and all but two them grow or fall exponentially.
## 5 Ostrogradskian derivation
My representation is related to the infinite $`N`$ limit of Ostrogradski’s through the Maclaurin series,
$$Q(s,t)=\underset{I=0}{\overset{\mathrm{}}{}}\frac{s^I}{I!}Q_{I+1}(t).$$
(62)
Note that differentiation with respect to the Ostrogradskian coordinates is realized by the functional chain rule,
$$\frac{}{Q_I(t)}=_0^{\mathrm{\Delta }t}𝑑s\left[\frac{Q(s,t)}{Q_I(t)}\right]\frac{\delta }{\delta Q(s,t)}=_0^{\mathrm{\Delta }t}𝑑s\frac{s^{I1}}{(I1)!}\frac{\delta }{\delta Q(s,t)},$$
(63)
where the functional derivative is defined by
$$\frac{\delta Q(r,t)}{\delta Q(s,t)}=\delta (rs),$$
(64)
and the ordinary rules of calculus. From (19) one obtains a useful formula for the higher derivative representation,
$$\frac{L[q](t)}{q^{(I)}(t)}=_0^{\mathrm{\Delta }t}𝑑s\frac{s^I}{I!}\frac{\delta L[q](t)}{\delta q(s+t)}.$$
(65)
The conjugate momentum $`P(s,t)`$ should depend linearly on the Ostrogradskian momenta,
$$P(s,t)=\underset{I=0}{\overset{\mathrm{}}{}}p_I(s)P_{I+1}(t).$$
(66)
The combination coefficients $`p_I(s)`$ can be determined by enforcing the canonical Poisson bracket (23),
$`\delta (rs)`$ $`=`$ $`{\displaystyle \underset{I=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{r^I}{I!}}{\displaystyle \underset{J=0}{\overset{\mathrm{}}{}}}p_J(s)\{Q_{I+1}(t),P_{J+1}(t)\},`$ (67)
$`=`$ $`{\displaystyle \underset{I=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{r^I}{I!}}p_I(s).`$ (68)
By acting $`(/r)^J`$ and then taking $`r0`$ one finds,
$$p_J(s)=\left(\frac{d}{ds}\right)^J\delta (s).$$
(69)
To obtain my formula (20) for the conjugate momenta note first that, for infinite $`N`$, the Ostrogradskian momenta are,
$`P_I(t)`$ $`=`$ $`{\displaystyle \underset{J=I}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{d}{dt}}\right)^{JI}{\displaystyle \frac{L[q](t)}{q^{(J)}(t)}},`$ (70)
$`=`$ $`{\displaystyle \underset{J=I}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{d}{dt}}\right)^{JI}{\displaystyle _0^{\mathrm{\Delta }t}}𝑑r{\displaystyle \frac{r^J}{J!}}{\displaystyle \frac{\delta L[q](t)}{\delta q(r+t)}}.`$ (71)
Now substitute this and (69) into (66),
$$P(s,t)=\underset{I=0}{\overset{\mathrm{}}{}}\left[\left(\frac{d}{ds}\right)^I\delta (s)\right]\underset{J=I+1}{\overset{\mathrm{}}{}}\left(\frac{d}{dt}\right)^{JI1}_0^{\mathrm{\Delta }t}𝑑r\frac{r^J}{J!}\frac{\delta L[q](t)}{\delta q(r+t)}.$$
(72)
Simplification is achieved by exploiting the identity,
$$\frac{r^J}{J!}=_0^r𝑑r^{}\frac{(rr^{})^I(r^{})^{JI1}}{I!(JI1)!},$$
(73)
to recognize the two sums as Taylor expansions of the shift operator,
$`P(s,t)`$ $`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑r{\displaystyle _0^r}𝑑r^{}{\displaystyle \underset{I=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(rr^{})^I}{I!}}\left({\displaystyle \frac{d}{ds}}\right)^I\delta (s)`$ (74)
$`\times {\displaystyle \underset{J=I+1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(r^{})^{JI1}}{(JI1)!}}({\displaystyle \frac{d}{dt}})^{JI1}{\displaystyle \frac{\delta L[q](t)}{\delta q(r+t)}},`$
$`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑r{\displaystyle _0^r}𝑑r^{}\delta (sr+r^{}){\displaystyle \frac{\delta L[q](tr^{})}{\delta q(r+tr^{})}},`$ (75)
$`=`$ $`{\displaystyle _s^{\mathrm{\Delta }t}}𝑑r{\displaystyle \frac{\delta L[q](s+tr)}{\delta q(s+t)}}.`$ (76)
The Hamiltonian follows similarly,
$`H(t)`$ $`=`$ $`{\displaystyle \underset{I=1}{\overset{\mathrm{}}{}}}P_I(t)Q_{I+1}(t)L[Q](t),`$ (77)
$`=`$ $`{\displaystyle \underset{I=1}{\overset{\mathrm{}}{}}}{\displaystyle _0^{\mathrm{\Delta }t}}𝑑s{\displaystyle \frac{s^{I1}}{(I1)!}}P(s,t)\left({\displaystyle \frac{d}{dr}}\right)^IQ(r,t)|{\displaystyle \frac{}{}}_{r=0}L[Q](t),`$ (78)
$`=`$ $`{\displaystyle _0^{\mathrm{\Delta }t}}𝑑sP(s,t){\displaystyle \frac{d}{ds}}Q(s,t)L[Q](t).`$ (79)
Its instability is manifest from the fact that it has been derived from Ostrogradski’s result in the limit that the number of derivatives becomes infinite.
## 6 Discussion
I have shown that Lagrangians with nonlocality of finite extent $`\mathrm{\Delta }t`$ can be treated as the limits of higher derivative Lagrangians. I have also given a canonical formalism that is somewhat more natural in which the canonical variables are labelled by a continuum parameter $`s`$, for $`0s\mathrm{\Delta }t`$. The canonical coordinates are just the dynamical variables at times $`t+s`$. A quantum mechanical state in such a system would be a functional of these coordinates. The conjugate momenta (20) are given by a simple integral of a functional derivative of the Lagrangian. With the canonical coordinates, the momenta allow one to reconstruct the dynamical variables at times $`ts`$.
There is no physical motivation for this exercise because all such models are virulently unstable. Indeed, the only point of the formalism is to remove any doubt about a possible phenomenological role for these Lagrangians. They have inherited the full Ostrogradskian instability: essentially half of the directions in the classical phase space access arbitarily negative energies. There is not even any barrier to decay. This is a nonperturbative result and, because it arises from a large region of phase space, it must survive quantization.
Negative results of such power and generality are to mathematically inclined physicists like a red flag is supposed to be to a bull. Nothing I can honestly add is likely to much discourage further attempts to carve out a physical niche for nonlocal Lagrangians but I do recommend that these efforts be preceded by sober reflection upon the following fact: in the long struggle of our species to understand the universe it has never once proven useful to invoke a theory which is nonlocal on the most fundamental level. Yet the subset of local Lagrangians containing no more than first derivatives is a minuscule fragment of the set of all functionals of the dynamical variable. The Ostrogradskian instability offers a simple and compelling explanation for the complete dominance of this tiny subset over its much larger whole. The only alternative would seem to be coincidence on a scale that makes even the worst fine tuning problem seem inconsequential.
Note Added: Shortly after posting this paper I learned of important work by Llosa and Vives on the problem of canonically formulating a general nonlocal Lagrangian. My work can be viewed as a specialization of their technique to the case of nonlocality of finite extent where the Euler-Lagrange equations are deterministic, where an explicit Poisson bracket structure can be determined and where the formalism can be derived from the infinite $`N`$ limit of Ostrogradski’s construction. (None of these features can be present in the general case.) Note should also be taken of the recent work of Gomis, Kamimura and Llosa on canonically formulating spacetime noncommutative theories .
Acknowledgments
I thank T. Jacobson for asking the question which stimulated me to carry out this exercise. This work was partially supported by DOE contract DE-FG02-97ER41029 and by the Institute for Fundamental Theory.
|
warning/0006/math0006223.html
|
ar5iv
|
text
|
# References
Arithmetic structure of CMSZ fake projective planes <sup>1</sup><sup>1</sup>1$`1991`$ Mathematics Subject Classification. Primary 14G35; Secondary 14J29, 11G15, 11Q25.
Fumiharu Kato and Hiroyuki Ochiai
We prove that the fake projective planes constructed from the diadic discrete group discovered by Cartwright, Mantero, Steger, and Zappa are connected Shimura varieties associated to a certain unitary group. The necessary Shimura data, as well as the unitary group, are explicitly described. We also give a field of definition of these fake projective planes.
1. Introduction
In \[CMSZ2\], Cartwright, Mantero, Steger, and Zappa discovered a unitary group in three variables with respect to the quadratic extension $`(\sqrt{15})/`$ whose integral model over the integer ring with the prime $`2`$ inverted gives rise to a diadic discrete group acting transitively on vertices of Bruhat-Tits building over $`_2`$. Inside the integral model are three subgroups to which the restricted action is simply transitive. Moreover, a slight inspection shows that two of them act freely also on simplicies of the other dimensions. Such a situation evokes that Mumford \[Mu79\] had also obtained a discrete group with the same properties (but in a different unitary group) to construct an algebraic surface with $`P_\mathrm{g}=q=0`$, $`c_1^2=3c_2=9`$, and with the ample canonical class, so-called, fake projective planes, which is among the most interesting classes of algebraic surfaces. Like that Mumford’s group produces such a surface through diadic uniformization, these two groups give rise to two fake projective planes. It has been shown in \[IK98\] that these three fake projective planes are not isomorphic to among others.
In \[Ka99\], in the mean time, it was proved that Mumford’s fake projective plane is a unitary Shimura variety. A clue to this result is that Mumford’s discrete group is, as it is clear from the construction, a congruence subgroup in the integral model of the unitary group, as well as that it is arithmetic. However, it is not clear whether the groups in \[CMSZ2\] are characterized by congruence conditions or not, because the definition of the groups therein only gives us generators in matrices, although they are certainly arithmetic. This obscurity should be resolved, for it is the key point to settle whether the other two fake projective planes also have such a nice arithmetic structure as Mumford’s one has.
This point is exactly what we will discuss in this paper. In the next section we will construct congruence subgroups in the unitary group characterized by conditions in (modulo $`3`$)-reduction, and will prove that they coincide up to scalars with the groups defined in \[CMSZ2\]. Properly speaking, we give another way of construction of these groups, and hence the main stream of our argument is independent from that of \[CMSZ2\], although we have been much motivated by it. From this we proceed to construct the Shimura varieties in the following section, which mimics the argument in \[Ka99\]. Our main theorem (Theorem 1) states that, for each of the two groups, there exists a Shimura variety with the reflex field $`(\sqrt{15})`$ having two geometrically connected components isomorphic to the corresponding fake projective plane; moreover, we will also see that these connected components have $`(\sqrt{3},\sqrt{5})`$ as a field of definition.
The last section is an appendix to the next section, in which we will concentrate on proving a technical result.
Notation and conventions. The notation $`AB`$ (absence of the base ring) only occurs when $`A`$ and $`B`$ are $``$-algebra, and the tensor is taken over $``$. For a field extension $`F/E`$, we denote by $`\mathrm{Res}_{F/E}`$ the Weil restriction. By an involution of a ring $`A`$ we always mean a homomorphism $`:AA^{\mathrm{op}}`$ such that $`=\mathrm{id}_A`$.
2. The CMSZ-group
2.1. Discrete group from unitary group. Let $`K`$ be an imaginary quadratic extension of $``$, and $`p`$ a rational prime subject to the prime decomposition of form $`(p)=𝔭\overline{𝔭}`$ in $`K`$ . We consider the unitary group $`𝐈=𝐈_Q`$ associated to a positive definite Hermitian matrix $`Q`$ with entries in $`K`$; the group $`𝐈`$ is the $``$-algebraic group such that
$$𝐈()=\{g\mathrm{GL}_nK|g^{}Qg=c(g)Q,c(g)^\times \}.$$
Let $``$ be the integral model of $`𝐈()`$ consisting of matrices $`Q`$-unitary similitudes lying in $`\mathrm{GL}_n𝒪_K[\frac{1}{p}]`$; viz.,
$$=𝐈()\mathrm{GL}_n𝒪_K[\frac{1}{p}].$$
Since $`p`$ decomposes on $`K`$, the group $`𝐈(_p)`$ is, once we choose $`K_p`$ continuous with respect to the $`𝔭`$-adic topology on $`K`$, identified with $`\mathrm{GL}_n_p`$, and thus we have a homomorphism $`𝐈()\mathrm{PGL}_n_p`$. Let $`\mathrm{\Gamma }`$ be the image of $``$ in $`\mathrm{PGL}_n_p`$. Since $`𝐈_{}\mathrm{GU}(3)`$, as is well-known, $`\mathrm{\Gamma }`$ is discrete and co-compact in $`\mathrm{PGL}_n_p`$.
2.2. CMSZ-situation. The so-called CMSZ-group is the group $``$ (or $`\mathrm{\Gamma }`$) as above in $`n=3`$, $`p=2`$, and the following setting: First we set $`K=(\sqrt{15})`$. The integer ring $`𝒪_K`$ is $`[\lambda ]`$ with $`\lambda =(1\sqrt{15})/2`$; the prime $`2`$ is decomposed as $`(2)=𝔭\overline{𝔭}`$ where $`𝔭=2+\lambda `$. The Hermitian matrix $`Q`$ is the one given by
$$Q=\left[\begin{array}{ccc}10& 2\left(\lambda +2\right)& \lambda +2\\ 2\left(\overline{\lambda }+2\right)& 10& 2\left(\lambda +2\right)\\ \overline{\lambda }+2& 2\left(\overline{\lambda }+2\right)& 10\end{array}\right].$$
The embedding $`K_2`$ is chosen to be the $`𝔭`$-adic completion; hence, $`\lambda /2`$ gives a uniformizing prime, and $`\overline{\lambda }=1\lambda `$ is a diadic unit. The significance of the group $`\mathrm{\Gamma }`$ thus obtained is the following fact:
2.3. Theorem (\[CMSZ2, §5\]). The group $`\mathrm{\Gamma }`$ acts transitively on the vertices of the Bruhat-Tits building $`\mathrm{\Delta }`$ attached to $`\mathrm{PGL}_3_2`$. $`\mathrm{}`$
We insert the proof herein for the reader’s convenience. Before doing it, we introduce two elements belonging to $``$:
$$\rho =\left[\begin{array}{ccc}0& 0& \frac{\lambda }{2}\\ 0& 1& 1+\frac{\lambda }{2}\\ 1& 1& 1\end{array}\right],\tau =\left[\begin{array}{ccc}0& 1& \frac{\lambda }{2}\\ 1& 1& 1+\frac{\lambda }{2}\\ 0& 0& 1\end{array}\right].$$
Note that $`\rho ^3=\frac{\lambda }{2}I_3`$ and $`\tau ^3=I_3`$.
Proof. Let $`\mathrm{\Lambda }_0`$ be the vertex which is the similarity class of the standard lattice $`(_2)^3`$. Since $`\mathrm{\Delta }`$ is a connected complex, it suffices to show that, for each vertex $`\mathrm{\Lambda }`$ adjacent to $`\mathrm{\Lambda }_0`$, there exists an element $`g`$ such that $`\mathrm{\Lambda }=g\mathrm{\Lambda }_0`$. There exist elements $`g_i`$ for $`i/7`$ with $`g_3=\rho `$ satisfying the following relations:
(1.1)
$$\tau ^1g_{2i}\tau =g_i,$$
(1.2)
$$\begin{array}{ccccccc}g_3g_3g_3\hfill & =& g_6g_6g_6\hfill & =& g_5g_5g_5\hfill & =& \hfill \frac{\lambda }{2}I_3\text{,}\\ g_1g_1g_0\hfill & =& g_2g_2g_0\hfill & =& g_4g_4g_0\hfill & =& \hfill I_3\text{,}\\ g_1g_3g_6\hfill & =& g_2g_6g_5\hfill & =& g_4g_5g_3\hfill & =& \hfill I_3\text{.}\end{array}$$
Then one can easily show that the set $`\{g_i^{\pm 1}\mathrm{\Lambda }_0|i/7\}`$ coincides with the set of vertices adjacent to $`\mathrm{\Lambda }_0`$. $`\mathrm{}`$
2.4. Remark (\[CMSZ1\]\[CMSZ2\]). Let $``$ be the subset of $`(/7)^3`$ given by
={(i,j,k)(/7)3|
Either gigjgk, gjgkgi, or gkgigj appears in
(1.2).}.conditional-set𝑖𝑗𝑘superscript73
Either gigjgk, gjgkgi, or gkgigj appears in
(1.2).\mathcal{F}=\left\{(i,j,k)\in(\mathbb{Z}/7\mathbb{Z})^{3}\,\bigg{|}\,\begin{minipage}{128.0374pt}
\small
Either $g_{i}g_{j}g_{k}$, $g_{j}g_{k}g_{i}$, or $g_{k}g_{i}g_{j}$ appears in
(\ref{thm-CMSZ}.2).
\end{minipage}\right\}.
Then $``$ consists of $`21=3+6\times 3`$ elements. Note that $`g_ig_jg_k`$ is either $`I_3`$ or $`\frac{\lambda }{2}I_3`$ for each $`(i,j,k)`$. This set has the following meaning: For each $`(i,j,k)`$, the set $`\{g_i\mathrm{\Lambda }_0,\mathrm{\Lambda }_0,g_k^1\mathrm{\Lambda }_0\}`$ forms a chamber in $`\mathrm{\Delta }`$, which we denote by $`C(i,j,k)`$; this gives rise to a bijection between $``$ and the set of all (non-oriented) chambers containing $`\mathrm{\Lambda }_0`$. The following facts are easy to see:
(1.1) $`g_i^1C(i,j,k)=C(j,k,i)`$; in particular, $`g_3`$ (resp. $`g_6`$, $`g_5`$) stabilizes the chamber $`C(3,3,3)`$ (resp. $`C(6,6,6)`$, $`C(5,5,5)`$).
(1.2) $`\tau C(i,j,k)=C(2i,2j,2k)`$ (this follows from (1.1) and the fact that $`\tau `$ fixes $`\mathrm{\Lambda }_0`$).
Using these facts, one can classify all the finite subgroups in $``$ (Theorem 1 below).
2.5. Proposition. Let $``$ act on $`\mathrm{\Delta }`$ through $`\mathrm{PGL}_3_2`$. Then the stabilizer $`\mathrm{Stab}\mathrm{\Lambda }_0`$ in $``$ of the vertex $`\mathrm{\Lambda }_0`$ is the subgroup generated by $`\tau `$ and scalar matrices. In particular, the group $``$ is generated by $`\rho `$, $`\tau `$, and scalars.
Proof. As the proof of Theorem 1 indicates, the subgroup of $``$ generated by $`\rho `$ and $`\tau `$ acts transitively on the vertices of $`\mathrm{\Delta }`$. Hence the second assertion follows from the first. Let $`g`$ be an element which fixes $`\mathrm{\Lambda }_0=[(_2)^3]`$. Multiplying by a scalar if necessary, one may assume $`g(_2)^3=(_2)^3`$. There, what we shall prove is:
2.6. Lemma. A matrix $`g`$ with $`g(_2)^3=(_2)^3`$ must be of form $`\pm \tau ^i`$ ($`i=0,1,2`$).
The proof is elementary, but technical, being partly carried out by computer search; we postpone it to Appendix. $`\mathrm{}`$
Recall that a labelling of $`\mathrm{\Delta }`$ is a map $`l`$ from the set of vertices to $`/3`$ which assigns $`\nu (detg)`$ mod $`3`$ to $`\mathrm{\Lambda }=[g(_2)^3]`$ with $`g\mathrm{PGL}_3_2`$, where $`\nu `$ is the normalized valuation $`\nu :_2^\times `$. The map $`l`$ restricted to each chamber is bijective, and hence $`l`$ gives rise to an orientation in $`\mathrm{\Delta }`$. Since $`l(g\mathrm{\Lambda })=l(\mathrm{\Lambda })+(\nu (detg)\text{mod}3)`$ for any $`g\mathrm{PGL}_3_2`$ and any $`\mathrm{\Lambda }`$, the action of $`\mathrm{PGL}_3_2`$ on $`\mathrm{\Delta }`$ preserves the orientation.
2.7. Proposition. Let us consider, for $`(i,j,k)`$, the stabilizer $`\mathrm{Stab}C(i,j,k)`$ in $``$ of the chamber $`C(i,j,k)`$. Then, if $`i=j=k=3`$, $`6`$, or $`5`$, $`\mathrm{Stab}C(i,j,k)`$ is the subgroup generated by $`g_i`$ and scalar matrices; otherwise, it consists only of scalars.
Proof. In view of (1.1) and (1.2), it suffices to show the proposition in the cases $`(i,j,k)=(3,3,3)`$, $`(1,1,0)`$, and $`(1,3,6)`$. Let $`g`$ stabilizes $`C(i,j,k)`$, and assume that $`g`$ is not a scalar. By (1.2) and Proposition 1, $`g`$ does not fix the vertex $`\mathrm{\Lambda }_0`$, but $`g^3`$ does, since $`g`$ preserves the orientation by the labelling. Then, taking inverse if necessary, one may assume $`g\mathrm{\Lambda }_0=g_i\mathrm{\Lambda }_0`$. By Proposition 1, we deduce $`g=g_i\tau ^jc`$ for some $`j\{0,1,2\}`$ and $`c(𝒪_K[\frac{1}{2}])^\times `$. Hence, from Proposition 1, it follows that $`(g_i\tau ^j)^9`$ must be a scalar. In case $`i=1`$, this can be shown to be impossible by direct calculation. If $`i=3`$, this is only the case when $`j=0`$ (hence $`g^3`$ is a scalar), as one can check directly. $`\mathrm{}`$
2.8. Theorem. Any non-trivial finite subgroup of $`\mathrm{\Gamma }`$ is conjugate to either $`\rho `$ or to $`\tau `$.
Proof. Let $`H`$ be a finite subgroup in $`\mathrm{\Gamma }`$. Then it is well-known that $`H`$ stabilizes a simplex in $`\mathrm{\Delta }`$; since $`\mathrm{\Gamma }`$ preserves the orientation, $`H`$ stabilizes either a vertex or a chamber. Since $`\mathrm{\Gamma }`$ acts on the vertices transitively, the result follows from Proposition 1 and Proposition 1. $`\mathrm{}`$
2.9. Reduction of unitary group. For an ideal $`𝔠`$ of $`𝒪_K[\frac{1}{2}]`$ which is stable under the complex conjugation and a non-negative integer $`n`$, the Artinian ring $`R_n=𝒪_K[\frac{1}{2}]/𝔠^{n+1}`$ has the induced involution, which we also denote by $`\alpha \overline{\alpha }`$. There is the obvious (modulo $`𝔠^{n+1}`$)-reduction morphism $`\pi _n:\mathrm{GL}_3R_n`$, whose image is a subgroup of the unitary group
$$U_n=\{g\mathrm{GL}_3R_n|g^{}Q_ng=c(g)Q_n,c(g)F_n^\times \},$$
where $`Q_n`$ is the (modulo $`𝔠^n`$)-reduction of $`Q`$ (note that our $`Q`$ has coefficients in $`𝒪_K[\frac{1}{2}]`$), and $`F_n`$ is the subring of $`R_n`$ consisting of elements fixed by the involution. Our particular interest is in the case $`𝔠=3+(\lambda +1)`$ and $`n=0`$ or $`1`$. Note that $`𝔠^2=(3)`$ gives the prime decomposition of $`3`$ on $`K`$. Our goal in this section is to construct subgroups in $``$ which acts freely on $`\mathrm{\Delta }`$ and transitively on the vertices in terms of congruence conditions in these reductions. To this end, we will first determine the subgroups $`\pi _i()`$ ($`i=0,1`$) and then, ask for nice subgroups of them of which the inverse images by $`\pi _i`$ keep the transitivity on vertices and rule out torsion elements.
Before carrying out this program, we change our presentation of matrices into more convenient form: Let us define a matrix
$$\mathrm{\Phi }=\left[\begin{array}{ccc}1& 2\lambda 1& 2\\ \frac{4\lambda 3}{2}& \frac{2\lambda +1}{2}& 1\\ \frac{2\lambda 1}{2}& \frac{1}{2}& 2\end{array}\right],$$
and the new Hermitian matrix $`Q^{}=\mathrm{\Phi }^{}Q\mathrm{\Phi }`$, which appears to be:
$$Q^{}=\left[\begin{array}{ccc}90& 2\overline{\lambda }1& 15\\ 2\lambda 1& 90& 15\left(2\lambda 1\right)\\ 15& 15\left(2\overline{\lambda }1\right)& 70\end{array}\right].$$
We set $`^{}=\mathrm{\Phi }^1\mathrm{\Phi }`$. (Note that, since $`det\mathrm{\Phi }=2^27`$, this twist is not admissible over $`𝒪_K[\frac{1}{2}]`$; however, it is harmless because we work only over the prime above $`3`$.) Set $`U_n^{}=\mathrm{\Phi }^1U_n\mathrm{\Phi }`$, and let $`\pi _n^{}:U_n^{}`$ be the composite of the reduction map $`\pi _n`$ followed by the isomorphism $`U_n\stackrel{}{}U_n^{}`$.
2.10. The unitary groups $`U_0^{}`$ and $`U_1^{}`$. First we note that $`R_0=F_0𝔽_3`$ and that the induced involution acts on it trivially. Since
$$Q_0^{}=\left[\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ 0& 0& 1\end{array}\right],$$
the group $`U_0^{}`$ modulo center is isomorphic to the Euclidean motion group on the affine plane over $`𝔽_3`$; more precisely,
(1.1)
$$U_0^{}=\left\{\left[\begin{array}{cc}A& B\\ 0& \pm 1\end{array}\right]\right|A\mathrm{SL}_2𝔽_3,B\mathrm{M}_{21}𝔽_3\}.$$
The order of $`U_0^{}`$ is therefore $`432=2^43^3`$.
The Artinian ring $`R_1`$ in turn is isomorphic to $`𝔽_3[t]/(t^2)`$, where $`t`$ is the image of $`12\lambda `$ ($`=\sqrt{15}`$), which is acted on by the involution $`tt`$; hence, $`F_1𝔽_3`$. The group $`U_1^{}`$ is the unitary group with respect to the Hermitian form
$$Q_1^{}=\left[\begin{array}{ccc}& t& \\ t& & \\ & & 1\end{array}\right],$$
whence
(1.2)
$$U_1^{}=\left\{\left[\begin{array}{cc}A& B\\ tC& d\end{array}\right]\right|A\mathrm{GL}_2R_1\text{}B\mathrm{M}_{21}R_1\text{ and }dR_1^\times \text{ with }(A\text{ mod }t)\mathrm{SL}_2𝔽_3\text{ and }{}_{}{}^{t}C=(d^1JA^1B\text{ mod }t)\text{ }\},$$
where $`J`$ denotes the standard symplectic matrix. The order of this group is calculated to be $`944784=2^43^{10}`$.
2.11. Special elements. Here we introduce some matrices in $`U_1^{}`$ which will play important roles in analyzing the structure of $`U_1^{}`$:
$$z=\left[\begin{array}{ccc}1+t& & \\ & 1+t& \\ & & 1+t\end{array}\right],u=\left[\begin{array}{ccc}1& 1& \\ & 1& \\ & & 1t\end{array}\right],w=\left[\begin{array}{ccc}& 1& \\ 1& & \\ & & 1\end{array}\right],$$
$$b_1=\left[\begin{array}{ccc}1& & t\\ & 1& \\ & & 1\end{array}\right],b_2=\left[\begin{array}{ccc}1& & \\ & 1& t\\ & & 1\end{array}\right],c_1=\left[\begin{array}{ccc}1& t& 1\\ & 1& \\ & t& 1\end{array}\right],c_2=\left[\begin{array}{ccc}1& & \\ t& 1& 1\\ t& & 1\end{array}\right],$$
$$d_1=\left[\begin{array}{ccc}1+t& & \\ & 1& \\ & & 1\end{array}\right],d_2=\left[\begin{array}{ccc}1& & \\ & 1+t& \\ & & 1\end{array}\right],d_3=\left[\begin{array}{ccc}1& t& \\ & 1& \\ & & 1\end{array}\right],d_4=\left[\begin{array}{ccc}1& & \\ t& 1& \\ & & 1\end{array}\right].$$
We set
$$T:=b_1,b_2,H:=z,c_1,c_2,M:=d_1,d_2,d_3,d_4,S:=u,w.$$
2.12. Notation. We hereafter make use of the following notation:
$`[\alpha ,\beta ]`$ $`=`$ $`\alpha ^1\beta ^1\alpha \beta ,`$
$`\alpha ^\delta `$ $`=`$ $`\delta ^1\alpha \delta .`$
2.13. Relations among special elements. Here we list up the relations among those elements defined above:
(1.1) All but $`w`$ are of order $`3`$, while the order of $`w`$ is $`4`$.
(1.2) $`z`$ lies in the center of $`U_1^{}`$, and $`[c_1,c_2]=z`$. Therefore, the subgroup $`H`$ is the Heisenberg group with the center $`z`$ and $`H/z𝔽_3^2`$.
(1.3) $`T`$ and $`M`$ are vector groups; $`T𝔽_3^2`$ and $`M𝔽_3^4`$. We moreover have $`[T,H]=[T,M]=1`$.
(1.4) In $`S`$ we have the relations $`[u,w^2]=1,wuw=u^1wu^1,wu^1w=uw^1u`$. Hence $`S`$ is isomorphic to $`\mathrm{SL}_2𝔽_3`$.
(1.5) $`T`$ is normalized by $`S`$; actually, we have $`b_1^u=b_1,b_2^u=b_1^1b_2,b_1^w=b_2^1,b_2^w=b_1`$.
(1.6) The action of $`S`$ on $`H`$ is given by $`c_1^u=b_1^1c_1,c_2^u=b_1^1c_1^1c_2,c_1^w=c_2^1,c_2^w=c_1`$. Hence, in particular, the $`T,H`$ is normalized by $`S`$.
(1.7) The action of $`H`$ on $`M`$ is by
$$\begin{array}{cccc}d_1^{c_1}=b_1d_1,\hfill & d_2^{c_1}=d_2,\hfill & d_3^{c_1}=d_3,\hfill & d_4^{c_1}=b_2d_4,\hfill \\ d_1^{c_2}=d_1,\hfill & d_2^{c_2}=b_2d_2,\hfill & d_3^{c_2}=b_1d_3,\hfill & d_4^{c_2}=d_4.\hfill \end{array}$$
(1.8) $`M`$ is normalized by $`S`$, since
$$\begin{array}{cccc}d_1^u=d_1d_3,\hfill & d_2^u=d_2d_3^1,\hfill & d_3^u=d_3,\hfill & d_4^u=d_1^1d_2d_3^1d_4,\hfill \\ d_1^w=d_2,\hfill & d_2^w=d_1,\hfill & d_3^w=d_4^1,\hfill & d_4^w=d_3^1.\hfill \end{array}$$
2.14. Proposition. The group $`U_1^{}`$ is generated by the subgroups $`T`$, $`H`$, $`M`$, $`S`$, and $`I_3`$. Moreover, there exists an isomorphism
$$U_1^{}\stackrel{}{}(((T\times M)H)S)\times 𝔽_3^\times \stackrel{}{}((𝔽_3^6H)\mathrm{SL}_2𝔽_3)\times 𝔽_3^\times .$$
Proof. Let $`U_1^+`$ be the subgroup generated by $`T`$, $`H`$, $`M`$, and $`S`$. By the appearance of the matrices in 1 and relations in 1, we easily see $`U_1^+((T\times M)H)S`$. But the order of the left-hand side already attains the half of that of $`U_1^{}`$, that is, $`2^33^{10}`$. Hence $`U_1^+`$ is precisely the kernel of the surjective homomorphism
(1.1)
$$(\text{det mod}t):U_1^{}𝔽_3^\times .$$
Clearly, this morphism has a cross-section with the image in the center of $`U_1^{}`$, thereby the assertion. $`\mathrm{}`$
2.15. Proposition. The image $`\pi _1^{}()`$ by the (modulo $`3`$)-reduction of $``$ is the subgroup of $`U_1^{}`$ generated by $`T`$, $`H`$, $`S`$, $`d_1d_2`$, and scalars (hence is of order $`2^43^7=34992`$). In particular, we have $`\pi _0^{}()=U_0^{}`$.
Proof. By Proposition 1, we know that $`\pi _1^{}()`$ is generated by $`(\mathrm{\Phi }^1\rho \mathrm{\Phi }\text{mod}3)`$, $`(\mathrm{\Phi }^1\tau \mathrm{\Phi }\text{mod}3)`$ (which we denote, in this proof, by $`\rho `$ and $`\tau `$, respectively, for simplicity), and scalars. As one checks easily,
(1.1)
$$\begin{array}{ccccc}\rho \hfill & =& \left[\begin{array}{ccc}0& 1+t& t\\ 1t& 1t& 1+t\\ 0& t& 1+t\end{array}\right]& =& b_1^1c_2wu^1(d_1d_2)^1,\hfill \\ \tau \hfill & =& \left[\begin{array}{ccc}1+t& 1t& 1+t\\ 0& 1+t& 0\\ 0& t& 1+t\end{array}\right]& =& z^1c_1u(d_1d_2)^1,\hfill \end{array}$$
and hence it follows that $`\pi _1^{}()T,H,S,d_1d_2`$. For the converse, one calculates
$$z=\rho ^3,w=\rho ^4(\tau \rho ^1\tau ^1\rho ^1)^2,$$
$$b_1=[\tau \rho \tau \rho ^2\tau \rho \tau ^1,\rho \tau \rho \tau ^1\rho \tau \rho \tau ],b_2=wb_1w^1,$$
$$c_1=\rho ^3(\tau ^1\rho \tau \rho \tau ^1)^{(\rho \tau \rho )^1}(\tau \rho )^2b_1b_2^1,c_2=wc_1w^1,$$
$$u=\tau \rho ^1\tau ^1\rho ^1\tau ^1b_1^1b_2c_2^1w^2,$$
whence $`\pi _1^{}()T,H,S,d_1d_2`$. $`\mathrm{}`$
Now, to find subgroups of $``$ as in 1, one has to first look for subgroups which does not contain $``$-conjugates of $`\tau `$ and $`\rho `$; moreover, the indices of such subgroups in $``$ are required to be $`3`$, since they have to act on the vertices of $`\mathrm{\Delta }`$ simply transitively. This leads us to the following kind of statement:
2.16. Lemma. Every subgroup of $`\pi _1^{}()`$ of index $`3`$ is conjugate to either one of the following subgroups:
$`𝒥_1`$ $`=`$ $`T,H,P,u,\text{scalars}`$
$`𝒥_2`$ $`=`$ $`T,H,P,d_1d_2,\text{scalars}`$
$`𝒥_3`$ $`=`$ $`T,H,P,ud_1d_2,\text{scalars}`$
$`𝒥_4`$ $`=`$ $`T,H,P,u(d_1d_2)^1,\text{scalars}\text{,}`$
where $`P`$ is the unique $`2`$-Sylow subgroup of $`S`$, generated by $`w`$ and $`w^u`$.
Proof. Let $`𝒥`$ be a subgroup of $`\pi _1^{}()`$ of index $`3`$. Since $`P,\text{scalars}P\times 𝔽_3^\times `$ is a $`2`$-Sylow subgroup of $`\pi _1^{}()`$, replacing it by a suitable conjugate, we may assume that $`𝒥`$ contains $`P,\text{scalars}`$. Let us consider the (modulo $`t`$)-reduction map $`\psi :\pi _1^{}()\pi _0^{}()`$, of which the kernel is $`K=T,z,d_1d_2`$. Since $`𝒥K`$ is a normal subgroup in $`𝒥`$, it is in particular a $`(P\times 𝔽_3^\times )`$-module by conjugation. Since the $`(P\times 𝔽_3^\times )`$-module $`K`$ is decomposed into irreducible factors as $`T\times z\times d_1d_2𝔽_3^2\times 𝔽_3\times 𝔽_3`$, we deduce that $`𝒥K`$ contains $`T`$;
(1.1)
$$T𝒥K𝒥.$$
Next we claim that
(1.2)
$$\psi (H)\psi (𝒥).$$
Consider the projection $`\pi _0^{}()\mathrm{SL}_2𝔽_3\times 𝔽_3^\times `$ which extracts the top-left $`2\times 2`$ component and the scalars. Since its kernel is $`\psi (H)`$ consisting of $`9`$ elements, and since $`\psi (𝒥)`$ is of index at most $`3`$ in $`\pi _0^{}()`$, we deduce $`\psi (𝒥)\psi (H)\{1\}`$. This intersection is, moreover, stable under $`P`$-conjugation. But, as one sees easily, the group $`\psi (H)`$ is an irreducible $`P`$-module, thereby the claim.
By (1.2), we know that, for each $`i=1,2`$, there exists $`k_iK`$ such that $`k_ic_i𝒥`$. Since $`K`$ is an abelian group, we deduce that $`z=[k_1,k_2][c_1,c_2]=[k_1c_1,k_2c_2]𝒥`$. Combining with (1.1), we get $`z,T𝒥K𝒥`$. By this and that $`K=T,z,d_1d_2`$, we may take $`k_i`$ from $`d_1d_2`$. Since $`[d_1d_2,w^2]=1`$, we get $`c_1=[c_1,w^2]=[k_1c_1,w^2]𝒥`$. Then we moreover have $`c_2=wc_1w^1𝒥`$. This implies that $`𝒥`$ contains also $`H`$, and hence:
(1.3)
$$T,H,P,\text{scalars}𝒥K𝒥.$$
The left-hand side of (1.3) is already a subgroup in $`\pi _1^{}()`$ of index $`9`$. Then it is now obvious that any subgroup of $`\pi _1^{}()`$ of index $`3`$ which contains it is one of $`𝒥_i`$ (i=1,2,3,4) as stated. $`\mathrm{}`$
2.17. Construction. Now, let us define the subgroup $`_i`$ ($`i=1,2,3,4`$) of $``$ by
$$_i=(\pi _1^{})^1(𝒥_i),$$
and let $`\mathrm{\Gamma }_i`$ be its image in $`\mathrm{PGL}_3_2`$. These are subgroups of index $`3`$. By (1.1) we know that the subgroup $`\mathrm{\Gamma }_3`$ contains $`\rho `$ and that $`\mathrm{\Gamma }_4`$ contains $`\tau `$; hence these are not torsion-free. The groups $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$, on the other hand, has no element of finite order except $`1`$, since, as one sees from the relations in 1, they does not have any conjugates of $`\tau `$ and $`\rho `$. Hence the discrete groups $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ act freely on $`\mathrm{\Delta }`$ and simply transitively on the vertices.
Let $`\mathrm{\Omega }`$ be the Drinfeld symmetric space over $`_2`$ of dimension $`2`$. By \[Mu79, 1\], we deduce that both $`\mathrm{\Gamma }_1\backslash \mathrm{\Omega }`$ and $`\mathrm{\Gamma }_2\backslash \mathrm{\Omega }`$ are, respectively, algebraized to fake projective planes $`{}_{}{}^{I}X_{\mathrm{CMSZ}}^{}`$ and $`{}_{}{}^{II}X_{\mathrm{CMSZ}}^{}`$ over $`_2`$, not isomorphic to each other, since our definition shows that $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ are not conjugate in $`\mathrm{PGL}_3_2`$ (cf. \[IK98\]).
2.18. Remark. One can check by direct calculation that our groups $`\mathrm{\Gamma }_i`$ for $`i=1,2,3`$ are exactly the $`\mathrm{\Phi }`$-conjugate of the groups $`\mathrm{\Gamma }_{B.i}`$ discovered by Cartwright, Mantero, Steger, and Zappa in their paper \[CMSZ2, pp. 182\]. Indeed, the elements $`\rho `$ and $`\tau `$ in 1 are exactly $`g_3s`$ and $`s`$ therein, respectively. The set $``$ defined in Remark 1, which was used to analyze the action of $``$ on the chambers of $`\mathrm{\Delta }`$, is nothing but the triangle presentation $`_{B.3}`$.
3. The Shimura varieties
3.1. Situation. In this section we regard the field $`K=(\lambda )`$ as a subfield of the 45th-cyclotomic field $`(\zeta _{45})`$ which is embedded in $``$ so that $`\zeta _{45}^{21}+\zeta _{45}^{33}+\zeta _{45}^{39}+\zeta _{45}^{42}`$ is mapped to $`\lambda `$ (cf. 1). Consider the intermediate field $`(\zeta _9)`$ with $`\zeta _9=\zeta _{45}^5`$, and let $`L`$ be the composite field of $`K`$ and the maximal real subfield of $`(\zeta _9)`$; $`L`$ is an extension of $`K`$ of degree $`3`$ generated $`K`$ by $`\zeta _9+\zeta _9^1`$. The Galois group $`\mathrm{Gal}(L/K)`$ is generated by $`\sigma :\zeta _9+\zeta _9^1\zeta _9^2+\zeta _9^2`$. We need to construct
$``$ a central division algebra $`D`$ over $`K`$,
$``$ a positive involution $``$ of $`D`$ of second kind,
$``$ a non-degenerate anti-Hermitian form $`\psi `$,
from which we will construct the associated Shimura variety. The construction of these data is basically parallel to that in \[Ka99, 2\], and here, we sometimes limit ourselves to sketchy presentation.
3.2. Division algebra with involutions. First we let
$$\mu =\lambda /\overline{\lambda }K.$$
The division algebra $`D`$ is defined by
$$D=\underset{i=0}{\overset{2}{}}L\mathrm{\Pi }^i;\mathrm{\Pi }^3=\mu ,\mathrm{\Pi }z=z^\sigma \mathrm{\Pi }\text{for }zL.$$
Then $`D_KL`$ is isomorphic to the matrix algebra $`M_3(L)`$; more explicitly, such an isomorphism is given by the morphism $`DM_3(L)`$ defined by
(1.1)
$$z\left[\begin{array}{ccc}z& & \\ & z^{\sigma ^2}& \\ & & z^\sigma \end{array}\right]\text{and}\mathrm{\Pi }\left[\begin{array}{ccc}& & \mu \\ 1& & \\ & 1& \end{array}\right].$$
By this, $`D`$ is regarded as a subalgebra over $`K`$ of $`M_3(L)`$. The Hasse invariants of $`D`$ are zero except for those at $`𝔭`$ and $`\overline{𝔭}`$, which are $`1/3`$ and $`1/3`$, respectively. The subring
$$𝒪_D=𝒪_L𝒪_L\overline{\lambda }\mathrm{\Pi }𝒪_L\overline{\lambda }\mathrm{\Pi }^2$$
is an order of $`D`$, maximal both at $`𝔭`$ and $`\overline{𝔭}`$. The involution $``$ on $`D`$ is defined by
$$\mathrm{\Pi }^{}=\overline{\mu }\mathrm{\Pi }^2\text{and}z^{}=\overline{z}\text{for}zL,$$
and the anti-Hermitian form $`\psi `$ by $`\psi (\alpha ,\beta )=\mathrm{tr}_{D/}(\alpha ^{}b\beta )`$, where
$$b=(\overline{\lambda }\lambda )\overline{\lambda }\mathrm{\Pi }+\overline{\lambda }\mathrm{\Pi }^2.$$
Since $`\mathrm{Nm}_{D|}^{}(b)=7(\overline{\lambda }\lambda )`$ is both $`𝔭`$-adic and $`\overline{𝔭}`$-adic unit, the pairing $`\psi `$ restricted to $`𝒪_D`$ is perfect. The element $`b`$ also induces another involution $`\mathrm{}`$ by $`\alpha ^{\mathrm{}}=b^1\alpha ^{}b`$.
3.3. Shimura variety. Let us consider the $``$-algebraic group $`𝐆`$ such that
$$𝐆()=\{\gamma D^\times |\gamma ^{\mathrm{}}\gamma ^\times \}.$$
By the identification (1.1) we easily see that $`𝐆_{}`$ is isomorphic to $`\mathrm{GU}(2,1)`$. The Shimura data is given by the homomorphism
$$h:\mathrm{Res}_/𝔾_{\mathrm{m}}^{}{}_{,}{}^{}𝐆_{}$$
which sends $`\sqrt{1}`$ to $`\mathrm{diag}(\sqrt{1},\sqrt{1},\sqrt{1})`$, and is identity on scalars. This gives rise to a Shimura variety
$$𝒮h_C=𝐆()\backslash X_{\mathrm{}}\times 𝐆(𝔸_\mathrm{f})/C,$$
where $`X_{\mathrm{}}`$ is the set of all conjugates of $`h`$, and $`C`$ is a (sufficently small) open compact subgroup of $`𝐆(𝔸_\mathrm{f})`$. This is a finite disjoint union of quasi-projective manifolds, which are arithmetic quotients of the complex unit-ball. The reflex field $`E`$ of $`𝒮h`$ is $`K`$.
3.4. Proposition. Let $`𝐈`$ and $`𝐆`$ be the $``$-algebraic groups defined in 1 and 1, respectively. Then we have the following:
(a) $`𝐈`$ is an inner form of $`𝐆`$.
(b) $`𝐈(𝔸_\mathrm{f}^2)𝐆(𝔸_\mathrm{f}^2)`$.
Moreover, the algebraic group $`𝐈`$ is uniquely determined up to $``$-isomorphisms by (a), (b) and the following condition:
(c) $`𝐈(_2)\{(x,y)\mathrm{GL}_3_2\times (\mathrm{GL}_3_2)^{\mathrm{op}}|xy_2\}`$.
Proof. The proof is, literally, almost the same as that of \[Ka99, 4.1\]; it differs only when we compare, in the proof of (b), two involutions induced from $`Q`$ and $`b`$. We should hereupon compare two Hermitian matrices $`3Q`$ and $`7b`$ (where $`b`$ is regarded as a matrix by (1.1)). The determinants of these are square numbers, thereby the assertion. $`\mathrm{}`$
3.5. CMSZ-level structures. We fix an isomorphism $`\varphi :𝐈(𝔸_\mathrm{f}^2)𝐆(𝔸_\mathrm{f}^2)`$. Let $`C_2^{\mathrm{max}}`$ be the maximal open compact subgroup of $`𝐆(_2)`$ consisting of elements which map $`𝒪_D_{}_2`$ into itself. For a finite set of rational primes $`W`$ with $`2W`$ (resp. a rational prime $`p2`$), we define $`C_{\mathrm{max}}^W`$ (resp. $`C_p^{\mathrm{max}}`$) to be the maximal open compact subgroup of $`𝐈(𝔸_\mathrm{f}^W)`$ (resp. $`𝐈(_p)`$) consisting of elements which map $`𝒪_K^3_{}\widehat{}^W`$ (resp. $`𝒪_K^3_{}_p`$) into itself, where $`\widehat{}^W`$ is the product of all $`_{\mathrm{}}`$’s for $`\mathrm{}`$ not belonging to $`W`$. We look at $`C_3^{\mathrm{max}}`$, which has the reduction map (also denoted by $`\pi _1^{}`$) to the finite group $`U_1^{}`$ (cf. 1). Set $`{}_{}{}^{I}C_{3}^{}=(\pi _1^{})^1(𝒥_1)`$ and $`{}_{}{}^{II}C_{3}^{}=(\pi _1^{})^1(𝒥_2)`$, where the groups $`𝒥_i`$ are those defined in 1. Now, we define
$${}_{}{}^{I}C_{\mathrm{CMSZ}}^{2}=C_{\mathrm{max}}^{2,3}{}_{}{}^{I}C_{3}^{}\text{and}{}_{}{}^{II}C_{\mathrm{CMSZ}}^{2}=C_{\mathrm{max}}^{2,3}{}_{}{}^{II}C_{3}^{}.$$
Obviously, the image of $`𝐈()(𝐈(_2)\times {}_{}{}^{I}C_{\mathrm{CMSZ}}^{2})`$ (resp. $`𝐈()(𝐈(_2)\times {}_{}{}^{II}C_{\mathrm{CMSZ}}^{2})`$) in $`𝐈_{\mathrm{ad}}(_2)\mathrm{PGL}_3_2`$ is nothing but $`\mathrm{\Gamma }_1`$ (resp. $`\mathrm{\Gamma }_2`$). We finally set
$${}_{}{}^{I}C=C_2^{\mathrm{max}}\varphi ({}_{}{}^{I}C_{\mathrm{CMSZ}}^{2})\text{and}{}_{}{}^{II}C=C_2^{\mathrm{max}}\varphi ({}_{}{}^{II}C_{\mathrm{CMSZ}}^{2}),$$
which are open compact subgroups of $`𝐆(𝔸_\mathrm{f})`$.
3.6. Theorem. The canonical model $`\mathrm{Sh}_{{}_{}{}^{I}C}`$ (resp. $`\mathrm{Sh}_{{}_{}{}^{II}C}`$) of the Shimura variety $`𝒮h_{{}_{}{}^{I}C}`$ (resp. $`𝒮h_{{}_{}{}^{II}C}`$) consists of two connected components defined over $`E^{}=(\sqrt{3},\sqrt{5})`$ which are permuted by the Galois action of $`\mathrm{Gal}(E^{}/E)`$. Moreover, the base change of each connected component to $`E_𝔭^{}_{2^2}`$ is isomorphic to the fake projective plane $`{}_{}{}^{I}X_{\mathrm{CMSZ}}^{}\times _{Spec_2}Spec_{2^2}`$ (resp. $`{}_{}{}^{II}X_{\mathrm{CMSZ}}^{}\times _{Spec_2}Spec_{2^2}`$) defined in 1.
Proof. The proof is parallel to that of \[Ka99, 4.3\]; the main part is to apply \[RZ96, 6.5\] to our situation. The only point which should be mentioned here is the computation of the morphism $`\vartheta :𝐆𝐓`$ by $`\gamma (\mathrm{Nm}_{D^{\mathrm{op}}|K}^{}(\gamma ),c(\gamma ))`$, where
$$𝐓=\{(k,f)\mathrm{Res}_{K/}𝔾_{\mathrm{m}}^{}{}_{,K}{}^{}\times 𝔾_{\mathrm{m}}^{}{}_{,}{}^{}|k\overline{k}=f^3\}\mathrm{Res}_{K/}𝔾_{\mathrm{m}}^{}{}_{,K}{}^{}.$$
We shall claim that the images $`\vartheta ({}_{}{}^{I}C)`$ and $`\vartheta ({}_{}{}^{II}C)`$ are maximal open compact subgroups of $`\mathrm{Res}_{K/}𝔾_{\mathrm{m}}^{}{}_{,K}{}^{}(𝔸_\mathrm{f})`$; once we prove this, the assertions on connected components and the field of definition follow from the fact that $`E^{}=(\sqrt{3},\sqrt{5})`$ is the Hilbert class field of $`E=K`$. To this end, we only have to look at the component at $`3`$. What to prove is that the homomorphism $`\vartheta \varphi `$ localized at $`3`$
$$\vartheta \varphi :{}_{}{}^{I}C_{3}^{}\text{or}{}_{}{}^{II}C_{3}^{}k\theta (k\overline{k})^1k^3det\theta \stackrel{~}{}_3^\times ,$$
where $`\stackrel{~}{}_3`$ is the integer ring of the quadratic ramified extension of $`_3`$, is surjective (here $`\theta `$ has been taken such that $`\theta ^{}Q\theta =Q`$).
Here we prove it for the group $`{}_{}{}^{I}C_{3}^{}`$; for the other one is similar, even easier. Since $`_3^\times `$ is contained in $`{}_{}{}^{I}C_{3}^{}`$, we have $`_3^\times \vartheta \varphi ({}_{}{}^{I}C_{3}^{})`$. Also, since $`\stackrel{~}{}_3^\times `$ is contained in $`{}_{}{}^{I}C_{3}^{}`$, we know the elememts of form $`k^3\overline{k}`$ for $`k\stackrel{~}{}_3^\times `$ belongs to $`\vartheta \varphi ({}_{}{}^{I}C_{3}^{})`$; but since $`k\overline{k}_3^\times `$ is in it, we deduce $`\{k^3|k\stackrel{~}{}_3^\times \}=\pm 1+3\stackrel{~}{}_3\vartheta \varphi ({}_{}{}^{I}C_{3}^{})`$. Hence it suffices to prove that $`(\vartheta \varphi ({}_{}{}^{I}C_{3}^{})`$ mod $`3)`$ is the whole $`(𝔽_3[t]/(t^2))^\times (\stackrel{~}{}_3/(3))^\times `$. But this can be checked easily; for instance, factoring through $`U_1^{}`$, we get the value $`1t`$ by the element $`u`$ (cf. 1), and hence get $`1+t`$ and $`1`$. Multiplying with $`1`$, we get the all. $`\mathrm{}`$
3.7. Corollary. Each connected component of the complex surface $`𝒮h_{{}_{}{}^{I}C}`$ and $`𝒮h_{{}_{}{}^{II}C}`$ is a fake projective plane, having a field of definition $`(\sqrt{3},\sqrt{5})`$. Moreover, it is an arithmetic quotient of the complex unit-ball. $`\mathrm{}`$
4. Appendix: Proof of Lemma 1
4.1. Let $`G_1=\{g|g(_2)^3=(_2)^3\}`$ and $`G_2=\{\pm \tau ^i|i=0,1,2\}`$. What to prove is the equality $`G_1=G_2`$. The inclusion $`G_2G_1`$ is clear. Let $`g`$ be an element in $`G_1`$. Then $`g`$ has entries in
$$𝒪_K[\frac{1}{2}]_2=[\lambda ,\frac{1}{2}]_2=[\overline{\lambda },\overline{\lambda }^1].$$
Since $`g`$ is $`Q`$-unitary, and since $`30Q^1`$ takes its entries in $`[\overline{\lambda },\overline{\lambda }^1]`$, they actually belongs to $`[\overline{\lambda },\overline{\lambda }^1]\frac{1}{30}[\lambda ,\lambda ^1]=+\frac{\lambda }{2}`$, which we denote by $`L`$. Set
$$V=\left\{vL^3\right|v^{}Qv=10\text{}\tau v\text{ and }\tau ^1v\text{ lie in }L^3\text{, and }v\frac{\lambda }{2}^3\text{ }\}.$$
Then each column vetcor of $`g`$ belongs to $`V`$; indeed, $`v^{}Qv=10`$ and $`\tau v`$, $`\tau ^1vL^3`$ are obvious, and, if it were in $`(\lambda /2)^3`$, then $`detg`$ would be a multiple of $`\lambda /2`$, which is not invertible in $`_2`$.
4.2. Claim. The set $`V`$ consists of $`24`$ vectors among
$$G_2\left[\begin{array}{c}\frac{\lambda }{2}\\ 1\\ 0\end{array}\right],G_2\left[\begin{array}{c}0\\ 0\\ 1\end{array}\right],G_2\left[\begin{array}{c}0\\ 1\\ 1\end{array}\right],\text{and}G_2\left[\begin{array}{c}1\\ 1\\ 0\end{array}\right].$$
The proof of the claim is done in the following several paragraphs:
4.3. For $`x=a+(\lambda /2)b`$ ($`a,b`$) we have $`x\overline{x}=a^2+(1/2)ab+b^2(3/4)(a^2+b^2)`$. Let $`v={}_{}{}^{t}(x_1,x_2,x_3)L^3`$ with $`x_i=a_i+(\lambda /2)b_i`$ ($`i=1,2,3`$). Since the characteristic polynomial of $`Q`$ is $`t^330t^3+210t300`$, one sees that the minimum eigenvalue of $`Q`$ is greater than $`1.92`$; therefore,
$`10=v^{}Qv`$ $``$ $`1.92|v|^2`$
$``$ $`1.92(3/4)((a_i^2+b_i^2)),`$
whence
(1.1)
$$(a_i^2+b_i^2)6.$$
In particular, we deduce $`|a_i|`$, $`|b_i|2`$. We are to strengthen this restraint into $`|a_i|`$, $`|b_i|1`$.
4.4. To this end, let us consider the $``$-quadratic form $`F(a_1,a_2,a_3,b_1,b_2,b_3)=v^{}Qv`$; more explicitly,
$`{\scriptscriptstyle \frac{1}{10}}F`$ $`=`$ $`a_1^2+a_2^2+a_3^2+b_1^2+b_2^2+b_3^2`$
$`a_1a_2a_2a_3+{\scriptscriptstyle \frac{1}{2}}a_3a_1b_1b_2b_2b_3+{\scriptscriptstyle \frac{1}{2}}b_3b_1`$
$`+{\scriptscriptstyle \frac{1}{2}}a_1b_1+{\scriptscriptstyle \frac{1}{2}}a_1b_2{\scriptscriptstyle \frac{1}{4}}a_1b_3a_2b_1`$
$`+{\scriptscriptstyle \frac{1}{2}}a_2b_2+{\scriptscriptstyle \frac{1}{2}}a_2b_3+{\scriptscriptstyle \frac{1}{2}}a_3b_1a_3b_2+{\scriptscriptstyle \frac{1}{2}}a_3b_3.`$
Then we see, writing $`(a_1,a_2,a_3,b_1,b_2,b_3)=(y_1,y_2,y_3,y_4,y_5,y_6)`$,
$${\scriptscriptstyle \frac{1}{10}}Fy_i^2_{i<j}|y_iy_j|,$$
where the right-hand side (which we denote by $`F^{}`$) is symmetric in $`y_i`$’s. Assume now that one of $`a_i`$’s or $`b_i`$’s has absolute value $`2`$. Due to (1.1) and the symmetry, the possible values of $`F^{}`$ are $`F^{}(2,0,0,0,0,0)=4`$, $`F^{}(2,1,0,0,0,0)=3`$, and $`F^{}(2,1,1,0,0,0)=1`$. Hence, in order that $`(1/10)F=1`$, $`(|a_1|,|a_2|,|a_3|,|b_1|,|b_2|,|b_3|)`$ has to be a permutation of $`(2,1,1,0,0,0)`$ and, moreover, the terms in $`F(y_i)`$ which survives to be non-zero must have coefficients $`\pm 1`$. But this is impossible, because coefficient of $`y_iy_j`$ is $`\pm 1`$ if and only if $`y_i`$ and $`y_j`$ is connected by an edge in
$$\begin{array}{ccccccc}y_1& \text{}& y_2& \text{}& y_3& & \\ & & |& & |& & \\ & & y_4& \text{}& y_5& \text{}& y_6\text{,}\end{array}$$
which does not have a cycle with three vertices. We therefore deduce that $`|a_i|`$, $`|b_i|1`$ when $`(1/10)F1`$.
4.5. Taking some auxiliary conditions into account, we get
$$V=\{v={}_{}{}^{t}(x_1,x_2,x_3)|a_i\text{}b_i\text{}|a_i|\text{}|b_i|1\text{}b_3=0\text{}(a_1,a_2,a_3)(0,0,0)\text{, and }F(a_i,b_j)=10\text{ }\}.$$
Indeed, $`\tau vL^3`$ implies that $`b_3`$ is even (hence is zero), and $`v(\lambda /2)^3`$ shows $`(a_1,a_2,a_3)(0,0,0)`$. Without the last condition $`F(a_i,b_j)=10`$, $`V`$ might have at most $`3^5=243`$ elements. The rest part of the proof of Claim 1 can be done by searching such vectors with $`F(a_i,b_j)=10`$, possibly by using computer.
4.6. Finally, we ask for, for any $`v_2V`$, two vectors $`v_1`$, $`v_3V`$ such that $`v_1^{}Qv_2=v_2^{}Qv_3=2(\lambda +2)`$ to get the matrix $`g=(v_1,v_2,v_3)`$ which is presumed to be $`Q`$-unitary. Due to Claim 1 it suffices to check the following four cases:
(1.1) If $`v_2={}_{}{}^{t}(\lambda /2,1,0)`$, there does not exist such $`v_1`$.
(1.2) If $`v_2={}_{}{}^{t}(0,0,1)`$, then $`v_3={}_{}{}^{t}(0,1,1)`$ and $`v_1`$ is either $`{}_{}{}^{t}(0,1,0)`$ or $`{}_{}{}^{t}(0,\lambda /2,1)`$; but in this case, $`(v_1,v_2,v_3)`$ is not an invertible matrix.
(1.3) If $`v_2={}_{}{}^{t}(0,1,1)`$, then $`v_1={}_{}{}^{t}(0,0,1)`$ and $`v_3={}_{}{}^{t}(0,1,0)`$ or $`{}_{}{}^{t}(\lambda /2,1+\lambda /2,1)`$. But, also in this case, $`det(v_1,v_2,v_3)`$ is not invertible in $`_2`$.
(1.4) If $`v_2={}_{}{}^{t}(0,1,0)`$, then $`v_1`$ is either $`{}_{}{}^{t}(1,0,0)`$, $`{}_{}{}^{t}(0,1,1)`$, or $`{}_{}{}^{t}(\overline{\lambda },1,0)`$, and $`v_3`$ is either $`{}_{}{}^{t}(0,0,1)`$ or $`{}_{}{}^{t}(1,1,0)`$. Among them, $`v_1^{}Qv_3=\lambda +2`$ is satisfied only when $`g=(v_1,v_2,v_3)=I_3`$.
All these are checked without so much pain. Since any member in $`G_2`$ is $`Q`$-unitary, we immediately see that the possible $`(v_1,v_2,v_3)`$ are only among elements in $`G_2`$. Therefore, we conclude $`G_1=G_2`$, as desired. $`\mathrm{}`$
Acknowledgments. The first author is grateful to Professor Yves André, to whom his viewpoint over this work owes much.
|
warning/0006/math0006166.html
|
ar5iv
|
text
|
# EFFECTIVE BIRATIONALITY OF PLURICANONICAL SYSTEMS
## 1 Introduction
Let $`X`$ be a smooth projective variety and let $`K_X`$ be the canonical bundle of $`X`$. $`X`$ is said to be of general type, if $`K_X`$ is big, i.e.,
$$\underset{m\mathrm{}}{lim\; sup}m^{dimX}dimH^0(X,𝒪_X(mK_X))>0$$
holds. The following problem is fundamental to study projective vareity of general type.
Problem Let $`X`$ be a smooth projective variety of general type. Find a positive integer $`m_0`$ such that for every $`mm_0`$, $`mK_X`$ gives a birational rational map from $`X`$ into a projective space.
If $`dimX=1`$, it is well known that $`3K_X`$ gives a projective embedding. In the case of smooth projective surfaces of general type, E. Bombieri showed that $`5K_X`$ gives a birational rational map from $`X`$ into a projective space (). In the case of $`dimX3`$, I have proved the following theorem.
###### Theorem 1.1
() There exists a positive integer $`\nu _n`$ which depends only on $`n`$ such that for every smooth projective $`n`$-fold $`X`$ of general type defined over complex numbers, $`mK_X`$ gives a birational rational map from $`X`$ into a projective space for every $`m\nu _n`$.
Theorem 1.1 is an affirmative answer to the problem. But it seems to be very hard to give an effective estimate of the number $`\nu _n`$ because the proof depends on the abstract facts of Hilbert scheme.
The main purpose of this article is to give the following weak effective answer to the problem.
###### Theorem 1.2
For every smooth projective $`n`$-fold $`X`$ of general type, one of the followings holds.
1. for every
$$m\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1,$$
$`mK_X`$ gives a birational rational map from $`X`$ into a projective space,
2. $`X`$ is dominated by a family of subvarieties of dimension $`d(1)`$ which are birational to subvarieties of degree less than or equal to $`(_{\mathrm{}=1}^n\sqrt[\mathrm{}]{2}\mathrm{}+1)^d`$ in a projective space by some pluricanonical system $`\alpha K_X`$.
###### Theorem 1.3
For every smooth projective $`n`$-fold $`X`$ of general type, one of the followings holds.
1. for every
$$m\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1,$$
$`mK_X`$ gives a birational rational map from $`X`$ into a projective space,
2. there exists a rational fibration
$$f:X\mathrm{}Y$$
such that a general fiber $`F`$ of $`f`$ is positive dimensional and is birational to a subvariety of degree less than or equal to $`(_{\mathrm{}=1}^n\sqrt[\mathrm{}]{2}\mathrm{}+1)^{d^2}d^d`$ in a projective space by some pluricanonical system $`\alpha K_X`$, where $`d`$ denotes the dimension of $`F`$.
The proofs of Theorem 1.2 and 1.3 are technically much easier than that of Theorem 1.1 (). But they are effective and clarify the essential obstruction to obtain the birationality of the pluricanonical map $`\mathrm{\Phi }_{mK_X}`$ with relatively small $`m`$. As one see in the proof, if we need very large $`m`$ to embed $`X`$ birationally into a projective space by $`mK_X`$, the image $`\mathrm{\Phi }_{mK_X}(X)`$ is distorted in the sense that $`X`$ is small in the fiber direction and large in the horizontal direction with respect to a rational fiber space structure. Such a phenomenon was first observed by E. Bombieri in his paper . Actually he found the existence of a genus 2 fibration is an obstruction to the birationality of $`2K_X`$ for some surfaces of general type (\[5, p. 173, Main Theorem (iv)\]). Of course the results above will not be optimal and more abstract in comparison with the case of surfaces.
In the case of 3-folds of general type, there were several results in this direction. But these results depend on the plurigenus formula for canonical 3-folds of general type and moreover their estimates depend on the apriori bound of $`\chi (X,𝒪_X)`$, hence it is even weaker than Theorem 1.1 in this respect and the bound is is so huge that they have only a theoretical interests.
I hope the estimates in Theorem 1.2 and Theorem 1.3 are acceptable at least for projective varieties of low dimension. And it seems to be more or less optimal in order of size, even in the case of arbitrary dimension. But I should say that even in the case of 3-folds, the exceptional cases seem to be very hard to classify.
As in , the main difficulty is the fact that $`K_X`$ is not ample in general. To overcome this difficulty we use a special singular hemitian metric on $`K_X`$ called AZD which was originated by the author (). By using AZD we can handle $`K_X`$ as if $`K_X`$ were nef and big.
## 2 Preliminaries
### 2.1 Multiplier ideal sheaves
In this section, we shall review the basic definitions and properties of multiplier ideal sheaves.
###### Definition 2.1
Let $`L`$ be a line bundle on a complex manifold $`M`$. A singular hermitian metric $`h`$ is given by
$$h=e^\phi h_0,$$
where $`h_0`$ is a $`C^{\mathrm{}}`$-hermitian metric on $`L`$ and $`\phi L_{loc}^1(M)`$ is an arbitrary function on $`M`$.
The curvature current $`\mathrm{\Theta }_h`$ of the singular hermitian line bundle $`(L,h)`$ is defined by
$$\mathrm{\Theta }_h:=\mathrm{\Theta }_{h_0}+\sqrt{1}\overline{}\phi ,$$
where $`\overline{}`$ is taken in the sense of a current. The $`L^2`$-sheaf $`^2(L,h)`$ of the singular hermitian line bundle $`(L,h)`$ is defined by
$$^2(L,h):=\{\sigma \mathrm{\Gamma }(U,𝒪_M(L))h(\sigma ,\sigma )L_{loc}^1(U)\},$$
where $`U`$ runs opens subsets of $`M`$. In this case there exists an ideal sheaf $`(h)`$ such that
$$^2(L,h)=𝒪_M(L)(h)$$
holds. We call $`(h)`$ the multiplier ideal sheaf of $`(L,h)`$. If we write $`h`$ as
$$h=e^\phi h_0,$$
where $`h_0`$ is a $`C^{\mathrm{}}`$ hermitian metric on $`L`$ and $`\phi L_{loc}^1(M)`$ is the weight function, we see that
$$(h):=^2(𝒪_M,e^\phi )$$
holds. We also denote $`^2(𝒪_M,e^\phi )`$ by $`(\phi )`$. Let $`(L,h)`$ be a singular hermitian line bundle on a smooth projective variety $`X`$ such that
$$\mathrm{\Theta }_h\omega $$
holds for some $`C^{\mathrm{}}`$ Kähler form $`\omega `$ on $`X`$. Then by \[12, p. 561\], we see that $`(h)`$ is a coherent sheaf of $`𝒪_X`$-ideal.
Similarly we obtain the sheaf
$$_{\mathrm{}}(h):=^{\mathrm{}}(𝒪_M,e^\phi )$$
and call it the $`L^{\mathrm{}}`$-multiplier ideal sheaf of $`(L,h)`$. We have the following vanishing theorem.
###### Theorem 2.1
(Nadel’s vanishing theorem \[12, p.561\]) Let $`(L,h)`$ be a singular hermitian line bundle on a compact Kähler manifold $`M`$ and let $`\omega `$ be a Kähler form on $`M`$. Suppose that $`\mathrm{\Theta }_h`$ is strictly positive, i.e., there exists a positive constant $`\epsilon `$ such that
$$\mathrm{\Theta }_h\epsilon \omega $$
holds. Then $`(h)`$ is a coherent sheaf of $`𝒪_M`$-ideal and for every $`q1`$
$$H^q(M,𝒪_M(K_M+L)(h))=0$$
holds.
### 2.2 Analytic Zariski decomposition
To study a big line bundle we introduce the notion of analytic Zariski decompositions. By using analytic Zariski decompositions, we can handle big line bundles like a nef and big line bundles.
###### Definition 2.2
Let $`M`$ be a compact complex manifold and let $`L`$ be a line bundle on $`M`$. A singular hermitian metric $`h`$ on $`L`$ is said to be an analytic Zariski decomposition, if the followings hold.
1. $`\mathrm{\Theta }_h`$ is a closed positive current,
2. for every $`m0`$, the natural inclusion
$$H^0(M,𝒪_M(mL)(h^m))H^0(M,𝒪_M(mL))$$
is isomorphim.
###### Remark 2.1
If an AZD exists on a line bundle $`L`$ on a smooth projective variety $`M`$, $`L`$ is pseudoeffective by the condition 1 above.
###### Theorem 2.2
( see also \[20, Section 2.2\]) Let $`L`$ be a big line bundle on a smooth projective variety $`M`$. Then $`L`$ has an AZD.
### 2.3 Volume of subvarieties
To measure the positivity of big line bundles on a projective variety we shall introduce a volume of a projective variety with respect to a line bundle.
###### Definition 2.3
Let $`L`$ be a line bundle on a compact complex manifold $`M`$ of dimension $`n`$. We define the $`L`$-volume of $`M`$ by
$$\mu (M,L):=n!\overline{lim}_m\mathrm{}m^ndimH^0(M,𝒪_M(mL)).$$
###### Definition 2.4
() Let $`(L,h)`$ be a singular hermitian line bundle on a smooth projective variety $`X`$ such that $`\mathrm{\Theta }_h0`$. Let $`Y`$ be a subvariety of $`X`$ of dimension $`r`$. We define the volume $`\mu (Y,L)`$ of $`Y`$ with respect to $`L`$ by
$$\mu (Y,L):=r!\overline{lim}_m\mathrm{}m^rdimH^0(Y,𝒪_Y(mL)(h^m)/tor),$$
where $`tor`$ denotes the torsion part of the sheaf $`𝒪_Y(mL)(h^m)`$.
## 3 Stratification of varieties by multiplier ideal sheaves
In this section we shall construct a stratification of a smooth projective variety $`X`$ of general type by using an AZD $`h`$ of $`K_X`$. We use the ideas in to construct the stratification. But since $`(K_X,h)`$ is not an ample line bundle, the argument is a little bit more involved.
### 3.1 Construction of a stratification
Let $`X`$ be a smooth projective $`n`$-fold of general type. Let $`h`$ be an AZD of $`K_X`$. Let us denote $`\mu (X,K_X)`$ by $`\mu _0`$. We set
$$X^{}=\{xXx\text{Bs}mK_X\text{and }\mathrm{\Phi }_{mK_X}\text{ is a biholomorphism}$$
$$\text{ on a neighbourhood of }x\text{ for some }m1\}.$$
Then $`X^{}`$ is a nonempty Zariski open subset of $`X`$.
###### Lemma 3.1
Let $`x,x^{}`$ be distinct points on $`X^{}`$. We set
$$_{x,x^{}}=_x_x^{}$$
Let $`\epsilon `$ be a sufficiently small positive number. Then
$$H^0(X,𝒪_X(mK_X)_{x,x^{}}^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m}{\sqrt[n]{2}}})0$$
for every sufficiently large $`m`$, where $`_x,_x^{}`$ denote the maximal ideal sheaf of the points $`x,x^{}`$ respectively.
Proof of Lemma 3.1. Let us consider the exact sequence:
$$0H^0(X,𝒪_X(mK_X)_{x,x^{}}^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m}{\sqrt[n]{2}}})H^0(X,𝒪_X(mK_X))$$
$$H^0(X,𝒪_X(mK_X)/_{x,x^{}}^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m}{\sqrt[n]{2}}}).$$
Since
$$n!\overline{lim}_m\mathrm{}m^ndimH^0(X,𝒪_X(mK_X)/_{x,x^{}}^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m}{\sqrt[n]{2}}})=\mu _0(1\epsilon )^n<\mu _0$$
hold, we see that Lemma 3.1 holds. Q.E.D.
Let us take a sufficiently large positive integer $`m_0`$ and let $`\sigma `$ be a general (nonzero) element of $`H^0(X,𝒪_X(m_0K_X)_{x,x^{}}^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m_0}{\sqrt[n]{2}}})`$. We define a singular hermitian metric $`h_0`$ on $`K_X`$ by
$$h_0(\tau ,\tau ):=\frac{\tau ^2}{\sigma ^{2/m_0}}.$$
Then
$$\mathrm{\Theta }_{h_0}=\frac{2\pi }{m_0}(\sigma )$$
holds, where $`(\sigma )`$ denotes the closed positive current defined by the divisor $`(\sigma )`$. Hence $`\mathrm{\Theta }_{h_0}`$ is a closed positive current. Let $`\alpha `$ be a positive number and let $`(\alpha )`$ denote the multiplier ideal sheaf of $`h_0^\alpha `$, i.e.,
$$(\alpha )=^2(𝒪_X,(\frac{h_0}{h_X})^\alpha ),$$
where $`h_X`$ is an arbitrary $`C^{\mathrm{}}`$-hermitian metric on $`K_X`$. Let us define a positive number $`\alpha _0(=\alpha _0(x,y))`$ by
$$\alpha _0:=inf\{\alpha >0(𝒪_X/(\alpha ))_x0\text{and}(𝒪_X/(\alpha ))_x^{}0\}.$$
Since $`(_{i=1}^nz_i^2)^n`$ is not locally integrable around $`O\text{C}^n`$, by the construction of $`h_0`$, we see that
$$\alpha _0\frac{n\sqrt[n]{2}}{\sqrt[n]{\mu _0}(1\epsilon )}$$
holds. Then one of the following two cases occurs.
Case 1.1: For every small positive number $`\delta `$, $`𝒪_X/(\alpha _0\delta )`$ has $`0`$-stalk at both $`x`$ and $`x^{}`$.
Case 1.2: For every small positive number $`\delta `$, $`𝒪_X/(\alpha _0\delta )`$ has nonzero-stalk at one of $`x`$ or $`x^{}`$ say $`x^{}`$.
We first consider Case 1.1. Let $`\delta `$ be a sufficiently small positive number and let $`V_1`$ be the germ of subscheme at $`x`$ defined by the ideal sheaf $`(\alpha _0+\delta )`$. By the coherence of $`(\alpha )(\alpha >0)`$, we see that if we take $`\delta `$ sufficiently small, then $`V_1`$ is independent of $`\delta `$. It is also easy to verify that $`V_1`$ is reduced if we take $`\delta `$ sufficiently small. In fact if we take a log resolution of $`(X,\frac{\alpha _0}{m_0}(\sigma ))`$, $`V_1`$ is the image of the divisor with discrepancy $`1`$ (for example cf. \[7, p.207\]). Let $`X_1`$ be a subvariety of $`X`$ which defines a branch of $`V_1`$ at $`x`$. We consider the following two cases.
Case 2.1: $`X_1`$ passes through both $`x`$ and $`x^{}`$,
Case 2.2: Otherwise
For the first we consider Case 2.1. Suppose that $`X_1`$ is not isolated at $`x`$. Let $`n_1`$ denote the dimension of $`X_1`$. Let us define the volume $`\mu _1`$ of $`X_1`$ with respect to $`K_X`$ by
$$\mu _1:=\mu (X_1,K_X).$$
Since $`xX^{}`$, we see that $`\mu _1>0`$ holds.
###### Lemma 3.2
Let $`\epsilon `$ be a sufficiently small positive number and let $`x_1,x_2`$ be distinct regular points on $`X_1X^{}`$. Then for a sufficiently large $`m>1`$,
$$H^0(X_1,𝒪_{X_1}(mK_X)(h^m)_{x_1,x_2}^{\sqrt[n_1]{\mu _1}(1\epsilon )\frac{m}{\sqrt[n_1]{2}}})0$$
holds.
The proof of Lemma 3.2 is identical as that of Lemma 3.1, since
$$(h^m)_{x_i}=𝒪_{X,x_i}(i=1,2)$$
hold for every $`m`$ by Proposition 2.1 and Lemma 2.1.
By Kodaira’s lemma there is an effective Q-divisor $`E`$ such that $`K_XE`$ is ample. Let $`\mathrm{}_1`$ be a sufficiently large positive integer which will be specified later such that
$$L_1:=\mathrm{}_1(K_XE)$$
is Cartier.
###### Lemma 3.3
If we take $`\mathrm{}_1`$ sufficiently large, then
$$\varphi _m:H^0(X,𝒪_X(mK_X+L_1)(h^m))H^0(X_1,𝒪_{X_1}(mK_X+L_1)(h^m))$$
is surjective for every $`m0`$.
Proof. Let us take a locally free resolution of the ideal sheaf $`_{X_1}`$ of $`X_1`$.
$$0_{X_1}_1_2\mathrm{}_k0.$$
Then by the trivial extension of the case of vector bundles, if $`\mathrm{}_1`$ is sufficiently large, we see that
$$H^q(X,𝒪_X(mK_X+L_1)(h^m)_j)=0$$
holds for every $`m1`$, $`q1`$ and $`1jk`$. In fact if we take $`\mathrm{}_1`$ sufficiently large, we see that for every $`j`$, $`𝒪_X(L_1K_X)_j`$ admits a $`C^{\mathrm{}}`$-hermitian metric $`g_j`$ such that
$$\mathrm{\Theta }_{g_j}\text{Id}_{E_j}\omega $$
holds, where $`\omega `$ is a Kähler form on $`X`$. By \[6, Theorem 4.1.2 and Lemma 4.2.2\] we have the desired vanishing. Hence we have:
###### Sublemma 3.1
$$H^1(X,𝒪_X(mK_X+L_1)(h^m)_j)=0$$
holds for every $`m0`$ and $`1jr`$
Let
$$p_m:X_mX$$
be a composition of successive blowing ups with smooth centers such that $`p_m^{}(h^m)`$ is locally free on $`X_m`$.
###### Sublemma 3.2
$$R^pp_m(𝒪_{X_m}(K_{X_m})(p_m^{}h^m))=0$$
holds for every $`p1`$ and $`m1`$.
We note that by the definition of multiplier ideal sheaves
$$p_m(𝒪_{X_m}(K_{X_m})(p_m^{}h^m))=𝒪(K_X)(h^m)$$
holds. Hence by Sublemma 3.1 and Sublemma 3.2 and the Leray spectral sequence, we see that
$$H^q(X_m,𝒪_{X_m}(K_{X_m}+p_m^{}(mK_X+L_1K_X))(p_m^{}h^m)p_m^{}_j)=0$$
holds for every $`q1`$ and $`m1`$. Hence every element of
$$H^0(X_m,𝒪_{X_m}(K_{X_m}+p_m^{}(mK_X+L_1K_X))(p_m^{}h^m)𝒪_{X_m}/p_m^{}_{X_1})$$
extends to an element of
$$H^0(X_m,𝒪_{X_m}(K_{X_m}+p_m^{}(mK_X+L_1K_X))(p_m^{}h^m)).$$
Also there exists a natural map
$$H^0(X_1,𝒪_{X_1}(mK_X+L_1)(h^m))$$
$$H^0(X_m,𝒪_{X_m}(K_{X_m}+p_m^{}(mK_X+L_1K_X))(p_m^{}h^m)𝒪_{X_m}/p_m^{}_{X_1}).$$
Hence we can extend every element of
$$p_m^{}H^0(X_1,𝒪_{X_1}(mK_X+L_1)(h^m))$$
to an element of
$$H^0(X_m,𝒪_{X_m}(K_{X_m}+p_m^{}(mK_X+L_1K_X))(p_m^{}h^m)).$$
Since
$$H^0(X_m,𝒪_{X_m}(K_{X_m}+p_m^{}(mK_X+L_1K_X))(p_m^{}h^m))$$
$$H^0(X,𝒪_X(mK_X+L_1)(h^m))$$
holds by the isomorphism
$$p_m(𝒪_{X_m}(K_{X_m})(p_m^{}h^m))=𝒪(K_X)(h^m),$$
this completes the proof of Lemma 3.3. Q.E.D.
Let $`\tau `$ be a general section in $`H^0(X,𝒪_X(L_1))`$.
Let $`m_1`$ be a sufficiently large positive integer and let $`\sigma _1^{}`$ be a general element of
$$H^0(X_1,𝒪_{X_1}(m_1K_X)(h^{m_1})_{x_1,x_2}^{\sqrt[n_1]{\mu _1}(1\epsilon )\frac{m_1}{\sqrt[n_1]{2}}}),$$
where $`x_1,x_2X_1`$ are distinct nonsingular points on $`X_1`$.
By Lemma 3.2, we may assume that $`\sigma _1^{}`$ is nonzero. Then by Lemma 3.3 we see that
$$\sigma _1^{}\tau H^0(X_1,𝒪_{X_1}(m_1K_X+L_1)(h^{m_1})_{x_1,x_2}^{\sqrt[n_1]{\mu _1}(1\epsilon )\frac{m_1}{\sqrt[n_1]{2}}})$$
extends to a section
$$\sigma _1H^0(X,𝒪_X((m_1+\mathrm{}_1)K_X)(h^{m+\mathrm{}_1}))$$
We may assume that there exists a neighbourhood $`U_{x,x^{}}`$ of $`\{x,x^{}\}`$ such that the divisor $`(\sigma _1)`$ is smooth on $`U_{x,x^{}}X_1`$ by Bertini’s theorem, if we take $`\mathrm{}_1`$ sufficiently large, since by Theorem 2.1,
$$H^0(X,𝒪_X(mK_X+L_1)(h^m))H^0(X,𝒪_X(mK_X+L_1)(h^m))/𝒪_X(X_1)_y)$$
is surjective for every $`yX`$ and $`m0`$, where $`𝒪_X(X_1)`$ is the ideal sheaf of $`X_1`$. We define a singular hermitian metric $`h_1`$ on $`K_X`$ by
$$h_1=\frac{1}{\sigma _1^{\frac{2}{m_1+\mathrm{}_1}}}.$$
Let $`\epsilon _0`$ be a sufficiently small positive number and let $`_1(\alpha )`$ be the multiplier ideal sheaf of $`h_0^{\alpha _0\epsilon _0}h_1^\alpha `$,i.e.,
$$_1(\alpha )=^2(𝒪_X,h_0^{\alpha _0\epsilon _0}h_1^\alpha /h_X^{(\alpha _0+\alpha \epsilon _0)}).$$
Suppose that $`x,x^{}`$ are nonsingular points on $`X_1`$. Then we set $`x_1=x,x_2=x^{}`$ and define $`\alpha _1(=\alpha _1(x,y))>0`$ by
$$\alpha _1:=inf\{\alpha (𝒪_X/_1(\alpha ))_x0\text{and}(𝒪_X/_1(\alpha ))_x^{}0\}.$$
By Lemma 3.3 we may assume that we have taken $`m_1`$ so that
$$\frac{\mathrm{}_1}{m_1}\epsilon _0\frac{\sqrt[n_1]{\mu _1}}{n_1\sqrt[n_1]{2}}$$
holds.
###### Lemma 3.4
$$\alpha _1\frac{n_1\sqrt[n_1]{2}}{\sqrt[n_1]{\mu _1}}+O(\epsilon _0)$$
holds.
To prove Lemma 3.4, we need the following elementary lemma.
###### Lemma 3.5
(\[21, p.12, Lemma 6\]) Let $`a,b`$ be positive numbers. Then
$$_0^1\frac{r_2^{2n_11}}{(r_1^2+r_2^{2a})^b}𝑑r_2=r_1^{\frac{2n_1}{a}2b}_0^{r_1^{2a}}\frac{r_3^{2n_11}}{(1+r_3^{2a})^b}𝑑r_3$$
holds, where
$$r_3=r_2/r_1^{1/a}.$$
Proof of Lemma 3.3. Let $`(z_1,\mathrm{},z_n)`$ be a local coordinate on a neighbourhood $`U`$ of $`x`$ in $`X`$ such that
$$UX_1=\{qUz_{n_1+1}(q)=\mathrm{}=z_n(q)=0\}.$$
We set $`r_1=(_{i=n_1+1}^nz_1^2)^{1/2}`$ and $`r_2=(_{i=1}^{n_1}z_i^2)^{1/2}`$. Then there exists a positive constant $`C`$ such that
$$\sigma _1^2C(r_1^2+r_2^{2\sqrt[n_1]{\mu _1}(1\epsilon )\frac{m_1}{\sqrt[n_1]{2}}})$$
holds on a neighbourhood of $`x`$, where $``$ denotes the norm with respect to $`h_X^{m_1+\mathrm{}_1}`$. We note that there exists a positive integer $`M`$ such that
$$\sigma ^2=O(r_1^M)$$
holds on a neighbourhood of the generic point of $`UX_1`$, where $``$ denotes the norm with respect to $`h_X^{m_0}`$. Then by Lemma 3.5, we have the inequality
$$\alpha _1(\frac{m_1+\mathrm{}_1}{m_1})\frac{n_1\sqrt[n_1]{2}}{\sqrt[n_1]{\mu _1}}+O(\epsilon _0)$$
holds. By using the fact that
$$\frac{\mathrm{}_1}{m_1}\epsilon _0\frac{\sqrt[n_1]{\mu _1}}{n_1\sqrt[n_1]{2}}$$
we obtain that
$$\alpha _1\frac{n_1\sqrt[n_1]{2}}{\sqrt[n_1]{\mu _1}}+O(\epsilon _0)$$
holds. Q.E.D.
If $`x`$ or $`x^{}`$ is a singular point on $`X_1`$, we need the following lemma.
###### Lemma 3.6
Let $`\phi `$ be a plurisubharmonic function on $`\mathrm{\Delta }^n\times \mathrm{\Delta }`$. Let $`\phi _t(t\mathrm{\Delta })`$ be the restriction of $`\phi `$ on $`\mathrm{\Delta }^n\times \{t\}`$. Assume that $`e^{\phi _t}`$ does not belong to $`L_{loc}^1(\mathrm{\Delta }^n,O)`$ for every $`t\mathrm{\Delta }^{}`$.
Then $`e^{\phi _0}`$ is not locally integrable at $`O\mathrm{\Delta }^n`$.
Lemma 3.6 is an immediate consequence of . Using Lemma 3.6 and Lemma 3.5, we see that Lemma 3.4 holds by letting $`x_1x`$ and $`x_2x^{}`$.
For the next we consider Case 1.2 and Case 2.2. We note that in Case 2.2 by modifying $`\sigma `$ a little bit , if necessary we may assume that $`(𝒪_X/(\alpha _0\epsilon ))_x^{}0`$ and $`(𝒪_X/(\alpha _0\epsilon ^{}))_x=0`$ hold for a sufficiently small positive number $`\epsilon ^{}`$. For example it is sufficient to replace $`\sigma `$ by the following $`\sigma ^{}`$ constructed below.
Let $`X_1^{}`$ be a subvariety which defines a branch of
$$\text{Spec}(𝒪_X/(\alpha +\delta ))$$
at $`x^{}`$. By the assumption (changing $`X_1`$, if necessary) we may assume that $`X_1^{}`$ does not contain $`x`$. Let $`m^{}`$ be a sufficiently large positive integer such that $`m^{}/m_0`$ is sufficiently small (we can take $`m_0`$ arbitrary large).
Let $`\tau _x^{}`$ be a general element of
$$H^0(X,𝒪_X(m^{}K_X)_{X_1^{}}),$$
where $`_{X_1^{}}`$ is the ideal sheaf of $`X_1^{}`$. If we take $`m^{}`$ sufficiently large, $`\tau _x^{}`$ is not identically zero. We set
$$\sigma ^{}=\sigma \tau _x^{}.$$
Then we see that the new singular hermitian metric $`h_0^{}`$ defined by $`\sigma ^{}`$ satisfies the desired property.
In these cases, instead of Lemma 3.2, we use the following simpler lemma.
###### Lemma 3.7
Let $`\epsilon `$ be a sufficiently small positive number and let $`x_1`$ be a smooth point on $`X_1`$. Then for a sufficiently large $`m>1`$,
$$H^0(X_1,𝒪_{X_1}(mK_X)(h^m)_{x_1}^{\sqrt[n_1]{\mu _1}(1\epsilon )m})0$$
holds.
Then taking a general $`\sigma _1^{}`$ in
$$H^0(X_1,𝒪_{X_1}(m_1K_X)(h^{m_1})_{x_1}^{\sqrt[n_1]{\mu _1}(1\epsilon )m_1}),$$
for a sufficiently large $`m_1`$. As in Case 1.1 and Case 2.1 we obtain a proper subvariety $`X_2`$ in $`X_1`$ also in this case.
Inductively for distinct points $`x,x^{}X^{}`$, we construct a strictly decreasing sequence of subvarieties
$$X=X_0X_1\mathrm{}X_rX_{r+1}=\{x\}R_x^{}\text{or}\{x^{}\}R_x,$$
where $`R_x^{}`$ (or $`R_x`$) is a subvariety such that $`x`$ deos not belong to $`R_x^{}`$ and $`x^{}`$ belongs to $`R_x^{}`$. and invariants (depending on small positive numbers $`\epsilon _0,\mathrm{},\epsilon _{r1}`$, large positive integers $`m_0,m_1,\mathrm{},m_r`$, etc.) :
$$\alpha _0,\alpha _1,\mathrm{},\alpha _r,$$
$$\mu _0,\mu _1,\mathrm{},\mu _r$$
and
$$n>n_1>\mathrm{}>n_r.$$
By Nadel’s vanishing theorem we have the following lemma.
###### Lemma 3.8
Let $`x,x^{}`$ be two distinct points on $`X^{}`$. Then for every $`m_{i=0}^r\alpha _i+1`$, $`\mathrm{\Phi }_{mK_X}`$ separates $`x`$ and $`x^{}`$.
Proof. Let us define the singular hermitian metric $`h_{x,x^{}}`$ of $`(m1)K_X`$ defined by
$$h_{x,x^{}}=(\underset{i=0}{\overset{r1}{}}h_i^{\alpha _i\epsilon _i})h_r^{\alpha _r+\epsilon _r}h^{(m1(_{i=0}^{r1}(\alpha _i\epsilon _i))(\alpha _r+\epsilon _r)\delta _L)}h_L^{\delta _L},$$
where $`h_L`$ is a $`C^{\mathrm{}}`$-hermitian metric on the Q-line bundle $`L:=K_XE`$ with strictly positive curvature and $`\delta _L`$ be a sufficiently small positive number. Then we see that $`(h_{x,x^{}})`$ defines a subscheme of $`X`$ with isolated support around $`x`$ or $`x^{}`$ by the definition of the invariants $`\{\alpha _i\}`$’s. By the construction the curvature current $`\mathrm{\Theta }_{h_{x,x^{}}}`$ is strictly positive on $`X`$. Then by Nadel’s vanishing theorem (Theorem 2.1) we see that
$$H^1(X,𝒪_X(mK_X)(h_{x,x^{}}))=0.$$
This implies that $`\mathrm{\Phi }_{mK_X}`$ separates $`x`$ and $`x^{}`$. Q.E.D.
### 3.2 Construction of the stratification as a family
In this subsection we shall construct the above stratification as a family.
We note that for a fixed pair $`(x,x^{})X^{}\times X^{}\mathrm{\Delta }_X`$, $`_{i=0}^r\alpha _i`$ depends on the choice of $`\{X_i\}`$’s, where $`\mathrm{\Delta }_X`$ denotes the diagonal of $`X\times X`$. Moving $`(x,x^{})`$ in $`X^{}\times X^{}\mathrm{\Delta }_X`$, we shall consider the above operation simultaneously. Let us explain the procedure. We set
$$B:=X^{}\times X^{}\mathrm{\Delta }_X.$$
Let
$$p:X\times BX$$
be the first projection and let
$$q:X\times BB$$
be the second projection. Let $`Z`$ be the subvariety of $`X\times B`$ defined by
$$Z:=\{(x_1,x_2,x_3):X\times Bx_1=x_2\text{or}x_1=x_3\}.$$
In this case we consider
$$q_{}𝒪_{X\times B}(m_0p^{}K_X)_Z^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m_0}{\sqrt[n]{2}}}$$
instead of
$$H^0(X,𝒪_X(m_0K_X)_{x,x^{}}^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m_0}{\sqrt[n]{2}}}),$$
where $`_Z`$ denotes the ideal sheaf of $`Z`$. Let $`\stackrel{~}{\sigma }_0`$ be a nonzero global meromorphic section of
$$q_{}𝒪_{X\times B}(m_0p^{}K_X)_Z^{\sqrt[n]{\mu _0}(1\epsilon )\frac{m_0}{\sqrt[n]{2}}}$$
on $`B`$ for a sufficiently large positive integer $`m_0`$. We define the singular hermitian metric $`\stackrel{~}{h}_0`$ on $`p^{}K_X`$ by
$$\stackrel{~}{h}_0:=\frac{1}{\stackrel{~}{\sigma }_0^{2/m_0}}.$$
We shall replace $`\alpha _0`$ by
$$\stackrel{~}{\alpha }_0:=inf\{\alpha >0\text{the generic point of}Z\text{Spec}(𝒪_{X\times B}/(h_0^\alpha ))\}.$$
Then for every $`0<\delta <<1`$, there exists a Zariski open subset $`U`$ of $`B`$ such that for every $`bU`$, $`\stackrel{~}{h}_0_{X\times \{b\}}`$ is well defined and
$$b\text{Spec}(𝒪_{X\times \{b\}}/(\stackrel{~}{h}_0^{\alpha _0\delta }_{X\times \{b\}})),$$
where we have identified $`b`$ with distinct two points in $`X`$. And also by Lemma 3.6, we see that
$$b\text{Spec}(𝒪_{X\times \{b\}}/(\stackrel{~}{h}_0^{\alpha _0}_{X\times \{b\}})),$$
holds for every $`bB`$. Let $`\stackrel{~}{X}_1`$ be an irreducible component of
$$\text{Spec}(𝒪_{X\times B}/(\stackrel{~}{h}_0^{\alpha _0}))$$
containing $`Z`$. We note that $`\stackrel{~}{X}_1q^1(b)`$ may not be irreducible even for a general $`bB`$. But if we take a suitable finite cover
$$\varphi _0:B_0B,$$
on the base change $`X\times _BB_0`$, $`\stackrel{~}{X}_1`$ defines a family of irreducible subvarieties
$$f_1:\widehat{X}_1U_0$$
of $`X`$ parametrized by a nonempty Zariski open subset $`U_0`$ of $`\varphi _0^1(U)`$. We set
$$\stackrel{~}{\mu }_1:=\underset{b_0U_0}{inf}\mu (f_1^1(b_0),(K_X,h)).$$
We note that by its definition the volume $`\mu (f_1^1(b_0),(K_X,h))`$ is constant on a nonempty open subset say $`U_0^{}`$ of $`U_0`$ with respect to countable Zariski topology. We denote the constant by $`\stackrel{~}{\mu }_0`$. Continueing this process we may construct a finite morphism
$$\varphi _r:B_rB$$
and a nonempty Zariski open subset $`U_r`$ of $`B_r`$ which parametrizes a family of stratification
$$XX_1X_2\mathrm{}X_rX_{r+1}=\{x\}R_x^{}(\text{resp.}\{x^{}\}R_x)$$
constructed as before, where $`R_x`$ (resp. $`R_x^{}`$) is a subvariety of $`X`$ which is disjoint from $`x^{}`$ (resp. $`x`$). And we also obtain invariants $`\{\stackrel{~}{\alpha }_0,\mathrm{},\stackrel{~}{\alpha }_r\}`$, $`\{\stackrel{~}{\mu }_0,\mathrm{},\stackrel{~}{\mu }_r\}`$, $`\{n=\stackrel{~}{n}_0\mathrm{},\stackrel{~}{n}_r\}`$. Hereafter we denote these invariants without $`\stackrel{~}{}`$ for simplicity. By the same proof as Lemma 3.4, we have the following lemma.
###### Lemma 3.9
$$\alpha _i\frac{n_i\sqrt[n_i]{2}}{\sqrt[n_i]{\mu _i}}+O(\epsilon _{i1})$$
hold for $`1ir`$.
By Lemma 3.8 we obtain the following proposition.
###### Proposition 3.1
For every
$$m>\underset{i=0}{\overset{r}{}}\alpha _i+1$$
$`mK_X`$ gives a birational rational map from $`X`$ into a projective space.
###### Lemma 3.10
Let $`X_i`$ be a strata of a very general member of the stratification parametrized by $`B_r`$. If $`\mathrm{\Phi }_{mK_X}_{X_i}`$ is birational rational map onto its image, then
$$\mathrm{deg}\mathrm{\Phi }_{mK_X}(X_i)m^{n_i}\mu _i$$
holds.
Proof. Let $`p:\stackrel{~}{X}X`$ be the resolution of the base locus of $`mK_X`$ and let
$$p^{}mK_X=P_m+F_m$$
be the decomposition into the free part $`P_m`$ and the fixed component $`F_m`$. Let $`p_i:\stackrel{~}{X}_iX_i`$ be the resolution of the base locus of $`\mathrm{\Phi }_{mK_X}_{X_i}`$ obtained by the restriction of $`p`$ on $`p^1(X_i)`$. Let
$$p_i^{}(mK_X_{X_i})=P_{m,i}+F_{m,i}$$
be the decomposition into the free part $`P_{m,i}`$ and the fixed part $`F_{m,i}`$. We have
$$\mathrm{deg}\mathrm{\Phi }_{mK_X}(X_i)=P_{m,i}^{n_i}$$
holds. Then by the ring structure of $`R(X,K_X)`$, we have an injection
$$H^0(\stackrel{~}{X},𝒪_{\stackrel{~}{X}}(\nu P_m))H^0(X,𝒪_X(m\nu K_X)(h^{m\nu }))$$
for every $`\nu 1`$, since the righthandside is isomorphic to $`H^0(X,𝒪_X(m\nu K_X))`$ by the definition of an AZD. We note that since $`𝒪_{\stackrel{~}{X}}(\nu P_m)`$ is globally generated on $`\stackrel{~}{X}`$, for every $`\nu 1`$ we have the injection
$$𝒪_{\stackrel{~}{X}}(\nu P_m)p^{}(𝒪_X(m\nu K_X)(h^{m\nu })).$$
Hence there exists a natural morphism
$$H^0(\stackrel{~}{X}_i,𝒪_{\stackrel{~}{X}_i}(\nu P_{m,i}))H^0(X_i,𝒪_{X_i}(m\nu K_X)(h^{m\nu })/\text{tor})$$
for every $`\nu 1`$. This morphism is clearly injective. This implies that
$$\mu _im^{n_i}\mu (\stackrel{~}{X}_i,P_{m,i})$$
holds. Since $`P_{m,i}`$ is nef and big on $`X_i`$ we see that
$$\mu (\stackrel{~}{X}_i,P_{m,i})=P_{m,i}^{n_i}$$
holds. Hence
$$\mu _im^{n_i}P_{m,i}^{n_i}$$
holds. This implies that
$$\mathrm{deg}\mathrm{\Phi }_{mK_X}(X_i)\mu _im^{n_i}$$
holds. Q.E.D.
## 4 Proof of Theorem 1.2
Let
$$XX_1X_2\mathrm{}X_rX_{r+1}=\{x\}R_x^{}(\text{resp.}\{x^{}\}R_x)$$
be a very general stratification constructed as in the last section.
Suppose that $`(_{\mathrm{}=1}^n\sqrt[\mathrm{}]{2}\mathrm{}+1)K_X`$ does not give a birational rational map from $`X`$ into a projective space. Then by Proposition 3.1, we see that
$$\underset{i}{\mathrm{max}}\frac{\alpha _i}{\sqrt[n_i]{2}n_i}1$$
holds. Let $`k`$ be the number such that
$$\frac{\alpha _k}{\sqrt[n_k]{2}n_k}=\underset{i}{\mathrm{max}}\frac{\alpha _i}{\sqrt[n_i]{2}n_i}.$$
By Lemma 3.9 we see that
$$\mu _k<1$$
holds. We set
$$\alpha :=\underset{i=0}{\overset{r}{}}\alpha _i+1.$$
Now we see that by Lemma 3.10 and Lemma 3.9
$$\text{deg}\mathrm{\Phi }_{\alpha K_X}(X_k)((\underset{\mathrm{}=1}{\overset{n}{}}\frac{\sqrt[\mathrm{}]{2}\mathrm{}}{\sqrt[n_k]{2}n_k})\alpha _k+1)^{n_k}\mu _k$$
$$((\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1)\frac{1}{\sqrt[n_k]{\mu _k}})^{n_k}\mu _k$$
$$(\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1)^{n_k}$$
hold. Since such $`\{X_k\}`$ form a dominant family of subvarieties on $`X`$ by the construction of the stratifications, this completes the proof of Theorem 1.2.
## 5 Proof of Theorem 1.3
Let $`X`$ be a smooth projective $`n`$-fold of general type and suppose that $`(_{\mathrm{}=1}^n\sqrt[\mathrm{}]{2}\mathrm{}+1)K_X`$ does not give a birational embedding. Then by Theorem 1.2 and its proof, there exists some $`1dn`$ such that there exists a dominant family of subvarieties
$$\varpi :𝒳_kS_k$$
which parametrizes the strata $`X_k`$ of a general stratification
$$XX_1X_2\mathrm{}X_rX_{r+1}=\{x\}R_x^{}(\text{resp.}\{x^{}\}R_x)$$
such that for $`\alpha :=_{i=0}^r\alpha _i+1`$
$$\mathrm{deg}\mathrm{\Phi }_{\alpha K_X}(X_k)(\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1)^d^{}$$
holds where $`d^{}=dimX_k`$. Let
$$p:𝒳_kX$$
be the natural morphism. Inductively we define a sequence of (possibly reducible) subvarieties $`F_i(i0)`$ by
$$F_0=X_k([X_k]S_k)$$
and for $`i0`$
$$F_{i+1}=\text{the closure of}p(\varpi ^1(\varpi (\pi ^1(\text{the generic points of }F_i)))).$$
Then $`\{F_i\}_{i0}`$ is increasing and by the Noetherian property, we see that there exists some $`\mathrm{}0`$ such that
$$dimF_{\mathrm{}}=dimF_{\mathrm{}^{}}$$
for every $`\mathrm{}^{}\mathrm{}`$. Let $`F_{\mathrm{},0}`$ be a maximal dimensional component of $`F_{\mathrm{}}`$.
If we start from a general $`[X_k]S_k`$ and choose $`F_{\mathrm{},0}`$ properly, we may assume that $`\{F_{\mathrm{},0}\}`$ form a family. We note that possibly $`F_{\mathrm{},0}`$ may not be determined only by $`X_k`$ because of a monodoromy phenomenon. Let
$$\varpi _0:𝒰_0T_0$$
be the family of such $`\{F_{\mathrm{},0}\}`$. Then it is again dominant. Let
$$p_0:𝒰_0X$$
be the natural morphism. We see that $`p_0`$ is birational, since for a general $`xX`$, $`p_0^1(x)`$ is a point (otherwise it contradicts to the maximality of $`dimF_{\mathrm{},0}`$). Hence $`\varpi _0`$ induces a rational fibration structure
$$f:X\mathrm{}Y.$$
Let $`F`$ be a general fiber of $`f`$. To completes the proof of Theorem 1.3 we need the following lemma.
###### Lemma 5.1
$$\mathrm{deg}\mathrm{\Phi }_{\alpha K_X}(F)(\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1)^{d^2}d^d$$
Proof. By taking a suitable modification of $`X`$, we may assume that the following conditions are satisfied :
1. $`\mathrm{\Phi }_{\alpha K_X}`$ is a morphism on $`X`$,
2. there exists a regular fibration
$$f:XY$$
induced by $`p_0:𝒰_0X`$ as above.
Let $`H`$ be the free part of $`\alpha K_X`$. We set
$$a:=(\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1)^d^{}.$$
Let $`F`$ be a general fiber of $`f`$. Suppose that
$$\mathrm{deg}\mathrm{\Phi }_{\alpha K_X}(F)>a^dd^d$$
holds.
###### Lemma 5.2
Let us fix an arbitrary point $`x_0`$ on $`F`$. Then for a sufficiently large $`m`$ there exists a section
$$\sigma \mathrm{\Gamma }(F,𝒪_F(mH))\{0\}$$
such that
$$\text{mult}_{x_0}(\sigma )>mad+1$$
holds.
Proof. Since
$$dim𝒪_F/_{x_0}^m=\left(\begin{array}{c}d+m1\\ d\end{array}\right)=\frac{1}{d!}m^d+O(m^{d1})$$
and
$$dimH^0(F,𝒪_X(mH))=\frac{1}{d!}a^dd^dm^d+O(m^{d1})$$
hold, the lemma is clear. Q.E.D.
Let
$$\varpi _F:𝒳_k(F)S_k(F)$$
be the family of the strata $`X_k`$ contained in $`F`$. Then there exists a subvariety $`S_k^{}(F)`$ such that
1. $`dimS_k^{}(F)=dimFdimX_k`$,
2. We set $`\stackrel{~}{F}:=\varpi _F^1(S_k^{}(F))`$. Then $`p_{\stackrel{~}{F}}:\stackrel{~}{F}F`$ is generically finite.
Then
$$\stackrel{~}{\varpi }_F:\stackrel{~}{F}S_k^{}(F)$$
is an algebraic fiber space. We set
$$W_j=\{\stackrel{~}{x}\stackrel{~}{F}\text{mult}_{\stackrel{~}{x}}p_{\stackrel{~}{F}}^{}(\sigma )1+maj\}.$$
We have a decending chain of subvarieties
$$W_0W_1\mathrm{}W_dp_{\stackrel{~}{F}}^1(x_0).$$
For each $`j\{0,\mathrm{},n\}`$, we choose an irreducible component of $`W_j`$ containing $`x`$, and denote this irreducible component by $`W_j^{}`$. We may assume that these irreducible components have been chosen so that
$$W_0^{}W_1^{}\mathrm{}W_d^{}p_{\stackrel{~}{F}}^1(x_0).$$
Since this chain has length greater than $`d=dimF`$, there exists some $`j`$ such that
$$W_j^{}=W_{j+1}^{}$$
holds. We set $`W=W_j^{}`$.
Let $`𝒞`$ denote the family of irreducible curves
$$C=H_1\mathrm{}H_{d^{}1}X_k^{}$$
on $`F`$ which is obtained as the intersection of $`(d^{}1)`$-members $`H_1,\mathrm{}H_{d^{}1}`$ of $`H`$ and a strata $`X_k^{}([X_k^{}]S_k)`$ which is contained in $`F`$. We note that for a general member $`C`$ of $`𝒞`$, the inverse image of $`C`$ in the normalization of $`X_k^{}`$ is smooth. Hence a general member of $`𝒞`$ is immersed in $`X`$ that means the differential of the natural morphism from the normalization of $`C`$ to $`X`$ is nowhere vanishing. By the construction $`𝒞`$ determines a dominant family of curves $`\stackrel{~}{𝒞}`$ on $`\stackrel{~}{F}`$. We note that by the construction of $`F`$ a member $`X_k^{}`$ of $`S_k`$ intersects $`F`$, then $`X_k^{}`$ should be contained in $`F`$. Now we note have that
$$HCa$$
holds.
Now we quote the following two theorems (the satements are slightly generalized, but the proofs are completely same).
###### Theorem 5.1
(\[13, p. 686,Theorem 2\]) Fix a positve integers $`\mathrm{}`$ and $`k`$. Let $`f:M\mathrm{\Delta }^k`$ be a smooth family of irreducible curves. Let $`L`$ be a holomorphic line bundle on $`M`$ such that the restriction of $`L`$ to $`f^1(0)`$ has degree $`\mathrm{}`$ and let $`sH^0(M,𝒪_M(L))`$. Let $`V_j(s)`$ denote the complex subspace of $`M`$ consinsting precisely those points at which the vanishing order of $`s`$ is at least $`j`$.
Then either
$$f^1(0)V_{j+\mathrm{}}(s)=\mathrm{}$$
or
$$f^1(0)V_j(s)$$
holds.
###### Theorem 5.2
(\[13, p.686,Theorem 3\]) Let $`M`$ be a complex manifold, let $`L`$ be a holomorphic line bundle on $`M`$ and let $`sH^0(M,𝒪_M(L))`$. Let $`𝒩`$ be an irreducible family of immersed curves thich is free and set $`\mathrm{}:=\mathrm{deg}_NL`$, where $`N`$ is a member of $`𝒩`$. For any integer $`j`$, either
$$NV_{j+\mathrm{}}(s)=\mathrm{}$$
or
$$NV_j(s)$$
holds.
###### Lemma 5.3
If $`\stackrel{~}{C}`$ be an immersed member of $`\stackrel{~}{𝒞}`$ which intersects $`W_j`$, then $`\stackrel{~}{C}`$ is contained in $`W_{j1}`$.
Proof. We note that $`\stackrel{~}{C}`$ is free by the construction. Then the lemma follows from Theorem 5.2. Q.E.D.
Now we fix a general $`W`$. Then by the definition of $`𝒞`$ and Lemma 5.1 implies that if for a fiber $`\stackrel{~}{X}_k`$ of
$$\stackrel{~}{\varpi }_F:\stackrel{~}{F}S_k^{}(F)$$
$$\stackrel{~}{X}_kW\mathrm{}$$
holds, then
$$\stackrel{~}{X}_kW$$
holds. We note that $`W`$ is defined by the pullback of the section of $`mH`$ on $`F`$. By the inductive construction of $`F`$, we see that if we take $`W`$ general, $`W=\stackrel{~}{F}`$ holds. This is the contradiction. Hence we see that
$$\mathrm{deg}\mathrm{\Phi }_{\alpha K_X}(F)a^dd^d$$
holds. Since
$$a^dd^d(\underset{\mathrm{}=1}{\overset{n}{}}\sqrt[\mathrm{}]{2}\mathrm{}+1)^{d^2}d^d$$
holds, this completes the proof of Theorem 1.3.
Author’s address
Hajime Tsuji
Department of Mathematics
Tokyo Institute of Technology
2-12-1 Ohokayama, Megro 152
Japan
e-mail address: tsuji@math.titech.ac.jp
|
warning/0006/hep-th0006219.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Recently there has been much interest in the study of non-commutative (nc) geometry in the context of quantum field theory (including gauge theories) and string theory There are several directions of exploration. To begin with let us cite some examples of nc geometry.
1. Witten’s open string field theory is formulated with a associative, non-commutative product .
2. The BFSS matrix model formulation of M-theory. Here space coordinates are matrices and hence there is a natural associative, non-commutative product. If we express the Hamiltonian of this model in terms of variables that describe 2-branes the associative, non-commutative matrix algebra is reflected in the 2-brane world volume being a 2-dim. phase space. .
3. World volume theories of brane systems in the presence of certain moduli fields like the Neveu-Schwarz $`B_{NS}`$ (in the Seiberg-Witten limit) become field theories where ordinary multiplication of fields is replaced by the Moyal star product in a phase space. .
4. In the study of the large $`k`$ limit of the $`SU(2)`$ WZW model describing open strings moving on a group manifold one encounters the coadjoint orbits of $`SU(2)`$. These are 2-spheres with a natural symplectic form (called fuzzy 2-spheres.) The radius of the coadjoint orbit is $`jk/2`$.
5. The fuzzy 2-sphere also occurs in the Polchinski-Strassler description of bulk geometry in the presence of relevant perturbations of the $`D_3`$ brane system .
6. The c=1 matrix model is exactly formulated as a particular representation of the $`W_{\mathrm{}}`$ algebra, that comes from 1-dim. non-relativistic fermions. The coadjoint representation is carried by the phase space density $`u`$ of fermions constrained by the equation $`uu=u`$, which is a quantum statement of fermi statistics . Collective field theory describes the classical geometry limit where the constraint involves the ordinary product $`u^2=u`$.
7. The gauge invariant description of 2-dim. $`U(N)`$ QCD with fermions also leads to a specific representation of the $`W_{\mathrm{}}`$ algebra. Here too the coadjoint representation is described by the equation $`MM=M`$ where M represents the gauge invariant Wilson line between two quarks. The Regge trajectory of mesons appear as solutions of the small fluctuation equation around the large N classical solution of $`MM=M`$ .
The examples cited above share one important feature in common: they are theories whose fields are valued in a $`\mathrm{𝑞𝑢𝑎𝑛𝑡𝑢𝑚}\mathrm{𝑝ℎ𝑎𝑠𝑒}\mathrm{𝑠𝑝𝑎𝑐𝑒}`$ rather than on a manifold. It is presently not clear how Witten’s string field theory fits into this framework. <sup>1</sup><sup>1</sup>1Witten has recently made progress in this direction. See the note added at the end of this paper.
The quantum phase space is specified by a operator algebra. By virtue of this, these fields become operators and act in an appropriate Hilbert space. If one considers the coadjoint representation of the operator algebra then one can have a correspondence between the operator valued fields and classical functions on the coadjoint orbit. For the simplest example of the Heisenberg algebra the coadjoint orbit corresponds to the familiar phase space. The composition law for functions on the phase space is the Moyal star product. This point is briefly reviewed in the next section and is called the Weyl-Moyal correspondence.
Besides the fact that space-time is non-commutative at the fundamental level, as suggested by the Matrix model formulation of M-theory, (see also ) this new type of field theory has many interesting properties. One very significant property is the IR/UV connection . Such a connection is indeed novel and seems to signal a breakdown of decoupling. As has been suggested it may be useful in understanding the cosmological constant problem . Another important property of a nc theory is the role it plays in the resolution of singularities . Another use can be in understanding the large N limit of gauge theories.
With these and other applications in mind it is important to understand these theories in various ways. In this note we formulate a gauge theory on a quantum phase space. Our main point is to give a formulation directly in the language of operators, using simple rules that physicists are familiar with. For simplicity we discuss the Heisenberg algebra and show that the operator formulation of the gauge theory corresponds to a formulation on the coadjoint orbit (phase space) in terms of the Moyal star product.
In section 2. we introduce the basic notion of a derivative operator which can translate operators. We also introduce the notion of operator valued forms, the exterior derivative, the wedge product and the Dirac operator. In section 3. we discuss the Weyl-Moyal correspondence. In section 4. we introduce the notion of parallel transport of operators using the Wilson line. We also introduce the corresponding gauge field , gauge transformations, field strength, action and instanton number. In section 5. we define the Wilson loop. In section 6. comment on the quantum theory. In section 7. we discuss the Higgs mechanism and in section 8. we make some remarks on the large N master field.
## 2 The Derivative Operator on a Quantum Phase Space
A quantum phase space is specified by an operator algebra. This algebra can have finite or infinite number of generators. Examples of finite number of generators are the Heisenberg algebra and Lie algebras of compact and non-compact groups. The algebras with infinite number of generators are eg. the Kac-Moody algebra and the Virasoro algebra. One can and should include super-algebras to this list.
To simplify matters we restrict ourselves to the Heisenberg algebra.
$$[X_i,X_j]=iI\theta _{i,j}$$
(1)
where , $`I`$ is the identity operator, $`i,j=1,2,..,2d`$, and $`\theta _{i,j}`$ is a real anti-symmetric, invertible matrix, with inverse $`\theta _{i,j}^1`$. We have to specify the Hilbert space on which these operators act. For our present purposes we can take this to be the space of delta-function normalizable functions in d-dimensions.
Using the basic operators (1) we can construct other operators in terms of polynomials of $`X_i`$ over the complex numbers. Instead of a polynomial basis we can more fruitfully use the Weyl basis defined by the exponential operators
$$g(\alpha )=\mathrm{exp}i\alpha _iX_i$$
(2)
In this way we can introduce complex, self-adjoint and unitary operators in the Hilbert space.
Since we would like to develop an operator calculus, the first thing that we should do is to define the derivative operator and the notion of translations in the space of operators. We define the derivative by,
$$\overline{}_i=\theta _{i,j}^1adX_j$$
(3)
where the adjoint action is defined by $`(adA)B=[A,B]`$. This derivative operator has a number of important properties which we list:
1. $`\overline{}_i`$ is anti-hermitian and linear,
$`\overline{}_i^{}`$ $`=`$ $`\overline{}_i`$
$`\overline{}_i(a_1𝒪_1+a_2𝒪_2)`$ $`=`$ $`a_1\overline{}_i𝒪_1+a_2\overline{}_i𝒪_2`$
2. $`\overline{}_i`$ satisfies the Leibniz rule,
$$\overline{}_i(𝒪_1𝒪_2)=(\overline{}_i𝒪_1)𝒪_2+𝒪_1(\overline{}_i𝒪_2)$$
(5)
3. $`\overline{}_i`$ commute amongst themselves,
$$[\overline{}_i,\overline{}_j]=0$$
(6)
4. The commutative property of the derivative enables us to introduce the notion of an exterior derivative that acts on operator valued forms. The operator valued n-form and its exterior derivative are defined by,
$`O^{(n)}`$ $`=`$ $`{\displaystyle 𝒪_{i_1,..,i_n}dy_{i_1}}\mathrm{}dy_{i_n}`$
$`dO^{(n)}`$ $`=`$ $`{\displaystyle \overline{}_{i_1}𝒪_{i_2,..,i_n}dy_{i_1}}\mathrm{}dy_{i_{n+1}}`$
In the above $`dy_i`$ are real 1-forms. The commutative property of the derivative clearly implies that $`d^2=0`$.
Using the above definition of the operator valued n-form we can introduce the notion of a non-commutative wedge product
$$O^{(n)}\overline{}O^{(m)}=C_{n,m}𝒪_{i_1,.,i_n}𝒪_{i_{n+1}},.,i_{n+m}dy_{i_1}\mathrm{}dy_{i_{n+m}}$$
(8)
$`C_{n,m}`$ is a normalization constant. The above definition of the wedge product is associative. One can also define $`{}_{}{}^{}O_{}^{(n)}`$ the Poincare dual of $`O^{(n)}`$ in the standard fashion using the totally antisymmetric $`ϵ`$ tensor in $`2d`$ dimensions.
5. $`\overline{}_i`$ is the generator of translations in the following sense,
$$\mathrm{exp}ia_i\overline{}_i𝒪(X)=𝒪(X+Ia)$$
(9)
6. The integral of an operator is defined by its trace in the corresponding Hilbert space. Then using the trace formula $`TrA[B,C]=Tr[A,B]C`$ we have the formula for integration by parts,
$$Tr𝒪_1(\overline{}_i𝒪_2)=Tr(\overline{}_i𝒪_1)𝒪_2$$
(10)
We note that there is no ‘surface term’.
7. We can introduce the Dirac operator $`\overline{)\overline{}}=\gamma _i\overline{}_i`$, where $`\gamma _i`$ are the standard Dirac gamma matrices. Note that $`(\overline{)\overline{}})^2=\overline{}_i\overline{}_i`$.
The Landau Condition
The defining equation for the derivative operator can be understood in a more physical setting by introducing additional momentum operators $`P_i`$ and extending the algebra,
$$[P_i,P_j]=0$$
(11)
$$[P_i,X_j]=i\delta _{ij}I$$
(12)
and introducing the constraint,
$$adP_i=\overline{}_i=\theta _{i,j}^1adX_j$$
(13)
On states in the Hilbert space this equation becomes the Landau constraint ,
$$(P_i\theta _{i,j}^1X_j)|\mathrm{\Psi }>=0$$
(14)
which implies for such states the uncertainty principle,
$$\delta X_i\theta _{i,j}^1\delta X_j1$$
(15)
## 3 Weyl-Moyal Correspondence
The Weyl-Moyal (WM) correspondence is best understood in terms of the operators $`g(\alpha )=\mathrm{exp}i\alpha _iX_i`$. Using (1) these operators satisfy the Heisenberg-Weyl algebra,
$$g(\alpha )g(\beta )=\mathrm{exp}(\frac{1}{2}\theta _{i,j}\alpha _i\beta _j)g(\alpha +\beta )$$
(16)
The WM correspondence is given by
$$𝒪(X)=d^{2d}\alpha g(\alpha ,X)\stackrel{~}{O}(\alpha )$$
(17)
$`\stackrel{~}{O}`$ is the Fourier transform of $`O`$. Using (16) we can derive correspondence between the operator product and the star product,
$`𝒪_1𝒪_2`$ $`=`$ $`{\displaystyle d^{2d}\alpha g(\alpha ,X)\stackrel{~}{O}_{12}(\alpha )}`$
$`O_{12}`$ $`=`$ $`O_1O_2`$
$`O_1O_2`$ $`=`$ $`\mathrm{exp}(i{\displaystyle \frac{1}{2}}\theta _{i,j}{\displaystyle \frac{}{\xi _i}}{\displaystyle \frac{}{\eta _j}})O_1(x+\xi )O_2(x+\eta )_{\xi =\eta =0}`$ (18)
Using the WM correspondence, we can easily prove the correspondence,
$`\overline{}_i𝒪(X)`$ $``$ $`_iO(x)`$
$`𝒪(X)_1𝒪(X+Ia)_2`$ $``$ $`O(x)_1O(x+a)_2`$
## 4 Parallel Transport of Operators, Wilson Lines and Gauge Fields
In section 2. we introduced the notion of translating operators. In this section we introduce the notion of parallel transport and connection. For convenience of presentation we will suppress the $`U(N)`$ indices carried by the various operators.
Consider the set of operators $`\mathrm{\Phi }(X)`$, which transform under the right action of the group of unitary operators $`\{\mathrm{\Omega }(X)\}`$
$$\mathrm{\Phi }(X)\mathrm{\Phi }(X)\mathrm{\Omega }(X)$$
(20)
Then clearly
$$\mathrm{\Phi }(X)\mathrm{\Phi }(X+Ia)^{}$$
(21)
is not gauge invariant under the gauge transformations
$`\mathrm{\Phi }(X)`$ $``$ $`\mathrm{\Phi }(X)\mathrm{\Omega }(X)`$ (22)
$`\mathrm{\Phi }(X+Ia)`$ $``$ $`\mathrm{\Phi }(X+Ia)\mathrm{\Omega }(X+Ia)`$ (23)
The standard way to form a gauge invariant operator is to introduce the Wilson line $`U(X,X+Ia)`$, with gauge transformation
$$U(X,X+Ia)\mathrm{\Omega }(X)^{}U(X,X+Ia)\mathrm{\Omega }(X+Ia)$$
(24)
and the property that $`U(X,X)=I`$ and $`U(X,X+Ia)^{}=U(X+Ia,X)`$. The operator
$$\mathrm{\Phi }(X)U(X,X+Ia)\mathrm{\Phi }(X+Ia)^{}$$
(25)
is gauge invariant.
Now a similar construction is possible for those operators $`\mathrm{\Psi }(X)`$ which transform under the left action of the group of unitary operators,
$$\mathrm{\Psi }(X)\mathrm{\Omega }(X)\mathrm{\Psi }(X)$$
(26)
In this case the gauge invariant operators are given by,
$$\mathrm{\Psi }(X)U(X,X+Ia)^{}\mathrm{\Psi }(X+Ia)^{}$$
(27)
The Connection:
Using the Wilson line for infinitesimal $`a_i=ϵ_i`$ we can introduce the definition of the operator valued connection,
$$U(X,X+Iϵ)=\mathrm{exp}(i𝒜_i(X)ϵ_i)$$
(28)
The operator $`𝒜_i(X)`$ is Hermitian and the gauge transformation of the Wilson line implies the gauge transformation of the operator valued gauge field.
$$𝒜_i(X)\mathrm{\Omega }(X)^{}(𝒜_i(X)i\overline{}_i)\mathrm{\Omega }(X)$$
(29)
Covariant Derivative and Field Strength:
The operator covariant derivative and the field strength are defined by,
$`𝒟_i`$ $`=`$ $`i\overline{}_i+𝒜_i(X)`$ (30)
$`_{ij}(X)`$ $`=`$ $`i[𝒟_i,𝒟_j]`$ (31)
Both the covariant derivative and the field strength are gauge covariant and the Jacobi identity for the covariant derivative implies the Bianchi identity for the field strength,
$$[𝒟_i,_{jk}]+[𝒟_j,_{ki}]+[𝒟_k,_{ij}]=0$$
(32)
The Action and Equations of Motion:
The gauge invariant action is given by
$`S`$ $`=`$ $`{\displaystyle \frac{1}{4g^2}}Tr((X)\overline{}^{})`$ (33)
$`=`$ $`{\displaystyle \frac{1}{4g^2}}Tr(_{ij}(X)_{ij}(X))`$
In the above we have chosen the euclidean metric. From the above action we can easily derive the equations of motion by requiring the action to be stationary w.r.t the variation $`𝒜_i(X)𝒜_i(X)+\delta 𝒜_i(X)`$,
$$𝒟_i_{i,j}(X)=0$$
(34)
We can also define the nc version of the instanton number
$`I`$ $`=`$ $`Tr((X)\overline{})`$ (35)
$`=`$ $`Tr(_{i,j}(X)_{k,l}(X)ϵ_{ijkl})`$
(In the above formulas the trace can also includes a trace over $`U(N)`$.)
It is also possible to introduce the following operator current,
$$𝒥_i=ϵ_{ijkl}(𝒜_j\overline{}_k𝒜_l+i\frac{2}{3}𝒜_j𝒜_k𝒜_l)$$
(36)
so that (35) can be written as,
$$I=4Tr(\overline{}_i𝒥_i)$$
(37)
Using the WM correspondence it is easy to prove that the operator formulation given above goes over into a gauge theory formulated in terms of a real connection, and the star product, e.g.
$`𝒜_i(X)`$ $``$ $`A_i(x)`$
$`_{ij}(X)`$ $``$ $`_iA_j_jA_i+i(A_iA_jA_jA_i)`$
In the operator formulation the gauge group is generated by all unitary operators. The generators of this gauge group are the set of all Hermitian operators: $`\{H(X)\}`$, and $`\{\mathrm{\Omega }(X)=\mathrm{exp}i(X)\}`$. Expressing the exponential as a series we can easily obtain the correspondence for the gauge group,
$$\mathrm{exp}i(X)(1+H(x)+\frac{1}{2}H(x)H(x)+\frac{1}{6}H(x)H(x)H(x)\mathrm{})$$
(39)
## 5 The Wilson Loop
Let us now present the expression for the Wilson loop operator. Consider a curve $`\mathrm{\Gamma }`$ in $`R^{2d}`$ and divide it into $`n\mathrm{}`$ infinitesimal segments each denoted by a tangent vector $`ϵ_i^m`$. $`i=0,1,2..n`$ and $`ϵ_i^0=0`$. Since the loop is closed we have $`i=1^nϵ_i^m=0`$. The Wilson loop is composed of a product of Wilson lines around the curve,
$`W(\mathrm{\Gamma })`$ $`=`$ $`Tr{\displaystyle \underset{m=0}{\overset{n}{}}}U(X+Iϵ^m,X+Iϵ^{m+1})`$ (40)
$`=`$ $`Tr{\displaystyle \underset{m=0}{\overset{n}{}}}\mathrm{exp}i(𝒜_i(X+Iϵ^{m+1})(ϵ_i^mϵ_i^{m+1}))`$
$`W(\mathrm{\Gamma })`$ is gauge invariant. It would be useful to see the connection of this formulation with the reduced model of lattice gauge theory .
## 6 The Quantum Theory
Until now we have been dealing with a classical theory whose fields are defined on a ‘quantum phase space’ specified by the matrix $`\theta _{i,j}`$. Quantization of this theory consists of studying fluctuations whose strength is controlled by the gauge coupling. One quantization procedure appeals to the WM correspondence and gives a path integral prescription in which one integrates over the histories of the gauge field $`A_i(x)`$. Another approach to quantization is to quantize the matrix elements of $`𝒜_i(X)`$ in a coherent state basis .
## 7 The Higgs Mechanism
We now discuss ‘matter fields’ in the nc gauge theory. For simplicity we discuss matter fields $`\mathrm{\Psi }(X)`$ with gauge transformation (only left action) $`\mathrm{\Psi }(X)\mathrm{\Omega }(X)\mathrm{\Psi }(X)`$. The gauge invariant action is given by
$$S=\frac{1}{4g^2}Tr(_{i,j}(X)_{i,j}(X))+\frac{1}{2}Tr(𝒟_i\mathrm{\Psi })^{}(𝒟_i\mathrm{\Psi })+\frac{1}{4}Tr(\mathrm{\Psi }^{}\mathrm{\Psi }a^2I)^2$$
(41)
To discuss the Higgs mechanism, we write $`\mathrm{\Psi }=U`$, where $`U`$ is unitary and $`H`$ is Hermitian, and perform the gauge transformation to the unitary gauge $`\mathrm{\Psi }U^{}\mathrm{\Psi }`$. In this gauge the potential term becomes $`V=\frac{1}{4}Tr(^2a^2I)^2`$ and the ground state is a solution of the operator equation,
$$^3=a^2$$
(42)
or equivalently
$$HHH=a^2H$$
(43)
If we assume that $`^1`$ exists then (42) reduces to
$$^2=a^2I$$
(44)
Such operator equations were originally discussed in They have also been recently discussed in the context of soliton solutions of nc field theories . See also for subsequent applications. The minimum energy solution is given by $`=|a|I`$, where we have absorbed a possible sign ambiguity by a gauge transformation. This leads to a ‘mass term’ for the gauge field $`|a|^2Tr(𝒜_i𝒜_i)`$. It would be interesting to look for vortex like solutions in these models.
## 8 The large N Master Field
Let us now consider the nc gauge theory with additional color gauge group $`U(N)`$. In ref. the perturbation expansion in the maximally non-commutative regime ($`\theta \mathrm{}`$ ) was discussed. In this limit the leading term in the perturbation expansion consists of planar graphs which have no theta or N dependence except for an overall phase involving $`\theta `$ and a multiplicative factor of $`N^2`$. Hence the problem of summing planar diagrams of the nc gauge theory is mapped onto the problem of solving the nc gauge theory in the $`\theta \mathrm{}`$ limit.
Let us choose $`\theta _{ij}=\theta \left(\begin{array}{cc}0& 1_d\\ 1_d& 0\end{array}\right)`$ and write the gauge theory in terms of the scaled operators $`\frac{X_i}{\sqrt{\theta }}`$ or after the WM correspondence, in terms of the scaled co-ordinates $`\frac{x_i}{\sqrt{\theta }}`$. If we require that the gauge field has the transformation $`A_i(\frac{x_i}{\sqrt{\theta }})=\sqrt{\theta }A_i(x)`$, then the action becomes
$$S=\frac{\theta ^{\frac{2d4}{2}}}{4g^2}d^{2d}xtr(F_{ij}F_{ij})$$
(45)
and the path integral is given by
$$Z=𝒟A_i(x)e^S$$
(46)
For $`2d>4`$ and the path integral is evaluated in the $`\theta \mathrm{}`$ limit by the saddle point $`\frac{\delta S}{\delta A_i(x)}=_iF_{ij}=0`$. Reverting back to the operator formalism, for $`2d>4`$, is given by four $`\theta \mathrm{}`$ operators $`𝒜_i(x)`$ which satisfy the n.c YM equations
$$𝒟_i_{ij}=0$$
(47)
It is reasonable that (38) are equations for the $`U(\mathrm{})`$ master field for $`2d>4`$.
The situation in the most interesting dimension is certainly more complicated and the Schwinger-Dyson equations approach may help. Finally it would be interesting to make a connection of the large N limit of nc gauge theory with non-commutative probability theory as applied to the problem of the large N limit by Gopakumar and Gross .
Note added:
After this work was completed we received where an approach similar to ours has been discussed. However the derivative operator discussed in this paper (equation 2.7 ) does not necessarily commute in different directions. Gross and Nekrasov have also presented the basic ingredients of the nc gauge theory in the operator formulation. Their derivative operator is identical to ours. Recently, Witten has shown the emergence of the Moyal product in the large B-field limit of string field theory.
It turns out that the geometrical construction of the non-commutative gauge theory presented here can be shown to be equivalent to the perturbation of the IKKT model around the brane configuration defined by (1). This point has been made in . We would like to thank R. Szabo and A. Dhar for pointing this out to us.
Acknowledgement:
One of us (SRW) would like to thank the CERN theory division for a visit during which this work was done. We would like to thank Farhad Ardalan, Hessam Arfei, Ali Chamseddine, Avinash Dhar, Gautam Mandal, Dileep Jatkar, Sandip Trivedi and K.P. Yogendran for useful discussions.
|
warning/0006/hep-ph0006125.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Quark production in electron positron annihilation provides us with additional information about the constants appearing in the standard model of the electroweak and strong interactions. This in particular holds for heavy flavour production. An example is the electroweak mixing angle $`\theta _W`$ which has been very accurately extracted from the forward-backward asymmetry measured for bottom quarks at LEP and SLC. Besides a more accurate determination of these constants heavy quark production can also reveal some signatures of new physics. They might be observable when linear colliders are built which are giving access to much larger energies than those available until now. These energies are large enough to observe top anti-top production which will be one of the issues under study. In particular of interest is the polarization of this quark because the spin can be easily measured from its decay products. This is possible because the mass of the top-quark is so big that it can decay electroweakly before it undergoes hadronization. The measurement of the polarization provides us with new tests of the standard model and it might even signal new interactions which are not predicted by this model. An example is given in where the effect of anomalous chromoelectric couplings of the gluon to the top quark on the longitudinal polarization is studied. Another important quantity is the polarization which is perpendicular to the plane spanned by the momenta of the incoming electron and outgoing quark. This so called normal polarization is a time reversal odd observable which has implications for the observation of CP-violation in the neutral current sector. Polarized heavy quark production has been studied on the Born level in , , . In most of these calculations the top quark spin is decomposed in the helicity basis. However one can also make other choices like the beamline basis and the off-diagonal basis. In one has shown that choosing the off-diagonal basis the top and anti-top quarks are produced in one unique spin configuration only which depends on the helicities of the incoming leptons. QCD corrections to quark production have been computed in several papers , -. The previous references deal with the production of heavy quarks only. For the decay process, among which the spin density matrix, see , . The corrections in the case of massive quarks are rather complicated even in first order. In the case of the helicity basis the most of them are done by one group only see - except for a few corrections which will be presented in this paper. The QCD corrections in the case of the off-diagonal basis have been calculated in as far as the soft and virtual gluon corrections are concerned. However the contribution due to hard gluon bremsstrahlung has not been calculated in this basis. Therefore the authors in choose an alternative by putting one of the quarks in a special spin configuration whereas the other (anti-) quark and the gluon were taken to be inclusive. Furthermore one has neglected the width of the Z-boson and the normal polarization was not considered. Because of the importance of polarized quark production one should have an independent check on the calculations in the literature. Therefore we want to repeat the computations done in - and include some order $`\alpha _s`$ corrections which were not considered before. Because of the complexity of the expressions for the first order corrections we try to write the results as compact as possible. The procedure is akin to the one followed in in which the forward-backward asymmetry and the shape parameter could be written in a compact form. The same was also achieved for the longitudinal and transverse cross section in . The main problem is to reduce the number of Spence functions because they lead to unnecessarily long expressions. Exploiting several relations like the Hill-identity one can reduce them to a minimum. Furthermore we will present the cross section in such a way that it holds for the longitudinal, transverse and normal polarization at the same time. Finally we concentrate on the second order corrections to the longitudinal polarization of massless quarks. They can be derived from the first moment of the timelike coefficient functions computed in -. In the case of unpolarized scattering , the massless quark approach can be also applied to heavy quarks provided the centre of mass (CM) energy is much larger than the mass of the quark. These second order estimates are very useful as long as the exact calculations are not available because the latter are very difficult to compute. Unfortunately the massless quark approach does not work for heavy flavour production in the case of polarized scattering. In this approach the coefficient functions are computed under the condition that the quark mass m is put to zero at the start of the calculation like in , (for the first order see , , ). One can also compute these functions in the so called massive quark approach with $`m0`$ and taking the limit $`m0`$ afterwards. In the calculation of the polarized coefficient functions both approaches lead to different results contrary to what we have seen for unpolarized scattering. In other words the zero mass limit does not commute with the integrations. This anomaly is due to chiral symmetry breaking when the quark becomes massive and it was discovered for the first Bjorken sum rule which is given by the first moment of the longitudinal spin structure function $`g_1(x,Q^2)`$ in , . For timelike processes like $`e^+e^{}`$ collisions it was discussed in , , . Here we would like to emphasize that this anomaly does not affect the longitudinal polarization of the light flavours even if we regularize the collinear divergences by giving the light quark a fictitious mass. It turns out that this anomaly also appears in the quark operator matrix element so that it will be removed by mass factorization (see ). Hence we obtain the same result as for $`m=0`$ where one can use n-dimensional regularization. However for heavy flavours, where the mass has a definite meaning, a subtraction via the quark operator matrix element is not justified so that these anomalous terms are retained in the QCD corrections. Therefore in the case of heavy quarks the polarized coefficient functions have to be calculated for non zero masses. These calculations are far from trivial because one cannot apply the tricks which were so successful for the calculation in of the order $`\alpha _s^2`$ corrections to $`\sigma _{tot}(e^+e^{}\mathrm{`}hadrons^{})`$ with massive quarks. This is because the timelike coefficient functions cannot be written as the imaginary part of a forward scattering amplitude which is an essential ingredient for the computations in . The paper will be organized as follows. In the section 2 we define the kinematics and give an outline of the calculation procedure. We also present the formulae for the spin dependent and spin independent parts of the cross section. In section 3 we show the results for the radiative corrections computed up to first order in the strong coupling constant $`\alpha _s`$ for massive quarks. Here we also include the contributions which are proportional to the width of the Z-boson. The corrections will be extended for the longitudinal polarization of massless quarks up to order $`\alpha _s^2`$. In section 4 we discuss the effects of these corrections on the longitudinal, transverse and normal polarization of the detected quark. The long expressions for the order $`\alpha _s`$ corrected quark structure functions $`W_i(x,Q^2,m^2)`$ are presented in Appendix A. After integrating these functions over the Bjorken variable $`x`$ we could express them into a compact form in Appendix B.
## 2 Kinematics for polarized $`e^+e^{}`$-annihilation
Polarized heavy quark production in electron-positron annihilation is given by the following process
$`e^+(q_{e^+},\lambda _{e^+})+e^{}(q_e^{},\lambda _e^{})V(q)\mathrm{H}(p,s)+\mathrm{"}\mathrm{X}\mathrm{"},`$ (2.1)
where $`\mathrm{H}`$ denotes the heavy quark and $`\mathrm{"}\mathrm{X}\mathrm{"}`$ represents any inclusive multi-partonic state containing the heavy anti-quark and the light (anti-) quarks and gluons. Further $`\lambda _i,q_i`$ denote the spin and momenta of the incoming leptons and $`s,p`$ stand for the spin and momentum of the heavy quark $`\mathrm{H}`$ detected in the final state. If we denote the momentum of the virtual vector boson $`V`$ ($`V=\gamma ,Z`$) by $`q`$, the centre of mass energy $`Q`$ is defined by
$`q^2=Q^2=(q_{e^+}+q_e^{})^2.`$ (2.2)
The differential cross section corresponding to reaction (2.1) equals
$`d\sigma `$ $`=`$ $`{\displaystyle \frac{1}{2Q^2}}dPS{\displaystyle \underset{V_1V_2}{}}_{\mu \nu }^{(V_1V_2)}𝒫^{(V_1V_2)}M^{\mu \nu ,(V_1V_2)}`$ (2.3)
$``$ $`{\displaystyle \frac{1}{2Q^2}}dPS[_{\mu \nu }^{(\gamma \gamma )}𝒫^{(\gamma \gamma )}M^{\mu \nu ,(\gamma \gamma )}+_{\mu \nu }^{(ZZ)}𝒫^{(ZZ)}M^{\mu \nu ,(ZZ)}`$
$`+_{\mu \nu }^{(\gamma Z)}𝒫^{(\gamma Z)}M^{\mu \nu ,(\gamma Z)}+_{\mu \nu }^{(\gamma Z)}𝒫^{(\gamma Z)}M^{\mu \nu ,(\gamma Z)}].`$
Here $`dPS`$ denotes the multi-parton phase space including the heavy quark and $`_{\mu \nu }^{(V_1V_2)}`$ and $`M_{\mu \nu }^{(V_1V_2)}`$ are the leptonic and partonic matrix elements respectively. Further $`𝒫^{(V_1V_2)}`$ represents the product of the propagators corresponding to the vector bosons $`V_1`$ and $`V_2`$. The leptonic tensors are given by
$`_{\mu \nu }^{(\gamma \gamma )}`$ $`=`$ $`4\pi \alpha Q_e^2\left[(1\lambda _{e^+}\lambda _e^{})l_{\mu \nu }+(\lambda _{e^+}\lambda _e^{})\stackrel{~}{l}_{\mu \nu }\right],`$
$`_{\mu \nu }^{(ZZ)}`$ $`=`$ $`{\displaystyle \frac{4\pi \alpha }{c_w^2s_w^2}}[(2g_e^Vg_e^A(\lambda _{e^+}\lambda _e^{})+(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(1\lambda _{e^+}\lambda _e^{}))l_{\mu \nu }`$
$`+(2g_e^Vg_e^A(1\lambda _{e^+}\lambda _e^{})+(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(\lambda _{e^+}\lambda _e^{}))\stackrel{~}{l}_{\mu \nu }],`$
$`_{\mu \nu }^{(\gamma Z)}`$ $`=`$ $`{\displaystyle \frac{4\pi \alpha }{c_ws_w}}Q_e[(g_e^V(1\lambda _{e^+}\lambda _e^{})+g_e^A(\lambda _{e^+}\lambda _e^{}))l_{\mu \nu }`$ (2.4)
$`+(g_e^V(\lambda _{e^+}\lambda _e^{})+g_e^A(1\lambda _{e^+}\lambda _e^{}))\stackrel{~}{l}_{\mu \nu }],`$
with
$`l^{\mu \nu }`$ $`=`$ $`q_{e^+}^\mu q_e^{}^\nu +q_{e^+}^\nu q_e^{}^\mu q_{e^+}.q_e^{}g^{\mu \nu },`$
$`\stackrel{~}{l}^{\mu \nu }`$ $`=`$ $`iϵ^{\mu \nu \alpha \beta }q_{e^+\alpha }q_{e^{}\beta }.`$ (2.5)
Notice that the polarizations of the positron and the electron indicated by $`\lambda _{e^+}`$ and $`\lambda _e^{}`$ respectively are defined by
$`v_{\lambda _{e^+}}(q_{e^+})\overline{v}_{\lambda _{e^+}}(q_{e^+})`$ $`=`$ $`{\displaystyle \frac{1}{2}}/q_{e^+}(1+\lambda _{e^+}\gamma _5),`$
$`u_{\lambda _e^{}}(q_e^{})\overline{u}_{\lambda _e^{}}(q_e^{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}/q_e^{}(1\lambda _e^{}\gamma _5),`$ (2.6)
where the incoming leptons are taken to be massless. In the case $`\lambda _i=1`$ the leptons are polarized along the direction of their momenta. When $`\lambda _i=1`$ the leptons are polarized opposite to the direction of their momenta.
On the Born level the vertex for the coupling of the vector boson $`V`$ to the fermions will be denoted by
$`\mathrm{\Gamma }_{a,\mu }^{V,(0)}`$ $`=`$ $`i(v_a^V+a_a^V\gamma _5)\gamma _\mu ,V=\gamma ,Z,`$
$`v_a^\gamma `$ $`=`$ $`eQ_a,a_a^\gamma =0,`$
$`v_a^Z`$ $`=`$ $`{\displaystyle \frac{e}{c_ws_w}}g_a^V,a_a^Z={\displaystyle \frac{e}{c_ws_w}}g_a^A.`$ (2.7)
The electroweak coupling constants are given by
$`\alpha `$ $`=`$ $`e^2/4\pi ,c_w=\mathrm{cos}\theta _W,s_w=\mathrm{sin}\theta _W,`$
$`g_a^V`$ $`=`$ $`{\displaystyle \frac{1}{2}}T_a^3s_w^2Q_a,g_a^A={\displaystyle \frac{1}{2}}T_a^3.`$ (2.8)
The electroweak charges for the leptons are equal to
$`Q_a`$ $`=`$ $`0,T_a^3={\displaystyle \frac{1}{2}},a=\nu _l,\overline{\nu }_l,l=e,\mu ,\tau ,`$
$`Q_a`$ $`=`$ $`1,T_a^3={\displaystyle \frac{1}{2}}a=e^{},\mu ^{},\tau ^{},`$ (2.9)
and for the quarks we obtain
$`Q_a`$ $`=`$ $`{\displaystyle \frac{2}{3}},T_3^a={\displaystyle \frac{1}{2}},a=u,c,t,`$
$`Q_a`$ $`=`$ $`{\displaystyle \frac{1}{3}},T_3^a={\displaystyle \frac{1}{2}},a=d,s,b.`$ (2.10)
The squared propagators $`𝒫^{(V_1V_2)}`$ are given by
$`𝒫^{\gamma \gamma }`$ $`=`$ $`{\displaystyle \frac{1}{Q^4}},𝒫^{ZZ}={\displaystyle \frac{1}{(Q^2M_Z^2)^2+M_Z^2\mathrm{\Gamma }_Z^2}},`$
$`𝒫^{\gamma Z}`$ $`=`$ $`{\displaystyle \frac{(Q^2M_Z^2)+iM_Z\mathrm{\Gamma }_Z}{Q^2\{(Q^2M_Z^2)^2+M_Z^2\mathrm{\Gamma }_Z^2\}}},`$ (2.11)
where we have introduced a finite width $`\mathrm{\Gamma }_Z`$ for the Z-boson.
The differential cross section for the inclusive reaction Eq. (2.1) can be written as
$`{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}`$ $`=`$ $`{\displaystyle \frac{1}{2Q^2}}{\displaystyle _\sqrt{\rho }^1}𝑑x{\displaystyle \underset{(V_1V_2)}{}}_{\mu \nu }^{(V_1V_2)}𝒫^{(V_1V_2)}W^{\mu \nu ,(V_1V_2)},`$
$`d\mathrm{\Omega }`$ $`=`$ $`d\mathrm{cos}\theta d\varphi ,x={\displaystyle \frac{2pq}{Q^2}},\rho ={\displaystyle \frac{4m^2}{Q^2}}.`$ (2.12)
Here $`\theta `$ is the polar angle of the outgoing quark with respect to the beam direction of the electron and $`m`$ denotes the mass of the heavy quark. The partonic tensor can be expressed into structure functions $`W_i^{(V_1V_2)}`$ ($`i=111`$) in the following way
$`W^{\mu \nu ,(V_1V_2)}`$ $`=`$ $`{\displaystyle \frac{N_c}{8\pi ^2}}Q^2(g^{\mu \nu }W_1^{(V_1V_2)}+{\displaystyle \frac{p^\mu p^\nu }{Q^2}}W_2^{(V_1V_2)}+mg^{\mu \nu }{\displaystyle \frac{sq}{Q^2}}W_3^{(V_1V_2)}`$ (2.13)
$`+mp^\mu p^\nu {\displaystyle \frac{sq}{Q^4}}W_4^{(V_1V_2)}+m{\displaystyle \frac{p^\mu s^\nu +s^\mu p^\nu }{Q^2}}W_5^{(V_1V_2)}`$
$`+{\displaystyle \frac{1}{Q^2}}iϵ^{\mu \nu \alpha \beta }p_\alpha q_\beta W_6^{(V_1V_2)}+miϵ^{\mu \nu \alpha \beta }p_\alpha q_\beta {\displaystyle \frac{sq}{Q^4}}W_7^{(V_1V_2)}`$
$`+{\displaystyle \frac{m}{Q^2}}iϵ^{\mu \nu \alpha \beta }p_\alpha s_\beta W_8^{(V_1V_2)}+{\displaystyle \frac{m}{Q^2}}iϵ^{\mu \nu \alpha \beta }q_\alpha s_\beta W_9^{(V_1V_2)}`$
$`+m{\displaystyle \frac{p^\mu s^\nu s^\mu p^\nu }{Q^2}}W_{10}^{(V_1V_2)}+{\displaystyle \frac{m}{Q^4}}i(p^\mu ϵ^{\nu \alpha \beta \gamma }p_\alpha q_\beta s_\gamma `$
$`+p^\nu ϵ^{\mu \alpha \beta \gamma }p_\alpha q_\beta s_\gamma )W_{11}^{(V_1V_2)}),`$
where $`N_c`$ denotes the number of colours (in QCD $`N_c=3`$). Finally we have to specify the spin of the quark $`\mathrm{H}(p,s)`$. In the rest frame i.e. $`p=(m,\stackrel{}{0})`$ the spin vector is given by
$`s=(0,\widehat{W}),\widehat{W}=(\widehat{W}^1,\widehat{W}^2,\widehat{W}^3),\text{with}\widehat{W}^2=1,`$ (2.14)
so that $`s^2=1`$ and $`sp=0`$. In the CM frame of the electron positron pair we introduce the notations
$`q_e^{}={\displaystyle \frac{Q}{2}}(1,0,0,1),q_{e^+}={\displaystyle \frac{Q}{2}}(1,0,0,1),`$
$`p=(E,|\stackrel{}{p}|\widehat{n})n=(\mathrm{sin}\theta \mathrm{cos}\varphi ,\mathrm{sin}\theta \mathrm{sin}\varphi ,\mathrm{cos}\theta ),`$
$`E={\displaystyle \frac{1}{2}}Qx,|\stackrel{}{p}|={\displaystyle \frac{1}{2}}Q\sqrt{x^2\rho }.`$ (2.15)
In this frame The spin four-vector of the quark can be found after an appropriate Lorentz transformation so that it becomes equal to
$`s=(|\stackrel{}{p}|{\displaystyle \frac{\widehat{n}\widehat{W}}{m}},\widehat{W}+{\displaystyle \frac{|\stackrel{}{p}|^2(\widehat{n}\widehat{W})\widehat{n}}{m(m+E)}}),\text{with}\widehat{n}^2=1.`$ (2.16)
Note that the spin four-vector in the CM frame depends on the energy of the quark and hence depends on the integration variable $`x`$ in Eq. (2). The computation of the cross section in Eq. (2) involves the contraction of the symmetric and antisymmetric parts of the leptonic tensor $`_{\mu \nu }^{(V_1V_2)}`$ with $`W^{\mu \nu ,(V_1V_2)}`$. The contraction with the symmetric part equals
$`{\displaystyle 𝑑xl_{\mu \nu }W^{\mu \nu ,(V_1V_2)}}`$ $`=`$ $`{\displaystyle \frac{N_cQ^4}{8\pi ^2}}[{\displaystyle \frac{1}{8}}(1+\mathrm{cos}^2\theta )(v_q^{V_1}v_q^{V_2}𝒯_1+a_q^{V_1}a_q^{V_2}𝒯_2)`$ (2.17)
$`+{\displaystyle \frac{1}{4}}\mathrm{sin}^2\theta \left(v_q^{V_1}v_q^{V_2}𝒯_3+a_q^{V_1}a_q^{V_2}𝒯_4\right)`$
$`+{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}+v_q^{V_2}a_q^{V_1}\right)𝒯_7+{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}v_q^{V_2}a_q^{V_1}\right)𝒯_8`$
$`+v_q^{V_1}v_q^{V_2}𝒯_9],`$
and for the antisymmetric parts we get
$`{\displaystyle 𝑑x\stackrel{~}{l}_{\mu \nu }W^{\mu \nu ,(V_1V_2)}}`$ $`=`$ $`{\displaystyle \frac{N_cQ^4}{8\pi ^2}}[{\displaystyle \frac{1}{4}}(v_q^{V_1}a_q^{V_2}+v_q^{V_2}a_q^{V_1})𝒯_5\mathrm{cos}\theta +{\displaystyle \frac{1}{4}}(v_q^{V_1}a_q^{V_2}`$ (2.18)
$`v_q^{V_2}a_q^{V_1})𝒯_6\mathrm{cos}\theta +v_q^{V_1}v_q^{V_2}𝒯_{10}+a_q^{V_1}a_q^{V_2}𝒯_{11}`$
$`+{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}+v_q^{V_2}a_q^{V_1}\right)𝒯_{12}`$
$`+{\displaystyle \frac{1}{2}}(v_q^{V_1}a_q^{V_2}v_q^{V_2}a_q^{V_1})𝒯_{13}],`$
In the expressions above $`𝒯_i`$ ($`i=16`$) represent the unpolarized structure functions. The polarized structure functions can be decomposed as follows
$`𝒯_7`$ $`=`$ $`𝒯_{7,T}W_T\mathrm{cos}\theta +W_L\left[𝒯_{7,L_1}(1+\mathrm{cos}^2\theta )+𝒯_{7,L_2}\mathrm{sin}^2\theta \right],`$
$`𝒯_8`$ $`=`$ $`𝒯_{8,T}W_T\mathrm{cos}\theta +𝒯_{8,L}W_L(1+\mathrm{cos}^2\theta ),`$
$`𝒯_9`$ $`=`$ $`𝒯_{9,N}W_N\mathrm{cos}\theta \mathrm{sin}\theta ,`$
$`𝒯_{10}`$ $`=`$ $`𝒯_{10,T}W_T+𝒯_{10,L}W_L\mathrm{cos}\theta ,`$
$`𝒯_{11}`$ $`=`$ $`𝒯_{11,T}W_T+𝒯_{11,L}W_L\mathrm{cos}\theta ,`$
$`𝒯_{12}`$ $`=`$ $`𝒯_{12,N}W_N\mathrm{sin}\theta ,`$
$`𝒯_{13}`$ $`=`$ $`𝒯_{13,N}W_N\mathrm{sin}\theta .`$ (2.19)
Notice that the polarized structure functions $`𝒯_i`$ ($`i=713`$) are expanded in terms of the longitudinal ($`W_L`$), the transverse $`W_T`$ and the normal polarization $`W_N`$ of the quark. They are defined by
$`W_L=\widehat{n}\widehat{W},W_T=\widehat{W}^3\widehat{n}\widehat{W}\mathrm{cos}\theta ,W_N=\widehat{W}^2\mathrm{cos}\varphi \widehat{W}^1\mathrm{sin}\varphi .`$ (2.20)
If we decompose the partonic structure functions as follows
$`W_i^{(V_1V_2)}`$ $`=`$ $`v_q^{V_1}v_q^{V_2}W_i^{v_q^2}+a_q^{V_1}a_q^{V_2}W_i^{a_q^2}+\left(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2}\right)W_i^{\{v_q,a_q\}}`$ (2.21)
$`+\left(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2}\right)W_i^{[v_q,a_q]},`$
one can express the quantities $`𝒯_i`$ into integrals over the partonic structure functions $`W_i`$. For the unpolarized structure functions we obtain
$`𝒯_1`$ $`=`$ $`4{\displaystyle _\sqrt{\rho }^1}𝑑xW_1^{v_q^2}(x,Q^2,m^2),`$ (2.22)
$`𝒯_2`$ $`=`$ $`4{\displaystyle _\sqrt{\rho }^1}𝑑xW_1^{a_q^2}(x,Q^2,m^2),`$ (2.23)
$`𝒯_3`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\alpha _x^2}{2}}W_2^{v_q^2}(x,Q^2,m^2)2W_1^{v_q^2}(x,Q^2,m^2)\right),`$ (2.24)
$`𝒯_4`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\alpha _x^2}{2}}W_2^{a_q^2}(x,Q^2,m^2)2W_1^{a_q^2}(x,Q^2,m^2)\right),`$ (2.25)
$`𝒯_5`$ $`=`$ $`2{\displaystyle _\sqrt{\rho }^1}𝑑x\alpha _xW_6^{\{v_q,a_q\}}(x,Q^2,m^2),`$ (2.26)
$`𝒯_6`$ $`=`$ $`2{\displaystyle _\sqrt{\rho }^1}𝑑x\alpha _xW_6^{[v_q,a_q]}(x,Q^2,m^2),`$ (2.27)
The results for the longitudinal polarized structure functions are given by
$`𝒯_{7,L_1}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\alpha _x}{2}}W_3^{\{v_q,a_q\}}(x,Q^2,m^2)\right),`$ (2.28)
$`𝒯_{7,L_2}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}dx({\displaystyle \frac{\alpha _x}{2}}W_3^{\{v_q,a_q\}}(x,Q^2,m^2)+{\displaystyle \frac{\alpha _x^3}{8}}W_4^{\{v_q,a_q\}}(x,Q^2,m^2)`$ (2.29)
$`+{\displaystyle \frac{\alpha _x}{2}}xW_5^{\{v_q,a_q\}}(x,Q^2,m^2)),`$
$`𝒯_{8,L}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\alpha _x}{2}}W_3^{[v_q,a_q]}(x,Q^2,m^2)\right),`$ (2.30)
$`𝒯_{10,L}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}dx({\displaystyle \frac{\alpha _x^2}{4}}W_7^{v_q^2}(x,Q^2,m^2){\displaystyle \frac{\rho }{4}}W_8^{v_q^2}(x,Q^2,m^2)`$ (2.31)
$`{\displaystyle \frac{x}{2}}W_9^{v_q^2}(x,Q^2,m^2)),`$
$`𝒯_{11,L}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}dx({\displaystyle \frac{\alpha _x^2}{4}}W_7^{a_q^2}(x,Q^2,m^2){\displaystyle \frac{\rho }{4}}W_8^{a_q^2}(x,Q^2,m^2)`$ (2.32)
$`{\displaystyle \frac{x}{2}}W_9^{a_q^2}(x,Q^2,m^2)).`$
Similar expressions are found for the transverse polarized structure functions
$`𝒯_{7,T}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\sqrt{\rho }}{2}}\alpha _xW_5^{\{v_q,a_q\}}(x,Q^2,m^2)\right),`$ (2.33)
$`𝒯_{8,T}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\sqrt{\rho }}{2}}\alpha _xW_5^{[v_q,a_q]}(x,Q^2,m^2)\right),`$ (2.34)
$`𝒯_{10,T}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\sqrt{\rho }}{4}}xW_8^{v_q^2}(x,Q^2,m^2){\displaystyle \frac{\sqrt{\rho }}{2}}W_9^{v_q^2}(x,Q^2,m^2)\right),`$ (2.35)
$`𝒯_{11,T}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{\sqrt{\rho }}{4}}xW_8^{a_q^2}(x,Q^2,m^2){\displaystyle \frac{\sqrt{\rho }}{2}}W_9^{a_q^2}(x,Q^2,m^2)\right),`$ (2.36)
and the results for the normal polarized structure functions can be expressed into the form
$`𝒯_{9,N}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{i\sqrt{\rho }}{8}}\alpha _x^2W_{11}^{v_q^2}(x,Q^2,m^2)\right),`$ (2.37)
$`𝒯_{12,N}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{i\sqrt{\rho }}{2}}\alpha _xW_{10}^{\{v_q,a_q\}}(x,Q^2,m^2)\right),`$ (2.38)
$`𝒯_{13,N}`$ $`=`$ $`{\displaystyle _\sqrt{\rho }^1}𝑑x\left({\displaystyle \frac{i\sqrt{\rho }}{2}}\alpha _xW_{10}^{[v_q,a_q]}(x,Q^2,m^2)\right),`$ (2.39)
with
$`\alpha _x=\sqrt{x^2\rho }.`$ (2.40)
Analogous to deep inelastic lepton hadron scattering the leading contributions to the unpolarized spin structure functions $`W_i`$ with $`i=1,2,6`$ in Eqs. (2.22)-(2.27) are of type twist two. The same holds for the polarized structure function $`W_3`$. However $`W_4`$, $`W_5`$ and $`W_7`$, $`W_8`$, $`W_9`$, $`W_{10}`$, $`W_{11}`$ contain twist two as well as twist three contributions. Notice that $`W_8`$ is due to the fact that the electroweak currents are not conserved. Furthermore the twist three part cancels in the combinations
$`xW_4+4W_5,xW_7+2W_9,`$ (2.41)
which means that the leading contributions in $`m^2/Q^2`$ to the longitudinal structure functions $`𝒯_{i,L}`$ in Eqs. (2.28)-(2.32) are of twist two only. Notice that $`W_8`$ in the equations above is multiplied by $`\rho =4m^2/Q^2`$ so that this term vanishes in the limit $`m0`$. The transverse parts $`𝒯_{i,T}`$ in Eqs. (2.33)-(2.36) receive contributions from twist two as well as twist three. The same also applies to the normal parts $`𝒯_{i,N}`$ in Eqs. (2.37)- (2.39). A second feature of the above equations is that in the limit $`m0`$ all transverse and normal parts vanish whereas the longitudinal parts $`𝒯_{i,L}`$ ($`i=7,10,11`$) and the unpolarized quantities $`𝒯_i`$ ($`i=15`$) tend to non zero values. After having carried out the integration over $`x`$ in Eq. (2) the differential cross section can be written as
$`{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W)={\displaystyle \frac{d\sigma _U}{d\mathrm{\Omega }}}+W_L{\displaystyle \frac{d\sigma _L}{d\mathrm{\Omega }}}+W_T{\displaystyle \frac{d\sigma _T}{d\mathrm{\Omega }}}+W_N{\displaystyle \frac{d\sigma _N}{d\mathrm{\Omega }}},`$ (2.42)
where $`U`$ represents the unpolarized cross section with respect to the outgoing quark. The four cross sections on the right hand side of Eq. (2.42) can be decomposed according to the vector bosons which appear in the intermediate state i.e.
$`{\displaystyle \frac{d\sigma _k}{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{})`$ $`=`$ $`N_c\alpha ^2\left[{\displaystyle \frac{d\sigma _k^{(\gamma \gamma )}}{d\mathrm{\Omega }}}+{\displaystyle \frac{d\sigma _k^{(ZZ)}}{d\mathrm{\Omega }}}+{\displaystyle \frac{d\sigma _k^{(\gamma Z)}}{d\mathrm{\Omega }}}\right],k=U,T,L,N.`$ (2.43)
The results for the photon-photon interference term can be written as <sup>1</sup><sup>1</sup>1Very often one replaces $`\alpha `$ by $`G_FM_Z^2\sqrt{2}sw^2cw^2/\pi `$ (see e.g. , ).
$`{\displaystyle \frac{d\sigma _U^{(\gamma \gamma )}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2𝒫^{\gamma \gamma }Q_e^2Q_q^2\left[(1\lambda _{e^+}\lambda _e^{})\left({\displaystyle \frac{1}{8}}(1+\mathrm{cos}^2\theta )𝒯_1+{\displaystyle \frac{1}{4}}\mathrm{sin}^2\theta 𝒯_3\right)\right],`$ (2.44)
$`{\displaystyle \frac{d\sigma _L^{(\gamma \gamma )}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2𝒫^{\gamma \gamma }Q_e^2Q_q^2\left[(\lambda _{e^+}\lambda _e^{})\mathrm{cos}\theta 𝒯_{10,L}\right],`$ (2.45)
$`{\displaystyle \frac{d\sigma _T^{(\gamma \gamma )}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2𝒫^{\gamma \gamma }Q_e^2Q_q^2\left[(\lambda _{e^+}\lambda _e^{})𝒯_{10,T}\right],`$ (2.46)
$`{\displaystyle \frac{d\sigma _N^{(\gamma \gamma )}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2𝒫^{\gamma \gamma }Q_e^2Q_q^2\left[(1\lambda _{e^+}\lambda _e^{})\mathrm{cos}\theta \mathrm{sin}\theta 𝒯_{9,N}\right].`$ (2.47)
The Z-Z interference term receives contributions from
$`{\displaystyle \frac{d\sigma _U^{(ZZ)}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2{\displaystyle \frac{𝒫^{ZZ}}{c_w^4s_w^4}}[\{2g_e^Vg_e^A(\lambda _{e^+}\lambda _e^{})+(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(1\lambda _{e^+}\lambda _e^{})\}`$ (2.48)
$`\times \left\{{\displaystyle \frac{1}{8}}(1+\mathrm{cos}^2\theta )(g_{q}^{V}{}_{}{}^{2}𝒯_1+g_{q}^{A}{}_{}{}^{2}𝒯_2)+{\displaystyle \frac{1}{4}}\mathrm{sin}^2\theta (g_{q}^{V}{}_{}{}^{2}𝒯_3+g_{q}^{A}{}_{}{}^{2}𝒯_4)\right\}`$
$`+\left\{2g_e^Vg_e^A(1\lambda _{e^+}\lambda _e^{})(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(\lambda _{e^+}\lambda _e^{})\right\}`$
$`\times \left\{{\displaystyle \frac{1}{2}}g_q^Vg_q^A\mathrm{cos}\theta 𝒯_5\right\}],`$
$`{\displaystyle \frac{d\sigma _L^{(ZZ)}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2{\displaystyle \frac{𝒫^{ZZ}}{c_w^4s_w^4}}[\{2g_e^Vg_e^A(\lambda _{e^+}\lambda _e^{})(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(1\lambda _{e^+}\lambda _e^{})\}`$ (2.49)
$`\times \left\{g_q^Vg_q^A\left((1+\mathrm{cos}^2\theta )𝒯_{7,L_1}+\mathrm{sin}^2\theta 𝒯_{7,L_2}\right)\right\}`$
$`+\left\{2g_e^Vg_e^A(1\lambda _{e^+}\lambda _e^{})+(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(\lambda _{e^+}\lambda _e^{})\right\}`$
$`\times \left\{\mathrm{cos}\theta (g_{q}^{V}{}_{}{}^{2}𝒯_{10,L}+g_{q}^{A}{}_{}{}^{2}𝒯_{11,L})\right\}],`$
$`{\displaystyle \frac{d\sigma _T^{(ZZ)}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2{\displaystyle \frac{𝒫^{ZZ}}{c_w^4s_w^4}}[\{2g_e^Vg_e^A(\lambda _{e^+}\lambda _e^{})(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(1\lambda _{e^+}\lambda _e^{})\}`$ (2.50)
$`\times \left\{g_q^Vg_q^A\mathrm{cos}\theta 𝒯_{7,T}\right\}`$
$`+\left\{2g_e^Vg_e^A(1\lambda _{e^+}\lambda _e^{})+(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(\lambda _{e^+}\lambda _e^{})\right\}`$
$`\times \{g_{q}^{V}{}_{}{}^{2}𝒯_{10,T}+g_{q}^{A}{}_{}{}^{2}𝒯_{11,T}\}],`$
$`{\displaystyle \frac{d\sigma _N^{(ZZ)}}{d\mathrm{\Omega }}}`$ $`=`$ $`Q^2{\displaystyle \frac{𝒫^{ZZ}}{c_w^4s_w^4}}[\{2g_e^Vg_e^A(\lambda _{e^+}\lambda _e^{})+(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(1\lambda _{e^+}\lambda _e^{})\}`$ (2.51)
$`\times \left\{g_{q}^{V}{}_{}{}^{2}\mathrm{cos}\theta \mathrm{sin}\theta 𝒯_{9,N}\right\}`$
$`+\left\{2g_e^Vg_e^A(1\lambda _{e^+}\lambda _e^{})(g_{e}^{V}{}_{}{}^{2}+g_{e}^{A}{}_{}{}^{2})(\lambda _{e^+}\lambda _e^{})\right\}`$
$`\times \left\{g_q^Vg_q^A\mathrm{sin}\theta 𝒯_{12,N}\right\}].`$
The photon-Z interference term consists of the following parts
$`{\displaystyle \frac{d\sigma _U^{(\gamma Z)}}{d\mathrm{\Omega }}}`$ $`=`$ $`2Q^2{\displaystyle \frac{\mathrm{Re}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(1\lambda _{e^+}\lambda _e^{})g_e^A(\lambda _{e^+}\lambda _e^{})\}`$ (2.52)
$`\times \left\{{\displaystyle \frac{1}{8}}(1+\mathrm{cos}^2\theta )g_q^V𝒯_1+{\displaystyle \frac{1}{4}}\mathrm{sin}^2\theta g_q^V𝒯_3\right\}`$
$`+\{g_e^V(\lambda _{e^+}\lambda _e^{})+g_e^A(1\lambda _{e^+}\lambda _e^{})\}{\displaystyle \frac{1}{4}}g_q^A\mathrm{cos}\theta 𝒯_5]`$
$`+2Q^2{\displaystyle \frac{\mathrm{Im}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(\lambda _{e^+}\lambda _e^{})g_e^A(1\lambda _{e^+}\lambda _e^{})\}`$
$`\times {\displaystyle \frac{1}{4}}g_q^A\mathrm{cos}\theta \mathrm{Im}𝒯_6],`$
$`{\displaystyle \frac{d\sigma _L^{(\gamma Z)}}{d\mathrm{\Omega }}}`$ $`=`$ $`2Q^2{\displaystyle \frac{\mathrm{Re}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(1\lambda _{e^+}\lambda _e^{})+g_e^A(\lambda _{e^+}\lambda _e^{})\}`$ (2.53)
$`\times \left\{{\displaystyle \frac{1}{2}}g_q^A\left((1+\mathrm{cos}^2\theta )𝒯_{7,L_1}+\mathrm{sin}^2\theta 𝒯_{7,L_2}\right)\right\}`$
$`+\{g_e^V(\lambda _{e^+}\lambda _e^{})g_e^A(1\lambda _{e^+}\lambda _e^{})\}g_q^V\mathrm{cos}\theta 𝒯_{10,L}]`$
$`+2Q^2{\displaystyle \frac{\mathrm{Im}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(1\lambda _{e^+}\lambda _e^{})g_e^A(\lambda _{e^+}\lambda _e^{})\}`$
$`\times \left\{{\displaystyle \frac{1}{2}}g_q^A(1+\mathrm{cos}^2\theta )\mathrm{Im}𝒯_{8,L}\right\}],`$
$`{\displaystyle \frac{d\sigma _T^{(\gamma Z)}}{d\mathrm{\Omega }}}`$ $`=`$ $`2Q^2{\displaystyle \frac{\mathrm{Re}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(1\lambda _{e^+}\lambda _e^{})+g_e^A(\lambda _{e^+}\lambda _e^{})\}`$ (2.54)
$`\times \left\{{\displaystyle \frac{1}{2}}g_q^A\mathrm{cos}\theta 𝒯_{7,T}\right\}`$
$`+\{g_e^V(\lambda _{e^+}\lambda _e^{})g_e^A(1\lambda _{e^+}\lambda _e^{})\}g_q^V𝒯_{10,T}\}]`$
$`+2Q^2{\displaystyle \frac{\mathrm{Im}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(1\lambda _{e^+}\lambda _e^{})g_e^A(\lambda _{e^+}\lambda _e^{})\}`$
$`\times \left\{{\displaystyle \frac{1}{2}}g_q^A\mathrm{cos}\theta \mathrm{Im}𝒯_{8,T}\right\}],`$
$`{\displaystyle \frac{d\sigma _N^{(\gamma Z)}}{d\mathrm{\Omega }}}`$ $`=`$ $`2Q^2{\displaystyle \frac{\mathrm{Re}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(1\lambda _{e^+}\lambda _e^{})g_e^A(\lambda _{e^+}\lambda _e^{})\}`$ (2.55)
$`\times \left\{g_q^V\mathrm{cos}\theta \mathrm{sin}\theta 𝒯_{9,N}\right\}`$
$`+\{g_e^V(\lambda _{e^+}\lambda _e^{})+g_e^A(1\lambda _{e^+}\lambda _e^{})\}{\displaystyle \frac{1}{2}}g_q^A\mathrm{sin}\theta 𝒯_{12,N}]`$
$`+2Q^2{\displaystyle \frac{\mathrm{Im}𝒫^{\gamma Z}}{c_w^2s_w^2}}Q_eQ_q[\{g_e^V(\lambda _{e^+}\lambda _e^{})g_e^A(1\lambda _{e^+}\lambda _e^{})\}`$
$`\times {\displaystyle \frac{1}{2}}g_q^A\mathrm{sin}\theta \mathrm{Im}𝒯_{13,N}].`$
The Born contributions (zeroth order in $`\alpha _s`$) to the unpolarized quantities denoted by $`𝒯_i^{(0)}`$ ($`i=1,2,3,5`$) can be found in , , . The longitudinal $`𝒯_{7,L_1}^{(0)}`$, $`𝒯_{i,L}^{(0)}`$ ($`i=10,11`$) and the transverse polarized structure functions $`𝒯_{7,T}^{(0)}`$ and $`𝒯_{10,T}^{(0)}`$ were calculated in . In the last reference one also finds the Born contribution to the normal polarized quantity given by $`\mathrm{Im}𝒯_{13,N}^{(0)}`$. Notice that the quantities $`𝒯_i`$ not mentioned above all vanish in the Born approximation. The first order QCD contributions to $`𝒯_{9,N}`$ and $`𝒯_{12,N}`$ are also computed in but the corrections to the other structure functions $`𝒯_i`$ were neglected. The latter are computed in , , (longitudinal) and (transverse and normal). The exceptions are the order $`\alpha _s`$ contributions to $`\mathrm{Im}𝒯_6`$ and $`\mathrm{Im}𝒯_{8,L}`$ which will be presented in this paper for the first time. The second order QCD corrections are not known yet but for light quarks like $`u,d,s`$ they can be computed for the longitudinal polarized structure functions $`𝒯_{i,L}`$ ($`i=7,10,11`$) and the results are shown in the next section. Notice that the transverse and normal polarized structure functions $`𝒯_{i,T}`$ and $`𝒯_{i,N}`$ vanish for massless quarks.
## 3 Computation of the corrections up to order $`\alpha _s^2`$.
The partonic structure tensor defined in Eq. (2.13) is represented by the following perturbation series
$`W_{\mu \nu }^{(V_1V_2)}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{\alpha _s(\mu ^2)}{4\pi }}\right)^kW_{\mu \nu }^{(V_1V_2),(k)},`$ (3.1)
where $`\mu `$ denotes the renormalization scale. The same expression for the perturbation series holds for the the structure functions $`W_i^{(V1V2)}`$ and the functions $`𝒯_i`$ in Eqs. (2.21)-(2.39). The structure tensor is determined by the following reaction
$`V(q)H(p)+\overline{H}(p_1)+l(p_2)\mathrm{}l(p_n),`$ (3.2)
where $`l(p_i)`$ ($`i=2,3,\mathrm{}n`$) represent the momenta of the light partons ((anti-)quarks and gluons) in the final state. If the matrix element of the process above is given by $`M^{\mu \nu ,(V1V2)}`$ then the partonic tensor is obtained by integration over the multi-partonic phase space which also includes the heavy anti-quark i.e.
$`W_{\mu \nu }^{(V1V2)}={\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{d^3p_i}{(2\pi )^32p_i^0}(2\pi )^4\delta ^{(4)}\left(qp\underset{j=1}{\overset{n}{}}p_j\right)M_{\mu \nu }^{(V1V2)}}.`$ (3.3)
The computation of the phase space integrals is carried out in the rest frame of the vector boson $`V`$ and proceeds in the way as is given in where the integrals are performed up to order $`\alpha _s^2`$ in the case of massless quarks. It can be easily extended for massive quarks up to order $`\alpha _s`$. However in second order the integrals for massive quarks become very tedious and we are only able to present the results for massless quarks.
In the Born approximation the zeroth order structure functions $`W_i^{(V1V2),(0)}`$ are determined by the following process
$`V(q)H(p)+\overline{H}(p_1).`$ (3.4)
In this case the phase space integral is trivial and the partonic structure functions, presented in Eq. (A), are proportional to $`\delta (1x)`$. From Eqs. (2.22)-(2.39) we obtain for the unpolarized structure functions
$`𝒯_1^{(0)}`$ $`=`$ $`\beta ,𝒯_2^{(0)}=\beta ^3,𝒯_3^{(0)}={\displaystyle \frac{1}{2}}\beta \rho ,𝒯_4^{(0)}=0,𝒯_5^{(0)}=\beta ^2,`$
$`\mathrm{Im}𝒯_6^{(0)}`$ $`=`$ $`0,\beta =\sqrt{1\rho }.`$ (3.5)
The longitudinal polarized quantities in the Born approximation are given by
$`𝒯_{7,L_1}^{(0)}`$ $`=`$ $`{\displaystyle \frac{\beta ^2}{4}},𝒯_{7,L_2}^{(0)}=0,\mathrm{Im}𝒯_{8,L}^{(0)}=0,𝒯_{10,L}^{(0)}={\displaystyle \frac{\beta }{4}},`$
$`𝒯_{11,L}^{(0)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\beta ^3.`$ (3.6)
The transverse polarized quantities equal
$`𝒯_{7,T}^{(0)}`$ $`=`$ $`{\displaystyle \frac{\beta ^2}{4}}\sqrt{\rho },\mathrm{Im}𝒯_{8,T}^{(0)}=0,𝒯_{10,T}^{(0)}={\displaystyle \frac{\beta }{4}}\sqrt{\rho },𝒯_{11,T}^{(0)}=0.`$ (3.7)
For the normal polarized quantities we get
$`𝒯_{9,N}^{(0)}`$ $`=`$ $`0,𝒯_{12,N}^{(0)}=0,\mathrm{Im}𝒯_{13,N}^{(0)}={\displaystyle \frac{\beta ^2}{4}}\sqrt{\rho },`$
$`\beta `$ $`=`$ $`\sqrt{1\rho }.`$ (3.8)
The order $`\alpha _s`$ corrections originate from the one-loop contributions to the Born reaction in Eq. (3.4) and the gluon bremsstrahlung process
$`V(q)H(p)+\overline{H}(p_1)+g(p_2).`$ (3.9)
To facilitate the calculation we split the partonic tensor into a virtual (VIRT), a soft gluon (SOFT) and a hard (HARD) gluon part i.e.
$`W_{\mu \nu }^{(V_1V_2),(1)}`$ $`=W_{\mu \nu }^{(V_1V_2),VIRT}+W_{\mu \nu }^{(V_1V_2),SOFT}+W_{\mu \nu }^{(V_1V_2),HARD},`$ (3.10)
Starting with the virtual contribution the order $`\alpha _s`$ corrected vector boson quark vertex reads
$`\mathrm{\Gamma }_{a,\mu }^{V,(1)}`$ $`=`$ $`i[\gamma _\mu (1+𝒞_1)v_a^V+\gamma _5\gamma _\mu (1+𝒞_1+2𝒞_2)a_a^V`$ (3.11)
$`+{\displaystyle \frac{(2p_\mu q_\mu )}{2m}}𝒞_2v_a^V+\gamma _5{\displaystyle \frac{q_\mu }{2m}}𝒞_3a_a^V],V=\gamma ,Z.`$
Here the functions $`C_i`$ are given by
$`\mathrm{Re}𝒞_1`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}C_F[({\displaystyle \frac{2\rho }{\beta }}\mathrm{ln}(t)+2)\mathrm{ln}\left({\displaystyle \frac{\lambda ^2}{m^2}}\right)43\beta \mathrm{ln}(t)`$
$`+{\displaystyle \frac{2\rho }{\beta }}\{{\displaystyle \frac{1}{2}}\mathrm{ln}^2(t)+2\mathrm{ln}(t)\mathrm{ln}(1t)+2i_2(t)+{\displaystyle \frac{2\pi ^2}{3}}\}],`$
$`\mathrm{Re}𝒞_2`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}C_F\left[{\displaystyle \frac{\rho }{\beta }}\mathrm{ln}(t)\right],`$
$`\mathrm{Re}𝒞_3`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}C_F\left[(\rho 3){\displaystyle \frac{\rho }{\beta }}\mathrm{ln}(t)2\rho \right],`$
$`\mathrm{Im}𝒞_1`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}\pi C_F[\left({\displaystyle \frac{2\rho }{\beta }}\right)\mathrm{ln}\left({\displaystyle \frac{\lambda ^2}{m^2}}\right)3\beta +{\displaystyle \frac{2\rho }{\beta }}\{\mathrm{ln}(t)`$
$`+2\mathrm{ln}(1t)\}],`$
$`\mathrm{Im}𝒞_2`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}\pi C_F\left[{\displaystyle \frac{\rho }{\beta }}\right],`$
$`\mathrm{Im}𝒞_3`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}\pi C_F\left[(3\rho ){\displaystyle \frac{\rho }{\beta }}\right],`$
$`t`$ $`=`$ $`{\displaystyle \frac{1\beta }{1+\beta }},`$ (3.12)
where $`C_F`$ denotes the colour factor which is equal to $`C_F=(N_c^21)/2N_c`$. In the equation above we have introduced a fictitious mass of the gluon $`\lambda `$ which is needed to regularize the infrared divergence. The ultraviolet divergences, which cancel in the expressions above, can be regularized with the cut-off method ($`^{\mathrm{\Lambda }^2}d^4k`$) which is known from old textbooks on QED. In this way one avoids the intricacies of the $`\gamma _5`$-matrix prescription which is characteristic of n-dimensional regularization. The mass renormalization is carried out in the pole mass scheme (on-shell renormalization). However one can also choose the $`\overline{MS}`$-scheme and carry out the analysis of the radiative corrections using the running mass (see e.g. ). The contributions from the Born reaction and the one-loop corrections are given by
$`W_{\mu \nu }^{(V_1V_2),(0)}+W_{\mu \nu }^{(V_1V_2),VIRT}`$ $`=`$ $`{\displaystyle \frac{N_c\beta }{32\pi ^2}}Tr\left[{\displaystyle \frac{(1+\gamma _5/s)}{2}}(/p+m)\mathrm{\Gamma }_{a,\mu }^{V_1,(1)}(/p_1m)\stackrel{~}{\mathrm{\Gamma }}_{a,\nu }^{V_2,(1)}\right]`$ (3.13)
$`\times \delta (1x),`$
with $`\stackrel{~}{\mathrm{\Gamma }}_\mu =\gamma _0\mathrm{\Gamma }_\mu ^{}\gamma _0`$. The soft gluon part of the partonic tensor is given by
$`W_{\mu \nu }^{(V_1V_2),SOFT}`$ $`=`$ $`{\displaystyle \frac{N_c\beta }{16\pi ^2}}S^{SOFT}Tr\left[{\displaystyle \frac{(1+\gamma _5/s)}{2}}(/p+m)\mathrm{\Gamma }_{a,\mu }^{V_1,(0)}(/p_1m)\mathrm{\Gamma }_{a,\nu }^{V_2,(0)}\right]\delta (1x),`$
with
$`S^{SOFT}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{8\pi ^2}}C_F{\displaystyle _0^\omega }{\displaystyle \frac{d^3p_2}{E_2}}\left({\displaystyle \frac{m^2}{(p_1p_2)^2}}+{\displaystyle \frac{m^2}{(pp_2)^2}}{\displaystyle \frac{2pp_1}{(p_1p_2)(pp_2)}}\right).`$ (3.15)
The above integral will be evaluated in the rest frame of the vector boson $`V`$ where $`\omega `$ is a cut off on the gluon energy $`E_2`$ which is taken to be much smaller than the quark mass $`m`$ (i.e. $`\omega m`$). This is the reason why like in the case of the contribution from the Born reaction and the virtual corrections the soft gluon part is proportional to $`\delta (1x)`$. The result is
$`S^{SOFT}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}C_F[2\mathrm{ln}\left({\displaystyle \frac{4\omega ^2}{\lambda ^2}}\right)+2{\displaystyle \frac{2\rho }{\beta }}\{\mathrm{ln}(t)2i_2(t)2i_2(t)`$ (3.16)
$`+\mathrm{ln}\left({\displaystyle \frac{4\omega ^2}{\lambda ^2}}\right)\mathrm{ln}(t)2\mathrm{ln}(t)\mathrm{ln}(1t)2\mathrm{ln}(t)\mathrm{ln}(1+t)`$
$`+\mathrm{ln}^2(t)+{\displaystyle \frac{\pi ^2}{6}}\}].`$
Addition of the virtual ($`V`$) and soft ($`S`$) gluon parts leads to the expressions $`W_i^{(V_1V_2),V+S}`$ in Eq. (A). Note that in the combination $`𝒞^{V+S}=𝒞_1+S^{SOFT}`$ the gluon regulator mass $`\lambda `$ vanishes i.e.
$`𝒞^{V+S}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{4\pi }}C_F[(2+{\displaystyle \frac{2\rho }{\beta }}\mathrm{ln}(t))\mathrm{ln}\left({\displaystyle \frac{4\omega ^2}{m^2}}\right){\displaystyle \frac{2\rho }{\beta }}\{4i_2(t)`$ (3.17)
$`2i_2(t)4\mathrm{ln}(t)\mathrm{ln}(1t)2\mathrm{ln}(t)\mathrm{ln}(1+t)+{\displaystyle \frac{3}{2}}\mathrm{ln}^2(t)`$
$`{\displaystyle \frac{\pi ^2}{2}}\}2{\displaystyle \frac{54\rho }{\beta }}\mathrm{ln}(t)].`$
The hard gluon part of the partonic tensor can be calculated in a straightforward way and the results are given in Eq. (A). Notice that the upper bound on the integral over $`x`$ in Eq. (2) for $`W_{\mu \nu }^{(V_1V_2),HARD}`$ is given by $`12m\omega /Q^2`$ so that the energy cut off $`\omega `$ is cancelled between $`W_i^{(V_1V_2),V+S}`$ and $`W_i^{(V_1V_2),HARD}`$. The results for $`𝒯_i`$ are given in Eqs. (B.1)-(B). We tried to shorten these expression as much as possible by minimizing the number of independent polylogarithms. Note that there is a difference in sign between $`𝒯_{12,N}`$ in Eq. (B.17) and the equivalent expression in Eq. (33) of . However substitution of $`𝒯_{9,N}`$ (Eq. (B.16)) and $`𝒯_{12,N}`$ (Eq. (B.17)) in the cross section of Eq. (2.51) leads to the same result as presented in Eq. (16) in .
The computation of the order $`\alpha _s^2`$ corrections for massive quarks is extremely tedious so that it has not been performed yet. The reason is that the partonic structure functions for timelike processes like $`e^+e^{}`$-collisions cannot be written as the imaginary part of an amplitude. This is in contrast to deep inelastic scattering where $`q^2`$ is spacelike or the total cross section $`\sigma _{tot}(e^+e^{}\mathrm{`}hadrons^{})`$. In the latter case the cross section can be written as the imaginary part of the hadronic vacuum polarization function so that one is able to apply advanced methods to compute these type of quantities (see e.g. ). However for massless quarks we can compute the second order corrections to the partonic structure functions contributing to the longitudinal polarization. This is feasible using conventional techniques as is shown in -. In the case of massless quarks the computation is facilitated because one has the following relations
$`W_1^{v_q^2}=W_1^{a_q^2}`$ $`=`$ $`{\displaystyle \frac{x}{2}}W_3^{v_qa_q},`$
$`W_2^{v_q^2}=W_2^{a_q^2}`$ $`=`$ $`{\displaystyle \frac{x}{2}}W_4^{v_qa_q}2W_5^{v_qa_q},`$
$`W_6^{v_qa_q}`$ $`=`$ $`{\displaystyle \frac{x}{2}}W_7^{v_q^2,a_q^2}+W_9^{v_q^2,a_q^2}.`$ (3.18)
which follow from Eq. (2.13) by putting $`s=p/m`$. Since $`m=0`$ the contributions coming from the vector currents $`v_q^2`$ lead to the same answer as those originating from the axial-vector currents given by $`a_q^2`$. In this case the matrix element only contains $`\gamma `$-matrices which anti-commute with the $`\gamma _5`$-matrix. It also explains why the components proportional to $`v_q^2`$ and $`a_q^2`$ which show up in the unpolarized structure functions are the same as the contributions multiplying $`v_qa_q`$ appearing in the polarized structure functions. The relations in Eq. (3) break down for $`m0`$ as will be discussed at the end of this section. The structure functions $`W_i`$ in Eq. (2.13) can be related to the fragmentation functions $`F_i^h(x,Q^2)`$ (unpolarized see , ) and $`g_i^h(x,Q^2)`$ (polarized see ) defined for the process $`e^+e^{}h+\mathrm{"}X\mathrm{"}`$. Here $`h`$ represents a hadron in the final state which originates from a light (anti-) quark. Using the relations in Eqs. (2.22)- (2.32), the quantities $`𝒯_i`$ can be expressed into the first moments of the non-singlet quark coefficient functions as follows
$`𝒯_1`$ $`=`$ $`𝒯_2={\displaystyle _0^1}𝑑x𝒞_{1,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}}),`$ (3.19)
$`𝒯_3`$ $`=`$ $`𝒯_4={\displaystyle \frac{1}{2}}{\displaystyle _0^1}𝑑x\left(𝒞_{1,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}})𝒞_{2,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}})\right),`$ (3.20)
$`𝒯_5`$ $`=`$ $`{\displaystyle _0^1}𝑑x𝒞_{3,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}}),`$ (3.21)
$`𝒯_{7,L_1}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle _0^1}𝑑x\mathrm{\Delta }𝒞_{5,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}}),`$ (3.22)
$`𝒯_{7,L_2}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle _0^1}𝑑x\left(\mathrm{\Delta }𝒞_{5,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}})\mathrm{\Delta }𝒞_{4,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}})\right),`$ (3.23)
$`𝒯_{10,L}`$ $`=`$ $`𝒯_{11,L}={\displaystyle \frac{1}{4}}{\displaystyle _0^1}𝑑x\mathrm{\Delta }𝒞_{1,q}(x,{\displaystyle \frac{Q^2}{\mu ^2}}),`$ (3.24)
Here $`\mu `$ represents the mass factorization scale as well as the renormalization scale. However since all coefficient functions above are of the non-singlet type the dependence on the factorization scale vanishes while taking the first moments so that $`𝒯_i`$ only depends on the renormalization scale. Because of the relations in Eq. (3) the coefficient functions above are not mutually independent. In , it was demonstrated that one could derive the following relations between the polarized $`\mathrm{\Delta }𝒞_{i,q}`$ ($`i=1,4,5`$) and the unpolarized $`𝒞_{i,q}`$ ($`i=1,2,3`$) coefficient functions. They are given by
$`\mathrm{\Delta }𝒞_{1,q}=𝒞_{3,q},\mathrm{\Delta }𝒞_{4,q}=𝒞_{2,q},\mathrm{\Delta }𝒞_{5,q}=𝒞_{1,q},`$ (3.25)
which means that in the case of massless quarks all quantities $`𝒯_i`$ are determined by three independent coefficient functions only. The order $`\alpha _s^2`$ contributions originate from the processes
$`VH+\overline{H}+g+g,`$
$`VH+\overline{H}+H+\overline{H},`$
$`VH+\overline{H}+q+q,`$ (3.26)
which also includes the two-loop corrections to reaction (3.4) and the one-loop corrections to process (3.9). The corresponding coefficient functions have been computed in - for massless quarks and their first moments are presented in . If we also include the contributions from the lower order processes in (3.4) and (3.9) (see also -) we obtain for massless quarks up to order $`\alpha _s^2`$
$`𝒯_1^{m=0}`$ $`=`$ $`𝒯_2^{m=0}=1+{\displaystyle \frac{\alpha _s(\mu ^2)}{4\pi }}C_F\left[1\right]+\left({\displaystyle \frac{\alpha _s(\mu ^2)}{4\pi }}\right)^2[C_F^2\left({\displaystyle \frac{7}{2}}\right)+C_AC_F({\displaystyle \frac{11}{3}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{\mu ^2}}\right)`$ (3.27)
$`+{\displaystyle \frac{347}{18}}44\zeta (3))+n_fC_FT_f({\displaystyle \frac{4}{3}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{\mu ^2}}\right){\displaystyle \frac{62}{9}}+16\zeta (3))],`$
$`𝒯_3^{m=0}`$ $`=`$ $`𝒯_4^{m=0}={\displaystyle \frac{\alpha _s(\mu ^2)}{4\pi }}C_F\left[2\right]+\left({\displaystyle \frac{\alpha _s(\mu ^2)}{4\pi }}\right)^2[C_F^2(5)+C_AC_F({\displaystyle \frac{22}{3}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{\mu ^2}}\right)`$ (3.28)
$`+{\displaystyle \frac{380}{9}})+n_fC_FT_f({\displaystyle \frac{8}{3}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{\mu ^2}}\right){\displaystyle \frac{136}{9}})],`$
$`𝒯_5^{m=0}`$ $`=`$ $`1+\left({\displaystyle \frac{\alpha _s(\mu ^2)}{4\pi }}\right)^2\left[C_AC_F\left(44\zeta (3)\right)+n_fC_FT_f\left(16\zeta (3)\right)\right],`$ (3.29)
$`𝒯_{7,L_1}^{m=0}`$ $`=`$ $`{\displaystyle \frac{1}{4}}𝒯_1^{m=0},`$ (3.30)
$`𝒯_{7,L_2}^{m=0}`$ $`=`$ $`{\displaystyle \frac{1}{2}}𝒯_3^{m=0},`$ (3.31)
$`𝒯_{10,L}^{m=0}`$ $`=`$ $`𝒯_{11,L}^{m=0}={\displaystyle \frac{1}{4}}𝒯_5,`$ (3.32)
where we have chosen the $`\overline{MS}`$-scheme for the coupling constant renormalization. In the above equations the colour factors are given by $`C_A=N_c`$ and $`T_f=1/2`$ (for $`C_F`$ see below Eq. (3). Furthermore $`n_f`$ is the number of light flavours and the scale $`\mu `$ represents the renormalization scale. Notice that the sum $`𝒯_1+𝒯_3`$ is equal to $`R_{e^+e^{}}`$ which is defined by $`\sigma _{tot}(e^++e^{}hadrons)/\sigma _{tot}(e^++e^{}\mu ^++\mu ^{})`$. Finally the results obtained for the spin dependent quantities $`𝒯_{7,L_1}`$, $`𝒯_{10,L}`$, $`𝒯_{11,L}`$ depend on the way the limit $`m0`$ is taken before or after the integration over the momenta. This is revealed by a comparison of the order $`\alpha _s`$ corrections in Eqs. (3.30)-(3.32) with those obtained for the same quantities in Eqs. (B), (B), (B.11). We find the following relation
$`𝒯_{7,L_1}^{(1),m0}+2𝒯_{7,L_1}^{(0),m0}=𝒯_{7,L_1}^{(1),m=0},𝒯_{i,L}^{(1),m0}+2𝒯_{i,L}^{(0),m0}=𝒯_{i,L}^{(1),m=0},i=10,11.`$
Therefore we expect that the same difference between the massless and massive quark approach will happen in higher order. This difference originates from the property that the $`\gamma _5`$-matrix commutes with the mass term in the trace (see e.g. Eq. (3.13)) contrary to the ordinary gamma-matrix with which it anti-commutes. Hence the relations between the polarized and the unpolarized partonic structure functions in Eq. (3) will break down for the subleading terms. The origin of this phenomenon is explained in <sup>2</sup><sup>2</sup>2A similar phenomenon has been observed in polarized deep inelastic lepton-hadron scattering for the structure function $`\mathrm{\Delta }g_1`$ see . . It does not show up in the relations between the $`v_q^2`$\- and $`a_q^2`$-parts of the unpolarized structure functions. Therefore only the coefficient functions $`\mathrm{\Delta }𝒞_i`$ ($`i=1,4,5`$) in Eqs. (3.22)-(3.24) will get an anomalous term while going from the massless to the massive quark approach. It also turns out that this term cancels in the combination $`\mathrm{\Delta }𝒞_5\mathrm{\Delta }𝒞_4`$ so that $`𝒯_{7,L_2}`$ in Eq. (3.23) will be unaffected at least up to order $`\alpha _s`$. It is unlikely that the latter will also hold in higher order of perturbation theory.
## 4 Results
In this section we will present the effect of the higher order QCD corrections to the polarization of the quark in $`e^+e^{}`$-collisions. When both the incoming leptons as well as the outgoing quark are unpolarized the differential cross section is given by
$`{\displaystyle \frac{d\overline{\sigma }}{d\mathrm{\Omega }}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{\lambda _{e^+},\lambda _e^{}}{}}\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W)+{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W)\right)={\displaystyle \frac{1}{2}}{\displaystyle \underset{\lambda _{e^+},\lambda _e^{}}{}}{\displaystyle \frac{d\sigma _U}{d\mathrm{\Omega }}}.`$ (4.1)
When the incoming leptons are unpolarized but the quark is polarized the asymmetry is defined by
$`{\displaystyle \frac{d\sigma _W}{d\mathrm{\Omega }}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{\lambda _{e^+},\lambda _e^{}}{}}\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W){\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W)\right)`$ (4.2)
$`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{\lambda _{e^+},\lambda _e^{}}{}}\left(W_L{\displaystyle \frac{d\sigma _L}{d\mathrm{\Omega }}}+W_T{\displaystyle \frac{d\sigma _T}{d\mathrm{\Omega }}}+W_N{\displaystyle \frac{d\sigma _N}{d\mathrm{\Omega }}}\right).`$
In the case the electron is polarized the asymmetry becomes equal to
$`{\displaystyle \frac{d\sigma _W(\lambda _e^{})}{d\mathrm{\Omega }}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{\lambda _{e^+}}{}}\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W){\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{},W)\right)`$ (4.3)
$`=`$ $`{\displaystyle \underset{\lambda _{e^+}}{}}\left(W_L{\displaystyle \frac{d\sigma _L}{d\mathrm{\Omega }}}+W_T{\displaystyle \frac{d\sigma _T}{d\mathrm{\Omega }}}+W_N{\displaystyle \frac{d\sigma _N}{d\mathrm{\Omega }}}\right).`$
When the positron is polarized and the electron is unpolarized we have the relation
$`{\displaystyle \frac{d\sigma _W(\lambda _{e^+})}{d\mathrm{\Omega }}}={\displaystyle \frac{d\sigma _W(\lambda _e^{})}{d\mathrm{\Omega }}}`$ (4.4)
The polarizations of the quark are defined by
$`P_k=W_k{\displaystyle \frac{\underset{\lambda _{e^+},\lambda _e^{}}{}d\sigma _k/d\mathrm{\Omega }}{_{\lambda _{e^+},\lambda _e^{}}d\sigma _U/d\mathrm{\Omega }}},P_k(\lambda _e^{})=W_k{\displaystyle \frac{\underset{\lambda _{e^+}}{}d\sigma _k/d\mathrm{\Omega }}{_{\lambda _{e^+}}d\sigma _U/d\mathrm{\Omega }}},`$
$`k=L,T,N.`$ (4.5)
For the longitudinally polarized quark we choose the following spin vector
$`\widehat{W}=\widehat{n},W_L=1,W_T=0,W_N=0.`$ (4.6)
For the transversely polarized quark the spin vector is chosen to be in the plane spanned by the electron and the quark momenta
$`\widehat{W}={\displaystyle \frac{\widehat{n}\times (\widehat{n}\times \widehat{q}_e^{})}{|\widehat{n}\times (\widehat{n}\times \widehat{q}_e^{})|}},W_L=0,W_T=\mathrm{sin}\theta ,W_N=0.`$ (4.7)
For the quark polarisation which is directed normal to the plan spanned by the electron and the quark momenta we make the choice
$`\widehat{W}={\displaystyle \frac{\widehat{n}\times \widehat{q}_e^{}}{|\widehat{n}\times \widehat{q}_e^{}|}},W_L=0,W_T=0,W_N=1.`$ (4.8)
Notice that $`\widehat{W}`$ is chosen in the same way as in but opposite the choice made in <sup>3</sup><sup>3</sup>3In the case the anti-quark $`\overline{H}`$ is detected in the final state we have $`P_k(\overline{H})=P_k(H)`$ for $`k=L,T`$ but $`P_N(\overline{H})=P_N(H)`$.. With the definitions above one can infer that on the Born level we have the following properties. In the case of longitudinally polarized quarks ($`W_T=0,W_N=0`$) one obtains for $`W_L=\pm 1`$
$`{\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{}=W_L,W_L,\mathrm{cos}\theta =1)={\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma _U}{d\mathrm{\Omega }}}(\mathrm{cos}\theta =1)`$
$`{\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{}=W_L,W_L,\mathrm{cos}\theta =1)=0`$
$`{\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{}=W_L,W_L,\mathrm{cos}\theta =1)={\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma _U}{d\mathrm{\Omega }}}(\mathrm{cos}\theta =1)`$
$`{\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{}=W_L,,W_L,\mathrm{cos}\theta =1)=0`$ (4.9)
The above relations hold for massive and massless quarks. From Eqs. (4.3) and (4) it follows
$`P_L(\lambda _e^{}=\pm 1,\mathrm{cos}\theta =\pm 1)=1P_L(\lambda _e^{}=\pm 1,\mathrm{cos}\theta =1)=1`$ (4.10)
For the transverse polarized cross section we also have a relation which only holds at threshold $`Q2m`$. It reads for $`\theta =\pi /2`$ ($`W_T=1`$)
$`{\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{}=W_T,W_T,\mathrm{cos}\theta =0){\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma _U}{d\mathrm{\Omega }}}(\mathrm{cos}\theta =0)`$
$`{\displaystyle \underset{\lambda _{e^+}}{}}{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}(\lambda _{e^+},\lambda _e^{}=W_T,W_T,\mathrm{cos}\theta =0)0`$ (4.11)
from which follows
$`\underset{Q2m}{\text{lim}}P_T(\lambda _e^{}=\pm 1,\mathrm{cos}\theta =0)=1`$ (4.12)
When $`Q>2m`$ we get $`|P_T|<1`$. Far away from threshold the transverse polarization becomes very small and tends to zero. All relations above will be modified by QCD corrections as we will see below.
We will now discuss the effect of the higher order QCD corrections on the polarizations of the quarks where we neglect any higher order effect coming from the electro-weak sector. Our results are obtained by choosing the parameters given in . The electro-weak constants are: $`M_Z=91.187\mathrm{GeV}/\mathrm{c}^2`$, $`\mathrm{\Gamma }_Z=2.490\mathrm{GeV}/\mathrm{c}^2`$ and $`\mathrm{sin}^2\theta _W=0.23116`$. For the strong parameters we adopt $`\mathrm{\Lambda }_{\overline{MS}}=237\mathrm{MeV}/\mathrm{c}`$ at $`n_f=5`$ so that the two-loop corrected running coupling constant equals $`\alpha _s(M_Z^2)=0.119`$. Further we take the renormalization scale $`\mu =Q`$. Furthermore we only study up-quark, bottom- and top quark production for which the following masses are chosen $`m_u=0`$, $`m_b=4.5\mathrm{GeV}/\mathrm{c}^2`$ and $`m_t=173.8\mathrm{GeV}/\mathrm{c}^2`$. In our plots we will show the polarizations computed in different orders of perturbation theory. Hence we follow the notation in Eq. (3.1) and define $`P_k^{(i)}`$ ($`k=L,T,N`$) to be the order $`\alpha _s^i`$ contribution to the polarization. Similarly we define $`P_{k,i}`$ to be the order $`\alpha _s^i`$ corrected polarization. Finally we only show figures where the electron beam is polarized. In the case the positron is polarized and the electron beam is unpolarized the figures for $`\lambda _{e^+}=\pm 1`$ are the same as those for $`\lambda _e^{}=1`$.
In Fig. 1 we have plotted the Born (zeroth order $`\alpha _s`$) and the higher order corrections to the longitudinal polarization of the up-quark at $`Q=M_Z`$. The computation is done in the massless quark approach were the anomalous terms are absent. Since the QCD corrections for unpolarized beams are very small we could only show the order $`\alpha _s^2`$ corrected result $`P_{L,2}`$. For this case we have plotted the ratios $`P_{L,1}/P_{L,0}`$ and $`P_{L,2}/P_{L,1}`$ in Fig. 2. From the latter figure we infer that the order $`\alpha _s`$ corrections do not exceed the 6 pro-mille level. The order $`\alpha _s^2`$ corrections are at most 2 pro-mille. In the case of polarized beams they become larger. For $`\mathrm{cos}\theta =\pm 1`$ the order $`\alpha _s`$ and order $`\alpha _s^2`$ corrections are $`8.5\%`$ and $`2.5\%`$ respectively. Similar features are shown in Fig. 3 by bottom production at the same CM energy which are computed in the massive quark approach so that the anomalous terms are implicitly present. Here the corrections for polarized beams are larger than those obtained for the up-quark and they amount to $`25\%`$ at $`\mathrm{cos}\theta =\pm 1`$ but the corrections for unpolarized beams are so small that they could not be shown in the figure. In Fig. 4 we have studied the validity of the massless quark approach ($`m=0`$) and the massive quark approach ($`m0`$) for $`P_L^{(1)}`$ in the case of bottom production at $`Q=M_Z`$ with unpolarized lepton beams. To that order we have plotted the ratios of the various approaches with respect to the exact result computed for $`m=m_b=4.5\mathrm{GeV}/\mathrm{c}^2`$. As expected the massless approach given by $`P_L^{(1)}(m=0)`$ does not work very well (see the dotted curve in Fig. 4) but also the massive approach $`P_L^{(1)}(m0)`$ is rather bad in particular near $`\mathrm{cos}\theta =\pm 1`$. It turns out that only for $`m_b<0.2\mathrm{GeV}/\mathrm{c}^2`$ the difference between $`P_L^{(1)}(m0)`$ and $`P_L^{(1)}(m=0.2)`$ is less than $`9\%`$. For a comparison we have also shown the ratio of the unpolarized cross sections $`d^2\sigma ^{(1)}(m0)/d^2\sigma ^{(1)}(m=4.5)`$ where no anomalous terms are present so that $`d^2\sigma ^{(1)}(m0)=d^2\sigma ^{(1)}(m=0)`$. In this case the massless quark approach works rather well except near $`\mathrm{cos}\theta =\pm 1`$. This is the justification for neglecting the bottom mass in the order $`\alpha _s^2`$ contributions to the forward backward asymmetry and the shape parameter in , . In Fig. 5 we also plotted the longitudinal polarization of the top-quark which is produced at $`Q=500\mathrm{GeV}`$. Like in the two previous cases the order $`\alpha _s`$ corrections to $`P_L`$ are extremely small even when the incoming leptons are polarized. Notice that for up-quark and bottom-quark production at $`Q=M_Z`$ the $`ZZ`$ interference term in Eq. (2.49) dominates the longitudinal polarization. However at $`Q=500\mathrm{GeV}`$ it turns out that the $`\gamma \gamma `$ (Eq. (2.45)) and $`\gamma Z`$ (Eq. (2.53)) interference terms become important too although they partially cancel each other because the latter is negative.
For an analysis of the transverse $`P_T`$ and normal $`P_N`$ polarization we have to limit ourselves to heavy quark production since these quantities vanish for massless quarks. In Fig. 6 we have shown $`P_T`$ at $`Q=M_Z`$ for bottom quarks. The QCD corrections are much larger than for $`P_L`$. This also holds for unpolarized beams. On the other hand the absolute values for $`P_T`$ are rather small which is mainly due to the fact that the bottom quark is produced far away from threshold $`Q=M_Z2m_b`$ so that $`P_T2m_b/M_Z1`$. For top-quark production at $`Q=500\mathrm{GeV}`$ (see Fig. 7) the situation is different. This quark is produced rather close to threshold and the ratio $`2m_t/Q0.7`$ is large enough so that for polarized beams the maximum value of $`|P_T|`$ is close to one. However due to the difference $`Q2m_t`$ there is a shift from $`\theta =\pi /2`$ to larger angles (here about $`\theta =2\pi /3`$) where $`P_T`$ attains its maximum ($`\lambda _e^{}=1`$) or its minimum ($`\lambda _e^{}=1`$).
The normal polarization $`P_N`$ is presented for the bottom quark ($`Q=M_Z`$) in Fig. 8. From this figure we infer that $`P_N`$ is about a factor of five smaller than $`P_T`$ in Fig. 6. Furthermore the difference between the Born approximation $`P_{N,0}`$, which is wholly due to the quantity $`𝒯_{13,N}^{(0)}`$ in $`d^2\sigma _N^{\gamma Z}`$ (Eq. (2.55)), and the QCD corrected result $`P_{N,1}`$ is much larger than observed for $`P_L`$ and $`P_T`$ . Notice that the QCD corrections to $`P_N`$ are not determined by the Z-peak in Eq. (2.51) only but they also receive contributions from the $`\gamma Z`$ interference term in Eq. (2.55). Actually it is the structure function $`𝒯_{13,N}^{(1)}`$ (see Eq. (B) which is mainly responsible for the difference in Fig. 8 between $`P_{N,0}`$ and $`P_{N,1}`$ around $`\mathrm{cos}\theta =0.8`$ for $`\lambda _e^{}=1`$. When one is off-resonance the contributions in Eqs. (2.52)-(2.55) which are proportional to $`\mathrm{Im}P^{\gamma Z}M_Z\mathrm{\Gamma }_Z`$ are heavily suppressed. This is the reason why for top production at $`Q=500\mathrm{GeV}`$ (see Fig. 9) the Born approximation to the normal polarization given by $`P_{N,0}`$ is almost zero. Therefore in this case neither $`𝒯_{13,N}^{(0)}`$ nor $`𝒯_{13,N}^{(1)}`$, both appearing in Eq. (2.55), play any role anymore. Hence $`P_N`$ in Fig. 9, which is smaller than $`P_T`$ in Fig. 7 by a factor of forty, is completely dominated by the QCD contributions coming from the structure functions $`𝒯_{9,N}^{(1)}`$ (Eq. (B.16)) and $`𝒯_{12,N}^{(1)}`$ (Eq. (B.17)). This observation is important because besides of these corrections above, $`P_N`$ can also receive contributions coming from CP-violating terms in the neutral current sector which is a signal of new physics. Hence it is important to compute the QCD corrections beyond order $`\alpha _s`$ which is not an easy task as we have discussed below Eq. (3.3). Finally we want to emphasize that for the top-quark all interference terms are equally important for the determination of $`P_T`$ and $`P_N`$. This observation was already mentioned for $`P_L`$ above. This is because $`t\overline{t}`$-production occurs far above the Z-boson resonance so that the dominance of the Z-propagator disappears.
As far as possible we have also made a comparison with some results obtained in the literature. First we agree with the results presented for the top-quark for polarized and unpolarized beams shown in Figs. 2, 9 in . We also found agreement with the plots presented for $`P_L`$ and $`P_T`$ in \- where the contributions $`𝒯_6^{(1)}`$ (Eq. (B.6)) and $`𝒯_{8,L}^{(1)}`$ (Eq. (B.9)) were not taken into account. However these contributions are negligible. On the other hand we disagree with the normal polarization $`P_N`$ plotted for the bottom quark (Fig. 2b) and the top-quark (Fig. 3b) in . This is due to the difference in minus sign between $`𝒯_{12,N}^{(1)}`$ and the equivalent expression in Eq. (33) of which we have mentioned below Eq. (3.17). Note that the signs for $`𝒯_{9,N}^{(1)}`$ (Eq. (B.16)) and $`𝒯_{12,N}^{(1)}`$ (Eq. (B.17)) are the same as the contributions appearing in Eq. (16) of .
Before finishing this section we would like to comment on some results which are obtained in and . Here one has expressed $`\widehat{W}`$ in Eq. (2.14) into the following orthonormal basis (see also )
$`\widehat{W}=\mathrm{sin}\xi \mathrm{cos}\psi \widehat{n}_1+\mathrm{sin}\xi \mathrm{sin}\psi \widehat{n}_2+\mathrm{cos}\xi \widehat{n}_3,`$ (4.13)
with
$`\widehat{n}_1={\displaystyle \frac{\widehat{n}\times (\widehat{n}\times e_3)}{|\widehat{n}\times (\widehat{n}\times e_3)|}},\widehat{n}_2={\displaystyle \frac{\widehat{n}\times e_3}{|\widehat{n}\times e_3|}},\widehat{n}_3=\widehat{n},`$
$`e_1=(1,0,0),e_2=(0,1,0),e_3=(0,0,1).`$ (4.14)
Notice that $`\widehat{n}`$ is the direction of the three-momentum of the observed quark in the CM frame of the electron-positron pair whereas $`\widehat{n}_3=\widehat{n}`$ points into the direction of the momentum $`q=q_{e^+}+q_e^{}`$ in the rest frame of the observed quark. On the basis $`e_i`$ ($`i=1,2,3`$) the vectors $`\widehat{n}_i`$ take the following form:
$`\widehat{n}_1`$ $`=`$ $`(\mathrm{cos}\theta \mathrm{cos}\varphi ,\mathrm{cos}\theta \mathrm{sin}\varphi ,\mathrm{sin}\theta ),`$
$`\widehat{n}_2`$ $`=`$ $`(\mathrm{sin}\varphi ,\mathrm{cos}\varphi ,0),`$
$`\widehat{n}_3`$ $`=`$ $`(\mathrm{sin}\theta \mathrm{cos}\varphi ,\mathrm{sin}\theta \mathrm{sin}\varphi ,\mathrm{cos}\theta ).`$ (4.15)
Substituting Eq. (4.13) in Eq. (2.16) and using the parameterization for the quark momentum in (2) the spin four-vector $`s`$ in the CM frame of the incoming leptons has the following components
$`s^0`$ $`=`$ $`{\displaystyle \frac{Q\alpha _x}{2m}}\mathrm{cos}\xi ,`$
$`s^1`$ $`=`$ $`\left({\displaystyle \frac{Qx}{2m}}\right)\mathrm{cos}\xi \mathrm{sin}\theta \mathrm{cos}\varphi +\mathrm{sin}\xi \mathrm{cos}\psi \mathrm{cos}\theta \mathrm{cos}\varphi +\mathrm{sin}\xi \mathrm{sin}\psi \mathrm{sin}\varphi ,`$
$`s^2`$ $`=`$ $`\left({\displaystyle \frac{Qx}{2m}}\right)\mathrm{cos}\xi \mathrm{sin}\theta \mathrm{sin}\varphi +\mathrm{sin}\xi \mathrm{cos}\psi \mathrm{cos}\theta \mathrm{sin}\varphi \mathrm{sin}\xi \mathrm{sin}\psi \mathrm{cos}\varphi ,`$
$`s^3`$ $`=`$ $`\left({\displaystyle \frac{Qx}{2m}}\right)\mathrm{cos}\xi \mathrm{cos}\theta \mathrm{sin}\xi \mathrm{cos}\psi \mathrm{sin}\theta .`$ (4.16)
Similarly, using Eq.(4.13) one finds that
$`W_L`$ $`=`$ $`\widehat{n}\widehat{W}=\mathrm{cos}\xi ,`$
$`W_T`$ $`=`$ $`\widehat{W}^3\widehat{n}.\widehat{W}\mathrm{cos}\theta =\mathrm{sin}\xi \mathrm{cos}\psi \mathrm{sin}\theta ,`$
$`W_N`$ $`=`$ $`\widehat{W}^2\mathrm{cos}\varphi \widehat{W}^1\mathrm{sin}\varphi =\mathrm{sin}\xi \mathrm{sin}\psi .`$ (4.17)
Comparing the equation above with our choices of the polarization vector in Eqs. (4.6) - (4.8) we obtain $`\xi =0,\pi ,\psi =arbitrary`$ (longitudinal), $`\xi =\pm \pi /2,\psi =0`$ (transverse), and $`\xi =\pm \pi /2,\psi =\pi /2`$ (normal) respectively. Notice that the same angles were also adopted in -. A different choice was made made in , where the values of the angles are taken at $`\psi =0`$ and $`\xi =arbitrary`$. In this case only the longitudinal and transverse polarization of the heavy quark can be studied since $`W_N=0`$. The reason for this choice is that the authors in , did not include the contributions coming from the imaginary parts of the virtual corrections and the $`Z`$-boson propagator so that the normal polarization vanishes. Furthermore in one has adopted three different bases for the spin vector which are called the helicity, the beamline and the off-diagonal bases. The helicity bases is defined by the condition
$`\mathrm{cos}\xi =\pm 1.`$ (4.18)
which is equivalent to our choice for the longitudinal polarization of the quark. The beamline basis is given by
$`\mathrm{cos}\xi ={\displaystyle \frac{\mathrm{cos}\theta +\beta }{1+\beta \mathrm{cos}\theta }},\beta =\sqrt{1\rho }.`$ (4.19)
This condition only applies to the Born and the soft plus virtual gluon contribution where there is no hard gluon emission. Here the spin of the quark is polarized along the positron momentum in the rest frame of the quark. The above equation is not valid when there is hard gluon emission. In the later case one obtains
$`\mathrm{cos}\xi ={\displaystyle \frac{x\mathrm{cos}\theta +\alpha _x}{x+\alpha _x\mathrm{cos}\theta }},\alpha _x=\sqrt{x^2\rho },`$ (4.20)
which reproduces the Eq.(4.19) in the limit $`x1`$ (soft gluon region) where $`\alpha _x\beta `$. Since $`\mathrm{cos}\xi `$ in Eq.(4.20) is a function of the integration variable $`x`$ in Eq. (2) the choice above cannot be used in those expressions where the $`x`$ integration is already done. Therefore in the case of hard gluon radiation the integral over $`x`$ is very complicated since this variable appears in the numerator as well as denominator of the expression for $`\mathrm{cos}\xi `$. This will lead to a non-trivial dependence of the cross section on $`\mathrm{cos}\theta `$. Note that the choice in Eq.(4.19) at the level of hard gluon emission is at variance with the interpretation that the spin of the top quark is polarized along the positron momentum direction in the rest frame of the quark. The third choice, called off-diagonal basis, corresponds to the case when the like-spin configuration of the quark anti-quark pair vanishes identically. As for the beamline basis this choice leads to an $`x`$-dependence of $`\mathrm{cos}\xi `$ which is non-trivial so that it can be only applied to the Born and soft plus virtual gluon contribution.
Summarizing the results obtained in this paper we have presented the complete first order QCD corrections to polarized (heavy) quark production. The most of these corrections were already calculated in -. We agree with these results except those obtained for the normal polarization in . Further we were able to compress the expressions as far as possible by minimizing the number of polylogarithms. Moreover we also computed the second order QCD corrections to the production of light quarks when the latter are longitudinally polarized. Furthermore we discovered that the massless quark approach which was so successful to compute the higher order corrections to unpolarized quantities in the kinematical regime $`Qm`$ failed in the case of the longitudinal polarization. As we have seen for bottom-quark production this was not only due to the appearance of anomalous terms but also to the slow convergence to the zero mass limit. Our results reveal that the QCD corrections to the longitudinal polarization are rather small for light as well as heavy quarks except for very small or very large values of the scattering angle $`\theta `$. These corrections become more important for the transverse polarization whereas in the case of the normal polarization they completely dominate the process. For this reason it is very important to compute the order $`\alpha _s^2`$ corrections to the normal polarization which is a difficult task.
ACKNOWLEDGMENTS
V. Ravindran would like to thank J. Blümlein, H.S. Mani and S.D. Rindani for discussions. The work of W.L. van Neerven was supported by the EC network ‘QCD and Particle Structure’ under contract No. FMRX–CT98–0194.
## Appendix A Appendix A
Since the leptonic current is conserved for massless leptons, which implies $`q_\mu ^{\mu \nu }=q_\nu ^{\mu \nu }=0`$, we present only those partonic tensors which contribute to the cross section (see Eqs. (2), (2.13)). In the Born approximation (3.4) we obtain
$`W_1^{(V_1V_2),(0)}`$ $`=`$ $`\left(v_q^{V_1}v_q^{V_2}\left[{\displaystyle \frac{1}{4}}\beta \right]+a_q^{V_1}a_q^{V_2}\left[{\displaystyle \frac{1}{4}}\beta ^3\right]\right)\delta (1x),`$
$`W_2^{(V_1V_2),(0)}`$ $`=`$ $`\left(v_q^{V_1}v_q^{V_2}+a_q^{V_1}a_q^{V_2}\right)\beta \delta (1x),`$
$`W_3^{(V_1V_2),(0)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2}\right)\beta \delta (1x),`$
$`W_4^{(V_1V_2),(0)}`$ $`=`$ $`0,`$
$`W_5^{(V_1V_2),(0)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2}\right)\beta \delta (1x),`$
$`W_6^{(V_1V_2),(0)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2}\right)\beta \delta (1x),`$
$`W_7^{(V_1V_2),(0)}`$ $`=`$ $`0,`$
$`W_8^{(V_1V_2),(0)}`$ $`=`$ $`\left(a_q^{V_1}a_q^{V_2}\right)\beta \delta (1x),`$
$`W_9^{(V_1V_2),(0)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(v_q^{V_1}v_q^{V_2}+a_q^{V_1}a_q^{V_2}\right),\beta \delta (1x)`$
$`W_{10}^{(V_1V_2),(0)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2}\right),\beta \delta (1x)`$
$`W_{11}^{(V_1V_2),(0)}`$ $`=`$ $`0,`$ (A.1)
with $`V_1,V_2=\gamma ,Z`$. For the order $`\alpha _s`$ correction (3) it is convenient to decompose the partonic structure functions $`W_i(i=19)`$ as follows
$`W_i^{(V_1V_2),(1)}`$ $`=`$ $`W_i^{(V_1V_2),V+S}+W_i^{(V_1V_2),HARD},`$
$`\text{with}W_i^{(V_1V_2),V+S}`$ $`=`$ $`W_i^{(V_1V_2),VIRT}+W_i^{(V_1V_2),SOFT}.`$ (A.2)
The soft plus virtual gluon contributions are given by
$`W_1^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{\beta }{2}}\left[v_q^{V_1}v_q^{V_2}𝒞^{V+S}+a_q^{V_1}a_q^{V_2}\beta ^2\left(2\mathrm{Re}𝒞_2+𝒞^{V+S}\right)\right]\delta (1x),`$
$`W_2^{(V_1V2)),V+S}`$ $`=`$ $`2\beta \left[v_q^{V_1}v_q^{V_2}\left(\mathrm{Re}𝒞_2+𝒞^{V+S}\right)+a_q^{V_1}a_q^{V_2}\left(2\mathrm{Re}𝒞_2+𝒞^{V+S}\right)\right]\delta (1x),`$
$`W_3^{(V_1V_2),V+S}`$ $`=`$ $`\beta [(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})(\mathrm{Re}𝒞_2+𝒞^{V+S})`$
$`(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2})i\mathrm{Im}𝒞_2]\delta (1x),`$
$`W_4^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{2\beta }{\rho }}[(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})\mathrm{Re}𝒞_2`$
$`+(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2})i\mathrm{Im}𝒞_2]\delta (1x),`$
$`W_5^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\beta [(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})({\displaystyle \frac{1}{\rho }}\mathrm{Re}𝒞_2+3\mathrm{Re}𝒞_2+2𝒞^{V+S})`$
$`+(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2})({\displaystyle \frac{1}{\rho }}i\mathrm{Im}𝒞_2i\mathrm{Im}𝒞_2)]\delta (1x),`$
$`W_6^{(V_1V_2),V+S}`$ $`=`$ $`\beta [(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})(\mathrm{Re}𝒞_2+𝒞^{V+S})`$
$`(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2})i\mathrm{Im}𝒞_2]\delta (1x),`$
$`W_7^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{\beta }{\rho }}\left[v_q^{V_1}v_q^{V_2}\mathrm{Re}𝒞_2+a_q^{V_1}a_q^{V_2}\mathrm{Re}𝒞_3\right]\delta (1x),`$
$`W_8^{(V_1V_2),V+S}`$ $`=`$ $`\beta \left[a_q^{V_1}a_q^{V_2}\left(4\mathrm{Re}𝒞_2{\displaystyle \frac{1}{\rho }}\mathrm{Re}𝒞_3+2𝒞^{V+S}\right)\right]\delta (1x),`$
$`W_9^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{\beta }{2}}[v_q^{V_1}v_q^{V_2}({\displaystyle \frac{1}{\rho }}\mathrm{Re}𝒞_2\mathrm{Re}𝒞_22𝒞^{V+S})`$
$`+a_q^{V_1}a_q^{V_2}(4\mathrm{Re}𝒞_2+{\displaystyle \frac{1}{\rho }}\mathrm{Re}𝒞_32𝒞^{V+S})]\delta (1x),`$
$`W_{10}^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\beta [(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2})({\displaystyle \frac{1}{\rho }}\mathrm{Re}𝒞_2+3\mathrm{Re}𝒞_2+2𝒞^{V+S})`$
$`+(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})({\displaystyle \frac{1}{\rho }}i\mathrm{Im}𝒞_2i\mathrm{Im}𝒞_2)]\delta (1x),`$
$`W_{11}^{(V_1V_2),V+S}`$ $`=`$ $`{\displaystyle \frac{2\beta }{\rho }}\left[v_q^{V_1}v_q^{V_2}i\mathrm{Im}𝒞_2\right]\delta (1x).`$ (A.3)
The hard gluon contributions are
$`W_1^{(V_1V_2),HARD}`$ $`=`$ $`C_Fv_q^{V_1}v_q^{V_2}[{\displaystyle \frac{1}{\alpha _x(44x+\rho )^2(1x)}}(20\rho 11\rho ^2\rho ^3`$
$`+8x+18\rho x2\rho ^2x2\rho ^3x+8x^2+58\rho x^2+23\rho ^2x^2`$
$`+\rho ^3x^244x^374\rho x^38\rho ^2x^3+32x^4+18\rho x^44x^5)`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{4\alpha _x^2(1x)}}(\rho \rho ^22\rho ^2x2x^2+7\rho x^2+\rho ^2x^2`$
$`2\rho x^32x^4)]`$
$`+C_Fa_q^{V_1}a_q^{V_2}[{\displaystyle \frac{1}{\alpha _x(44x+\rho )^2(1x)}}(20\rho +25\rho ^2+16\rho ^3`$
$`+2\rho ^4+8x+10\rho x80\rho ^2x20\rho ^3x+8x^2+50\rho x^2`$
$`+55\rho ^2x^244x^334\rho x^3+2\rho ^2x^3+32x^46\rho x^44x^5)`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{4\alpha _x^2(1x)}}(\rho +2\rho ^38\rho ^2x2x^2+7\rho x^2+2\rho x^32x^4)],`$
$`W_2^{(V_1V_2),HARD}`$ $`=`$ $`C_Fv_q^{V_1}v_q^{V_2}[{\displaystyle \frac{4}{\alpha _x^3(44x+\rho )^2(1x)}}(4\rho +35\rho ^2+17\rho ^3+2\rho ^4`$
$`+24x82\rho x110\rho ^2x22\rho ^3x40x^2+186\rho x^2+85\rho ^2x^2`$
$`+\rho ^3x^2+4x^3126\rho x^38\rho ^2x^3+16x^4+18\rho x^44x^5)`$
$`{\displaystyle \frac{\mathrm{ln}(\xi )}{\alpha _x^4(1x)}}(\rho \rho ^22\rho ^3+10\rho ^2x+2x^213\rho x^2\rho ^2x^2`$
$`+2\rho x^3+2x^4)]`$
$`+C_Fa_q^{V_1}a_q^{V_2}[{\displaystyle \frac{4}{\alpha _x^3(44x+\rho )^2(1x)}}(4\rho +31\rho ^2+18\rho ^3+2\rho ^4`$
$`+24x106\rho x108\rho ^2x24\rho ^3x40x^2+274\rho x^2`$
$`+95\rho ^2x^2+2\rho ^3x^2+4x^3242\rho x^318\rho ^2x^3+16x^4+82\rho x^4`$
$`+2\rho ^2x^44x^512\rho x^5)`$
$`{\displaystyle \frac{\mathrm{ln}(\xi )}{\alpha _x^4(1x)}}(\rho 2\rho ^22\rho ^3+12\rho ^2x+2x^215\rho x^22\rho ^2x^2`$
$`+6\rho x^3+2x^42\rho x^4)],`$
$`W_3^{(V_1V_2),HARD}`$ $`=`$ $`{\displaystyle \frac{1}{2}}C_F(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})[{\displaystyle \frac{2}{\alpha _x^3(44x+\rho )^2(1x)}}(68\rho ^2+35\rho ^3`$
$`+4\rho ^4128\rho x208\rho ^2x45\rho ^3x+48x^2+332\rho x^2`$
$`+172\rho ^2x^2+5\rho ^3x^2112x^3304\rho x^352\rho ^2x^33\rho ^3x^3`$
$`+104x^4+156\rho x^4+24\rho ^2x^464x^556\rho x^5+24x^6)`$
$`{\displaystyle \frac{\mathrm{ln}(\xi )}{2\alpha _x^4(1x)}}(3\rho ^2+4\rho ^32\rho x21\rho ^2x+18\rho x^2+5\rho ^2x^2`$
$`4x^32\rho x^33\rho ^2x^3+6\rho x^44x^5)],`$
$`W_4^{(V_1V_2),HARD}`$ $`=`$ $`{\displaystyle \frac{1}{2}}C_F(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})[{\displaystyle \frac{8}{\alpha _x^7(44x+\rho )^2}}(64\rho ^2+124\rho ^3`$
$`+45\rho ^4+4\rho ^5528\rho ^2x404\rho ^3x63\rho ^4x+112\rho x^2`$
$`+976\rho ^2x^2+319\rho ^3x^2+6\rho ^4x^2+96\rho x^3432\rho ^2x^318\rho ^3x^3`$
$`176x^4764\rho x^4158\rho ^2x^49\rho ^3x^4+432x^5+756\rho x^5`$
$`+75\rho ^2x^5336x^6206\rho x^6\rho ^2x^6+80x^7+6\rho x^7)`$
$`{\displaystyle \frac{2\mathrm{ln}(\xi )}{\alpha _x^6}}(13\rho ^2+4\rho ^318\rho x39\rho ^2x+58\rho x^2+10\rho ^2x^2`$
$`12x^321\rho x^3+4x^4+\rho x^4)],`$
$`W_5^{(V_1V_2),HARD}`$ $`=`$ $`{\displaystyle \frac{1}{2}}C_F(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})[{\displaystyle \frac{2}{\alpha _x^3(44x+\rho )^2(1x)}}(80\rho +96\rho ^2`$
$`+37\rho ^3+4\rho ^496x440\rho x272\rho ^2x46\rho ^3x+400x^2`$
$`+736\rho x^2+193\rho ^2x^2+\rho ^3x^2568x^3446\rho x^314\rho ^2x^3`$
$`+320x^4+76\rho x^4+\rho ^2x^456x^56\rho x^5)`$
$`{\displaystyle \frac{\mathrm{ln}(\xi )}{2\alpha _x^4(1x)}}(4\rho +5\rho ^2+4\rho ^326\rho x22\rho ^2x+8x^2`$
$`+49\rho x^2+\rho ^2x^216x^38\rho x^3+\rho x^4)],`$
$`W_6^{(V_1V_2),HARD}`$ $`=`$ $`{\displaystyle \frac{1}{2}}C_F(v_q^{V_1}a_q^{V_2}+a_q^{V_1}v_q^{V_2})[{\displaystyle \frac{8}{\alpha _x(44x+\rho )^2(1x)}}(8+12\rho `$
$`+8\rho ^2+\rho ^3+12x31\rho x9\rho ^2x2x^2+18\rho x^2+\rho x^32x^4)`$
$`+{\displaystyle \frac{2\mathrm{ln}(\xi )}{\alpha _x^2(1x)}}(\rho ^2+x3\rho x+x^3)],`$
$`W_7^{(V_1V_2),HARD}`$ $`=`$ $`C_Fv_q^{V_1}v_q^{V_2}[{\displaystyle \frac{2}{\alpha _x^3(44x+\rho )^2}}(64\rho 36\rho ^24\rho ^3+96x`$
$`+172\rho x+39\rho ^2x200x^2102\rho x^2+\rho ^2x^2+104x^36\rho x^3)`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{2\alpha _x^4}}(4\rho 4\rho ^2+15\rho x8x^2+\rho x^2)]`$
$`+C_Fa_q^{V_1}a_q^{V_2}[{\displaystyle \frac{2}{\alpha _x^3(44x+\rho )^2}}(64\rho 16\rho ^2\rho ^3+96x`$
$`+132\rho x+\rho ^2x3\rho ^3x200x^246\rho x^2+19\rho ^2x^2`$
$`+120x^322\rho x^316x^4)`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{2\alpha _x^4}}(4\rho \rho ^2+15\rho x3\rho ^2x8x^2+\rho x^2)],`$
$`W_8^{(V_1V_2),HARD}`$ $`=`$ $`C_Fa_q^{V_1}a_q^{V_2}[{\displaystyle \frac{2}{\alpha _x(44x+\rho )^2(1x)}}(52\rho +31\rho ^2+4\rho ^356x`$
$`130\rho x34\rho ^2x+112x^2+72\rho x^2\rho ^2x^256x^3+6\rho x^3)`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{2\alpha _x^2(1x)}}(\rho 4\rho ^2+10\rho x8x^2+\rho x^2)],`$
$`W_9^{(V_1V_2),HARD}`$ $`=`$ $`C_Fv_q^{V_1}v_q^{V_2}[{\displaystyle \frac{1}{\alpha _x(44x+\rho )(1x)}}(817\rho 4\rho ^2`$
$`+22x+18\rho x4x^2+3\rho x^210x^3)]`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{4\alpha _x^2(1x)}}(\rho +4\rho ^24x10\rho x+8x^23\rho x^2+4x^3)]`$
$`+C_Fa_q^{V_1}a_q^{V_2}[{\displaystyle \frac{1}{\alpha _x(44x+\rho )^2(1x)}}(324\rho +\rho ^2120x`$
$`+62\rho x+30\rho ^2x+4\rho ^3x+120x^2144\rho x^235\rho ^2x^2`$
$`8x^3+86\rho x^324x^4)`$
$`+{\displaystyle \frac{\mathrm{ln}(\xi )}{4\alpha _x^2(1x)}}(\rho 4x2\rho x+4\rho ^2x+8x^2`$
$`11\rho x^2+4x^3)],`$
$`W_{10}^{(V_1V_2),HARD}`$ $`=`$ $`{\displaystyle \frac{1}{2}}C_F(v_q^{V_1}a_q^{V_2}a_q^{V_1}v_q^{V_2})[{\displaystyle \frac{2}{\alpha _x(44x+\rho )^2(1x)}}(76\rho +33\rho ^2`$
$`+4\rho ^372x190\rho x38\rho ^2x+144x^2+120\rho x^2+\rho ^2x^2`$
$`72x^36\rho x^3)+{\displaystyle \frac{\mathrm{ln}(\xi )}{2\alpha _x^2(1x)}}(\rho +4\rho ^214\rho x+8x^2+\rho x^2)],`$
$`W_{11}^{(V_1V_2),HARD}`$ $`=`$ $`0.`$ (A.4)
In the expressions above we have introduced the following notations
$`\alpha _x=\sqrt{x^2\rho },\xi ={\displaystyle \frac{\rho 2x2\alpha _x}{\rho 2x+2\alpha _x}}.`$ (A.5)
## Appendix B Appendix B
In this appendix we present the order $`\alpha _s`$ contributions to the structure functions $`𝒯_i`$ Eqs. (2.21)-(2.32) computed in section 3. The unpolarized parts are given by
$`𝒯_1^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{1}{2}}\rho (1+\rho )F_1+\sqrt{\rho }(13\rho )F_2+2(85\rho \rho ^2)F_3+2\beta F_4`$ (B.1)
$`+{\displaystyle \frac{1}{2}}\beta (2+13\rho )+{\displaystyle \frac{1}{2}}(6439\rho 7\rho ^2)i_2(t)+{\displaystyle \frac{1}{4}}(48+36\rho 5\rho ^2)\mathrm{ln}(t)`$
$`+2(43\rho \rho ^2)\mathrm{ln}(t)\mathrm{ln}(1+t)],`$
$`𝒯_2^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{1}{2}}\rho (1+2\rho )F_1+\sqrt{\rho }(14\rho )F_2+2(813\rho +2\rho ^2)F_3+2\beta (1\rho )F_4`$ (B.2)
$`+{\displaystyle \frac{1}{2}}\beta (2+\rho )+{\displaystyle \frac{1}{2}}(64103\rho +18\rho ^2)i_2(2)+{\displaystyle \frac{1}{4}}(48+60\rho 9\rho ^2)\mathrm{ln}(t)`$
$`+2(47\rho )\mathrm{ln}(t)\mathrm{ln}(1+t)],`$
$`𝒯_3^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{1}{2}}\rho (1+\rho )F_1+\sqrt{\rho }(1+3\rho )F_2+2\rho (5\rho )F_3+\beta \rho F_4+2\beta (1\rho )`$ (B.3)
$`+{\displaystyle \frac{1}{2}}\rho (399\rho )i_2(t)+\rho (7+3\rho )\mathrm{ln}(t)+6\rho \mathrm{ln}(t)\mathrm{ln}(1+t)],`$
$`𝒯_4^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{1}{2}}\rho (1+2\rho )F_1+\sqrt{\rho }(1+4\rho )F_2+2\rho (1+2\rho )F_3`$ (B.4)
$`+{\displaystyle \frac{1}{4}}\beta (838\rho +3\rho ^2)+{\displaystyle \frac{7}{2}}\rho (1+2\rho )i_2(t)+{\displaystyle \frac{1}{8}}\rho (32+8\rho 3\rho ^2)\mathrm{ln}(t)`$
$`+2\rho (1+2\rho )\mathrm{ln}(t)\mathrm{ln}(1+t)],`$
$`𝒯_5^{(1)}`$ $`=`$ $`C_F[4\beta (2+\rho )G_1+2(45\rho )G_24\sqrt{\rho }+4\rho +8(1+\rho )\mathrm{ln}(1+t)`$ (B.5)
$`+8\mathrm{ln}(1+t\sqrt{t})+16(1+\rho )\mathrm{ln}(1\sqrt{t})`$
$`+2(24\rho +3\beta \rho 2\beta )\mathrm{ln}(t)],`$
$`\mathrm{Im}𝒯_6^{(1)}`$ $`=`$ $`C_F\left[2\pi \rho \beta \right].`$ (B.6)
The longitudinal polarized structure functions are equal to
$`𝒯_{7,L_1}^{(1)}`$ $`=`$ $`C_F[\beta (2\rho )G_1{\displaystyle \frac{1}{4}}(8+2\rho +3\rho ^2)G_2{\displaystyle \frac{\sqrt{\rho }}{8}}(8+29\rho )+{\displaystyle \frac{1}{8}}(2+35\rho )`$
$`+2(1\rho )\mathrm{ln}(1+t)+{\displaystyle \frac{1}{16}}(32+60\rho 17\rho ^2)\mathrm{ln}(1+t\sqrt{t})`$
$`+4(1\rho )\mathrm{ln}(1\sqrt{t})+{\displaystyle \frac{1}{32}}(32+4\rho +32\beta +72\rho \beta +17\rho ^2)\mathrm{ln}(t)],`$
$`𝒯_{7,L_2}^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{\rho }{2}}(10+3\rho )G_2+{\displaystyle \frac{\sqrt{\rho }}{2}}(8+13\rho ){\displaystyle \frac{1}{2}}(2+19\rho )`$ (B.8)
$`+{\displaystyle \frac{\rho }{4}}(24+7\rho )\mathrm{ln}(1+t\sqrt{t})+{\displaystyle \frac{\rho }{8}}(2452\beta 7\rho )\mathrm{ln}(t)],`$
$`\mathrm{Im}𝒯_{8,L}^{(1)}`$ $`=`$ $`C_F\left[{\displaystyle \frac{\pi }{2}}\rho \beta \right],`$ (B.9)
$`𝒯_{10,L}^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{5\rho }{8}}F_1{\displaystyle \frac{\sqrt{\rho }}{4}}(4+\rho )F_2{\displaystyle \frac{1}{2}}(8+\rho )F_3{\displaystyle \frac{\beta }{2}}F_4+{\displaystyle \frac{\beta }{2}}`$
$`{\displaystyle \frac{1}{8}}(64+3\rho )i_2(t)+{\displaystyle \frac{1}{4}}(123\rho )\mathrm{ln}(t){\displaystyle \frac{1}{2}}(4+3\rho )\mathrm{ln}(t)\mathrm{ln}(1+t)],`$
$`𝒯_{11,L}^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{\rho }{8}}(5\rho )F_1\sqrt{\rho }F_2+{\displaystyle \frac{1}{2}}(8+7\rho 3\rho ^2)F_3+{\displaystyle \frac{\beta }{2}}(1+\rho )F_4`$ (B.11)
$`+{\displaystyle \frac{\beta }{2}}(1+3\rho )+{\displaystyle \frac{1}{8}}(64+61\rho 25\rho ^2)i_2(t)+{\displaystyle \frac{1}{4}}(129\rho +\rho ^2)\mathrm{ln}(t)`$
$`+{\displaystyle \frac{1}{2}}(4+\rho \rho ^2)\mathrm{ln}(t)\mathrm{ln}(1+t)].`$
The transverse polarized structure functions are
$`𝒯_{7,T}^{(1)}`$ $`=`$ $`C_F[\sqrt{\rho }\beta (2\rho )G_1{\displaystyle \frac{\sqrt{\rho }}{4}}(16+7\rho )G_2+{\displaystyle \frac{\sqrt{\rho }}{8}}(48+17\rho ){\displaystyle \frac{\rho }{8}}(62+3\rho )`$ (B.12)
$`+2\sqrt{\rho }(1\rho )\mathrm{ln}(1+t)+{\displaystyle \frac{\sqrt{\rho }}{16}}(40+2\rho 3\rho ^2)\mathrm{ln}(1+t\sqrt{t})`$
$`+4\sqrt{\rho }(1\rho )\mathrm{ln}(1\sqrt{t})+{\displaystyle \frac{\sqrt{\rho }}{32}}(104+62\rho +168\beta +16\rho \beta `$
$`+3\rho ^2)\mathrm{ln}(t)].`$
$`\mathrm{Im}𝒯_{8,T}^{(1)}`$ $`=`$ $`C_F\left[{\displaystyle \frac{\pi }{4}}\sqrt{\rho }\beta (1+\rho )\right],`$ (B.13)
$`𝒯_{10,T}^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{\sqrt{\rho }}{16}}(4+\rho )F_1{\displaystyle \frac{5\rho }{8}}F_2+{\displaystyle \frac{\sqrt{\rho }}{4}}(20+7\rho )F_3{\displaystyle \frac{1}{2}}\sqrt{\rho }\beta F_4{\displaystyle \frac{3}{4}}\sqrt{\rho }\beta `$ (B.14)
$`+{\displaystyle \frac{\sqrt{\rho }}{16}}(156+57\rho )i_2(t)+{\displaystyle \frac{\sqrt{\rho }}{8}}(2613\rho )\mathrm{ln}(t)`$
$`+{\displaystyle \frac{\sqrt{\rho }}{4}}(12+3\rho )\mathrm{ln}(t)\mathrm{ln}(1+t)],`$
$`𝒯_{11,T}^{(1)}`$ $`=`$ $`C_F[{\displaystyle \frac{\sqrt{\rho }}{4}}F_1+{\displaystyle \frac{\rho }{8}}(5+\rho )F_2\sqrt{\rho }F_3+{\displaystyle \frac{1}{8}}\sqrt{\rho }\beta (143\rho )`$ (B.15)
$`{\displaystyle \frac{7\sqrt{\rho }}{4}}i_2(t)+{\displaystyle \frac{\sqrt{\rho }}{16}}(2012\rho +3\rho ^2)\mathrm{ln}(t)\sqrt{\rho }\mathrm{ln}(t)\mathrm{ln}(1+t)].`$
The normal polarized structure functions are
$`𝒯_{9,N}^{(1)}`$ $`=`$ $`C_F\left[{\displaystyle \frac{\pi }{4}}\sqrt{\rho }(1\rho )\right],`$ (B.16)
$`𝒯_{12,N}^{(1)}`$ $`=`$ $`C_F\left[{\displaystyle \frac{\pi }{4}}\sqrt{\rho }\beta (1+\rho )\right],`$ (B.17)
$`\mathrm{Im}𝒯_{13,N}^{(1)}`$ $`=`$ $`C_F[\sqrt{\rho }\beta (\rho 2)G_1+{\displaystyle \frac{\sqrt{\rho }}{4}}(813\rho )G_2+{\displaystyle \frac{\sqrt{\rho }}{8}}(20+9\rho ){\displaystyle \frac{\rho }{8}}(3\rho +26)`$
$`2\sqrt{\rho }(1\rho )\mathrm{ln}(1+t)+{\displaystyle \frac{\sqrt{\rho }}{16}}(242\rho 3\rho ^2)\mathrm{ln}(1+t\sqrt{t})`$
$`4\sqrt{\rho }(1\rho )\mathrm{ln}(1\sqrt{t})+{\displaystyle \frac{\sqrt{\rho }}{32}}(4062\rho 8\beta +48\rho \beta +3\rho ^2)\mathrm{ln}(t)],`$
where $`i_2(x)`$ is defined in . Furthermore the functions $`F_i`$ ($`i=14`$) and $`G_i`$ ($`i=1,2`$) appearing in the expressions above are given by
$`F_1`$ $`=`$ $`i_2(t^3)+4\zeta (2)+{\displaystyle \frac{1}{2}}\mathrm{ln}^2(t)+3\mathrm{ln}(t)\mathrm{ln}(1+t+t^2),`$
$`F_2`$ $`=`$ $`i_2\left(t^{3/2}\right)i_2\left(t^{3/2}\right)+i_2\left(t^{1/2}\right)i_2\left(t^{1/2}\right)+3\zeta (2)`$
$`+2\mathrm{ln}(t)\mathrm{ln}(1+\sqrt{t})2\mathrm{ln}(t)\mathrm{ln}(1\sqrt{t})+{\displaystyle \frac{3}{2}}\mathrm{ln}(t)\mathrm{ln}(1+t\sqrt{t})`$
$`{\displaystyle \frac{3}{2}}\mathrm{ln}(t)\mathrm{ln}(1+t+\sqrt{t}),`$
$`F_3`$ $`=`$ $`i_2(t)+\mathrm{ln}(t)\mathrm{ln}(1t),`$
$`F_4`$ $`=`$ $`6\mathrm{ln}(t)8\mathrm{ln}(1t)4\mathrm{ln}(1+t),`$
$`G_1`$ $`=`$ $`i_2\left(t^{3/2}\right)3i_2\left(t^{1/2}\right)4i_2\left(t^{1/2}\right)i_2\left(t\right)`$
$`{\displaystyle \frac{1}{2}}\zeta (2){\displaystyle \frac{1}{8}}\mathrm{ln}^2(t),`$
$`G_2`$ $`=`$ $`i_2\left({\displaystyle \frac{\sqrt{t}}{1+t}}\right){\displaystyle \frac{1}{8}}\mathrm{ln}^2(t){\displaystyle \frac{1}{2}}\mathrm{ln}(t)\mathrm{ln}(1+t)+{\displaystyle \frac{1}{2}}\mathrm{ln}^2(1+t){\displaystyle \frac{1}{2}}\zeta (2),`$ (B.19)
where the variable t is defined in (3). Further we are interested in the values taken by $`𝒯_i`$ when $`m0`$. In the unpolarized case they become
$`𝒯_1^{(1),m0}`$ $`=`$ $`𝒯_2^{(1),m0}=1,𝒯_3^{(1),m0}=𝒯_4^{(1),m0}=2,`$ (B.20)
$`𝒯_5^{(1),m0}=0i,𝒯_6^{(1),m0}=0,`$
whereas the longitudinal structure functions tend to the limits
$`𝒯_{7,L_1}^{(1),m0}`$ $`=`$ $`{\displaystyle \frac{1}{4}},𝒯_{7,L_2}^{(1),m0}=1,𝒯_{8,L}^{(1),m0}=0,`$ (B.21)
$`𝒯_{10,L}^{(1),m0}=𝒯_{11,L}^{(1),m0}={\displaystyle \frac{1}{2}}.`$
All transverse and normal polarized quantities become zero in the limit $`m0`$.
FIGURE CAPTIONS
Longitudinal polarization $`P_L`$ of the up-quark up to second order in $`\alpha _s`$ at $`Q=M_Z`$ for polarized ($`\lambda _e^{}=\pm 1`$) and unpolarized (unpol) electrons. The positron beam is unpolarized. (a) Born $`P_{L,0}`$ (dashed line); (b) order $`\alpha _s`$ corrected $`P_{L,1}`$ (dashed dotted line); (c) order $`\alpha _s^2`$ corrected $`P_{L,2}`$ (solid line).
Ratios $`R_L`$ of the higher order and lower order corrected longitudinal polarization of the up-quark at $`Q=M_Z`$ for unpolarized electrons and positrons. (a) Ratio of first order corrected and the Born contribution to the polarization $`P_{L,1}/P_{L,0}`$ (dashed line); (b) Ratio of the second order corrected and the first order corrected polarization $`P_{L,1}/P_{L,0}`$ (solid line).
Longitudinal polarization $`P_L`$ of the bottom-quark up to first order in $`\alpha _s`$ at $`Q=M_Z`$ for polarized ($`\lambda _e^{}=\pm 1`$) and unpolarized (unpol) electrons. The positron beam is unpolarized. (a) Born $`P_{L,0}`$ (dashed line); (b) order $`\alpha _s`$ corrected $`P_{L,1}`$ (solid line).
Ratios $`R_L`$ of order $`\alpha _s`$ contributions to various quantities for unpolarized electrons and positrons for bottom production at $`Q=M_Z`$. (a) Ratio of the longitudinal polarization in the massless quark approach with $`m_b=0`$ and the exact longitudinal polarization $`m_b=4.5`$ (dotted line); (b) Ratio of the unpolarized cross section $`d^2\sigma /d\mathrm{\Omega }`$ for $`m_b0`$ and $`m_b=4.5`$ (dashed dotted line); (c) Ratio of the longitudinal polarization in the massive quark approach with $`m_b0`$ and the exact longitudinal polarization for $`m_b=4.5`$ (dashed line).
Same as in Fig. 3 but now for top-quark production at $`Q=500\mathrm{GeV}`$.
Transverse polarization $`P_T`$ of the bottom-quark up to first order in $`\alpha _s`$ at $`Q=M_Z`$ for polarized ($`\lambda _e^{}=\pm 1`$) and unpolarized (unpol) electrons. The positron beam is unpolarized. (a) Born $`P_{T,0}`$ (dashed line); (b) order $`\alpha _s`$ corrected $`P_{T,1}`$ (solid line).
The same as in Fig. 6 but now top-quark production at $`Q=500\mathrm{GeV}`$.
Normal polarization $`P_N`$ of the bottom-quark up to first order in $`\alpha _s`$ at $`Q=M_Z`$ for polarized ($`\lambda _e^{}=\pm 1`$) and unpolarized (unpol) electrons. The positron beam is unpolarized. (a) Born approximation ($`\gamma Z`$-interference term) $`P_{N,0}`$ (dashed line); (b) order $`\alpha _s`$ (QCD) corrected results $`P_{N,1}`$ (solid line).
Normal polarization $`P_N`$ of the top-quark up to first order in $`\alpha _s`$ at $`Q=500\mathrm{GeV}`$ for polarized ($`\lambda _e^{}=\pm 1`$) and unpolarized (unpol) electrons. The positron beam is unpolarized. (a) $`\lambda _e^{}=1`$ (dashed line); (b) unpolarized electrons (solid line); (c) $`\lambda _e^{}=1`$ (dotted line). The lower curves represent the Born approximation ($`\gamma Z`$-interference term) $`P_{N,0}`$ whereas the upper curves stand for the order $`\alpha _s`$ (QCD) corrected results $`P_{N,1}`$.
Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
|
warning/0006/hep-ph0006178.html
|
ar5iv
|
text
|
# CP Violation in Λ→𝑝𝜋⁻: SM vs New Physics
## Introduction
In non-leptonic hyperon decays such as $`\mathrm{\Lambda }p\pi ^{}`$ it is possible to search for CP violation by comparing the decay with the corresponding anti-hyperon decay op . The Fermilab experiment E871 is currently searching for CP violation in such a decay and is sensitive to certain types of physics beyond the standard model. The observable provides information that is complementary to that obtained from the measurement of $`ϵ^{}/ϵ`$.
The reaction of interest is the decay of a polarized $`\mathrm{\Lambda }`$, with known polarization $`\stackrel{}{w}`$, into a proton (whose polarization is not measured) and a $`\pi ^{}`$ with momentum $`q`$. The final $`p\pi ^{}`$ state can be in an S-wave or a P-wave, and in an $`I=1/2`$ or $`I=3/2`$ state. The observables are the total decay rate and a correlation in the decay distribution of the form
$$\frac{d\mathrm{\Gamma }}{d\mathrm{\Omega }}1+\alpha \stackrel{}{w}\stackrel{}{q}$$
(1)
The branching ratio for this mode is $`63.9\%`$ and the parameter $`\alpha `$ has been measured to be $`\alpha =0.64`$ pdb . The CP violation in question involves a comparison of the parameter $`\alpha `$ with the corresponding parameter $`\overline{\alpha }`$ for the reaction $`\overline{\mathrm{\Lambda }}\overline{p}\pi ^+`$.
It is standard to write the amplitudes in terms of their isospin components in the form
$`S`$ $`=`$ $`S_1e^{i\delta _1^S}+S_3e^{i\delta _3^S}`$
$`P`$ $`=`$ $`P_1e^{i\delta _1^P}+P_3e^{i\delta _3^P}`$ (2)
A $`\mathrm{\Delta }I=1/2`$ rule is observed experimentally, $`S_3/S_10.026`$ and $`P_3/P_1=0.03\pm 0.03`$ over . The strong $`\pi N`$ scattering phases have been measured for the $`I=1/2`$ channel, $`\delta _1^S6^o`$ and $`\delta _1^P1^o`$ roper . The $`I=3/2`$ scattering phases have been measured with large errors but are not needed here.
To discuss CP violation, we allow the amplitudes in Eq. 2 to have a CP violating weak phase, $`S_iS_i\mathrm{exp}(i\varphi _i^S)`$ and $`P_iP_i\mathrm{exp}(i\varphi _i^P)`$ and compare the pair of CP conjugate reactions. CP symmetry predicts that $`\mathrm{\Gamma }=\overline{\mathrm{\Gamma }}`$ and that $`\overline{\alpha }=\alpha `$. One therefore defines the CP-odd observables
$`\mathrm{\Delta }`$ $``$ $`{\displaystyle \frac{\mathrm{\Gamma }\overline{\mathrm{\Gamma }}}{\mathrm{\Gamma }+\overline{\mathrm{\Gamma }}}}\sqrt{2}{\displaystyle \frac{S_3}{S_1}}\mathrm{sin}(\delta _3^S\delta _1^S)\mathrm{sin}(\varphi _3^S\varphi _1^S)`$
$`A(\mathrm{\Lambda }_{}^0)`$ $``$ $`{\displaystyle \frac{\alpha +\overline{\alpha }}{\alpha \overline{\alpha }}}\mathrm{sin}(\delta _1^P\delta _1^S)\mathrm{sin}(\varphi _1^P\varphi _1^S)0.12\mathrm{sin}(\varphi _1^P\varphi _1^S)`$ (3)
The partial rate asymmetry is very small, being suppressed by three small factors, $`S_3/S_1`$, strong phases, and weak phases. It represents an interference between amplitudes with $`\mathrm{\Delta }I=1/2`$ and $`\mathrm{\Delta }I=3/2`$. The asymmetry $`A(\mathrm{\Lambda }_{}^0)`$, on the other hand, is not suppressed by the $`\mathrm{\Delta }I=1/2`$ rule, as it originates in an interference of S and P-waves within the $`\mathrm{\Delta }I=1/2`$ transition. For this reason, the observable $`A(\mathrm{\Lambda }_{}^0)`$ is qualitatively different from $`ϵ^{}/ϵ`$.
The experiment E871 at Fermilab produces the polarized $`\mathrm{\Lambda }`$ from the weak decay $`\mathrm{\Xi }^{}\mathrm{\Lambda }\pi ^{}`$ and for this reason what they measure is actually the combination $`A(\mathrm{\Lambda }_{}^0)+A(\mathrm{\Xi }_{}^{})`$. Their expected sensitivity is $`10^4`$. The weak phases in $`\mathrm{\Xi }^{}`$ decay (within the standard model) have been estimated to be about two times smaller than those in $`\mathrm{\Lambda }`$ decay dhp . Similarly, the strong phases in $`\mathrm{\Xi }^{}`$ decay are estimated to be of order $`1^o`$ lsw ; dp and therefore five times smaller than the strong phase difference in $`\mathrm{\Lambda }`$ decay. For these two reasons we expect that the E871 measurement will be dominated by $`A(\mathrm{\Lambda }_{}^0)`$.
## Standard Model
Within the standard model one writes the $`|\mathrm{\Delta }S|=1`$ effective weak Hamiltonian as a sum of four-quark operators multiplied by Wilson coefficients in the usual way,
$$H=\frac{G_F}{\sqrt{2}}V_{ud}^{}V_{us}\underset{i=1}{\overset{12}{}}c_i(\mu )Q_i(\mu )$$
(4)
This is, of course, the same effective Hamiltonian responsible for Kaon non-leptonic decays and is very well known. In particular the Wilson coefficients, $`c_i(\mu )`$ have been calculated in detail by Buras and his collaborators bbh . The remaining problem is to calculate the matrix elements of the four-quark operators between hadronic states. This problem has not been resolved yet, and there is large theoretical uncertainty in these matrix elements. The usual way to proceed (which is the same as in kaon physics) is to take the real part of the matrix element from experiment (assuming CP conservation) and to use the calculated imaginary parts.
Unlike the case of $`ϵ^{}`$, where both $`\mathrm{\Delta }I=1/2,3/2`$ amplitudes are important, $`A(\mathrm{\Lambda }_{}^0)`$ is dominated by CP violation in $`\mathrm{\Delta }I=1/2`$ amplitudes. One expects that the asymmetry will be dominated by the penguin operator with small corrections from other operators. A detailed study using vacuum saturation to estimate the matrix elements supports the view that $`Q_6`$ is dominantly responsible for $`A(\mathrm{\Lambda }_{}^0)`$ hsv .
Once we have determined that only $`Q_6`$ is important, the strategy is to calculate the matrix elements of the form $`<B^{}|Q_6|B>`$ using a model, and then use these results to treat the non-leptonic hyperon decay at leading order in chiral perturbation theory as sketched in Figure 1. Equivalently, the S-waves are obtained with a soft-pion theorem and the P-waves with baryon poles. At present, the baryon to baryon matrix elements are taken from the MIT bag model calculation of Ref. dghp .
It is difficult to quantify the theoretical error in this calculation. There are the obvious uncertainties in the short distance parameters as well as errors in the value of the strong phases. However, of greater concern is the issue of assigning an error to the hadronic matrix elements. Even if we assume that the baryon to baryon matrix elements calculated in the MIT bag model are exact, we know from the study of CP conserving amplitudes that non-leading order terms in chiral perturbation theory can be as large as the leading order amplitudes. For example, the s-wave imaginary part calculated in vacuum saturation, is a higher order correction to the bag-model plus soft pion theorem amplitude outlined above, but it is larger hsv . To get an idea for the impact of this error we assign an overall error of a factor of two to the calculated matrix elements plus an overall 30% uncorrelated error between S and P-waves. Combining all this results in,
$$A(\mathrm{\Lambda }_{}^0)=(3.0\pm 2.6)\times 10^5.$$
(5)
## Beyond the Standard Model
There have been several estimates of $`A(\mathrm{\Lambda }_{}^0)`$ beyond the standard model. For the most part these studies discuss specific models, concentrating on one or a few operators and normalizing the strength of CP violation by fitting $`ϵ`$. Some of these results (which have not been updated to incorporate current constraints on model parameters) are:
$$A(\mathrm{\Lambda }_{}^0)=\{\begin{array}{cc}2\times 10^5\hfill & \text{SM }\text{dhp}\hfill \\ 2\times 10^5\hfill & \text{3 Higgs }\text{dhp}\hfill \\ 0\hfill & \text{Superweak}\hfill \\ 6\times 10^4\hfill & \text{LR }\text{chp}\hfill \end{array}$$
(6)
Perhaps a more interesting question is whether it is possible to have large CP violation in hyperon decays in view of what is known about $`ϵ`$ and $`ϵ^{}`$. This question has been addressed in a model independent way by considering all the CP violating operators that can be constructed at dimension 6 that are compatible with the symmetries of the standard model hvop . With this general formalism one can compute the contributions of each new CP violating phase to $`ϵ,ϵ^{}`$, and $`A(\mathrm{\Lambda }_{}^0)`$. Of course, there is the caveat that the hadronic matrix elements cannot be computed reliably. Nevertheless, one finds in general, that parity even operators generate a weak phase $`\varphi _1^P`$ and do not contribute to $`ϵ^{}`$. Their strength can be bound from the long distance contributions to $`ϵ`$ that they induce. Similarly, the parity-odd operators generate a weak phase $`\varphi _1^S`$ and contribute to $`ϵ^{}`$ (but not to $`ϵ`$).
The constraints from $`ϵ^{}`$ turn out to be much more stringent than those from $`ϵ`$, and, therefore, the only natural way (without invoking fine cancellations between different operators) to obtain a large $`A(\mathrm{\Lambda }_{}^0)`$ given what we know about $`ϵ^{}`$ is with new CP-odd, P-even interactions. Within the model independent analysis, one can identify a few new operators with the required properties, that can lead to hvop ,
$$A(\mathrm{\Lambda }_{}^0)5\times 10^4\mathrm{P}\mathrm{even},\mathrm{CP}\mathrm{odd}$$
(7)
This possibility has been revisited recently, motivated in part by the observation of $`ϵ^{}`$. The average value $`ϵ^{}/ϵ=(21.2\pm 4.6)\times 10^4`$ epp appears to be larger than the standard model central prediction with simplistic models for the hadronic matrix elements. This has motivated searches for new sources of CP violation that can give large contributions to $`ϵ^{}`$, in particular, within supersymmetric theories. One such scenario generates a large $`ϵ^{}`$ through an enhanced gluonic dipole operator mm . The effective Hamiltonian is of the form
$`H_{eff}`$ $`=`$ $`(\delta _{12}^d)_{LR}C_g\overline{d}\sigma _{\mu \nu }t^a(1+\gamma _5)sG^{a\mu \nu }`$ (8)
$`+`$ $`(\delta _{12}^d)_{RL}C_g\overline{d}\sigma _{\mu \nu }t^a(1\gamma _5)sG^{a\mu \nu }`$
The quantity $`C_g`$ is a known loop factor, and the $`(\delta _{12}^d)_{LR,RL}`$ originate in the supersymmetric theory ggms . Depending on the correlation between the value of $`(\delta _{12}^d)_{LR}`$ and $`(\delta _{12}^d)_{RL}`$ one gets different scenarios for $`ϵ^{}`$ and $`A(\mathrm{\Lambda }_{}^0)`$ as shown in Figure 2 hmpv .
For example, if only $`(\delta _{12}^d)_{LR}`$ is non-zero, there can be a large $`ϵ^{}`$ mm , but $`A(\mathrm{\Lambda }_{}^0)`$ is small as in the 3-Higgs model of dhp . However, in models in which $`\mathrm{Im}(\delta _{12}^d)_{LR}=\mathrm{Im}(\delta _{12}^d)_{RL}`$ the CP violating operator is parity-even. In this case there is no contribution to $`ϵ^{}`$ and $`A(\mathrm{\Lambda }_{}^0)`$ can be as large as $`10^3`$ hmpv . It is interesting that this type of model is not an ad-hoc model to give a large $`A(\mathrm{\Lambda }_{}^0)`$, but is a type of model originally designed to naturally reproduce the relation $`\lambda =\sqrt{m_d/m_s}`$, as in Ref. bdh , for example.
## Conclusion and Comments
E871 is expected to reach a sensitivity of $`10^4`$ for the observable $`A(\mathrm{\Lambda }_{}^0)+A(\mathrm{\Xi }_{}^{})`$.
* $`A(\mathrm{\Lambda }_{}^0)`$ is likely to be significantly larger than $`A(\mathrm{\Xi }_{}^{})`$.
* $`A(\mathrm{\Lambda }_{}^0)=(3.0\pm 2.6)\times 10^5`$ is our current best guess for the standard model and the theoretical uncertainty is dominated by our inability to calculate hadronic matrix elements reliably. For this reason, the error assigned to this quantity is no more than an educated guess.
* $`A(\mathrm{\Lambda }_{}^0)`$ can be much larger if CP violation originates in P-even new physics. A specific realization of this scenario is possible in supersymmetric theories leading to $`A(\mathrm{\Lambda }_{}^0)`$ as large as $`10^3`$.
I conclude that a non-zero measurement by E871 is not only possible but that it would provide valuable complementary information to what we already know from $`ϵ^{}`$.
Finally I would like to mention two related issues. A search for $`\mathrm{\Delta }S=2`$ hyperon non-leptonic decays is also a useful enterprise as it provides information that is complementary to what we know from $`K\overline{K}`$ mixing hvds . A CP violating rate asymmetry in $`\mathrm{\Omega }\mathrm{\Xi }\pi `$ decay can be as large as $`2\times 10^5`$ within the standard model (and up to ten times larger beyond), much larger than the corresponding rate asymmetries in octet-hyperon decay tv .
This work was supported by DOE under contract number DE-FG02-92ER40730. This talk summarizes work done in collaboration with John Donoghue, Xiao-Gang He, Hitoshi Murayama, Sandip Pakvasa, Herbert Steger and Jusak Tandean.
|
warning/0006/hep-th0006185.html
|
ar5iv
|
text
|
# On covariant limit-from𝜅-symmetry fixing and the relation between the NSR string and the Type II GS superstring
## 1 Introduction
In Refs., was proved the classical equivalence between the massless $`N=1`$ Brink-Schwarz superparticle and the massless spinning particle , possessing the $`n=1`$ local worldsheet supersymmetry. However, their first-quantized spectra of states are different. On the other hand, it is well known that in $`D=10`$ the GSO-projected NSR string , and the GS superstring , describe the same set of quantum states. Thus, there naturally arises a question of establishing a classical relation between these models in the manifestly $`SO(1,9)`$ Lorentz covariant manner<sup>1</sup><sup>1</sup>1In such an equivalence between the NSR string and the Type II GS superstring was proved by imposing the $`SO(1,9)`$ Lorentz noncovariant light-cone gauge to fix $`\kappa `$symmetry.. Although some interesting results towards a solution of this problem were obtained in ,, this issue is far from being clear.
The crucial point in solving this problem is to find a Cartan-Penrose-type relation between the NSR string and the Type II GS superstring variables $`\psi _\pm ^m`$ and $`\theta ^{\alpha 1,2}`$. We suggest such a relation. It is a direct generalization of that of Ref., for particles and involves commumting $`D=10`$ MW spinors $`\lambda _\pm ^\alpha `$, which are the superpartners of $`\theta ^{\alpha 1,2}`$ with respect to the $`n=(1|1)`$ local worldsheet supersymmetry. It is this $`n=(1|1)`$ local worldsheet supersymmetry that the NSR string possesses.
Another issue which we concern in this paper is the manifestly $`SO(1,9)`$ Lorentz covariant $`\kappa `$symmetry fixing for the Type II GS superstring model. This allows one to deal only with the physical variables and becomes especially important when trying to simplify the GS superstring action in curved backgrounds, in particular, in the intensively studied now AdS ones . To gauge away $`\kappa `$symmetry in the manifestly covariant way we use twistor-like Lorentz-harmonic variables, parametrizing the coset space $`SO(1,9)/SO(1,1)\times SO(8)`$ - <sup>2</sup><sup>2</sup>2The concept of harmonic variables was originally introduced in Ref. to describe gauge theories with extended supersymmetry. and decompose Grassmannian spinors $`\theta ^{\alpha 1,2}`$ on them. We also relate $`\kappa `$symmetry fixed variables $`\theta _{\dot{A}}^{}`$ and $`\theta _A^+`$ to the NSR string physical variables $`\phi _\pm ^i`$, which are the orthogonal to the worldsheet components of $`\psi _\pm ^m`$.
## 2 Relation between the NSR string and the Type II GS superstring
As it is known from the superembedding approach<sup>3</sup><sup>3</sup>3The first models invariant under both the local worldsheet supersymmetry and the local target-space one were proposed in and paved the way for the so-called spinning superparticle and spinning superstring models . They correspond to the unresricted embedding and describe more physical states than the conventional superparticle and superstring theories. (for review see ), an embedding of a supersurface into a target-superspace is governed by the so called geometrodynamical equation ,,-, which asserts that the pullback of the target-superspace supervielbein bosonic components onto the supersurface Grassmannian directions has to vanish, i.e.
$$𝒟_{\widehat{q}}Z^ME_M^a(Z)=0,$$
(1)
where $`Z^M`$ is the condensed notation for the superspace coordinates, considered as the worldvolume scalar superfields, $`E_M^a(Z)`$ are the tangent space vector components of the target-superspace supervielbein 1-form and $`𝒟_{\widehat{q}}`$ is the supersurface Grassmannian covariant derivative. Index $`\widehat{q}`$ stands for the direct product of the supersurface Lorentz group $`SO(1,p)`$ spinor index and the one corresponding to the fundamental representation(s) of the automorphisms group of the extended supersymmetry on the worldvolume. Written in such a form superemdedding equation is valid for all known types of branes coupled to corresponding supergravity backgrounds. We, however, will concentrate on the Type II superstrings embedded into flat $`D=10`$ target-superspace with the bosonic metrics $`\eta ^{mn}=(+,,\mathrm{},)`$. Then (1) reads
$$\mathrm{\Pi }_{\pm q}^m=𝒟_{\pm q}X^mi\left(𝒟_{\pm q}\mathrm{\Theta }^{\alpha 1}\sigma _{\alpha \beta }^m\mathrm{\Theta }^{\beta 1}+𝒟_{\pm q}\mathrm{\Theta }_\alpha ^2\stackrel{~}{\sigma }^{m\alpha \beta }\mathrm{\Theta }_\beta ^2\right)=0,$$
(2)
for the Type IIA case and
$$\mathrm{\Pi }_{\genfrac{}{}{0pt}{}{+q}{\dot{q}}}^m=𝒟_{\genfrac{}{}{0pt}{}{+q}{\dot{q}}}X^mi\left(𝒟_{\genfrac{}{}{0pt}{}{+q}{\dot{q}}}\mathrm{\Theta }^{\alpha 1}\sigma _{\alpha \beta }^m\mathrm{\Theta }^{\beta 1}+𝒟_{\genfrac{}{}{0pt}{}{+q}{\dot{q}}}\mathrm{\Theta }^{\alpha 2}\sigma _{\alpha \beta }^m\mathrm{\Theta }^{\beta 2}\right)=0,$$
(3)
for the Type IIB case, where Grassmannian superfields $`\mathrm{\Theta }^{\alpha 1}(\mathrm{\Theta }_\alpha ^2)(\sigma ^{\pm 2},\eta ^{\pm q})`$ (Type IIA) or $`\mathrm{\Theta }^{\alpha 1,2}(\sigma ^{\pm 2},\eta ^{\genfrac{}{}{0pt}{}{+q}{\dot{q}}})`$ (Type IIB) are the worldsheet scalars and $`D=10`$ MW spinors, bosonic superfield $`X^m(\sigma ^{\pm 2},\eta ^{\pm q})`$ (Type IIA) or $`X^m(\sigma ^{\pm 2},\eta ^{\genfrac{}{}{0pt}{}{+q}{\dot{q}}})`$ (Type IIB) is also the worldsheet scalar, but $`D=10`$ vector. The superworldsheet is parametrized by 2 bosonic light-cone coordinates $`\sigma ^{\pm 2}`$ and $`2n`$ Grassmannian ones $`\eta ^{\pm q}`$ or $`\eta ^{\genfrac{}{}{0pt}{}{+q}{\dot{q}}}`$ $`(q,\dot{q}=1,\mathrm{},n)`$, that are $`d=2`$ MW spinors with different chiralities. So $`n`$, that can take values from 1 to 8, denotes the number of (anti)chiral local worlsheet supersymmetries. Here we will consider the Type IIB case as a basic one, although all the essential differences concerning the Type IIA case will be outlined.
It is known that in all classically allowed dimensions, except $`D=3`$, this superemedding equation yields the superstring equations of motion, thus unabling the construction of the conventional doubly supersymmetric superfield actions for the Type II superstrings in all cases but $`D=3`$ , . Nonetheless, doubly supersymmetric superfield actions were found for the superparticles - and the heterotic superstrings - in diverse dimensions with various numbers of local worldsheet supersymmetries, as well as, for the nullstrings and the $`D=4N=1`$ supermembrane .
In order to establish a relation between the NSR string and the Type II superstring we need to analyse this superembedding equation for the $`n=(1|1)`$ worldsheet superspace, parametrized by only two Grassmannian variables $`\eta ^\pm `$, which corresponds to the NSR string type worldsheet supersymmetry. Note, that the $`n=(1|1)`$ worldsheet supergravity can be considered superconformally flat . Thus, component expansions of the worldsheet superfields $`X^m`$, $`\mathrm{\Theta }^{\alpha 1,2}`$ acquire the simplest form
$$X^m(\sigma ^{\pm 2},\eta ^\pm )=X^m(\sigma ^{\pm 2})+\frac{i}{\sqrt{8}}\eta ^+\psi _+^m(\sigma ^{\pm 2})+\frac{i}{\sqrt{8}}\eta ^{}\psi _{}^m(\sigma ^{\pm 2})+i\eta ^+\eta ^{}F^m(\sigma ^{\pm 2}),$$
(4)
$$\mathrm{\Theta }^{\alpha 1,2}(\sigma ^{\pm 2},\eta ^\pm )=\theta ^{\alpha 1,2}(\sigma ^{\pm 2})+\eta ^+\lambda _+^{\alpha 1,2}(\sigma ^{\pm 2})+\eta ^{}\lambda _{}^{\alpha 1,2}(\sigma ^{\pm 2})+i\eta ^+\eta ^{}\rho ^{\alpha 1,2}(\sigma ^{\pm 2}).$$
(5)
$`X^m`$ and $`\theta ^{\alpha 1,2}`$ are ordinary GS variables, $`\psi _\pm ^m`$ are the NSR Grassmannian variables, $`\lambda _\pm ^{\alpha 1,2}`$ are the stringy twistor-like variables, $`F^m`$ and $`\rho ^{\alpha 1,2}`$ are redundant auxiliary ones. Covariant derivatives then look like
$$𝒟_\pm =E^{1/2}D_\pm ^0=E^{1/2}\left(\frac{}{\eta ^\pm }i\eta ^\pm _{\pm 2}\right),$$
(6)
where superscript <sup>0</sup> corresponds to the flat superworldsheet. They satisfy the following algebra
$$D_+^0D_+^0=i_{+2},D_{}^0D_{}^0=i_2,\{D_+^0,D_{}^0\}=0.$$
(7)
Fixing the superconformal gauge we impose the chirality conditions on $`\mathrm{\Theta }^{\alpha 1,2}`$ superfields
$$𝒟_{}\mathrm{\Theta }^{\alpha 1}=𝒟_+\mathrm{\Theta }^{\alpha 2}=0,$$
(8)
which on component level are equivalent to
$$\lambda _{}^{\alpha 1}=\rho ^{\alpha 1}=0,\lambda _+^{\alpha 2}=\rho ^{\alpha 2}=0;$$
(9)
$$_2\theta ^{\alpha 1}=_2\lambda _+^{\alpha 1}_2\lambda _+^\alpha =0,_{+2}\theta ^{\alpha 2}=_{+2}\lambda _{}^{\alpha 2}_{+2}\lambda _{}^\alpha =0.$$
(10)
Conditions (8) contain equations of motion and thus put the variables on the mass shell. Upon utilization of (9) superembedding equation yields
$$\psi _{}^m=\sqrt{8}\lambda _{}^\alpha \sigma _{\alpha \beta }^m\theta ^{\beta 2},\psi _+^m=\sqrt{8}\lambda _+^\alpha \sigma _{\alpha \beta }^m\theta ^{\beta 1};$$
(11)
$$\mathrm{\Pi }_{\pm 2}^m=_{\pm 2}X^mi_{\pm 2}\theta ^1\sigma ^m\theta ^1i_{\pm 2}\theta ^2\sigma ^m\theta ^2=\lambda _\pm \sigma ^m\lambda _\pm ;$$
(12)
$$F^m=0.$$
(13)
Applying further equation (10) one recovers the NSR string fermionic equations of motion
$$_{+2}\psi _{}^m=0,_2\psi _+^m=0.$$
(14)
Let us discuss some properties of the obtained formulae. Reprsentation (11) connects in a natural way the NSR string and the Type II GS superstring fermionic variables. The NSR string Grassmannian vectors $`\psi _\pm ^m`$ contain $`9+9`$ components as a result of the two supercurrent constraints. However, as will be seen below, proper solution of these constraints ensures dropping out of the two extra components of $`\psi _\pm ^m`$ from the NSR string action. So, actually only their $`8+8`$ physical components contribute to the action. On the other hand, among $`16+16`$ components of the two MW spinors $`\theta ^{\alpha 1,2}`$ after explicit fixing of $`\kappa `$symmetry (see Sec.3) there remain $`8+8`$ components. Thus, on the constraint shell there is the same number of the Grassmannian degrees of freedom in both formulations of the string theory, as it should be. In the next Section we will find manifest expressions for the physical variables in the NSR and the GS models and relate them to each other. Representation (12) solves the Type II GS superstring Virasoro constraints since the vectors $`\lambda _+\sigma ^m\lambda _+`$ and $`\lambda _{}\sigma ^m\lambda _{}`$ are light-like due to the famous $`10D`$ permutation relation
$$\sigma _{\alpha \beta }^m\sigma _{m\gamma \delta }+\sigma _{\alpha \delta }^m\sigma _{m\beta \gamma }+\sigma _{\alpha \gamma }^m\sigma _{m\delta \beta }=0.$$
(15)
The NSR string and the GS superstring equations of motion are satisfied by virtue of (10-12).
Let us consider the NSR string constraints. The supercurrent constraints
$$\psi _+^m_{+2}X_m=\psi _{}^m_2X_m=0$$
(16)
after substitution of the representations (10,11,12) give rise to the following equations
$$_{+2}\theta ^{\alpha 1}=d_+\lambda _+^\alpha +e_{+2}\theta ^{\alpha 1},_2\theta ^{\alpha 2}=d_{}\lambda _{}^\alpha +e_2\theta ^{\alpha 2}.$$
(17)
The fact that equations (17) contain worldsheet superreparametrization-like terms with only two arbitrary functions $`d_\pm (\sigma ^{\pm 2})`$, as was shown in , amounts to all but two $`\kappa `$symmetry parameteres being fixed. The rest are identified with the $`n=(1|1)`$ worldsheet superreparametrization transformations in order to establish the relation with the NSR string. The stress-tensor constraints
$$_{\pm 2}X^m_{\pm 2}X_m\frac{i}{2}\psi _\pm _{\pm 2}\psi _\pm =0$$
(18)
after substitution of the representations for $`_{\pm 2}X^m`$ (12) reduce to
$$\psi _\pm _{\pm 2}\psi _\pm =0,$$
(19)
since the vectors $`_{\pm 2}X^m`$ are light-like as a result of the equations of motion for $`\theta ^{1,2}`$ (10,17) and the permutation formula for the $`D=10`$ $`\sigma `$matrices (15). From the constraints (19) there follow the equations of motion for the commuting spinors $`\lambda _\pm ^\alpha `$
$$_{+2}\lambda _+^\alpha =f_{+2}\lambda _+^\alpha +g_{+3}\theta ^{\alpha 1},_2\lambda _{}^\alpha =f_2\lambda _{}^\alpha +g_3\theta ^{\alpha 2}.$$
(20)
Equations (17,20) lead to the following expressions for $`_{\pm 2}\psi _\pm ^m`$:
$$_{+2}\psi _+^m=(e_{+2}+f_{+2})\psi _+^m+d_+\lambda _+\sigma ^m\lambda _+,_2\psi _{}^m=(e_2+f_2)\psi _{}^m+d_{}\lambda _{}\sigma ^m\lambda _{}.$$
(21)
After taking into account (11, 12) it is possible to establish the connection between the NSR and the Type II GS string actions
$$S_{GS}=S_{NSR}+\mathrm{\Delta }S,$$
(22)
where
$$S_{NSR}=\frac{2}{c\alpha ^{}}_{+2}X^m_2X_m+\frac{i}{c\alpha ^{}}\left(\psi _+^m_2\psi _{m+}+\psi _{}^m_{+2}\psi _m\right),$$
(23)
$$\begin{array}{c}\mathrm{\Delta }S=\frac{4}{c\alpha ^{}}_2\theta ^1\sigma ^m\theta ^1_{+2}\theta ^2\sigma _m\theta ^2\frac{2}{c\alpha ^{}}\left(_{+2}\theta ^1\sigma ^m\theta ^1_2\theta ^1\sigma _m\theta ^1+_{+2}\theta ^2\sigma ^m\theta ^2_2\theta ^2\sigma _m\theta ^2\right)\\ \frac{8i}{c\alpha ^{}}\left(\lambda _+\sigma ^m\theta ^1_2\lambda _+\sigma _m\theta ^1+\lambda _{}\sigma ^m\theta ^2_{+2}\lambda _{}\sigma _m\theta ^2\right).\end{array}$$
(24)
$`\mathrm{\Delta }S`$ vanishes on the mass shell (10).
## 3 Manifestly $`SO(1,9)`$ Lorentz covariant $`\kappa `$symmetry fixing
Now let us concern the issue of $`\kappa `$symmetry fixing for the Type II GS superstring. To this end let us consider the classically equivalent twistor-like Lorentz harmonic formulation for the Type IIB GS superstring :
$$\begin{array}{c}S=e\left((4\alpha ^{})^{1/2}\left(e^{\mu [+2]}u_m^{[2]}+e^{\mu [2]}u_m^{[+2]}\right)\omega _\mu ^m+c\right)\\ \frac{1}{c\alpha ^{}}ϵ^{\mu \nu }\left[i\omega _\mu ^m\left(_\nu \theta ^1\sigma _m\theta ^1_\nu \theta ^2\sigma _m\theta ^2\right)+_\mu \theta ^1\sigma _m\theta ^1_\nu \theta ^2\sigma _m\theta ^2\right].\end{array}$$
(25)
In the Type IIA case one should replace $`_\nu \theta ^{\alpha 2}\sigma _{m\alpha \beta }\theta ^{\beta 2}`$ with $`_\nu \theta _\alpha ^2\stackrel{~}{\sigma }_m^{\alpha \beta }\theta _\beta ^2`$. In addition to the variables which are present in the standart GS superstring formulation it contains the worldsheet zweinbein $`e^{\mu [\pm 2]}`$ and the light-like Lorentz frame vectors $`u^{m[\pm 2]}`$ tangent to the string worldsheet. These light-like Lorentz frame vectors together with the orthogonal to the worldsheet ones $`u^{m(i)}`$ $`((i)=1,\mathrm{},8)`$ constitute a complete orthonormal basis one can use to expand any $`D=10`$ Minkowski vector. Lorentz frame vectors can be presented as the bilinear combinations of the spinor harmonics $`v_\alpha ^a=(v_{\alpha \dot{A}}^{},v_{\alpha A}^+)`$ or their inverse $`(v^1)_a^\alpha =(v_A^\alpha ,v_{\dot{A}}^{\alpha +})`$ ($`(v^1)_a^\alpha v_\alpha ^b=\delta _a^b`$):
$$u_m^{[+2]}=\frac{1}{8}(v_{\alpha A}^+\stackrel{~}{\sigma }_m^{\alpha \beta }v_{\beta A}^+)=\frac{1}{8}(v_{\dot{A}}^{\alpha +}\sigma _{m\alpha \beta }v_{\dot{A}}^{\beta +}),$$
(26)
$$u_m^{[2]}=\frac{1}{8}(v_{\alpha \dot{A}}^{}\stackrel{~}{\sigma }_m^{\alpha \beta }v_{\beta \dot{A}}^{})=\frac{1}{8}(v_A^\alpha \sigma _{m\alpha \beta }v_A^\beta ),$$
(27)
$$u_m^{(i)}=\frac{1}{8}(v_{\alpha A}^+\stackrel{~}{\sigma }_m^{\alpha \beta }v_{\beta \dot{A}}^{})\gamma _{A\dot{A}}^i=\frac{1}{8}(v_A^\alpha \sigma _{m\alpha \beta }v_{\dot{A}}^{\beta +})\gamma _{A\dot{A}}^i.$$
(28)
They are orthonormal
$$u^{[2]}u^{[\pm 2]}=0,u^{[+2]}u^{[2]}=2,u^{[\pm 2]}u^{(i)}=0,u^{(i)}u^{(j)}=\delta ^{(i)(j)}$$
(29)
as a result of certain harmonicity conditions imposed on the spinor harmonics that reduce the number of the independent variables in the spinor harmonics to the dimension of the $`SO(1,9)`$ Lorentz group equal to 45 . Lorentz frame vector harmonics satisfy the following differential equations
$$\begin{array}{c}_{\pm 2}u_m^{[+2]}=\mathrm{\Omega }_{\pm 2}^{(0)}u_m^{[+2]}+\mathrm{\Omega }_{\pm 2}^{[+2](i)}u_m^{(i)},\\ _{\pm 2}u_m^{[2]}=\mathrm{\Omega }_{\pm 2}^{(0)}u_m^{[2]}+\mathrm{\Omega }_{\pm 2}^{[2](i)}u_m^{(i)},\\ _{\pm 2}u_m^{(i)}=\frac{1}{2}\mathrm{\Omega }_{\pm 2}^{[+2](i)}u_m^{[2]}+\frac{1}{2}\mathrm{\Omega }_{\pm 2}^{[2](i)}u_m^{[+2]}+\mathrm{\Omega }_{\pm 2}^{(i)(j)}u_m^{(j)}.\end{array}$$
(30)
These are the only possible equations compatible with the orthonormality conditions (29). Coefficients $`\mathrm{\Omega }`$ in (30) are the $`SO(1,1)\times SO(8)`$ decomposed $`SO(1,9)`$ Cartan forms. From the embedding theory point of view ,,,,, $`\mathrm{\Omega }_{\pm 2}^{(i)(j)}`$ can be identified with the torsion (third fundamental form) components, $`\mathrm{\Omega }_{\pm 2}^{[+2](i)}`$ and $`\mathrm{\Omega }_2^{[2](i)}`$ with the second fundamental form components and $`\mathrm{\Omega }_{\pm 2}^{(0)}`$ with the 2d spin connection. Integrability conditions of Eqs.(30) are Gauss, Peterson-Kodacci and Ricci equations ,.
Action (25) is invariant under the following $`\kappa `$symmetry gauge transformations with the local parameters $`\kappa _A^+`$ and $`\kappa _{\dot{A}}^{}`$:
$$\begin{array}{c}\delta \theta ^{\alpha 1}=v_A^\alpha \kappa _A^+,\delta \theta ^{\alpha 2}=v_{\dot{A}}^{\alpha +}\kappa _{\dot{A}}^{},\\ \delta X^m=i\left(\kappa _A^+v_A^\alpha \sigma _{\alpha \beta }^m\theta ^{\beta 1}+\kappa _{\dot{A}}^{}v_{\dot{A}}^{\alpha +}\sigma _{\alpha \beta }^m\theta ^{\beta 2}\right),\\ \delta \left(ee^{\mu [+2]}\right)=\frac{4i}{c(\alpha )^{1/2}}\kappa _A^+\epsilon ^{\mu \nu }_\nu \theta ^{\alpha 1}v_{\alpha A}^+,\delta \left(ee^{\mu [2]}\right)=\frac{4i}{c(\alpha )^{1/2}}\kappa _{\dot{A}}^{}\epsilon ^{\mu \nu }_\nu \theta ^{\alpha 2}v_{\alpha \dot{A}}^{},\\ \delta u_m^{[+2]}=\frac{2i}{c(\alpha )^{1/2}}e^{\mu [+2]}w_\mu ^{(i)}u_m^{(i)},\delta u_m^{[2]}=\frac{2i}{c(\alpha )^{1/2}}e^{\mu [2]}w_\mu ^{(i)}u_m^{(i)},\end{array}$$
(31)
where $`w_\mu ^{(i)}=\left(\kappa _B^+\gamma _{B\dot{B}}^i_\mu \theta ^{\alpha 1}v_{\alpha \dot{B}}^{}\kappa _{\dot{B}}^{}\stackrel{~}{\gamma }_{\dot{B}B}^i_\mu \theta ^{\alpha 2}v_{\alpha B}^+\right)`$ <sup>6</sup><sup>6</sup>6For the Type IIA case $`\kappa `$symmetry transformations read:
$$\begin{array}{c}\delta \theta ^{\alpha 1}=v_A^\alpha \kappa _A^+,\delta \theta _\alpha ^2=v_{\alpha A}^+\kappa _A^{},\\ \delta X^m=i\left(\kappa _A^+v_A^\alpha \sigma _{\alpha \beta }^m\theta ^{\beta 1}+\kappa _A^{}v_{\alpha A}^+\stackrel{~}{\sigma }^{m\alpha \beta }\theta _\beta ^2\right),\\ \delta \left(ee^{\mu [+2]}\right)=\frac{4i}{c(\alpha ^{})^{1/2}}\kappa _A^+\epsilon ^{\mu \nu }_\nu \theta ^{\alpha 1}v_{\alpha A}^+,\delta \left(ee^{\mu [2]}\right)=\frac{4i}{c(\alpha ^{})^{1/2}}\kappa _A^{}\epsilon ^{\mu \nu }_\nu \theta _\alpha ^2v_A^\alpha ,\\ \delta u_m^{[+2]}=\frac{2i}{c(\alpha ^{})^{1/2}}e^{\mu [+2]}w_\mu ^{(i)}u_m^{(i)},\delta u_m^{[2]}=\frac{2i}{c(\alpha ^{})^{1/2}}e^{\mu [2]}w_\mu ^{(i)}u_m^{(i)},\end{array}$$
(32) where $`w_\mu ^{(i)}=\left(\kappa _B^+\gamma _{B\dot{B}}^i_\mu \theta ^{\alpha 1}v_{\alpha \dot{B}}^{}+\kappa _B^{}\gamma _{B\dot{B}}^i_\mu \theta _\alpha ^2v_{\dot{B}}^{\alpha +}\right)`$..
To fix $`\kappa `$symmetry gauge it is useful to expand Grassmannian variables using the spinor harmonics:
$$\theta ^{\alpha 1(2)}=v_A^\alpha \theta _A^{1(2)+}+v_{\dot{A}}^{\alpha +}\theta _{\dot{A}}^{1(2)}.$$
(33)
$`\kappa `$Symmetry transformations for the introduced variables look as follows:
$$\delta \theta _{\dot{A}}^1=\frac{i}{c(\alpha ^{})^{1/2}}\theta _A^{1+}\gamma _{A\dot{A}}^ie^{\mu [2]}w_\mu ^{(i)},\delta \theta _A^{1+}=\kappa _A^++\frac{i}{c(\alpha ^{})^{1/2}}\theta _{\dot{A}}^1\stackrel{~}{\gamma }_{\dot{A}A}^ie^{\mu [+2]}w_\mu ^{(i)},$$
(34)
$$\delta \theta _A^{2+}=\frac{i}{c(\alpha ^{})^{1/2}}\theta _{\dot{A}}^2\stackrel{~}{\gamma }_{\dot{A}A}^ie^{\mu [+2]}w_\mu ^{(i)},\delta \theta _{\dot{A}}^2=\kappa _{\dot{A}}^{}\frac{i}{c(\alpha ^{})^{1/2}}\theta _A^{2+}\gamma _{A\dot{A}}^ie^{\mu [2]}w_\mu ^{(i)}.$$
(35)
$`\theta _A^{1+}`$ and $`\theta _{\dot{A}}^2`$ are pure gauge variables, so we are able to impose the following $`\kappa `$symmetry fixing conditions:
$$\theta _A^{1+}=0,\theta _{\dot{A}}^2=0.$$
(36)
In this gauge the remaining variables $`\theta _{\dot{A}}^1\theta _{\dot{A}}^{}`$ and $`\theta _A^{2+}\theta _A^+`$ are $`\kappa `$invariant as is seen from (34),(35). Note, that they are the worldsheet MW spinors, whereas original variables $`\theta ^{\alpha 1,2}`$ were the worldsheet scalars. $`\kappa `$Symmetry fixed action (25) written in these new variables acquires the form
$$\begin{array}{c}S_{fixed}=e\left[(\alpha ^{})^{1/2}e_{[2]}^\mu (D_\mu x^{[2]}2i\stackrel{~}{D}_\mu \theta ^{}\theta ^{})(\alpha ^{})^{1/2}e_{[+2]}^\mu \left(D_\mu x^{[+2]}2iD_\mu \theta ^+\theta ^+\right)+c\right]\\ \frac{i}{c\alpha ^{}}ϵ^{\mu \nu }[(D_\mu x^{[+2]}2iD_\mu \theta ^+\theta ^+)\stackrel{~}{D}_\nu \theta ^{}\theta ^{}(D_\mu x^{[2]}2i\stackrel{~}{D}_\mu \theta ^{}\theta ^{})D_\nu \theta ^+\theta ^+2i\stackrel{~}{D}_\mu \theta ^{}\theta ^{}D_\nu \theta ^+\theta ^+\\ \frac{1}{2}(D_\mu x^{(i)}\frac{1}{4}(\theta ^{}\stackrel{~}{\gamma }^{ij}\theta ^{})\mathrm{\Omega }_\mu ^{[+2](j)}\frac{1}{4}(\theta ^+\gamma ^{ij}\theta ^+)\mathrm{\Omega }_\mu ^{[2](j)})((\theta ^{}\stackrel{~}{\gamma }^{ij}\theta ^{})\mathrm{\Omega }_\nu ^{[+2](j)}(\theta ^+\gamma ^{ij}\theta ^+)\mathrm{\Omega }_\nu ^{[2](j)})].\end{array}$$
(37)
We expanded bosonic coordinates $`x^m`$ on the vector harmonics
$$x^m=u^{m(n)}x_{(n)}\frac{1}{2}u^{m[+2]}x^{[2]}+\frac{1}{2}u^{m[2]}x^{[+2]}u_m^{(i)}x^{(i)}$$
(38)
in order to get rid of harmonics in the action and also introduced vector covariant derivatives $`D_\mu x^{(n)}=_\mu x^{(n)}+\mathrm{\Omega }_\mu ^{(n)(l)}x_{(l)},`$ whose decompositions read
$`D_\mu x^{[\pm 2]}=_\mu x^{[\pm 2]}\mathrm{\Omega }_\mu ^{(0)}x^{[\pm 2]}\mathrm{\Omega }_\mu ^{[\pm 2](i)}x^{(i)},`$ (39)
$`D_\mu x^{(i)}=_\mu x^{(i)}\mathrm{\Omega }_\mu ^{(i)(j)}x^j{\displaystyle \frac{1}{2}}\mathrm{\Omega }_\mu ^{[+2](i)}x^{[2]}{\displaystyle \frac{1}{2}}\mathrm{\Omega }_\mu ^{[2](i)}x^{[+2]},`$ (40)
and the spinor ones
$`D_\mu \theta _A^+=_\mu \theta _A^+{\displaystyle \frac{1}{2}}\mathrm{\Omega }_\mu ^{(0)}\theta _A^+{\displaystyle \frac{1}{4}}\mathrm{\Omega }_\mu ^{(i)(j)}\gamma _{AB}^{ij}\theta _B^+,`$ (41)
$`\stackrel{~}{D}_\mu \theta _{\dot{A}}^{}=_\mu \theta _{\dot{A}}^{}+{\displaystyle \frac{1}{2}}\mathrm{\Omega }_\mu ^{(0)}\theta _{\dot{A}}^{}{\displaystyle \frac{1}{4}}\mathrm{\Omega }_\mu ^{(i)(j)}\stackrel{~}{\gamma }_{\dot{A}\dot{B}}^{ij}\theta _{\dot{B}}^{}.`$ (42)
Action (37) contains the following light-cone-like terms quadratic in $`\theta ^\pm `$
$$S_{l.c.}=\frac{2i}{(\alpha ^{})^{1/2}}e[(1+\frac{2}{c(\alpha ^{})^{1/2}}D_{[+2]}x^{[+2]})\stackrel{~}{D}_{[2]}\theta ^{}\theta ^{}+(1+\frac{2}{c(\alpha ^{})^{1/2}}D_{[2]}x^{[2]})D_{[+2]}\theta ^+\theta ^+]$$
(43)
Note, that the form of the action (37) resembles that of Ref. for superparticles.
Let us turn to the problem of establishing a connection between the Lorentz harmonic variables and the commuting spinors $`\lambda _\pm ^\alpha `$ of the previous section. For this note that the variation of the action (25) with respect to the zweinbeins and harmonics produces the superstring embedding equation
$$\mathrm{\Pi }_\mu ^m=\frac{c(\alpha ^{})^{1/2}}{2}\left(e_\mu ^{[2]}u^{m[+2]}+e_\mu ^{[+2]}u^{m[2]}\right).$$
(44)
It can be simplified by choosing the conformal gauge for the zweinbein $`e_\mu ^f=e^\varphi \delta _\mu ^f`$:
$$\mathrm{\Pi }_{\pm 2}^m=\frac{c(\alpha )^{1/2}}{2}e^\varphi u^{m[2]}.$$
(45)
Equation (45) coincides with (12) only if
$$\lambda _\pm \sigma ^m\lambda _\pm =\frac{c(\alpha ^{})^{1/2}}{2}e^\varphi u^{m[2]},$$
(46)
thus establishing the bridge with the discussion of the preceding section. To analyse the consequences of (46) let us expand $`\lambda _\pm ^\alpha `$ on the spinor harmonics
$$\lambda _+^\alpha =v_A^\alpha \lambda _A+v_{\dot{A}}^{\alpha +}\lambda _{+2\dot{A}},\lambda _{}^\alpha =v_A^\alpha \lambda _{2A}+v_{\dot{A}}^{\alpha +}\lambda _{\dot{A}}.$$
(47)
Then one finds that $`\lambda _{+2\dot{A}}=\lambda _{2A}=0`$ and $`\lambda _A{}_{}{}^{2}=\frac{c(\alpha )^{1/2}}{2}e^\varphi ,\lambda _{\dot{A}}{}_{}{}^{2}=\frac{c(\alpha )^{1/2}}{2}e^\varphi `$. The integrability conditions of equations (45) after using (10) lead to the following expressions for the $`2d`$ spin connection
$$\mathrm{\Omega }_{+2}^{(0)}=_{+2}\varphi ,\mathrm{\Omega }_2^{(0)}=_2\varphi ,$$
(48)
and the minimality conditions for the components of the second fundamental form
$$\mathrm{\Omega }_2^{[2](i)}=\mathrm{\Omega }_{+2}^{[+2](i)}=0.$$
(49)
The connection between $`\kappa `$symmetry fixed GS variables and the NSR variables can be established upon the substitution of
$$\theta ^{\alpha 1}=v_{\dot{A}}^{\alpha +}\theta _{\dot{A}}^{},\theta ^{\alpha 2}=v_A^\alpha \theta _A^+,\lambda _+^\alpha =v_A^\alpha \lambda _A,\lambda _{}^\alpha =v_{\dot{A}}^{\alpha +}\lambda _{\dot{A}}$$
(50)
into (11) and the expansion of $`\psi _\pm ^m`$ on the vector harmonics
$$\psi _\pm ^m=u^{m(n)}\phi _{\pm (n)}=\frac{1}{2}u^{m[+2]}\phi _\pm ^{[2]}+\frac{1}{2}u^{m[2]}\phi _\pm ^{[+2]}u_m^{(i)}\phi _\pm ^{(i)}.$$
(51)
As a result we obtain
$$\phi _\pm ^{[+2]}=\phi _\pm ^{[2]}=0,\phi _+^i=\sqrt{8}\lambda _A\gamma _{A\dot{B}}^i\theta _{\dot{B}}^{},\phi _{}^i=\sqrt{8}\lambda _{\dot{A}}\stackrel{~}{\gamma }_{\dot{A}B}^i\theta _B^+.$$
(52)
The equations of motion for $`\kappa `$symmetry fixed Grassmannian variables $`\theta _{\dot{A}}^{}`$ and $`\theta _A^+`$ and the commuting spinors $`\lambda _A`$ and $`\lambda _{\dot{A}}`$ one obtaines after the substitution of (50) into (10,17,20) and taking into account the equations of motion for the spinor harmonics following from (30)
$$D_{+2}\theta _A^+=0,D_2\theta _A^+=e_2\theta _A^+,\stackrel{~}{D}_{+2}\theta _{\dot{A}}^{}=e_{+2}\theta _{\dot{A}}^{},\stackrel{~}{D}_2\theta _{\dot{A}}^{}=0.$$
(53)
Note, that these equations can be considered from the $`2d`$ field theory point of view as the Dirac equations for spinors interacting with the $`SO(1,1)\times SO(8)`$ Yang-Mills connection. There also appear the following equations for the nonzero components of the second fundamental form
$$\mathrm{\Omega }_{+2}^{[2](i)}\stackrel{~}{\gamma }_{\dot{B}A}^i\theta _A^+=0,\mathrm{\Omega }_2^{[+2](i)}\gamma _{A\dot{B}}^i\theta _{\dot{B}}^{}=0$$
(54)
and for the coefficients in (17)
$$d_+=d_{}=0.$$
(55)
As was noted before, coefficients $`d_+`$ and $`d_{}`$ correspond to the two unfixed $`\kappa `$symmetry trnasformations which, when establishing the relation with the NSR string, are identified with the $`n=(1|1)`$ worldsheet superreparametrizations. Their nullification signifies that we have entirely fixed $`\kappa `$symmetry.
Analogously equations for $`\lambda _A`$ and $`\lambda _{\dot{A}}`$ are
$$D_{+2}\lambda _A=f_{+2}\lambda _A,D_2\lambda _A=0,\stackrel{~}{D}_2\lambda _{\dot{A}}=f_2\lambda _{\dot{A}},D_{+2}\lambda _{\dot{A}}=0.$$
(56)
There also appear new equations for the components of the second fundamental form
$$\mathrm{\Omega }_{+2}^{[2](i)}\stackrel{~}{\gamma }_{\dot{B}A}^i\lambda _A=g_{+3}\theta _{\dot{B}}^{},\mathrm{\Omega }_2^{[+2](i)}\gamma _{A\dot{B}}^i\lambda _{\dot{B}}=g_3\theta _A^+.$$
(57)
To solve equations (54) we suggest that
$$\mathrm{\Omega }_{+2}^{[2](i)}=\stackrel{~}{\mathrm{\Omega }}_{+2}^{[10](i)}(\theta ^+)^8,\mathrm{\Omega }_2^{[+2](i)}=\stackrel{~}{\mathrm{\Omega }}_2^{[+10](i)}(\theta ^{})^8,$$
(58)
where $`(\theta ^{})^8\epsilon _{\dot{A}_1\mathrm{}\dot{A}_8}\theta _{\dot{A}_1}^{}\mathrm{}\theta _{\dot{A}_8}^{}`$ and $`(\theta ^+)^8\epsilon _{A_1\mathrm{}A_8}\theta _{A_1}^+\mathrm{}\theta _{A_8}^+`$. Further, the following representation
$$\stackrel{~}{\mathrm{\Omega }}_{+2}^{[10](i)}=(\theta ^{}\stackrel{~}{\gamma }^{ij}\theta ^{})q_{+10}^j,\stackrel{~}{\mathrm{\Omega }}_2^{[+10](i)}=(\theta ^+\gamma ^{ij}\theta ^+)q_{10}^j$$
(59)
allows to define $`g_{\pm 3}=8\phi _\pm ^iq_{\pm 10}^i(\theta ^\pm )^8`$ and to reformulate (57) as
$$(\theta ^{}\stackrel{~}{\gamma }^{ij}\theta ^{})q_{+10}^k\gamma _{\dot{B}A}^{ijk}\lambda _A=0,(\theta ^+\gamma ^{ij}\theta ^+)q_{10}^k\gamma _{B\dot{A}}^{ijk}\lambda _{\dot{A}}=0.$$
(60)
The system (60) has the rank equal to 4 so it admits nontrivial solutions.
The integrability conditions for equations (53,56) look like
$`_{+2}\mathrm{\Omega }_2^{(i)(j)}_2\mathrm{\Omega }_{+2}^{(i)(j)}+\mathrm{\Omega }_{+2}^{(k)(i)}\mathrm{\Omega }_2^{(k)(j)}\mathrm{\Omega }_2^{(k)(i)}\mathrm{\Omega }_{+2}^{(k)(j)}=0,`$ (61)
$`_{\pm 2}e_2=_{+2}_2\varphi `$ (62)
$`_{\pm 2}f_2=_{+2}_2\varphi .`$ (63)
Equation (61) coincides with Ricci equation after taking into account introduced representations for the second fundamental form components. Then Gauss equation reduces to $`_{+2}_2\varphi =0`$. Thus, $`e_{\pm 2}=e_{\pm 2}(\sigma ^{\pm 2})`$ is an arbitrary function and $`f_{\pm 2}=_{\pm 2}\varphi `$ as follows from the normalization conditions for $`\lambda _A`$, $`\lambda _{\dot{A}}`$.
The general solutions to equations (53,56) read:
$`\theta _A^+=\mathrm{exp}(\varphi /2+{\displaystyle e_2(\sigma ^2)𝑑\sigma ^2})\alpha _{AB}^1\theta _{0B}^+,\theta _{\dot{A}}^{}=\mathrm{exp}(\varphi /2+{\displaystyle e_{+2}(\sigma ^{+2})𝑑\sigma ^{+2}})\beta _{\dot{A}\dot{B}}^1\theta _{0\dot{B}}^{},`$ (64)
$`\lambda _A=\mathrm{exp}\left(\varphi /2\right)\alpha _{AB}^1\lambda _{0B},\lambda _{\dot{A}}=\mathrm{exp}\left(\varphi /2\right)\beta _{\dot{A}\dot{B}}^1\lambda _{0\dot{B}},`$ (65)
where the nondegenerate matrices $`\alpha _{AB}`$ and $`\beta _{\dot{A}\dot{B}}`$ satisfy the following equations
$$_{\pm 2}\alpha _{AB}=\frac{1}{4}\alpha _{AC}\mathrm{\Omega }_{\pm 2}^{(i)(j)}\gamma _{CB}^{(i)(j)},_{\pm 2}\beta _{\dot{A}\dot{B}}=\frac{1}{4}\beta _{\dot{A}\dot{C}}\mathrm{\Omega }_{\pm 2}^{(i)(j)}\stackrel{~}{\gamma }_{\dot{C}\dot{B}}^{(i)(j)}.$$
(66)
The integrability conditions for these equations coincide with (61).
Equations of motion for the NSR string physical variables $`\phi _\pm ^i`$ can be obtained either by differentiating (52) and substituting the equations of motion for $`\theta _A^+`$, $`\theta _{\dot{A}}^{}`$, $`\lambda _A`$, $`\lambda _{\dot{A}}`$ or by substitution of (52) into (14,21)
$`_{+2}\phi _+^{(i)}\mathrm{\Omega }_{+2}^{(i)(j)}\phi _+^{(j)}=\stackrel{~}{e}_{+2}\phi _+^{(i)},_2\phi _+^{(i)}\mathrm{\Omega }_2^{(i)(j)}\phi _+^{(i)}=0,`$ (67)
$`_2\phi _{}^{(i)}\mathrm{\Omega }_2^{(i)(j)}\phi _{}^{(j)}=\stackrel{~}{e}_2\phi _{}^{(i)},_{+2}\phi _{}^{(i)}\mathrm{\Omega }_{+2}^{(i)(j)}\phi _{}^{(i)}=0,`$ (68)
where $`\stackrel{~}{e}_{\pm 2}(\sigma ^{\pm 2})=e_{\pm 2}_{\pm 2}\varphi .`$ Their solutions read
$$\phi _+^i=\mathrm{exp}\left(\stackrel{~}{e}_{+2}(\sigma ^{+2})𝑑\sigma ^{+2}\right)A^{1ij}\phi _{0+}^j,\phi _{}^i=\mathrm{exp}\left(\stackrel{~}{e}_2(\sigma ^2)𝑑\sigma ^2\right)A^{1ij}\phi _0^j,$$
(69)
where invertible matrix $`A^{ij}`$ satisfies the following system
$$_2A^{(i)(j)}=A^{(i)(k)}\mathrm{\Omega }_2^{(k)(j)},_{+2}A^{(i)(j)}=A^{(i)(k)}\mathrm{\Omega }_{+2}^{(k)(j)}.$$
(70)
After substitution of $`\phi _\pm ^i`$ into the NSR string action its fermionic part acquires the form
$$S_{NSR}^{ferm}=\frac{i}{c\alpha ^{}}\left(\phi _+^i_2\phi _+^i\phi _+^i\mathrm{\Omega }_2^{ij}\phi _+^j\right)\frac{i}{c\alpha ^{}}\left(\phi _{}^i_{+2}\phi _{}^i\phi _{}^i\mathrm{\Omega }_{+2}^{ij}\phi _{}^j\right).$$
(71)
Taking into account the fact that (52) contains $`8d`$ $`\sigma `$matrices and performing corresponding Fierz rearrengements one obtains another form of the action (71)
$$\begin{array}{c}S_{NSR}^{fermmod}=\frac{32i}{\alpha _{}^{}{}_{}{}^{1/2}}e^\varphi \left(\theta _A^+D_{+2}\theta _A^++\theta _{\dot{A}}^{}\stackrel{~}{D}_2\theta _{\dot{A}}^{}\right)\\ \frac{i}{3!c\alpha ^{}}\underset{\pm }{}\left[\psi _\pm ^{ijk}\left(_2\psi _\pm ^{ijk}\mathrm{\Omega }_2^{im}\psi _\pm ^{mjk}\mathrm{\Omega }_2^{jm}\psi _\pm ^{imk}\mathrm{\Omega }_2^{km}\psi _\pm ^{ijm}\right)\right],\end{array}$$
(72)
where the first part is the covariatized version of the light-cone action<sup>10</sup><sup>10</sup>10 It coincides with (43), written in the conformal gauge, provided that the following equations are satisfied $`D_{+2}x^{[+2]}=D_2x^{[2]}=17c(\alpha ^{}){}_{}{}^{1/2}e_{}^{\varphi }`$. , the second part involves new variables, absent in the original formulations of both the NSR string and the GS superstring theories, namely, the Grassmannian 3-forms $`\psi _+^{ijk}=\sqrt{8}\lambda _A\gamma _{A\dot{A}}^{ijk}\theta _{\dot{A}}^{}`$ and $`\psi _{}^{ijk}=\sqrt{8}\lambda _{\dot{A}}\stackrel{~}{\gamma }_{\dot{A}A}^{ijk}\theta _A^+`$.
## 4 Conclusions
We have considered the $`n=(1|1)`$ superembedding equation for the Type II superstring. It was shown to contain the relation (11) between the NSR string and the Type II GS superstring variables, as well as the solution to the super-Virasoro constraints. Upon manifestly $`SO(1,9)`$ Lorentz covariant fixation of $`\kappa `$symmetry using the twistor-like Lorentz harmonic variables, which amounts to covariantizing the light-cone gauge, (11) reduces to the relation between $`\kappa `$symmetry fixed Type II GS superstring variables and the transverse physical NSR string variables $`\phi _\pm ^i`$. The equations of motion for the gauge fixed variables were obtained and solved. It was demonstrated that the gauge fixed Type II GS and NSR actions contain covariantized light-cone terms (43,72).
## 5 Acknowledgements
The author would like to thank A.A. Zheltukhin for the numerous valuable discussions, I.A. Bandos and D. Polyakov for the interest to the work and stimulating discussions, and the Abdus Salam ICTP, where the part of this work was done, for the warm hospitality. The work was supported by Ukrainian State Foundation for Fundamental Research.
|
warning/0006/cond-mat0006290.html
|
ar5iv
|
text
|
# Observability of quantum phase fluctuations in cuprate superconductors
## Abstract
We study the order parameter phase fluctuation effects in cuprate superconductors near $`T=0`$, using a quasi-two-dimensional $`d`$-wave BCS model. An effective phason theory is obtained which is used to estimate the strength of the fluctuations, the fluctuation correction to the in-plane penetration depth, and the pair-field susceptibility. We find that while the phase fluctuation effects are difficult to observe in the renormalization of the superfluid phase stiffness, they may be observed in a pair tunneling experiment which measures the pair-field susceptibility.
Underdoped cuprates exhibit large deviations from the predictions of BCS mean field theory, including a large gap $`\mathrm{\Delta }`$ which does not scale with the transition temperature $`T_c`$, and a “pseudogap” feature in the normal state. These facts, together with the empirical scaling of the superfluid phase stiffness (SPS) and $`T_c`$ with doping, have motivated a picture of the cuprates in which the discrepancies are attributed to strong phase fluctuations . The estimated phase fluctuation energy scale is smaller than the gap (pairing) energy scale in these materials, and within this picture $`T_c`$ is determined by the small SPS rather than the gap energy. The pseudogap is ascribed to a precursor pairing amplitude whose phase coherence is destroyed above $`T_c`$ .
As many other explanations for these effects have been put forward, it is of great interest to devise tests of the phase fluctuation scenario which distinguish it from others. Some evidence for thermal phase fluctuation effects was provided by Corson et al. , who observed unusual conductivity resonances in underdoped $`\mathrm{Bi}_2\mathrm{Sr}_2\mathrm{CaCu}_2\mathrm{O}_{8+\delta }`$ (Bi2212) near $`T_c`$ and analyzed their data using two-dimensional Kosterlitz-Thouless-Berezinskii dynamics in the terahertz range. It has been claimed that phase fluctuations are the dominant excitations even at low temperatures $`TT_c`$ , determining the well-known linear-$`T`$ dependence of the penetration depth. However, Millis et al. have shown that quantum phase fluctuations cannot account for this behavior. Quantum phase fluctuations do have important consequences for the superconductor-insulator transition, and the $`c`$-axis optical conductivity. The above probes provide indirect observation of the phase fluctuation effects through electronic observables which are modified by the phase fluctuations. In this paper, we search for more direct probes of phase fluctuations within a model of quasi-two-dimensional $`d`$-wave BCS superconductors with interlayer Josephson coupling near optimal doping.
The in-plane SPS can be expressed as $`D_{ab}=n_{s,ab}\mathrm{}^2/4md`$ where $`n_{s,ab}`$ is the planar superfluid electron density, $`m`$ is the effective mass of the quasiparticle, and $`d`$ is the interplanar spacing. Because $`D_{ab}`$ is determined by the quasiparticle properties which are not strongly renormalized by the phase fluctuations, we find that although the renormalization of the Debye-Waller factor $`e^{i\varphi }`$ is relatively strong, both the SPS at $`T=0`$ and its temperature corrections are weakly renormalized, in contrast to the case of a Josephson junction array (JJA) model. We also consider an experiment to measure the excess current in a tunnel junction, which can be directly related to the pair-field susceptibility $`\chi =i\theta (t)[\mathrm{\Delta }(𝐫,t),\mathrm{\Delta }^{}(\mathrm{𝟎},0)]`$. For the cuprates we find that this current is experimentally observable, due to the combination of large phase fluctuations and a low-lying $`c`$-axis plasmon mode. We predict a pronounced peak in the in the excess current at the $`c`$-axis Josephson plasma frequency $`\omega _c`$.
We begin with a continuum BCS model with $`d`$-wave pairing symmetry in an isolated two-dimensional layer at temperature $`T=1/\beta `$ in the superconducting state:
$`S_{2\mathrm{D}}={\displaystyle _0^\beta }d\tau \{{\displaystyle }d^2x{\displaystyle \underset{\sigma }{}}c_\sigma ^{}(_\tau {\displaystyle \frac{^2}{2m_0}}\mu )c_\sigma `$ (1)
$`+{\displaystyle }d^2Rd^2r[\mathrm{\Delta }(𝐑,𝐫,\tau )c_{}^{}(𝐑+𝐫/2,\tau )c_{}^{}(𝐑𝐫/2,\tau )`$ (2)
$`+\mathrm{h}.\mathrm{c}.+{\displaystyle \frac{1}{g}}|\mathrm{\Delta }(𝐑,𝐫,\tau )|^2]\},`$ (3)
with the interaction strength $`g>0`$. Here and throughout the paper we set $`\mathrm{}=c=k_B=1`$ for convenience except when numerical values are estimated. The gap has the property that $`d^2re^{i𝐩𝐫}\mathrm{\Delta }(𝐑,𝐫,\tau )=\mathrm{\Delta }(𝐑,\tau )(p_x^2p_y^2)/p^2`$. Then we factor the pairing field as $`\mathrm{\Delta }(𝐱,\tau )=|\mathrm{\Delta }(𝐱,\tau )|e^{i\varphi (𝐱,\tau )}`$; in what follows we will assume that the amplitude of the order parameter is constant, $`|\mathrm{\Delta }(𝐱,\tau )|=\mathrm{\Delta }`$, and focus on the phase degree of freedom, $`\varphi `$. In order to decouple the $`\varphi `$ field from the order parameter amplitude, we perform a singular gauge transformation $`\psi _\sigma (𝐱,\tau )=c_\sigma (𝐱,\tau )e^{i\varphi (𝐱,\tau )/2}`$, with $`\psi _\sigma `$ the field operators for the transformed quasiparticle. The phase-quasiparticle coupling terms are then
$$S_I=d^2x_0^\beta 𝑑\tau \left[\frac{\varphi }{2}\widehat{\psi }^{}\frac{i}{m}\widehat{\psi }+\frac{(\varphi )^2}{8m}\widehat{\psi }^{}\widehat{\tau }_3\widehat{\psi }\right]$$
(4)
where $`\widehat{\psi }`$ is the Nambu spinor. The wavevector of the phase fluctuations has an upper cutoff $`\mathrm{\Lambda }_c`$ of order $`\xi _0^1\mathrm{\Delta }/v_F`$ since beyond this momentum scale the mean-field assumption breaks down. Here we consider only the effect of longitudinal phase fluctuations since the production of vortex pairs is energetically unfavorable near $`T=0`$ and far away from the insulator transition. In order to study a realistic model, we consider such layers of two-dimensional superconductors with an interlayer distance $`d`$ and a weak Josephson tunneling ($`J`$) between adjacent layers, and a three-dimensional Coulomb interaction $`V(𝐪)=4\pi e^2/ϵ_bq^2`$, where $`𝐪=(𝐪_{},q_{})`$, with $`𝐪_{}`$ and $`q_{}`$ the in-plane and $`c`$-axis components of $`𝐪`$, and $`ϵ_b`$ the background dielectric constant. After integrating out the fermions we can obtain the effective phase-only action; the Gaussian term is
$`S^{(2)}[\varphi ]`$ $`=`$ $`T{\displaystyle \underset{\omega _n,𝐪}{}}{\displaystyle \frac{\omega _n^2+\omega _p^2(𝐪)}{8V(𝐪)}}\varphi (𝐪,\omega _n)\varphi (𝐪,\omega _n)`$ (5)
with $`\omega _n=2\pi nT`$ the bosonic Matsubara frequencies. The plasma frequency $`\omega _p`$ is defined through $`\omega _p^2(𝐪)=(\omega _{ab}^2q_{}^2+\omega _c^2q_{}^2)/(q_{}^2+q_{}^2)`$ where $`\omega _{ab}=\sqrt{4\pi n_{ab}e^2/ϵ_bmd}`$, $`n_{ab}`$ is the planar charge-carrier density at the plasma resonance frequency, and $`\omega _c=\sqrt{4\pi Jd^2e^2/ϵ_b}`$ is the $`c`$-axis Josephson plasma energy. Equation (5) gives the correct plasma spectrum for a layered superconductor, with $`\omega _{ab}`$ the planar plasma frequency at $`T=0`$ which is 0.44–1.4 eV, and $`\omega _c`$ is the $`c`$-axis plasma frequency which is about 0.6 meV in Bi2212 (larger in other materials), with an interlayer distance $`d1.5\mathrm{nm}`$ and a dielectric constant of $`ϵ_b10`$. The in-plane SPS $`D_{ab}=n_{s,ab}/4md`$ can be read off from Eq. (5) as the coefficient of $`(_{ab}\varphi )^2/2`$ in the $`\omega 0`$ and $`|𝐪|0`$ limit. The quasiparticle damping term, which will broaden the plasma mode in Eq. (5), has been omitted, but the effect is known to be small even in the case of the $`d`$-wave quasiparticle spectrum , and it will not affect the order of magnitude estimate of the correction to the SPS in what follows.
We first estimate the strength of the phase fluctuation from Eq. (5). One measure of the strength is the Debye-Waller factor $`\alpha =e^{\varphi ^2}`$, which for instance can be determined from the $`c`$-axis optical conductivity . For our model, at $`T=0`$ we have
$$\varphi ^2=\underset{𝐪}{}\frac{2V(𝐪)}{\mathrm{}\omega _p(𝐪)}\frac{2\mathrm{\Lambda }_ce^2/ϵ_b}{\mathrm{}\omega _{ab}}.$$
(6)
From this expression we see that the size of the quantum phase fluctuations is determined by the ratio of the Coulomb energy of a Cooper pair to the plasma energy; in the cuprates the short coherence lengths and small superfluid densities conspire to enhance these fluctuations. For instance, assuming $`\xi _020\mathrm{\AA }`$ and with the parameters given above, we estimate from Eq. (6) that $`\varphi ^2`$ ranges from $`0.1`$ to $`1`$, which is a sizable number compared to, for instance, $`10^3`$ in Pb. In this paper we study the effect of these strong phase fluctuations in the BCS model given in Eq. (3), which is more appropriate near optimal doping, far away from the insulator transition.
Since we are interested in the renormalization of the SPS, we need to go beyond the quadratic expansion in Eq. (5). For instance, in the JJA model, the effective SPS can be substantially renormalized due to the non-trivial potential of the form $`\mathrm{cos}(\varphi _i\varphi _j)`$ . In our model, higher-order terms can be determined by expanding Eq. (4) and integrating out the fermions with a $`d`$-wave gap. Each $`n`$-point vertex of the $`\varphi `$-field is a fermion loop (see Fig. 1). Therefore, the renormalization of the SPS and its temperature dependence is determined by the magnitudes of the fermion loops. From the new effective theory of the phase fields thus obtained,
$`S_{\mathrm{eff}}[\varphi ]`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle d^4x_1\mathrm{}d^4x_{2n}\mathrm{\Gamma }^{(2n)}(x_1,x_2,\mathrm{},x_{2n})}`$ (8)
$`\times _1\varphi (x_1)_2\varphi (x_2)\mathrm{}_{2n}\varphi (x_{2n}),`$
where $`\mathrm{\Gamma }^{(2n)}`$ are the 2$`n`$th-order phason vertices, we can estimate the renormalization of the in-plane SPS by using the one loop expansion in $`\varphi `$ as in Fig. 2.
At $`T=0`$, we can show that the diagrams in Fig. 2 (a), (b) and (c) cancel one another in the limit of zero external momentum and frequency by using the identity $`i_\omega \widehat{𝒢}(𝐩,\omega )=\widehat{𝒢}^2(𝐩,\omega )`$ and performing the integration by parts in $`\omega `$. Consequently, the only contribution comes from (d). The correction to the SPS is found to be
$`{\displaystyle \frac{\delta D_{ab}}{D_{ab}}}{\displaystyle \frac{e^2\mathrm{\Lambda }_c^3}{6n_{s,ab}ϵ_b\mathrm{}\omega _{ab}}}.`$ (9)
Assuming that the in-plane penetration depth is $`\lambda _{ab}=\sqrt{mc^2d/4\pi n_{s,ab}e^2}2000\mathrm{\AA }`$, we estimate that $`\delta D_{ab}/D_{ab}10^1`$. This should be compared to a 40% reduction obtained using the JJA model . In our model, the phason vertices are determined by the $`d`$-wave quasiparticle fermion loops, which are smaller than the vertices of a JJA model; consequently the renormalization of the SPS is smaller.
Next we study the temperature dependence of the SPS. The BCS theory gives a linear temperature dependence due to thermal excitations of quasiparticles, $`D_{ab}(T)D_{ab}(0)aT`$ where $`a=v_F\mathrm{ln}(2)/4\pi v_\mathrm{\Delta }d`$. Here $`v_\mathrm{\Delta }`$ is the slope of the gap at the node in momentum space. We find that the one-loop correction to $`a`$ is $`\delta a/a\frac{e^2v_F\mathrm{\Lambda }_c^2}{\pi ϵ_b\mathrm{}\omega _{ab}^2}10^2`$. The increase in the slope $`a`$ due to phase fluctuations is therefore hardly a measurable quantity, in agreement with the JJA result . Additional temperature dependence can be obtained from considering classical phase fluctuations or by coupling the phason to a heat bath. However, classical phase fluctuation effects are not relevant to our model and the coupling to the heat bath leads to only a sub-leading $`T^2`$ correction to the SPS.
Unlike quasiparticle properties represented by the SPS as discussed above, the phase fluctuation effects on Cooper pair properties can be significant. Here we propose an experiment which can measure the strength and the form of the phase fluctuations via the pair-field susceptibility. We consider a $`c`$-axis tunnel junction between two cuprate superconductors as illustrated in Fig. 3. The Josephson coupling between the phases (denoted by $`\varphi `$ and $`\varphi ^{}`$) of the two superconductors will lead to the usual Josephson current oscillating at a frequency of $`2eV/\mathrm{}`$ if there is a potential difference $`V`$ across the junction. There will also be a quasiparticle tunneling current. In addition, an excess current will flow due to the Josephson coupling of the superconducting pair-field of one superconducting electrode to the fluctuating pair-field of the other. To isolate the excess current a small magnetic field is applied parallel to the junction to suppress the Josephson current, and the quasiparticle tunneling current must be modeled and subtracted . The excess current is interesting because it can be related to the pair-field susceptibility at a frequency $`2eV/\mathrm{}`$ , and can thus provide information about the spectrum of phase fluctuations.
To specialize the experiment to our case, we suppose that both of the electrodes are identical with a gap $`\mathrm{\Delta }`$ and ignore the fluctuations in the amplitude of the order parameter assuming that $`2eV\mathrm{\Delta }`$. For simplicity, we consider a junction in the $`ab`$ plane (at $`z=0`$) of dimensions $`L_x\times L_y`$, with a magnetic field $`H_y`$ in the $`b`$-direction, and we will work at zero temperature. If we assume that the thickness of the electrodes is larger than $`\lambda _c`$, the $`c`$-axis penetration depth, the Josephson coupling Hamiltonian for a phase difference $`\delta \varphi (𝐫,t)\varphi (𝐫,t)\varphi ^{}(𝐫,t)`$ is
$`H_J`$ $`=`$ $`{\displaystyle \frac{E_J}{2S}}e^{i\omega t}{\displaystyle d^3re^{iq_xx}e^{i\delta \varphi (𝐫,t)}\delta (z)}+\mathrm{h}.\mathrm{c}.,`$ (10)
where $`E_J`$ is the Josephson coupling energy, $`S=L_xL_y`$ is the junction contact area, $`q_x4eH_y\lambda _c/\mathrm{}c`$ , and $`\omega =2eV/\mathrm{}`$. The current through the junction is
$$I=\frac{2eE_J}{\mathrm{}S}\mathrm{Im}e^{i\omega t}d^3re^{iq_xx}e^{i\delta \varphi (𝐫,t)}\delta (z).$$
(11)
If we calculate the current to zeroth order in $`H_J`$, and carry out the averages with respect to the Gaussian action in Eq. (5), we obtain the Josephson current with a critical current $`\stackrel{~}{I}_c=(2eE_J/\mathrm{})\alpha `$ which is renormalized by phase fluctuations through the Debye-Waller factor $`\alpha `$ (there is no quasiparticle current in our model). By calculating the current to first order in $`H_J`$ (linear response), we obtain the excess current,
$$I_{\mathrm{ex}}(\omega ,q_x)=\frac{eE_J^2}{S\mathrm{}^2}\mathrm{Im}D^R(q_x,q_y=0,\omega ;z=0),$$
(12)
where the retarded pair field susceptibility is
$$D^R(𝐫,t)=i\theta (t)[e^{i\delta \varphi (𝐫,t)},e^{i\delta \varphi (\mathrm{𝟎},0)}].$$
(13)
For $`\omega >0`$ we have $`\mathrm{Im}D^R(𝐪,\omega )=\mathrm{Im}D(𝐪,\omega )`$, where $`D(𝐪,\omega )`$ is the Fourier transform of the time-ordered correlation function, which for a Gaussian action is
$`D(𝐫,t)`$ $`=`$ $`i\alpha ^2e^{2T[\varphi (𝐫,t)\varphi (\mathrm{𝟎},0)]}`$ (14)
$``$ $`i\alpha ^2\left\{1+2T[\varphi (𝐫,t)\varphi (\mathrm{𝟎},0)]\right\},`$ (15)
where the factor of two comes from the two sides of the junction and $`T`$ is the time-ordering operator. Equation (15) assumes an expansion in the small parameter $`\mathrm{ln}\alpha `$. If we neglect the boundary effects near the junction, we can obtain the propagator for the phase fields from the action in Eq. (5); after an analytic continuation, we have
$$\varphi (𝐪,\omega )\varphi (𝐪,\omega )=\frac{4iV(𝐪)}{\omega ^2\omega _p^2(𝐪)+i0^+}.$$
(16)
For an isotropic plasma frequency $`\omega _p`$ it can be shown that the excess current consists of a series of $`\delta `$-functions at integer multiples of $`\omega _p`$. This result is somewhat academic since in the known isotropic superconductors the plasma energy is much larger than the gap energy, and we would expect amplitude fluctuations and quasiparticle damping to completely obliterate this effect. For an anisotropic plasma frequency these resonances become broadened—since the junction is localized at $`z=0`$, we must integrate over $`q_{}`$, which results in a plasma frequency that ranges from $`\omega _c`$ (when $`q_x/q_{}0`$) to $`\omega _{ab}`$ (when $`q_{}/q_x0`$). The result of the calculation is
$`I_{\mathrm{ex}}(\omega ,q_x)`$ $``$ $`\left({\displaystyle \frac{2\pi e\stackrel{~}{I}_c^2}{q_xϵ_bL_yL_x\mathrm{}}}\right){\displaystyle \frac{\theta (\omega \omega _c)\theta (\omega _{ab}\omega )}{\sqrt{(\omega ^2\omega _c^2)(\omega _{ab}^2\omega ^2)}}}.`$ (17)
To quench the background Josephson current, a field may be chosen such that $`q_x=2\pi /L_x`$; a result is shown in Fig. 3.
The excess current exhibits a sharp onset at $`\omega =\omega _c`$ with a peak of the form $`(\omega ^2\omega _c^2)^{1/2}`$. This peak would be rounded by the small quasiparticle damping which we have not considered here. This experiment would serve as a direct observation of the phase fluctuations and as an alternative way to measure the $`c`$-axis Josephson plasma energy lying below the maximum gap. In addition, it would provide a measure of the strength of the phase fluctuations. The gapless collective modes studied in Refs. are due to the order parameter fluctuations which are rendered visible near the transition temperature, which are in principle observable in any superconductors, whereas the quantum phason modes that we studied are observable as a result of the intrinsically strong phase fluctuations and the quasi-two-dimensionality of cuprate materials. Therefore, the result shown in Fig. 3 is a special zero-temperature property of cuprate superconductors.
We have shown that in the case of quasi-two-dimensional $`d`$-wave BCS model that we used here, the resulting correction to the absolute value and the temperature-dependence of the in-plane SPS is minute despite the strong phase fluctuations. Since quasiparticle interaction effects can obscure the fluctuation corrections to the SPS, it may be difficult to observe the effects of phase fluctuations on the penetration depth. As a more direct measurement, we have proposed a pair tunneling experiment which can probe the strength and spectrum of the quantum phase fluctuations. We expect that the pair-field susceptibility will show a pronounced peak at $`\omega =\omega _c`$. It will be also interesting to explore the role of the order parameter fluctuations at the superconductor-insulator transition in the underdoped regime via the suggested experiment. In the underdoped regime, the simple BCS model fails, especially at the superconductor-insulator transition which is one extreme example of the renormalization of the SPS, and the physics of doped Mott insulators needs to be taken into account.
We gratefully acknowledge discussions with A. J. Millis and E. H. Hwang. This work was supported by the NHMFL, NSF DMR-9978547 (ATD), NSF DMR-9974396 (PJH), NSF DMR-9815094, and the Packard Foundation (HJK).
|
warning/0006/math0006165.html
|
ar5iv
|
text
|
# From slightly coloured noises to unitless product systems
## Introduction
The white noise is a Gaussian stationary generalized random process whose restrictions to adjacent intervals $`(a,b)`$ and $`(b,c)`$ are independent. In contrast, for a continuous process, its restrictions to $`(a,b)`$ and $`(b,c)`$ are heavily dependent via the value at $`b`$. Such a dependence cannot be described by a probability density; the joint distribution is singular w.r.t. the product of marginal distributions. By a slightly coloured noise I mean a Gaussian stationary generalized random process such that the distribution of its restriction to $`(a,c)`$ is absolutely continuous w.r.t. the product of the distributions of its restrictions to $`(a,b)`$ and $`(b,c)`$, whenever $`\mathrm{}<a<b<c<+\mathrm{}`$.
In the Hilbert space of all linear functionals of the white noise, every interval $`(a,b)`$ determines a subspace $`G_{a,b}`$ satisfying $`G_{a,b}G_{b,c}=G_{a,c}`$; the whole space is a direct integral (a continuous direct sum). For arbitrary (not just linear) functionals the relation is multiplicative rather than additive:
$$H_{a,b}H_{b,c}=H_{a,c};$$
the whole space is a continuous tensor product, in other words, a product system. The constant $`1`$ may be treated as an element $`\mathrm{𝟏}_{a,b}H_{a,b}`$, satisfying
$$\mathrm{𝟏}_{a,b}\mathrm{𝟏}_{b,c}=\mathrm{𝟏}_{a,c};$$
such a multiplicative family is called a unit (of a product system). There exist product systems with many units, with a single unit (up to a natural equivalence), and unitless (with no unit). However, the theory of unitless product systems suffers from lack of rich sources of examples. Slightly coloured noises are such a source, rich enough for producing a continuum of nonisomorphic unitless product systems.
> “We believe that there should be a natural way of constructing such product systems, and we offer that as a basic unsolved problem. The fact that we do not yet know how to solve it shows how poorly understood continuous tensor products are today.”
> Arveson 1994 \[3, p. 5\].
After producing a continuum of nonisomorphic product systems with units I was asked by Arveson (private communication, January 2000) about a continuum of nonisomorphic unitless product systems. His question is answered here by using a construction that was outlined by Tsirelson and Vershik \[13, Sect. 1c\] with no proofs.
> “This example may be considered as a commutative (bosonic) counterpart of Power’s noncommutative (fermionic) example of a non-Fock factorization over $``$.”
> Tsirelson and Vershik 1998 \[13, p. 91\].
Probabilistic aspects
Probabilistic results are summarized here in probabilistic language (while the rest of the paper is written in rather analytical language).
Consider a Gaussian stationary generalized random process $`(\xi _t)_t`$ whose covariation function $`B(t)=\mathrm{Cov}(\xi _s,\xi _{s+t})`$ is positive, decreasing and convex on $`(0,\mathrm{})`$, and
$$B(t)=\frac{1}{|t|\mathrm{ln}^\alpha (1/|t|)}\text{for all }t\text{ small enough};$$
here $`\alpha (1,\mathrm{})`$ is a parameter.
Then the joint distribution of random variables
$$X_k=_0^{2\pi }e^{ikt}\xi _t𝑑t(k=\mathrm{},2,1,0,1,2,\mathrm{})$$
has a density (finite and strictly positive almost everywhere) w.r.t. the product of corresponding one-dimensional distributions $`N(𝔼X_k,\mathrm{Var}X_k)`$.
The same holds for a larger family $`(\mathrm{},X_1,Y_1,X_0,Y_0,X_1,Y_1,X_2,Y_2,\mathrm{})`$ of random variables, where $`X_k`$ are as before, and $`Y_k=_{2\pi }^0e^{ikt}\xi _t𝑑t`$.
Another result. Take some $`\epsilon _n(0,1)`$ and consider random variables
$$Z_n=\frac{1}{\epsilon _n}\underset{k=0}{\overset{n1}{}}_{k/n}^{(k+\epsilon _n)/n}\xi _t𝑑t,Z=_0^1\xi _t𝑑t.$$
If $`\epsilon _n\mathrm{ln}^{\alpha 1}n\mathrm{}`$ then $`Z_nZ`$ in $`L_2`$.
If $`\epsilon _n\mathrm{ln}^{\alpha 1}n0`$ then $`\mathrm{Var}(Z_n)\mathrm{}`$ and the correlation coefficient $`\mathrm{Corr}(Z_n,Z)0`$.
If $`lim_n(\epsilon _n\mathrm{ln}^{\alpha 1}n)(0,\mathrm{})`$ then $`lim_n\mathrm{Var}(Z_n)(0,\mathrm{})`$ and $`lim_n\mathrm{Corr}(Z_n,Z)(0,1)`$.
For detail see Sect. 10. The reader interested just in these probabilistic statements may skip sections 1–9 in which case, however, he/she should bypass some points of analytical nature in Sect. 10, and restore some proofs omitted since they are not necessary from the analytical viewpoint.
Quantal aspects
Acquaintance with quantum theory is not needed for reading the rest of the paper, but should help to understand the idea as explained here.
A product system may be thought of as a local quantum field over the one-dimensional space $``$ (just space, no time at all; see also Arveson for a better, dynamical interpretation of product systems). The whole field is a quantum system, and its restriction to $`(0,1)`$ is a subsystem. A unit vector of $`H_{0,1}`$ describes a pure state of the subsystem (though in general the subsystem and the rest of the system are entangled).
Local Hilbert spaces (and algebras) are ascribed to intervals, as well as to more general regions, consisting of a finite number of intervals. Introduce a region $`E_n`$ consisting of $`n`$ small equidistant intervals of equal length,
$$E_n=(0,\frac{\epsilon _n}{n})(\frac{1}{n},\frac{1+\epsilon _n}{n})\mathrm{}(\frac{n1}{n},\frac{n1+\epsilon _n}{n})$$
and consider the corresponding subsystem, described by its Hilbert space $`H_{E_n}`$. For a fixed $`n`$, the subsystem may be entangled or not (with the rest). If the product system has a unit then there is a ‘white state’ with no spatial correlations. It makes $`H_{E_n}`$ disentangled for all $`n`$ simultaneously. That is the case for a product system constructed out of the white noise.
A unitless product system is constructed out of the slightly coloured noise mentioned in ‘Probabilistic aspects’. Spatial correlations inherent to the noise are inherited by quantum states. Subsystems $`H_{E_n}`$ can be disentangled for finitely many $`n`$, but not for all $`n`$ simultaneously.
Asymptotic behavior of subsystems $`H_{E_n}`$ depends crucially on $`lim(\epsilon _n\mathrm{ln}^{\alpha 1}n)`$, as is suggested by properties of random variables $`Z_n`$ (see ‘Probabilistic aspects’). If $`lim(\epsilon _n\mathrm{ln}^{\alpha 1}n)=0`$ then $`E_n`$ are a tail sequence in the sense that the mixed state of the subsystem (the density matrix on $`H_{E_n}`$) has a universal asymptotics, irrespective of the state of the system. The situation is different if $`lim(\epsilon _n\mathrm{ln}^{\alpha 1}n)0`$. This is why product systems for different $`\alpha `$ are nonisomorphic.
Aspects of functional analysis
A classical construction going back to Fock may be outlined as follows. One starts with a Hilbert space $`G`$. One identifies $`G`$ with the space of all measurable linear functionals over a Gaussian measure $`\gamma `$. One gets another Hilbert space $`H=L_2(\gamma )`$ that may be denoted $`H=\mathrm{Exp}G`$, since $`G=G_1G_2`$ implies $`H=H_1H_2`$ (via $`\gamma =\gamma _1\gamma _2`$). A continuous direct sum, $`G=^{}G_\zeta 𝑑\zeta `$ (basically the same as a projection-valued measure well-known in spectral theory) leads to a continuous tensor product. Classical product systems, obtained this way, contain units.
My modification of the classical construction is rather innocent (but surprisingly powerful). Basically, the subspaces $`G_1,G_2`$ are allowed to be slightly nonorthogonal, without destroying the relation $`H=H_1H_2`$. In terms of Gaussian measures, the relation $`\gamma =\gamma _1\gamma _2`$ is generalized to $`\gamma =p(\gamma _1\gamma _2)`$ where $`p`$ is a density (that is, Radon-Nikodym derivative). The density $`p`$ is inserted properly into the formula for $`h_1h_2H`$.
The classical language (of Hilbert spaces, Gaussian measures, spaces $`L_2(\gamma )`$ etc.) is not well-suited to the modified construction. A modified language, used in the paper, stipulates larger invariance (symmetry) groups. Namely, the ‘Hilbert space’ structure corresponds to the group of all unitary operators, $`U^{}U=I`$. I use a larger group consisting of all invertible $`U`$ such that $`U^{}UI`$ is a Hilbert-Schmidt operator. (These $`U`$ are called ‘equivalence operators’ in Feldman’s well-known paper on equivalence of Gaussian measures). The corresponding structure, weaker than ‘Hilbert space’ structure but stronger than ‘linear topological space’ structure, is defined in Sect. 2 under the name ‘FHS-space’. An equivalence class (in Feldman’s sense) of norms is used rather than a single norm.
Accordingly, a measure space $`(\mathrm{\Omega },,P)`$ is replaced with a ‘measure type space’ $`(\mathrm{\Omega },,𝒫)`$. An equivalence class $`𝒫`$ of measures is used rather than a single measure $`P`$. See Sect. 1. It appears that $`L_2(\mathrm{\Omega },,𝒫)`$ can be defined naturally (in addition to the usual $`L_2(\mathrm{\Omega },,P)`$); see Sect. 1.
Also, a ‘Gaussian type space’ defined in Sect. 2 stipulates an equivalence class of Gaussian measures rather than a single Gaussian measure.
For a technical reason we need Borel measurability of several natural constructions. For example, the orthogonal projection of a vector to a subspace (in a Hilbert space) is a jointly Borel measurable function of the point and the subspace, provided that the set of all subspaces is equipped with its natural Borel structure. Similarly, the conditional expectation is a jointly Borel measurable function of a random variable and a sub-$`\sigma `$-field. See Sect. 6 for detail.
## 1 Measure type spaces and square roots of measures
Let $`\mathrm{\Omega }`$ be a nonempty set, $``$ a $`\sigma `$-field of its subsets, and $`\mu ,\nu `$ measures<sup>2</sup><sup>2</sup>2I assume always, that every measure space $`(\mathrm{\Omega },,\mu )`$ is a Lebesgue-Rokhlin space (that is, isomorphic $`mod0`$ to an interval with Lebesgue measure, to a finite or countable set of atoms, or a combination of both). However, the assumption is not really used in this work. We deal with such objects as $`L_2(\mathrm{\Omega },,\mu )`$; the latter must be separable; other properties of $`(\mathrm{\Omega },,\mu )`$ do not matter. I assume also that all considered $`\sigma `$-fields ($``$ itself, and its sub-$`\sigma `$-fields) contain all negligible sets. Everything is treated $`mod0`$, that is, up to negligible sets. (real-valued, finite, positive) on $``$. One says that $`\mu ,\nu `$ are *equivalent,* if they are mutually absolutely continuous. Let $`𝒫`$ be an equivalence class of probability measures on $`(\mathrm{\Omega },)`$. That is, every $`P𝒫`$ is a measure on $`(\mathrm{\Omega },)`$ satisfying $`P(\mathrm{\Omega })=1`$, and every $`P_1,P_2𝒫`$ are equivalent, and $`𝒫`$ contains every probability measure equivalent to a measure of $`𝒫`$ (and of course, $`𝒫`$ is nonempty). One says that $`𝒫`$ is a type of measure, and $`(\mathrm{\Omega },,𝒫)`$ is a *measure type space.*
The linear topological (metrizable, but not locally convex) space $`L_0(\mathrm{\Omega },,𝒫)`$ consists of all (equivalence classes of) measurable functions on $`\mathrm{\Omega }`$. Its topology corresponds to convergence in measure (in probability), irrespective of the choice of a measure $`P𝒫`$.
Hilbert spaces $`L_2(\mathrm{\Omega },,P_1)`$ and $`L_2(\mathrm{\Omega },,P_2)`$ for $`P_1,P_2𝒫`$ differ (unless $`\epsilon \epsilon P_1P_2(1/\epsilon )P_1`$). However, they are in a *natural* unitary correspondence. Namely, $`\psi _1L_2(\mathrm{\Omega },,P_1)`$ corresponds to $`\psi _2L_2(\mathrm{\Omega },,P_2)`$ when $`\psi _2=\sqrt{\frac{P_1}{P_2}}\psi _1`$; here and henceforth $`\frac{P_1}{P_2}`$ stands for the Radon-Nikodym density, and I write $`P_1=\frac{P_1}{P_2}P_2`$. We’ll glue all $`L_2(\mathrm{\Omega },,P)`$ together, forming $`L_2(\mathrm{\Omega },,𝒫)`$ that contains $`\sqrt{P}`$ for all $`P𝒫`$, as we’ll see soon.
There are several reasons for introducing ‘square roots of measures’. One reason. In quantum mechanics, Schrödinger’s wave function $`\psi (x)`$ determines a probability distribution $`|\psi (x)|^2dx`$. However, for infinitely many degrees of freedom we have no Lebesgue measure ($`dx`$). It could be convenient to describe a quantum state by an object that combines a measure and phases, something like $`\psi (x)\sqrt{dx}`$.
Old-fashioned tensor analysis stipulates a notion of a relative tensor, in particular, a relative scalar of a given weight (see for instance \[6, item 156 on pp. 345–346\]). A density of a measure on a manifold is a relative scalar field of weight $`1`$. Relative scalar fields of weight $`1/2`$ are smooth finite-dimensional ‘square roots of measures’ considered below.
Define $`L_2(\mathrm{\Omega },,𝒫)`$, denoted also by $`L_2(𝒫)`$ for short, as the Hilbert space of all families $`\psi =(\psi _P)_{P𝒫}`$ such that $`\psi _PL_2(\mathrm{\Omega },,P)`$ for every $`P𝒫`$, and
(1.1)
$$\psi _{P_2}=\sqrt{\frac{P_1}{P_2}}\psi _{P_1}$$
for all $`P_1,P_2𝒫`$. Linear operations and scalar product are defined naturally; for every $`P𝒫`$, the map $`L_2(𝒫)\psi \psi _PL_2(P)`$ is unitary.
Given $`P𝒫`$, we define $`\sqrt{P}L_2(𝒫)`$ by $`(\sqrt{P})_P=1`$ (identically on $`\mathrm{\Omega }`$). More generally, given $`P𝒫`$ and $`fL_2(P)`$, we define $`f\sqrt{P}L_2(𝒫)`$ by $`(f\sqrt{P})_P=f`$. Thus, $`\psi =\psi _P\sqrt{P}`$ for all $`P𝒫`$, $`\psi L_2(𝒫)`$. Now we may replace the notation $`\psi _P`$ by a more expressive notation
$$\psi _P=\frac{\psi }{\sqrt{P}}.$$
Given $`\psi ^{},\psi ^{\prime \prime }L_2(𝒫)`$, their scalar product is $`\psi ^{},\psi ^{\prime \prime }=\frac{\psi ^{}}{\sqrt{P}}\frac{\psi ^{\prime \prime }}{\sqrt{P}}𝑑P`$; the latter does not depend on $`P𝒫`$. Heuristically we could write $`\psi ^{},\psi ^{\prime \prime }=\psi ^{}(\omega )\psi ^{\prime \prime }(\omega )`$, however, I prefer the notation
$$\psi ^{},\psi ^{\prime \prime }=\frac{\psi ^{}(\omega )}{\sqrt{P(d\omega )}}\frac{\psi ^{\prime \prime }(\omega )}{\sqrt{P(d\omega )}}P(d\omega )=\psi ^{}(\omega )\psi ^{\prime \prime }(\omega )\frac{d\omega }{d\omega };$$
being rather awkward and illogical,<sup>3</sup><sup>3</sup>3Neither $`\psi (\omega )`$ nor $`\psi (d\omega )`$ is a logical notation for an object of the form $`f(\omega )\sqrt{P(d\omega )}`$. it still helps. Given $`\psi ^{},\psi ^{\prime \prime }L_2(𝒫)`$, we define the signed measure $`\psi ^{}\psi ^{\prime \prime }`$ by $`\frac{\psi ^{}\psi ^{\prime \prime }}{P}=\frac{\psi ^{}}{\sqrt{P}}\frac{\psi ^{\prime \prime }}{\sqrt{P}}`$, thus $`\psi ^{},\psi ^{\prime \prime }=(\psi ^{}\psi ^{\prime \prime })(\mathrm{\Omega })`$, the measure of the whole space. Note also that $`\sqrt{P_1},\sqrt{P_2}=\sqrt{P_1(d\omega )}\sqrt{P_2(d\omega )}`$ was used in Kakutani’s well-known work about equivalence of product measures. Two natural metrics on $`𝒫`$ define the same topology (the only one used here); I mean the variation distance, $`P_1P_2=\left|\frac{P_1}{P}\frac{P_2}{P}\right|𝑑P`$, and the angle in $`L_2(𝒫)`$, $`\mathrm{arccos}\sqrt{P_1},\sqrt{P_2}`$; the proof is left to the reader.<sup>4</sup><sup>4</sup>4In fact, $`\left(\frac{1}{2}P_1P_2\right)^21\sqrt{P_1},\sqrt{P_2}\frac{1}{2}P_1P_2`$, which is not used here.
## 2 Gaussian measures, quadratic norms, FHS
One may introduce a Gaussian measure as a probability measure on a Banach (or Hilbert, or Frechet, etc) space, such that every (continuous) linear functional has a normal distribution. Such a viewpoint is convenient for heuristic thinking but not for a formal presentation, since topological structures on the linear space with measure are in fact irrelevant. Following an old advice of A. Vershik, I prefer discarding any topological structure on the (Banach, etc.) space and even keeping implicit its linear structure. Everything can be formulated in terms of two different spaces, a probability space and a Hilbert space; the latter plays the role of the space of all measurable linear functionals, as well as its dual, the space of all admissible shifts.
Let $`(\mathrm{\Omega },,P)`$ be a probability space and $`GL_2(\mathrm{\Omega },,P)`$ a (closed) linear subspace, containing constants and generating the whole $`\sigma `$-field $``$. We call $`G`$ a *Gaussian space,* if every $`gG`$ has a normal distribution; the latter is therefore $`N(𝔼g,\mathrm{Var}g)=N(g𝑑P,g^2𝑑P(g𝑑P)^2)`$. Up to isomorphism, there is exactly one Gaussian space in every dimension ($`0,1,2,\mathrm{}`$ or $`\mathrm{}`$). That is, if $`G_1L_2(\mathrm{\Omega }_1,_1,P_1)`$ and $`G_2L_2(\mathrm{\Omega }_2,_2,P_2)`$ are Gaussian spaces and $`dimG_1=dimG_2`$, then there exists an isomorphism (invertible measure preserving map) between the probability spaces $`(\mathrm{\Omega }_1,_1,P_1)`$ and $`(\mathrm{\Omega }_2,_2,P_2)`$ that induces an isometry betwen $`G_1`$ and $`G_2`$. Moreover, each isometry between $`G_1`$ and $`G_2`$ corresponds to exactly one isomorphism of probability spaces.
The standard model is the space $`^{\mathrm{}}=\times \times \mathrm{}`$ of all sequences of real numbers, equipped with the product measure $`\gamma ^{\mathrm{}}=\gamma ^1\gamma ^1\mathrm{}`$ where $`\gamma ^1`$ is the standard normal distribution $`N(0,1)`$. Here $`G`$ consists of all measurable linear functionals $`^{\mathrm{}}(x_1,x_2,\mathrm{})c_kx_k`$ with $`c_k^2<\mathrm{}`$; that is, $`l_2`$ is the standard model of $`G`$. Any other model is necessarily isomorphic to the standard model, as far as $`G`$ is of infinite dimension; straightforward modifications for a finite dimension are left to the reader.
The same $`G`$ over $`^{\mathrm{}}`$ is also a Gaussian space w.r.t. many other measures on $`^{\mathrm{}}`$ equivalent to the standard measure $`\gamma ^{\mathrm{}}`$. In particular, consider the product measure $`\gamma `$ on $`^{\mathrm{}}`$ whose $`k`$-th factor is the normal distribution $`N(m_k,\sigma _k^2)`$ with given parameters $`m_k(\mathrm{},+\mathrm{})`$, $`\sigma _k(0,+\mathrm{})`$. The well-known S. Kakutani’s theorem on equivalence of infinite product measures shows that $`\gamma `$ is equivalent to $`\gamma ^{\mathrm{}}`$ if and only if
$$\underset{k}{}(\sigma _k1)^2<\mathrm{}\text{and}\underset{k}{}m_k^2<\mathrm{}.$$
Replacing coordinate axes by arbitrary orthogonal (in $`l_2`$) axes one gets all Gaussian measures on $`^{\mathrm{}}`$ equivalent to $`\gamma ^{\mathrm{}}`$, which is basically the well-known criterion (Feldman , Hajek , Segal ).
In terms of a Gaussian space $`GL_2(\mathrm{\Omega },,P)`$ we introduce the set $`𝒫`$ of all probability measures on $`(\mathrm{\Omega },)`$ equivalent to $`P`$, and its subset $`𝒫_G`$ consisting of all $`\gamma 𝒫`$ such that $`G`$ is also a Gaussian space in $`L_2(\mathrm{\Omega },,\gamma )`$; these measures will be called *Gaussian measures w.r.t. $`G`$.* Now we forget the initial measure $`P`$; each measure of $`𝒫_G`$ may serve as $`P`$ equally well. Up to isomorphism, the structure $`(\mathrm{\Omega },,𝒫,G)`$ is uniquely determined by $`dimG`$. I call $`(\mathrm{\Omega },,𝒫,G)`$ a *Gaussian type space.*
We introduce a quotient space $`G_0=G/\mathrm{Const}`$ (where $`\mathrm{Const}`$ is the one-dimensional space of constants) and its dual space $`G^0`$ (consisting of all continuous linear functionals on $`G_0`$). Every $`\gamma \mathrm{\Gamma }`$ determines a Hilbert norm $`_\gamma `$ on $`G_0`$ (that is, $`(G_0,_\gamma )`$ is a Hilbert space) via
$$g_\gamma ^2=\mathrm{Var}_\gamma (g)=g^2𝑑\gamma \left(g𝑑\gamma \right)^2,$$
which is insensitive to any constant added to $`gG`$. Each norm $`_\gamma `$ defines the same topology on $`G_0`$. However, the norms are equivalent in a stronger sense, namely, for any $`\gamma _1,\gamma _2\mathrm{\Gamma }`$ there exists a basis $`(g_k)`$ of $`G_0`$, orthogonal for both norms and such that the numbers $`\lambda _k=g_k_{\gamma _2}/g_k_{\gamma _1}`$ satisfy $`\lambda _k>0`$ and $`(\lambda _k1)^2<\mathrm{}`$ (which can be reformulated in terms of Hilbert-Schmidt operators; namely, the unit operator $`(G_0,_{\gamma _1})gg(G_0,_{\gamma _2})`$ must be an equivalence operator, as defined by Feldman \[7, Def. 1\]). I’ll call such norms *FHS-equivalent.*<sup>5</sup><sup>5</sup>5You may interpret FHS as Feldman-Hajek-Segal, or alternatively as Feldman-Hilbert-Schmidt. Of course, it is an equivalence relation \[7, Lemma 2\]. Numbers $`\lambda _k`$ determine the distance between $`\gamma _1,\gamma _2`$ (provided that $`\gamma _1,\gamma _2`$ have the same mean, that is, $`g𝑑\gamma _1=g𝑑\gamma _2`$ for all $`gG`$); namely, a simple (basically, one-dimensional) calculation gives
$$\sqrt{\gamma _1},\sqrt{\gamma _2}=\underset{k}{}\left(\frac{\lambda _k^{1/2}+\lambda _k^{1/2}}{2}\right)^{1/2};$$
of course, convergence of the product is equivalent to convergence of the series $`_k(\lambda _k1)^2`$.
###### 2.1 Definition.
An *FHS-space* is a pair $`(H,𝒩)`$ of a linear space $`H`$ and a set $`𝒩`$ of norms on $`H`$ such that
(a) every norm of $`𝒩`$ turns $`H`$ into a separable Hilbert space;
(b) all norms of $`𝒩`$ are pairwise FHS-equivalent;
(c) every norm FHS-equivalent to a norm of $`𝒩`$ belongs to $`𝒩`$.
Norms belonging to $`𝒩`$ will be called *admissible norms* on $`H`$.
Isomorphisms of FHS-spaces are basically the same as Feldman’s equivalence operators \[7, Def. 1\]. Up to isomorphism, there is exactly one FHS-space in every dimension ($`0,1,2,\mathrm{}`$ or $`\mathrm{}`$); just the same situation as for (separable) Hilbert spaces.
Every separable Hilbert space is isometric to a Gaussian space (over some probability space), and this superstructure brings no arbitrariness, as far as it is considered up to isomorphism. (One speaks about *the* isonormal random process on any given Hilbert space.) Similarly, every FHS-space may be identified with $`G/\mathrm{Const}`$ for *the* Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$ of the corresponding dimension. Thus, in principle one may prove a purely geometric statements about an FHS-space via Gaussian measures, and this way is indeed used in the next section. The correspondence between $`𝒩`$ (admissible norms) and $`𝒫_G`$ (Gaussian measures) transfers to $`𝒩`$ the natural topology of $`𝒫_G`$ (inherited from $`𝒫`$; recall the end of Sect. 1). In terms of the numbers $`\lambda _k`$, a basis of neighborhoods of a given admissible norm may be written as $`_k(\lambda _k1)^2<\epsilon `$.
The (additive group of) space $`G^0`$, dual to $`G_0`$, acts on $`(\mathrm{\Omega },,𝒫)`$ by automorphisms (invertible measurable transformations preserving $`𝒫`$) $`U_x`$, $`xG^0`$, such that
$$g(U_x\omega )g(\omega )=g,x\text{for almost all }\omega \mathrm{\Omega }$$
for all $`gG`$. Clearly, $`U_x`$ is uniquely determined. Existence of these $`U_x`$ may be checked just for the standard model, which makes it evident: $`G^0=l_2`$ acts on $`^{\mathrm{}}`$ by shifts, $`U_x(\omega )=\omega +x`$. The map $`U_x`$ sends each Gaussian measure $`\gamma _1𝒫_G`$ into another Gaussian measure $`\gamma _2𝒫_G`$ such that $`_{\gamma _1}=_{\gamma _2}`$ on $`G_0`$, and $`g𝑑\gamma _2g𝑑\gamma _1=g,x`$ for all $`gG`$. A simple (basically, one-dimensional) calculation gives
$$\sqrt{\gamma _1},\sqrt{\gamma _2}=\mathrm{exp}\left(\frac{1}{8}x^2\right),$$
where $`=_{\gamma _1}=_{\gamma _2}`$. In fact, $`\sqrt{\gamma _1}\sqrt{\gamma _2}=\mathrm{exp}\left(\frac{1}{8}x^2\right)\gamma `$, where $`\gamma `$ is the image of $`\gamma _1`$ under $`U_{x/2}`$.
Turn to the Hilbert space $`L_2(\mathrm{\Omega },,𝒫)`$ of ‘square roots of measures’. Transformations $`U_x`$ induce unitary operators on $`L_2(𝒫)`$; I denote them by $`U_x`$, too. Namely, for every $`\psi L_2(𝒫)`$ and $`P𝒫`$,
$$\frac{\psi }{\sqrt{P}}(\omega )=\frac{U_x\psi }{\sqrt{U_xP}}(U_x\omega ),$$
where $`U_xP`$ (denoted also by $`PU_x^1`$) is the image of $`P`$ under $`U_x`$. Thus, the FHS-space $`G^0`$ acts unitarily on the Hilbert space $`L_2(𝒫)`$.
There is also a natural projective action of the other FHS-space, $`G_0`$, on $`L_2(𝒫)`$. Namely, every $`gG`$ determines a unitary operator $`V_g:L_2(𝒫)L_2(𝒫)`$,
$$V_g\psi =e^{ig}\psi ,$$
the multiplication by the function $`\omega e^{ig(\omega )}`$. Given $`yG_0`$, we get $`V_y`$ determined up to a phase factor, which means a *projective* action. In fact, one can choose phase factors getting a unitary representation (which is evident for the standard model). Anyway,
$$V_yU_x=e^{ix,y}U_xV_y,$$
the well-known Weyl form of Canonical Commutation Relations. In this context, vectors $`\psi L_2(𝒫)`$ of the form $`\psi =V_y\sqrt{\gamma }`$, $`\gamma \mathrm{\Gamma }`$, $`yG_0`$, are known as coherent states, or quasi-free pure states, or Gaussian pure states. It is easy to see that
$$V_y\sqrt{\gamma },\sqrt{\gamma }=\mathrm{exp}\left(\frac{1}{2}y_\gamma ^2\right)$$
for $`yG_0`$.
## 3 Some geometry via measure theory
Many statements of this section are purely geometric and probably could be proved within the Hilbert space geometry, but I find it easier to prove them via measure theory. Usually, the corresponding properties of Gaussian measures hold for arbitrary (non-Gaussian) measures as well.
The formula $`H=H_1H_2`$ has several interpretations. It always implies that $`H_1,H_2`$ are subspaces of $`H`$ and every vector $`hH`$ has a unique representation as $`h_1+h_2`$ where $`h_1H_1`$, $`h_2H_2`$. However, when $`H`$ is a Hilbert space, one usually stipulates that $`H_1,H_2`$ are orthogonal. When $`H`$ is an FHS-space, we treat
$$H=H_1H_2$$
as follows: there exists an *admissible*<sup>6</sup><sup>6</sup>6Recall Definition 2.1. norm on $`H`$ that makes $`H_1,H_2`$ orthogonal (and of course, they span the whole $`H`$). Similarly, $`H=H_1\mathrm{}H_n`$ means existence of an admissible norm that makes $`H_1,\mathrm{},H_n`$ orthogonal (also, they span $`H`$).
###### 3.1 Proposition.
Let $`H`$ be an FHS-space, and $`H_1,H_{12},H_{23},H_3`$ its subspaces such that $`H_1H_{12}`$, $`H_{23}H_3`$, and
$$H_{12}H_3=H=H_1H_{23}.$$
Then the subspace $`H_2=H_{12}H_{23}`$ satisfies
$$H_1H_2H_3=H,H_1H_2=H_{12},H_2H_3=H_{23}.$$
Note. Given that $`H_1H_2H_3=H`$, other relations $`H_1H_2=H_{12}`$, $`H_2H_3=H_{23}`$ may be treated in the topological sense.
###### Proof.
The decomposition $`H=H_1H_{23}`$ determines a projection $`P_1:HH`$ such that $`P_1H=H_1`$ and $`1P_1=P_{23}`$ is also a projection, $`P_{23}H=H_{23}`$. The same for $`P_3`$ and $`P_{12}=1P_3`$. The inclusion $`H_1H_{12}`$ gives $`P_1=P_{12}P_1=(1P_3)P_1`$, that is, $`P_3P_1=0`$; similarly, $`P_1P_3=0`$. So, our projections commute with each other. Introducing
$$P_2=P_{12}P_{23}=P_{23}P_{12}=(1P_1)(1P_3)=1P_1P_3,$$
we have $`P_2^2=P_{12}P_{23}P_{12}P_{23}=P_{12}^2P_{23}^2=P_{12}P_{23}=P_2`$; that is, $`P_2`$ is also a projection, and $`P_1+P_2+P_3=1`$. It follows that $`P_2H=\left((P_1+P_2)H\right)\left((P_2+P_3)H\right)=H_{12}H_{23}=H_2`$. So, the relation $`H=H_1H_2H_3`$ holds in the topological sense (that is, when $`H`$ is treated as a linear topological space). The following lemma completes the proof. ∎
###### 3.2 Lemma.
Let $`H`$ be an FHS-space, $`H_1,H_2,H_3`$ its subspaces such that $`H=H_1H_2H_3`$ in the topological sense, and $`(H_1+H_2)H_3=H=H_1(H_2+H_3)`$ in the FHS sense (the bracketed sums being topological). Then $`H=H_1H_2H_3`$ in the FHS sense.
###### Proof.
We introduce a Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$ and identify $`H`$ with $`G_0`$. Subspaces $`H_1,H_2,H_3`$ generate sub-$`\sigma `$-fields $`_1,_2,_3`$. Note that $`H_1+H_2`$ generates $`_1_2`$, the least $`\sigma `$-field containing both $`_1`$ and $`_2`$. The following two lemmas (and one definition) complete the proof. ∎
###### 3.3 Definition.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, and $`_0,_1,\mathrm{},_n`$ sub-$`\sigma `$-fields. We write $`_0=_1\mathrm{}_n`$, if $`_1\mathrm{}_n=_0`$ and there exists $`P𝒫`$ making $`_1,\mathrm{},_n`$ independent.<sup>7</sup><sup>7</sup>7Which means $`P(A_1\mathrm{}A_n)=P(A_1)\mathrm{}P(A_n)`$ for all $`A_1_1,\mathrm{},A_n_n`$.
###### 3.4 Lemma.
Let $`(\mathrm{\Omega },,𝒫,G)`$ be a Gaussian type space, $`H=G_0`$ the corresponding FHS-space, $`H_1,\mathrm{},H_nH`$ subspaces, and $`_1,\mathrm{},_n`$ corresponding sub-$`\sigma `$-fields. Then
$$H=H_1\mathrm{}H_n\text{if and only if}=_1\mathrm{}_n.$$
###### Proof.
‘Only if’: take a Gaussian measure $`\gamma \mathrm{\Gamma }`$ such that $`_\gamma `$ makes $`H_1,\mathrm{},H_n`$ orthogonal, then $`\gamma `$ makes $`_1,\mathrm{},_n`$ independent.
‘If’: some $`P𝒫`$ makes $`_1,\mathrm{},_n`$ independent; however, $`P`$ need not be Gaussian. We take any Gaussian measure $`\gamma _0\mathrm{\Gamma }`$ and introduce $`_k`$-measurable densities
$$f_k=\frac{\gamma _0|__k}{P|__k}\text{for }k=1,\mathrm{},n.$$
The measure $`\gamma =f_1\mathrm{}f_nP`$ still makes $`_1,\mathrm{},_n`$ independent, and $`\gamma |__k=f_k(P|__k)=\gamma _0|__k`$. So, w.r.t. $`\gamma `$ the spaces $`H_1,\mathrm{},H_n`$ are *independent* Gaussian spaces. Therefore their sum $`H`$ is Gaussian w.r.t. $`\gamma `$, that is, $`\gamma \mathrm{\Gamma }`$. ∎
###### 3.5 Lemma.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, and $`_1,_2,_3`$ sub-$`\sigma `$-fields such that
$$(_1_2)_3==_1(_2_3).$$
Then
$$=_1_2_3.$$
###### Proof.
We have $`P,Q𝒫`$ such that $`_1_2`$ and $`_3`$ are $`P`$-independent, while $`_1`$ and $`_2_3`$ are $`Q`$-independent. Consider the $`_1_2`$-measurable density<sup>8</sup><sup>8</sup>8In fact, it is a conditional expectation w.r.t. $`P`$, $`f_{12}=𝔼\left(\frac{Q}{P}|_1_2\right)`$.
$$f_{12}=\frac{Q|_{_1_2}}{P|_{_1_2}}$$
and the measure $`R=f_{12}P𝒫`$, then $`R|_{_1_2}=Q|_{_1_2}`$. The $`P`$-independence of $`_1_2`$ and $`_3`$ implies their $`R`$-independence. For every $`A_1`$, $`B_2`$, $`C_3`$
$$\begin{array}{c}R(ABC)=R(AB)R(C)=Q(AB)R(C)=\hfill \\ \hfill =Q(A)Q(B)R(C)=R(A)R(B)R(C).\end{array}$$
The proof of Proposition 3.1 is now complete. If you find the proof of Lemma 3.5 rather tricky, consider the following calculation as a clue. Let $`\psi _kL_2(\mathrm{\Omega },_k,𝒫)`$ ($`k=1,2,3`$), then (writing for short $`_{12}=_1_2`$ etc.),
$$\begin{array}{c}(\psi _1\psi _2)\psi _3=\frac{\psi _1\psi _2}{\sqrt{P|_{_{12}}}}\frac{\psi _3}{\sqrt{P|__3}}\sqrt{P}=\hfill \\ \hfill =\frac{\psi _1}{\sqrt{Q|__1}}\frac{\psi _2}{\sqrt{Q|__2}}\sqrt{\frac{Q|_{_{12}}}{P|_{_{12}}}}\frac{\psi _3}{\sqrt{P|__3}}\sqrt{P},\end{array}$$
and the equality $`(\psi _1\psi _2)\psi _3^2=\psi _1^2\psi _2^2\psi _3^2`$ turns into the following equality for functions $`f_1=(\psi _1/\sqrt{Q|__1})^2`$, $`f_2=(\psi _2/\sqrt{Q|__2})^2`$, $`f_3=(\psi _3/\sqrt{P|__3})^2`$:
$$f_1f_2f_3\frac{Q|_{_{12}}}{P|_{_{12}}}𝑑P=\left(f_1𝑑Q|__1\right)\left(f_2𝑑Q|__2\right)\left(f_3𝑑P|__3\right).$$
The argument is easily generalized for $`_1,\mathrm{},_n`$. Now we turn to infinite sequences of subspaces.
###### 3.6 Definition.
Let $`X`$ be a metrizable topological space and $`X_1,X_2,\mathrm{}X`$ closed subsets. We define $`lim\; inf_n\mathrm{}X_n`$ as the set of limits of all convergent sequences $`x_1,x_2,\mathrm{}`$ such that $`x_1X_1,x_2X_2,\mathrm{}`$
The set $`lim\; infX_n`$ is always closed. If $`X`$ is a linear topological space and $`X_n`$ are linear subspaces, then $`lim\; infX_n`$ is a linear subspace. If $`X`$ is the $`\sigma `$-field $``$ of a measure type space $`(\mathrm{\Omega },,𝒫)`$ and $`X_n`$ are sub-$`\sigma `$-fields, then $`lim\; infX_n`$ is a sub-$`\sigma `$-field.<sup>9</sup><sup>9</sup>9The natural topology of $``$ is defined by a metric $`\mathrm{dist}(A,B)=P(AB)+P(BA)`$; the metric depends on $`P𝒫`$, but the topology does not. Of course, $`X`$ is $`mod0`$ rather than $``$ itself. Proofs of these facts are left to the reader.
###### 3.7 Proposition.
Let $`H`$ be an FHS-space, $`E_n,F_nH`$ subspaces ($`n=1,2,\mathrm{}`$), and $`lim\; infE_n=H`$. For each $`n`$ denote by $`𝒩_n`$ the set of all admissible norms on $`H`$ that make $`E_n,F_n`$ orthogonal. Then the set $`lim\; inf𝒩_n`$ is either the empty set, or the whole $`𝒩`$.
###### Proof.
We identify $`H`$ with $`G_0`$ of a Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$ and use the natural homeomorphism $`\gamma _\gamma `$ between the space $`𝒩`$ of admissible norms and the space $`𝒫_G`$ of Gaussian measures. Subspaces $`E_n,F_n`$ generate corresponding sub-$`\sigma `$-fields $`_n,_n`$. The set $`𝒩_n𝒩`$ corresponds to $`𝒫_n𝒫_G`$, where $`𝒫_n`$ consists of all measures making $`_n,_n`$ independent. The following two lemmas complete the proof. ∎
###### 3.8 Lemma.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, $`E,E_1,E_2,\mathrm{}L_0(𝒫)`$ closed linear subspaces, and $`,_1,_2,\mathrm{}`$ the sub-$`\sigma `$-fields generated by the subspaces. Then
$$lim\; infE_n=E\text{implies}lim\; inf_n.$$
###### Proof.
We have to prove that every $`flim\; infE_n`$ is measurable w.r.t. $`lim\; inf_n`$. It suffices to prove that the set $`A=\{\omega :f(\omega )a\}`$ belongs to $`lim\; inf_n`$ for every $`a`$ such that the set $`\{\omega :f(\omega )=a\}`$ is negligible (indeed, such $`a`$ are dense in $``$). We take $`f_nE_n`$ such that $`f_nf`$ in $`L_0(𝒫)`$, consider sets $`A_n=\{\omega :f_n(\omega )a\}`$ and note that $`A_n_n`$ and $`A_nA`$. ∎
Note. In general, $`lim\; inf_n`$ need not be equal to $``$; it may happen that $`_n=`$ but $`lim\; infE_n=\{0\}`$ (even for one-dimensional $`E_n`$). However, in the Gaussian case (that is, when $`(\mathrm{\Omega },,𝒫,G)`$ is a Gaussian type space, and each $`E_n`$ is a Gaussian space) the equality $`lim\; inf_n=`$ holds, which is neither proved nor used here.
###### 3.9 Lemma.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, $`_n,_n`$ sub-$`\sigma `$-fields ($`n=1,2,\mathrm{}`$), and $`lim\; inf_n=`$. For each $`n`$ denote by $`𝒫_n`$ the set of all $`P𝒫`$ such that $`_n`$ and $`_n`$ are $`P`$-independent. Then
(a) the set $`lim\; inf𝒫_n`$ is either the empty set or the whole $`𝒫`$;
(b) if $`lim\; inf𝒫_n=𝒫`$ then $`(PQ)|__n0`$ for all $`P,Q𝒫`$ (here $``$ means the total variation).
###### Proof.
Assume that $`lim\; inf𝒫_n`$ is nonempty; we have $`P_n,P𝒫`$ such that $`P_nP`$ and $`_n,_n`$ are $`P_n`$-independent. Let $`K(0,\mathrm{})`$ be a finite set and $`f:\mathrm{\Omega }K`$ an $``$-measurable function satisfying $`f𝑑P=1`$. Take $`_n`$-measurable functions $`f_n:\mathrm{\Omega }K`$ such that $`f_nf`$ in $`L_0(𝒫)`$ and $`f_n𝑑P=1`$. The $`P_n`$-independent $`\sigma `$-fields $`_n,_n`$ are also $`Q_n`$-independent, where $`Q_n=f_nP_nfP=Q`$ (since $`f_nP_nf_nPf_n_{\mathrm{}}P_nP0`$). Thus $`Q_n𝒫_n`$ and $`Qlim\; inf𝒫_n`$. However, such measures $`Q`$ (for all $`f`$ and $`K`$) are dense in $`𝒫`$. So, $`lim\; inf𝒫_n=𝒫`$, which is (a). Also, the $`P_n`$-independence of $`_n,_n`$ implies $`Q_n|__n=P_n|__n`$. However, $`(P_nP)|__nP_nP0`$, and similarly $`(Q_nQ)|__n0`$. So, $`(PQ)|__n0`$ for all $`Q`$ of a dense set, therefore for all $`Q`$, which is (b). ∎
The proof of Proposition 3.7 is now complete.
###### 3.10 Definition.
(a) Let $`H`$ be an FHS-space, $`E_n,F_nH`$ subspaces, and $`lim\; infE_n=H`$. We say that $`F_n`$ is asymptotically orthogonal to $`E_n`$, if there exists a convergent<sup>10</sup><sup>10</sup>10In the space $`𝒩`$ of all admissible norms, whose topology is defined in Sect. 2. sequence of admissible norms $`_n`$ such that for each $`n`$, $`F_n`$ is orthogonal to $`E_n`$ w.r.t. $`_n`$.
(b) Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, $`_n,_n`$ sub-$`\sigma `$-fields, and $`lim\; inf_n=`$. We say that $`_n`$ is asymptotically independent of $`_n`$, if there exists a convergent<sup>11</sup><sup>11</sup>11In the space $`𝒫`$ whose topology is defined in Sect. 1. sequence of measures $`P_n𝒫`$ such that for each $`n`$, $`_n`$ is independent of $`_n`$ w.r.t. $`P_n`$.
Proposition 3.7 states that the norms $`_n`$ can be chosen so as to converge to any given admissible norm, provided that $`F_n`$ is asymptotically orthogonal to $`E_n`$. Similarly, if $`_n`$ is asymptotically independent of $`_n`$, then the measures $`P_n`$ can be chosen so as to converge to any given $`P𝒫`$ due to 3.9(a), and $`(PQ)|__n0`$ for all $`P,Q𝒫`$ due to 3.9(b). Also, in the case of a Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$ and $`H=G_0`$, asymptotical orthogonality of subspaces is equivalent to asymptotical independence of the corresponding sub-$`\sigma `$-fields.
###### 3.11 Lemma.
Let $`H`$ be an FHS-space, $`F_nH`$ subspaces, then the following conditions are equivalent.
(a) For every $`f_1F_1`$, $`f_2F_2`$, …, if the sequence $`(f_n)`$ is bounded then $`f_n0`$ weakly.<sup>12</sup><sup>12</sup>12That is, $`f_n,h0`$ for every $`hH`$.
(b) For every finite-dimensional subspaces $`E_1,E_2,\mathrm{}`$ such that $`lim\; infE_n=H`$ there exist integers $`k_1k_2\mathrm{}`$ such that $`k_n\mathrm{}`$ and $`F_n`$ is asymptotically orthogonal to $`E_{k_n}`$.
###### Proof.
We choose an admissible norm on $`H`$, thus turning $`H`$ into a Hilbert space. Condition (a) becomes
$$hH\underset{fF_n,f1}{sup}f,h\underset{n\mathrm{}}{\overset{}{}}0,$$
that is,
$$hH\mathrm{}(h,F_n)\underset{n\mathrm{}}{\overset{}{}}\frac{\pi }{2},$$
where the angle is defined by $`\mathrm{cos}\mathrm{}(h,F_n)=sup\{h,f:fF_n,f1\}`$. It is equivalent to
$$E\mathrm{}(E,F_n)\underset{n\mathrm{}}{\overset{}{}}\frac{\pi }{2},$$
where $`E`$ runs over finite-dimensional subspaces, and $`\mathrm{cos}\mathrm{}(E,F_n)=sup\{e,f:eE,fF_n,e1,f1\}`$. The following lemma completes the proof, provided that $`k_n`$ tends to $`\mathrm{}`$ slowly enough. Namely, in terms of $`\delta (,)`$ introduced there, it suffices that $`\delta (\mathrm{}(E_{k_n},F_n),dimE_{k_n})\underset{n\mathrm{}}{\overset{}{}}0`$. ∎
###### 3.12 Lemma.
Let $`H`$ be an FHS-space, $`E,FH`$ subspaces, $`dim(E)<\mathrm{}`$, $`EF=\{0\}`$. Then for every admissible norm $`_1`$ there exists an admissible norm $`_2`$ such that $`E,F`$ are orthogonal w.r.t. $`_2`$, and
$$\mathrm{dist}(_1,_2)\delta (\mathrm{}(E,F),dimE)$$
for some function $`\delta :[0,\frac{\pi }{2}]\times \{0,1,2,\mathrm{}\}(0,\mathrm{})`$ such that for every $`n`$, $`\delta (\alpha ,n)0`$ for $`\alpha \frac{\pi }{2}`$.<sup>13</sup><sup>13</sup>13Of course, $`\delta `$ does not depend on $`H,E,F`$.
###### Proof.
We equip $`H`$ with the norm $`_1`$, thus turning $`H`$ into a Hilbert space, and consider orthogonal projections $`Q_E,Q_F`$ onto $`E,F`$ respectively. Introduce subspaces $`EF^{}`$, $`E^{}F`$, $`E^{}F^{}`$ (here $`E^{}`$ is the orthogonal complement of $`E`$); the subspaces are orthogonal to each other, and invariant under both $`Q_E`$ and $`Q_F`$. Therefore
$$H=H_0(EF^{})(E^{}F)(E^{}F^{}),$$
where $`H_0`$ is another subspace invariant under $`Q_E,Q_F`$ (since these operators are Hermitian). Introduce $`E_0=EH_0`$, $`F_0=FH_0`$, then $`Q_Eh_0=Q_{E_0}h_0`$ for all $`h_0H_0`$ (since $`Q_E`$ commutes with $`Q_{H_0}`$), and $`Q_Fh_0=Q_{F_0}h_0`$. We may get rid of $`H_0^{}`$ by letting
$$h_0+h_1_2^2=h_0_2^2+h_1_1^2\text{for all }h_0H_0\text{}h_1H_0^{}\text{.}$$
In other words, we’ll construct $`_2`$ on $`H_0`$ while preserving both the given norm on $`H_1`$ and the orthogonality of $`H_0,H_1`$. Now we forget about $`H_1`$, assuming that $`H=H_0`$, $`E=E_0`$, $`F=F_0`$.
So, we have $`EF=\{0\}`$, $`EF^{}=\{0\}`$, $`E^{}F=\{0\}`$, $`E^{}F^{}=\{0\}`$. The latter implies $`dim(F^{})\mathrm{codim}(E^{})=dimE`$. Similarly, $`dimFdimE`$. Therefore $`H`$ is finite-dimensional, $`dimH2dimE`$.
Both $`Q_E`$ and $`Q_F`$ commute with the Hermitian operator $`C=\frac{1}{2}(2Q_E1)(2Q_F1)+\frac{1}{2}(2Q_F1)(2Q_E1)`$. The spectrum of $`C`$ consists of some numbers $`\mathrm{cos}2\phi _k`$ of multiplicity 2 (though, some $`\phi _k`$ may coincide), and $`0<\phi _k<\frac{\pi }{2}`$ (the case $`\phi _k=0`$ is excluded by $`EF=\{0\}`$; the case $`\phi _k=\pi /2`$ is excluded by $`EF^{}=\{0\}`$, $`E^{}F=\{0\}`$, $`E^{}F^{}=\{0\}`$). Accordingly, $`H_0`$ decomposes into the (orthogonal) direct sum of planes, $`H=H_1\mathrm{}H_d`$, $`dimH_k=2`$, invariant under $`Q_E,Q_F`$. Subspaces $`E_k=EH_k`$, $`F_k=FH_k`$ are two lines on the plane $`H_k`$, and $`\mathrm{}(E_k,F_k)=\phi _k`$; $`k=1,\mathrm{},d`$; $`ddimE`$. Clearly,
$$\mathrm{}(E,F)=\mathrm{min}(\phi _1,\mathrm{},\phi _d).$$
We construct $`_2`$ on each $`H_k`$ separately, while preserving their orthogonality. Elementary 2-dimensional geometry shows that the corresponding numbers $`\lambda _k^{},\lambda _k^{\prime \prime }`$ (two numbers for each plane) are, in the optimal case,
$$\lambda _k^{}=\left(\mathrm{tan}\frac{\phi }{2}\right)^{1/2},\lambda _k^{\prime \prime }=\left(\mathrm{tan}\frac{\phi }{2}\right)^{1/2}.$$
The corresponding angle $`\beta `$ between $`\sqrt{\gamma _1}`$ and $`\sqrt{\gamma _2}`$ is given by
$$\mathrm{cos}\beta =\underset{k=1}{\overset{d}{}}\left(\frac{\mathrm{tan}^{1/4}\frac{\phi }{2}+\mathrm{tan}^{1/4}\frac{\phi }{2}}{2}\right)^1,$$
therefore
$$\beta \mathrm{arccos}\left(\frac{\mathrm{tan}^{1/4}\frac{\alpha }{2}+\mathrm{tan}^{1/4}\frac{\alpha }{2}}{2}\right)^{dimE};$$
here $`\beta =\mathrm{}(\sqrt{\gamma _1},\sqrt{\gamma _2})=\mathrm{dist}(_1,_2)`$, $`\alpha =\mathrm{}(E,F)`$. ∎
###### 3.13 Definition.
Let $`H`$ be an FHS-space, $`F_nH`$ subspaces. We write $`lim\; supF_n=\{0\}`$, if the sequence $`(F_n)`$ satisfies equivalent conditions (a), (b) of Lemma 3.11.
Note. More generally, one could define $`lim\; supF_n`$ as the set of limits of all weakly convergent subsequences of all bounded sequences $`f_1,f_2,\mathrm{}`$ such that $`f_1F_1,f_2F_2,\mathrm{}`$ It is in general not a linear space, but anyway, $`(lim\; supE_n^{})^{}=lim\; infE_n`$ in a Hilbert space. For an FHS-space, as well as a separable Banach space, $`F_n`$ should be situated in the dual space. However, all that is not needed here.
###### 3.14 Theorem.
Let $`H`$ be an FHS-space, $`E_n,F_nH`$ subspaces ($`n=1,2,\mathrm{}`$) such that $`lim\; infE_n=H`$, and $`lim\; supF_n=\{0\}`$, and $`H=E_nF_n`$ (in the FHS sense) for all $`n`$. Then there exist subspaces $`G_n,H_nH`$ such that
$$E_n=G_nH_n\text{and}H=G_nH_nF_n$$
(both in the FHS sense), and $`lim\; infG_n=H`$, and $`H_nF_n`$ is asymptotically orthogonal to $`G_n`$.
###### Proof.
We choose an admissible norm on $`H`$, thus turning $`H`$ into a Hilbert space. Let $`LH`$ be a finite-dimensional subspace, $`L\{0\}`$. For any given $`n`$ consider the pair $`L,E_n`$. Its geometry may be described (similarly to the proof of Lemma 3.12) via angles $`\phi _1^{(n)},\mathrm{},\phi _{d_n}^{(n)}[0,\frac{\pi }{2})`$, $`d_ndimL`$. This time, zero angles are allowed, since $`LE_n`$ need not be $`\{0\}`$. It may happen that $`d_n<dimL`$, since $`LE_n^{}`$ need not be $`\{0\}`$. However,
$$\underset{xL,x0}{sup}\mathrm{}(x,E_n)=\alpha _n0\text{for }n\mathrm{};$$
for large $`n`$ we have $`\alpha _n<\pi /2`$ which implies $`d_n=d=dimL`$ and $`\mathrm{max}(\phi _1^{(n)},\mathrm{},\phi _d^{(n)})=\alpha _n`$. We may send $`L`$ into $`E_n`$ rotating it by $`\phi _1^{(n)},\mathrm{},\phi _d^{(n)}`$. In other words, there is a rotation $`U_n:HH`$ such that
$$U_n(L)E_n\text{and}U_n12\mathrm{sin}\frac{\alpha _n}{2}\underset{n\mathrm{}}{\overset{}{}}0.$$
We choose subspaces $`L_kH`$ such that $`dimL_k=k`$ and $`lim\; infL_k=H`$.<sup>14</sup><sup>14</sup>14Of course, one may take $`L_1L_2\mathrm{}`$ Introduce
$$\alpha _{k,n}=\underset{xL_k,x0}{sup}\mathrm{}(x,E_n),$$
then $`\alpha _{k,n}\underset{n\mathrm{}}{\overset{}{}}0`$ for each $`k`$. On the other hand, introduce
$$\beta _{k,n}=\frac{\pi }{2}\mathrm{}(L_k,F_n).$$
Similarly to the proof of Lemma 3.11 we have $`\beta _{k,n}\underset{n\mathrm{}}{\overset{}{}}0`$ for each $`k`$, therefore<sup>15</sup><sup>15</sup>15Recall that $`\delta (,)`$ is introduced in Lemma 3.12. $`\delta (\frac{\pi }{2}\beta _{k_n,n},k_n)\underset{n\mathrm{}}{\overset{}{}}0`$ if $`k_n`$ tends to $`\mathrm{}`$ slowly enough. However, we choose $`k_1k_2\mathrm{}`$, $`k_n\mathrm{}`$ so as to satisfy a stronger condition:
$$\delta (\frac{\pi }{2}\alpha _{k_n,n}\beta _{k_n,n},k_n)\underset{n\mathrm{}}{\overset{}{}}0.$$
We take
$$G_n=U_n(L_{k_n}),$$
where rotations $`U_n`$ satisfy $`U_n(L_{k_n})E_n`$ and $`U_n12\mathrm{sin}(\frac{1}{2}\alpha _{k_n,n})0`$. Then $`lim\; infG_n=H`$, and
$$\frac{\pi }{2}\mathrm{}(G_n,F_n)\alpha _{k_n,n}+\beta _{k_n,n};$$
due to Lemma 3.12, $`F_n`$ is asymptotically orthogonal to $`G_n`$. We take admissible norms $`_n`$ such that $`F_n`$ is orthogonal to $`G_n`$ w.r.t. $`_n`$. Consider the orthogonal complement $`M_n`$ of $`G_n`$ w.r.t. $`_n`$; clearly, $`M_n`$ is asymptotically orthogonal to $`G_n`$. We have $`F_nM_n`$ and $`H=G_nM_n`$ (in the FHS-sense). On the other hand, $`G_nE_n`$ and $`H=E_nF_n`$. Proposition 3.1 states that the subspace
$$H_n=E_nM_n$$
satisfies $`G_nH_nF_n=H`$ and $`G_nH_n=E_n`$ (and also $`H_nF_n=M_n`$). ∎
## 4 Density matrices
Recall the notion of a density matrix (borrowed from quantum theory). Let $`H_1,H_2`$ be Hilbert spaces, $`H=H_1H_2`$, and $`\psi H`$, $`\psi =1`$. Every unit vector $`\xi H_1`$ determines a subspace $`\xi H_2H`$ and the corresponding projection operator $`Q_{\xi H_2}=Q_\xi \mathrm{𝟏}_{H_2}`$; here $`Q_\xi :H_1H_1`$, $`Q_\xi x=(x,\xi )\xi `$ is a one-dimensional projection, and $`\mathrm{𝟏}_{H_2}:H_2H_2`$, $`\mathrm{𝟏}_{H_2}y=y`$ the unit operator. The function
$$\xi Q_{\xi H_2}\psi ^2$$
is a quadratic form on $`H_1`$. The corresponding operator $`\rho _\psi :H_1H_1`$ satisfies
$$\rho _\psi \xi ,\xi =Q_{\xi H_2}\psi ^2=(Q_\xi \mathrm{𝟏}_{H_2})\psi ,\psi $$
(that is, $`\rho _\psi _\xi =Q_\xi \mathrm{𝟏}_{H_2}_\psi `$) for all $`\xi H_1`$; one calls $`\rho _\psi `$ the *density matrix* of $`\psi `$ (on $`H_1`$). In terms of an orthonormal basis $`(e_k)`$ of $`H_2`$,
$$\psi =\underset{k}{}\psi _ke_k\text{for some }\psi _kH_1;$$
$$(Q_\xi \mathrm{𝟏}_{H_2})\psi =\underset{k}{}(Q_\xi \psi _k)e_k=\underset{k}{}\psi _k,\xi \xi e_k;$$
$$\rho _\psi \xi ,\xi =\underset{k}{}|\psi _k,\xi |^2.$$
Note that
$$\rho _\psi 0;\mathrm{Tr}(\rho _\psi )=1;$$
$$\mathrm{Tr}\left((\rho _{\psi _1}\rho _{\psi _2})A\right)2\psi _1\psi _2A$$
for all unit vectors $`\psi _1,\psi _2H`$ and operators $`A:H_1H_1`$. The inequality may be proven as follows: $`\mathrm{Tr}(\rho _\psi Q_\xi )=\mathrm{Tr}\left((Q_\xi \mathrm{𝟏}_{H_2})Q_\psi \right)`$; $`\mathrm{Tr}(\rho _\psi A)=\mathrm{Tr}\left((A\mathrm{𝟏}_{H_2})Q_\psi \right)`$; $`\mathrm{Tr}\left((\rho _{\psi _1}\rho _{\psi _2})A\right)=\mathrm{Tr}\left((A\mathrm{𝟏}_{H_2})(Q_{\psi _1}Q_{\psi _2})\right)A\mathrm{𝟏}_{H_2}Q_{\psi _1}Q_{\psi _2}A2\psi _1\psi _2`$.
Note also that
$$\psi ^{}=(U_1U_2)\psi \text{implies}\rho _\psi ^{}=U_1\rho _\psi U_1^{}$$
for all unitary operators $`U_1:H_1H_1`$, $`U_2:H_2H_2`$. Proof: $`\rho _\psi ^{}\xi ,\xi =(Q_\xi \mathrm{𝟏}_{H_2})\psi ^{},\psi ^{}=(U_1U_2)^{}(Q_\xi \mathrm{𝟏}_{H_2})(U_1U_2)\psi ,\psi =(U_1^{}Q_\xi U_1U_2^{}\mathrm{𝟏}_{H_2}U_2)\psi ,\psi =(Q_{U_1^{}\xi }\mathrm{𝟏}_{H_2})\psi ,\psi =\rho _\psi U_1^{}\xi ,U_1^{}\xi =U_1\rho _\psi U_1^{}\xi ,\xi `$ for all unit vectors $`\xi H_1`$, $`\psi H`$.
Assume in addition that $`H_1=L_2(\mathrm{\Omega }_1,_1,P_1)`$, $`H_2=L_2(\mathrm{\Omega }_2,_2,P_2)`$, then (after the usual identification) $`H=L_2(\mathrm{\Omega },,P)`$ where $`(\mathrm{\Omega },,P)=(\mathrm{\Omega }_1,_1,P_1)(\mathrm{\Omega }_2,_2,P_2)`$. We have
$$\rho _\psi \xi ,\xi =\xi (\omega ^{})\overline{\xi (\omega ^{\prime \prime })}\rho _\psi (\omega ^{},\omega ^{\prime \prime })P_1(d\omega ^{})P_1(d\omega ^{\prime \prime }),$$
where $`\rho _\psi `$ is an element of $`L_2\left((\mathrm{\Omega }_1,_1,P_1)(\mathrm{\Omega }_1,_1,P_1)\right)`$ (the kernel of the operator $`\rho _\psi :H_1H_1`$), namely,
$$\rho _\psi (\omega ^{},\omega ^{\prime \prime })=\overline{\psi (\omega ^{},\omega _2)}\psi (\omega ^{\prime \prime },\omega _2)P_2(d\omega _2).$$
In terms of a basis,
$$\psi (\omega _1,\omega _2)=\underset{k}{}\psi _k(\omega _1)e_k(\omega _2);$$
$$\rho _\psi (\omega ^{},\omega ^{\prime \prime })=\underset{k}{}\overline{\psi _k(\omega ^{})}\psi _k(\omega ^{\prime \prime }).$$
The proof is basically a calculation:
$$\begin{array}{c}\rho _\psi (\omega ^{},\omega ^{\prime \prime })=\overline{\psi (\omega ^{},\omega _2)}\psi (\omega ^{\prime \prime },\omega _2)P_2(d\omega _2)=\hfill \\ \hfill =\overline{\underset{k}{}\psi _k(\omega ^{})e_k(\omega _2)}\underset{l}{}\psi _l(\omega ^{\prime \prime })e_l(\omega _2)P_2(d\omega _2)=\\ \hfill =\underset{k,l}{}\overline{\psi _k(\omega ^{})}\psi _l(\omega ^{\prime \prime })\overline{e_k(\omega _2)}e_l(\omega _2)P_2(d\omega _2)=\underset{k}{}\overline{\psi _k(\omega ^{})}\psi _k(\omega ^{\prime \prime });\end{array}$$
$$\begin{array}{c}\rho _\psi \xi ,\xi =\underset{k}{}|\psi _k,\xi |^2=\underset{k}{}\left|\psi _k(\omega )\overline{\xi (\omega )}P_1(d\omega )\right|^2=\hfill \\ \hfill =\underset{k}{}\left(\overline{\psi _k(\omega ^{})}\xi (\omega ^{})P_1(d\omega ^{})\psi _k(\omega ^{\prime \prime })\overline{\xi (\omega ^{\prime \prime })}P_1(d\omega ^{\prime \prime })\right)=\\ \hfill =\xi (\omega ^{})\overline{\xi (\omega ^{\prime \prime })}\left(\underset{k}{}\overline{\psi _k(\omega ^{})}\psi _k(\omega ^{\prime \prime })\right)P_1(d\omega ^{})P_1(d\omega ^{\prime \prime })=\\ \hfill =\xi (\omega ^{})\overline{\xi (\omega ^{\prime \prime })}\rho _\psi (\omega ^{},\omega ^{\prime \prime })P_1(d\omega ^{})P_1(d\omega ^{\prime \prime }).\end{array}$$
We turn to measure type spaces $`(\mathrm{\Omega },,𝒫)=(\mathrm{\Omega }_1,_1,𝒫_1)(\mathrm{\Omega }_2,_2,𝒫_2)`$ (which means $`P_1P_2𝒫`$ for some, therefore all, $`P_1𝒫_1`$ and $`P_2𝒫_2`$), and the corresponding Hilbert spaces; $`H=L_2(\mathrm{\Omega },,𝒫)=L_2(\mathrm{\Omega }_1,_1,𝒫_1)L_2(\mathrm{\Omega }_2,_2,𝒫_2)=H_1H_2`$ under the natural identification
$$\frac{\psi _1\psi _2}{\sqrt{P_1P_2}}(\omega _1,\omega _2)=\frac{\psi _1}{\sqrt{P_1}}(\omega _1)\frac{\psi _2}{\sqrt{P_2}}(\omega _2),$$
that is,
$$\psi _1\psi _2=\left(\frac{\psi _1}{\sqrt{P_1}}\frac{\psi _2}{\sqrt{P_2}}\right)\sqrt{P_1P_2}L_2(𝒫)$$
for all $`\psi _1L_2(𝒫_1)`$, $`\psi _2L_2(𝒫_2)`$, $`P_1𝒫_1`$, $`P_2𝒫_2`$. An element $`\psi H`$, $`\psi =1`$, determines an operator $`\rho _\psi :H_1H_1`$, whose kernel (denoted also by $`\rho _\psi `$) belongs to $`L_2\left((\mathrm{\Omega }_1,_1,𝒫_1)(\mathrm{\Omega }_1,_1,𝒫_1)\right)`$;
$$\rho _\psi \xi ,\xi =\frac{\xi }{\sqrt{P_1}}(\omega ^{})\frac{\xi }{\sqrt{P_1}}(\omega ^{\prime \prime })\frac{\rho _\psi }{\sqrt{P_1P_1}}(\omega ^{},\omega ^{\prime \prime })P_1(d\omega ^{})P_1(d\omega ^{\prime \prime }),$$
$$\frac{\rho _\psi }{\sqrt{P_1P_1}}(\omega ^{},\omega ^{\prime \prime })=\overline{\frac{\psi }{\sqrt{P_1P_2}}(\omega ^{},\omega _2)}\frac{\psi }{\sqrt{P_1P_2}}(\omega ^{\prime \prime },\omega _2)P_2(d\omega _2)$$
for all $`\xi L_2(𝒫_1)`$, $`\xi =1`$, and $`P_1𝒫_1`$, $`P_2𝒫_2`$.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space and $`_1,_2`$ sub-$`\sigma `$-fields such that $`=_1_2`$ (as defined by 3.3), then
$$L_2()=L_2(_1)L_2(_2).$$
Here $`L_2()=L_2(\mathrm{\Omega },,𝒫)`$ and $`L_2(_1)=L_2(\mathrm{\Omega },_1,𝒫|__1)`$, the same for $`_2`$,<sup>16</sup><sup>16</sup>16Of course, $`𝒫|__1=\{P|__1:P𝒫\}`$ consists of restricted measures. Alternatively one may introduce quotient spaces $`\mathrm{\Omega }_1=\mathrm{\Omega }/_1`$, $`\mathrm{\Omega }_2=\mathrm{\Omega }/_2`$ and identify $`\mathrm{\Omega }`$ with $`\mathrm{\Omega }_1\times \mathrm{\Omega }_2`$. and the natural identification is made, namely,
$$\left(f_1\sqrt{P|__1}\right)\left(f_2\sqrt{P|__2}\right)=(f_1f_2)\sqrt{P}$$
whenever $`f_1L_2(\mathrm{\Omega },_1,P)`$, $`f_2L_2(\mathrm{\Omega },_2,P)`$, and $`P𝒫`$ makes $`_1,_2`$ independent. We need a counterpart of Lemma 3.9.
###### 4.1 Lemma.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, $`_n,_n`$ sub-$`\sigma `$-fields ($`n=1,2,\mathrm{}`$), $`=_n_n`$ for each $`n`$, and $`lim\; inf_n=`$, and $`_n`$ is asymptotically independent of $`_n`$. Then for every $`P𝒫`$
$$lim\; inf\left(L_2(_n)\sqrt{P|__n}\right)=L_2().$$
That is, for every $`\psi L_2()`$ there exist $`\xi _nL_2(_n)`$ such that $`\psi \xi _n\sqrt{P|__n}0`$ when $`n\mathrm{}`$.
###### Proof.
Similarly to the proof of 3.9, we take $`P_n𝒫`$ such that $`_n,_n`$ are $`P_n`$-independent and $`P_nP`$. We consider an arbitrary finite set $`K`$, an arbitrary $``$-measurable function $`f:\mathrm{\Omega }K`$, and the corresponding vector $`\psi =f\sqrt{P}L_2()`$. We construct $`_n`$-measurable functions $`f_n:\mathrm{\Omega }K`$ such that $`f_nf`$ in $`L_0(𝒫)`$, and corresponding vectors $`\psi _n=f_n\sqrt{P_n}L_2()`$. Independence of $`_n,_n`$ w.r.t. $`P_n`$ means that $`\sqrt{P_n}=\sqrt{P_n|__n}\sqrt{P_n|__n}`$, therefore $`\psi _n=\xi _n\sqrt{P_n|__n}L_2(_n)\sqrt{P_n|__n}`$, where $`\xi _n=f_n\sqrt{P_n|__n}L_2(_n)`$. However, $`\psi _n\psi `$, since $`f_n\sqrt{P_n}f_n\sqrt{P}f_n_{\mathrm{}}\sqrt{P_n}\sqrt{P}0`$. Also $`P_n|__nP|__nP_nP0`$, therefore $`\sqrt{P_n|__n}\sqrt{P|__n}0`$. So, $`\psi \xi _n\sqrt{P|__n}\psi \psi _n+\xi _n\sqrt{P_n|__n}\xi _n\sqrt{P|__n}0`$, and $`\psi lim\; inf\left(L_2(_n)\sqrt{P|__n}\right)`$. It remains to note that such vectors $`\psi `$ (for all $`f`$ and $`K`$) are dense in $`L_2()`$. ∎
## 5 Fock spaces and tail density matrices
Recall the correspondence (described in Sect. 2) between Gaussian type spaces $`(\mathrm{\Omega },,𝒫,G)`$ and FHS-spaces $`G_0=G/\mathrm{Const}`$. We know that any FHS-space $`G_0`$ determines (up to isomorphism) the corresponding Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$, which in turn determines the Hilbert space $`H=L_2(\mathrm{\Omega },,𝒫)`$. I denote the relation by
$$H=\mathrm{Exp}(G_0),$$
and call $`H`$ the Fock exponential of $`G_0`$.<sup>17</sup><sup>17</sup>17By choosing an admissible norm on $`G_0`$ one turns $`G_0`$ into a Hilbert space, in which case $`\mathrm{Exp}(G_0)`$ contains a special element, ‘ground state vector’ $`\sqrt{\gamma }`$ (where $`\gamma `$ is the Gaussian measure corresponding to the chosen norm); that is the classical Fock construction. Why call it ‘exponential’? Since
$$G_0=G_1G_2\text{implies}\mathrm{Exp}(G_0)=\mathrm{Exp}(G_1)\mathrm{Exp}(G_2)$$
in the following sense. Assume that $`G_1,G_2G_0`$ are subspaces such that $`G_0=G_1G_2`$ (in the FHS sense, as defined in Sect. 3). Then the sub-$`\sigma `$-fields $`_1,_2`$, generated by $`G_1,G_2`$ respectively, satisfy $`=_1_2`$, therefore $`L_2()=L_2(_1)L_2(_2)`$, as explained in Sect. 4. So, every decomposition of $`G_0`$ into a direct sum determines a decomposition of $`\mathrm{Exp}(G_0)`$ into a tensor product.
Given such a decomposition $`G_0=G_1G_2`$, every unit vector $`\psi \mathrm{Exp}(G_0)`$ determines a density matrix $`\rho _\psi `$ on $`\mathrm{Exp}(G_2)`$. Do not think, however, that $`\rho _\psi `$ is uniquely determined by $`\psi `$ and $`G_2`$; also $`G_1`$ influences $`\rho _\psi `$.
Consider an infinite sequence of decompositions, $`G_0=E_nF_n`$, of a single FHS-space $`G_0`$. We want to know, whether or not $`\rho _n(\psi _1)\rho _n(\psi _2)0`$ when $`n\mathrm{}`$ for all unit vectors $`\psi _1,\psi _2\mathrm{Exp}(G_0)`$; here $`\rho _n(\psi )`$ is the density matrix on $`\mathrm{Exp}(F_n)`$ that corresponds to $`\psi \mathrm{Exp}(E_n)\mathrm{Exp}(F_n)`$, and the trace norm is used,
$$\rho _n(\psi _1)\rho _n(\psi _2)=\underset{A1}{sup}\mathrm{Tr}(\rho _n(\psi _1)\rho _n(\psi _2))A);$$
here $`A`$ runs over Hermitian operators on $`\mathrm{Exp}(F_n)`$. In general we have only $`\rho _n(\psi _1)\rho _n(\psi _2)2\psi _1\psi _2`$.
Assume that $`lim\; infE_n=G_0`$. If $`F_n`$ is asymptotically orthogonal to $`E_n`$, then $`\rho _n(\psi _1)\rho _n(\psi _2)0`$, which follows easily from Lemma 4.1. Namely, the lemma represents $`\psi `$ as the limit of $`\xi _n\sqrt{P|__n}`$; however, $`\rho _n(\xi _n\sqrt{P|__n})`$ is the one-dimensional projection $`Q_n`$ onto $`\sqrt{P|__n}`$, therefore $`\rho _n(\psi )Q_n2\psi \xi _n\sqrt{P|__n}0`$, and so, $`\rho _n(\psi _1)\rho _n(\psi _2)\rho _n(\psi _1)Q_n+\rho _n(\psi _2)Q_n0`$. However, the asymptotical orthogonality condition may be dropped, as we’ll see now.
###### 5.1 Proposition.
Let $`G_0`$ be an FHS-space, $`E_n,F_nG_0`$ subspaces such that $`lim\; infE_n=G_0`$, and $`lim\; supF_n=\{0\}`$, and $`G_0=E_nF_n`$ (in the FHS sense) for all $`n`$. Then $`\rho _n(\psi _1)\rho _n(\psi _2)0`$ when $`n\mathrm{}`$, for all unit vectors $`\psi _1,\psi _2\mathrm{Exp}(G_0)`$.
###### Proof.
Theorem 3.14 gives us $`G_n,H_nG_0`$ such that $`E_n=G_nH_n`$ and $`G_0=G_nH_nF_n`$ (both in the FHS sense), and $`lim\; infG_n=G_0`$, and $`H_nF_n`$ is asymptotically orthogonal to $`G_n`$. Lemma 4.1 gives us a representation
$$\psi =\underset{n\mathrm{}}{lim}(\xi _n\chi _n)$$
for an arbitrary unit vector $`\psi \mathrm{Exp}(G_0)`$; here $`\xi _n`$ are unit vectors of $`\mathrm{Exp}(G_n)`$, $`\chi _n`$ are unit vectors of $`\mathrm{Exp}(H_nF_n)`$, and these $`\chi _n`$ (unlike $`\xi _n`$) do not depend on $`\psi `$. We have $`\rho _n(\psi )\rho _n(\xi _n\chi _n)2\psi \xi _n\chi _n0`$. However, the density matrix $`\rho _n(\xi _n\chi _n)`$ on $`\mathrm{Exp}(F_n)`$ that corresponds to the vector $`\xi _n\chi _n\mathrm{Exp}(G_n)\mathrm{Exp}(H_n)\mathrm{Exp}(F_n)`$ is the same as the density matrix on $`\mathrm{Exp}(F_n)`$ that corresponds to the vector $`\chi _n\mathrm{Exp}(H_n)\mathrm{Exp}(F_n)`$. The vector does not depend on $`\psi `$, therefore $`\rho _n(\xi _n\chi _n)`$ does not depend on $`\psi `$. So, $`\rho _n(\psi _1)\rho _n(\psi _2)\rho _n(\psi _1)\rho _n(\xi _n\chi _n)+\rho _n(\psi _2)\rho _n(\xi _n\chi _n)0`$. ∎
Consider a decomposition $`G_0=G_1G_2`$ of an FHS-space $`G_0`$ into the direct sum (in the FHS sense) of subspaces $`G_1,G_2G_0`$, and the corresponding decomposition $`G^0=G^1G^2`$ of its dual FHS-space $`G^0`$; that is, $`G^1=G_2^{}`$ is the annihilator of $`G_2`$ in $`G^0`$, and $`G^2=G_1^{}`$. Of course, also $`G_1=(G^2)^{}`$ and $`G_2=(G^1)^{}`$. Introduce an admissible norm $``$ on $`G_0`$ (note that $`G_1,G_2`$ need not be orthogonal w.r.t. $``$), and its dual norm on $`G^0`$ (denoted by $``$ as well). Consider
$`\mathrm{dist}(g,G_2)`$ $`=\underset{g_2G_2}{inf}gg_2=\underset{x_1G^1,x_11}{sup}g,x_1,`$
$`\mathrm{dist}(x,G^2)`$ $`=\underset{x_2G^2}{inf}xx_2=\underset{g_1G_1,g_11}{sup}g_1,x`$
for any $`gG_0`$, $`xG^0`$.
Introduce the corresponding Hilbert spaces $`H=\mathrm{Exp}(G_0)`$, $`H_1=\mathrm{Exp}(G_1)`$, $`H_2=\mathrm{Exp}(G_2)`$; we have $`H=H_1H_2`$. Recall the operators $`U_x`$ and $`V_g`$ for $`xG^0`$, $`gG_0`$, satisfying the Canonical Commutation Relations $`V_gU_x=e^{ig,x}U_xV_g`$. Let $`(\mathrm{\Omega },,𝒫,G)`$ be the corresponding Gaussian type space (thus, $`H=L_2(\mathrm{\Omega },,𝒫)`$), and $`\gamma 𝒫_G`$ a Gaussian measure such that $`_\gamma =`$. We know that the vector $`\psi =\sqrt{\gamma }H`$ satisfies $`U_x\psi ,\psi =\mathrm{exp}\left(\frac{1}{8}x^2\right)`$, $`V_g\psi ,\psi =\mathrm{exp}\left(\frac{1}{2}g^2\right)`$ for all $`xG^0`$, $`gG_0`$. Denote by $`\rho (\psi )`$ the corresponding density matrix on $`H_1`$. In the following lemma, $`\rho (\psi ^{})`$ for some other $`\psi ^{}`$ are the corresponding density matrices on $`H_1`$, and a function $`M:[0,\mathrm{})[0,\mathrm{})`$ is defined by<sup>18</sup><sup>18</sup>18The exact form of the function is of no importance here; we only need to know that $`M(r_n)0`$ implies $`r_n0`$.
$$M(r)=\underset{\phi [0,\pi ]}{\mathrm{max}}\left(\mathrm{exp}\left(\frac{\phi ^2}{2r^2}\right)2\mathrm{sin}\frac{\phi }{2}\right).$$
###### 5.2 Lemma.
For all $`xG^0`$, $`gG_0`$
$`\rho (\psi )\rho (U_x\psi )`$ $`M\left(\mathrm{dist}(x,G^2)\right),`$
$`\rho (\psi )\rho (V_g\psi )`$ $`M\left(2\mathrm{dist}(g,G_2)\right).`$
###### Proof.
Let $`x=y+z`$, $`yG^1`$, $`zG^2`$, then $`U_x=U_y^{(1)}U_z^{(2)}`$ (the notation being self-explanatory), which implies $`\rho (U_x\psi )=U_y^{(1)}\rho (\psi )U_y^{(1)}`$. For every $`gG_1`$
$$\mathrm{Tr}\left(\rho (U_x\psi )V_g^{(1)}\right)=\mathrm{Tr}\left(U_y^{(1)}\rho (\psi )U_y^{(1)}V_g^{(1)}\right)=\mathrm{Tr}\left(\rho (\psi )U_y^{(1)}V_g^{(1)}U_y^{(1)}\right);$$
however, $`V_g^{(1)}U_y^{(1)}=\mathrm{exp}(ig,y)U_y^{(1)}V_g^{(1)}`$ and $`g,y=g,x`$ (since $`g(G^2)^{}`$); we have $`\mathrm{Tr}\left(\rho (U_x\psi )V_g^{(1)}\right)=\mathrm{exp}(ig,x)\mathrm{Tr}\left(\rho (\psi )V_g^{(1)}\right)`$. Note that $`|\mathrm{Tr}\left((\rho (U_x\psi )\rho (\psi ))V_g^{(1)}\right)|\rho (U_x\psi )\rho (\psi )`$, since $`\mathrm{Re}(e^{i\alpha }\mathrm{Tr}\left((\rho (U_x\psi )\rho (\psi ))V_g^{(1)}\right)=\mathrm{Tr}\left((\rho (U_x\psi )\rho (\psi ))\mathrm{Re}(e^{i\alpha }V_g^{(1)})\right)\rho (U_x\psi )\rho (\psi )`$ for all $`\alpha `$. We have
$$\rho (U_x\psi )\rho (\psi )|1e^{ig,x}||\mathrm{Tr}\left(\rho (\psi )V_g^{(1)}\right)|;$$
$$|1e^{ig,x}|=2\left|\mathrm{sin}\frac{g,x}{2}\right|;$$
$$\mathrm{Tr}\left(\rho (\psi )V_g^{(1)}\right)=\mathrm{Tr}\left((V_g^{(1)}\mathrm{𝟏}_{H_2})Q_\psi \right)=(V_g^{(1)}\mathrm{𝟏}_{H_2})\psi ,\psi =$$
$$=V_g\psi ,\psi =\mathrm{exp}\left(\frac{1}{2}g^2\right);$$
so,
$$\rho (U_x\psi )\rho (\psi )\underset{gG_1}{sup}\left(\mathrm{exp}\left(\frac{1}{2}g^2\right)2\left|\mathrm{sin}\frac{g,x}{2}\right|\right).$$
Denote $`r=\mathrm{dist}(x,G^2)`$. We need only one ray of vectors $`gG_1`$ such that $`g,x=rg`$. For every $`\phi [0,\pi ]`$ there exists such $`g`$, satisfying $`g,x=\phi `$ and $`g=\phi /r`$, which gives $`\rho (U_x\psi )\rho (\psi )\mathrm{exp}\left(\frac{\phi ^2}{2r^2}\right)2\mathrm{sin}\frac{\phi }{2}`$; the supremum over $`\phi `$ gives the first inequality.
For the second inequality, the proof is quite similar. Only $`U_y^{(1)}`$ and $`V_g^{(1)}`$ change places, and $`\mathrm{exp}\left(\frac{1}{8}x^2\right)`$ appears instead of $`\mathrm{exp}\left(\frac{1}{2}g^2\right)`$, which leads to $`M(2r)`$ instead of $`M(r)`$. ∎
###### 5.3 Theorem.
Let $`G_0`$ be an FHS-space, $`E_n,F_nG_0`$ subspaces such that $`G_0=E_nF_n`$ (in the FHS sense) for all $`n`$. For every unit vector $`\psi \mathrm{Exp}(G_0)`$ let $`\rho _n(\psi )`$ denote the density matrix on $`F_n`$ that corresponds to $`\psi `$. Then the following two conditions are equivalent.
(a) $`lim\; infE_n=G_0`$ and $`lim\; supF_n=\{0\}`$.
(b) $`\rho _n(\psi _1)\rho _n(\psi _2)0`$ when $`n\mathrm{}`$, for all unit vectors $`\psi _1,\psi _2\mathrm{Exp}(G_0)`$.
###### Proof.
Proposition 5.1 gives (a) $``$ (b). Assume (b); we have to prove (a). We choose a Gaussian measure $`\gamma `$ and apply Lemma 5.2 to $`\psi =\sqrt{\gamma }`$:
$`M\left(\mathrm{dist}(x,F_n^{})\right)\rho _n(\psi )\rho _n(U_x\psi )0,`$
$`M\left(2\mathrm{dist}(g,E_n)\right)\rho _n(\psi )\rho _n(V_g\psi )0,`$
which implies that $`\mathrm{dist}(x,F_n^{})0`$ and $`\mathrm{dist}(g,E_n)0`$ (when $`n\mathrm{}`$) for all $`gG_0`$, $`xG^0`$ (the dual to $`G_0`$). The latter, $`\underset{eE_n}{inf}ge0`$, shows that $`lim\; infE_n=G_0`$. The former, $`\underset{fF_n,f1}{sup}f,x0`$, shows that $`lim\; supF_n=\{0\}`$. ∎
## 6 Borel measurability of it all
Recall some notions and results about Borel measurability (see \[5, Chapter 3\], \[11, Chapter 3\]). A *Borel space* is a set equipped with a $`\sigma `$-field (of subsets). The subsets belonging to the $`\sigma `$-field are called Borel measurable sets (or ‘measurable sets’, or ‘Borel sets’). A subset of a Borel space is naturally a Borel space. The product of two Borel spaces is naturally a Borel space. A Borel measurable map (or ‘measurable map’, or ‘Borel map’) is a map from one Borel space into another, such that the inverse image of every measurable set is a measurable set. A Borel isomorphism between two Borel spaces is an invertible measurable map whose inverse is also measurable. (Note that no measure (type) is given, and so, no subset is negligible; every single point counts.)
A *Polish space* is a topological space which is homeomorphic to a separable complete metric space. A Polish space is naturally equipped with the $`\sigma `$-field generated by all open sets, thus, it is a Borel space. Surprisingly, the Borel space does not depend (up to isomorphism) on the Polish space, as far as it is uncountable. That is, every two uncountable Polish spaces are Borel isomorphic. Moreover, all uncountable Borel sets in Polish spaces are Borel isomorphic. A Borel space is called *standard,* if it is isomorphic to a Borel subset of a Polish space. Up to isomorphism, there is a single uncountable standard Borel space, a single countable (infinite) one, and for each (finite) $`n`$, a single $`n`$-point one.
Let $`X`$ be a Polish space and $`𝐅(X)`$ the set of all nonempty closed subsets of $`X`$. There is a natural Borel structure on $`𝐅(X)`$, namely, the $`\sigma `$-field generated by sets of the form
$$\{F𝐅(X):FU\mathrm{}\}$$
where $`U`$ varies over open sets in $`X`$. Thus $`𝐅(X)`$ is a Borel space; it is called the Effros Borel space of $`X`$, and is standard (see \[11, Th. 3.3.10\]). There is a sequence $`(f_n)`$ of Borel measurable maps $`f_n:𝐅(X)X`$ such that every $`F𝐅(X)`$ is the closure of the countable (or finite) set $`\{f_1(F),f_2(F),\mathrm{}\}`$; it is called Castaing’s theorem (see \[11, Prop. 5.2.7\]). Therefore, for any Borel space $`T`$, a general form of a measurable map $`f:T𝐅(X)`$ is
(6.1)
$$f(t)=\mathrm{closure}\left(\{f_1(t),f_2(t),\mathrm{}\}\right),$$
where $`f_1,f_2,\mathrm{}:TX`$ are measurable maps. Note that the disjoint union of all $`F𝐅(X)`$, defined as the set of all pairs $`(F,x)`$ such that $`F𝐅(X)`$ and $`xF`$, is naturally a standard Borel space, since it is a Borel subset of $`𝐅(X)\times X`$.
Now we apply all that to our matter. Let $`H`$ be a (separable) Hilbert space and $`𝐋(H)`$ the set of all (closed) linear subspaces of $`H`$. Then $`H`$ is a Polish space, $`𝐋(H)𝐅(H)`$, and we get measurable maps $`f_n:𝐋(H)H`$ such that every $`L𝐋(H)`$ is spanned by (and even the closure of) $`\{f_1(L),f_2(L),\mathrm{}\}`$. Applying the usual orthogonalization process to the sequence $`\left(f_n(L)\right)`$ we get a new sequence, denote it again by $`\left(f_n(L)\right)`$, such that
(6.2)
$$\{f_n(L):1n<1+\mathrm{dim}L\}\text{is an orthonormal basis of }L$$
for every $`L𝐋(H)`$, and still, $`f_n:𝐋(H)H`$ are Borel measurable, and in addition, $`f_n(L)=0`$ when $`n1+\mathrm{dim}L`$. The same argument shows also that $`\mathrm{dim}L\{0,1,2,\mathrm{};\mathrm{}\}`$ is a measurable function of $`L`$.
There is a natural map, $`F\mathrm{span}F`$, from $`𝐅(H)`$ to $`𝐋(H)`$; namely, $`\mathrm{span}F`$ is the (closed) subspace spanned by $`F`$. The map is measurable, which follows from (6.1). Indeed, if $`F`$ is the closure of $`\{f_n(F):n=1,2,\mathrm{}\}`$, then $`\mathrm{span}F`$ is the closure of
(6.3)
$$\{\alpha _1f_1(F)+\mathrm{}+\alpha _nf_n(F):\alpha _1,\mathrm{},\alpha _n,n=1,2,\mathrm{}\},$$
still a countable set of measurable functions (indexed by finite sequences $`(\alpha _1,\mathrm{},\alpha _n)`$ of rational numbers).
###### 6.4 Lemma.
Linear subspaces of a Hilbert space are a standard Borel space.<sup>19</sup><sup>19</sup>19*Closed* linear subspaces of a *separable* Hilbert space, of course.
###### Proof.
$`𝐋(H)=\{F𝐅(H):\mathrm{span}(F)=F\}`$, therefore $`𝐋(H)`$ is a Borel subset of the standard Borel space $`𝐅(H)`$.<sup>20</sup><sup>20</sup>20Indeed, $`F(F,\mathrm{span}F)`$ is a Borel map $`𝐅(H)𝐅(H)\times 𝐅(H)`$, and the diagonal is a Borel subset of $`𝐅(H)\times 𝐅(H)`$.
###### 6.5 Lemma.
The closure of $`L_1+L_2`$ is a jointly Borel measurable function of linear subspaces $`L_1,L_2`$ of a Hilbert space.
The proof is left to the reader. Hint: Similar to (6.3) but simpler.<sup>21</sup><sup>21</sup>21An alternative way: $`L_1L_2`$ is measurable in $`L_1,L_2`$ by \[11, Exercise 3.3.11(ii)\], thus $`\mathrm{span}(L_1L_2)`$ is also measurable. It is a good luck that we need unions, not intersections; see the note after Exercise 3.3.11 in .
###### 6.6 Lemma.
The orthogonal projection of $`x`$ to $`L`$ is jointly measurable in $`xH`$ and $`L𝐋(H)`$.
###### Proof.
The projection is the limit (for $`n\mathrm{}`$) of the orthogonal projection of $`x`$ to $`\mathrm{span}\{f_k(L):kn\}`$. Measurability of the latter implies that of the former. ∎
Note also that the disjoint union of all $`L𝐋(H)`$ is a standard Borel space, and linear operations are measurable in the following sense: $`(L,h_1+h_2)`$ is jointly Borel measurable in $`(L,h_1)`$ and $`(L,h_2)`$ on the domain consisting of all pairs $`((L_1,h_1),(L_2,h_2))`$ where $`L_1,L_2𝐋(X)`$, $`h_1L_1`$, $`h_2L_2`$ satisfy $`L_1=L_2`$. The same for other linear combinations, and for the scalar product.
Let $`H`$ be an FHS-space (rather than a Hilbert space). Lemmas 6.4 and 6.5 still hold.
###### 6.7 Lemma.
The set of all pairs $`(L_1,L_2)`$ such that $`H=L_1L_2`$ (in the FHS sense, see Sect. 3) is a Borel subset of $`𝐋(H)\times 𝐋(H)`$.
###### Proof.
The relation $`H=L_1L_2`$ means that, first, $`L_1,L_2`$ are orthogonal in some admissible norm, and second, $`L_1+L_2`$ is dense in $`H`$. The latter condition defines a measurable set of pairs $`(L_1,L_2)`$ due to Lemma 6.5. The former condition may be expressed in terms of the infinite matrix
$$M(L_1,L_2)=\left(f_k(L_1),f_l(L_2)\right)_{k,l}$$
where $`f_n`$ are as in (6.2). The relevant set of matrices is Borel measurable. ∎
We turn to $`\sigma `$-fields. Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space and $`𝐀()`$ the set of all sub-$`\sigma `$-fields of $``$.<sup>22</sup><sup>22</sup>22As before, each $`\sigma `$-field must contain all $`𝒫`$-negligible sets. Then $`_0=mod0`$ is a complete Boolean algebra and a Polish space (recall Footnote 8), and $`𝐀()`$ may be identified with the set $`𝐀(_0)`$ of all closed subalgebras of $`_0`$. Thus, $`𝐀()=𝐀(_0)𝐅(_0)`$, and we get measurable maps $`f_n:𝐀(_0)_0`$ such that every $`𝒜𝐀(_0)`$ is generated by (and even the closure of) $`\{f_1(𝒜),f_2(𝒜),\mathrm{}\}`$.
Striving to a counterpart of (6.2), recall the space $`L_0(\mathrm{\Omega },,𝒫)`$ of all equivalence classes of measurable maps $`\mathrm{\Omega }`$. Every $`XL_0(\mathrm{\Omega },,𝒫)`$ generates a $`\sigma `$-field $`\sigma (X)𝐀(_0)`$. Every $`𝒜𝐀(_0)`$ is $`\sigma (X)`$ for some $`XL_0(\mathrm{\Omega },,𝒫)`$; the set of all such $`X`$ (for a given $`𝒜`$) is usually large and non-closed. Nevertheless a selection is constructed below.
###### 6.8 Lemma.
There exists a Borel map $`𝒜X_𝒜`$ from $`𝐀(_0)`$ to $`L_0(\mathrm{\Omega },,𝒫)`$ such that $`\sigma (X_𝒜)=𝒜`$ for all $`A𝐀(_0)`$.
###### Proof.
One may take
$$X_𝒜(\omega )=\underset{n=1}{\overset{\mathrm{}}{}}\frac{2}{3^n}\mathrm{𝟏}_{f_n(𝒜)}(\omega );$$
here $`\mathrm{𝟏}_{f_n(𝒜)}(\omega )`$ is equal to $`1`$ if $`\omega f_n(𝒜)`$ and $`0`$ otherwise. ∎
A $`\sigma `$-field $`𝒜`$ is nonatomic if and only if $`X_𝒜`$ is nonatomic (that is, $`X_𝒜^1(\{x\})`$ is negligible for every $`x`$). Nonatomic elements of $`L_0(\mathrm{\Omega },,𝒫)`$ are a Borel set. Therefore, nonatomic $`\sigma `$-fields are a Borel set (in $`𝐀(_0)`$). Similarly, the number of atoms ($`0,1,2,\mathrm{}`$ or $`\mathrm{}`$) is a Borel function of $`𝒜`$, as well as their (ordered) probabilities.
There is a natural map, $`F\sigma (F)`$, from $`𝐅(_0)`$ to $`𝐀(_0)`$; namely, $`\sigma (F)`$ is the $`\sigma `$-field generated by $`F`$. The map is measurable, which follows from (6.1) similarly to (6.3).<sup>23</sup><sup>23</sup>23Arbitrary combinations of Boolean operations (union, intersection, complement) are used here instead of the linear combinations used in (6.2).
Proofs of the following Lemmas 6.96.11 are left to the reader, since they are similar to 6.46.6 respectively. Especially for 6.11 a hint: $`{\displaystyle \frac{Q|_𝒜}{P|_𝒜}}=\underset{n\mathrm{}}{lim}{\displaystyle \frac{Q|_{𝒜_n}}{P|_{𝒜_n}}}`$ where $`𝒜_n`$ is generated by $`f_1(𝒜),\mathrm{},f_n(𝒜)`$.
###### 6.9 Lemma.
Sub-$`\sigma `$-fields of $``$ are a standard Borel space.
###### 6.10 Lemma.
The $`\sigma `$-field $`\sigma (𝒜_1𝒜_2)`$ generated by $`𝒜_1,𝒜_2`$ is a jointly measurable function of sub-$`\sigma `$-fields $`𝒜_1,𝒜_2`$.
The set $`𝒫`$ of equivalent probability measures on $`(\mathrm{\Omega },)`$ is also a standard Borel space, since it is homeomorphic (in fact, isometric) to a subset of $`L_1(P)`$ (the choice of $`P𝒫`$ does not matter); the whole $`L_1(P)`$ is a Polish space, and the subset, consisting of all strictly positive functions whose integral is equal to $`1`$, is a Borel subset.
###### 6.11 Lemma.
The Radon-Nikodym density
$$\frac{Q|_𝒜}{P|_𝒜}$$
(treated as an element of $`L_2(\mathrm{\Omega },,𝒫)`$ that belongs in fact to $`L_0(\mathrm{\Omega },𝒜,𝒫)`$) is jointly measurable in $`P,Q𝒫`$ and $`𝒜𝐀()`$.
Given $`P𝒫`$ and $`XL_0(\mathrm{\Omega },,𝒫)`$, we get a probability measure on $``$, namely, $`SP\left(\{\omega :X(\omega )S\}\right)`$ for Borel sets $`S`$. The measure will be called the distribution of $`X`$ w.r.t. $`P`$ and denoted by $`X(P)`$. Note that $`X(P)`$ is jointly measurable in $`X`$ and $`P`$, that is, $`(X,P)X(P)`$ is a Borel map from $`L_0(𝒫)\times 𝒫`$ to the space of probability distributions on $``$. Indeed, if $`\phi `$ is a bounded continuous function $``$ then $`\phi d(X(P))=\phi (X())𝑑P`$ is continuous in $`(X,P)`$.
###### 6.12 Lemma.
(a) The set $`𝒟`$ of all pairs $`(𝒜_1,𝒜_2)`$ such that $`𝒜_1𝒜_2`$ is well-defined,<sup>24</sup><sup>24</sup>24It means existence of $`P𝒫`$ that makes $`𝒜_1,𝒜_2`$ independent (recall Def. 3.3). When defined, $`𝒜_1𝒜_2`$ is just $`\sigma (𝒜_1𝒜_2)`$. is a Borel subset of $`𝐀()\times 𝐀()`$, and the map $`(𝒜_1,𝒜_2)𝒜_1𝒜_2`$ from $`𝒟`$ to $`𝐀()`$ is Borel measurable.
(b) For every $`P𝒫`$, the map
$$(𝒜_1,𝒜_2)\frac{P|_{𝒜_1𝒜_2}}{P|_{𝒜_1}P|_{𝒜_2}}$$
from $`𝒟`$ to $`L_0(𝒫)`$ is Borel measurable.
###### Proof.
(a) The condition $`(𝒜_1,𝒜_2)𝒟`$ may be expressed in terms of the measure
$$\mu _{𝒜_1,𝒜_2}=(X_{𝒜_1},X_{𝒜_2})(P)$$
on $`^2`$. (The choice of $`P𝒫`$ does not matter.) That is the joint probability distribution of random variables $`X_{𝒜_1},X_{𝒜_2}`$ on $`(\mathrm{\Omega },,P)`$, and its marginal distributions are $`X_{𝒜_1}(P),X_{𝒜_2}(P)`$. Clearly, $`\mu _{𝒜_1,𝒜_2}`$ is a product measure if and only if $`𝒜_1,𝒜_2`$ are independent w.r.t. $`P`$. Thus, the relevant condition on $`\mu _{𝒜_1,𝒜_2}`$ says that $`\mu _{𝒜_1,𝒜_2}`$ must be equivalent to a product measure. The set of all such measures on $`^2`$ is Borel measurable. The map $`(𝒜_1,𝒜_2)𝒜_1𝒜_2`$ is the restriction to $`𝒟`$ of the map $`(𝒜_1,𝒜_2)\sigma (𝒜_1𝒜_2)`$ measurable by Lemma 6.10.
(b) For such $`𝒜_1,𝒜_2`$ the map $`(X_{𝒜_1},X_{𝒜_2})`$ is an isomorphism ($`mod0`$) between $`(\mathrm{\Omega },𝒜_1𝒜_2,P)`$ and $`(^2,\mathrm{},\mu )`$, where $`\mu =\mu _{𝒜_1,𝒜_2}`$ (and the $`\sigma `$-field of $`\mu `$-measurable subsets of $`^2`$ is suppressed in the notation). Therefore,
$$\frac{P}{P|_{𝒜_1}P|_{𝒜_2}}(\omega )=\frac{\mu }{\mu _1\mu _2}(X_{𝒜_1}(\omega ),X_{𝒜_2}(\omega )),$$
where $`\mu _1=\mu _{𝒜_1}=X_{𝒜_1}(P)`$, $`\mu _2=\mu _{𝒜_2}=X_{𝒜_2}(P)`$ are the marginals of $`\mu `$.
I assume in addition that the $`\sigma `$-fields $`𝒜_1,𝒜_2`$ are nonatomic (atoms are left to the reader); then an additional transformation $``$ turns $`\mu _1,\mu _2`$ into Lebesgue measure on $`(0,1)`$.<sup>25</sup><sup>25</sup>25One may use the cumulative distribution function $`F_𝒜(x)=P\left(\{\omega :X_𝒜(\omega )x\}\right)`$; it is continuous (due to the nonatomicity), and the random variable $`\omega F_𝒜(X_𝒜(\omega ))`$ is distributed uniformly on $`(0,1)`$. The density of $`\mu `$ (w.r.t. Lebesgue measure $`\mu _1\mu _2`$ on $`(0,1)\times (0,1)`$) is a Borel function of $`\mu `$ (take an increasing (refining) sequence of finite partitions of $`(0,1)\times (0,1)`$). The following lemma (or rather, its straightforward two-dimensional generalization) completes the proof. ∎
###### 6.13 Lemma.
Let $`(\mathrm{\Omega },,P)`$ be a nonatomic probability space. Consider the set $`𝒰L_0(\mathrm{\Omega },,P)`$ of all random variables $`U:\mathrm{\Omega }`$ distributed uniformly on $`(0,1)`$ (in other words, measure preserving transformations from $`(\mathrm{\Omega },,P)`$ to $`(0,1)`$ with Lebesgue measure). For each $`U𝒰`$ and $`fL_0(0,1)`$ consider the composition $`fU`$ (that is, $`f(U())`$) as an element of $`L_0(\mathrm{\Omega },,P)`$. Then
(a) $`U`$ is a Borel measurable subset of $`L_0(\mathrm{\Omega },,P)`$;
(b) $`fU`$ is jointly Borel measurable in $`fL_0(0,1)`$ and $`U𝒰`$.
###### Proof.
(a) Follows immediately from measurability of $`U(P)`$ in $`U`$.
(b) Let $`Q`$ be another probability measure on $`(\mathrm{\Omega },)`$, equivalent to $`P`$. Consider the distribution
$$(fU)(Q)=f(U(Q));$$
we know that $`f(\mu )`$ is jointly measurable in $`f`$ and $`\mu `$, and $`U(Q)`$ is measurable in $`U`$, therefore $`(fU)(Q)`$ is jointly measurable in $`f`$ and $`U`$. However, distributions $`X(Q)`$ for all $`Q`$ determine $`XL_0(P)`$ uniquely, and moreover, they generate the Borel $`\sigma `$-field on $`L_0(P)`$. ∎
###### 6.14 Proposition.
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space, and $`=\{(𝒜,\psi ):𝒜𝐀(),\psi L_2(\mathrm{\Omega },𝒜,𝒫)\}`$ the disjoint union of Hilbert spaces $`L_2(\mathrm{\Omega },𝒜,𝒫)`$ over all sub-$`\sigma `$-fields $`𝒜`$. Then
(a) $``$ is (naturally) a standard Borel space;
(b) the set $`𝒟_1`$ of all pairs $`((𝒜_1,\psi _1),(𝒜_2,\psi _2))\times `$ satisfying $`𝒜_1=𝒜_2`$ is a Borel subset of $`\times `$; the map $`((𝒜,\psi _1),(𝒜,\psi _2))(𝒜,\psi _1+\psi _2)`$ from $`𝒟_1`$ to $``$ is Borel measurable; the map $`(\alpha ,(𝒜,\psi ))(𝒜,\alpha \psi )`$ from $`\times `$ to $``$ is Borel measurable; and the map $`((𝒜,\psi _1),(𝒜,\psi _2))\psi _1,\psi _2`$ from $`𝒟_1`$ to $``$ is Borel measurable;
(c) the set $`𝒟_2`$ of all pairs $`((𝒜_1,\psi _1),(𝒜_2,\psi _2))\times `$ such that $`𝒜_1𝒜_2`$ is well-defined<sup>26</sup><sup>26</sup>26See Lemma 6.12. is a Borel subset of $`\times `$; and the map $`((𝒜_1,\psi _1),(𝒜_2,\psi _2))(𝒜_1𝒜_2,\psi _1\psi _2)`$ from $`𝒟_2`$ to $``$ is Borel measurable.
###### Proof.
We choose some $`P𝒫`$ and replace each $`L_2(\mathrm{\Omega },𝒜,𝒫)`$ with the corresponding $`L_2(\mathrm{\Omega },𝒜,P)`$ according to their unitary correspondence determined by $`P`$,
$$L_2(\mathrm{\Omega },𝒜,𝒫)\psi \frac{\psi }{\sqrt{P|_𝒜}}L_2(\mathrm{\Omega },𝒜,P).$$
The disjoint union of $`L_2(\mathrm{\Omega },𝒜,P)`$ over all $`𝒜`$ is naturally a standard Borel space, since it is a Borel subset of the disjoint union of all subspaces of $`L_2(\mathrm{\Omega },,P)`$.<sup>27</sup><sup>27</sup>27You see, $`L_2(\mathrm{\Omega },𝒜,𝒫)`$ is not (naturally identified with) a subspace of $`L_2(\mathrm{\Omega },,𝒫)`$. However, $`L_2(\mathrm{\Omega },𝒜,P)`$ is a subspace of $`L_2(\mathrm{\Omega },,P)`$. Thus, all $`L_2(\mathrm{\Omega },𝒜,𝒫)`$ become embedded into $`L_2(\mathrm{\Omega },,𝒫)`$, but the embedding depends on $`P`$. It may be written as $`L_2(\mathrm{\Omega },𝒜,𝒫)\psi {\displaystyle \frac{\psi }{\sqrt{P|_𝒜}}}\sqrt{P}L_2(\mathrm{\Omega },,P)`$. Thus, a Borel structure appears also on the disjoint union of $`L_2(\mathrm{\Omega },𝒜,𝒫)`$, and the Borel structure does not depend on the choice of $`P`$ (since $`{\displaystyle \frac{Q|_𝒜}{P|_𝒜}}`$ is measurable in $`𝒜`$, see Lemma 6.11). Item (b) is easily transferred from the disjoint union of all subspaces of $`L_2(\mathrm{\Omega },,P)`$ to the disjoint union of $`L_2(\mathrm{\Omega },𝒜,𝒫)`$. It remains to prove (c).
Lemma 6.12(a) gives us measurability of the set $`𝒟_2`$, and measurability of $`𝒜_1𝒜_2`$ in $`(𝒜_1,𝒜_2)`$. It remains to verify measurability of $`\psi _1\psi _2`$. We know (recall Sect. 4) that
$$\psi _1\psi _2=\left(\frac{\psi _1}{\sqrt{Q_1}}\frac{\psi _2}{\sqrt{Q_2}}\right)\sqrt{Q_1Q_2}L_2(\mathrm{\Omega },𝒜_1𝒜_2,𝒫);$$
that is, if $`Q𝒫`$ is a measure that makes $`𝒜_1,𝒜_2`$ independent, then
$$\frac{\psi _1\psi _2}{\sqrt{Q|_{𝒜_1𝒜_2}}}=\frac{\psi _1}{\sqrt{Q|_{𝒜_1}}}\frac{\psi _2}{\sqrt{Q|_{𝒜_2}}}\text{for }\psi _1L_2(\mathrm{\Omega },𝒜_1,𝒫),\psi _2L_2(\mathrm{\Omega },𝒜_2,𝒫);$$
the product in the right-hand side is just a pointwise product of two functions.<sup>28</sup><sup>28</sup>28It is, at the same time, their tensor product, since they are independent (w.r.t. $`Q`$). Therefore
$$\frac{\psi _1\psi _2}{\sqrt{P|_{𝒜_1𝒜_2}}}=\sqrt{\frac{Q|_{𝒜_1𝒜_2}}{P|_{𝒜_1𝒜_2}}}\sqrt{\frac{P|_{𝒜_1}}{Q|_{𝒜_1}}}\sqrt{\frac{P|_{𝒜_2}}{Q|_{𝒜_2}}}\frac{\psi _1}{\sqrt{P|_{𝒜_1}}}\frac{\psi _2}{\sqrt{P|_{𝒜_2}}}.$$
However, the independence of $`𝒜_1`$ and $`𝒜_2`$ under $`Q`$ means that
$$\frac{P|_{𝒜_1𝒜_2}}{Q|_{𝒜_1𝒜_2}}=\frac{P|_{𝒜_1𝒜_2}}{Q|_{𝒜_1}Q|_{𝒜_2}}=\frac{P|_{𝒜_1𝒜_2}}{P|_{𝒜_1}P|_{𝒜_2}}\frac{P|_{𝒜_1}}{Q|_{𝒜_1}}\frac{P|_{𝒜_2}}{Q|_{𝒜_2}},$$
so,
$$\frac{\psi _1\psi _2}{\sqrt{P|_{𝒜_1𝒜_2}}}=\sqrt{\frac{P|_{𝒜_1}P|_{𝒜_2}}{P|_{𝒜_1𝒜_2}}}\frac{\psi _1}{\sqrt{P|_{𝒜_1}}}\frac{\psi _2}{\sqrt{P|_{𝒜_2}}}.$$
By Lemma 6.12(b), $`{\displaystyle \frac{P|_{𝒜_1𝒜_2}}{P|_{𝒜_1}P|_{𝒜_2}}}`$ is Borel measurable in $`(𝒜_1,𝒜_2)`$, therefore $`{\displaystyle \frac{\psi _1\psi _2}{\sqrt{P|_{𝒜_1𝒜_2}}}}`$ is measurable in $`(𝒜_1,𝒜_2,{\displaystyle \frac{\psi _1}{\sqrt{P|_{𝒜_1}}}},{\displaystyle \frac{\psi _2}{\sqrt{P|_{𝒜_2}}}})`$, which means that $`\psi _1\psi _2`$ is measurable in $`((𝒜_1,\psi _1),(𝒜_2,\psi _2))`$. ∎
For any Hilbert spaces $`H_1,H_2`$ denote by $`\stackrel{~}{}(H_1,H_2)`$ the set of all isomorphisms between $`H_1`$ and $`H_2`$, that is, linear isometric invertible maps $`H_1H_2`$ (of course, $`\stackrel{~}{}(H_1,H_2)`$ is empty if $`H_1,H_2`$ are of different dimension). For any Hilbert space $`H`$ denote by $`(H)`$ the disjoint union of sets $`\stackrel{~}{}(L_1,L_2)`$ over all subspaces $`L_1,L_2𝐋(H)`$. That is, $`(H)`$ consists of all triples $`(L_1,L_2,U)`$ where $`L_1,L_2H`$ are subspaces and $`U:L_1L_2`$ is an isomorphism. However, we may identify each $`U`$ with its graph, and $`L_1,L_2`$ with projections of the graph, $`L_1(U)=\{x:(x,y)U\}`$, $`L_2(U)=\{y:(x,y)U\}`$. Now $`(H)`$ consists of all subspaces $`U𝐋(HH)`$ such that $`x=y`$ whenever $`xH`$, $`yH`$, $`(x,y)U`$. Clearly, $`(H)`$ is a Borel subset of $`𝐋(HH)`$, therefore, a standard Borel space.
###### 6.15 Lemma.
Let $`H`$ be a Hilbert space. Then
(a) the set $`𝒟`$ of all pairs $`((L_1,L_2,U),x)`$ such that $`(L_1,L_2,U)(H)`$ and $`xL_1`$ is a Borel subset of $`(H)\times H`$;
(b) the map $`((L_1,L_2,U),x)U(x)`$ from $`𝒟`$ to $`H`$ is Borel measurable.
###### Proof.
Treating $`U`$ as a subspace of $`HH`$ we choose measurable maps $`f_n:𝐋(HH)HH`$ such that every $`U`$ is spanned by $`\{f_1(U),f_2(U),\mathrm{}\}`$. We have $`f_n(U)=(g_n(U),h_h(U))`$ where $`g_n(U)H`$, $`h_n(U)H`$. Applying the orthogonalization process we ensure that $`g_n(U)`$ form an orthogonal basis of $`L_1(U)`$. Introducing Borel functions $`c_n(x,U)=x,g_n(U)`$ for $`xH`$ we have
$$(U,x)𝒟\underset{n}{}|c_n(x,U)|^2=x^2,$$
$$(U,x)𝒟U(x)=\underset{n}{}c_n(x,U)h_n(U).$$
For any FHS space $`G`$ we define $`(G)`$ as consisting of all triples $`(L_1,L_2,U)`$ where $`L_1,L_2G`$ are subspaces and $`U:L_1L_2`$ is an FHS-isomorphism. Alternatively, $`(G)`$ may be thought of as a subset of $`𝐋(HH)`$; we’ll see (Lemma 6.18) that it is a Borel subset, therefore, a standard Borel space.
For any measure type space $`(\mathrm{\Omega },,𝒫)`$ we define $`()`$ as consisting of all triples $`(𝒜_1,𝒜_2,U)`$ where $`𝒜_1,𝒜_2𝐀(mod0)`$ and $`U:𝒜_1𝒜_2`$ is an isomorphism of complete Boolean algebras (it means $`mod0`$ isomorphism between quotient spaces $`\mathrm{\Omega }/𝒜_1`$ and $`\mathrm{\Omega }/𝒜_2`$, provided that $`(\mathrm{\Omega },𝒜,𝒫)`$ is a Lebesgue-Rokhlin space). The graph of $`U`$ is a subset of $`(mod0)(mod0)=()mod0`$, where $``$ is the natural $`\sigma `$-field on the union $`\mathrm{\Omega }\mathrm{\Omega }`$ of two disjoint copies of $`\mathrm{\Omega }`$. We identify $`U`$ with its graph; it is not just a subset but a $`\sigma `$-field; so, $`()𝐀()`$.
###### 6.16 Lemma.
$`()`$ is a Borel subset of $`𝐀()`$.
###### Proof.
A sub-$`\sigma `$-field $``$ belongs to $`()`$ if and only if $`\epsilon \delta (A,B)(P(A)<\delta P(B)\epsilon )`$; here an element of $``$ is identified with a pair $`(A,B)`$ of elements of $``$. (The choice of a measure $`P𝒫`$ does not matter.) For such $`\epsilon `$ and $`\delta `$, $``$ must be disjoint to the *open* set $`\{(A,B):P(A)<\delta ,P(B)>\epsilon \}`$. ∎
Let $`(\mathrm{\Omega },,𝒫)`$ be a measure type space. For any closed set $`FL_0(\mathrm{\Omega },,𝒫)`$ denote by $`\sigma (F)`$ the sub-$`\sigma `$-field generated by $`F`$, that is, the least $`\sigma `$-field $`^{}`$ such that $`FL_0(\mathrm{\Omega },^{},𝒫)`$.
###### 6.17 Lemma.
The map $`F\sigma (F)`$ from $`𝐅\left(L_0(\mathrm{\Omega },,𝒫)\right)`$ to $`𝐀(\mathrm{\Omega },,𝒫)`$ is Borel measurable.
###### Proof.
Due to (6.1) it suffices to prove measurability of the map $`f\sigma (f)`$ from $`L_0(\mathrm{\Omega },,𝒫)`$ to $`𝐀(\mathrm{\Omega },,𝒫)`$; of course, $`\sigma (f)`$ means $`\sigma (\{f\})`$. We know (see the proof of Lemma 3.8) that $`f_nf`$ implies $`\sigma (f)lim\; inf\sigma (f_n)`$. Therefore, for any open set $`V𝐀(\mathrm{\Omega },,𝒫)`$ the set of all $`f`$ such that $`\sigma (f)V=\mathrm{}`$ is closed. ∎
###### 6.18 Lemma.
Let $`G`$ be an FHS space, then $`(G)`$ is a Borel subset of $`𝐋(GG)`$.
###### Proof.
We take a Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$<sup>29</sup><sup>29</sup>29Suppressing in the notation the distinction between $`G`$ and $`G/\mathrm{Const}`$. and consider $``$, so that $`GGL_0()`$. Every subspace $`U𝐋(GG)`$ generates a sub-$`\sigma `$-field $`\sigma (U)𝐀()`$. It is easy to see that $`U(G)`$ if and only if $`\sigma (U)()`$. However, $`\sigma (U)`$ is measurable in $`U`$ by Lemma 6.17, and $`()`$ is measurable by Lemma 6.16. ∎
For any measure type space $`(\mathrm{\Omega },,𝒫)`$ we define $`(𝒫)`$<sup>30</sup><sup>30</sup>30Sorry for the clumsy notation: $`()`$ for $`\sigma `$-fields, but $`(𝒫)`$ for square roots of measures. as consisting of all triples $`(𝒜_1,𝒜_2,U)`$ where $`𝒜_1,𝒜_2`$ are sub-$`\sigma `$-fields and $`U\stackrel{~}{}(L_2(\mathrm{\Omega },𝒜_1,𝒫),L_2(\mathrm{\Omega },𝒜_2,𝒫))`$ is a linear isometry. Spaces $`L_2(\mathrm{\Omega },𝒜,𝒫)`$ are not naturally embedded into $`L_2(\mathrm{\Omega },,𝒫)`$; however, we may choose some measure $`P𝒫`$ and embed all $`L_2(\mathrm{\Omega },𝒜,𝒫)`$ into $`L_2(\mathrm{\Omega },,P)`$ by
$$L_2(\mathrm{\Omega },𝒜,𝒫)\psi \frac{\psi }{\sqrt{P|_𝒜}}L_2(\mathrm{\Omega },𝒜,P)L_2(\mathrm{\Omega },,P).$$
We have a bijective correspondence between $`(𝒫)`$ and $`\left(L_2(\mathrm{\Omega },,P)\right)`$, which turns $`(𝒫)`$ into a standard Borel space. Its Borel structure does not depend on the choice of $`P𝒫`$ (recall Lemma 6.11), but the correspondence depends on $`P`$.
Take an element $`(𝒜_1,𝒜_2,U)`$ of $`()`$ (this time we prefer a triple to a graph). The isomorphism $`U`$ between $`\sigma `$-fields $`𝒜_1,𝒜_2`$ induces naturally an isomorphism (linear isometry) between Hilbert spaces $`L_2(\mathrm{\Omega },𝒜_1,𝒫)`$ and $`L_2(\mathrm{\Omega },𝒜_2,𝒫)`$. Namely, if measures $`P_1𝒫|_{𝒜_1}`$, $`P_2𝒫|_{𝒜_2}`$ satisfy $`P_2(U(A))=P_1(A)`$ for all $`A𝒜_1`$, then the vector $`\psi _2=\sqrt{P_2}L_2(\mathrm{\Omega },𝒜_2,𝒫)`$ corresponds to the vector $`\psi _1=\sqrt{P_1}L_2(\mathrm{\Omega },𝒜_1,𝒫)`$. (Such vectors are not a linear set, but span the Hilbert spaces.) So, we have a map from $`()`$ to $`(𝒫)`$.
###### 6.19 Lemma.
The map from $`()`$ to $`(𝒫)`$ is Borel measurable.
###### Proof.
Consider some $`U()`$ treated as a sub-$`\sigma `$-field of $``$. Given some $`P_1𝒫`$, we introduce on $`U`$ a measure $`P(A,B)=P_1(A)`$ for $`(A,B)U`$; here, as before, an element of $``$ is represented by a pair $`(A,B)`$ where $`A,B`$. There is also a measure $`P_2`$ on the $`\sigma `$-field $`𝒜_2=\{B:(A,B)U\}`$ such that $`P_2(B)=P(A,B)=P_1(A)`$ whenever $`(A,B)U`$. The pair $`(\sqrt{P_1},\sqrt{P_2})`$ belongs to the graph $`U_1(𝒫)`$ that corresponds to $`U`$. Such pairs for all $`P_1𝒫`$ (or for a countable dense subset) span the graph $`U_1`$. It remains to prove that, for a given $`P_1`$ and arbitrary $`U`$, the pair $`(\sqrt{P_1},\sqrt{P_2})`$ is measurable in $`U`$ (you see, $`P_2`$ depends implicitly on $`U`$). According to our definition of the Borel structure on $`(𝒫)`$, we have to prove measurability in $`U`$ of the density $`{\displaystyle \frac{P_2}{P_1|_{𝒜_2}}}`$. The latter is the restriction (to the second copy of $`\mathrm{\Omega }`$) of a density on the doubled space, $`\mathrm{\Omega }\mathrm{\Omega }`$, namely, $`{\displaystyle \frac{P^{}|_U}{P^{\prime \prime }|_U}}`$, where measures $`P^{},P^{\prime \prime }`$ on $``$ are defined (irrespective of $`U`$) by $`P^{}(A,B)=P_1(A)`$, $`P^{\prime \prime }(A,B)=P_1(B)`$. We apply Lemma 6.11 to $`P^{},P^{\prime \prime }`$ on $`\mathrm{\Omega }\mathrm{\Omega }`$. Though, $`P^{}`$ and $`P^{\prime \prime }`$ are not equivalent, but one can consider, say, $`{\displaystyle \frac{2P^{}+P^{\prime \prime }}{P^{}+P^{\prime \prime }}}`$. ∎
###### 6.20 Proposition.
Let $`(\mathrm{\Omega },,𝒫,G)`$ be a Gaussian type space, and $`G_0=G/\mathrm{Const}`$ the corresponding FHS space. Then the natural map from $`(G_0)`$ to $`(𝒫)`$ is Borel measurable.
###### Proof.
We know that every $`U(G)`$ generates $`\sigma (U)()`$ (see the proof of Lemma 6.18), and the map $`U\sigma (U)`$ is measurable. The transition from $`\sigma (U)`$ to the corresponding element of $`(𝒫)`$ is measurable by Lemma 6.19. ∎
## 7 Sum systems and product systems
###### 7.1 Definition.
A *sum system* consists of a two-parameter family $`(G_{a,b})`$ of FHS-spaces $`G_{a,b}`$, given for $`\mathrm{}<a<b<+\mathrm{}`$, embedded into a single linear space $`G_{\mathrm{},+\mathrm{}}`$, and a one-parameter group $`(U_t)`$ of linear maps $`U_t:G_{\mathrm{},+\mathrm{}}G_{\mathrm{},+\mathrm{}}`$ for $`t`$, such that
(a) $`G_{a,c}=G_{a,b}G_{b,c}`$ (in the FHS sense) whenever $`\mathrm{}<a<b<c<+\mathrm{}`$;<sup>31</sup><sup>31</sup>31Thus, $`H_{a,b}`$ and $`H_{b,c}`$ must be linear subspaces of $`H_{a,c}`$; and their FHS structures must be inherited from $`H_{a,c}`$; and they must be orthogonal in some admissible norm. Note that the norm may depend on $`b`$.
(b) $`U_t:G_{a,b}G_{a+t,b+t}`$ is an isomorphism of FHS spaces, whenever $`\mathrm{}<a<b<+\mathrm{}`$ and $`\mathrm{}<t<+\mathrm{}`$.<sup>32</sup><sup>32</sup>32Thus, $`U_t`$ must map $`G_{a,b}`$ onto $`G_{a+t,b+t}`$ and send an admissible norm into an admissible norm.
(c) $`(U_t)`$ is strongly continuous in the sense that $`U_txx0`$ when $`t0`$, whenever $`a<b<c<d`$, $`xG_{b,c}`$, and the norm is taken in $`G_{a,b}`$ (which is correct for $`t`$ small enough).
The structure is local in the sense that the global space $`G_{\mathrm{},+\mathrm{}}`$ is equipped with a linear structure only, not an FHS structure, nor even a topology. One may assume that $`G_{\mathrm{},+\mathrm{}}`$ is just the union of all $`G_{a,b}`$, since anyway, only the local spaces $`G_{a,b}`$ will be used.
Given a sum system $`((G_{a,b}),(U_t))`$, we may introduce (as explained in Sect. 5) Hilbert spaces $`H_{a,b}=\mathrm{Exp}(G_{a,b})`$ satisfying (under the usual identification)
$$H_{a,b}H_{b,c}=H_{a,c}$$
whenever $`\mathrm{}<a<b<c<+\mathrm{}`$. Given $`a<b<c<d`$, $`\psi _1H_{a,b}`$, $`\psi _2H_{b,c}`$, $`\psi _3H_{c,d}`$, we may calculate $`\psi _1\psi _2\psi _3`$ as $`(\psi _1\psi _2)\psi _3`$ or $`\psi _1(\psi _2\psi _3)`$, which is the same (recall Lemma 3.5). The two-parameter families may be reduced to one-parameter families using $`(U_t)`$; namely, for $`s,t(0,\mathrm{})`$,
(7.2) $`G_{0,s+t}`$ $`=G_{0,s}U_sG_{0,t},`$
$`H_{0,s+t}`$ $`=H_{0,s}\left(\mathrm{Exp}(U_s|_{G_{0,t}})\right)H_{0,t},`$
where $`\left(\mathrm{Exp}(U_s|_{G_{0,t}})\right):H_{0,t}H_{s,s+t}`$ is the unitary operator corresponding to the FHS isomorphism $`U_s|_{H_{0,t}}:H_{0,t}H_{s,s+t}`$. The binary operation of tensor product, $`(s,\psi _1),(t,\psi _2)(s+t,\psi _1(\mathrm{Exp}(U_s|_{G_{0,t}}))\psi _2)`$ (for $`\psi _1H_{0,s}`$, $`\psi _2H_{0,t}`$) is associative. That is the algebraic part of the ‘product system’ structure defined by W. Arveson \[4, Def. 1.4\]. It is not the whole story, since some measurability in $`s,t`$ is needed. Namely, the disjoint union of spaces $`H_{0,s}`$ must be a standard Borel space, and the tensor product must be a measurable binary operation.
The disjoint union is the set of all pairs $`(s,\psi )`$ such that $`s(0,\mathrm{})`$ and $`\psi H_{0,s}`$. We take some $`T(0,\mathrm{})`$; it is enough to consider $`s(0,T)`$ rather than $`(0,\mathrm{})`$. We have a Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$ such that $`G/\mathrm{Const}=G_{0,T}`$, and sub-$`\sigma `$-fields $`_{a,b}`$ such that $`L_2(\mathrm{\Omega },_{a,b},𝒫)=H_{a,b}`$ whenever $`(a,b)(0,T)`$.
###### 7.3 Lemma.
The map $`(a,b)G_{a,b}`$ from the triangle $`\{(a,b):0a<bT\}`$ to the Borel space $`𝐋(G_{0,T})`$ of subspaces of $`G_{0,T}`$ is Borel measurable.
###### Proof.
It is enough to consider the case $`0a<t<bT`$ for an arbitrary $`t(0,T)`$. The equality $`G_{a,b}=G_{a,t}G_{t,b}`$, in combination with Lemma 6.5, reduces the problem to measurability of $`G_{a,t}`$ in $`a`$, and $`G_{t,b}`$ in $`b`$. However, a *monotone* function is always measurable. ∎
So, $`G_{a,b}`$ is measurable in $`(a,b)`$. It follows by Lemma 6.17 that $`_{a,b}`$ is measurable in $`(a,b)`$. Proposition 6.14 gives us a Borel structure on the disjoint union of spaces $`H_{a,b}`$ (over all $`a,b`$ satisfying $`0a<bT`$); it is compatible with linear operations and scalar product; and the map
$$(((a,b),\psi _1),((b,c),\psi _2))((a,c),\psi _1\psi _2)$$
is Borel measurable (here $`\psi _1H_{a,b}`$, $`\psi _2H_{b,c}`$, $`0a<b<cT`$).
Given $`a,b`$ such that $`0a<bT`$, we may treat the restriction $`U_{ba}|_{G_{0,a}}`$ as an FHS isomorphism $`G_{0,a}G_{ba,b}`$, therefore an element of the Borel space $`(G_{0,T})`$ introduced in Sect. 6.
###### 7.4 Lemma.
The map $`(a,b)U_{ba}|_{G_{0,a}}`$ from the triangle $`\{(a,b):0a<bT\}`$ to $`(G_{0,T})`$ is Borel measurable.
###### Proof.
By Lemma 7.3, $`G_{0,a}`$ is measurable in $`a`$. According to (6.1) there are $`f_n(a)G_{0,a}`$, measurable in $`a`$, such that every $`G_{0,a}`$ is spanned by $`\{f_1(a),f_2(a),\mathrm{}\}`$. Therefore the graph of $`U_{ba}|_{G_{0,a}}`$ is spanned by pairs $`(f_n(a),U_{ba}f_n(a))`$. Each pair is measurable in $`a`$ and continuous in $`b`$ (recall 7.1(c)), therefore, measurable in $`(a,b)`$ (see \[11, Th. 3.1.30\]). ∎
So, $`U_{ba}|_{G_{0,a}}`$ is measurable in $`(a,b)`$. It follows by Proposition 6.20 that $`\mathrm{Exp}(U_{ba}|_{G_{0,a}})`$ is also measurable in $`(a,b)`$. Lemma 6.16 shows that $`\mathrm{Exp}(U_{ba}|_{G_{0,a}})\psi `$ is jointly measurable in $`\psi H_{0,a}`$ and $`a,b`$. It follows that $`\psi _1\left(\mathrm{Exp}(U_{ba}|_{G_{0,a}})\right)\psi _2`$ is jointly measurable in $`a`$, $`b`$, $`\psi _1H_{ba}`$, $`\psi _2H_a`$, which proves the following result.
###### 7.5 Theorem.
If $`((G_{a,b}),(U_t))`$ is a sum system, then Hilbert spaces $`H_{a,b}=\mathrm{Exp}(G_{a,b})`$ with the natural identification $`H_{0,s+t}=H_{0,s}\left(\mathrm{Exp}(U_s|_{G_{0,t}})\right)H_{0,t}`$ form a product system.
The product system may be called the exponential of the given sum system.
## 8 The invariant
An isomorphism of two product systems is defined \[4, p. 6\] as a family $`(V_t)`$ of linear isomorphisms $`V_t:H_{0,t}^{}H_{0,t}^{\prime \prime }`$ between corresponding Hilbert spaces that respects the two structures on the disjoint union of the Hilbert spaces, namely, the binary operation of tensor multiplication, and the Borel $`\sigma `$-field. Assuming that the two product systems are exponentials of two given sum systems $`((G_{a,b}^{}),(U_t^{}))`$ and $`((G_{a,b}^{\prime \prime }),(U_t^{\prime \prime }))`$, we may redefine equivalently an isomorphism as a two-parameter family $`(V_{a,b})`$ of unitary operators that satisfy
$$V_{a,b}:H_{a,b}^{}H_{a,b}^{\prime \prime }\text{ unitarily,}$$
$$V_{a,c}=V_{a,b}V_{b,c},$$
$$\mathrm{Exp}(U_t^{\prime \prime }|_{G_{a,b}^{\prime \prime }})V_{a,b}=V_{a+t,b+t}\mathrm{Exp}(U_t|_{G_{a,b}^{}}),$$
whenever $`\mathrm{}<a<b<c<+\mathrm{}`$ and $`\mathrm{}<t<+\mathrm{}`$ (as before, $`H_{a,b}^{}=\mathrm{Exp}(G_{a,b}^{})`$, $`H_{a,b}^{\prime \prime }=\mathrm{Exp}(G_{a,b}^{\prime \prime })`$), and respects the Borel structure on the disjoint union of Hilbert spaces.
For now $`H_E`$, as well as $`G_E`$ and $`_E`$, are defined only when $`E`$ is an interval, $`E=(a,b)`$. However, they may be defined for any elementary set $`E`$, that is, a union of a finite number of intervals. Given $`\mathrm{}<a<b<c<d<+\mathrm{}`$, we have
$$\underset{H_{(a,d)}}{\underset{}{H_{a,d}}}=H_{a,b}H_{b,c}H_{c,d}=\underset{H_{(a,b)(c,d)}}{\underset{}{(H_{a,b}H_{c,d})}}\underset{H_{(b,c)}}{\underset{}{H_{b,c}}}.$$
The same for any finite number of intervals. We get $`H_{E_1E_2}=H_{E_1}H_{E_2}`$ when $`E_1E_2=\mathrm{}`$. Dealing with elementary sets we neglect boundary points, treating, say, $`(a,b)(b,c)`$ as $`(a,c)`$. Also, $`H_E=H_{E_1}\mathrm{}H_{E_n}`$ whenever $`E_1,\mathrm{},E_n`$ are pairwise disjoint and $`E=E_1\mathrm{}E_n`$. Similarly, $`G_E=G_{E_1}\mathrm{}G_{E_n}`$ (in the FHS sense), and $`_E=_{E_1}\mathrm{}_{E_n}`$ (recall 3.23.5).
###### 8.1 Proposition.
Let two sum systems $`((G_{a,b}^{}),(U_t^{}))`$ and $`((G_{a,b}^{\prime \prime }),(U_t^{\prime \prime }))`$ be such that the corresponding product systems are isomorphic. Let $`E_1,E_2,\mathrm{}(0,1)`$ be elementary sets. Then the following two conditions are equivalent.
(a) $`lim\; infG_{(0,1)E_n}^{}=G_{(0,1)}^{}`$ and $`lim\; supG_{E_n}^{}=\{0\}`$;
(b) $`lim\; infG_{(0,1)E_n}^{\prime \prime }=G_{(0,1)}^{\prime \prime }`$ and $`lim\; supG_{E_n}^{\prime \prime }=\{0\}`$.
(The FHS spaces are treated as subspaces of $`G_{(0,1)}^{}`$, $`G_{(0,1)}^{\prime \prime }`$ respectively.)
###### Proof.
Theorem 5.3 allows us to reformulate the conditions in terms of density matrices in product systems, thus making explicit their invariance under isomorphisms. ∎
## 9 Slightly coloured noises
Consider a scalar product of the form
(9.1)
$$f,g=f(s)g(t)B(st)𝑑s𝑑t$$
assuming that $`B:\{0\}`$ is continuous outside of the origin, and $`B(t)=B(t)`$ for all $`t(0,\mathrm{})`$, and $`_0^1|B(t)|𝑑t<\mathrm{}`$. The scalar product is well-defined whenever $`f,gL_2(M,M)`$, $`M(0,\mathrm{})`$. We assume that $`B`$ is positively definite in the sense that $`f,f0`$ for all such $`f`$.
Denote by $`G_{a,b}`$ the completion of $`L_2(a,b)`$ w.r.t. the scalar product (9.1), then $`G_{a,b}`$ is a Hilbert space. We introduce operators $`U_t`$ by $`(U_tf)(s)=f(st)`$ for $`fL_2(a,b)`$ and extend $`U_t`$ by continuity to any $`G_{a,b}`$; thus, $`U_t:G_{a,b}G_{a+t,b+t}`$ is a unitary operator, and $`U_sU_t=U_{s+t}`$, and $`U_tff_{G_{a,d}}0`$ for $`t0`$, if $`fG_{b,c}`$ and $`a<b<c<d`$ (since it holds for the dense subset of continuous functions $`f`$).
In order to get a sum system (as defined by 7.1) we need to ensure that $`G_{a,c}=G_{a,b}G_{b,c}`$ in the FHS sense. The property will be verified for $`B`$ such that
(9.2)
$$\begin{array}{c}\epsilon >0t(0,\epsilon )B(t)=\frac{1}{t\mathrm{ln}^\alpha (1/t)};\\ \text{on }(0,\mathrm{})\text{ the function }B()\text{ is positive, decreasing and convex;}\end{array}$$
here $`\alpha (1,\mathrm{})`$ is a parameter. Such $`B`$ is positively definite, since it is an integral combination (with positive weights) of ‘triangle’ functions of the form $`t\mathrm{max}(0,a|t|)`$, and maybe a positive constant function.
We consider $`G_{T,0}`$ and $`G_{0,T}`$ for an arbitrary $`T(0,\mathrm{})`$. To this end we introduce $`X_kG_{0,T}`$ and $`Y_kG_{T,0}`$ by
(9.3)
$$\begin{array}{cc}\hfill X_k(t)& =\mathrm{𝟏}_{(0,T)}(t)\mathrm{exp}(2\pi ikt/T),\hfill \\ \hfill Y_k(t)& =\mathrm{𝟏}_{(T,0)}(t)\mathrm{exp}(2\pi ikt/T)\hfill \end{array}$$
for $`k`$; of course, $`\mathrm{𝟏}_{(a,b)}`$ is the indicator of $`(a,b)`$. Clearly, $`X_k`$ span $`G_{0,T}`$, and $`Y_k`$ span $`G_{T,0}`$. An elementary calculation gives
(9.4)
$$\begin{array}{c}X_k,X_k=Y_k,Y_k=2_0^T(Tt)B(t)\mathrm{cos}(2\pi kt/T)𝑑t,\\ X_k,X_l=Y_k,Y_l=\frac{T}{\pi (kl)}_0^TB(t)\left(\mathrm{sin}(2\pi kt/T)\mathrm{sin}(2\pi lt/T)\right)𝑑t,\\ X_k,Y_k=_0^{2T}\mathrm{min}(t,2Tt)B(t)\mathrm{exp}(2\pi ikt/T)𝑑t,\\ X_k,Y_l=i\frac{T}{2\pi (kl)}_0^{2T}B(t)\mathrm{sgn}(Tt)\left(\mathrm{exp}(2\pi ikt/T)\mathrm{exp}(2\pi ilt/T)\right)𝑑t\end{array}$$
for $`kl`$. We want to estimate $`X_k,X_l`$ and $`X_k,Y_l`$ from above. These are increments of Fourier transforms, thus we want to differentiate these Fourier transforms. The singularity of $`B`$ at the origin contributes a term that decays slowly (near $`\mathrm{}`$) and is monotone. Jumps outside the origin (at $`T`$ and $`2T`$) contribute terms that decay much faster, but oscillate. After differentiation, these oscillating terms dominate the monotone term. However, we need Fourier transforms only on the lattice $`(2\pi /T)`$, thus we have a freedom to change the given functions without changing their Fourier transforms on the lattice. We’ll use the freedom for eliminating the jumps.
Note that $`e^{i\lambda t}(U_sf)(t)𝑑t=e^{i\lambda s}e^{i\lambda t}f(t)𝑑t`$, therefore $`e^{i\lambda t}(U_Tff)(t)𝑑t=0`$ for $`\lambda (2\pi /T)`$. We use a piecewise linear $`f`$ for correcting $`B`$; namely, we define
$$\begin{array}{cc}\hfill b_1(t)& =\{\begin{array}{cc}B(t)B(T)\frac{Tt}{T}\hfill & \text{for }0<tT,\hfill \\ B(T)\frac{2Tt}{T}\hfill & \text{for }Tt2T,\hfill \\ 0\hfill & \text{for other }t,\hfill \end{array}\hfill \\ \hfill \widehat{b}_1(\lambda )& =_0^{\mathrm{}}e^{i\lambda t}b_1(t)𝑑t,\hfill \end{array}$$
then $`b_1`$ is continuous on $`(0,\mathrm{})`$, and
$$X_k,X_l=Y_k,Y_l=\frac{T}{\pi (kl)}\mathrm{Im}\left(\widehat{b}_1(2\pi k/T)\widehat{b}_1(2\pi l/T)\right)$$
for $`kl`$. Similarly,
$$\begin{array}{cc}\hfill b_2(t)& =\{\begin{array}{cc}B(t)B(t+T)(B(T)B(2T))\frac{Tt}{T}\hfill & \text{for }0<tT,\hfill \\ (B(T)B(2T))\frac{2Tt}{T}\hfill & \text{for }Tt2T,\hfill \\ 0\hfill & \text{for other }t,\hfill \end{array}\hfill \\ \hfill \widehat{b}_2(\lambda )& =_0^{\mathrm{}}e^{i\lambda t}b_2(t)𝑑t;\hfill \\ \hfill X_k,Y_l& =i\frac{T}{2\pi (kl)}\left(\widehat{b}_2(2\pi k/T)\widehat{b}_2(2\pi l/T)\right)\text{for }kl.\hfill \end{array}$$
It is easy to check that both $`b_1`$ and $`b_2`$ satisfies the conditions of the following lemma, provided that $`B`$ satisfies (9.2).<sup>33</sup><sup>33</sup>33Finite variation of $`(tB(t))^{}`$ on any $`[\epsilon ,1/\epsilon ]`$ follows from increase of $`(tB(t))^{}2B(t)`$.
###### 9.5 Lemma.
Assume that $`\alpha (1,\mathrm{})`$, and a function $`b:(0,\mathrm{})`$ is continuous, and the difference $`b(t){\displaystyle \frac{\mathrm{𝟏}_{(0,\epsilon )}(t)}{t\mathrm{ln}^\alpha (1/t)}}`$ is of finite variation on $`(0,\mathrm{})`$ for some (therefore, every) $`\epsilon (0,1)`$, and $`b(t)=0`$ for all $`t`$ large enough. Assume also that the function $`ttb(t)`$ is absolutely continuous on $`(0,\mathrm{})`$, and the difference $`\left(tb(t)\right)^{}\alpha {\displaystyle \frac{\mathrm{𝟏}_{(0,\epsilon )}(t)}{t\mathrm{ln}^{\alpha +1}(1/t)}}`$ is of finite variation on $`(0,\mathrm{})`$ for some (therefore, every) $`\epsilon (0,1)`$. Then the function $`\widehat{b}(\lambda )=_0^{\mathrm{}}e^{i\lambda t}b(t)𝑑t`$ satisfies
$$\begin{array}{cc}\hfill \widehat{b}(\lambda )& =\frac{1}{\alpha 1}\frac{1}{\mathrm{ln}^{\alpha 1}\lambda }+O\left(\frac{1}{\mathrm{ln}^\alpha \lambda }\right)\text{for }\lambda +\mathrm{},\hfill \\ \hfill \frac{d}{d\lambda }\widehat{b}(\lambda )& =\frac{1}{\lambda \mathrm{ln}^\alpha \lambda }+O\left(\frac{1}{\lambda \mathrm{ln}^{\alpha +1}\lambda }\right)\text{for }\lambda +\mathrm{}.\hfill \end{array}$$
###### Proof.
Choosing any $`\epsilon (0,1)`$ we have for large $`\lambda `$
$$\begin{array}{c}\widehat{b}(\lambda )=\underset{O(1/\lambda )}{\underset{}{_0^{\mathrm{}}e^{i\lambda t}\left(b(t)\frac{\mathrm{𝟏}_{(0,\epsilon )}(t)}{t\mathrm{ln}^\alpha (1/t)}\right)𝑑t}}+_0^\epsilon e^{i\lambda t}\frac{1}{t\mathrm{ln}^\alpha (1/t)}𝑑t=\hfill \\ \hfill =\left(_0^{1/\lambda }+_{1/\lambda }^\epsilon \right)e^{i\lambda t}\frac{1}{t\mathrm{ln}^\alpha (1/t)}dt+O(1/\lambda )=\\ \hfill =_0^{1/\lambda }\frac{1}{t\mathrm{ln}^\alpha (1/t)}𝑑t+_0^{1/\lambda }\frac{e^{i\lambda t}1}{t\mathrm{ln}^\alpha (1/t)}𝑑t+_{1/\lambda }^\epsilon e^{i\lambda t}\frac{1}{t\mathrm{ln}^\alpha (1/t)}𝑑t+O(1/\lambda );\end{array}$$
$$\begin{array}{cc}& _0^{1/\lambda }\frac{1}{\mathrm{ln}^\alpha (1/t)}\frac{dt}{t}=\frac{1}{\alpha 1}\frac{1}{\mathrm{ln}^{\alpha 1}\lambda };\hfill \\ & \left|_0^{1/\lambda }\frac{e^{i\lambda t}1}{t\mathrm{ln}^\alpha (1/t)}𝑑t\right|_0^{1/\lambda }\frac{\lambda t}{t\mathrm{ln}^\alpha (1/t)}𝑑t=\lambda _0^{1/\lambda }\frac{dt}{\mathrm{ln}^\alpha (1/t)}\frac{1}{\mathrm{ln}^\alpha \lambda };\hfill \\ & _{1/\lambda }^\epsilon \frac{e^{i\lambda t}}{t}\frac{1}{\mathrm{ln}^\alpha (1/t)}𝑑t=_{1/\lambda }^\epsilon \frac{1}{\mathrm{ln}^\alpha (1/t)}d\left(\mathrm{ci}(\lambda t)+i\mathrm{si}(\lambda t)\right),\hfill \end{array}$$
where $`\mathrm{ci}(t)={\displaystyle _t^{\mathrm{}}}{\displaystyle \frac{\mathrm{cos}u}{u}}𝑑u`$, $`\mathrm{si}(t)={\displaystyle _t^{\mathrm{}}}{\displaystyle \frac{\mathrm{sin}u}{u}}𝑑u`$. Taking into account that $`\mathrm{ci}(t)=O(1/t)`$ and $`\mathrm{si}(t)=O(1/t)`$, we get
$$\begin{array}{c}_{1/\lambda }^\epsilon e^{i\lambda t}\frac{1}{t\mathrm{ln}^\alpha (1/t)}𝑑t=\underset{O(1/\mathrm{ln}^\alpha \lambda )}{\underset{}{\frac{\mathrm{ci}(\lambda t)+i\mathrm{si}(\lambda t)}{\mathrm{ln}^\alpha (1/t)}|_{1/\lambda }^\epsilon }}\hfill \\ \hfill \underset{O(_{1/\lambda }^\epsilon \frac{1}{\lambda t}\frac{dt}{t\mathrm{ln}^{\alpha +1}(1/t)})}{\underset{}{_{1/\lambda }^\epsilon \left(\mathrm{ci}(\lambda t)+i\mathrm{si}(\lambda t)\right)\frac{\alpha }{t\mathrm{ln}^{\alpha +1}(1/t)}𝑑t}};\end{array}$$
$$\begin{array}{c}\frac{1}{\lambda }_{1/\lambda }^\epsilon \frac{dt}{t^2\mathrm{ln}^{\alpha +1}(1/t)}=\frac{1}{\lambda }\left(_{1/\lambda }^{1/\sqrt{\lambda }}+_{1/\sqrt{\lambda }}^\epsilon \right)\frac{dt}{t^2\mathrm{ln}^{\alpha +1}(1/t)}\hfill \\ \hfill \frac{1}{\lambda }\frac{1}{\mathrm{ln}^{\alpha +1}\sqrt{\lambda }}\underset{\lambda }{\underset{}{_{1/\lambda }^{\mathrm{}}\frac{dt}{t^2}}}+\frac{1}{\lambda }\frac{1}{\mathrm{ln}^{\alpha +1}(1/\epsilon )}\underset{\sqrt{\lambda }}{\underset{}{_{1/\sqrt{\lambda }}^{\mathrm{}}\frac{dt}{t^2}}}=O\left(\frac{1}{\mathrm{ln}^{\alpha +1}\lambda }\right).\end{array}$$
So,
$$\widehat{b}(\lambda )=\frac{1}{\alpha 1}\frac{1}{\mathrm{ln}^{\alpha 1}\lambda }+O\left(\frac{1}{\mathrm{ln}^\alpha \lambda }\right)\text{for }\lambda +\mathrm{},$$
which is the first claim of the lemma. In order to prove the second claim we note that the only properties of the function $`b(t)`$ used till now are the finite variation of $`b(t)\frac{\mathrm{𝟏}_{(0,\epsilon )}(t)}{t\mathrm{ln}^\alpha (1/t)}`$, and $`b(t)=0`$ for large $`t`$. Therefore the same argument may be applied to the function $`\frac{1}{\alpha }\left(tb(t)\right)^{}`$ w.r.t. $`\alpha +1`$:
$$_0^{\mathrm{}}e^{i\lambda t}\frac{1}{\alpha }\left(tb(t)\right)^{}𝑑t=\frac{1}{\alpha }\frac{1}{\mathrm{ln}^\alpha \lambda }+O\left(\frac{1}{\mathrm{ln}^{\alpha +1}\lambda }\right).$$
Hence
$$\begin{array}{c}\frac{d}{d\lambda }\widehat{b}(\lambda )=_0^{\mathrm{}}e^{i\lambda t}itb(t)𝑑t=\frac{1}{\lambda }_0^{\mathrm{}}tb(t)(e^{i\lambda t})^{}𝑑t=\hfill \\ \hfill =\frac{1}{\lambda }_0^{\mathrm{}}\left(tb(t)\right)^{}e^{i\lambda t}𝑑t=\frac{1}{\lambda \mathrm{ln}^\alpha \lambda }+O\left(\frac{1}{\lambda \mathrm{ln}^{\alpha +1}\lambda }\right).\end{array}$$
###### 9.6 Lemma.
Let $`B`$ satisfy (9.2) and $`X_k,Y_k`$ be defined by (9.3); then
$$\underset{m,n:mn}{}\frac{|X_m,X_n|^2}{X_m^2X_n^2}<\mathrm{},\underset{m,n}{}\frac{|X_m,Y_n|^2}{X_m^2Y_n^2}<\mathrm{}.$$
###### Proof.
First, the function $`tB(t)`$ on $`[0,T]`$ is of finite variation, thus, using (9.4) and Lemma 9.5,
$$\begin{array}{c}X_k,X_k=2T\underset{\widehat{b}_1(2\pi k/T)}{\underset{}{_0^TB(t)\mathrm{cos}(2\pi kt/T)𝑑t}}2\underset{O(1/|k|)}{\underset{}{_0^TtB(t)\mathrm{cos}(2\pi kt/T)𝑑t}}\hfill \\ \hfill 2T\frac{1}{\alpha 1}\frac{1}{\mathrm{ln}^{\alpha 1}(2\pi |k|/T)}\frac{2T}{\alpha 1}\frac{1}{\mathrm{ln}^{\alpha 1}|k|};\end{array}$$
$$\frac{1}{X_k^2}=O\left(\mathrm{ln}^{\alpha 1}|k|\right)$$
for $`k\pm \mathrm{}`$.
Second, $`X_k,Y_k=_0^{2T}\mathrm{min}(t,2Tt)B(t)\mathrm{exp}(2\pi ikt/T)=O(1/|k|)`$, since $`\mathrm{min}(t,2Tt)B(t)`$ is of finite variation on $`[0,2T]`$. Hence $`|X_n,Y_n|^2/(X_n^2Y_n^2)=O\left(\frac{1}{n^2}\mathrm{ln}^{2\alpha 2}|n|\right)`$ and
$$\underset{n}{}\frac{|X_n,Y_n|^2}{X_n^2Y_n^2}<\mathrm{}.$$
Third, $`|X_n,X_n|=O(1/|n|)`$ and $`|X_n,Y_n|=O(1/|n|)`$, hence
$$\underset{n}{}\frac{|X_n,X_n|^2}{X_n^2X_n^2}<\mathrm{},\underset{n}{}\frac{|X_n,Y_n|^2}{X_n^2Y_n^2}<\mathrm{}.$$
It is enough to prove that<sup>34</sup><sup>34</sup>34You see, $`\mathrm{ln}|n|`$ is replaced with $`\mathrm{ln}(|n|+2)`$ in order to cover the small values, $`n=1,0,1`$.
$$\underset{m,n:m\pm n0}{}\frac{\mathrm{ln}^{\alpha 1}(|m|+2)\mathrm{ln}^{\alpha 1}(|n|+2)}{(mn)^2}|\widehat{b}(2\pi m/T)\widehat{b}(2\pi n/T)|^2<\mathrm{}$$
for every function $`b`$ as in Lemma 9.5. Taking into account that $`\widehat{b}(\lambda )=\overline{\widehat{b}(\lambda )}`$ we transform it into
$$\begin{array}{c}\underset{m,n:\mathrm{\hspace{0.17em}0}m<n}{}\mathrm{ln}^{\alpha 1}(m+2)\mathrm{ln}^{\alpha 1}(n+2)(\frac{|\widehat{b}(2\pi m/T)\widehat{b}(2\pi n/T)|^2}{(nm)^2}+\hfill \\ \hfill +\frac{|\overline{\widehat{b}(2\pi m/T)}\widehat{b}(2\pi n/T)|^2}{(n+m)^2})<\mathrm{};\end{array}$$
$$\begin{array}{c}\underset{m,n:\mathrm{\hspace{0.17em}0}m<n}{}\mathrm{ln}^{\alpha 1}(m+2)\mathrm{ln}^{\alpha 1}(n+2)\hfill \\ \hfill ((\mathrm{Re}\widehat{b}(2\pi m/T)\mathrm{Re}\widehat{b}(2\pi n/T))^2(\frac{1}{(nm)^2}+\frac{1}{(n+m)^2})+\\ \hfill +\left(\mathrm{Im}\widehat{b}(2\pi m/T)\mathrm{Im}\widehat{b}(2\pi n/T)\right)^2\frac{1}{(nm)^2}+\\ \hfill +(\mathrm{Im}\widehat{b}(2\pi m/T)+\mathrm{Im}\widehat{b}(2\pi n/T))^2\frac{1}{(n+m)^2})<\mathrm{};\end{array}$$
it is enough to prove that
(9.7) $`{\displaystyle \underset{m,n:\mathrm{\hspace{0.17em}0}m<n}{}}{\displaystyle \frac{\mathrm{ln}^{\alpha 1}(m+2)\mathrm{ln}^{\alpha 1}(n+2)}{(nm)^2}}|\widehat{b}(2\pi m/T)\widehat{b}(2\pi n/T)|^2<\mathrm{},`$
(9.8) $`{\displaystyle \underset{m,n:\mathrm{\hspace{0.17em}0}m<n}{}}{\displaystyle \frac{\mathrm{ln}^{\alpha 1}(m+2)\mathrm{ln}^{\alpha 1}(n+2)}{(n+m)^2}}\left(\mathrm{Im}\widehat{b}(2\pi m/T)+\mathrm{Im}\widehat{b}(2\pi n/T)\right)^2<\mathrm{}.`$
We treat separately two cases, $`0m<\sqrt{n}`$ and $`\sqrt{n}m<n`$. The first case, $`0m<\sqrt{n}`$, is simple; just using boundedness of $`\widehat{b}`$ we have for (9.7) and (9.8) as well,
$$\begin{array}{c}\underset{m,n:\mathrm{\hspace{0.17em}0}m<\sqrt{n}}{}\mathrm{}\mathrm{const}\underset{m,n:0m<\sqrt{n}}{}\frac{\mathrm{ln}^{2\alpha 2}(n+2)}{n^2}\hfill \\ \hfill \mathrm{const}\underset{n}{}\sqrt{n}\frac{\mathrm{ln}^{2\alpha 2}(n+2)}{n^2}<\mathrm{}.\end{array}$$
We turn to the other case, $`\sqrt{n}m<n`$. Now $`\mathrm{ln}(n+2)=O(\mathrm{ln}(m+2))`$. Lemma 9.5 gives $`\mathrm{Im}\widehat{b}(\lambda )=O\left(\frac{1}{\mathrm{ln}^\alpha \lambda }\right)`$, hence
$$\begin{array}{c}\frac{\mathrm{ln}^{\alpha 1}(m+2)\mathrm{ln}^{\alpha 1}(n+2)}{(n+m)^2}\left(\mathrm{Im}\widehat{b}(2\pi m/T)+\mathrm{Im}\widehat{b}(2\pi n/T)\right)^2=\hfill \\ \hfill =O\left(\frac{\mathrm{ln}^{\alpha 1}(m+2)\mathrm{ln}^{\alpha 1}(n+2)}{n^2}\left(\frac{1}{\mathrm{ln}^{2\alpha }(m+2)}+\frac{1}{\mathrm{ln}^{2\alpha }(n+2)}\right)\right)=\\ \hfill =O\left(\frac{\mathrm{ln}^{2\alpha 2}(n+2)}{n^2\mathrm{ln}^{2\alpha }(m+2)}\right)=O\left(\frac{1}{n^2\mathrm{ln}^2(n+2)}\right);\end{array}$$
summing over $`m`$ gives $`O\left(\frac{1}{n\mathrm{ln}^2(n+2)}\right)`$, a convergent series in $`n`$, which proves (9.8).
It remains to prove the most delicate case, (9.7) for $`\sqrt{n}m<n`$. Neglecting a finite number of terms, we get $`m`$ large enough for using the asymptotic relation of Lemma 9.5:
$$\widehat{b}(2\pi m/T)\widehat{b}(2\pi n/T)=O\left(_{2\pi m/T}^{2\pi n/T}\frac{d\lambda }{\lambda \mathrm{ln}^\alpha \lambda }\right)=O\left(\frac{1}{\mathrm{ln}^{\alpha 1}m}\frac{1}{\mathrm{ln}^{\alpha 1}n}\right).$$
However, $`(\mathrm{ln}m)^{(\alpha 1)}(\mathrm{ln}n)^{(\alpha 1)}(\alpha 1)(\mathrm{ln}m)^\alpha (\mathrm{ln}n\mathrm{ln}m)`$ and $`(\mathrm{ln}m)^\alpha (\frac{1}{2}\mathrm{ln}n)^\alpha `$, therefore
$$\widehat{b}(2\pi m/T)\widehat{b}(2\pi n/T)=O\left(\frac{\mathrm{ln}n\mathrm{ln}m}{\mathrm{ln}^\alpha n}\right).$$
It is enough to prove that
$$\underset{m,n:\sqrt{n}m<n}{}\frac{\mathrm{ln}^{\alpha 1}m\mathrm{ln}^{\alpha 1}n}{(nm)^2}\left(\frac{\mathrm{ln}n\mathrm{ln}m}{\mathrm{ln}^\alpha n}\right)^2<\mathrm{}$$
or, equivalently,
$$\underset{n}{}\frac{1}{n^2\mathrm{ln}^2n}\underset{m:\sqrt{n}m<n}{}\left(\frac{\mathrm{ln}\frac{n}{m}}{1\frac{m}{n}}\right)^2<\mathrm{}.$$
It remains to note that
$$\frac{1}{n}\underset{m:\sqrt{n}m<n}{}\left(\frac{\mathrm{ln}\frac{n}{m}}{1\frac{m}{n}}\right)^2_0^1\left(\frac{\mathrm{ln}(1/u)}{1u}\right)^2𝑑u<\mathrm{}.$$
###### 9.9 Proposition.
If $`B`$ satisfies (9.2) then $`X_k`$ (defined by (9.3)) are orthogonal w.r.t. some admissible norm on the FHS-space $`G_{0,T}`$.
###### Proof.
Here is an equivalent formulation: there is an operator $`A:l_2G_{0,T}`$ such that
$$A(c_1,c_2,\mathrm{})=\underset{k}{}\frac{c_k}{X_k}X_k\text{for all }(c_1,c_2,\mathrm{})l_2,$$
and $`A`$ is an FHS-isomorphism, in other words, an equivalence operator in the sense of Feldman \[7, Def. 1\]. It means that $`A`$ is one-to-one onto, has a bounded inverse, and $`\sqrt{A^{}A}I𝒮`$ (the Hilbert-Schmidt class of operators). The latter is equivalent to $`A^{}AI𝒮`$, see \[7, Lemma 1(b)\].<sup>35</sup><sup>35</sup>35Though, his formulation of the lemma is incorrect, see the review 21#1546 in Mathematical Reviews.
Matrix elements of $`A^{}A`$ are $`{\displaystyle \frac{|X_m,X_n|}{X_mX_n}}`$; Lemma 9.6 shows that $`A^{}AI𝒮`$ and, of course, $`A`$ is bounded. It remains to prove that $`A`$ has a bounded inverse. The range of $`A`$ being evidently dense, we have to prove that $`Ax\epsilon x`$ for some $`\epsilon `$, that is, $`0`$ does not belong to the spectrum of $`A^{}A`$. The spectrum accumulates to $`1`$ only (since $`A^{}AI𝒮`$); we have to prove that $`0`$ is not an eigenvalue, that is,
$$\underset{k}{}\frac{c_k}{X_k}X_k=0c_1=c_2=\mathrm{}=0$$
for all $`(c_1,c_2,\mathrm{})l_2`$. It is enough to prove that the following formula is a correct definition of (continuous) linear functionals $`X^1,X^2,\mathrm{}`$ on $`G_{0,T}`$:
$$X^k(g)=\frac{1}{T}_0^Tg(t)\mathrm{exp}(2\pi ikt/T)𝑑t\text{for }gG_{0,T};$$
indeed, it will follow that
$$c_k=X_kX^k\left(\underset{l}{}\frac{c_l}{X_l}X_l\right).$$
The norm on $`G_{0,T}`$, defined in terms of $`B()`$, uses $`B(t)`$ for $`t[T,T]`$ only. Therefore we may assume that $`B(t)`$ vanishes outside of some bounded interval (and still satisfies (9.2)). For every $`gL_2(0,T)G_{0,T}`$,
$$g_{G_{0,T}}^2=_{\mathrm{}}^+\mathrm{}\widehat{B}(\lambda )|\widehat{g}(\lambda )|^2𝑑\lambda ;$$
here $`\widehat{B}`$ is the Fourier transform (normalized as to be unitary) of $`B`$, and $`\widehat{g}`$ — of $`g`$. The formula $`(Zg)(\lambda )=\sqrt{\widehat{B}(\lambda )}\widehat{g}(\lambda )`$ defines a linear isometric embedding $`Z:G_{0,T}L_2()`$ on the dense subset $`L_2(0,T)G_{0,T}`$; we may extend $`Z`$ to the whole $`G_{0,T}`$ by continuity. Every $`\phi L_2()`$ gives a linear functional on $`G_{0,T}`$, namely, $`g\sqrt{\widehat{B}(\lambda )}\widehat{g}(\lambda )\phi (\lambda )𝑑\lambda `$. In order to get $`X^k`$, we take $`\phi `$ such that $`\sqrt{\widehat{B}(\lambda )}\phi (\lambda )`$ is the Fourier transform of $`(1/T)\mathrm{exp}(2\pi ikt/T)\mathrm{𝟏}_{(0,T)}`$; it remains to verify that such $`\phi `$ belongs to $`L_2()`$.
The function $`(1/T)\mathrm{exp}(2\pi ikt/T)\mathrm{𝟏}_{(0,T)}`$ is of finite variation; its Fourier transform is $`O\left(1/\sqrt{1+\lambda ^2}\right)`$; it remains to check that
$$\frac{1}{(1+\lambda ^2)\widehat{B}(\lambda )}𝑑\lambda <\mathrm{}.$$
However, the continuous function $`\widehat{B}`$ never vanishes, and $`\widehat{B}(\lambda ){\displaystyle \frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}|\lambda |}}`$ for $`\lambda \pm \mathrm{}`$ by Lemma 9.5. ∎
###### 9.10 Proposition.
If $`B`$ satisfies (9.2) then $`G_{T,T}=G_{T,0}G_{0,T}`$ (in the FHS sense).
###### Proof.
Vectors $`X_k,Y_k`$ (defined by (9.3)) are orthogonal w.r.t. some admissible norm on the FHS-space $`G_{T,T}`$; the proof is quite similar to the proof of Proposition 9.9. ∎
Combining Proposition 9.10 with elementary properties of spaces $`G_{a,b}`$ and operators $`U_t`$ mentioned in the beginning of the section, we get the following result.
###### 9.11 Theorem.
If $`B`$ satisfies (9.2) for some $`\alpha (1,\mathrm{})`$, then $`((G_{a,b}),(U_t))`$ is a sum system (as defined by 7.1).
## 10 The product systems are nonisomorphic and unitless
Sum systems given by Theorem 9.11 depend on the parameter $`\alpha (1,\mathrm{})`$.<sup>36</sup><sup>36</sup>36That is, each such system corresponds to some $`\alpha `$. However, for a given $`\alpha `$ there is some freedom when choosing $`B`$ satisfying (9.2). Their exponentials, given by Theorem 7.5, are product systems. Such product systems for different $`\alpha `$ are nonisomorphic, which will be shown using Proposition 8.1. Accordingly, we consider a sequence of elementary sets $`E_n(0,1)`$, and we want to know, which sequences $`(E_n)`$ satisfy $`lim\; infG_{(0,1)E_n}=G_{(0,1)}`$ and $`lim\; supG_{E_n}=\{0\}`$; here $`G_{E_n}`$ (as well as $`G_{(0,1)E_n}`$) correspond (as explained in Sect. 8) to the sum system given by Theorem 9.11; recall that $`lim\; inf`$ was defined by 3.6, and $`lim\; sup`$ by 3.13.
###### 10.1 Lemma.
If $`\mathrm{mes}E_n0`$ then $`lim\; infG_{(0,1)E_n}=G_{(0,1)}`$. (Here $`\mathrm{mes}E_n`$ is Lebesgue measure of $`E_n`$.)
###### Proof.
We have to represent an arbitrary vector $`gG_{(0,1)}`$ as $`limg_n`$ for some $`g_nG_{(0,1)E_n}`$. It is enough to consider a dense set of vectors $`g`$ (since $`lim\; inf`$ is always closed). Let $`gL_2(0,1)G_{(0,1)}`$ and $`g_n=g\mathrm{𝟏}_{(0,1)E_n}G_{(0,1)E_n}L_2(0,1)`$. Clearly, $`\mathrm{mes}E_n0`$ implies $`g\mathrm{𝟏}_{E_n}0`$ in $`L_2(0,1)`$, hence $`g_ng`$ in $`L_2(0,1)`$, therefore $`g_ng`$ in $`G_{0,1}`$. ∎
###### 10.2 Lemma.
Assume that $`B`$ satisfies (9.2) for a given $`\alpha (1,\mathrm{})`$, and $`\mathrm{mes}E_n=o(1/\mathrm{ln}^{\alpha 1}n)`$, and $`E_n`$ consists of (no more than) $`n`$ intervals. Then $`lim\; supG_{E_n}=\{0\}`$.
###### Proof.
Let $`g_nG_{E_n}`$, $`g_n1`$; we have to prove that $`g_n0`$ weakly. Introduce
$$h_n(t)=c\frac{n^2}{\mathrm{ln}^{(\alpha 1)/2}n}\mathrm{𝟏}_{(1/n^2,+1/n^2)}$$
where $`c>0`$ is chosen such that for all $`n`$ large enough,
$$Bh_nh_n\text{is positively definite;}$$
in terms of Fourier transform it means that
$$\left(\widehat{h}_n(\lambda )\right)^2B(\lambda )\text{for all }\lambda ;$$
such $`c`$ exists due to the asymptotic relation $`\widehat{B}(\lambda )\frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}|\lambda |}`$ for large $`|\lambda |`$ (recall 9.5). The positive definiteness means that
$$gh_n_{L_2()}g_{G_{0,1}}$$
for all $`gL_2(0,1)G_{0,1}`$ and then (extending the convolution operator by continuity) for all $`gG_{0,1}`$.
Denote by $`E_n^{}`$ the $`(1/n^2)`$-neighborhood of $`E_n`$, then $`\mathrm{mes}E_n^{}(2n/n^2)+\mathrm{mes}E_n=o(1/\mathrm{ln}^{\alpha 1}n)`$, and $`gh_nL_2(E_n^{})L_2()`$.
We have to prove that $`\phi (g_n)0`$ for every linear functional $`\phi `$ on $`G_{0,1}`$. We may restrict ourselves to a dense subset of functionals $`\phi `$ (since $`g_n1`$). In particular, we may consider only functionals $`\phi _k`$ defined by
$$\phi _k(g)=g(t)\mathrm{exp}(2\pi ikt)𝑑t.$$
Taking into account that $`\phi _k(gh_n)={\displaystyle \frac{2c}{\mathrm{ln}^{(\alpha 1)/2}n}}{\displaystyle \frac{\mathrm{sin}(2\pi k/n^2)}{2\pi k/n^2}}\phi _k(g)`$ we see that the following would be enough:
$$\frac{1}{2c}\left(\mathrm{ln}^{(\alpha 1)/2}n\right)\phi _k(g_nh_n)0\text{when }n\mathrm{}$$
for every $`k`$.
Recalling that $`g_nh_nL_2(E_n^{})`$ and $`g_nh_n_{L_2()}g_n_{G_{0,1}}1`$ we have
$$|\phi _k(g_nh_n)|\sqrt{\mathrm{mes}E_n^{}}g_nh_n_{L_2()}=o\left(\frac{1}{\mathrm{ln}^{(\alpha 1)/2}n}\right).$$
###### 10.3 Lemma.
Assume that $`B`$ satisfies (9.2) for a given $`\alpha (1,\mathrm{})`$. Then there exists a sequence of elementary sets $`E_n(0,1)`$ such that $`\mathrm{mes}E_n=O(1/\mathrm{ln}^{\alpha 1}n)`$, and $`E_n`$ consists of $`n`$ intervals, and the relation $`lim\; supG_{E_n}=\{0\}`$ is violated.
###### Proof.
The construction is straightforward, just $`n`$ equidistant intervals of equal length:
$$\begin{array}{cc}\hfill E_n& =\underset{k=1}{\overset{n}{}}(\frac{1}{n}\left(k\frac{1}{2}\frac{1}{n\mathrm{ln}^{\alpha 1}n}\right),\frac{1}{n}\left(k\frac{1}{2}+\frac{1}{n\mathrm{ln}^{\alpha 1}n}\right));\hfill \\ \hfill \mathrm{mes}E_n& =\frac{2}{\mathrm{ln}^{\alpha 1}n}.\hfill \end{array}$$
We have to find $`g_nG_{E_n}`$ and $`gG_{0,1}`$ such that
(10.4)
$$\underset{n}{sup}g_n_{G_{0,1}}<\mathrm{},$$
(10.5)
$$\underset{n}{lim\; sup}|g_n,g_{G_{0,1}}|>0.$$
Still, the construction is straightforward:
$$\begin{array}{cc}\hfill g_n& =(1/\mathrm{mes}E_n)\mathrm{𝟏}_{E_n}=\frac{1}{2}\mathrm{ln}^{\alpha 1}n\mathrm{𝟏}_{E_n},\hfill \\ \hfill g& =\mathrm{𝟏}_{(0,1)};\hfill \end{array}$$
the proof of (10.4), (10.5) is more complicated. Fourier transform $`f\widehat{f}`$ will be used, $`\widehat{f}(\lambda )=e^{i\lambda t}f(t)𝑑t`$. Recall that
$$2\pi f_1,f_2_{G_{0,1}}=_{\mathrm{}}^+\mathrm{}\widehat{B}(\lambda )\widehat{f}_1(\lambda )\overline{\widehat{f}_2(\lambda )}𝑑\lambda $$
for all $`f_1,f_2G_{0,1}`$, and $`\widehat{B}(\lambda )\frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}|\lambda |}`$ for large $`|\lambda |`$. An elementary calculation gives
$$\begin{array}{cc}\hfill \widehat{g}(\lambda )& =\frac{e^{i\lambda }1}{i\lambda };\hfill \\ & \underset{k=0}{\overset{n1}{}}\mathrm{exp}(ik\lambda /n)=\frac{1\mathrm{exp}(i\lambda )}{1\mathrm{exp}(i\lambda /n)};\hfill \\ \hfill \widehat{g}_n(\lambda )& =(\mathrm{ln}n)^{\alpha 1}\mathrm{exp}\left(i\frac{\lambda }{2n}\right)\frac{1\mathrm{exp}(i\lambda )}{1\mathrm{exp}(i\lambda /n)}\frac{1}{\lambda }\mathrm{sin}\frac{\lambda }{n\mathrm{ln}^{\alpha 1}n};\hfill \\ \hfill |\widehat{g}_n(\lambda )|& =(\mathrm{ln}n)^{\alpha 1}\frac{|\mathrm{sin}\frac{\lambda }{2}|}{|\mathrm{sin}\frac{\lambda }{2n}|}\frac{1}{|\lambda |}\left|\mathrm{sin}\frac{\lambda }{n\mathrm{ln}^{\alpha 1}n}\right|.\hfill \end{array}$$
Note that
$$\frac{1}{2\pi n}_0^{2\pi n}\frac{\mathrm{sin}^2\frac{\lambda }{2}}{\mathrm{sin}^2\frac{\lambda }{2n}}𝑑\lambda =\frac{1}{2\pi n}_0^{2\pi n}\left|\underset{k=0}{\overset{n1}{}}\mathrm{exp}(ik\lambda /n)\right|^2𝑑\lambda =n.$$
We have
$$2\pi g_n^2=_{\mathrm{}}^+\mathrm{}\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda =2\left(_0^n+_n^{\mathrm{}}\right)\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2d\lambda .$$
For $`\lambda (0,n)`$ we note that
$$\begin{array}{cc}& \widehat{B}(\lambda )\widehat{B}(0),\hfill \\ & |\widehat{g}_n(\lambda )|\frac{1}{n}\frac{|\mathrm{sin}\frac{\lambda }{2}|}{|\mathrm{sin}\frac{\lambda }{2n}|},\hfill \\ & _0^n|\widehat{g}_n(\lambda )|^2𝑑\lambda \frac{1}{n^2}_0^n\frac{\mathrm{sin}^2\frac{\lambda }{2}}{\mathrm{sin}^2\frac{\lambda }{2n}}𝑑\lambda 2\pi ,\hfill \end{array}$$
therefore
$$_0^n\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda 2\pi \widehat{B}(0)\text{for all }n.$$
For $`\lambda (n,\mathrm{})`$ we note that $`\widehat{B}(\lambda )\frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}n}`$ and change the scale, introducing $`u={\displaystyle \frac{\lambda }{n\mathrm{ln}^{\alpha 1}n}}`$:
$$\begin{array}{c}_n^{\mathrm{}}\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda \frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}n}(\mathrm{ln}n)^{2(\alpha 1)}_n^{\mathrm{}}\frac{\mathrm{sin}^2\frac{\lambda }{2}}{\mathrm{sin}^2\frac{\lambda }{2n}}\frac{1}{\lambda ^2}\mathrm{sin}^2\frac{\lambda }{n\mathrm{ln}^{\alpha 1}n}d\lambda \hfill \\ \hfill \frac{\mathrm{const}}{n}_0^{\mathrm{}}\frac{\mathrm{sin}^2(n\omega _nu)}{\mathrm{sin}^2(\omega _nu)}\frac{\mathrm{sin}^2u}{u^2}𝑑u,\end{array}$$
where $`\omega _n=\frac{1}{2}\mathrm{ln}^{\alpha 1}n\mathrm{}`$. On each period, $`u(\frac{\pi }{\omega _n}k,\frac{\pi }{\omega _n}(k+1))`$, we substitute $`\frac{\mathrm{sin}^2u}{u^2}`$ by its maximal value:
$$\begin{array}{c}_0^{\mathrm{}}\frac{\mathrm{sin}^2(n\omega _nu)}{\mathrm{sin}^2(\omega _nu)}\frac{\mathrm{sin}^2u}{u^2}𝑑u\hfill \\ \hfill (1+o(1))\left(_0^{\mathrm{}}\frac{\mathrm{sin}^2u}{u^2}𝑑u\right)\underset{=n}{\underset{}{\left(\frac{\omega _n}{\pi }_0^{\pi /\omega _n}\frac{\mathrm{sin}^2(n\omega _nu)}{\mathrm{sin}^2(\omega _nu)}𝑑u\right)}}.\end{array}$$
Hence $`_n^{\mathrm{}}\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda =O(1)`$. So, (10.4) is verified.
Further,
$$2\pi g_n,g=_{\mathrm{}}^+\mathrm{}\widehat{B}(\lambda )\widehat{g}_n(\lambda )\overline{\widehat{g}(\lambda )}𝑑\lambda =\left(_{|\lambda |<M}+_{|\lambda |>M}\right)\widehat{B}(\lambda )\widehat{g}_n(\lambda )\overline{\widehat{g}(\lambda )}d\lambda $$
for any $`M(0,\mathrm{})`$. It is easy to see that
$$\widehat{g}_n(\lambda )\underset{n\mathrm{}}{\overset{}{}}\widehat{g}(\lambda )\text{uniformly in }\lambda [M,M],$$
which implies
$$_{|\lambda |<M}\widehat{B}(\lambda )\widehat{g}_n(\lambda )\overline{\widehat{g}(\lambda )}𝑑\lambda _{|\lambda |<M}\widehat{B}(\lambda )|\widehat{g}(\lambda )|^2𝑑\lambda \text{for }n\mathrm{}.$$
On the other hand,
$$\begin{array}{c}\left|_{|\lambda |>M}\widehat{B}(\lambda )\widehat{g}_n(\lambda )\overline{\widehat{g}(\lambda )}𝑑\lambda \right|\hfill \\ \hfill \underset{sup_ng_n}{\underset{}{\left(_{|\lambda |>M}\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda \right)^{1/2}}}\underset{0\text{ for }M\mathrm{}}{\underset{}{\left(_{|\lambda |>M}\widehat{B}(\lambda )|\widehat{g}(\lambda )|^2𝑑\lambda \right)^{1/2}}},\end{array}$$
it tends to $`0`$ uniformly in $`n`$, when $`M\mathrm{}`$. Choose $`M`$ and $`\epsilon >0`$ such that
$$\underset{n}{sup}\left|_{|\lambda |>M}\widehat{B}(\lambda )\widehat{g}_n(\lambda )\overline{\widehat{g}(\lambda )}𝑑\lambda \right|\frac{1}{2}\epsilon \text{and}_{|\lambda |<M}\widehat{B}(\lambda )|\widehat{g}(\lambda )|^2𝑑\lambda \epsilon ,$$
then
$$\underset{n}{lim\; inf}\mathrm{Re}g_n,g\frac{1}{2}\frac{1}{2\pi }\epsilon ,$$
which implies (10.5). ∎
###### 10.6 Theorem.
Let<sup>37</sup><sup>37</sup>37Here $`B^{}`$ does not mean the derivative of $`B`$, sorry. $`B^{},B^{\prime \prime }`$ satisfy (9.2) for some $`\alpha ^{},\alpha ^{\prime \prime }`$ respectively, $`\alpha ^{},\alpha ^{\prime \prime }(1,\mathrm{})`$, $`\alpha ^{}\alpha ^{\prime \prime }`$. Then the corresponding product systems are nonisomorphic.
###### Proof.
Suppose that $`\alpha ^{}<\alpha ^{\prime \prime }`$. Lemma 10.3 gives elementary sets $`E_n(0,1)`$ such that $`\mathrm{mes}E_n=O(1/\mathrm{ln}^{\alpha ^{\prime \prime }1}n)`$, and $`E_n`$ consists of $`n`$ intervals, and the relation $`lim\; supG_{E_n}^{\prime \prime }=\{0\}`$ is violated. Taking into account that $`O(1/\mathrm{ln}^{\alpha ^{\prime \prime }1}n)=o(1/\mathrm{ln}^{\alpha ^{}1}n)`$ we get $`lim\; supG_{E_n}^{}=\{0\}`$ by Lemma 10.2. Also, $`lim\; infG_{(0,1)E_n}^{}=G_{(0,1)}^{}`$ and $`lim\; infG_{(0,1)E_n}^{\prime \prime }=G_{(0,1)}^{\prime \prime }`$ by Lemma 10.1. So, $`E_n`$ satisfy 8.1(a) but violate 8.1(b). By Proposition 8.1 the product systems are nonisomorphic. ∎
###### 10.7 Lemma.
Assume that $`B`$ satisfies (9.2) for a given $`\alpha (1,\mathrm{})`$, and
$$\begin{array}{cc}& E_n=\underset{k=1}{\overset{n}{}}(\frac{1}{n}\left(k\frac{1}{2}\frac{\mathrm{mes}E_n}{2n}\right),\frac{1}{n}\left(k\frac{1}{2}+\frac{\mathrm{mes}E_n}{2n}\right)),\hfill \\ & (\mathrm{ln}n)^{\alpha 1}\mathrm{mes}E_n\mathrm{}\text{for }n\mathrm{},\hfill \\ & g_n=(1/\mathrm{mes}E_n)\mathrm{𝟏}_{E_n},\hfill \\ & g=\mathrm{𝟏}_{(0,1)}.\hfill \end{array}$$
Then $`g_ng_{G_{0,1}}0`$ for $`n\mathrm{}`$.
###### Proof.
Similarly to the proof of Lemma 10.3 we have for any $`M(0,\mathrm{})`$
$$\widehat{g}_n(\lambda )\underset{n\mathrm{}}{\overset{}{}}\widehat{g}(\lambda )\text{uniformly on }\lambda [M,M].$$
This time, however, the following compactness property holds:
$$_{|\lambda |>M}\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda \underset{M\mathrm{}}{\overset{}{}}0\text{uniformly in }n,$$
which ensures $`g_ng0`$. In order to prove the compactness property we estimate integrals similarly to the proof of 10.3. We have
$$|\widehat{g}_n(\lambda )|=\frac{2}{\mathrm{mes}E_n}\frac{|\mathrm{sin}\frac{\lambda }{2}|}{|\mathrm{sin}\frac{\lambda }{2n}|}\frac{1}{|\lambda |}\left|\mathrm{sin}\frac{\lambda \mathrm{mes}E_n}{2n}\right|.$$
For $`\lambda (M,n)`$ we note that<sup>38</sup><sup>38</sup>38Assuming that $`M`$ is large enough.
$$\begin{array}{cc}& \widehat{B}(\lambda )\frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}M},\hfill \\ & |\widehat{g}_n(\lambda )|\frac{1}{n}\frac{|\mathrm{sin}\frac{\lambda }{2}|}{|\mathrm{sin}\frac{\lambda }{2n}|},\text{(the same as in }\text{10.3}\text{)}\hfill \\ & _M^n\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda \frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}M}_0^n|\widehat{g}_n(\lambda )|^2𝑑\lambda \frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}M}2\pi \underset{M\mathrm{}}{\overset{}{}}0.\hfill \end{array}$$
For $`\lambda (n,\mathrm{})`$, introducing $`u={\displaystyle \frac{\mathrm{mes}E_n}{2n}}\lambda `$ and $`\omega _n=1/\mathrm{mes}E_n\mathrm{}`$, we have
$$\begin{array}{c}_n^{\mathrm{}}\widehat{B}(\lambda )|\widehat{g}_n(\lambda )|^2𝑑\lambda \frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}n}\left(\frac{2}{\mathrm{mes}E_n}\right)^2_n^{\mathrm{}}\frac{\mathrm{sin}^2\frac{\lambda }{2}}{\mathrm{sin}^2\frac{\lambda }{2n}}\frac{1}{\lambda ^2}\mathrm{sin}^2\frac{\lambda \mathrm{mes}E_n}{2n}d\lambda \hfill \\ \hfill \underset{0}{\underset{}{\frac{\mathrm{const}}{\mathrm{ln}^{\alpha 1}n}\frac{2}{\mathrm{mes}E_n}}}\frac{1}{n}\underset{O(n)}{\underset{}{_0^{\mathrm{}}\frac{\mathrm{sin}^2(n\omega _nu)}{\mathrm{sin}^2(\omega _nu)}\frac{\mathrm{sin}^2u}{u^2}𝑑u}}\underset{n\mathrm{}}{\overset{}{}}0.\end{array}$$
###### 10.8 Proposition.
If $`B`$ satisfies (9.2) for some $`\alpha (1,\mathrm{})`$ then the corresponding product system is unitless.
###### Proof.
Assume the contrary, then there exist $`\psi _{a,b}H_{a,b}=\mathrm{Exp}G_{a,b}`$ such that $`\psi _{a,b}=1`$ and $`\psi _{a,b}\psi _{b,c}=\psi _{a,c}`$ whenever $`a<b<c`$. We’ll use $`\psi _{a,b}`$ for $`0a<b1`$ only. Introduce a Gaussian type space $`(\mathrm{\Omega },,𝒫,G)`$ such that $`G_{0,1}=G/\mathrm{Const}`$; we have $`H_{a,b}=L_2(\mathrm{\Omega },_{a,b},𝒫|_{_{a,b}})`$ where $`_{a,b}`$ is the $`\sigma `$-field generated by the Gaussian subspace $`G_{a,b}G_{0,1}`$. Consider measures $`\mu _{a,b}=|\psi _{a,b}|^2`$; as was explained in Sect. 1, $`\mu _{a,b}`$ is a probability measure on the $`\sigma `$-field $`_{a,b}`$ such that
$$\frac{\mu _{a,b}}{P|_{_{a,b}}}=\left|\frac{\psi _{a,b}}{\sqrt{P|_{_{a,b}}}}\right|^2\text{for all }P𝒫.$$
The relation $`\psi _{a,b}\psi _{b,c}=\psi _{a,c}`$ implies
$$\mu _{a,b}\mu _{b,c}=\mu _{a,c}.$$
In other words, $`\sigma `$-fields $`_{a,b}`$ and $`_{b,c}`$ are independent w.r.t. the measure $`\mu =\mu _{0,1}`$ on $``$, and $`\mu _{a,b}`$ is just the restriction of $`\mu `$ to $`_{a,b}`$.<sup>39</sup><sup>39</sup>39Clearly, $`\mu `$ is a probability measure on $`(\mathrm{\Omega },)`$, absolutely continuous w.r.t. any $`P𝒫`$. However, $`P`$ need not be absolutely continuous w.r.t. $`\mu `$; that is, $`\mu `$ need not belong to $`𝒫`$. Moreover, $`\sigma `$-fields $`_{t_0,t_1},_{t_1,t_2},\mathrm{},_{t_{n1},t_n}`$ are $`\mu `$-independent whenever $`0t_0<t_1<\mathrm{}<t_n1`$. Every elementary set $`E(0,1)`$ determines its sub-$`\sigma `$-field $`_E`$; as explained in Sect. 8,
$$E=(t_0,t_1)(t_2,t_3)\mathrm{}(t_{2n},t_{2n+1})_E=_{t_0,t_1}\mathrm{}_{t_{2n},t_{2n+1}}$$
whenever $`0t_0<t_1<\mathrm{}<t_{2n+1}1`$. Note that $`_{E_1},_{E_2}`$ are $`\mu `$-independent whenever $`E_1E_2=\mathrm{}`$.
Consider elementary sets
$$\begin{array}{cc}\hfill E_n& =(0,\frac{1}{2n})(\frac{2}{2n},\frac{3}{2n})\mathrm{}(\frac{2n2}{2n},\frac{2n1}{2n}),\hfill \\ \hfill \mathrm{mes}E_n& =\frac{1}{2},\hfill \end{array}$$
and vectors
$$\begin{array}{cc}\hfill g_n& =2\mathrm{𝟏}_{E_n}G_{E_n}G_{0,1},\hfill \\ \hfill g& =\mathrm{𝟏}_{(0,1)}G_{0,1},\hfill \\ \hfill h_n& =2\mathrm{𝟏}_{(0,1)E_n}G_{(0,1)E_n}G_{0,1}.\hfill \end{array}$$
Lemma 10.7 shows<sup>40</sup><sup>40</sup>40Shifting the set $`E_n`$ of Lemma 10.7 does not invalidate the lemma. that $`g_ng`$ and $`h_ng`$ in $`G_{0,1}`$. Though, $`G_{0,1}`$ is not a subspace but a quotient space $`G/\mathrm{Const}`$ of $`G`$; anyway, we may choose elements of $`G`$, denoted again by $`g_n,g,h_n`$, such that
$$\begin{array}{cc}& g_nL_0(\mathrm{\Omega },_{E_n},𝒫),\hfill \\ & gL_0(\mathrm{\Omega },,𝒫),\hfill \\ & h_nL_0(\mathrm{\Omega },_{(0,1)E_n},𝒫),\hfill \\ & g_ng\text{and}h_ng\text{in }L_0(\mathrm{\Omega },,𝒫).\hfill \end{array}$$
The natural map<sup>41</sup><sup>41</sup>41Generally, non-invertible, since $`\mu `$ need not belong to $`𝒫`$. $`L_0(\mathrm{\Omega },,𝒫)L_0(\mathrm{\Omega },,\mu )`$ allows us to treat $`g_n,g,h_n`$ as elements of $`L_0(\mathrm{\Omega },,\mu )`$. Now they are random variables; $`g_ng`$ and $`h_ng`$ in probability. On the other hand, for every $`n`$, the two random variables $`g_n,h_n`$ are independent (since $`_{E_n}`$ and $`_{(0,1)E_n}`$ are $`\mu `$-independent). It follows that $`g`$ is independent of itself, that is, $`g`$ is constant $`\mu `$-almost sure.<sup>42</sup><sup>42</sup>42Proof: let $`\phi :`$ be a continuous and bounded function, then $`\phi (g_n(\omega ))\phi (h_n(\omega ))\mu (d\omega )=\left(\phi (g_n(\omega ))\mu (d\omega )\right)\left(\phi (h_n(\omega ))\mu (d\omega )\right)`$ due to independence. The limit for $`n\mathrm{}`$ gives $`\left(\phi (g(\omega ))\right)^2\mu (d\omega )=\left(\phi (g(\omega ))\mu (d\omega )\right)^2`$, which is impossible unless $`g=\mathrm{const}`$.
Consider a Gaussian measure $`\gamma 𝒫_G`$. Though, $`g`$ need not be constant $`\gamma `$-almost sure; however, $`g`$ must be constant on a set of positive probability w.r.t. $`\gamma `$. On the other hand, the distribution of $`g`$ w.r.t. $`\gamma `$ is normal (Gaussian); it cannot have an atom unless it is degenerate, which means that $`g_{G_{0,1}}`$ must vanish. However, it does not vanish, which is evident when using Fourier transform. A contradiction. ∎
School of Mathematics, Tel Aviv Univ., Tel Aviv 69978, Israel
tsirel@math.tau.ac.il
http://www.math.tau.ac.il/$``$tsirel/
|
warning/0006/gr-qc0006019.html
|
ar5iv
|
text
|
# Negative Norm States in de Sitter Space and QFT without Renormalization Procedure
## 1 Introduction
Antoniadis, Iliopoulos and Tomaras have shown that the pathological large-distance behavior (infrared divergence) of the graviton propagator on a de Sitter background does not manifest itself in the quadratic part of the effective action in the one-loop approximation. This means that the pathological behavior of the graviton propagator may be gauge dependent and so should not appear in an effective way as a physical quantity. The linear gravity (the traceless rank-2 “massless” tensor field) on de Sitter space is indeed built up from copies of the minimally coupled scalar field . It has been shown that one can construct a covariant quantization of the “massless” minimally coupled scalar field in de Sitter space-time, which is causal and free of any infrared divergence. The essential point of that paper is the unavoidable presence of the negative norm states. Although they do not propagate in the physical space, they play a renormalizing role. In the forthcoming paper , we shall show that this is also true for linear gravity (the traceless rank-2 “massless” tensor field). These questions have recently been studied by several authors (for minimally coupled scalar field see and for linear gravity see ).
The auxiliary states (the negative norm states) appear to be necessary for obtaining a fully covariant quantization of the free minimally coupled scalar field in de Sitter space-time, which is free of any infrared divergence. It has been shown that these auxiliary states automatically renormalize the infrared divergence in the two-point function and removes the ultraviolet divergence in the stress tensor . The crucial point about the minimally coupled scalar field lies in the fact that there is no de Sitter invariant decomposition
$$=_+_{},$$
where $`_+`$ and $`_{}`$ are Hilbert and anti-Hilbert spaces respectively. For this reason our states contain negative frequency solutions and consequently the use of a Krein space (i.e. Hilbert $``$ anti-Hilbert space) is necessitated. For the scalar massive field where such a decomposition exists as a de Sitter invariant, $`_+`$ as the usual physical state space $`(_{}=_+^{})`$ suffices . It has been also shown that if this method is applied to the free “massive” scalar field in de Sitter space, automatically covariant renormalization of the vacuum energy divergence is obtained . We would like to generalize this method (adding the negative frequency solutions) to the interacting quantum scalar field in Minkowski space-time. These auxiliary states once again, automatically renormalize the problem to the one-loop approximation. In other words, introducing negative frequency solutions plays the key role in the renormalization procedure.
## 2 de Sitter scalar field
Let us briefly describe our quantization of the minimally coupled massless scalar field. It is defined by
$$\mathrm{}_H\varphi (x)=0,$$
where $`\mathrm{}_H`$ is the Laplace-Beltrami operator on de Sitter space. As proved by Allen , the covariant canonical quantization procedure with positive norm states fails in this case. The Allen’s result can be reformulated in the following way: the Hilbert space generated by a complete set of modes (named here the positive modes, including the zero mode) is not de Sitter invariant,
$$=\{\underset{k0}{}\alpha _k\varphi _k;\underset{k0}{}|\alpha _k|^2<\mathrm{}\}.$$
This means that it is not closed under the action of the de Sitter group. Nevertheless, one can obtain a fully covariant quantum field by adopting a new construction . In order to obtain a fully covariant quantum field, we add all the conjugate modes to the previous ones. Consequently, we have to deal with an orthogonal sum of a positive and negative inner product space, which is closed under an indecomposable representation of the de Sitter group. The negative values of the inner product are precisely produced by the conjugate modes: $`\varphi _k^{},\varphi _k^{}=1`$, $`k0`$. We do insist on the fact that the space of solution should contain the unphysical states with negative norm. Now, the decomposition of the field operator into positive and negative norm parts reads
$$\varphi (x)=\frac{1}{\sqrt{2}}\left[\varphi _p(x)+\varphi _n(x)\right],$$
(1)
where
$$\varphi _p(x)=\underset{k0}{}a_k\varphi _k(x)+H.C.,\varphi _n(x)=\underset{k0}{}b_k\varphi ^{}(x)+H.C..$$
(2)
The positive mode $`\varphi _p(x)`$ is the scalar field as was used by Allen. The crucial departure from the standard QFT based on CCR lies in the following requirement on commutation relations:
$$a_k|0>=0,[a_k,a_k^{}^{}]=\delta _{kk^{}},b_k|0>=0,[b_k,b_k^{}^{}]=\delta _{kk^{}}.$$
(3)
A direct consequence of these formulas is the positivity of the energy i.e.
$$\stackrel{}{k}|T_{00}|\stackrel{}{k}0,$$
for any physical state $`|\stackrel{}{k}`$ (those built from repeated action of the $`a_k^{}`$’s on the vacuum). This quantity vanishes if and only if $`|\stackrel{}{k}=|0`$. Therefore the “normal ordering” procedure for eliminating the ultraviolet divergence in the vacuum energy, which appears in the usual QFT is not needed . Another consequence of this formula is a covariant two-point function, which is free of any infrared divergence .
This result is the same as that of de Vega et al. where flat coordinate modes solutions were employed. For calculating the Schwinger commutator function, they have not used the two-point function since in their construction it would result in appearance of a divergence. They calculated the commutator function directly, which resulted in disappearance of the infrared divergence due to the sign of the divergence term. In our alternative method, the Schwinger commutator function was calculated from the finite and covariant two point function . It has been also shown that the Schwinger commutator functions in both methods are one and the same.
## 3 Minkowskian free quantum scalar field
Let us first recall elementary facts about Minkowskian QFT. A classical scalar field $`\varphi (x)`$, which is defined in the $`4`$-dimensional Minkowski space-time, satisfies the field equation
$$(\mathrm{}+m^2)\varphi (x)=0=(\eta ^{\mu \nu }_\mu _\nu +m^2)\varphi (x),\eta ^{\mu \nu }=\text{diag}(1,1,1,1).$$
(4)
Inner or Klein-Gordon product and related norm are defined by
$$(\varphi _1,\varphi _2)=i_{t=\text{cons.}}\varphi _1(x)\underset{t}{\overset{}{}}\varphi _2^{}(x)d^3x.$$
(5)
Two sets of solutions of $`(4)`$ are given by:
$$u_p(k,x)=\frac{e^{i\stackrel{}{k}.\stackrel{}{x}iwt}}{\sqrt{(2\pi )^32w}}=\frac{e^{ik.x}}{\sqrt{(2\pi )^32w}},u_n(k,x)=\frac{e^{i\stackrel{}{k}.\stackrel{}{x}+iwt}}{\sqrt{(2\pi )^32w}}=\frac{e^{ik.x}}{\sqrt{(2\pi )^32w}},$$
(6)
where $`w(\stackrel{}{k})=k^0=(\stackrel{}{k}.\stackrel{}{k}+m^2)^{\frac{1}{2}}0`$. These $`u(k,x)`$ modes are orthogonal and normalized in the sense of $`(5)`$:
$$(u_p(k,x),u_p(k^{},x))=\delta ^3(\stackrel{}{k}\stackrel{}{k}^{}),$$
$$(u_n(k,x),u_n(k^{},x))=\delta ^3(\stackrel{}{k}\stackrel{}{k}^{}),$$
$$(u_p(k,x),u_n(k^{},x))=0.$$
(7)
$`u_p`$ modes are positive norm states and the $`u_n`$’s are negative norm states. The general classical field solution is
$$\varphi (x)=d^3\stackrel{}{k}[a(\stackrel{}{k})u_p(k,x)+b(\stackrel{}{k})u_n(k,x)],$$
where $`a(\stackrel{}{k})`$ and $`b(\stackrel{}{k})`$ are two independent coefficients. The usual quantization of this field is based on the positive norm states only. In the Minkowskian case this choice leads to a covariant quantization (covariant under the proper orthochronous Poincaré group). However, it is well known that an ultraviolet divergence appears in the vacuum energy. This divergence is eliminated with the aid of a “normal ordering” operation.
In the above, we have seen that, in the case of the minimally coupled scalar field in de Sitter space, one cannot construct a covariant quantization of this field with only positive norm states (this fact was proved by Allen in ). Also there appears an infrared divergence in the two-point function built from the positive norm states. For obtaining a covariant quantization and eliminating the infrared divergence the two sets of solutions (positive and negative norms states) are necessary . It has been also shown that the commutator function, which is calculated by these two different methods, is the same . Therefore there exists another possibility for defining the field operator, which satisfies the same commutation relation (locality condition). In contrast to the usual quantization, the field operator acts on the Krein space (positive and negative norms states). Let us show now that, if we use this new method of quantization for the free scalar field in Minkowski space, the ultraviolet divergence in the vacuum energy disappears and there is no need for use the “normal ordering” operation. In Krein QFT the quantum field is defined as follows
$$\varphi (x)=\frac{1}{\sqrt{2}}[\varphi _p(x)+\varphi _n(x)],$$
(8)
where
$$\varphi _p(x)=d^3\stackrel{}{k}[a(\stackrel{}{k})u_p(k,x)+a^{}(\stackrel{}{k})u_p^{}(k,x)],$$
$$\varphi _n(x)=d^3\stackrel{}{k}[b(\stackrel{}{k})u_n(k,x)+b^{}(\stackrel{}{k})u_n^{}(k,x)],$$
where $`a(\stackrel{}{k})`$ and $`b(\stackrel{}{k})`$ are two independent operators. The positive mode $`\varphi _p`$ is the scalar field as was used in the usual QFT. Creation and annihilation operators are constrained to obey the following commutation rules
$$[a(\stackrel{}{k}),a(\stackrel{}{k}^{})]=0,[a^{}(\stackrel{}{k}),a^{}(\stackrel{}{k}^{})]=0,,[a(\stackrel{}{k}),a^{}(\stackrel{}{k}^{})]=\delta (\stackrel{}{k}\stackrel{}{k}^{}),$$
(9)
$$[b(\stackrel{}{k}),b(\stackrel{}{k}^{})]=0,[b^{}(\stackrel{}{k}),b^{}(\stackrel{}{k}^{})]=0,,[b(\stackrel{}{k}),b^{}(\stackrel{}{k}^{})]=\delta (\stackrel{}{k}\stackrel{}{k}^{}),$$
(10)
$$[a(\stackrel{}{k}),b(\stackrel{}{k}^{})]=0,[a^{}(\stackrel{}{k}),b^{}(\stackrel{}{k}^{})]=0,,[a(\stackrel{}{k}),b^{}(\stackrel{}{k}^{})]=0,[a^{}(\stackrel{}{k}),b(\stackrel{}{k}^{})]=0.$$
(11)
The vacuum state $`0>`$ is then defined by
$$a^{}(\stackrel{}{k})0>=1_\stackrel{}{k}>;a(\stackrel{}{k})0>=0,\stackrel{}{k},$$
(12)
$$b^{}(\stackrel{}{k})0>=\overline{1}_\stackrel{}{k}>;b(\stackrel{}{k})0>=0,\stackrel{}{k},$$
(13)
$$b(\stackrel{}{k})1_\stackrel{}{k}>=0;a(\stackrel{}{k})\overline{1}_\stackrel{}{k}>=0,\stackrel{}{k},$$
(14)
where $`1_\stackrel{}{k}>`$ is called a one particle state and $`\overline{1}_\stackrel{}{k}>`$ is called a one “unparticle state”. These commutation relations, together with the normalization of the vacuum
$$<00>=1,$$
lead to positive (resp. negative) norms on the physical (resp. unphysical) sector:
$$<1_\stackrel{}{k}^{}1_\stackrel{}{k}>=\delta (\stackrel{}{k}\stackrel{}{k}^{}),<\overline{1}_\stackrel{}{k}^{}\overline{1}_\stackrel{}{k}>=\delta (\stackrel{}{k}\stackrel{}{k}^{}).$$
(15)
If we calculate the energy operator in terms of these Fourier modes, we have
$$H=d^3\stackrel{}{k}k^0[a^{}(\stackrel{}{k})a(\stackrel{}{k})+b^{}(\stackrel{}{k})b(\stackrel{}{k})+a^{}(\stackrel{}{k})b^{}(\stackrel{}{k})+a(\stackrel{}{k})b(\stackrel{}{k})].$$
(16)
This energy for the vacuum state is zero and it is not needed to use the “normal ordering” operation. It is also positive for any particles state or physical state $`N_\stackrel{}{k}>`$ (those built from repeated action of the $`a^{}(\stackrel{}{k})`$’s on the vacuum)
$$<N_\stackrel{}{k}^{}HN_\stackrel{}{k}^{}>=<N_\stackrel{}{k}^{}a^{}(\stackrel{}{k})a(\stackrel{}{k})N_\stackrel{}{k}^{}>k^0d^3\stackrel{}{k}0,0_\stackrel{}{k}^{}>0>.$$
We shall attempt to generalize this method to the interacting quantum field in the next section. At this stage we consider various Green’s functions fundamental to the interacting case. Within the framework of our approach, the two-point function is the imaginary part the usual Wightman two-point function, which is built from the positive norm states
$$𝒲(x,x^{})=<0\varphi (x)\varphi (x^{})0>=\frac{1}{2}[𝒲_p(x,x^{})+𝒲_n(x,x^{})]=i\mathrm{}𝒲_p(x,x^{}),$$
(17)
where $`𝒲_n=𝒲_p^{}`$. The commutator and anticommutator of the field are defined respectively by
$$iG(x,x^{})=<0[\varphi (x),\varphi (x^{})]0>=2i\mathrm{}𝒲(x,x^{})=2i\mathrm{}𝒲_p(x,x^{})=iG_p(x,x^{}),$$
(18)
$$G^1(x,x^{})=<0\{\varphi (x),\varphi (x^{})\}0>=0.$$
(19)
Retarded and advanced Green’s functions are defined respectively by
$$G^{ret}(x,x^{})=\theta (tt^{})G(x,x^{})=G_p^{ret}(x,x^{}),$$
(20)
$$G^{adv}(x,x^{})=\theta (t^{}t)G(x,x^{})=G_p^{adv}(x,x^{}).$$
(21)
The Schwinger commutator function, retarded and advanced Green’s functions is the same in the two formalism. The “Feynman” propagator or the Time-ordered product propagator is defined
$$iG_T(x,x^{})=<0T\varphi (x)\varphi (x^{})0>=\theta (tt^{})𝒲(x,x^{})+\theta (t^{}t)𝒲(x^{},x).$$
(22)
In this case we obtain
$$G_T(x,x^{})=\frac{1}{2}[G_F^p(x,x^{})+(G_F^p(x,x^{}))^{}]=\mathrm{}G_F^p(x,x^{}).$$
(23)
where the positive norm state is
$$G_F^p(x,x^{})=\frac{d^4k}{(2\pi )^4}e^{ik.(xx^{})}\stackrel{~}{G}^p(k)=\frac{d^4k}{(2\pi )^4}\frac{e^{ik.(xx^{})}}{k^2m^2+iϵ}.$$
(24)
Using the Bessel functions it is also written in the following form
$$G_T(x,x^{})=\mathrm{}G_F^p(x,x^{})=\frac{m^2}{8\pi }\theta (\sigma )\frac{J_1(\sqrt{2m^2\sigma })}{\sqrt{2m^2\sigma }}\frac{1}{8\pi }\delta (\sigma ),$$
(25)
where
$$G_F^p(x,x^{})=\frac{1}{8\pi }\delta (\sigma )+\frac{m^2}{8\pi }\theta (\sigma )\frac{J_1(\sqrt{2m^2\sigma })iN_1(\sqrt{2m^2\sigma })}{\sqrt{2m^2\sigma }}$$
$$\frac{im^2}{4\pi ^2}\theta (\sigma )\frac{K_1(\sqrt{2m^2\sigma })}{\sqrt{2m^2\sigma }},\sigma =\frac{1}{2}(xx^{})^2.$$
The ultraviolet singularity appears in the imaginary part of the Feynman propagator $`G_F^p(x,x)`$ (equation $`(9.52)`$ in ),
$$G_F^p(x,x)\underset{n4}{lim}\frac{2i}{(4\pi )^2}\frac{m^2}{(n4)}+G_F^{\text{finite}}(x,x),$$
where $`G_F^{\text{finite}}(x,x)`$ is finite as $`n4`$. Then the Green’s function $`G_T`$ is convergence in the ultraviolet limit
$$G_T(x,x)=\mathrm{}G_F^p(x,x)=\mathrm{}G_F^{\text{finite}}(x,x).$$
(26)
This Green’s function for every time-like separated pair $`(x,x^{})`$ is
$$\underset{xx^{}}{lim}G_T(x,x^{})=\frac{m^2}{16\pi },$$
(27)
and for space-like separated pair $`(x,x^{})`$ is zero
$$\underset{xx^{}}{lim}G_T(x,x^{})=0.$$
In the momentum space for this propagator we have
$$\stackrel{~}{G}(k)=\frac{1}{2}[\stackrel{~}{G}^p(k)+\stackrel{~}{G}^p(k)^{}]=\frac{1}{2}\left[\frac{1}{k^2m^2+iϵ}+\frac{1}{k^2m^2iϵ}\right]=PP\frac{1}{k^2m^2},$$
(28)
where $`PP`$ is the principal part symbol.
## 4 The interaction QFT
In the interaction case the S matrix elements, which describes the scattering of the i states into the f states ($`S_{fi}=<\text{out, f}|\text{in, i}>`$) are the most important quantities to be calculated. The S matrix elements can be written in terms of the time order product of the two free field operator $`(22)`$ by applying the reduction formulas, Wick’s theorem and time evolution operator . As this two point function is convergence in the ultraviolet and infrared limit $`(26)`$, this method may be renormalized automatically. Here the $`\lambda \varphi ^4`$ interaction field in Minkowski space is studied to the one-loop approximation.
The tree order S-matrix elements do not change when it is applied to the Krein QFT since the unphysical state disappear in the external line due to the conditions $`(11)`$ and $`(14)`$ and the internal propagator in the two cases are the same $`(28)`$.
In the one-loop approximation case, two primitive divergent integrals appear, which can be written in the following form
$$i\mathrm{\Sigma }^p=\frac{\lambda }{2}\frac{d^4k}{(2\pi )^4}\frac{1}{k^2m^2+iϵ}=\frac{\lambda }{2}\underset{xx^{}}{lim}G_F^p(x,x^{};m^2),$$
(29)
$$\mathrm{\Gamma }^4(s,t,u)=i\lambda +\mathrm{\Gamma }(s)+\mathrm{\Gamma }(t)+\mathrm{\Gamma }(u),$$
(30)
where $`s,t`$ and $`u`$ are the Mandelstam variable and
$$\mathrm{\Gamma }(s)=\frac{\lambda ^2}{2}_0^1𝑑l\frac{d^4k}{(2\pi )^4}\frac{1}{[k^2m^2+sl(1l)+iϵ]^2}.$$
(31)
If we define $`M^2=m^2sl(1l)`$, this integral can be written in terms of the first one
$$\mathrm{\Gamma }(s)=\frac{\lambda ^2}{2}_0^1𝑑l\frac{}{M^2}\frac{d^4k}{(2\pi )^4}\frac{1}{k^2M^2+iϵ}$$
$$=\frac{\lambda ^2}{2}_0^1𝑑l\frac{}{M^2}\underset{xx^{}}{lim}G_F^p(x,x^{};M^2),M^20.$$
By using the equation $`(26)`$ we see simply that if the Green function $`G_F^p`$ is replaced by the Green function $`G_T`$ the integrals is convergent.
Now the “self energy” graph is explicitly considered. By using the above discussion (equations $`(26)`$ and $`(27)`$) the self energy terms of the two point function in the new approach is given by
$$i\mathrm{\Sigma }=\frac{\lambda }{2}\underset{xx^{}}{lim}G_T(x,x^{};m^2)=\frac{\lambda m^2}{32\pi },m^20.$$
(32)
The positive norm states and the full propagator in the one loop correction are given respectively by
$$i\mathrm{\Delta }(k)=\frac{i}{k^2m^2\frac{\lambda }{32\pi }m^2+iϵ},PP\frac{i}{k^2m^2\frac{\lambda }{32\pi }m^2}.$$
(33)
Then in the one loop correction the mass is replaced by
$$m^2(\lambda )=m^2(0)[1+\frac{\lambda }{32\pi }+\mathrm{}\mathrm{}],$$
(34)
where $`m(0)=m`$. Due to the interaction, the effective mass of the particle $`m(\lambda )`$, which determines its response to an externally applied force, is certainly different from the mass of the particle without interaction $`m(0)`$. In this case, $`m(0)`$ and $`m(\lambda )`$ are both measurable.
Finally we calculate explicitly the transition amplitude of the state $`|q_1,q_2;\text{ in}>`$ to the state $`|p_1,p_2;\text{ out}>`$ for s-channel contribution in the one-loop approximation. It is given by
$$𝒯<p_1,p_2;\text{ out}|q_1,q_2;\text{ in}>_s=d^4y_1d^4y_2d^4x_1d^4x_2e^{ip_1.y_1+ip_2.y_2iq_1.x_1iq_2.x_2}$$
$$(\mathrm{}_{y_1}+m^2)(\mathrm{}_{y_2}+m^2)(\mathrm{}_{x_1}+m^2)(\mathrm{}_{x_2}+m^2)\frac{(i\lambda )^2}{2!}d^4z_1d^4z_2[iG_T(z_1z_2)]^2$$
$$G_T(y_1z_2)G_T(y_2z_1)G_T(x_1z_2)G_T(x_2z_2),$$
where the Feynman Green function $`G_F^p`$ is replaced by the Time-order product Green function $`G_T`$. We obtain
$$𝒯=\frac{\lambda ^2}{2}d^4z_1d^4z_2e^{i(p_1+p_2).z_1i(q_1+q_2).z_2}[G_T(z_1z_2)]^2$$
$$=\frac{\lambda ^2}{2}(2\pi )^4\delta ^4(p_1+p_2q_1q_2)d^4ze^{i(p_1+p_2).z}\left(\frac{m^2}{8\pi }\theta (z^2)\frac{J_1(\sqrt{m^2z^2})}{\sqrt{m^2z^2}}\frac{1}{4\pi }\delta (z^2)\right)^2,$$
(35)
where $`2\sigma =(z_1z_2)^2=z^2`$. The integral for the space-like separated pair $`(z_1,z_2)`$ is zero. That means the interaction between the intermediate states do not exist for a space-like separated pair. Then the causality or locality principle is preserved for the intermediate states. By using the equation $`(26)`$ we obtained that the integral for the light-like separated pair $`(z_1,z_2)`$ is also zero. That means an internal “particle”, which propagate in the intermediate states, is “massive” and it can not propagate on the light-cone.
The integral for the time-like separated pair $`(z_1,z_2)`$ is finite
$$d^4ze^{i(p_1+p_2).z}\left(\theta (z^2)\frac{J_1(\sqrt{m^2z^2})}{\sqrt{m^2z^2}}\right)^2=\text{finite}.$$
(36)
Therefore the transition amplitude is finite in the one-loop approximation.
## 5 Conclusion
We recall that the negative frequency solutions of the field equation are needed for quantizing in a correct way the minimally coupled scalar field in de Sitter space. Contrary to the Minkowski space, the elimination of de Sitter negative norm in the minimally coupled states breaks the de Sitter invariance. Then for restoring the de Sitter invariance, one needs to take into account the negative norm states i.e. the Krein space quantization. It provides a natural tool for renormalization technique . In this paper the $`\lambda \varphi ^4`$ theory in Minkowski space-time has been studied to the one-loop approximation in the Krein space quantization. It is found that the theory is automatically renormalized in this approximation. The main questions, which naturally arise, are: dose the natural renormalizeability preserve in the higher-loop expansion? does this construction affect the physical world ?
Acknowlegements: The author would like to thank S. Rouhani for very useful discussions and S. Teymourpoor and M. Oveisy for their interest in this work.
|
warning/0006/quant-ph0006049.html
|
ar5iv
|
text
|
# A Simple Algorithm for Local Conversion of Pure States
## 1 Introduction
Consider the case where two parties, Alice and Bob, share an entangled state of two $`N`$-level particles. M. A. Nielsen introduced in \[N\] an algorithm for converting one such pure bipartite state into another using only local operations and classical communication. He gave an explicit condition for when such a conversion is possible: We write the first state in Schmidt form
$$|\varphi =\underset{i=1}{\overset{N}{}}\sqrt{\alpha ^i}|i|i,$$
(1)
where the $`\alpha ^i`$ are non-negative real numbers satisfying $`_i\alpha ^i=1`$, and similarly for the target state
$$|\psi =\underset{i=1}{\overset{N}{}}\sqrt{\beta ^i}|i|i.$$
(2)
Then $`|\varphi `$ may be converted to $`|\psi `$ iff the vector $`(\beta ^i)`$ majorizes $`(\alpha ^i)`$, i.e.,
$$k:\underset{i=1}{\overset{k}{}}{}_{}{}^{}\beta _{}^{i}\underset{i=1}{\overset{k}{}}{}_{}{}^{}\alpha _{}^{i}$$
(3)
with equality for $`k=N`$; here $`{}_{}{}^{}\beta _{}^{i}`$ are the $`\beta ^i`$ arranged in descending order, and similarly for the $`{}_{}{}^{}\alpha _{}^{i}`$. This is again equivalent (theorem II.1.10 of \[Bh\]) to the existence of a doubly stochastic matrix $`D`$ such that $`\alpha =D\beta `$ (a doubly stochastic matrix has non-negative real entries, and all its rows and columns sum to one).
Nielsen’s algorithm uses several rounds of individual measurements and classical communication. Although it is known that such a sequence of operations can be replaced by one involving only a single measurement \[LP\], the proof of this is non-constructive and not easily applied to Nielsen’s algorithm. Our simpler algorithm may be useful for practical applications and for the analysis of local pure-state conversion in quantum cryptography \[B, JS\]. A similar result using a different method has been obtained earlier by Hardy \[H\].
## 2 Example
In this section we illustrate how the algorithm works by an example. We consider the case $`\beta ^T=(\beta ^i)^T=(3/5,3/10,1/10)`$ and $`\alpha ^T=(\alpha ^i)^T=(2/5,1/4,7/20)`$ (for typographic reasons, we write the transpose of the column vectors); note that $`\alpha `$ is not sorted: as we shall see, this doesn’t matter. We check that $`\beta \alpha `$. Using the algorithm from (the proof of) theorem II.1.10 in \[Bh\], we find a doubly stochastic matrix that maps $`\beta `$ to $`\alpha `$:
$$\left(\begin{array}{ccc}1/3& 2/3& 0\\ 1/6& 1/3& 1/2\\ 1/2& 0& 1/2\end{array}\right)\left(\begin{array}{c}3/5\\ 3/10\\ 1/10\end{array}\right)=\left(\begin{array}{c}2/5\\ 1/4\\ 7/20\end{array}\right)$$
(4)
From this matrix we now derive the set of unitary transformations, which turn out to be permutations. We start by finding a set of non-zero entries with no two entries in the same row or in the same column, i.e., corresponding to some permutation matrix. We first choose positions $`\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 1\end{array}\right)`$, corresponding to the permutation $`(12)`$. The smallest entry in the doubly stochastic matrix in these positions is $`1/6`$, so we subtract $`1/6`$ times the permutation matrix. Then we get
$$\left(\begin{array}{ccc}1/3& 1/2& 0\\ 0& 1/3& 1/2\\ 1/2& 0& 1/3\end{array}\right)$$
(5)
Again we look for a set of non-zero entries with no two in the same row or column; let’s pick the identity matrix this time. The smallest entry on the main diagonal of (5) is $`1/3`$, so we subtract $`1/3`$ times the identity matrix, and finally we are left with $`1/2`$ times a permutation matrix corresponding to the permutation $`(132)`$.
Using the above decomposition, we can now rewrite (4) as
$$\left(\begin{array}{ccc}3/10& 3/5& 3/10\\ 3/5& 3/10& 1/10\\ 1/10& 1/10& 3/5\end{array}\right)\left(\begin{array}{c}1/6\\ 1/3\\ 1/2\end{array}\right)=\left(\begin{array}{c}2/5\\ 1/4\\ 7/20\end{array}\right)$$
(6)
We may imagine that the columns in the matrix are indexed by the permutations we found above, i.e., $`(12)`$, $`()`$, and $`(132)`$, respectively. The entries of column $`\sigma `$ are then $`\beta `$ permuted by $`\sigma ^1`$; we permute the vector by shuffling the components: $`(\beta ^1,\beta ^2,\beta ^3).(123)=(\beta ^2,\beta ^3,\beta ^1)`$ (this corresponds to the last row because $`(132)^1=(123)`$). Thus, in the column corresponding to $`\sigma `$, the $`i`$th row has the entry $`\beta ^{\sigma ^1(i)}`$. The LHS vector in (6) is $`(p^\sigma )`$, where the $`p^\sigma `$ is the weight corresponding to $`\sigma `$ that we found above when decomposing the doubly stochastic matrix. The RHS vector is, of course, $`\alpha `$.
Now we need to find the POVM. To do this, we find three diagonal matrices $`A_\sigma `$, each defined by taking column $`\sigma `$ and dividing the $`i`$th entry (in the $`i`$th row) with $`\alpha ^i`$, and multiplying with $`p^\sigma `$. Thus
$`A_{(12)}`$ $`={\displaystyle \frac{1}{6}}\mathrm{diag}({\displaystyle \frac{3/10}{2/5}},{\displaystyle \frac{3/5}{1/4}},{\displaystyle \frac{1/10}{7/20}})`$ $`=\mathrm{diag}(1/8,2/5,1/21)`$ (7)
$`A_{()}`$ $`={\displaystyle \frac{1}{3}}\mathrm{diag}({\displaystyle \frac{3/5}{2/5}},{\displaystyle \frac{3/10}{1/4}},{\displaystyle \frac{1/10}{7/20}})`$ $`=\mathrm{diag}(1/2,2/5,2/21)`$ (8)
$`A_{(123)}`$ $`={\displaystyle \frac{1}{2}}\mathrm{diag}({\displaystyle \frac{3/10}{2/5}},{\displaystyle \frac{1/10}{1/4}},{\displaystyle \frac{3/5}{7/20}})`$ $`=\mathrm{diag}(3/8,1/5,6/7)`$ (9)
We observe that each $`A_\sigma `$ is a positive matrix, and that the sum of the $`A_\sigma `$ is the identity matrix. The $`A_\sigma `$ thus define a POVM. All we now need to see is what happens when the corresponding operations are applied to the state described by $`\alpha `$.
The POVM is performed locally by one of the parties, say Alice. In our example, the post-measurement state corresponding to the outcome $`\sigma `$ is a pure state with Schmidt coefficients obtained by normalizing the vector $`A_\sigma \alpha `$:
* $`(12)`$: $`(1/20,1/10,1/60)(3/10,3/5,1/10)`$
* $`()`$: $`(1/5,1/10,1/30)(3/5,3/10,1/10)`$
* $`(123)`$: $`(3/20,1/20,3/10)(3/10,1/10,3/5)`$
where the $``$ indicates the normalization. The post-measurement states are thus given by the columns of the matrix in equation (6). To reach the target state, Alice permutes the bases according to the permutation $`\sigma `$, and she communicates $`\sigma `$ to Bob who then performs the same permutation on his basis.
## 3 Formal Description and Proof
### 3.1 Description of the algorithm
Find the doubly stochastic matrix
Given vectors of Schmidt coefficients $`\alpha `$ and $`\beta `$ such that $`\beta \alpha `$, we first use the algorithm of \[Bh\], II.1.10, to find a doubly stochastic matrix $`D`$ such that $`\alpha =D\beta `$.
Decompose the doubly stochastic matrix
The Birkhoff-von Neumann theorem (\[Bh\], chapter 2) states that we can find permutations $`\sigma \mathrm{\Sigma }𝔖_N`$ and positive numbers $`p^\sigma `$ such that
$$D=\underset{\sigma \mathrm{\Sigma }}{}p^\sigma P_\sigma $$
(10)
where $`P_\sigma `$ is the permutation matrix associated to $`\sigma `$, i.e., it maps the fundamental vector $`e_i`$ (with a one in the $`i`$th place and zeros otherwise) to $`e_{\sigma (i)}`$.
To explicitly find the decomposition of the density matrix, we may use that many of the proofs of the Birkhoff-von Neumann theorem are constructive. The approach we sketch here is similar to the proof in \[P\].
It will be useful to define an arrangement (or diagonal) of a $`n\times n`$ matrix $`M=(m_j^i)`$ as a set of entries $`\{m_i^{\sigma (i)}\}`$ for some permutation $`\sigma 𝔖_n`$, i.e., where no two entries are taken from the same row or the same column. It follows from the König-Frobenius theorem (\[Bh\], chapter 2) that any doubly stochastic matrix has an arrangement with all entries non-zero.
There is a polynomial-time algorithm for finding such arrangements in a matrix: Consider the matrix $`\stackrel{~}{D}`$ which is $`D`$ with each non-zero entry replaced by $`1`$. Let $`R=\{r_1,\mathrm{},r_n\}`$ be the set of row indices of $`\stackrel{~}{D}`$ and $`C=\{c_1,\mathrm{},c_n\}`$ the set of column indices. Then $`\stackrel{~}{D}`$ defines a balanced bipartite graph $`G`$ with vertices $`RC`$ (disjoint union): we have an edge from $`r_i`$ to $`c_j`$ iff $`\stackrel{~}{D}_j^i=1`$. By the König-Frobenius theorem, $`G`$ has no isolated nodes. We can thus easily find a perfect matching in the graph; there are polynomial-time algorithms for doing this, see e.g., chapter 3, section 3 of \[P\]. A perfect matching defines a unique permutation $`\sigma 𝔖_n`$ through its edges $`r_{\sigma (i)}c_i`$, $`i=1,\mathrm{},n`$. It is easy to see that the matching also defines an arrangement of $`D`$ with all entries non-zero.
Constructing the POVM
Now we assume that we have the decomposition (10). Hence
$$i:\underset{\sigma }{}\beta ^{\sigma ^1(i)}p^\sigma =\alpha ^i$$
(11)
Indeed, $`\alpha ^i=(D\beta )^i=[(_\sigma p^\sigma P_\sigma )\beta ]^i=_\sigma p^\sigma \beta ^{\sigma ^1(i)}`$.
Next, we define matrices
$$A_\sigma =p^\sigma \mathrm{diag}_i(\frac{\beta ^{\sigma ^1(i)}}{\alpha ^i})$$
(12)
where $`\mathrm{diag}_i(f(i))`$ denotes a diagonal matrix whose $`(i,i)`$ entry is $`f(i)`$.
The $`\{A_\sigma \sigma \mathrm{\Sigma }\}`$ define a POVM, since each $`A_\sigma `$ is a positive matrix, and
$`{\displaystyle \underset{\sigma }{}}A_\sigma `$ $`=`$ $`{\displaystyle \underset{\sigma }{}}p^\sigma \mathrm{diag}_i\left({\displaystyle \frac{\beta ^{\sigma ^1(i)}}{\alpha ^i}}\right)`$
$`=`$ $`\mathrm{diag}_i\left({\displaystyle \frac{1}{\alpha ^i}}{\displaystyle \underset{\sigma }{}}p^\sigma \beta ^{\sigma ^1(i)}\right)`$
$`=`$ $`\mathrm{diag}_i1`$
Furthermore,
$$\mathrm{tr}(A_\sigma \mathrm{diag}_i(\alpha ^i))=p^\sigma $$
(13)
so that $`p^\sigma `$ is the probability of outcome $`\sigma `$.
The measurement
Now we turn to the measurement itself. When Alice performs the measurement on her side, we need only consider her reduced density matrix, $`\rho `$, which has eigenvalues $`\alpha ^i`$. We choose the operations corresponding to the POVM $`A_\sigma `$ such that, if the outcome of the measurement is $`\sigma `$, then the reduced density matrix after the measurement is
$$\rho _\sigma =\frac{1}{p^\sigma }\sqrt{A_\sigma }\rho \sqrt{A_\sigma }=\frac{1}{p^\sigma }A_\sigma \rho .$$
(14)
The eigenvalues of $`\rho _\sigma `$ are $`\beta ^{\sigma ^1(i)}`$, $`i=1,\mathrm{},N`$, which sum to one. Alice then performs the permutation $`\sigma `$ on her side, and communicates $`\sigma `$ to Bob so that he can do the same: since $`_i\beta ^{\sigma ^1(i)}|i|i=_i\beta ^i|\sigma (i)|\sigma (i)`$, the operation should map $`|\sigma (i)`$ to $`|i`$ for all $`i`$. After the unitary operations, the particles are in the pure state $`|\psi `$ (2) with Schmidt coefficients $`\beta ^i`$.
### 3.2 The converse
We now turn to the converse: given $`\alpha `$ and $`\beta `$ and a POVM $`\{A_\sigma \sigma \mathrm{\Sigma }\}`$ with $`\mathrm{\Sigma }𝔖_n`$ such that for all $`\sigma \mathrm{\Sigma }`$,
$$(A_\sigma \alpha ).\sigma =p^\sigma \beta ,$$
(15)
for some positive constants $`p^\sigma `$, we show that $`\beta \alpha `$ (of course, this also follows directly from Nielsen’s theorem \[N\]). We assume that each $`A_\sigma `$ is given by a diagonal matrix.
Define a matrix $`\mathrm{\Gamma }=(\gamma _j^i)`$ by
$$\gamma _j^i=\underset{\{\sigma \sigma (j)=i\}}{}p^\sigma $$
(16)
Then we have
$`j:{\displaystyle \underset{i}{}}\gamma _j^i`$ $`=`$ $`{\displaystyle \underset{i}{}}{\displaystyle \underset{\{\sigma \sigma (j)=i\}}{}}p^\sigma `$
$`=`$ $`{\displaystyle \underset{\sigma \mathrm{\Sigma }}{}}p^\sigma =1;`$
$`i:{\displaystyle \underset{j}{}}\gamma _j^i`$ $`=`$ $`{\displaystyle \underset{j}{}}{\displaystyle \underset{\{\sigma \sigma (j)=i\}}{}}p^\sigma `$
$`=`$ $`{\displaystyle \underset{\sigma \mathrm{\Sigma }}{}}p^\sigma =1,`$
so $`\mathrm{\Gamma }`$ is doubly stochastic. Now equation (15) implies $`A_\sigma \alpha =p^\sigma \beta .(\sigma ^1)`$, which again implies that $`A_\sigma `$ must be of the form (12). Finally,
$`i:(\mathrm{\Gamma }\beta )^i`$ $`=`$ $`{\displaystyle \underset{j}{}}\gamma _j^i\beta ^j`$ (17)
$`=`$ $`{\displaystyle \underset{j}{}}{\displaystyle \underset{\{\sigma \sigma (j)=i\}}{}}p^\sigma \beta ^{\sigma ^1(i)}`$
$`=`$ $`{\displaystyle \underset{\sigma \mathrm{\Sigma }}{}}p^\sigma \beta ^{\sigma ^1(i)}`$
$`=`$ $`\alpha ^i.`$
Thus $`\mathrm{\Gamma }\beta =\alpha `$, which means that $`\beta \alpha `$.
## 4 Acknowledgment
This work was motivated by a discussion in the Quantum Dynamics group at Royal Holloway between N. Lütkenhaus and the authors. This work is supported by the UK Engineering and Physical Sciences Research Council (EPSRC).
|
warning/0006/hep-lat0006007.html
|
ar5iv
|
text
|
# I. INTRODUCTION
## I. INTRODUCTION
The dual superconductivity of the vacuum in gauge theories to explain color confinement has been proposed since long time by G. ’t Hooft and S. Mandelstam . These authors proposed that the confining vacuum behaves as a coherent state of color magnetic monopoles. In other words the confining vacuum is a magnetic (dual) superconductor. This fascinating proposal offers a picture of confinement whose physics can be clearly extracted. Indeed, the dual Meissner effect causes the formation of chromoelectric flux tubes between chromoelectric charges leading to a linear confining potential.
Following Ref. let us consider gauge theories without matter fields. In order to realize gauge field configurations which describe magnetic monopoles we need a scalar Higgs field . In the ’t Hooft’s scheme the role of the scalar field is played by any operator which transforms in the adjoint representation of the gauge group. Let $`X(x)`$ be an operator in the adjoint representation, then one fixes the gauge by diagonalizing $`X(x)`$ at each point. This choice does not fix completely the gauge, for it leaves as residual invariance group the maximal Abelian (Cartan) subgroup of the gauge group. This procedure is known as Abelian projection . The world line of the monopoles can be identified as the lines where two eigenvalues of the operator $`X(x)`$ are equal. Thus, the dual superconductor idea is realized if these Abelian monopole condense. Due to the gauge invariance we expect that the monopole condensation should manifest irrespective to the gauge fixing. In other words all the Abelian projections are physically equivalent. However, it is conceivable that the dual superconductor scenario could manifest clearly with a clever choice of the operator $`X(x)`$. It is remarkable that, if one adopts the so called maximally Abelian projection , then it seems that the Abelian projected links retain the information relevant to the confinement .
It turns out that the Abelian projection can be implemented on the lattice , so that one can analyze the dynamics of the Abelian projected gauge fields by means of non perturbative numerical simulations. Indeed, the first direct evidence of the dual Abrikosov vortex joining two static quark-antiquark pair has been obtained in lattice simulations of gauge theories . In particular in Ref. we considered the pure gauge $`SU(2)`$ lattice theory and found evidence of the dual Meissner effect both in the maximally Abelian gauge and without gauge fixing. Moreover we showed that the London penetration length is a physical gauge invariant quantity.
An alternative and more direct method to detect the dual superconductivity relies upon the very general assumption that the dual superconductivity of the ground state is realized if there is condensation of Abelian monopoles. Thus, according to Ref. it suffices to measure a disorder parameter defined as the vacuum expectation value of a nonlocal operator with non zero magnetic charge and non vanishing vacuum expectation value in the confined phase. However, in the case of non Abelian gauge theories, the disorder parameter is expected to break a non Abelian symmetry, while the dual superconductivity is realized by condensation of Abelian monopoles. As we have already argued, the Abelian monopole charge can be associated to each operator in the adjoint representation by the so-called Abelian projection . Indeed, the authors of Ref. introduced on the lattice a disorder parameter describing condensation of monopoles within a particular Abelian projection. On the other hand, recent results show that the Abelian monopoles defined through several Abelian projection condense, suggesting that the monopole condensation does not depend on the adjoint operator used in the Abelian projection procedure. This is in accordance with the theoretical expectation that monopole condensation should occur irrespective of the gauge fixing procedure. However, a gauge invariant evidence of the Abelian monopole condensation is still lacking.
The aim of the present paper is to investigate the Abelian monopole condensation in pure lattice gauge SU(2) and SU(3) theories in a gauge-invariant way . To do this we introduce a disorder parameter defined in terms of a gauge-invariant thermal partition functional in presence of an external background field.
The plan of the paper is as follows. In Section II we introduce the thermal partition functional, built up using the lattice Schrödinger functional . In Section III we study the Abelian monopole condensation for finite temperature SU(2) lattice gauge theory. Section IV is devoted to the case of SU(3) gauge theory at finite temperature, where, according to the choice of the Abelian subgroup, different kinds of Abelian monopoles can be defined. Our conclusions are drawn in Section V.
## II. THE THERMAL PARTITION FUNCTIONAL
To investigate the dynamics of the vacuum at zero temperature we introduced the gauge-invariant effective action for external static (i.e. time-independent) background field defined by means of the lattice Schrödinger functional:
$$𝒵[U_\mu ^{\mathrm{ext}}]=𝒟Ue^{S_W},$$
(2.1)
where $`S_W`$ is the standard Wilson action. The functional integration is extended over links on a lattice with the hypertorus geometry and satisfying the constraints
$$U_\mu (x)|_{x_4=0}=U_\mu ^{\mathrm{ext}}(\stackrel{}{x}).$$
(2.2)
In Equations (2.1) and (2.2) $`U_\mu ^{\mathrm{ext}}(\stackrel{}{x})`$ is the lattice version of the external continuum gauge field $`\stackrel{}{A}^{\mathrm{ext}}(\stackrel{}{x})=\stackrel{}{A}_a^{\mathrm{ext}}(\stackrel{}{x})\lambda _a/2`$:
$$U_\mu ^{\text{ext}}(\stackrel{}{x})=\mathrm{P}\mathrm{exp}\left\{iag_0^1𝑑tA_{a,\mu }^{\text{ext}}(\stackrel{}{x}+t\widehat{\mu })\frac{\lambda _a}{2}\right\},$$
(2.3)
where $`\mathrm{P}`$ is the path-ordering operator and $`g`$ the gauge coupling constant.
The lattice effective action for the external static background field $`\stackrel{}{A}^{\mathrm{ext}}(\stackrel{}{x})`$ is given by
$$\mathrm{\Gamma }[\stackrel{}{A}^{\mathrm{ext}}]=\frac{1}{L_4}\mathrm{ln}\left\{\frac{𝒵[\stackrel{}{A}^{\mathrm{ext}}]}{𝒵(0)}\right\},$$
(2.4)
where $`L_4`$ is the extension in Euclidean time and $`𝒵(0)`$ is the lattice Schrödinger functional, Eq. (2.1), without the external background field ($`U_\mu ^{\mathrm{ext}}=\mathrm{𝟏}`$). It can be shown that in the continuum limit $`\mathrm{\Gamma }[\stackrel{}{A}^{\mathrm{ext}}]`$ is the vacuum energy in presence of the background field $`\stackrel{}{A}^{\mathrm{ext}}(\stackrel{}{x})`$.
We want now to extend our definition of lattice effective action to gauge systems at finite temperature. In this case the relevant quantity is the thermal partition function. In the continuum we have:
$$\text{Tr}\left[e^{\beta _TH}\right]=𝒟\stackrel{}{A}\stackrel{}{A}\left|e^{\beta _TH}𝒫\right|\stackrel{}{A},$$
(2.5)
where $`\beta _T`$ is the inverse of the physical temperature, $`H`$ is the Hamiltonian, and $`𝒫`$ projects onto the physical states. As is well known, the thermal partition function can be written as :
$$\text{Tr}\left[e^{\beta _TH}\right]=_{A_\mu (\beta _T,\stackrel{}{x})=A_\mu (0,\stackrel{}{x})}𝒟A_\mu (x_4,\stackrel{}{x})e^{_0^{\beta _T}𝑑x_4{\scriptscriptstyle d^3\stackrel{}{x}_{YM}(\stackrel{}{x},x_4)}}.$$
(2.6)
On the lattice we have:
$$\text{Tr}\left[e^{\beta _TH}\right]=_{U_\mu (\beta _T,\stackrel{}{x})=U_\mu (0,\stackrel{}{x})=U_\mu (\stackrel{}{x})}𝒟U_\mu (x_4,\stackrel{}{x})e^{S_W}.$$
(2.7)
Comparing Eq. (2.7) with Eqs. (2.1) and (2.2), we get:
$$\text{Tr}\left[e^{\beta _TH}\right]=𝒟U_\mu (\stackrel{}{x})𝒵[U_\mu (\stackrel{}{x})],$$
(2.8)
where $`𝒵[U_\mu (\stackrel{}{x})]`$ is the Schrödinger functional Eq. (2.1) defined on a lattice with $`L_4=\beta _T`$, with “external” links $`U_\mu (\stackrel{}{x})`$ at $`x_4=0`$.
We are interested in the thermal partition function in presence of a given static background field $`\stackrel{}{A}^{\mathrm{ext}}(\stackrel{}{x})`$. In the continuum this can be obtained by splitting the gauge field into the background field $`\stackrel{}{A}^{\mathrm{ext}}(\stackrel{}{x})`$ and the fluctuating fields $`\eta (x)`$. So that we could write formally for the thermal partition function $`𝒵_T[\stackrel{}{A}^{\mathrm{ext}}]`$:
$$𝒵_T[\stackrel{}{A}^{\mathrm{ext}}]=𝒟\stackrel{}{\eta }\stackrel{}{A}^{\mathrm{ext}},\stackrel{}{\eta }\left|e^{\beta _TH}𝒫\right|\stackrel{}{A}^{\mathrm{ext}},\stackrel{}{\eta }.$$
(2.9)
The lattice implementation of Eq. (2.9) can be obtained from Eq. (2.7) if we write
$$U_k(\beta _T,\stackrel{}{x})=U_k(0,\stackrel{}{x})=U_k^{\text{ext}}(\stackrel{}{x})\stackrel{~}{U}_k(\stackrel{}{x}),$$
(2.10)
where $`U_k^{\text{ext}}(\stackrel{}{x})`$ is given by Eq. (2.3) and the $`\stackrel{~}{U}_k(\stackrel{}{x})`$’s are the fluctuating links. Thus we get
$$𝒵_T[\stackrel{}{A}^{\mathrm{ext}}]=𝒟\stackrel{~}{U}_k(\stackrel{}{x})𝒟U_4(\stackrel{}{x})𝒵[U_k^{\text{ext}}(\stackrel{}{x}),\stackrel{~}{U}_k(\stackrel{}{x})],$$
(2.11)
where we integrate over the fluctuating links $`\stackrel{~}{U}_k(\stackrel{}{x})`$, while the $`U_k^{\text{ext}}`$ links are fixed. Note that in Eq. (2.11) only the spatial links belonging to the hyperplane $`x_4=0`$ are written as the product of the external link $`U_k^{\text{ext}}(\stackrel{}{x})`$ and the fluctuating links $`\stackrel{~}{U}_k(\stackrel{}{x})`$. The temporal links $`U_4(x_4=0,\stackrel{}{x})`$ are left freely fluctuating. It follows that the temporal links $`U_4(x)`$ satisfy the usual periodic boundary conditions. We stress that the periodic boundary conditions in the temporal direction are crucial to retain the physical interpretation that the functional $`𝒵_T[\stackrel{}{A}^{\text{ext}}]`$ is a thermal partition function. In the following the spatial links belonging to the time-slice $`x_4=0`$ will be called “frozen links”, while the remainder will be the “dynamical links”.
From the physical point of view we are considering the gauge system at finite temperature in interaction with a fixed external background field. As a consequence, in the Wilson action $`S_W`$ we keep only the plaquettes built up with the dynamical links or with dynamical and frozen links. With these limitations it is easy to see that in Eq. (2.11) we have
$$𝒵[U_k^{\text{ext}}(\stackrel{}{x}),\stackrel{~}{U}_k(\stackrel{}{x})]=𝒵\left[U_k^{\text{ext}}(\stackrel{}{x})\right].$$
(2.12)
Indeed, let us consider an arbitrary frozen link $`U_k^{\text{ext}}(\stackrel{}{x})\stackrel{~}{U}_k(\stackrel{}{x})`$. This link enters in the modified Wilson action by means of the plaquette:
$$P_{k4}(x_4=0,\stackrel{}{x})=\text{Tr}\left\{U_k^{\text{ext}}(\stackrel{}{x})\stackrel{~}{U}_k(\stackrel{}{x})U_4(0,\stackrel{}{x}+\widehat{k})U_k^{}(1,\stackrel{}{x})U_4^{}(0,\stackrel{}{x})\right\}.$$
(2.13)
Now we observe that the link $`U_4(0,\stackrel{}{x}+\widehat{k})`$ in Eq. (2.13) is a dynamical one, i.e. we are integrating over it. So that, by using the invariance of the Haar measure we obtain
$$P_{k4}(x_4=0,\stackrel{}{x})=\text{Tr}\left\{U_k^{\text{ext}}(\stackrel{}{x})U_4(0,\stackrel{}{x}+\widehat{k})U_k^{}(1,\stackrel{}{x})U_4^{}(0,\stackrel{}{x})\right\}.$$
(2.14)
It is evident that Eq. (2.14) in turns implies Eq. (2.12). Then, we see that in Eq. (2.11) the integration over the fluctuating links $`\stackrel{~}{U}(\stackrel{}{x})`$ gives an irrelevant multiplicative constant. So that we have:
$$𝒵_T\left[\stackrel{}{A}^{\text{ext}}\right]=_{U_k(\beta _T,\stackrel{}{x})=U_k(0,\stackrel{}{x})=U_k^{\text{ext}}(\stackrel{}{x})}𝒟Ue^{S_W},$$
(2.15)
where the integrations are over the dynamical links with periodic boundary conditions in the time direction. As concerns the boundary conditions at the spatial boundaries, we keep the fixed boundary conditions $`U_k(\stackrel{}{x},x_4)=U_k^{\text{ext}}(\stackrel{}{x})`$ used in the Schrödinger functional Eq.(2.1). Thus we see that, if we send the physical temperature to zero, then the thermal functional Eq. (2.15) reduces to the zero-temperature Schrödinger functional Eq. (2.1) with the constraints $`U_k(x)|_{x_4=0}=U_k^{\mathrm{ext}}(\stackrel{}{x})`$ instead of Eq. (2.2). In our previous study we checked that in the thermodynamic limit both conditions agree as concerns the zero-temperature effective action Eq. (2.4).
## III. ABELIAN MONOPOLE CONDENSATION: SU(2)
Let us consider the SU(2) pure gauge theory at finite temperature. We are interested in the thermal partition function Eq. (2.15) in presence of an Abelian monopole field. In the case of SU(2) gauge theory the maximal Abelian group is an Abelian U(1) group. Thus, in the continuum the Abelian monopole field turns out to be:
$$g\stackrel{}{b}^a(\stackrel{}{x})=\delta ^{a,3}\frac{n_{\mathrm{mon}}}{2}\frac{\stackrel{}{x}\times \stackrel{}{n}}{|\stackrel{}{x}|(|\stackrel{}{x}|\stackrel{}{x}\stackrel{}{n})}.$$
(3.1)
where $`\stackrel{}{n}`$ is the direction of the Dirac string and, according to the Dirac quantization condition, $`n_{\text{mon}}`$ is an integer. The lattice links corresponding to the Abelian monopole field Eq. (3.1) can be readily obtained as:
$$U_k^{\text{ext}}(\stackrel{}{x})=\text{P}\mathrm{exp}\left\{ig_0^1𝑑t\frac{\sigma _a}{2}b_k^a(\stackrel{}{x}+t\widehat{x}_k)\right\},$$
(3.2)
where the $`\sigma _a`$’s are the Pauli matrices. By choosing $`\stackrel{}{n}=x_3`$ we get:
$$\begin{array}{cc}\hfill U_{1,2}^{\text{ext}}(\stackrel{}{x})& =\mathrm{cos}[\theta _{1,2}(\stackrel{}{x})]+i\sigma _3\mathrm{sin}[\theta _{1,2}(\stackrel{}{x})],\hfill \\ \hfill U_3^{\text{ext}}(\stackrel{}{x})& =\mathrm{𝟏},\hfill \end{array}$$
(3.3)
with
$$\begin{array}{cc}\hfill \theta _1(\stackrel{}{x})& =\frac{n_{\text{mon}}}{4}\frac{(x_2X_2)}{|\stackrel{}{x}_{\text{mon}}|}\frac{1}{|\stackrel{}{x}_{\text{mon}}|(x_3X_3)},\hfill \\ \hfill \theta _2(\stackrel{}{x})& =+\frac{n_{\text{mon}}}{4}\frac{(x_1X_1)}{|\stackrel{}{x}_{\text{mon}}|}\frac{1}{|\stackrel{}{x}_{\text{mon}}|(x_3X_3)}.\hfill \end{array}$$
(3.4)
In Equation (3.4) $`(X_1,X_2,X_3)`$ are the monopole coordinates and $`\stackrel{}{x}_{\text{mon}}=(\stackrel{}{x}\stackrel{}{X})`$. In the numerical simulations we put the lattice Dirac monopole at the center of the time slice $`x_4=0`$. To avoid the singularity due to the Dirac string we locate the monopole between two neighboring sites. We have checked that the numerical results are not too sensitive to the precise position of the magnetic monopole.
According to the discussion in the previous Section we are interested in the thermal partition function $`𝒵_T[\stackrel{}{A}^{\text{ext}}]`$ given by Eq. (2.15). Note that we do not need to fix the gauge due to the gauge invariance of the thermal partition functional against gauge transformations of the external background field. On the lattice the physical temperature $`T_{\text{phys}}`$ is given by
$$\frac{1}{T_{\text{phys}}}=\beta _T=L_t,$$
(3.5)
where $`L_t=L_4`$ is the lattice linear extension in the time direction. In order to approximate the thermodynamic limit the spatial extension $`L_s`$ should satisfy
$$L_sL_t.$$
(3.6)
To this end we performed our numerical simulations on lattices such that
$$\frac{L_t}{L_s}4.$$
(3.7)
In the numerical simulations we impose periodic boundary conditions in the time direction. As already discussed, at the spatial boundaries the links are fixed according to Eq. (3.3). This last condition corresponds to the requirement that the fluctuations over the background field vanish at infinity.
Following the suggestion of Ref. we introduce the gauge-invariant disorder parameter for confinement
$$\mu =e^{F_{\text{mon}}/T_{\text{phys}}}=\frac{𝒵_T[n_{\text{mon}}]}{𝒵_T[0]},$$
(3.8)
where $`𝒵_T[0]`$ is the thermal partition function without monopole field (i.e. with $`n_{\text{mon}}=0`$).
From Eq. (3.8) it is clear that $`F_{\text{mon}}`$ is the free energy to create an Abelian monopole. If there is monopole condensation, then $`F_{\text{mon}}=0`$ and $`\mu =1`$. To avoid the problem of measuring a partition function we focus on the derivative of the monopole free energy:
$$F_{\text{mon}}^{}=\frac{}{\beta }F_{\text{mon}}.$$
(3.9)
It is straightforward to see that $`F_{\text{mon}}^{}`$ is given by the difference between the average plaquette
$$F_{\text{mon}}^{}=V\left[<P>_{n_{\text{mon}}=0}<P>_{n_{\text{mon}}0}\right],$$
(3.10)
where $`V`$ is the spatial volume.
We use the over-relaxed heat-bath algorithm to update the gauge configurations. Simulations have been performed by means of the APE100/Quadrics computer facility in Bari. Since we are measuring a local quantity such as the plaquette, a low statistics (from 2000 up to 12000 configurations) is required in order to get a good estimation of $`F_{\mathrm{mon}}^{}`$.
In Figure 1 we display the derivative of the monopole free energy versus $`\beta `$ for $`n_{\text{mon}}=10`$ on a lattice with $`L_t=4`$ and $`L_s=24`$. We see that $`F_{\text{mon}}^{}`$ vanishes at strong coupling and displays a rather sharp peak near $`\beta 2.13`$. We expect that the peak corresponds to the finite temperature deconfinement transition. In Figure 1 we also display the absolute value of the Polyakov loop:
$$|P|=<|\frac{1}{V}\underset{\stackrel{}{x}}{}\frac{1}{2}\mathrm{Tr}[\underset{t=1}{\overset{L_t}{}}U_4(\stackrel{}{x},t)]|>,$$
(3.11)
and, indeed, we see that the peak corresponds to the rise of Polyakov loop.
In the weak coupling region the plateau in $`F_{\text{mon}}^{}`$ indicates that the monopole free energy tends to the classical monopole action which behaves linearly in $`\beta `$. To see this, we observe that deeply in the weak coupling region the lattice action should reduce to the classical action. In the naive continuum limit the classical action reads :
$$S_{\text{class}}=\frac{1}{2}_0^{\beta _T}𝑑x_4d^3\stackrel{}{x}\stackrel{}{B}^a(\stackrel{}{x})\stackrel{}{B}^a(\stackrel{}{x})$$
(3.12)
where $`\stackrel{}{B}^a(\stackrel{}{x})`$ is the classical Abelian monopole magnetic field. Introducing an ultraviolet cutoff $`\mathrm{\Lambda }=\alpha /a`$, with $`\alpha `$ a constant and $`a`$ the lattice spacing, and performing in Eq. (3.12) the spatial integral over the volume $`V=L_s^3`$, we get:
$$S_{\text{class}}\frac{\pi \alpha \beta }{8T_{\mathrm{phys}}}n_{\text{mon}}^2+\text{O}(1/L_sa).$$
(3.13)
So that in the weak coupling region we have :
$$F_{\mathrm{mon}}^{}\frac{\pi }{8}\alpha n_{\text{mon}}^2.$$
(3.14)
From Figure 1 we see that Eq. (3.14) with $`\alpha 1.2`$ ( dashed line) describes quite well the numerical data in the relevant region.
In order to determine the critical parameters and the order of the transition, we need to perform the finite size scaling analysis. We plan to do this in a future work. In this paper we restrict ourself to a preliminary qualitative analysis. In Figure 2 we compare the derivative of the monopole free energy on lattices with $`L_t=4`$ and $`L_s=24,48`$. We see that in the strong coupling $`F_{\mathrm{mon}}^{}`$ agrees for the two lattices. On the other hand, in the weak coupling region the different values of the plateaus can be ascribed to finite volume effects. In the critical region we see that the peak increases.
With the aim of obtaining the disorder parameter $`\mu `$ (Eq. (3.8)) we perform the numerical integration of the monopole free energy derivative
$$F_{\text{mon}}(\beta )=_0^\beta 𝑑\beta ^{}F_{\text{mon}}^{}(\beta ^{}).$$
(3.15)
In Figure 3 we show the disorder parameter $`\mu `$ versus $`\beta `$ for lattices with $`L_t=4`$ and $`L_s=24,48`$. We see clearly that $`\mu =1`$ in the confined phase. In other words the monopoles condense in the vacuum. On the other hand, it seems that $`\mu 0`$ in the thermodynamic limit when $`\beta `$ reaches the critical value . Indeed, by increasing the spatial volume of the lattice, the disorder parameter $`\mu `$ decreases faster toward zero. Moreover we see that the finite volume behavior of our disorder parameter is consistent with a second order deconfinement phase transition.
It worthwhile to comment on the finite volume effects. As a matter of fact, it appears that, even though the spatial volume of our larger lattice looks enormous, we gain a rather small increase in the peak value of the monopole free energy derivative. This can be understood by observing that, due to our peculiar conditions at the spatial boundaries, the dynamical volume is smaller than the geometrical one. Moreover, it is well known that the fixed boundary conditions for the gauge fields lead to more severe finite volume effects with respect to the usual periodic boundary conditions. So that, to reach the thermodynamic limit we must simulate our gauge system on lattices with very large spatial volumes. We stress again that the precise determination of the critical parameters requires a finite size scaling which will be presented elsewhere.
## IV. ABELIAN MONOPOLE CONDENSATION: SU(3)
In the case of SU(3) gauge theory, the maximal Abelian group is U(1)$`\times `$U(1). Therefore we have two different types of Abelian monopole. Let us consider, firstly, the Abelian monopole field given by Eq. (3.1), which we call the $`T_3`$ Abelian monopole. The lattice links are given by
$$U_{1,2}^{\text{ext}}(\stackrel{}{x})=\left[\begin{array}{ccc}e^{i\theta _{1,2}(\stackrel{}{x})}& 0& 0\\ 0& e^{i\theta _{1,2}(\stackrel{}{x})}& 0\\ 0& 0& 1\end{array}\right],$$
(4.1)
with $`\theta _{1,2}(\stackrel{}{x})`$ defined in Eq. (3.4). The second type of independent Abelian monopole can be obtained by considering the diagonal generator $`T_8`$. In this case we have the $`T_8`$ Abelian monopole:
$$U_{1,2}^{\text{ext}}(\stackrel{}{x})=\left[\begin{array}{ccc}e^{i\theta _{1,2}(\stackrel{}{x})}& 0& 0\\ 0& e^{i\theta _{1,2}(\stackrel{}{x})}& 0\\ 0& 0& e^{2i\theta _{1,2}(\stackrel{}{x})}\end{array}\right],$$
(4.2)
with
$$\begin{array}{cc}\hfill \theta _1(\stackrel{}{x})& =\frac{1}{\sqrt{3}}\left[\frac{n_{\text{mon}}}{4}\frac{(x_2X_2)}{|\stackrel{}{x}_{\text{mon}}|}\frac{1}{|\stackrel{}{x}_{\text{mon}}|(x_3X_3)}\right],\hfill \\ \hfill \theta _2(\stackrel{}{x})& =\frac{1}{\sqrt{3}}\left[+\frac{n_{\text{mon}}}{4}\frac{(x_1X_1)}{|\stackrel{}{x}_{\text{mon}}|}\frac{1}{|\stackrel{}{x}_{\text{mon}}|(x_3X_3)}\right].\hfill \end{array}$$
(4.3)
Obviously, the lattice links Eq. (4.2) corresponds now to the continuum gauge field
$$g\stackrel{}{b}^a(\stackrel{}{x})=\delta ^{a,8}\frac{n_{\mathrm{mon}}}{2}\frac{\stackrel{}{x}\times \stackrel{}{n}}{|\stackrel{}{x}|(|\stackrel{}{x}|\stackrel{}{x}\stackrel{}{n})}.$$
(4.4)
The other Abelian monopoles can be generated by considering the linear combination of the $`T_3`$ and $`T_8`$ generators. In particular we have considered the $`T_{3a}`$ Abelian monopole corresponding to the following linear combination of $`\lambda _3/2`$ and $`\lambda _8/2`$:
$$T_{3a}=\frac{1}{2}\frac{\lambda _3}{2}+\frac{\sqrt{3}}{2}\frac{\lambda _8}{2}=\left[\begin{array}{ccc}0& 0& 0\\ 0& \frac{1}{2}& 0\\ 0& 0& \frac{1}{2}\end{array}\right].$$
(4.5)
In Figure 4 we compare the free energy monopole derivative for the $`T_3`$, $`T_{3a}`$ and $`T_8`$ Abelian monopoles for the lattice with $`L_t=4`$ and $`L_s=32`$. We see that the $`T_3`$ and $`T_{3a}`$ Abelian monopoles agree within statistical errors in the whole range of $`\beta `$. On the other hand the $`T_8`$ Abelian monopole displays a signal about a factor two higher in the peak region. This is at variance of previous studies which find out that the disorder parameters for the three Abelian monopoles defined by means of the Polyakov projection coincide within statistical errors. This result is quite interesting, for it suggests that in the pattern of dynamical symmetry breaking due to the Abelian monopole condensation the color direction $`8`$ is slightly preferred.
Let us consider, now, in detail the $`T_8`$ Abelian monopole. In Figure 5 we report the derivative of the monopole free energy versus $`\beta `$ for the lattice with $`L_t=4`$ and $`L_s=32`$. We also display the absolute value of the Polyakov loop:
$$|P|=<|\frac{1}{V}\underset{\stackrel{}{x}}{}\frac{1}{3}\mathrm{Tr}[\underset{t=1}{\overset{L_t}{}}U_4(\stackrel{}{x},t)]|>,$$
(4.6)
We see that $`F_{\text{mon}}^{}`$ behaves like in the SU(2) case. Indeed, the free energy monopole derivative is zero within errors in the strong coupling region, while it display a sharp peak in correspondence of the rise of the Polyakov loop. In the weak coupling region $`F_{\text{mon}}^{}`$ is almost constant. The value of the plateau correspond to the classical action Eq. (3.12) which in the present case gives:
$$S_{\text{class}}\frac{\alpha \pi \beta }{12T_{\mathrm{phys}}}n_{\text{mon}}^2+\text{O}(1/L_s\text{a}),$$
(4.7)
so that
$$F_{\mathrm{mon}}^{}=\alpha \frac{\pi }{12}n_{\mathrm{mon}}^2.$$
(4.8)
The dashed line in Figure 5 in the weak coupling region corresponds to Eq. (4.8) with $`\alpha 2.0`$.
As in the $`SU(2)`$ theory we find that by increasing the spatial volume the peak increases (see Figure 6). Our data do not show a measurable shift of the peak. We feel that this is a manifestation of the first order nature of the $`SU(3)`$ deconfinement transition. This is confirmed if we look at the disorder parameter $`\mu `$. In Figure 7 we show the disorder parameter $`\mu `$ versus $`\beta `$ for the $`L_t=4`$ and $`L_s=32,48`$ lattices. Again we see that the disorder parameter $`\mu `$ is different from zero in the confined phase and decreases towards zero in the thermodynamic limit when we approach the critical coupling. Moreover our numerical results suggest that by increasing the spatial volume the two curves cross. This is precisely the finite volume behavior expected for the order parameter in the case of a first order phase transition .
## V. CONCLUSIONS
In this paper we have investigated the Abelian monopole condensation in the finite temperature SU(2) and SU(3) lattice gauge theories. By means of the lattice thermal partition functional we introduce a disorder parameter which signals the Abelian monopole condensation in the confined phase. By construction our definition of the disorder parameter is gauge invariant, so that we do not need to perform the Abelian projection. Our numerical results suggest that the disorder parameter $`\mu `$ is different from zero in the confined phase and tends to zero when approaching the critical coupling in the thermodynamic limit. We point out that in our approach the precise determination of the critical parameters could be obtained by means of a finite size scaling analysis. However, our results are consistent with a second order deconfining phase transition in the case of the $`SU(2)`$ gauge theory. On the other hand, in the case of $`SU(3)`$ the disorder parameter $`\mu `$ displays the finite-size behavior expected for a first order transition. It is clear that the finite size analysis in the critical region requires a separate study with both better statistic and larger lattice volumes. Remarkably, in the case of SU(3) gauge theory, where there are two independent Abelian monopole fields related to the two diagonal generators of the gauge algebra, we find that the non perturbative vacuum reacts moderately strongly in the case of the $`T_8`$ Abelian monopole. We feel that this last result should be useful in the theoretical efforts to understand the pattern of symmetry breaking in the deconfined phase of QCD.
In conclusion we stress that our approach, while keeping the gauge invariance, can be readily extended to incorporate the dynamical fermions. We hope to present results in this direction in a future study.
|
warning/0006/hep-ph0006267.html
|
ar5iv
|
text
|
# 1/𝑚_𝑄 and 1/𝑁_𝑐 Expansions for Excited Heavy Baryons with Light Quarks in the Spin-Flavor Symmetric Representation
## Abstract
The mass spectrum of the $`L=1`$ orbitally excited heavy baryons with light quarks in the spin-flavor symmetric representation is studied by the $`1/N_c`$ expansion method in the framework of the heavy quark effective theory. The mixing effect from the baryons in the mixed representation is considered. The general pattern of the spectrum is predicted which will be verified by the experiments in the near future. The $`1/m_Q`$ and SU(3) corrections are also considered. Mass relations for the baryons $`\mathrm{\Lambda }_{c1}^{()}`$, $`\mathrm{\Sigma }_{c1}^{()}`$, $`\mathrm{\Xi }_{c1}^{(^{})()}`$, and $`\mathrm{\Omega }_{c1}^{()}`$ are derived.
preprint: ASITP-2000-002 SNUTP 00-005
A lot of data for orbitally excited heavy baryons have been accumulating experimentally . Understanding them will extend our ability in the application of QCD. The heavy quark effective theory (HQET) provides a systematic way to investigate hadrons containing a single heavy quark. To obtain detailed prediction, however, some nonperturbative QCD methods have to be used. In this paper, $`1/N_c`$ expansion is applied in the analysis. Within this framework, the masses of $`L=1`$ orbitally excited heavy baryons with light quarks in both the spin-flavor symmetric and mixed representation have been analyzed . By the HQET sum rule, masses of lowest state of the excited baryons have also been calculated . They were studied in other approaches, too, for example in quark models , in the chiral Lagrangian formalism and in the Skyrme model or the large $`N_c`$ <sup>§</sup><sup>§</sup>§We make distinction between $`1/N_c`$ expansion and the large $`N_c`$ limit. Because the former is essentially based on the light quark spin-flavor symmetry in the baryon sector, the leading order result of it is not that of $`N_c\mathrm{}`$. See Manohar, in Ref.. HQET . In constituent quark models , the classification of the baryons according to the light quark spin-flavor symmetry is taken to be physical. In the treatment of the baryons with light quarks in the spin-flavor symmetric representation in Ref., it was erroneous to take only one light quark being excited. In fact, it is the heavy quark that is orbitally excited. Note that the orbital excitation of the heavy quark is not suppressed by the mass of the heavy quark. Or relatively speaking, it is the light quark pair as a whole in which the two light quarks have zero relative orbital angular momentum, that is $`L=1`$ excited . This paper reconsiders the excited heavy baryons with light quarks in the spin-flavor symmetric representation in the approach of $`1/N_c`$ expansion within the framework of the HQET. The results are very simple. Furthermore, the mixing between the two kinds of representations will also be discussed by $`1/N_c`$ expansion, which is argued being small. Therefore our results are physical and predictive.
In the HQET, many features of heavy hadrons have been analyzed. In the heavy quark limit, the heavy quark spin decouples from the strong interaction. The mass of a heavy hadron $`H`$ is expanded as
$$M_H=m_Q+\overline{\mathrm{\Lambda }}_H\frac{\lambda _1^H}{2m_Q}+\frac{\lambda _2^H}{2m_Q}+O(\frac{1}{m_Q^2}),$$
(1)
where $`m_Q`$ is the heavy quark mass, the parameter $`\overline{\mathrm{\Lambda }}_H`$ is independent of the heavy quark spin and flavor, and describes mainly the contribution of the light degrees of freedom in the baryon. $`\lambda _1^H`$ and $`\lambda _2^H`$ are the kinetic and chromomagnetic matrix elements, respectively,
$`\lambda _1^H`$ $`=`$ $`H(v)|\overline{h}_v(iD)^2h_v|H(v),`$ (2)
$`\lambda _2^H`$ $`=`$ $`H(v)|\overline{h}_v{\displaystyle \frac{g_s}{2}}G_{\mu \nu }\sigma ^{\mu \nu }h_v|H(v),`$ (3)
with $`h_v`$ denoting the heavy quark field with velocity $`v`$. The quantities $`\overline{\mathrm{\Lambda }}_H`$, $`\lambda _1^H`$ and $`\lambda _2^H`$ should be calculated by nonperturbative HQET.
At this stage, the $`1/N_c`$ expansion is applied in the analysis. It is one of the most important and model-independent methods of nonperturbative QCD. Nonperturbative properties of mesons can be observed from the analysis of the planar diagrams, and baryons from the Hartree-Fock picture. For the ground state baryons, it has been found that there is a contracted SU(2$`N_f`$) light quark spin-flavor symmetry in the large $`N_c`$ limit . This makes a $`1/N_c`$ expansion based on the spin-flavor structure possible for the baryons. Many quantitative predictions and further extensions of the above result have been made .
Before we go on, two remarks should be made. First, the above mentioned $`1/N_c`$ expansion applies to the s- or p-wave states of low spin in the baryon multiplet. The states with spin of order $`N_c/2`$ are considerably modified by spin-spin and spin-orbit interactions . Second, It is actually $`N_c1`$, which is $`2`$ in real World, that will be taken as a large number, because heavy quark is distinguished. This is an improvement compared to the $`1/N_c`$ for the excited heavy baryons with light quarks in the mixed representation. In that case, the expansion parameter is $`N_c2`$ .
The quantum numbers which describe the hadrons are angular momentum $`J`$ and isospin $`I`$. For the heavy hadrons, the total angular momentum of the light degrees of freedom $`J^l`$ becomes a good quantum number in the HQET. In the light quark spin-flavor symmetric representation, the light degrees of freedom in $`H`$ look like a collection of $`N_c1`$ light quarks without orbital angular momentum excitation. This picture for the light quarks is essentially the same as that of the ground state heavy baryons. The spin-flavor decomposition rule is $`I=S^l`$ for the non-strange baryons, where $`S^l`$ is the total spin of the light quark system. Note that the light quark system as a whole has $`L=1`$ orbital angular momentum. In other words, the heavy quark now is $`L=1`$ excited in this case. In real World $`N_c`$ is fixed to be 3, so there are only two light quarks in the heavy baryon. The spin-flavor structure of them is quite simple, $`(I,S^l)=(0,0)`$ and $`(1,1)`$. All possible states of excited heavy baryons are listed in Table I. In the table, except the third state, the other six states form three pairs. Each pair is a doublet under the heavy quark spin symmetry. We adopt the Hartree-Fock picture to study $`\overline{\mathrm{\Lambda }}_H`$ where in the baryon $`H`$, the light quarks are in the spin-flavor symmetric representation. One of the essential points of the $`1/N_c`$ expansion is the $`N_c`$ counting rules of the relevant Feynman diagrams.
In the Hartree–Fock picture of the baryons, the $`N_c`$ counting rules require us to include many-body interactions in the analysis, instead of including only one- or two-body interactions. However, a large part of these interactions are spin-flavor irrelevant. Namely this part contributes in the order $`N_c\mathrm{\Lambda }_{\mathrm{QCD}}`$ universally to all the baryons with different spin-flavor structure in Table I. This makes us arrive in an $`1/N_c`$ expansion based on the light quark spin-flavor structure of the baryons. The mass splittings among the baryons in the same light quark spin-flavor representation can be obtained. For the purely light quark contribution to $`\overline{\mathrm{\Lambda }}_H`$, the $`1/N_c`$ analysis goes the same as that to the ground state heavy baryons . There is a light quark spin-flavor symmetry at the leading order of the $`1/N_c`$ expansion. $`\overline{\mathrm{\Lambda }}_H`$ is trivially $`N_c\mathrm{\Lambda }_{\mathrm{QCD}}`$ at this order. The mass splitting due to the light quark spin-flavor symmetry violation started from $`S_{}^{l}{}_{}{}^{2}/N_c`$ . However, different from the ground state baryons, formally the orbital angular momentum of the heavy quark has more dominant contribution to $`\overline{\mathrm{\Lambda }}_H`$ than $`O(1/N_c)`$. This is because of the orbital-light-quark-spin interactions. After summing up all the relevant many-body interactions, this order $`O(1)`$ contribution is $`\stackrel{}{L}\stackrel{}{S^l}f({\displaystyle \frac{S_{}^{l}{}_{}{}^{2}}{N_c^2}})`$, where $`f`$ is a general function which can be Taylor expanded. The mass $`\overline{\mathrm{\Lambda }}_H`$ can be written simply as
$$\overline{\mathrm{\Lambda }}_H^0=N_c\stackrel{~}{c_0}+\stackrel{~}{c_1}\stackrel{}{L}\stackrel{}{S^l}+\stackrel{~}{c_2}\frac{S_{}^{l}{}_{}{}^{2}}{N_c}+O\left(\frac{1}{N_c^2}\right),$$
(4)
where coefficients $`\stackrel{~}{c_i}\mathrm{\Lambda }_{\mathrm{QCD}}`$ ($`i=0,1,2`$). There should be also term proportional to $`L^2`$ in the above equation, which gives constant contribution to $`\overline{\mathrm{\Lambda }}_H^0`$ for a given light quark representation, and therefore has been absorbed into the leading term. The term $`\stackrel{}{L}\stackrel{}{S^l}`$ can be rewritten as $`J_{}^{l}{}_{}{}^{2}S_{}^{l}{}_{}{}^{2}`$ with $`\stackrel{}{J^l}`$ being defined as $`\stackrel{}{J^l}=\stackrel{}{L}+\stackrel{}{S^l}`$. Therefore
$$\overline{\mathrm{\Lambda }}_H^0=N_cc_0+c_1(J_{}^{l}{}_{}{}^{2}S_{}^{l}{}_{}{}^{2})+c_2\frac{S_{}^{l}{}_{}{}^{2}}{N_c}+O\left(\frac{1}{N_c^2}\right),$$
(5)
where coefficients $`c_i\mathrm{\Lambda }_{\mathrm{QCD}}`$ need to be determined from experiments.
The numerical results are also given on the right-handed side of Table I. Because the mass formula of Eq. (5) is rather simple, some features of the spectrum can still be discussed. The parameters $`c_0`$ and $`c_2`$ are naturally expected to be positive. However $`c_1`$ can have both signs. If $`c_1>0`$, we see that the singlet state ($`J`$,$`I`$) $`=`$ ($`\frac{1}{2}`$,1) could be the lowest state. By requiring the first doublet to be the lowest, we must have $`c_2>2N_cc_1`$. The resulting spectrum will be
$$M(\frac{1}{2}(\frac{3}{2}),0,1,0)<M(\frac{1}{2},1,0,1)<M(\frac{1}{2}(\frac{3}{2}),1,1,1)<M(\frac{3}{2}(\frac{5}{2}),1,2,1)$$
(6)
with the quantum numbers denoting $`J`$, $`I`$, $`J^l`$ and $`S^l`$, respectively. On the other hand, if $`c_1<0`$, the first doublet is the lowest states only if $`c_2>N_cc_1`$. In this case, the singlet is the heaviest, and the spectrum is
$$M(\frac{1}{2}(\frac{3}{2}),0,1,0)<M(\frac{3}{2}(\frac{5}{2}),1,2,1)<M(\frac{1}{2}(\frac{3}{2}),1,1,1)<M(\frac{1}{2},1,0,1).$$
(7)
None of the above discussed spectrum pattern is consistent with the quark model prediction . It should be noted that our analysis neglected the $`1/N_c^3`$ correction (compared to the leading order) which is expected to be not significant.
The conditions for $`c_2`$ are not satisfactory, although they are not unreasonable considering that in real World $`N_c`$ is not large. In fact, this unsatisfactory point can be avoided if the mixing effect from the baryons in the mixed representation is considered.
It is necessary to consider the mixing between the baryons with light quarks in the spin-flavor symmetric and mixed representations. When they have same quantum numbers of ($`J`$, $`I`$, $`J^l`$), there is no physical way to distinguish them. This consideration will give the physical spectrum. Because of the light quark spin-flavor symmetry at the leading order of $`1/N_c`$ expansion, the baryons with same ($`J`$, $`I`$, $`J^l`$) quantum numbers but in different representations do not mix. The mixing occurs at the sub-leading order. The classification of baryons by the spin-flavor symmetry is therefore physical at the leading order . For the physical spectrum, the mixing results in a deviation from $`\overline{\mathrm{\Lambda }}_H^0`$. By denoting the mixing mass as $`\stackrel{~}{m}`$ which is of $`O(1)`$, the mass matrix for the baryons with same ($`J,I,J^l`$) is written as
$$\left(\begin{array}{cc}\overline{\mathrm{\Lambda }}_H^0& \stackrel{~}{m}\\ \stackrel{~}{m}& \overline{\mathrm{\Lambda }}_H^{}^0\end{array}\right),$$
(8)
where $`H^{}`$ is the corresponding baryon in the mixed representation. $`\overline{\mathrm{\Lambda }}_H^{}^0`$ was given in Ref. . The mass difference $`\overline{\mathrm{\Lambda }}_H^0\overline{\mathrm{\Lambda }}_H^{}^0`$ is $`O(1)`$. Taking $`\stackrel{~}{m}<\overline{\mathrm{\Lambda }}_H^{}^0\overline{\mathrm{\Lambda }}_H^0`$ for illustration, the physical mass are corrected to be
$$\begin{array}{ccc}\overline{\mathrm{\Lambda }}_H\hfill & \hfill & \overline{\mathrm{\Lambda }}_H^0\frac{\stackrel{~}{m}^2}{\overline{\mathrm{\Lambda }}_H^{}^0\overline{\mathrm{\Lambda }}_H^0},\hfill \\ \overline{\mathrm{\Lambda }}_H^{}\hfill & \hfill & \overline{\mathrm{\Lambda }}_H^{}^0+\frac{\stackrel{~}{m}^2}{\overline{\mathrm{\Lambda }}_H^{}^0\overline{\mathrm{\Lambda }}_H^0}.\hfill \end{array}$$
(9)
The mixing effect $`{\displaystyle \frac{\stackrel{~}{m}^2}{\overline{\mathrm{\Lambda }}_H^{}^0\overline{\mathrm{\Lambda }}_H^0}}`$ is positive. It reduces the predictive power of Eq. (5) for the mass spectrum. The $`1/N_c`$ expansion of $`\stackrel{~}{m}`$ is parameterized as
$$\stackrel{~}{m}=\stackrel{~}{m}_0+O(1/N_c),$$
(10)
where $`\stackrel{~}{m}_0`$ is universal due to the light quark spin-flavor symmetry. To the order of $`O(1)`$, the spectrum is given as follows explicitly.
$$\begin{array}{ccc}\overline{\mathrm{\Lambda }}_{(\frac{1}{2}(\frac{3}{2}),0,1)}\hfill & =\hfill & N_cc_0+2c_1\frac{\stackrel{~}{m}_0^2}{kc_{LS}\frac{1}{6}\overline{c_1}\frac{1}{4}\overline{c_2}2c_1},\hfill \\ \overline{\mathrm{\Lambda }}_{(\frac{1}{2},1,0)}\hfill & =\hfill & N_cc_02c_1,\hfill \\ \overline{\mathrm{\Lambda }}_{(\frac{1}{2}(\frac{3}{2}),1,1)}\hfill & =\hfill & N_cc_0\frac{\stackrel{~}{m}_0^2}{k},\hfill \\ \overline{\mathrm{\Lambda }}_{(\frac{3}{2}(\frac{5}{2}),1,2)}\hfill & =\hfill & N_cc_0+4c_1,\hfill \end{array}$$
(11)
where $`k`$ is an $`O(1)`$ constant that remains after the $`\overline{\mathrm{\Lambda }}_H^{}^0`$ and $`\overline{\mathrm{\Lambda }}_H^0`$ cancellation, $`\overline{\mathrm{\Lambda }}_H^{}^0`$ is parameterized by $`c_{LS}`$, $`\overline{c_1}`$ and $`\overline{c_2}`$ which are around $`\mathrm{\Lambda }_{\mathrm{QCD}}`$, and can be found in the Table II of Ref. (where $`\overline{c_1}`$ and $`\overline{c_2}`$ were denoted as $`c_1`$ and $`c_2`$, respectively). Note that the masses of the states $`({\displaystyle \frac{1}{2}},1,0)`$ and $`({\displaystyle \frac{3}{2}}({\displaystyle \frac{5}{2}}),1,2)`$ are not affected by the mixing, because there are no physical states with the same good quantum numbers in the mixed representation. From the above spectrum, we see that $`c_1>0`$. The states $`({\displaystyle \frac{3}{2}}({\displaystyle \frac{5}{2}}),1,2)`$ is always the highest states. They are heavier than the other states at least by $`4c_1`$ through requiring the states $`({\displaystyle \frac{1}{2}}({\displaystyle \frac{3}{2}}),0,1)`$ to be the lowest. If $`2c_1>{\displaystyle \frac{\stackrel{~}{m}_0^2}{k}}`$, the requirement implies
$$\frac{\stackrel{~}{m}_0^2}{kc_{LS}\frac{1}{6}\overline{c_1}\frac{1}{4}\overline{c_2}2c_1}>4c_1.$$
(12)
In this case, the spectrum pattern is
$$M(\frac{1}{2}(\frac{3}{2}),0,1)<M(\frac{1}{2},1,0)<M(\frac{1}{2}(\frac{3}{2}),1,1)<M(\frac{3}{2}(\frac{5}{2}),1,2).$$
(13)
On the other hand, if $`2c_1<{\displaystyle \frac{\stackrel{~}{m}_0^2}{k}}`$, the requirement is
$$\stackrel{~}{m}_0^2\left(\frac{1}{kc_{LS}\frac{1}{6}\overline{c_1}\frac{1}{4}\overline{c_2}2c_1}\frac{1}{k}\right)>2c_1,$$
(14)
which gives the spectrum
$$M(\frac{1}{2}(\frac{3}{2}),0,1)<M(\frac{1}{2}(\frac{3}{2}),1,1)<M(\frac{1}{2},1,0)<M(\frac{3}{2}(\frac{5}{2}),1,2).$$
(15)
Experimentally, the excited charmed baryons $`\mathrm{\Lambda }_{c1}(\frac{1}{2})`$ and $`\mathrm{\Lambda }_{c1}(\frac{3}{2})`$ have been found which correspond to the $`({\displaystyle \frac{1}{2}}({\displaystyle \frac{3}{2}}),0,1)`$ states. More data are needed to fix the unknown parameters $`c_i`$’s, $`\overline{c_i}`$’s, $`k`$ and $`c_{LS}`$. In the near future, experiments will check the above predicted spectrum. Hopefully one of the above mass patterns will be picked out. It will be a check for the validity of our method, if the parameters are in the reasonable range ($`\mathrm{\Lambda }_{\mathrm{QCD}}`$) and meanwhile satisfy the relations given above.
For a complete analysis of the heavy hadron masses, $`1/m_Q`$ corrections have to be considered. The general expression of the corrections have been given in Eqs. (1) and (3). The quantities $`\lambda _1^H`$ and $`\lambda _2^H`$ can be analyzed by the $`1/N_c`$ expansion in the similar way as $`\overline{\mathrm{\Lambda }}_H`$. In the leading order of $`1/N_c`$, $`\lambda _1^H`$ is independent of the light quark structure and scales as unity. Therefore we have the following expansion ,
$`\lambda _1^H`$ $`=`$ $`\stackrel{~}{c_0}^{}+\stackrel{~}{c_1}^{}\stackrel{}{L}{\displaystyle \frac{\stackrel{}{S^l}}{N_c}}+O\left({\displaystyle \frac{1}{N_c^2}}\right)`$ (16)
$`=`$ $`c_0^{}+c_1^{}{\displaystyle \frac{J_{}^{l}{}_{}{}^{2}S_{}^{l}{}_{}{}^{2}}{N_c}}+O\left({\displaystyle \frac{1}{N_c^2}}\right).`$ (17)
The mixing effect also affects $`\lambda _1^H`$. Its $`1/N_c`$ expansion is that the non-vanishing contribution begins at $`O(1/N_c)`$. And at this order, the contribution is constant which can be absorbed into $`c_0^{}`$. The parameters $`c_0^{}/2m_Q`$ and $`c_1^{}/2m_Q`$ can be absorbed into $`c_0`$ and $`c_1`$ in Eq. (5), respectively. The inclusion of $`\lambda _1^H`$ corrects the masses of the baryons at the order of $`1/m_Q`$ which is expected to be not significant. It does not change the mass pattern given above to the order of $`O(1/m_cN_c)`$.
The degeneracy in the spectrum due to the heavy quark spin symmetry is lifted by $`\lambda _2^H`$. According to the definition in Eq. (3), $`\lambda _2^H`$ is heavy baryon spin dependent. It is convenient to extract this dependence explicitly,
$$\lambda _2^H=d_H\lambda _2$$
(18)
where $`d_H=2j^l`$ for $`H`$ with $`J=j^l+\frac{1}{2}`$, and $`d_H=2j^l2`$ for $`H`$ with $`J=j^l\frac{1}{2}`$. The new defined heavy quark hadronic matrix element $`\lambda _2`$ is heavy baryon spin independent. It is also independent of the light quark structure and scales as unity in the leading order $`1/N_c`$ expansion. Like $`\lambda _1^H`$, the $`1/N_c`$ expansion for $`\lambda _2^H`$ is
$$\lambda _2^H=d_H\left[c_0^{\prime \prime }+c_1^{\prime \prime }\frac{J_{}^{l}{}_{}{}^{2}S_{}^{l}{}_{}{}^{2}}{N_c}+O\left(\frac{1}{N_c^2}\right)\right].$$
(19)
The mixing effect for $`\lambda _2^H`$ is that the leading nonzero contribution is $`O(1/N_c)`$ which is constant and therefore can be absorbed into $`c_0\mathrm{"}`$. The parameters $`c_i^{\prime \prime }`$ should be determined by the experimental data. If we work to the accuracy of $`\mathrm{\Lambda }_{\mathrm{QCD}}/(m_QN_c)10\%`$, $`c_0^{\prime \prime }`$ can be fixed from the mass splitting of $`\mathrm{\Lambda }_{c1}(\frac{3}{2})`$ and $`\mathrm{\Lambda }_{c1}(\frac{1}{2})`$,
$$c_0^{\prime \prime }=\frac{m_c}{3}\left[M_{\mathrm{\Lambda }_{c1}(\frac{3}{2})}M_{\mathrm{\Lambda }_{c1}(\frac{1}{2})}\right](128\mathrm{MeV})^2,$$
(20)
by taking $`m_c1.5`$ GeV. Note that $`c_0^{\prime \prime }`$ is positive. The mass splittings of the other degenerate states listed in Table I are predicted to be
$`M({\displaystyle \frac{5}{2}},1,2,1)M({\displaystyle \frac{3}{2}},1,2,1)`$ $`=`$ $`{\displaystyle \frac{5c_0^{\prime \prime }}{m_c}}55\mathrm{MeV},`$ (21)
$`M({\displaystyle \frac{3}{2}},1,1,1)M({\displaystyle \frac{1}{2}},1,1,1)`$ $`=`$ $`{\displaystyle \frac{3c_0^{\prime \prime }}{m_c}}33\mathrm{MeV},`$ (22)
to the accuracy of $`c_0^{\prime \prime 2}/(m_cN_c)`$ which is about $`5`$ MeV. These predictions can be checked with the experiments in the near future.
Finally, let us consider the case of the excited heavy baryons with light quarks including the strange quark. Very recently, there are experimental evidence of the charmed-strange analogs of $`\mathrm{\Lambda }_{c1}(\frac{3}{2})`$, $`\mathrm{\Xi }_{c1}(\frac{3}{2})`$ particles . The above framework can be easily extended to include the charmed-strange baryons by taking strangeness as perturbation to the light quark flavor symmetry. The relevant baryon mass is then expressed as
$$M_H=m_Q+N_cc_0+c_1(J_{}^{l}{}_{}{}^{2}S_{}^{l}{}_{}{}^{2})+\mathrm{mixing}+c_3(s)+O\left(\frac{1}{N_c}\right),$$
(23)
where $`s`$ is the heavy baryon strangeness number which can be $`0`$, $`1`$, or $`2`$. The parameter $`c_3`$ stands for the leading order of SU(3) correction to the $`\overline{\mathrm{\Lambda }}_H`$ given in Eq. (5). It is fixed by the mass difference of $`\mathrm{\Xi }_{c1}(\frac{3}{2})`$ and $`\mathrm{\Lambda }_{c1}(\frac{3}{2})`$,
$$c_3190\mathrm{MeV}.$$
(24)
The mass of $`\mathrm{\Xi }_{c1}(\frac{1}{2})`$ is then predicted to be 190 MeV higher than $`\mathrm{\Lambda }_{c1}(\frac{1}{2})`$,
$`M_{\mathrm{\Xi }_{c1}(\frac{1}{2})}`$ $`=`$ $`M_{\mathrm{\Lambda }_{c1}(\frac{1}{2})}+M_{\mathrm{\Xi }_{c1}(\frac{3}{2})}M_{\mathrm{\Lambda }_{c1}(\frac{3}{2})}`$ (25)
$``$ $`2784\mathrm{MeV}.`$ (26)
Note that this prediction is only subject to a small uncertainty which is about $`c_3^2/(m_cN_c)10`$ MeV. The future experiments may find that particles $`\mathrm{\Sigma }_{c1}^{()}`$, which are the lowest charmed $`L=1`$ states with isospin $`1`$, are just the state pair $`[\frac{1}{2}(\frac{3}{2}),1,1,1]`$ in Table I. Their strange analogs $`\mathrm{\Xi }_{c1}^{()}`$ and $`\mathrm{\Omega }_{c1}^{()}`$ will then be predicted from the similar relation rather precisely,
$`M_{\mathrm{\Xi }_{c1}^{}(\frac{3}{2})}M_{\mathrm{\Sigma }_{c1}(\frac{3}{2})}`$ $`=`$ $`M_{\mathrm{\Xi }_{c1}^{}(\frac{1}{2})}M_{\mathrm{\Sigma }_{c1}(\frac{1}{2})}+O\left({\displaystyle \frac{\mathrm{\Lambda }_{\mathrm{QCD}}}{m_cN_c}}\right),`$ (27)
$`M_{\mathrm{\Omega }_{c1}(\frac{3}{2})}M_{\mathrm{\Sigma }_{c1}(\frac{3}{2})}`$ $`=`$ $`M_{\mathrm{\Omega }_{c1}(\frac{1}{2})}M_{\mathrm{\Sigma }_{c1}(\frac{1}{2})}+O\left({\displaystyle \frac{\mathrm{\Lambda }_{\mathrm{QCD}}}{m_cN_c}}\right).`$ (28)
To the accuracy of $`s^2/N_c30\%`$,
$`M_{\mathrm{\Omega }_{c1}^{()}}M_{\mathrm{\Xi }_{c1}^{()}}`$ $`=`$ $`M_{\mathrm{\Xi }_{c1}^{()}}M_{\mathrm{\Sigma }_{c1}^{()}}+O\left({\displaystyle \frac{s^2}{N_c}}\right)`$ (29)
$``$ $`(190\pm 70)\mathrm{MeV}.`$ (30)
In summary, we have applied the $`1/N_c`$ expansion method to study the mass spectrum of the $`L=1`$ orbitally excited heavy baryons with light quarks being in the spin-flavor symmetric representation within the framework of the HQET. The analysis is very simple compared to that for the heavy baryons with light quarks in the mixed representation in Ref. . The simplicity is an unique feature in this case. It can be seen from the following point of view, namely the light quark system is in the ground state and it is the heavy quark that is orbitally excited. However the mixing effect due to the baryon states in the mixed representation corrects the spectrum pattern in the subleading order of $`1/N_c`$ expansion. The effect is important to get the realistic spectrum at this order. The general pattern of the baryon spectrum has been given, which will be verified by the experiments in the near future. The $`1/m_Q`$ and SU(3) corrections have also been considered. Certain mass relations for the baryons $`\mathrm{\Lambda }_{c1}^{()}`$, $`\mathrm{\Sigma }_{c1}^{()}`$, $`\mathrm{\Xi }_{c1}^{(^{})()}`$, and $`\mathrm{\Omega }_{c1}^{()}`$ have been derived. The same analysis can be applied to the bottom baryons.
Acknowledgments
We would like to thank Chao-Hsi Chang and Chao-shang Huang for helpful discussions. This work was supported in part by the BK21 Program of the Korean Ministry of Education.
|
warning/0006/cond-mat0006006.html
|
ar5iv
|
text
|
# Localization of acoustic waves in 1D random liquid media
\[
## Abstract
We study acoustic propagation in one dimensional water ducts containing many air-filled blocks. The acoustic band structures for the periodic arrangements of the blocks is calculated, whereas the transmission for various random configurations of the blocks is computed by the transfer matrix method. The results show that while all waves are localized for any given amount of disorders, there is no genuine scaling behavior for the system. The results also reveal a distinct collective behaviour for localized waves, a feature useful for distinguishing the localization from the residual absorption effect.
PACS numbers: 43.20., 71.55J, 03.40K
\]
The fact that the electronic localization in disordered systems is of wave nature has led to suggestion that classical waves could be similarly localized in random systems. The effort in searching for localization of classical waves such as acoustic and electro-magnetic waves is tremendous. It has drawn intensive attentions from both theorists and experimentalists.
Localization of waves in one dimensional (1D) systems has attracted particular interest from scientists because in higher dimensions the interaction between waves and scatterers is so complicated that the theoretical computation is rather involved and most solutions require a series of approximations which are not always justified, making it difficult to relate theoretical predictions to experimental observations. Yet wave localization in one dimension (1D) poses a more manageable problem which can be tackled in an exact manner by the transfer matrix method. Moreover, results from 1D can provide insight to the problem of wave localization in general and are suitable for testing various ideas. Indeed, over the past decades considerable progress has been made in understanding the localization behavior in 1D disordered systems. However, a number of important issues remained untouched. These issues include, for example, how waves are localized inside the media and whether there is a distinct feature for wave localization which would allow to differentiate the localization from residual absorption effect without ambiguity. Results from the statistical analysis of the scaling behavior in 1D random media is not conclusive. A further question could be whether the localized state is a phase state which would accommodate a more systematic interpretation from the view of a symmetry breaking and collective behavior, in analogy to phase states such as superconductivity. This Letter attempts to provide insight to these questions.
Here we study the problem of acoustic wave propagation in one dimensional water ducts containing many air blocks either regularly or randomly but on average regularly distributed inside the ducts. The frequency band structures and wave transmission are computed numerically. We show that while our results affirm the previous claim that all waves are localized inside an 1D medium with any amount of disorder, there are, however, a few distinctive features in our results. Among them, in contrast to optical cases, there is no universal scaling behavior in the present system. In addition, when waves are localized, a collective behaviour of the system emerges.
Assume that $`N`$ air blocks of identical thickness $`a`$ are placed regularly or randomly in a water duct with length $`L`$ measured from the left boundary of the duct (LB). The air fraction is $`\beta =Na/L`$, the average distance between two adjacent water layers is $`d=L/N=a/\beta `$, and the average thickness of water layers is $`b=(_{j=1}^Nb_j)/N`$. The degree of randomness for the system is controlled by a parameter $`\mathrm{\Delta }`$ in such a way that the thickness of the $`j`$-th water layer is $`b_j=b(1+\delta _j)`$ with $`\delta _j`$ being a random number within the interval $`[\mathrm{\Delta },\mathrm{\Delta }]`$; the regular case corresponds to $`\mathrm{\Delta }=0`$. An acoustic source placed at LB generates monochromatic waves with an oscillation $`v(t)=ve^{i\omega t}`$. Transmitted waves propagate through the $`N`$ air layers and travel to the right infinity. In order to avoid unnecessary confusion, possible effects from surface tension, viscosity or any absorption are neglected. For convenience, we use the dimensionless quantity $`kb`$ to measure the frequency, where $`k=\omega /c`$ is the wave number and $`c`$ is the sound speed in water. Similarly, $`k_g`$ and $`c_g`$ represent the wave number and sound speed in the air blocks respectively.
The wave propagation in such a system can be solved using the transfer matrix method. After dropping out the time factor $`e^{i\omega t}`$, the wave propagation is governed by two equations. The first is the Helmholtz equation,
$$p_m^{\prime \prime }(x)+k_m^2p_m(x)=0,$$
(1)
in which $`p_m(x)`$ is the pressure field, and the subscript $`m`$ refers to the medium that can be either water or air, depending on where $`x`$ is located. Within any layer, Eq. (1) warrants two solutions: $`A_me^{ik_mx}`$ represents the wave transmitted away from the source to the right and $`B_me^{ik_mx}`$ the wave reflected towards the source. The total wave is therefore $`p_m(x)=A_me^{ik_mx}+B_me^{ik_mx}`$. The second equation relates the oscillation velocity and the pressure field,
$$u_m(x)=\frac{1}{i\omega \rho _m}p_m^{}(x)=(A_me^{ik_mx}B_me^{ik_mx})/\rho _mc_m,$$
(2)
where $`\rho _m`$ refers to the equilibrium mass density of medium $`m`$.
The coefficients $`A_m`$ and $`B_m`$ in any two adjacent layers are connected by a transfer matrix. By invoking the condition that the pressure and velocity fields are continuous across the interfaces separating water and air and the condition that there is no reflected wave to the right end of the system, the matrix elements for the transfer matrices linking all interfaces can be deduced, and waves in any particular block is therefore completely determined. The ratio between the amplitude of the outgoing wave at the right boundary and that of the transmitted wave at the source defines the transmission rate.
Before going further, a general discussion on wave propagation is in place. When waves propagate through media alternated with different material compositions, multiple scattering of waves is established by an infinite recursive pattern of rescattering. In terms of wave fields, the energy flow in the system is calculated from $`J\text{Re}[i(p^{}(x)_xp(x)]`$. Writing $`p(x)=|A(x)|e^{i\theta (x)}`$ with $`|A|`$ and $`\theta `$ being the amplitude and phase respectively, the energy flow becomes $`J|A|^2_x\theta `$. Obviously, the energy flow will come to a complete halt and the waves could be localized in space when phase $`\theta `$ is constant and $`|A|`$ does not equal zero.
First consider the case with the periodic placement of air blocks (with spatial period $`d`$, water layer thickness $`b`$). According to Bloch theorem, and the wave field $`p`$ can be written as
$$p(x)=A(x)e^{iKx},$$
(3)
where $`A(x)`$ is a periodic function satisfying $`A(x+d)=A(x)`$, and $`K`$ is the Bloch wave number represented in the dispersion relation
$$\mathrm{cos}Kd=\mathrm{cos}k_ga\mathrm{cos}kb\mathrm{cosh}2\eta \mathrm{sin}k_ga\mathrm{sin}kb.$$
(4)
Here $`k_g=k/h`$ and $`\eta =\mathrm{ln}q`$ with $`q^2=gh`$ and $`g=\rho _g/\rho `$, $`h=c_g/c`$.
The band structures for four air-fractions are shown in Fig. 1. The ranges covered by the dispersion curves refer to the passing bands, while the areas sandwiched by any two curves to the complete band gaps. Within the gap, waves are evanescent and cannot propagate. Since the air blocks are strong scatterers due to the large contrast in acoustic impedance between air and water, we see that even a small air-fraction can lead to wide band gaps as seen in Fig. 1(a). With increasing $`\beta `$, the passing bands are narrowed and become streaks. When the air-fraction is reduced, however, the band gaps gradually disappear, as shown by Fig. 1(d).
Wave propagation properties can be significantly affected by varying air-fraction or adding randomness. Fig. 2(a) presents the typical results of the transmission rate as a function of $`kb`$ for various $`\beta `$ at a given randomness. At frequencies for which the wavelength is smaller than the averaged distance between air blocks, the transmission is significantly reduced by increasing air fraction.
Fig. 2(b) illustrates the effect of the randomness $`\mathrm{\Delta }`$ on transmission for a given air-fraction. For comparison, the transmission in the corresponding regular array ($`\mathrm{\Delta }=0`$) is also plotted. The gaps are located between $`kb/\pi =0.46`$ and $`1`$, $`1.21`$ and $`2`$, $`2.12`$ and $`3`$, and so on. We find that for frequencies located inside the band gaps of the corresponding regular array, the disorder-induced localization effect competes yet reduces the band gap effect. To characterize wave localization in this case, both the band gap and the disorder effects should be considered, supporting the two parameter scaling theory. However, increasing disorder tends to smear out the band structures. When exceeding a certain amount, the effect from the disorder suppresses the band gap effect completely, and there is no distinction between the localization at frequencies within and outside the band gaps.
Fig. 2 shows that the localization behaviour depends crucially on whether the wavelength ($`\lambda `$) is greater than the average distance between air blocks. When $`b/\lambda `$ is less than one, the localization effect is weak. We also observe that with the added disorder, the transmission is enhanced in the middle of the gaps. Similar enhancement due to disorder has also been reported recently. Differing from , however, the transmission at frequencies within the gaps of the corresponding periodic arrays in the present system is not always enhanced by disorder. Instead the transmission is reduced further by the disorder near the band edges.
Fig. 3 presents the results for Lyapunov exponent (LE) and its variance as a function of non-dimensional frequency $`kb/\pi `$. At low disorders, the variance of LE inside the gaps is small. Contrast to the optical case, there are no double maxima inside the gap. With increasing disorder, double peaks appears inside the allow bands. When exceeding a certain critical value, however, the double peaks emerge. The higher frequency, the lower is the critical value. For example, the double peaks are still visible in the first allow band (c. f. Fig. 3(b)), while there is only one peak inside the higher passing bands. Meanwhile, the increasing disorder reduces the band gap effect and smears LE, in accordance with Fig. 2. We also plot LE-variance relation in Fig. 3. With increasing disorder, we do not observe genuine linear dependence between LE and its variance, as expected from the single parameter scaling theory.
In the past the localization phenomenon in 1D is usually characterized by LE, here we propose to use a phase behavior of the waves to characterize the localization. We compute the energy density inside the sample from
$$E_m(x)=\frac{\rho _m}{4}\left(|u_m|^2+\frac{|p_m|^2}{\rho _m^2c_m^2}\right),$$
(5)
where $`m`$ refers to either water or air. Meanwhile, the phases of $`p(x)`$ and $`u(x)`$ are recorded. Expressing $`p(x)`$ and $`u(x)`$ as $`A_p(x)e^{i\theta _p(x)}`$ and $`A_u(x)e^{i\theta _u(x)}`$, we construct unit phase vectors $`\stackrel{}{v}_p=\mathrm{cos}\theta _p\widehat{e}_x+\mathrm{sin}\theta _p\widehat{e}_y`$ and $`\stackrel{}{v}_u=\mathrm{cos}\theta _u\widehat{e}_x+\mathrm{sin}\theta _u\widehat{e}_y`$. Then the behavior of the phase vectors along the path are investigated. Physically, these phase vectors represent the oscillation behavior of the system.
Typical results for the spatial distribution of the energy and the phase behaviour for the given disorder and air fraction are shown in Fig. 4. Here for the sake of convenience, only the phase vectors at the interfaces between air and water are shown. First, we note that the energy density is constant in each individual block. This is a special feature of 1D classical systems, and can be verified by a deduction from Eq. (5). We find that when the sample size is sufficiently large, waves are always localized for any given amount of randomness. When localized, the waves are trapped inside the medium, but not necessarily confined at the site of the source, unless the band gap effect is dominant. The energy distribution does not follow an exponential decay along the path. This differs from situations in higher dimensions. It is also shown that the energy stored in the medium can be tremendous (c. f. Figs. 4(a) and (b)). With increasing sample size, the peak amplitude may grow, pointing to the stochastic resonance, in agreement with . When disorder is weak and for frequencies within the gaps, such a stochastic resonance behavior disappears.
It is also observed that for all frequencies, there is a collective behavior for the phase vectors. In Fig. 4(c) we plot the phase vectors in three spatial regimes, namely, near the transmitting site, in the middle of the duct, and at the far end from the source. Symbols $`p^L`$, $`p^R`$, $`u^L`$, $`u^R`$ appearing in Fig. 4(c) denote respectively the phase vectors for the pressure and the velocity fields on the left and right side of the air blocks. It is clear that when waves are localized, all the phase vectors of the pressure field are pointing to either $`\pi /2`$ or $`\pi /2`$, and perpendicular to the phase vectors of the velocity field. The pressure at the two sides of any air block varies in phase. Mostly, the two sides also oscillate in phase. Different from higher dimensional cases in which all phase vectors of localized fields point to the same direction, the present phase vectors are constant by domains; this ensures no energy flows. The velocity field in neighboring domains oscillate exactly out of phase. The phase vector domains are sensitive to the arrangement of the air blocks. We stress that such a phase ordering not only exists for the boundaries of the air blocks, but also appears inside the whole medium.
The coherence behavior is a unique feature for wave localization, which could be verified by the cross correlation measurement. At the far end of the sample, however, the phase vectors become gradually disoriented, implying that the energy can leak out only at the boundary due to the finite sample size. This boundary effect vanishes exponentially as the sample size is increased. The fact that the phase vectors are constant in domains indicates that once waves are localized, no more energy can be pumped into the system. When localization is evident, increasing the sample size by adding more air blocks will not change the patterns of the energy distribution and phase vectors. Therefore the energy localization and the phase behavior are not caused by the boundary effect.
In summary, we have demonstrated a phase transition in acoustic propagation in an 1D random liquid medium. The results have shown that waves are always confined in a finite spatial region. The disorder leads a significant energy storage in the system. It is also indicated that the wave localization is related to a collective behavior of the system in the presence of multiple scattering, also observed for higher dimensions. The appearance of such a collective phenomenon may be regarded as an indication of a kind of Goldstone modes in the context of the field theory.
The work received support from National Science Council (No. NSC89-2611-M008-002 and NSC89-2112-M008-008).
|
warning/0006/math0006063.html
|
ar5iv
|
text
|
# Identification of Berezin-Toeplitz deformation quantization
## 1. Introduction
In the seminal work Bayen, Flato, Fronsdal, Lichnerowicz and Sternheimer drew the attention of both physical and mathematical communities to a well posed mathematical problem of describing and classifying up to some natural equivalence the formal associative differential deformations of the algebra of smooth functions on a manifold. The deformed associative product is traditionally denoted $``$ and called star-product.
If the manifold carries a Poisson structure, or a symplectic structure (i.e. a non-degenerate Poisson structure) or even more specific if the manifold is a Kähler manifold with symplectic structure coming from the Kähler form one naturally asks for a deformation of the algebra of smooth functions in the “direction” of the given Poisson structure. According to this deformation is treated as a quantization of the corresponding Poisson manifold.
Due to work of De Wilde and Lecomte , Fedosov , and Omori, Maeda and Yoshioka it is known that every symplectic manifold admits a deformation quantization in this sense. The deformation quantizations for a fixed symplectic structure can be classified up to equivalence by formal power series with coefficients in two-dimensional cohomology of the underlying manifold, see , , , , . Kontsevich showed that every Poisson manifold admits a deformation quantization and that the equivalence classes of deformation quantizations on a Poisson manifold can be parametrized by the formal deformations of the Poisson structure.
Despite the general existence and classification theorems it is of importance to study deformation quantization for manifolds with additional geometric structure and ask for deformation quantizations respecting in a certain sense this additional structure. Examples of this additional structure are the structure of a complex manifold or symmetries of the manifold.
Another natural question in this context is how some naturally defined deformation quantizations fit into the classification of all deformation quantizations.
In this article we will deal with Kähler manifolds. Quantization of Kähler manifolds via symbol algebras was considered by Berezin in the framework of his quantization program developed in ,. In this program Berezin considered symbol algebras with the symbol product depending on a small parameter $`\mathrm{}`$ which has a prescribed semi-classical behavior as $`\mathrm{}0`$. To this end he introduced the covariant and contravariant symbols on Kähler manifolds. However, in order to study quantization via symbol algebras on Kähler manifolds he, as well as most of his successors, was forced to consider Kähler manifolds which satisfy very restrictive analytic conditions. These conditions were shown to be met by certain classes of homogeneous Kähler manifolds, e.g., $`^n`$, generalized flag manifolds, Hermitian symmetric domains etc. The deformation quantization obtained from the asymptotic expansion in $`\mathrm{}`$ as $`\mathrm{}0`$ of the product of Berezin’s covariant symbols on these classes of Kähler manifolds was studied in a number of papers by Moreno, Ortega-Navarro (, ); Cahen, Gutt, Rawnsley (, , ); see also . This deformation quantization is differential and respects the separation of variables into holomorphic and anti-holomorphic ones in the sense that left star-multiplication (i.e. the multiplication with respect to the deformed product) with local holomorphic functions is pointwise multiplication, and right star-multiplication with local anti-holomorphic functions is also point-wise multiplication, see Section 2 for the precise definition. It was shown in that such deformation quantizations ”with separation of variables” exist for every Kähler manifold. Moreover, a complete classification (not only up to equivalence) of all differential deformation quantizations with separation of variables was given. They are parameterized by formal closed forms of type $`(1,1)`$. The basic results are sketched in Section 2 below. Independently a similar existence theorem was proven by Bordemann and Waldmann along the lines of Fedosov’s construction. The corresponding classifying (1,1)-form was calculated in . Yet another construction was given by Reshetikhin and Takhtajan in . They directly derive it from Berezin’s integral formulas which are treated formally, i.e., with the use of the formal method of stationary phase. The classifying form of deformation quantization from can be easily obtained by the methods developed in this paper.
In Engliš obtained asymptotic expansion of Berezin transform on a quite general class of complex domains which do not satisfy the conditions imposed by Berezin.
For general compact Kähler manifolds $`(M,\omega _1)`$ which are quantizable, i.e. admit a quantum line bundle $`L`$ it was shown by Bordemann, Meinrenken and Schlichenmaier that the correspondence between the Berezin-Toeplitz operators and their contravariant symbols associated to $`L^m`$ has the correct semi-classical behavior as $`m\mathrm{}`$. Moreover, it was shown in ,, that it is possible to define a deformation quantization via this correspondence. For this purpose one can not use the product of contravariant symbols since in general it can not be correctly defined.
The approach of was based on the theory of generalized Toeplitz operators due to Boutet de Monvel and Guillemin , which was also used by Guillemin in his proof of the existence of deformation quantizations on compact symplectic manifolds.
The deformation quantization obtained in ,, which we call the Berezin-Toeplitz deformation quantization, is defined in a natural way related to the complex structure. It fulfils the condition to be ‘null on constants’ (i.e. $`1g=g1=g`$), it is self-adjoint (i.e. $`\overline{fg}=\overline{g}\overline{f}`$), and admits a trace of certain type (see for details).
As one of the results of this article we will show that the Berezin-Toeplitz deformation quantization is differential and has the property of separation of variables, though with the roles of holomorphic and antiholomorphic variables swapped. To comply with the conventions of we consider the opposite to the Berezin-Toeplitz deformation quantization (i.e., the deformation quantization with the opposite star-product) which is a deformation quantization with separation of variables in the usual sense.
We will show how the Berezin-Toeplitz deformation quantization fits into the classification scheme of . Namely, we will show that the classifying formal (1,1)-form of its opposite deformation quantization is
(1.1)
$$\stackrel{~}{\omega }=\frac{1}{\nu }\omega _1+\omega _{can},$$
where $`\nu `$ is the formal parameter, $`\omega _1`$ is the Kähler form we started with and $`\omega _{can}`$ is the closed curvature (1,1)-form of the canonical line bundle of $`M`$ with the Hermitian fibre metric determined by the symplectic volume. Using and (1.1) we will calculate the classifying cohomology class (classifying up to equivalence) of the Berezin-Toeplitz deformation quantization. This class was first calculated by E. Hawkins in by K-theoretic methods with the use of the index theorem for deformation quantization (, ).
In deformation quantization with separation of variables an important role is played by the formal Berezin transform $`fI(f)`$ (see ). In this paper we associate to a deformation quantization with separation of variables also a non-associative ”formal twisted product” $`(f,g)Q(f,g)`$. Here the images are always in the formal power series over the space $`C^{\mathrm{}}(M)`$. In the compact Kähler case by considering all tensor powers $`L^m`$ of the line bundle $`L`$ and with the help of Berezin-Rawnsley’s coherent states , it is possible to introduce for every level $`m`$ the Berezin transform $`I^{(m)}`$ and also some ”twisted product” $`Q^{(m)}`$. The key result of this article is that the analytic asymptotic expansions of $`I^{(m)}`$, resp. of $`Q^{(m)}`$ define formal objects which coincide with $`I`$ and $`Q`$ for some deformation quantization with separation of variables whose classifying form $`\omega `$ is completely determined in terms of the form $`\stackrel{~}{\omega }`$ (Theorem 5.9). To prove this we use the integral representation of the Szegö kernel on a strictly pseudoconvex domain obtained by Boutet de Monvel and Sjöstrand in and a theorem by Zelditch based on . We also use the method of stationary phase and introduce its formal counterpart which we call ”formal integral”.
Since the analytic Berezin transform $`I^{(m)}`$ has the asymptotics given by the formal Berezin transform it follows also that the former has the expansion
(1.2)
$$I^{(m)}=\text{id}+\frac{1}{m}\mathrm{\Delta }+O(\frac{1}{m^2}),$$
where $`\mathrm{\Delta }`$ is the Laplace-Beltrami operator on $`M`$.
It is worth mentioning that the above formal form $`\omega `$ is the formal object corresponding to the asymptotic expansion of the pullback of the Fubini-Study form via Kodaira embedding of $`M`$ into the projective space related to $`L^m`$ as $`m+\mathrm{}`$. This asymptotic expansion was obtained by Zelditch in as a generalization of a theorem by Tian .
The article is organized as follows. In Section 2 we recall the basic notions of deformation quantization and the construction of the deformation quantization with separation of variables given by a formal deformation of a (pseudo-)Kähler form.
In Section 3 formal integrals are introduced. Certain basic properties, like uniqueness are shown.
In Section 4 the covariant and contravariant symbols are introduced. Using Berezin-Toeplitz operators the transformation $`I^{(m)}`$ and the twisted product $`Q^{(m)}`$ are introduced. Integral formulas for them using 2-point, resp. cyclic 3-point functions defined via the scalar product of coherent states are given.
Section 5 contains the key result that $`I^{(m)}`$ and $`Q^{(m)}`$ admit a well-defined asymptotic expansion and that the formal objects corresponding to these expansions are given by $`I`$ and $`Q`$ respectively.
Finally in Section 6 the Berezin-Toeplitz star product is identified with the help of the results obtained in Section 5.
Acknowledgements. We would like to thank Boris Fedosov for interesting discussions and Mirsolav Engliš for bringing the work of Zelditch to our attention. A.K. thanks the Alexander von Humboldt foundation and the DFG for support and the Department of Mathematics at the University of Mannheim for a warm hospitality.
## 2. Deformation quantizations with separation of variables
Given a vector space $`V`$, we call the elements of the space of formal Laurent series with a finite principal part $`V[\nu ^1,\nu ]]`$ formal vectors. In such a way we define formal functions, differential forms, differential operators, etc. However we shall often call these formal objects just functions, operators, and so on, omitting the word formal.
Now assume that $`V`$ is a Hausdorff topological vector space and $`v(m),m,`$ is a family of vectors in $`V`$ which admits an asymptotic expansion as $`m\mathrm{},v(m)_{rr_0}(1/m^r)v_r`$, where $`r_0`$. In order to associate to such asymptotic families the corresponding formal vectors we use the ”formalizer” $`𝔽:v(m)_{rr_0}\nu ^rv_rV[\nu ^1,\nu ]]`$.
Let $`(M,\omega _1)`$ be a real symplectic manifold of dimension $`2n`$. For any open subset $`UM`$ denote by $`(U)=C^{\mathrm{}}(U)[\nu ^1,\nu ]]`$ the space of formal smooth complex-valued functions on $`U`$. Set $`=(M)`$. Denote by $`𝕂=[\nu ^1,\nu ]]`$ the field of formal numbers.
A deformation quantization on $`(M,\omega _1)`$ is an associative $`𝕂`$-algebra structure on $``$, with the product $``$ (named star-product) given for $`f=\nu ^jf_j,g=\nu ^kg_k`$ by the following formula:
(2.1)
$$fg=\underset{r}{}\nu ^r\underset{i+j+k=r}{}C_i(f_j,g_k).$$
In (2.1) $`C_r,r=0,1,\mathrm{}`$, is a sequence of bilinear mappings $`C_r:C^{\mathrm{}}(M)\times C^{\mathrm{}}(M)C^{\mathrm{}}(M)`$ where $`C_0(\phi ,\psi )=\phi \psi `$ and $`C_1(\phi ,\psi )C_1(\psi ,\phi )=i\{\phi ,\psi \}`$ for $`\phi ,\psi C^{\mathrm{}}(M)`$ and $`\{,\}`$ is the Poisson bracket corresponding to the form $`\omega _1`$.
Two deformation quantizations $`(,_1)`$ and $`(,_2)`$ on $`(M,\omega _1)`$ are called equivalent if there exists an isomorphism of algebras $`B:(,_1)(,_2)`$ of the form $`B=1+\nu B_1+\nu ^2B_2+\mathrm{}`$, where $`B_k`$ are linear endomorphisms of $`C^{\mathrm{}}(M)`$.
We shall consider only those deformation quantizations for which the unit constant 1 is the unit in the algebra $`(,)`$.
If all $`C_r,r0`$, are local, i.e., bidifferential operators, then the deformation quantization is called differential. The equivalence classes of differential deformation quantizations on $`(M,\omega _1)`$ are bijectively parametrized by the formal cohomology classes from $`(1/i\nu )[\omega _1]+H^2(M,[[\nu ]])`$. The formal cohomology class parametrizing a star-product $``$ is called the characteristic class of this star-product and denoted $`cl()`$.
A differential deformation quantization can be localized on any open subset $`UM`$. The corresponding star-product on $`(U)`$ will be denoted also $``$.
For $`f,g`$ denote by $`L_f,R_g`$ the operators of left and right multiplication by $`f,g`$ respectively in the algebra $`(,)`$, so that $`L_fg=fg=R_gf`$. The associativity of the star-product $``$ is equivalent to the fact that $`L_f`$ commutes with $`R_g`$ for all $`f,g`$. If a deformation quantization is differential then $`L_f,R_g`$ are formal differential operators.
Now let $`(M,\omega _1)`$ be pseudo-Kähler, i.e., a complex manifold such that the form $`\omega _1`$ is of type (1,1) with respect to the complex structure. We say that a differential deformation quantization $`(,)`$ is a deformation quantization with separation of variables if for any open subset $`UM`$ and any holomorphic function $`a`$ and antiholomorphic function $`b`$ on $`U`$ the operators $`L_a`$ and $`R_b`$ are the operators of point-wise multiplication by $`a`$ and $`b`$ respectively, i.e., $`L_a=a`$ and $`R_b=b`$.
A formal form $`\omega =(1/\nu )\omega _1+\omega _0+\nu \omega _1+\mathrm{}`$ is called a formal deformation of the form $`(1/\nu )\omega _1`$ if the forms $`\omega _r,r0`$, are closed but not necessarily nondegenerate (1,1)-forms on $`M`$.
It was shown in that all deformation quantizations with separation of variables on a pseudo-Kähler manifold $`(M,\omega _1)`$ are bijectively parametrized by the formal deformations of the form $`(1/\nu )\omega _1`$.
Recall how the star-product with separation of variables $``$ on $`M`$ corresponding to the formal form $`\omega =(1/\nu )\omega _1+\omega _0+\nu \omega _1+\mathrm{}`$ is constructed. For an arbitrary contractible coordinate chart $`UM`$ with holomorphic coordinates $`\{z^k\}`$ let $`\mathrm{\Phi }=(1/\nu )\mathrm{\Phi }_1+\mathrm{\Phi }_0+\nu \mathrm{\Phi }_1+\mathrm{}`$ be a formal potential of the form $`\omega `$ on $`U`$, i.e., $`\omega =i\overline{}\mathrm{\Phi }`$ (notice that in \- a potential $`\mathrm{\Phi }`$ of a closed (1,1)-form $`\omega `$ is defined via the formula $`\omega =i\overline{}\mathrm{\Phi }`$).
The star-product corresponding to the form $`\omega `$ is such that $`L_{\mathrm{\Phi }/z^k}=\mathrm{\Phi }/z^k+/z^k`$ and $`R_{\mathrm{\Phi }/\overline{z}^l}=\mathrm{\Phi }/\overline{z}^l+/\overline{z}^l`$ on $`U`$. The set $`(U)`$ of all left multiplication operators on $`U`$ is completely described as the set of all formal differential operators commuting with the point-wise multiplication operators by antiholomorphic coordinates $`R_{\overline{z}^l}=\overline{z}^l`$ and the operators $`R_{\mathrm{\Phi }/\overline{z}^l}=\mathrm{\Phi }/\overline{z}^l+/\overline{z}^l`$. One can immediately reconstruct the star-product on $`U`$ from the knowledge of $`(U)`$. The local star-products agree on the intersections of the charts and define the global star-product $``$ on $`M`$.
One can express the characteristic class $`cl()`$ of the star-product with separation of variables $``$ parametrized by the formal form $`\omega `$ in terms of this form (see ). Unfortunately, there were wrong signs in the formula for $`cl()`$ in which should be read as follows:
(2.2)
$$cl()=(1/i)([\omega ]\epsilon /2),$$
where $`\epsilon `$ is the canonical class of the complex manifold $`M`$, i.e., the first Chern class of the canonical holomorphic line bundle on $`M`$.
Given a deformation quantization with separation of variables $`(,)`$ on the pseudo-Kähler manifold $`(M,\omega _1)`$, one can introduce the formal Berezin transform $`I`$ as the unique formal differential operator on $`M`$ such that for any open subset $`UM`$, holomorphic function $`a`$ and antiholomorphic function $`b`$ on $`U`$ the relation $`I(ab)=ba`$ holds (see ). One can check that $`I=1+\nu \mathrm{\Delta }+\mathrm{}`$, where $`\mathrm{\Delta }`$ is the Laplace-Beltrami operator corresponding to the pseudo-Kähler metric on $`M`$. The dual star-product $`\stackrel{~}{}`$ on $`M`$ defined for $`f,g`$ by the formula $`f\stackrel{~}{}g=I^1(IgIf)`$ is a star-product with separation of variables on the pseudo-Kähler manifold $`(M,\omega _1)`$. For this deformation quantization the formal Berezin transform equals $`I^1`$, and thus the dual to $`\stackrel{~}{}`$ is again $``$.
Denote by $`\stackrel{~}{\omega }=(1/\nu )\omega _1+\stackrel{~}{\omega }_0+\nu \stackrel{~}{\omega }_1+\mathrm{}`$ the formal form parametrizing the star-product $`\stackrel{~}{}`$. The opposite to the dual star-product, $`^{}=\stackrel{~}{}^{op}`$, given by the formula $`f^{}g=I^1(IfIg)`$, also defines a deformation quantization with separation of variables on $`M`$ but with the roles of holomorphic and antiholomorphic variables swapped. Differently said, $`(,^{})`$ is a deformation quantization with separation of variables on the pseudo-Kähler manifold $`(\overline{M},\omega _1)`$ where $`\overline{M}`$ is the manifold $`M`$ with the opposite complex structure. The formal Berezin transform $`I`$ establishes an equivalence of deformation quantizations $`(,)`$ and $`(,^{})`$.
Introduce the following non-associative operation $`Q(,)`$ on $``$. For $`f,g`$ set $`Q(f,g)=IfIg=I(f^{}g)=I(g\stackrel{~}{}f)`$. We shall call it formal twisted product. The importance of the formal twisted product will be revealed later.
A trace density of a deformation quantization $`(,)`$ on a symplectic manifold $`M`$ is a formal volume form $`\mu `$ on $`M`$ for which the functional $`\kappa (f)=_Mf\mu ,f,`$ has the trace property, $`\kappa (fg)=\kappa (gf)`$ for all $`f,g`$ where at least one of the functions $`f,g`$ has compact support. It was shown in that on a local holomorphic chart $`(U,\{z^k\})`$ any formal trace density $`\mu `$ can be represented in the form $`c(\nu )\mathrm{exp}(\mathrm{\Phi }+\mathrm{\Psi })dzd\overline{z}`$, where $`c(\nu )𝕂`$ is a formal constant, $`dzd\overline{z}=dz^1\mathrm{}dz^nd\overline{z}^1\mathrm{}d\overline{z}^n`$ is the standard volume on $`U`$ and $`\mathrm{\Phi }=(1/\nu )\mathrm{\Phi }_1+\mathrm{},\mathrm{\Psi }=(1/\nu )\mathrm{\Psi }_1+\mathrm{}`$ are formal potentials of the forms $`\omega ,\stackrel{~}{\omega }`$ respectively such that the relations
(2.3)
$$\mathrm{\Phi }/z^k=I(\mathrm{\Psi }/z^k),\mathrm{\Phi }/\overline{z}^l=I(\mathrm{\Psi }/\overline{z}^l),\text{ and }\mathrm{\Phi }_1+\mathrm{\Psi }_1=0$$
hold. Vice versa, any such form is a formal trace density.
## 3. Formal integrals, jets, and almost analytic functions
Let $`\varphi =(1/\nu )\varphi _1+\varphi _0+\nu \varphi _1+\mathrm{}`$ and $`\mu =\mu _0+\nu \mu _1+\mathrm{}`$ be, respectively, a smooth complex-valued formal function and a smooth formal volume form on an open set $`U^n`$. Assume that $`xU`$ is a nondegenerate critical point of the function $`\varphi _1`$ and $`\mu _0`$ does not vanish at $`x`$. We call a $`𝕂`$-linear functional $`K`$ on $`(U)`$ such that
* $`K=K_0+\nu K_1+\mathrm{}`$ is a formal distribution supported at the point $`x`$;
* $`K_0=\delta _x`$ is the Dirac distribution at the point $`x`$;
* $`K(1)=1`$ (normalization condition);
* for any vector field $`\xi `$ on $`U`$ and $`f(U)K\left(\xi f+(\xi \varphi +\mathrm{div}_\mu \xi )f\right)=0`$,
a (normalized) formal integral at the point $`x`$ associated to the pair $`(\varphi ,\mu )`$.
It is clear from the definition that a formal integral at a point $`x`$ is independent of a particular choice of the neighborhood $`U`$ and is actually associated to the germs of $`(\varphi ,\mu )`$ at $`x`$. Usually we shall consider a contractible neighborhood $`U`$ such that $`\mu _0`$ vanishes nowhere on $`U`$.
We shall prove that a formal integral at the point $`x`$ associated to the pair $`(\varphi ,\mu )`$ is uniquely determined. One can also show the existence of such a formal integral, but this fact will neither be used nor proved in what follows.
We call two pairs $`(\varphi ,\mu )`$ and $`(\varphi ^{},\mu ^{})`$ equivalent if there exists a formal function $`u=u_0+\nu u_1+\mathrm{}`$ on $`U`$ such that $`\varphi ^{}=\varphi u,\mu ^{}=e^u\mu `$.
Since the expression $`\xi \varphi +\mathrm{div}_\mu \xi `$ remains invariant if we replace the pair $`(\varphi ,\mu )`$ by an equivalent one, a formal integral is actually associated to the equivalence class of the pair $`(\varphi ,\mu )`$. This means that a formal integral actually depends on the product $`e^\varphi \mu `$ which can be thought of as a part of the integrand of a ”formal oscillatory integral”. In the sequel it will be shown that one can directly produce formal integrals from the method of stationary phase.
Notice that if $`K`$ is a formal integral associated to a pair $`(\varphi ,\mu )`$ it is then associated to any pair $`(\varphi ,c(\nu )\mu )`$, where $`c(\nu )`$ is a nonzero formal constant.
It is easy to show that it is enough to check condition (d) for the coordinate vector fields $`/x^k`$ on $`U`$. Moreover, if $`U`$ is contractible and such that $`\mu _0`$ vanishes nowhere on it, one can choose an equivalent pair of the form $`(\varphi ^{},dx)`$, where $`dx=dx^1\mathrm{}dx^n`$ is the standard volume form.
###### Proposition 3.1.
A formal integral $`K=K_0+\nu K_1+`$ at a point $`x`$, associated to a pair $`(\varphi =(1/\nu )\varphi _1+\varphi _0+\nu \varphi _1+\mathrm{},\mu )`$ is uniquely determined.
###### Proof.
We assume that $`K`$ is defined on a coordinate chart $`(U,\{x^k\}),\mu =dx`$, and take $`fC^{\mathrm{}}(U)`$. Since $`\mathrm{div}_{dx}(/x^k)=0`$, the last condition of the definition of a formal integral takes the form
(3.1)
$$K\left(f/x^k+(\varphi /x^k)f\right)=0.$$
Equating to zero the coefficient at $`\nu ^r,r0`$, of the l.h.s. of (3.1) we get $`K_r(f/x^k)+_{s=0}^{r+1}K_s\left((\varphi _{rs}/x^k)f\right)=0`$, which can be rewritten as a recurrent equation
(3.2)
$$K_{r+1}\left((\varphi _1/x^k)f\right)=\text{r.h.s. depending on }K_j,jr.$$
Since $`x`$ is a nondegenerate critical point of $`\varphi _1`$, the functions $`\varphi _1/x^k`$ generate the ideal of functions vanishing at $`x`$. Taking into account that $`K_{r+1}(1)=0`$ for $`r0`$ we see from (3.2) that $`K_{r+1}`$ is determined uniquely. Thus the proof proceeds by induction. ∎
Let $`V`$ be an open subset of a complex manifold $`M`$ and $`Z`$ be a relatively closed subset of $`V`$. A function $`fC^{\mathrm{}}(V)`$ is called almost analytic at $`Z`$ if $`\overline{}f`$ vanishes to infinite order there.
Two functions $`f_1,f_2C^{\mathrm{}}(V)`$ are called equivalent at $`Z`$ if $`f_1f_2`$ vanishes to infinite order there.
Consider open subsets $`U^n`$ and $`\stackrel{~}{U}^n`$ such that $`U=\stackrel{~}{U}^n`$, and a function $`fC^{\mathrm{}}(U)`$. A function $`\stackrel{~}{f}C^{\mathrm{}}(\stackrel{~}{U})`$ is called an almost analytic extension of $`f`$ if it is almost analytic at $`U`$ and $`\stackrel{~}{f}|_U=f`$.
It is well known that every $`fC^{\mathrm{}}(U)`$ has an almost analytic extension uniquely determined up to equivalence.
Fix a formal deformation $`\omega =(1/\nu )\omega _1+\omega _0+\nu \omega _1+\mathrm{}`$ of the form $`(1/\nu )\omega _1`$ on a pseudo-Kähler manifold $`(M,\omega _1)`$. Consider the corresponding star-product with separation of variables $``$, the formal Berezin transform $`I`$ and the formal twisted product $`Q`$ on $`M`$. We are going to show that for any point $`xM`$ the functional $`K_x^I(f)=(If)(x)`$ on $``$ and the functional $`K_x^Q`$ on $`(M\times M)`$ such that $`K_x^Q(fg)=Q(f,g)(x)`$ can be represented as formal integrals.
Let $`UM`$ be a contractible coordinate chart with holomorphic coordinates $`\{z^k\}`$. Given a smooth function $`f(z,\overline{z})`$ on $`U`$, where $`U`$ is considered as the diagonal of $`\stackrel{~}{U}=U\times \overline{U}`$, one can choose its almost analytic extension $`\stackrel{~}{f}(z_1,\overline{z}_1,z_2,\overline{z}_2)`$ on $`\stackrel{~}{U}`$, so that $`\stackrel{~}{f}(z,\overline{z},z,\overline{z})=f(z,\overline{z})`$. It is a substitute of the holomorphic function $`f(z_1,\overline{z}_2)`$ on $`\stackrel{~}{U}`$ which in general does not exist.
Let $`\mathrm{\Phi }=(1/\nu )\mathrm{\Phi }_1+\mathrm{\Phi }_0+\nu \mathrm{\Phi }_1+\mathrm{}`$ be a formal potential of the form $`\omega `$ on $`U`$ and $`\stackrel{~}{\mathrm{\Phi }}`$ its almost analytic extension on $`\stackrel{~}{U}`$. In particular, $`\stackrel{~}{\mathrm{\Phi }}(x,x)=\mathrm{\Phi }(x)`$ for $`xU`$. Introduce an analogue of the Calabi diastatic function on $`U\times U`$ by the formula $`D(x,y)=\stackrel{~}{\mathrm{\Phi }}(x,y)+\stackrel{~}{\mathrm{\Phi }}(y,x)\mathrm{\Phi }(x)\mathrm{\Phi }(y)`$. We shall also use the notation $`D_k(x,y)=\stackrel{~}{\mathrm{\Phi }}_k(x,y)+\stackrel{~}{\mathrm{\Phi }}_k(y,x)\mathrm{\Phi }_k(x)\mathrm{\Phi }_k(y)`$ so that $`D=(1/\nu )D_1+D_0+\nu D_1+\mathrm{}`$.
Let $`\stackrel{~}{\omega }`$ be the formal form corresponding to the dual star-product $`\stackrel{~}{}`$ of the star-product $``$. Choose a formal potential $`\mathrm{\Psi }`$ of the form $`\stackrel{~}{\omega }`$ on $`U`$, satisfying equation (2.3), so that $`\mu _{tr}=e^{\mathrm{\Phi }+\mathrm{\Psi }}dzd\overline{z}`$ is a formal trace density of the star-product $``$ on $`U`$.
###### Theorem 3.2.
For any point $`xU`$ the functional $`K_x^I(f)=(If)(x)`$ on $`(U)`$ is the formal integral at $`x`$ associated to the pair $`(\varphi ^x,\mu _{tr})`$, where $`\varphi ^x(y)=D(x,y)`$.
###### Remark.
In the proof of the theorem we use the notion of jet of order $`N`$ of a formal function $`f=\nu ^rf_r`$ at a given point. It is also a formal object, the formal series of jets of order $`N`$ of the functions $`f_r`$.
###### Proof.
The condition that $`x`$ is a nondegenerate critical point of the function $`\varphi _1^x(y)=D_1(x,y)`$ directly follows from the fact that $`\mathrm{\Phi }_1`$ is a potential of the non-degenerate (1,1)-form $`\omega _1`$. The conditions (a-c) of the definition of formal integral are trivially satisfied. It remains to check the condition (d). Replace the pair $`(\varphi ^x,\mu _{tr})`$ by the equivalent pair $`(\varphi ^x+\mathrm{\Phi }+\mathrm{\Psi },dzd\overline{z})=(\stackrel{~}{\mathrm{\Phi }}(x,y)+\stackrel{~}{\mathrm{\Phi }}(y,x)\mathrm{\Phi }(x)+\mathrm{\Psi }(y),dzd\overline{z})`$. Put $`x=(z_0,\overline{z}_0),y=(z,\overline{z})`$. For $`\xi =/z^k`$ the condition (d) takes the form
$$I\left(f/z^k+(/z^k)\left(\stackrel{~}{\mathrm{\Phi }}(z_0,\overline{z}_0,z,\overline{z})+\stackrel{~}{\mathrm{\Phi }}(z,\overline{z},z_0,\overline{z}_0)+\mathrm{\Psi }(z,\overline{z})\right)f\right)(z_0,\overline{z}_0)=0.$$
We shall check it by showing that
(i) $`I\left(f/z^k+(\mathrm{\Psi }/z^k)f\right)=If(\mathrm{\Phi }/z^k)`$;
(ii) $`I\left(\left(\stackrel{~}{\mathrm{\Phi }}(z_0,\overline{z}_0,z,\overline{z})/z^k\right)f\right)(z_0,\overline{z}_0)=0`$;
(iii) $`I\left(\left(\stackrel{~}{\mathrm{\Phi }}(z,\overline{z},z_0,\overline{z}_0)/z^k\right)f\right)(z_0,\overline{z}_0)=\left(If\mathrm{\Phi }/z^k\right)(z_0,\overline{z}_0)`$.
First, $`I\left(f/z^k+(\mathrm{\Psi }/z^k)f\right)=I\left((\mathrm{\Psi }/z^k)\stackrel{~}{}f\right)=IfI(\mathrm{\Psi }/z^k)=If(\mathrm{\Phi }/z^k)`$, which proves (i).
The function $`\psi (z,\overline{z})=\stackrel{~}{\mathrm{\Phi }}(z_0,\overline{z}_0,z,\overline{z})`$ is almost antiholomorphic at the point $`z=z_0`$. Thus, the full jet of the function $`\psi /z^k`$ at the point $`z=z_0`$ is equal to zero, which proves (ii).
The function $`\theta (z,\overline{z})=\stackrel{~}{\mathrm{\Phi }}(z,\overline{z},z_0,\overline{z}_0)/z^k`$ is almost holomorphic at the point $`z=z_0`$. For a holomorphic function $`a`$ we have $`I(af)=I(a\stackrel{~}{}f)=IfIa=Ifa`$. Since $`I(\theta f)(z_0,\overline{z}_0)`$ and $`(If\theta )(z_0,\overline{z}_0)`$ considered modulo $`\nu ^N`$ depend on the jets of finite order of the functions $`\theta `$ and $`f`$ at the point $`z_0`$ taken modulo $`\nu ^N^{}`$ for sufficiently big $`N^{}`$, we can approximate $`\theta `$ by a formal holomorphic function $`a`$ making sure that the jets of sufficiently high order of $`\theta `$ and $`a`$ at the point $`z_0`$ coincide modulo $`\nu ^N^{}`$. Then $`I(\theta f)(z_0,\overline{z}_0)I(af)(z_0,\overline{z}_0)(Ifa)(z_0,\overline{z}_0)(If\theta )(z_0,\overline{z}_0)(mod\nu ^N)`$. Since $`N`$ is arbitrary, $`I(\theta f)(z_0,\overline{z}_0)=(If\theta )(z_0,\overline{z}_0)`$ identically. The functions $`\mathrm{\Phi }/z^k`$ and $`\theta `$ have identical holomorphic parts of jets at the point $`z_0`$, i.e., all the holomorphic partial derivatives (of any order) of these functions at the point $`z_0`$ coincide. Since a left star-multiplication operator of deformation quantization with separation of variables differentiates its argument only in holomorphic directions, we get that $`(If\theta )(z_0,\overline{z}_0)=(If(\mathrm{\Phi }/z^k))(z_0,\overline{z}_0)`$. This proves (iii).
The check for $`\xi =/\overline{z}^l`$ is similar, which completes the proof of the theorem. ∎
The following lemma and theorem can be proved by the same methods as Theorem 3.2.
###### Lemma 3.3.
For any vector field $`\xi `$ on $`U`$ and $`xUI(\xi _x\varphi ^x)(x)=0`$, where $`\varphi ^x(y)=D(x,y)`$.
($`\xi _x\varphi ^x`$ denotes differentiation of $`\varphi ^x`$ w.r.t. the parameter $`x`$.)
Introduce a 3-point function $`T`$ on $`U\times U\times U`$ by the formula $`T(x,y,z)=\stackrel{~}{\mathrm{\Phi }}(x,y)+\stackrel{~}{\mathrm{\Phi }}(y,z)+\stackrel{~}{\mathrm{\Phi }}(z,x)\mathrm{\Phi }(x)\mathrm{\Phi }(y)\mathrm{\Phi }(z)`$.
###### Theorem 3.4.
For any point $`xU`$ the functional $`K_x^Q`$ on $`(U\times U)`$ such that $`K_x^Q(fg)=Q(f,g)(x)`$ is the formal integral at the point $`(x,x)U\times U`$ associated to the pair $`(\psi ^x,\mu _{tr}\mu _{tr})`$, where $`\psi ^x(y,z)=T(x,y,z)`$.
## 4. Covariant and contravariant symbols
In the rest of the paper let $`(M,\omega _1)`$ be a compact Kähler manifold. Assume that there exists a quantum line bundle $`(L,h)`$ on $`M`$, i.e., a holomorphic hermitian line bundle with fibre metric $`h`$ such that the curvature of the canonical connection on $`L`$ coincides with the Kähler form $`\omega _1`$.
Let $`m`$ be a non-negative integer. The metric $`h`$ induces the fibre metric $`h^m`$ on the tensor power $`L^m=L^m`$. Denote by $`L^2(L^m)`$ the Hilbert space of square-integrable sections of $`L^m`$ with respect to the norm $`s^2=h^m(s)\mathrm{\Omega }`$, where $`\mathrm{\Omega }=(1/n!)(\omega _1)^n`$ is the symplectic volume form on $`M`$. The Bergman projector $`B_m`$ is the orthogonal projector in $`L^2(L^m)`$ onto the space $`H_m=\mathrm{\Gamma }_{hol}(L^m)`$ of holomorphic sections of $`L^m`$.
Denote by $`k`$ the metric on the dual line bundle $`\tau :L^{}M`$ induced by $`h`$. It is a well known fact that $`D=\{\alpha L^{}|k(\alpha )<1\}`$ is a strictly pseudoconvex domain in $`L^{}`$. Its boundary $`X=\{\alpha L^{}|k(\alpha )=1\}`$ is a $`S^1`$-principal bundle.
The sections of $`L^m`$ are identified with the $`m`$-homogeneous functions on $`L^{}`$ by means of the mapping $`\gamma _m:s\psi _s`$, where $`\psi _s(\alpha )=\alpha ^m,s(x)`$ for $`\alpha L_x^{}`$. Here $`,`$ denotes the bilinear pairing between $`(L^{})^m`$ and $`L^m`$.
There exists a unique $`S^1`$-invariant volume form $`\stackrel{~}{\mathrm{\Omega }}`$ on $`X`$ such that for every $`fC^{\mathrm{}}(M)`$ the equality $`_X(\tau ^{}f)\stackrel{~}{\mathrm{\Omega }}=_Mf\mathrm{\Omega }`$ holds.
The mapping $`\gamma _m`$ maps $`L^2(L^m)`$ isometrically onto the weight subspace of $`L^2(X,\stackrel{~}{\mathrm{\Omega }})`$ of weight $`m`$ with respect to the $`S^1`$-action. The Hardy space $`L^2(X,\stackrel{~}{\mathrm{\Omega }})`$ of square integrable traces of holomorphic functions on $`L^{}`$ splits up into weight spaces, $`=_{m=0}^{\mathrm{}}_m`$, where $`_m=\gamma _m(H_m)`$.
Denote by $`S`$ and $`\widehat{B}_m`$ the Szegö and Bergman orthogonal projections in $`L^2(X,\stackrel{~}{\mathrm{\Omega }})`$ onto $``$ and $`_m`$ respectively. Thus $`S=_{m=0}^{\mathrm{}}\widehat{B}_m`$. The Bergman projection $`\widehat{B}_m`$ has a smooth integral kernel $`_m=_m(\alpha ,\beta )`$ on $`X\times X`$.
For each $`\alpha L^{}0`$ (’$`0`$’ means the zero section removed) one can define a coherent state $`e_\alpha ^{(m)}`$ as the unique holomorphic section of $`L^m`$ such that for each $`sH_ms,e_\alpha ^{(m)}=\psi _s(\alpha )`$ where $`,`$ is the hermitian scalar product on $`L^2(L^m)`$ antilinear in the second argument.
Since the line bundle $`L`$ is positive it is known that there exists a constant $`m_0`$ such that for $`m>m_0dimH_m>0`$ and all $`e_\alpha ^{(m)},\alpha L^{}0`$, are nonzero vectors. From now on we assume that $`m>m_0`$ unless otherwise specified.
The coherent state $`e_\alpha ^{(m)}`$ is antiholomorphic in $`\alpha `$ and for a nonzero $`ce_{c\alpha }^{(m)}=\overline{c}^me_\alpha ^{(m)}`$. Notice that in coherent states are parametrized by the points of $`L0`$.
For $`sL^2(L^m)s,e_\alpha ^{(m)}=s,B_me_\alpha ^{(m)}=B_ms,e_\alpha ^{(m)}=\psi _{B_ms}(\alpha )`$. The mapping $`\gamma _m`$ intertwines the Bergman projectors $`B_m`$ and $`\widehat{B}_m`$, for $`sL^2(L^m)\psi _{B_ms}=\widehat{B}_m\psi _s`$. Thus, on the one hand, $`s,e_\alpha ^{(m)}=\widehat{B}_m\psi _s(\alpha )=_X_m(\alpha ,\beta )\psi _s(\beta )\stackrel{~}{\mathrm{\Omega }}(\beta )`$. On the other hand, $`s,e_\alpha ^{(m)}=\psi _s,\psi _{e_\alpha ^{(m)}}=_X\psi _s(\beta )\overline{\psi _{e_\alpha ^{(m)}}(\beta )}\stackrel{~}{\mathrm{\Omega }}(\beta )`$. Taking into account that $`e_\beta ^{(m)},e_\alpha ^{(m)}=\psi _{e_\beta ^{(m)}}(\alpha )=\overline{\psi _{e_\alpha ^{(m)}}(\beta )}`$ we finally get that $`e_\beta ^{(m)},e_\alpha ^{(m)}=\psi _{e_\beta ^{(m)}}(\alpha )=_m(\alpha ,\beta )`$. In particular, one can extend the kernel $`_m(\alpha ,\beta )`$ from $`X\times X`$ to a holomorphic function on $`(L^{}0)\times (\overline{L^{}0})`$ such that for nonzero $`c,d`$
(4.1)
$$_m(c\alpha ,d\beta )=(c\overline{d})^m_m(\alpha ,\beta ).$$
For $`\alpha ,\beta L^{}0`$ the following inequality holds.
(4.2)
$$\left|_m(\alpha ,\beta )\right|=\left|e_\alpha ^{(m)},e_\beta ^{(m)}\right|e_\alpha ^{(m)}e_\beta ^{(m)}=(_m(\alpha ,\alpha )_m(\beta ,\beta ))^{\frac{1}{2}}.$$
The covariant symbol of an operator $`A`$ in the space $`H_m`$ is the function $`\sigma (A)`$ on $`M`$ such that
$$\sigma (A)(x)=\frac{Ae_\alpha ^{(m)},e_\alpha ^{(m)}}{e_\alpha ^{(m)},e_\alpha ^{(m)}}$$
for any $`\alpha L_x^{}0`$.
Denote by $`M_f`$ the multiplication operator by a function $`fC^{\mathrm{}}(M)`$ on sections of $`L^m`$. Define the Berezin-Toeplitz operator $`T_f^{(m)}=B_mM_fB_m`$ in $`H_m`$. If an operator in $`H_m`$ is represented in the form $`T_f^{(m)}`$ for some function $`fC^{\mathrm{}}(M)`$ then the function $`f`$ is called its contravariant symbol.
With these symbols we associate two important operations on $`C^{\mathrm{}}(M)`$, the Berezin transform $`I^{(m)}`$ and a non-associative binary operation $`Q^{(m)}`$ which we call twisted product, as follows. For $`f,gC^{\mathrm{}}(M)I^{(m)}f=\sigma (T_f^{(m)}),Q^{(m)}(f,g)=\sigma (T_f^{(m)}T_g^{(m)})`$.
We are going to show in Section 5 that both $`I^{(m)}`$ and $`Q^{(m)}`$ have asymptotic expansions in $`1/m`$ as $`m+\mathrm{}`$, such that if the asymptotic parameter $`1/m`$ in these expansions is replaced by the formal parameter $`\nu `$ then we get the formal Berezin transform $`I`$ and the formal twisted product $`Q`$ corresponding to some deformation quantization with separation of variables on $`(M,\omega _1)`$ which can be completely identified. We shall mainly be interested in the opposite to its dual deformation quantization. The goal of this paper is to show that it coincides with the Berezin-Toeplitz deformation quantization obtained in ,.
In order to obtain the asymptotic expansions of $`I^{(m)}`$ and $`Q^{(m)}`$ we need their integral representations. To calculate them it is convenient to work on $`X`$ rather than on $`M`$. We shall use the fact that for $`fC^{\mathrm{}}(M),s\mathrm{\Gamma }(L^m),\psi _{M_fs}=(\tau ^{}f)\psi _s`$. For $`xM`$ denote by $`X_x`$ the fibre of the bundle $`X`$ over $`x`$, $`X_x=\tau ^1(x)X`$. For $`x,y,zM`$ choose $`\alpha X_x,\beta X_y,\gamma X_z`$ and set
(4.3)
$$\begin{array}{c}u_m(x)=_m(\alpha ,\alpha ),v_m(x,y)=_m(\alpha ,\beta )_m(\beta ,\alpha ),\hfill \\ \hfill w_m(x,y,z)=_m(\alpha ,\beta )_m(\beta ,\gamma )_m(\gamma ,\alpha ).\end{array}$$
It follows from (4.1) that $`u_m(x),v_m(x,y),w_m(x,y,z)`$ do not depend on the choice of $`\alpha ,\beta ,\gamma `$ and thus relations (4.3) correctly define functions $`u_m,v_m,w_m`$. The function $`w_m`$ is the so called cyclic 3-point function studied in . Notice that $`u_m(x)=_m(\alpha ,\alpha )=e_\alpha ^{(m)}^2>0,v_m(x,y)=_m(\alpha ,\beta )_m(\beta ,\alpha )=|_m(\alpha ,\beta )|^20`$ and
(4.4)
$$|w_m(x,y,z)|^2=v_m(x,y)v_m(y,z)v_m(z,x).$$
It follows from (4.2) that
(4.5)
$$v_m(x,y)u_m(x)u_m(y).$$
For $`\alpha X_x`$ we have
(4.6)
$$\begin{array}{c}\left(I^{(m)}f\right)(x)=\sigma \left(T_f^{(m)}\right)(x)=\frac{T_f^{(m)}e_\alpha ^{(m)},e_\alpha ^{(m)}}{e_\alpha ^{(m)},e_\alpha ^{(m)}}=\frac{B_mM_fB_me_\alpha ^{(m)},e_\alpha ^{(m)}}{_m(\alpha ,\alpha )}=\hfill \\ \hfill \frac{M_fe_\alpha ^{(m)},e_\alpha ^{(m)}}{_m(\alpha ,\alpha )}=\frac{(\tau ^{}f)\psi _{e_\alpha ^{(m)}},\psi _{e_\alpha ^{(m)}}}{_m(\alpha ,\alpha )}=\frac{1}{_m(\alpha ,\alpha )}_X(\tau ^{}f)\psi _{e_\alpha ^{(m)}}(\beta )\overline{\psi _{e_\alpha ^{(m)}}(\beta )}\stackrel{~}{\mathrm{\Omega }}(\beta )=\\ \hfill \frac{1}{_m(\alpha ,\alpha )}_X_m(\alpha ,\beta )_m(\beta ,\alpha )(\tau ^{}f)(\beta )\stackrel{~}{\mathrm{\Omega }}(\beta )=\frac{1}{u_m(x)}_Mv_m(x,y)f(y)\mathrm{\Omega }(y).\end{array}$$
Similarly we obtain that
(4.7)
$$\begin{array}{c}Q^{(m)}(f,g)(x)=\hfill \\ \hfill \frac{1}{_m(\alpha ,\alpha )}_{X\times X}_m(\alpha ,\beta )_m(\beta ,\gamma )_m(\gamma ,\alpha )(\tau ^{}f)(\beta )(\tau ^{}g)(\gamma )\stackrel{~}{\mathrm{\Omega }}(\beta )\stackrel{~}{\mathrm{\Omega }}(\gamma )=\\ \hfill \frac{1}{u_m(x)}_{M\times M}w_m(x,y,z)f(y)g(z)\mathrm{\Omega }(y)\mathrm{\Omega }(z).\end{array}$$
## 5. Asymptotic expansion of the Berezin transform
In a microlocal description of the integral kernel $`𝒮`$ of the Szegö projection $`S`$ was given. The results in were obtained for a strictly pseudoconvex domain with a smooth boundary in $`^{n+1}`$. However, according to the concluding remarks in , these results are still valid for the domain $`D`$ in $`L^{}`$ (see also , ).
It was proved in that the Szegö kernel $`𝒮`$ is a generalized function on $`X\times X`$ singular on the diagonal of $`X\times X`$ and smooth outside the diagonal. The Szegö kernel $`𝒮`$ can be expressed via the Bergman kernels $`_m`$ as follows, $`𝒮=_{m0}_m`$, where the sum should be understood as a sum of generalized functions.
For $`(\alpha ,\beta )X\times X`$ and $`\theta `$ set $`r_\theta (\alpha ,\beta )=(e^{i\theta }\alpha ,\beta )`$. Since each $`_m`$ is a weight space of the $`S^1`$-action in the Hardy space $``$, one can recover $`_m`$ from the Szegö kernel,
(5.1)
$$_m=\frac{1}{2\pi }_0^{2\pi }e^{im\theta }r_\theta ^{}𝒮𝑑\theta .$$
This equality should be understood in the weak sense.
Let $`E_1,E_2`$ be closed disjoint subsets of $`M`$. Set $`F_i=\tau ^1(E_i)X,i=1,2`$. Thus $`F_1,F_2`$ are closed disjoint subsets of $`X`$ or, equivalently, $`F_1\times F_2`$ is a closed subset of $`X\times X`$ which does not intersect the diagonal. For $`𝒮`$ and $`_m`$ considered as smooth functions outside the diagonal of $`X\times X`$ equality (5.1) holds in the ordinary sense, from whence it follows immediately that
(5.2)
$$\underset{F_1\times F_2}{sup}|_m|=O\left(\frac{1}{m^N}\right)$$
for any $`N`$.
Now let $`E`$ be a closed subset of $`M`$ and $`xME`$. Then (5.2) implies that
(5.3)
$$\underset{yE}{sup}v_m(x,y)=O\left(\frac{1}{m^N}\right)$$
for any $`N`$.
In Zelditch proved that the function $`u_m`$ on $`M`$ expands in the asymptotic series $`u_mm^n_{r0}(1/m^r)b_r`$ as $`m+\mathrm{}`$, where $`b_0=1`$ ($`n=(1/2)dim_{}M`$). More precisely, he proved that for any $`k,N`$
(5.4)
$$\left|u_m\underset{r=0}{\overset{N1}{}}m^{nr}b_r\right|_{C^k}=O(m^{nN}).$$
Therefore
(5.5)
$$\underset{M}{sup}\frac{1}{u_m}=O\left(\frac{1}{m^n}\right).$$
Using (4.6),(5.3) and (5.5) it is easy to prove the following proposition.
###### Proposition 5.1.
Let $`fC^{\mathrm{}}(M)`$ be a function vanishing in a neighborhood of a point $`xM`$. Then $`|(I^{(m)}f)(x)|=O(1/m^N)`$ for any $`N`$, i.e., $`(I^{(m)}f)(x)`$ is rapidly decreasing as $`m+\mathrm{}`$.
Thus for arbitrary $`fC^{\mathrm{}}(M)`$ and $`xM`$ the asymptotics of $`(I^{(m)}f)(x)`$ as $`m+\mathrm{}`$ depends only on the germ of the function $`f`$ at the point $`x`$.
Let $`E`$ be a closed subset of $`M`$. Fix a point $`xME`$. The function $`w_m(x,y,z)`$ with $`yE`$ can be estimated using (4.4) and (4.5) as follows.
(5.6)
$$|w_m(x,y,z)|^2v_m(x,y)u_m(x)u_m(y)\left(u_m(z)\right)^2.$$
Using (5.3), (5.4) and (5.6) we obtain that for any $`N`$
(5.7)
$$\underset{yE,zM}{sup}\left|w_m(x,y,z)\right|=O\left(\frac{1}{m^N}\right).$$
Similarly,
(5.8)
$$\underset{yM,zE}{sup}\left|w_m(x,y,z)\right|=O\left(\frac{1}{m^N}\right)$$
for any $`N`$.
Using (4.7), (5.5), (5.7) and (5.8) one can readily prove the following proposition.
###### Proposition 5.2.
For $`xM`$ and arbitrary functions $`f,gC^{\mathrm{}}(M)`$ such that $`f`$ or $`g`$ vanishes in a neighborhood of $`xQ^{(m)}(f,g)(x)`$ is rapidly decreasing as $`m+\mathrm{}`$.
This statement can be reformulated as follows. For arbitrary $`f,gC^{\mathrm{}}(M)`$ and $`xM`$ the asymptotics of $`Q^{(m)}(f,g)(x)`$ as $`m+\mathrm{}`$ depends only on the germs of the functions $`f,g`$ at the point $`x`$.
We are going to show how formal integrals can be obtained from the method of stationary phase.
Let $`\varphi `$ be a smooth function on an open subset $`UM`$ such that (i) $`\mathrm{Re}\varphi 0`$; (ii) there is only one critical point $`x_cU`$ of the function $`\varphi `$, which is moreover a nondegenerate critical point; (iii) $`\varphi (x_c)=0`$.
Consider a classical symbol $`\rho (x,m)S^0(U\times )`$ (see for definition and notation) which has an asymptotic expansion $`\rho _{r0}(1/m^r)\rho _r(x)`$ such that $`\rho _0(x_c)0`$, and a smooth nonvanishing volume form $`dx`$ on $`U`$. Set $`\mu (m)=\rho (m,x)dx`$.
We can apply the method of stationary phase with a complex phase function (see and ) to the integral
(5.9)
$$S_m(f)=_Ue^{m\varphi }f\mu (m),$$
where $`fC_0^{\mathrm{}}(U)`$. Notice that the phase function in (5.9) is $`(1/i)\varphi `$ so that the condition $`\mathrm{Im}\left((1/i)\varphi \right)0`$ is satisfied.
Taking into account that $`dim_{}M=2n`$ and $`\varphi (x_c)=0`$ we obtain that $`S_m(f)`$ expands to an asymptotic series $`S_m(f)_{r=0}^{\mathrm{}}(1/m^{n+r})\stackrel{~}{K}_r(f)`$ as $`m+\mathrm{}`$. Here $`\stackrel{~}{K}_r,r0,`$ are distributions supported at $`x_c`$ and $`\stackrel{~}{K}_0=c_n\delta _{x_c}`$, where $`c_n`$ is a nonzero constant. Thus $`𝔽(S_m(f))=\nu ^n\stackrel{~}{K}(f)`$, where $`𝔽`$ is the ”formalizer” introduced in Section 2 and $`\stackrel{~}{K}`$ is the functional defined by the formula $`\stackrel{~}{K}=_{r0}\nu ^r\stackrel{~}{K}_r`$. Consider the normalized functional $`K(f)=\stackrel{~}{K}(f)/\stackrel{~}{K}(1)`$, so that $`K(1)=1`$. Then $`𝔽(S_m(f))=c(\nu )K(f)`$, where $`c(\nu )=\nu ^nc_n+\mathrm{}`$ is a formal constant.
###### Proposition 5.3.
For $`fC_0^{\mathrm{}}(U)S_m(f)`$ given by (5.9) expands in an asymptotic series in $`1/m`$ as $`m+\mathrm{}`$. $`𝔽\left(S_m(f)\right)=c(\nu )K(f)`$, where $`K`$ is the formal integral at the point $`x_c`$ associated to the pair $`((1/\nu )\varphi ,𝔽(\mu ))`$ and $`c(\nu )`$ is a nonzero formal constant.
###### Proof.
Conditions (a-c) of the definition of formal integral are satisfied. It remains to check condition (d). Let $`\xi `$ be a vector field on $`U`$. Denote by $`𝐋_\xi `$ the corresponding Lie derivative. We have $`0=_U𝐋_\xi (e^{m\varphi }f\mu (m))=_Ue^{m\varphi }\left(\xi f+(m\xi \varphi +\mathrm{div}_\mu \xi )f\right)\mu (m)`$. Applying $`𝔽`$ we obtain that $`0=𝔽\left(_Ue^{m\varphi }\left(\xi f+(m\xi \varphi +\mathrm{div}_\mu \xi )f\right)\mu (m)\right)=c(\nu )K\left(\xi f+\left(\xi ((1/\nu )\varphi )+\mathrm{div}_{𝔽(\mu )}\xi \right)f\right)`$, which concludes the proof. ∎
Our next goal is to get an asymptotic expansion of the Bergman kernel $`_m`$ in a neighborhood of the diagonal of $`X\times X`$ as $`m+\mathrm{}`$. An asymptotic expansion of $`_m`$ on the diagonal of $`X\times X`$ was obtained in (see (5.4)). As in , we use the integral representation of the Szegö kernel $`𝒮`$ given by the following theorem. We denote $`n=dim_{}M`$.
###### Theorem 5.4.
(L. Boutet de Monvel and J. Sjöstrand, , Theorem 1.5. and $`𝓍`$ 2.c) Let $`𝒮(\alpha ,\beta )`$ be the Szegö kernel of the boundary $`X`$ of the bounded strictly pseudoconvex domain $`D`$ in the complex manifold $`L^{}`$. There exists a classical symbol $`aS^n(X\times X\times ^+)`$ which has an asymptotic expansion
$$a(\alpha ,\beta ,t)\underset{k=0}{\overset{\mathrm{}}{}}t^{nk}a_k(\alpha ,\beta )$$
so that
(5.10)
$$𝒮(\alpha ,\beta )=_0^{\mathrm{}}e^{it\phi (\alpha ,\beta )}a(\alpha ,\beta ,t)𝑑t,$$
where the phase $`\phi (\alpha ,\beta )C^{\mathrm{}}(L^{}\times L^{})`$ is determined by the following properties:
* $`\phi (\alpha ,\alpha )=(1/i)\left(k(\alpha )1\right)`$;
* $`\overline{}_\alpha \phi `$ and $`_\beta \phi `$ vanish to infinite order along the diagonal;
* $`\phi (\alpha ,\beta )=\overline{\phi (\beta ,\alpha )}`$.
The phase function $`\phi `$ is thus almost analytic at the diagonal of $`L^{}\times \overline{L^{}}`$. It is determined up to equivalence at the diagonal.
Fix an arbitrary point $`x_0M`$. Let $`s`$ be a local holomorphic frame of $`L^{}`$ over a contractible open neighborhood $`UM`$ of the point $`x_0`$ with local holomorphic coordinates $`\{z^k\}`$. Then $`\alpha (x)=s(x)/\sqrt{k(s(x))}`$ is a smooth section of $`X`$ over $`U`$. Set $`\mathrm{\Phi }_1(x)=\mathrm{log}k\left(s(x)\right)`$, so that
(5.11)
$$\alpha (x)=e^{(1/2)\mathrm{\Phi }_1(x)}s(x).$$
It follows from the fact that $`L`$ is a quantum line bundle (i.e., that $`\omega _1`$ is the curvature form of the Hermitian holomorphic line bundle $`L`$) that $`\mathrm{\Phi }_1`$ is a potential of the form $`\omega _1`$ on $`U`$.
Let $`\stackrel{~}{\mathrm{\Phi }}_1(x,y)C^{\mathrm{}}(U\times \overline{U})`$ be an almost analytic extension of the potential $`\mathrm{\Phi }_1`$ from the diagonal of $`U\times \overline{U}`$. Denote $`D_1(x,y):=\stackrel{~}{\mathrm{\Phi }}_1(x,y)+\stackrel{~}{\mathrm{\Phi }}_1(y,x)\mathrm{\Phi }_1(x)\mathrm{\Phi }_1(y)`$. Since $`\stackrel{~}{\mathrm{\Phi }}_1(x,x)=\mathrm{\Phi }_1(x)`$, we have $`D_1(x,x)=0`$. In local coordinates
(5.12)
$$D_1(x,y)=Q_{x_0}(xy)+O(|xy|^3),$$
where
$$Q_{x_0}(z)=\frac{^2\mathrm{\Phi }_1}{z^k\overline{z}^l}(x_0)z^k\overline{z}^l$$
is a positive definite quadratic form (since $`\omega _1`$ is a Kähler form).
The following statement is an immediate consequence of (5.12).
###### Lemma 5.5.
There exists a neighborhood $`U^{}U`$ of the point $`x_0`$ such that for any two different points $`x,yU^{}`$ one has $`\mathrm{Re}D_1(x,y)<0`$.
Taking, if necessary, $`(1/2)\left(\stackrel{~}{\mathrm{\Phi }}_1(x,y)+\overline{\stackrel{~}{\mathrm{\Phi }}_1(y,x)}\right)`$ instead of $`\stackrel{~}{\mathrm{\Phi }}_1(x,y)`$ choose $`\stackrel{~}{\mathrm{\Phi }}_1`$ such that $`\stackrel{~}{\mathrm{\Phi }}_1(y,x)=\overline{\stackrel{~}{\mathrm{\Phi }}_1(x,y)}`$. Replace $`U`$ by a smaller neighborhood (retaining for it the notation $`U`$) such that $`\mathrm{Re}D_1(x,y)<0`$ for any different $`x,y`$ from this neighborhood.
For a point $`\alpha `$ in the restriction $`L^{}|_U`$ of the line bundle $`L^{}`$ to $`U`$ represented in the form $`\alpha =vs(x)`$ with $`v,xU`$ one has $`k(\alpha )=|v|^2k\left(s(x)\right)`$.
One can choose the phase function $`\phi (\alpha ,\beta )`$ in (5.10) of the form
(5.13)
$$\phi (\alpha ,\beta )=(1/i)\left(v\overline{w}e^{\stackrel{~}{\mathrm{\Phi }}_1(x,y)}1\right),$$
where $`\alpha =vs(x),\beta =ws(y)L^{}|_U`$.
Denote $`\chi (x,y):=\stackrel{~}{\mathrm{\Phi }}_1(x,y)(1/2)\mathrm{\Phi }_1(x)(1/2)\mathrm{\Phi }_1(y)`$. Notice that $`\chi (x,x)=0`$.
The following theorem is a slight generalization of Theorem 1 from .
###### Theorem 5.6.
There exists an asymptotic expansion of the Bergman kernel $`_m(\alpha (x),\alpha (y))`$ on $`U\times U`$ as $`m+\mathrm{}`$, of the form
(5.14)
$$_m(\alpha (x),\alpha (y))m^ne^{m\chi (x,y)}\underset{r0}{}(1/m^r)\stackrel{~}{b}_r(x,y)$$
such that (i) for any compact $`EU\times U`$ and $`N`$
(5.15)
$$\underset{(x,y)E}{sup}\left|_m(\alpha (x),\alpha (y))m^ne^{m\chi (x,y)}\underset{r=0}{\overset{N1}{}}(1/m^r)\stackrel{~}{b}_r(x,y)\right|=O(m^{nN});$$
(ii) $`\stackrel{~}{b}_r(x,y)`$ is an almost analytic extension of $`b_r(x)`$ from the diagonal of $`U\times U`$, where $`b_r,r0,`$ are given by (5.4); in particular, $`\stackrel{~}{b}_0(x,x)=1`$.
###### Proof.
Using integral representations (5.1) and (5.10) one gets for $`x,yU`$
(5.16)
$$_m(\alpha (x),\alpha (y))=\frac{1}{2\pi }_0^{2\pi }_0^{\mathrm{}}e^{im\theta }e^{it\phi (r_\theta \alpha (x),\alpha (y))}a(r_\theta \alpha (x),\alpha (y),t)𝑑\theta 𝑑t.$$
Changing variables $`tmt`$ in (5.16) gives
(5.17)
$$_m(\alpha (x),\alpha (y))=\frac{m}{2\pi }_0^{2\pi }_0^{\mathrm{}}e^{im(t\phi (r_\theta \alpha (x),\alpha (y))\theta )}a(r_\theta \alpha (x),\alpha (y),mt)𝑑\theta 𝑑t.$$
In order to apply the method of stationary phase to the integral in (5.17) the following preparations should be made.
Using (5.13) and (5.11) express the phase function of the integral in (5.17) as follows:
(5.18)
$$Z(t,\theta ;x,y):=t\phi (r_\theta \alpha (x),\alpha (y))\theta =(t/i)\left(e^{i\theta }e^{\chi (x,y)}1\right)\theta .$$
In order to find the critical points of the phase $`Z`$ (with respect to the variables $`(t,\theta )`$; the variables $`(x,y)`$ are parameters) consider first the equation
(5.19)
$$_tZ(t,\theta ;x,y)=(1/i)\left(e^{i\theta }e^{\chi (x,y)}1\right)=0.$$
It follows from $`\stackrel{~}{\mathrm{\Phi }}_1(y,x)=\overline{\stackrel{~}{\mathrm{\Phi }}_1(x,y)}`$ that $`\mathrm{Re}\chi (x,y)=(1/2)D_1(x,y)`$. Since $`D_1(x,y)<0`$ for $`xy`$ one has $`|e^{\chi (x,y)}|=e^{\mathrm{Re}\chi (x,y)}<1`$ for $`xy`$ whence it follows that (5.19) holds only if $`x=y`$ and thus $`Z`$ has critical points only if $`x=y`$. Since $`\chi (x,x)=0`$ one gets that $`_tZ(t,\theta ;x,x)=(1/i)(e^{i\theta }1)`$ and $`_\theta Z(t,\theta ;x,x)=te^{i\theta }1`$. As in the proof of Theorem 1 from , one shows that for each $`xU`$ the only critical point of the phase function $`Z(t,\theta ;x,x)`$ is $`(t=1,\theta =0)`$. It does not depend on $`x`$ and, moreover, is nondegenerate.
One has $`\mathrm{Im}Z(t,\theta ;x,y)=\mathrm{Im}\left((t/i)(e^{i\theta }e^{\chi (x,y)}1)\theta \right)=t\left(1\mathrm{Re}(e^{i\theta }e^{\chi (x,y)})\right)0`$ since $`|e^{\chi (x,y)}|1`$.
Finally, a simple calculation shows that the germs of the functions $`Z(t,\theta ;x,y)`$ and $`(1/i)\chi (x,y)`$ at the point $`(t=1,\theta =0,x=x_0,y=x_0)`$ are equal modulo the ideal generated by $`_tZ`$ and $`_\theta Z`$.
Applying now the method of stationary phase to the integral in (5.17) one obtains the expansion (5.14) satisfying (5.15).
It follows from (5.4) and (4.3) that $`\stackrel{~}{b}_r(x,x)=b_r(x)`$ and $`\stackrel{~}{b}_0(x,x)=b_0(x)=1`$. It remains to show that all $`\stackrel{~}{b}_r,r0,`$ are almost analytic along the diagonal of $`U\times \overline{U}`$. One has
$$_m(\alpha (x),\alpha (y))=e^{(m/2)(\mathrm{\Phi }_1(x)+\mathrm{\Phi }_1(y))}_m(s(x),s(y)).$$
The function $`_m(s(x),s(y))`$ is holomorphic on $`U\times \overline{U}`$. Let $`\xi `$ and $`\eta `$ be arbitrary holomorphic and antiholomorphic vector fields on $`U`$, respectively. Then $`\xi _y_m(s(x),s(y))=0`$ and $`\eta _x_m(s(x),s(y))=0`$ (the subscripts $`x,y`$ show in which variable the vector field acts). Thus
$$\left(\eta _x+\frac{m}{2}\eta _x\mathrm{\Phi }_1(x)\right)_m(\alpha (x),\alpha (y))=e^{(m/2)\mathrm{\Phi }_1(x)}\eta _xe^{(m/2)\mathrm{\Phi }_1(x)}_m(\alpha (x),\alpha (y))=0.$$
Analogously, $`\left(\xi _y+(m/2)\xi _y\mathrm{\Phi }_1(x)\right)_m(\alpha (x),\alpha (y))=0.`$ Let $`A_N`$ be a product of $`N`$ derivations on $`U\times U`$. Then, using integral representation (5.17), expand $`0=A_N\left(\eta _x+(m/2)\eta _x\mathrm{\Phi }_1(x)\right)_m(\alpha (x),\alpha (y))`$ to the asymptotic series
(5.20)
$$A_N\left(\eta _x+\frac{m}{2}\eta _x\mathrm{\Phi }_1(x)\right)\left(m^ne^{m\chi (x,y)}\underset{r0}{}(1/m^r)\stackrel{~}{b}_r(x,y)\right)=e^{m\chi (x,y)}\underset{rr_0}{}(1/m^r)c_r(x,y)$$
for some $`c_rC^{\mathrm{}}(U\times U)`$ and $`r_0`$, and with the norm estimate of the partial sums in the r.h.s. term in (5.20) analogous to (5.15). Since $`\chi (x,x)=0`$ one gets that all $`c_r(x,x)=0`$. From this fact one can prove by induction over $`N`$ that $`\eta _x\stackrel{~}{b}_r`$ vanishes to infinite order at the diagonal of $`U\times U`$. Similarly, $`\xi _y\stackrel{~}{b}_r`$ vanishes to infinite order at the diagonal. Thus $`\stackrel{~}{b}_r`$ is almost analytic along the diagonal. ∎
Choose a symbol $`b(x,y,m)S^0((U\times U)\times )`$ such that it has the asymptotic expansion $`b_{r=0}^{\mathrm{}}(1/m^r)\stackrel{~}{b}_r`$. Then $`_m(\alpha (x),\alpha (y))`$ is asymptotically equivalent to $`m^ne^{m\chi (x,y)}b(x,y,m)`$ on $`U\times U`$. One has $`\chi (x,y)+\chi (y,x)=\stackrel{~}{\mathrm{\Phi }}_1(x,y)+\stackrel{~}{\mathrm{\Phi }}_1(y,x)\mathrm{\Phi }_1(x)\mathrm{\Phi }_1(y)=D_1(x,y)`$ and $`\chi (x,y)+\chi (y,z)+\chi (z,x)=\stackrel{~}{\mathrm{\Phi }}_1(x,y)+\stackrel{~}{\mathrm{\Phi }}_1(y,z)+\stackrel{~}{\mathrm{\Phi }}_1(z,x)\mathrm{\Phi }_1(x)\mathrm{\Phi }_1(y)\mathrm{\Phi }_1(z)=T_1(x,y,z)`$ (the last equality is the definition of $`T_1`$). Thus the functions
$$\begin{array}{c}v_m(x,y)=_m(\alpha (x),\alpha (y))_m(\alpha (y),\alpha (x))\text{ and }\hfill \\ \hfill w_m(x,y,z)=_m(\alpha (x),\alpha (y))_m(\alpha (y),\alpha (z))_m(\alpha (z),\alpha (x))\end{array}$$
are asymptotically equivalent to
$$m^{2n}e^{mD_1(x,y)}b(x,y,m)b(y,x,m)\text{ and }m^{3n}e^{mT_1(x,y,z)}b(x,y,m)b(y,z,m)b(z,x,m)$$
respectively. It is easy to show that for the functions $`\varphi _1^x(y)=D_1(x,y)`$ and $`\psi _1^x(y,z)=T_1(x,y,z)`$ the points $`y=x`$ and $`(y,z)=(x,x)`$ respectively are nondegenerate critical ones.
Since $`\stackrel{~}{b}_0(x,x)=1`$ one can take a smaller contractible neighborhood $`VU`$ of $`x_0`$ such that $`\stackrel{~}{b}_0(x,y)`$ does not vanish on the closure of $`V\times V`$. One can choose $`V`$ such that for any $`xV`$ the only critical points of the functions $`\varphi _1^x(y)`$ on $`V`$ and $`\psi _1^x(y,z)`$ on $`V\times V`$ are $`y=x`$ and $`(y,z)=(x,x)`$ respectively.
The identity $`T_1(x,y,z)=(1/2)(D_1(x,y)+D_1(y,z)+D_1(z,x))`$ implies that
$`\mathrm{Re}T_1(x,y,z)0`$ for $`x,y,zV`$.
The symbol $`b(x,y,m)`$ does not vanish on $`V\times V`$ for sufficiently big values of $`m`$. It follows from (5.4) that $`1/u_m(x)`$ and $`(m^nb(x,x,m))^1`$ are asymptotically equivalent for $`xV`$. Denote
(5.21)
$$\mu _x(m)=\frac{b(x,y,m)b(y,x,m)}{b(x,x,m)}\mathrm{\Omega }(y),\stackrel{~}{\mu }_x(m)=\frac{b(x,y,m)b(y,z,m)b(z,x,m)}{b(x,x,m)}\mathrm{\Omega }(y)\mathrm{\Omega }(z).$$
Taking into account (4.6) we get for $`f,gC_0^{\mathrm{}}(V)`$ and $`xV`$ the following asymptotic equivalences,
(5.22)
$$\left(I^{(m)}f\right)(x)m^n_Ve^{m\varphi _1^x}f\mu _x(m)\text{ and }Q^{(m)}(f,g)(x)m^{2n}_{V\times V}e^{m\psi _1^x}(fg)\stackrel{~}{\mu }_x(m).$$
(In (5.22) $`\left(fg\right)(y,z)=f(y)g(z)`$.)
Applying Proposition 5.3 to the first integral in (5.22) we obtain that $`𝔽\left(\left(I^{(m)}f\right)(x)\right)=c(\nu ,x)L_x^I(f)`$, where the functional $`L_x^I`$ on $`(V)`$ is the formal integral at the point $`x`$ associated to the pair $`((1/\nu )\varphi _1^x,𝔽(\mu _x))`$ and $`c(\nu ,x)`$ is a formal function. It is easy to show that $`c(\nu ,x)`$ is smooth.
Similarly we obtain from (5.22) that $`𝔽\left(Q^{(m)}(f,g)(x)\right)=d(\nu ,x)L_x^Q(fg)`$ where the functional $`L_x^Q`$ on $`(V\times V)`$ is the formal integral at the point $`(x,x)`$ associated to the pair $`((1/\nu )\psi _1^x,𝔽(\stackrel{~}{\mu }_x))`$ and $`d(\nu ,x)`$ is a smooth formal function.
Since the unit constant $`1`$ is a contravariant symbol of the unit operator $`\mathrm{𝟏}`$, $`T_1^{(m)}=\mathrm{𝟏}`$, and $`\sigma (\mathrm{𝟏})=1`$, we have $`I^{(m)}1=1,Q^{(m)}(1,1)=1,`$ and thus $`𝔽(I^{(m)}1)=1`$ and $`𝔽\left(Q^{(m)}(1,1)\right)=1`$. Taking the functions $`f,g`$ in (5.22) to be equal to $`1`$ in a neighborhood of $`x`$ and applying Proposition 5.1 and Proposition 5.2 we get that $`c(\nu ,x)=1`$ and $`d(\nu ,x)=1`$.
Since $`b_0(x,y)`$ does not vanish on $`V\times V`$ we can find a formal function $`\stackrel{~}{s}(x,y)`$ on $`V\times V`$ such that $`𝔽(b(x,y,m))=e^{\stackrel{~}{s}(x,y)}`$. Set $`s(x)=\stackrel{~}{s}(x,x)`$. In these notations
(5.23)
$$\begin{array}{c}𝔽(\mu _x)=\mathrm{exp}(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(y,x)s(x))\mathrm{\Omega }(y)\text{ and }\hfill \\ \hfill 𝔽(\stackrel{~}{\mu }_x)=\mathrm{exp}(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(y,z)+\stackrel{~}{s}(z,x)s(x))\mathrm{\Omega }(y)\mathrm{\Omega }(z).\end{array}$$
It follows from Theorem 5.6 that $`\stackrel{~}{s}`$ is an almost analytic extension of the function $`s`$ from the diagonal of $`V\times V`$. According to (5.4), $`𝔽(u_m)=(1/\nu ^n)e^s`$.
Denote $`\stackrel{~}{\mathrm{\Phi }}=(1/\nu )\stackrel{~}{\mathrm{\Phi }}_1+\stackrel{~}{s},\mathrm{\Phi }=(1/\nu )\mathrm{\Phi }_1+s,D(x,y)=\stackrel{~}{\mathrm{\Phi }}(x,y)+\stackrel{~}{\mathrm{\Phi }}(y,x)\mathrm{\Phi }(x)\mathrm{\Phi }(y)=(1/\nu )D_1(x,y)+(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(y,x)s(x)s(y)),T(x,y,z)=\stackrel{~}{\mathrm{\Phi }}(x,y)+\stackrel{~}{\mathrm{\Phi }}(y,z)+\stackrel{~}{\mathrm{\Phi }}(z,x)\mathrm{\Phi }(x)\mathrm{\Phi }(y)\mathrm{\Phi }(z)`$. The pair $`((1/\nu )\varphi _1^x,𝔽(\mu _x))=((1/\nu )\varphi _1^x,\mathrm{exp}(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(y,x)s(x))\mathrm{\Omega }(y))`$ is then equivalent to the pair $`(\varphi ^x,e^s\mathrm{\Omega })`$, where $`\varphi ^x(y)=D(x,y)`$. Similarly, the pair $`((1/\nu )\psi _1^x,𝔽(\stackrel{~}{\mu }_x))`$ is equivalent to the pair $`(\psi ^x,e^s\mathrm{\Omega }e^s\mathrm{\Omega })`$, where $`\psi ^x(y,z)=T(x,y,z)`$.
Thus we arrive at the following proposition.
###### Proposition 5.7.
For $`f,gC_0^{\mathrm{}}(V),xV,\left(I^{(m)}f\right)(x)`$ and $`Q^{(m)}(f,g)(x)`$ expand in asymptotic series in $`1/m`$ as $`m+\mathrm{}`$. $`𝔽\left(\left(I^{(m)}f\right)(x)\right)=L_x^I(f)`$ and $`𝔽\left(Q^{(m)}(f,g)(x)\right)=L_x^Q(fg)`$, where the functional $`L_x^I`$ on $`(V)`$ is the formal integral at the point $`x`$ associated to the pair $`(\varphi ^x,e^s\mathrm{\Omega })`$ and the functional $`L_x^Q`$ on $`(V\times V)`$ is the formal integral at the point $`(x,x)`$ associated to the pair $`(\psi ^x,e^s\mathrm{\Omega }e^s\mathrm{\Omega })`$.
Now let $``$ denote the star-product with separation of variables on $`(V,\omega _1)`$ corresponding to the formal deformation $`\omega =i\overline{}\mathrm{\Phi }`$ of the form $`(1/\nu )\omega _1`$, so that $`\mathrm{\Phi }`$ is a formal potential of $`\omega `$. Let $`I`$ be the corresponding formal Berezin transform, $`\stackrel{~}{\omega }`$ the formal form parametrizing the dual star-product $`\stackrel{~}{}`$ and $`\mathrm{\Psi }`$ the solution of (2.3) so that $`\mu _{tr}=e^{\mathrm{\Phi }+\mathrm{\Psi }}dzd\overline{z}`$ is a formal trace density for the star-product $``$.
Choose a classical symbol $`\rho (x,m)S^0(V\times )`$ which has an asymptotic expansion $`\rho _{r0}(1/m^r)\rho _r`$ such that
(5.24)
$$𝔽(\rho )e^s\mathrm{\Omega }=\mu _{tr}.$$
Clearly, (5.24) determines $`𝔽(\rho )`$ uniquely.
For $`fC_0^{\mathrm{}}(V)`$ and $`xV`$ consider the following integral
(5.25)
$$(P_mf)(x)=m^n_Ve^{m\varphi _1^x}f\rho \mu _x,$$
where $`\varphi _1^x(y)=D_1(x,y)`$ and $`\mu _x`$ is given by (5.21).
###### Proposition 5.8.
For $`fC_0^{\mathrm{}}(V)`$ and $`xV`$ $`(P_mf)(x)`$ has an asymptotic expansion in $`1/m`$ as $`m+\mathrm{}`$. $`𝔽\left((P_mf)(x)\right)=c(\nu )(If)(x)`$, where $`c(\nu )`$ is a nonzero formal constant.
###### Proof.
It was already shown that the phase function $`(1/i)\varphi _1^x`$ of integral (5.25) satisfies the conditions required in the method of stationary phase. Thus Proposition 5.3 can be applied to (5.25). We get that $`𝔽\left((P_mf)(x)\right)=c(\nu ,x)K_x(f)`$, where $`K_x`$ is a formal integral at the point $`x`$ associated to the pair $`((1/\nu )\varphi _1^x,𝔽(\rho \mu _x))`$ and $`c(\nu ,x)`$ is a nonvanishing formal function on $`V`$. It follows from (5.23) and (5.24) that $`𝔽(\rho \mu _x)=𝔽(\rho )𝔽(\mu _x)=𝔽(\rho )\mathrm{exp}(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(x,y)s(x))\mathrm{\Omega }(y)=\mathrm{exp}(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(x,y)s(x)s(y))\mu _{tr}=\mathrm{exp}(D(x,y)(1/\nu )D_1(x,y))\mu _{tr}=\mathrm{exp}(\varphi ^x(1/\nu )\varphi _1^x)\mu _{tr}`$, where $`\varphi ^x(y)=D(x,y)`$. The pair $`((1/\nu )\varphi _1^x,𝔽(\rho \mu _x))`$ is thus equivalent to the pair $`(\varphi ^x,\mu _{tr})`$. Applying Theorem 3.2 we get that
(5.26)
$$𝔽\left((P_mf)(x)\right)=c(\nu ,x)\left(If\right)(x).$$
It remains to show that $`c(\nu ,x)`$ is actually a formal constant. Let $`x_1`$ be an arbitrary point of $`V`$. Choose a function $`ϵC_0^{\mathrm{}}(V)`$ such that $`ϵ=1`$ in a neighborhood $`WV`$ of $`x_1`$. Let $`\xi `$ be a vector field on $`V`$. Then, using (5.23), we obtain
(5.27)
$$\begin{array}{c}\frac{1}{\nu }\xi _x\varphi _1^x(y)+𝔽\left(\frac{\xi _x\mu _x}{\mu _x}(y)\right)=\hfill \\ \hfill \frac{1}{\nu }\xi _xD_1(x,y)+\xi _x(\stackrel{~}{s}(x,y)+\stackrel{~}{s}(x,y)s(x))=\xi _xD(x,y)=\xi _x\varphi ^x.\end{array}$$
On the one hand, taking into account (5.27) we get for $`xW`$ that
(5.28)
$$\begin{array}{c}𝔽\left((\xi P_mϵ)(x)\right)=𝔽\left(m^n\xi _Ve^{m\varphi _1^x}ϵ\rho \mu _x\right)=𝔽\left(m^n_Ve^{m\varphi _1^x}\left(m\xi _x\varphi _x+\frac{\xi _x\mu _x}{\mu _x}\right)ϵ\rho \mu _x\right)=\hfill \\ \hfill c(\nu ,x)I\left(\frac{1}{\nu }\xi _x\varphi _1^x(y)+𝔽\left(\frac{\xi _x\mu _x}{\mu _x}(y)\right)\right)=c(\nu ,x)I(\xi _x\varphi ^x)=0.\end{array}$$
The last equality in (5.28) follows from Lemma 3.3. On the other hand, for $`xW`$ we have from (5.26) that $`𝔽\left((P_mϵ)(x)\right)=c(\nu ,x)`$, from whence $`𝔽\left((\xi P_mϵ)(x)\right)=\xi 𝔽\left((P_mϵ)(x)\right)=\xi c(\nu ,x)`$. Thus we get from (5.28) that $`\xi c(\nu ,x)=0`$ on $`W`$ for an arbitrary vector field $`\xi `$, from which the Proposition follows. ∎
It follows from (5.22) and (5.25) that for $`fC_0^{\mathrm{}}(V)`$ $`\left(I^{(m)}(f\rho )\right)(x)`$ is asymptotically equivalent to $`(P_mf)(x)`$. Passing to formal asymptotic series we get from Proposition 5.7 and Proposition 5.8 that $`c(\nu )(If)(x)=𝔽\left((P_mf)(x)\right)=𝔽\left(\left(I^{(m)}(f\rho )\right)(x)\right)=L_x^I(f𝔽(\rho ))`$, where $`L_x^I`$ is the formal integral at the point $`x`$ associated to the pair $`(\varphi ^x,e^s\mathrm{\Omega })`$. Thus
(5.29)
$$c(\nu )(If)(x)=L_x^I(f𝔽(\rho )).$$
The formal function $`𝔽(\rho )`$ is invertible (see (5.24)). Setting $`f=1/𝔽(\rho )`$ in (5.29) we get $`c(\nu )\left(I(1/𝔽(\rho ))\right)(x)=L_x^I(1)=1`$ for all $`xV`$. Since the formal Berezin transform is invertible and $`I(1)=1`$, we finally obtain that
(5.30)
$$𝔽(\rho )=c(\nu ).$$
Now (5.24) can be rewritten as follows,
(5.31)
$$c(\nu )e^s\mathrm{\Omega }=d\mu _{tr}=e^{\mathrm{\Phi }+\mathrm{\Psi }}dzd\overline{z}.$$
In local holomorphic coordinates the symplectic volume $`\mathrm{\Omega }`$ can be expressed as follows, $`\mathrm{\Omega }=e^\theta dzd\overline{z}`$. The closed (1,1)-form $`\omega _{can}=i\overline{}\theta `$ does not depend on the choice of local holomorphic coordinates and is defined globally on $`M`$. The form $`\omega _{can}`$ is the curvature form of the canonical connection of the canonical holomorphic line bundle on $`M`$ equipped with the Hermitian fibre metric determined by the volume form $`\mathrm{\Omega }`$. Its de Rham class $`\epsilon =[\omega _{can}]`$ is the first Chern class of the canonical holomorphic line bundle on $`M`$ and thus depends only on the complex structure on $`M`$. The class $`\epsilon `$ is called the canonical class of the complex manifold $`M`$.
One can see from (5.31) that $`c(\nu )=c_0+\nu c_1+\mathrm{}`$, where $`c_00`$. Thus there exists a formal constant $`d(\nu )`$ such that $`e^{d(\nu )}=c(\nu )`$ and $`d(\nu )+s+\theta =\mathrm{\Phi }+\mathrm{\Psi }`$. Therefore the formal potential $`\mathrm{\Psi }`$ of the form $`\stackrel{~}{\omega }`$ is expressed explicitly, $`\mathrm{\Psi }=d(\nu )(1/\nu )\mathrm{\Phi }_1+\theta `$, from whence it follows that
(5.32)
$$\stackrel{~}{\omega }=(1/\nu )\omega _1+\omega _{can}.$$
Formula (5.32) defines $`\stackrel{~}{\omega }`$ globally on $`M`$. Thus the corresponding star-product $`\stackrel{~}{}`$ and therefore its dual star-product $`\omega `$ are also globally defined.
Theorem 3.2, Theorem 3.4, Proposition 5.1, Proposition 5.2 Proposition 5.7, formulas (5.29), (5.30) and (5.31) imply the following theorem, which is the central technical result of the paper.
###### Theorem 5.9.
For any $`f,gC^{\mathrm{}}(M)`$ and $`xM(I^{(m)}f)(x)`$ and $`Q^{(m)}(f,g)(x)`$ expand to asymptotic series in $`1/m`$ as $`m+\mathrm{}`$. $`𝔽\left(\left(I^{(m)}f\right)(x)\right)=\left(If\right)(x)`$ and $`𝔽\left(Q^{(m)}(f,g)(x)\right)=Q(f,g)(x)`$, where $`I`$ and $`Q`$ are the formal Berezin transform and the formal twisted product corresponding to the star-product with separation of variables $``$ on $`(M,\omega _1)`$ whose dual star-product $`\stackrel{~}{}`$ on $`(M,\omega _1)`$ is parametrized by the formal form $`\stackrel{~}{\omega }=(1/\nu )\omega _1+\omega _{can}.`$
###### Remark.
As shown in we have the following chain of inequalities
(5.33)
$$|I^{(m)}(f)|_{\mathrm{}}=|\sigma (T_f^{(m)})|_{\mathrm{}}T_f^{(m)}|f|_{\mathrm{}}.$$
Here $`||..||`$ denotes the operator norm with respect to the norm of the sections of $`L^m`$ and $`|..|_{\mathrm{}}`$ the sup-norm on $`C^{\mathrm{}}(M)`$. Choose as $`x_eM`$ a point with $`|f(x_e)|=|f|_{\mathrm{}}`$. From Theorem 5.9 and the fact that the formal Berezin transform has as leading term the identity it follows that $`|(I^{(m)}f)(x_e)f(x_e)|A/m`$ with a suitable constant $`A`$. This implies $`\left||f(x_e)||(I^{(m)}f)(x_e)|\right|A/m`$ and hence
(5.34)
$$|f|_{\mathrm{}}\frac{A}{m}=|f(x_e)|\frac{A}{m}|(I^{(m)}f)(x_e)||(I^{(m)}f)|_{\mathrm{}}.$$
Putting (5.33) and (5.34) together we obtain
(5.35)
$$|f|_{\mathrm{}}\frac{A}{m}T_f^{(m)}|f|_{\mathrm{}}.$$
This provides another proof of , Theorem 4.1.
## 6. The identification of the Berezin-Toeplitz star-product
In this section $``$ will denote the star-product with separation of variables on $`(M,\omega _1)`$ whose dual $`\stackrel{~}{}`$ is the star-product with separation of variables on $`(M,\omega _1)`$ parametrized by the formal form $`\stackrel{~}{\omega }=(1/\nu )\omega _1+\omega _{can}`$.
Let $`I=1+\nu I_1+\nu ^2I_2+\mathrm{}`$ and $`Q=Q_0+\nu Q_1+\mathrm{}`$ denote the formal Berezin transform and the formal twisted product corresponding to $``$. Theorem 5.9 asserts that for given $`f,gC^{\mathrm{}}(M),r𝐍,xM`$ there exist constants $`A,B`$ such that for sufficiently big values of $`m`$ the following inequalities hold:
(6.1)
$$\left|\left(I^{(m)}f\right)(x)\underset{i=0}{\overset{r1}{}}\frac{1}{m^i}I_i(f)(x)\right|\frac{A}{m^r},$$
(6.2)
$$\left|Q^{(m)}(f,g)(x)\underset{i=0}{\overset{r1}{}}\frac{1}{m^i}Q_i(f,g)(x)\right|\frac{B}{m^r}.$$
It was proved in , that Berezin-Toeplitz quantization on a compact Kähler manifold $`M`$ gives rise to a star-product on $`M`$. This star-product $`^{BT}`$ is given by a sequence of bilinear operators $`\{C_k\},k0,`$ on $`C^{\mathrm{}}(M)`$ satisfying the following conditions. For $`f,gC^{\mathrm{}}(M)`$ and any $`r𝐍`$ there exists a constant $`C`$ such that
(6.3)
$$T_f^{(m)}T_g^{(m)}T_{f_{[r]}g}^{(m)}C/m^r,$$
where $`f_{[r]}g=_{k=0}^{r1}(1/m^k)C_k(f,g)`$. The conditions (6.3) determine the star-product $`^{BT}`$ uniquely. We call $`^{BT}`$ the Berezin-Toeplitz star-product.
Recall that for $`f,gC^{\mathrm{}}(M)\sigma (T_f^{(m)})=I^{(m)}(f),\sigma (T_f^{(m)}T_g^{(m)})=Q^{(m)}(f,g)`$.
Passing from operators to their covariant symbols in (6.3) and using the inequality $`|\sigma (A)|A`$ we get that
(6.4)
$$\left|Q^{(m)}(f,g)(x)I^{(m)}(f_{[r]}g)(x)\right|C/m^r.$$
It follows from (6.1) that
(6.5)
$$\left|\frac{1}{m^k}I^{(m)}\left(C_k(f,g)\right)(x)\underset{i=0}{\overset{rk1}{}}\frac{1}{m^{i+k}}I_i\left(C_k(f,g)\right)(x)\right|\frac{A_k}{m^r}.$$
Summing up inequalities (6.2) and (6.5) for $`k=0,1,\mathrm{},r1`$, we obtain that
(6.6)
$$\begin{array}{c}|(Q^{(m)}(f,g)(x)I^{(m)}\left(f_{[r]}g\right)(x))\hfill \\ \hfill \underset{i=0}{\overset{r1}{}}\frac{1}{m^i}(Q_i(f,g)(x)\underset{j+k=i}{}I_j\left(C_k(f,g)\right)(x))|\frac{D}{m^r}.\end{array}$$
for some constant $`D`$. It follows from (6.4) and (6.6) that
$$\left|\underset{i=0}{\overset{r1}{}}\frac{1}{m^i}\left(Q_i(f,g)(x)\underset{j+k=i}{}I_j\left(C_k(f,g)\right)(x)\right)\right|\frac{E}{m^r},$$
for some constant $`E`$, which infers that for $`i=0,1,\mathrm{}`$
(6.7)
$$Q_i(f,g)=\underset{j+k=i}{}I_j(C_k(f,g)).$$
Equalities (6.7) mean that $`Q(f,g)=I(f^{BT}g)`$. Since $`I`$ is invertible we immediately obtain that the star-products $`^{}`$ and $`^{BT}`$ coincide. Thus the Berezin-Toeplitz deformation quantization is completely identified as the deformation quantization with separation of variables on $`(\overline{M},\omega _1)`$ whose star-product $`^{BT}`$ is opposite to $`\stackrel{~}{}`$.
Using (2.2) we can calculate the characteristic class $`cl(^{BT})`$ of the Berezin-Toeplitz star-product $`^{BT}`$.
It follows from (2.2) and (5.32) that the characteristic class of the star-product $`\stackrel{~}{}`$ equals to $`cl(\stackrel{~}{})=(1/i)\left([(1/\nu )\omega _1]+\epsilon /2\right)`$. It is easy to show that the characteristic class of the opposite star-product $`^{}`$ is equal to $`cl(\stackrel{~}{})`$. Since $`^{BT}=^{}`$, we finally get that the characteristic class of the Berezin-Toeplitz deformation quantization is given by the formula $`cl(^{BT})=(1/i)\left([(1/\nu )\omega _1]\epsilon /2\right)`$.
The characteristic class of the Berezin-Toeplitz deformation quantization was first calculated by Eli Hawkins in by K-theoretic methods.
As a concluding remark we would like to draw the readers attention to the fact that the classifying form $`\omega `$ of the star-product $``$ is the formal object corresponding to the asymptotic expansion as $`m+\mathrm{}`$ of the pullback $`\omega ^{(m)}`$ of the Fubini-Study form on the projective space $`(H_m^{})`$ via Kodaira embedding of $`M`$ into $`(H_m^{})`$. Here $`H_m^{}`$ denotes the Hilbert space dual to $`H_m=\mathrm{\Gamma }_{hol}(L^m)`$ (see Section 4). It was proved by Zelditch that $`\omega ^{(m)}`$ admits a complete asymptotic expansion in $`1/m`$ as $`m+\mathrm{}`$. As an easy consequence of the results obtained in this article one can show that $`𝔽(\omega ^{(m)})=\omega `$.
|
warning/0006/nucl-th0006064.html
|
ar5iv
|
text
|
# A six-body calculation of the alpha-deuteron radiative capture cross section
## I Introduction
Radiative capture of deuterons on alpha particles is the only process by which <sup>6</sup>Li is produced in standard primordial nucleosynthesis models . Because the low-energy cross section for this process is so small ($`<10^2`$ nb), it has long been held that no measurable amount of <sup>6</sup>Li can be made in big bang nucleosynthesis (BBN) without recourse to exotic physics (baryon-inhomogeneous scenarios, hadronically-evaporating black holes, etc.). However, there has been new interest in <sup>6</sup>Li as a cosmological probe in recent years, for two reasons. First, <sup>6</sup>Li is more sensitive to destruction in stars than is <sup>7</sup>Li. There has consequently been an attempt to set limits on <sup>7</sup>Li depletion in halo stars by determining how much <sup>6</sup>Li they have destroyed over their lifetimes. Second, the sensitivity of searches for <sup>6</sup>Li has been increasing as a result, so that there are now two claimed detections in metal-poor halo stars, and two more detections in disk stars . The explanation of these data in terms of chemical-evolution models remains an open question . They are presently at a level exceeding even optimistic estimates of how much <sup>6</sup>Li could have been made in standard BBN , and the observed <sup>6</sup>Li is believed to have been created in interactions between cosmic rays and the interstellar medium. However, it remains interesting from the point of view of understanding these observations to remove as many remaining uncertainties as possible from the standard scenario. The main such uncertainty arising in the standard BBN calculation is the cross section for $`d(\alpha ,\gamma )^6\mathrm{Li}`$.
At the same time, electroweak processes in light nuclei are interesting from the point of view of few-body nuclear physics. The advent of realistic two- and three-nucleon potentials has enabled calculation of nuclear wave functions and energy observables of systems with up to $`A=8`$ . Calculation of electroweak observables allows both comparison of these wave functions with experiment (e.g., electron scattering form factors ), and estimates of effects not observable in the laboratory (e.g., weak capture cross sections ). The radiative capture of deuterons on alpha particles provides both kinds of opportunities: there are data on the direct process at 700 keV (all energies are in center of mass) and above , and there are indirect data from Coulomb breakup experiments corresponding to 70–400 keV . There is also a need for extrapolation to lower energies for application to nucleosynthesis.
The $`\alpha d`$ capture problem is made even more interesting by the fact that $`S`$\- and $`P`$-wave captures are strongly inhibited by quasi-orthogonality between the initial and final states and by an isospin selection rule, respectively. As a result, the dominant process in all experiments performed to date has been electric quadrupole ($`E2`$) capture from $`D`$-wave scattering states. The small remaining $`E1`$ contribution from $`P`$-wave initial states has been observed at about 2 MeV, but its magnitude has not been successfully explained by theoretical treatments; it is generally expected to contribute half of the cross section at 100 keV. The $`S`$-wave capture induced by $`M1`$ has been neglected in most calculations because of the quasi-orthogonality mentioned above, which makes the associated matrix element identically zero in two-body treatments of the process. The energy dependences of the various capture mechanisms ($`E2`$, $`E1`$, $`M1`$) are such that even $`E1`$ and $`M1`$ captures with small amplitudes may become important at low ($`<200`$ keV) energies. Low-energy behavior is particularly important for standard BBN: the primordial <sup>6</sup>Li yield is only sensitive to the capture cross section between 20 and 200 keV, with the strongest sensitivity at 60 keV. This is demonstrated in Fig. 1, where we show the fractional change in $`{}_{}{}^{6}\mathrm{Li}`$ yields resulting from changing the cross section over narrow bins in energy. Since it is the energy integral of this quantity which determines the sensitivity, and it is shown on a logarithmic axis, the sensitivity function $`g_6`$ is defined to include a factor of energy so that relative areas may be accurately gauged. (A more extensive discussion of these functions appears in Ref. .) The $`{}_{}{}^{6}\mathrm{Li}`$ yield is directly proportional to the $`d(\alpha ,\gamma )^6\mathrm{Li}`$ cross section at the energies indicated by this function, so that an increase in the cross section by a given factor over the whole energy range produces an increase of the $`{}_{}{}^{6}\mathrm{Li}`$ yield by the same factor.
We have carried out a calculation of the alpha-deuteron radiative capture cross section, treating it as a six-body problem. The remainder of this paper describes this calculation, and it is organized as follows: In Section II, we describe wave functions used to compute the $`d(\alpha ,\gamma )^6\mathrm{Li}`$ cross section. In Section III, we describe the electromagnetic current operators and the methods used to calculate their matrix elements. In Section IV, we describe our results for the cross section, thermal reaction rates and nucleosynthesis yields, and in Section V, we summarize our results and discuss their implications.
## II wave functions
### A $`{}_{}{}^{2}\mathrm{H}`$, $`{}_{}{}^{4}\mathrm{He}`$, and $`{}_{}{}^{6}\mathrm{Li}`$ ground states
The wave functions $`|\psi _d^{m_d}`$, $`|\psi _\alpha `$, and $`|\psi _{\mathrm{Li}}^{m_6}`$ used in our calculation are the ground states derived from the Argonne $`v_{18}`$ two-nucleon and Urbana IX three-nucleon potentials, henceforth denoted as the AV18/UIX model. The deuteron wave function is a direct numerical solution, while the variational Monte Carlo technique described in Refs. is used to generate the $`{}_{}{}^{4}\mathrm{He}`$ and $`{}_{}{}^{6}\mathrm{Li}`$ wave functions. Here $`m_d`$ and $`m_6`$ denote spin orientation.
The variational trial functions for light nuclei are constructed from correlation operators acting on a Jastrow wave function:
$$|\mathrm{\Psi }_T=\left[1+\underset{i<j<k}{}\stackrel{~}{U}_{ijk}^{TNI}\right]\left[𝒮\underset{i<j}{}(1+U_{ij})\right]|\mathrm{\Psi }_J,$$
(1)
where $`U_{ij}`$ and $`\stackrel{~}{U}_{ijk}^{TNI}`$ are two- and three-body correlation operators that include significant spin and isospin dependence and $`𝒮`$ is a symmetrization operator, needed because the $`U_{ij}`$ do not commute. For $`{}_{}{}^{4}\mathrm{He}`$, the Jastrow part takes a relatively simple form:
$$|\mathrm{\Psi }_J=\underset{i<j<k4}{}f_{ijk}\underset{i<j4}{}f(r_{ij})|\mathrm{\Phi }_\alpha (0000)_{1234},$$
(2)
where $`f(r_{ij})`$ and $`f_{ijk}`$ are pair and triplet functions of relative position only, and $`\mathrm{\Phi }_\alpha (0000)`$ is a determinant in the spin-isospin space of the four particles.
For $`{}_{}{}^{6}\mathrm{Li}`$ the Jastrow part has a considerably more complicated structure due to the need to place the fifth and sixth particles in the p-shell:
$`|\mathrm{\Psi }_J`$ $`=`$ $`𝒜\{{\displaystyle \underset{i<j<k4}{}}f_{ijk}^{sss}{\displaystyle \underset{n4}{}}f_{n56}^{spp}{\displaystyle \underset{i<j4}{}}f_{ss}(r_{ij}){\displaystyle \underset{k4}{}}f_{sp}(r_{k5})f_{sp}(r_{k6})f_{pp}(r_{56})`$ (4)
$`{\displaystyle \underset{LS}{}}(\beta _{LS}|\mathrm{\Phi }_6(LSJMTT_3)_{1234:56})\}.`$
The $`𝒜`$ is an antisymmetrization operator over all partitions of the six particles into groups of four plus two. The central pair and triplet correlations $`f_{xy}(r_{ij})`$ and $`f_{ijk}^{xyz}`$ have been generalized, with the $`xyz`$ denoting whether the particles are in the s- or p-shell. The wave function $`|\mathrm{\Phi }_6(LSJMTT_3)`$ has orbital angular momentum $`L`$ and spin $`S`$ coupled to total angular momentum $`J`$, projection $`M`$, isospin $`T`$, and charge state $`T_3`$, and is explicitly written as
$`|\mathrm{\Phi }_6(LSJMTT_3)_{1234:56}=|\mathrm{\Phi }_\alpha (0000)_{1234}\varphi ^{LS}_p(R_{\alpha 5})\varphi ^{LS}_p(R_{\alpha 6})`$ (5)
$`\left\{[Y_{1m_l}(\mathrm{\Omega }_{\alpha 5})Y_{1m_l^{}}(\mathrm{\Omega }_{\alpha 6})]_{LM_L}\times [\chi _5({\displaystyle \frac{1}{2}}m_s)\chi _6({\displaystyle \frac{1}{2}}m_s^{})]_{SM_S}\right\}_{JM}`$ (6)
$`\times [\nu _5({\displaystyle \frac{1}{2}}t_3)\nu _6({\displaystyle \frac{1}{2}}t_3^{})]_{TT_3}.`$ (7)
Particles 1–4 are placed in the s-shell core with only spin-isospin degrees of freedom, while particles 5–6 are placed in p-wave orbitals $`\varphi _p^{LS}(R_{\alpha k})`$ that are functions of the distance between the center of mass of the core and the particle. Different amplitudes $`\beta _{LS}`$ are mixed to obtain an optimal wave function; for the $`J^\pi ,T=1^+,0`$ ground state of <sup>6</sup>Li the p-shell can have $`\beta _{01}`$, $`\beta _{21}`$, and $`\beta _{10}`$ terms.
The two-body correlation operator $`U_{ij}`$ is defined as:
$$U_{ij}=\underset{p=2,6}{}\left[\underset{ki,j}{}f_{ijk}^p(𝐫_{ik},𝐫_{jk})\right]u_p(r_{ij})O_{ij}^p,$$
(8)
where the $`O_{ij}^{p=2,6}`$ = $`𝝉_i𝝉_j`$, $`𝝈_i𝝈_j`$, $`𝝈_i𝝈_j𝝉_i𝝉_j`$, $`S_{ij}`$, and $`S_{ij}𝝉_i𝝉_j`$, and the $`f_{ijk}^p`$ is an operator-independent three-body correlation. The six radial functions $`f_{ss}(r)`$ and $`u_{p=2,6}(r)`$ are obtained from two-body Euler-Lagrange equations with variational parameters as discussed in detail in Ref. . Here we take them to be the same as in $`{}_{}{}^{4}\mathrm{He}`$, except that the $`u_{p=2,6}(r)`$ are forced to go to zero at large distance by multiplying in a cutoff factor, $`\left[1+\mathrm{exp}[R_u/a_u]\right]/\left[1+\mathrm{exp}[(rR_u)/a_u]\right]`$, with $`R_u`$ and $`a_u`$ as variational parameters. The $`f_{sp}`$ correlation is constructed to be similar to $`f_{ss}`$ for small separations, but goes to a constant of order unity at large distances:
$$f_{sp}(r)=\left[a_{sp}+\frac{b_{sp}}{1+\mathrm{exp}[(rR_{sp})/a_{sp}]}\right]f_{ss}(r)+c_{sp}(1\mathrm{exp}[(r/d_{sp})^2]),$$
(9)
where $`a_{sp}`$, $`b_{sp}`$, etc. are additional variational parameters. The $`f_{pp}(r)`$ correlation is given by:
$$f_{pp}(r)=u_d(r)/[13u_2(r)+u_3(r)3u_4(r)]r,$$
(10)
where $`u_d(r)`$ is the $`S`$-wave part of the exact deuteron wave function. Then $`U_{ij}`$ acting on the pair of nucleons in the p-shell will regenerate the correct deuteron $`S`$-wave and an effective $`D`$-wave $`w_d(r)=f_{pp}(r)[u_5(r)3u_6(r)]\sqrt{8}r`$ from the tensor components. While the $`D`$-wave is not exact, we find that the resulting effective deuteron energy is $`2.16`$ MeV, quite close to the correct value.
These choices for $`f_{ss}`$, $`f_{sp}`$, $`f_{pp}`$, and $`u_{p=2,6}`$ guarantee that when the two p-shell particles are far from the s-shell core, the overall wave function factorizes as:
$$\mathrm{\Psi }_T[\varphi _p^{LS}(R_{\alpha d})]^2\psi _\alpha \psi _d,$$
(11)
where $`\psi _\alpha `$ is the variational $`{}_{}{}^{4}\mathrm{He}`$ wave function and $`\psi _d`$ is (almost) the exact deuteron wave function.
The single-particle functions $`\varphi _p^{LS}(R_{\alpha n})`$ describe correlations between the s-shell core and the p-shell nucleons, and have been taken in previous work to be solutions of a radial Schrödinger equation for a Woods-Saxon potential and unit angular momentum, with energy and Woods-Saxon parameters determined variationally. It is important for low-energy radiative captures that these functions reproduce faithfully the large-separation behavior of the wave function, because the matrix elements receive large contributions at cluster separations greater than 10 fm. In fact, below 400 keV, more than 15% of the electric quadrupole operator comes from cluster separations beyond 30 fm. We have therefore modified the $`{}_{}{}^{6}\mathrm{Li}`$ wave function for the capture calculation to enforce cluster-like behavior when the two p-shell nucleons are both far from the s-shell core.
In general, for light p-shell nuclei with an asymptotic two-cluster structure, such as $`\alpha d`$ in $`{}_{}{}^{6}\mathrm{Li}`$ or $`\alpha t`$ in <sup>7</sup>Li, we want the large separation behavior to be
$$[\varphi _p^{LS}(r\mathrm{})]^nW_{km}(2\gamma r)/r,$$
(12)
where $`W_{km}(2\gamma r)`$ is the Whittaker function for bound-state wave functions in a Coulomb potential (see below) and $`n`$ is the number of p-shell nucleons. We achieve this by solving the equation
$$\left[\frac{\mathrm{}^2}{2\mu _{41}}\left(\frac{d^2}{dr^2}\frac{\mathrm{}(\mathrm{}+1)}{r^2}\right)+V(r)+\mathrm{\Lambda }(r)\right]r\varphi _p^{LS}(r)=0,$$
(13)
with $`\mathrm{}=1`$, $`\mu _{41}`$ the reduced mass of one nucleon against four, and $`V(r)`$ a parametrized Woods-Saxon potential plus Coulomb term:
$$V(r)=\frac{V_0}{1+\mathrm{exp}[(rR_0)/a_0]}+\frac{2(Z2)}{n}\frac{e^2}{r}F(r).$$
(14)
Here $`V_0`$, $`R_0`$, and $`a_0`$ are variational parameters, $`(Z2)/n`$ is the average charge of a p-shell nucleon, and F(r) is a form factor obtained by folding $`\alpha `$ and proton charge distributions together. The $`\mathrm{\Lambda }(r)`$ is a Lagrange multiplier that enforces the asymptotic behavior at large $`r`$, but is cut off at small $`r`$ by means of a variational parameter $`c_0`$:
$$\mathrm{\Lambda }(r)=\lambda (r)\left[1\mathrm{exp}\left((r/c_0)^2\right)\right].$$
(15)
The $`\lambda (r)`$ is given by
$$\lambda (r)=\frac{\mathrm{}^2}{2\mu _{41}}\left[\frac{1}{u_L}\frac{d^2u_L}{dr^2}\frac{2}{r^2}\right]\frac{2(Z2)}{n}\frac{e^2}{r},$$
(16)
where $`u_L`$ is directly related to the Whittaker function:
$$u_L/r=(W_{km}(2\gamma r)/r)^{1/n}.$$
(17)
Here $`\gamma ^2=2\mu _{4n}B_{4n}/\mathrm{}^2`$, with $`\mu _{4n}`$ and $`B_{4n}`$ the appropriate two-cluster effective mass and binding energy, $`k=2(Z2)e^2\mu _{4n}/\mathrm{}^2\gamma `$, and $`m=L+\frac{1}{2}`$.
For $`{}_{}{}^{6}\mathrm{Li}`$, $`B_{42}=1.47`$ MeV and $`L=0`$, 2 corresponding to the asymptotic $`S`$\- and $`D`$-waves of the $`{}_{}{}^{6}\mathrm{Li}`$ ground state, or amplitudes $`\beta _{01}`$ and $`\beta _{21}`$ in Eq.(4). There is no asymptotic $`\alpha d`$ cluster corresponding to the $`L=1`$ amplitude $`\beta _{10}`$. Nevertheless, the energy of the $`{}_{}{}^{6}\mathrm{Li}`$ ground state is variationally improved $`0.1`$ MeV by including such a component in the wave function, so we set the asymptotic behavior of $`\varphi _p^{10}(r)`$ to a binding energy of 3.70 MeV, which is the threshold for $`{}_{}{}^{6}\mathrm{Li}^4\mathrm{He}+p+n`$ breakup.
In Refs. the $`f_{ijk}^{sss}`$ three-body correlation of Eq.(4) was a valuable and inexpensive improvement to the trial function, but no $`f_{ijk}^{ssp}`$ or $`f_{ijk}^{spp}`$ correlations could be found that were of any benefit. However, in the present work we find the correlation
$$f_{n56}^{spp}=1+q_1[f_{ss}(r_{56})/f_{pp}(r_{56})1]\mathrm{exp}[q_2(r_{n5}+r_{n6})],$$
(18)
with $`q_{1,2}`$ as variational parameters, to be very useful. It effectively alters the central pair correlation between the two p-shell nucleons from their asymptotic, deuteron-like form to be more like the pair correlations within the s-shell when the two particles are close to the core. This correlation improves the binding energy by $`0.5`$ MeV.
In Refs. we reported energies for the trial function $`\mathrm{\Psi }_T`$ of Eq.(1), and for a more sophisticated variational wave function, $`\mathrm{\Psi }_V`$, which adds two-body spin-orbit and three-body spin- and isospin-dependent correlation operators. The $`\mathrm{\Psi }_V`$ gives improved binding compared to $`\mathrm{\Psi }_T`$ in both $`{}_{}{}^{4}\mathrm{He}`$ and $`{}_{}{}^{6}\mathrm{Li}`$, but is significantly more expensive to construct because of the numerical derivatives required for the spin-orbit correlations. In the case of an energy calculation the derivatives are also needed for the evaluation of $`L`$-dependent terms in AV18, so the cost is only a factor of two in computation. However for the evaluation of other expectation values the relative cost increase is $`6A`$. Thus in the present work we choose to use $`\mathrm{\Psi }_T`$ for non-energy evaluations; this proved quite adequate in our studies of $`{}_{}{}^{6}\mathrm{Li}`$ form factors .
The variational Monte Carlo (VMC) energies and point proton rms radii obtained with $`\mathrm{\Psi }_T`$ and $`\mathrm{\Psi }_V`$ are shown in Table I along with the results of essentially exact Green’s function Monte Carlo (GFMC) calculations and the experimental values. We note that the underbinding of $`{}_{}{}^{6}\mathrm{Li}`$ in the GFMC calculation is the fault of the AV18/UIX model and not the many-body method; it can be corrected by the introduction of more sophisticated three-nucleon potentials .
The present variational trial functions with the imposed proper asymptotic behavior in fact give slightly better energies than the older shell-model-like correlations of Refs. . Unfortunately, because the variational $`{}_{}{}^{6}\mathrm{Li}`$ energy is not below that of separated alpha and deuteron clusters, it is possible to lower the energy significantly by making the wave function more and more diffuse. Therefore, the variational parameters were constrained to give a reasonable rms radius, as seen in Table I. Although the variational $`{}_{}{}^{6}\mathrm{Li}`$ wave function remains unbound relative to alpha-deuteron breakup, it has the correct energy at large particle separations, as shown in Fig. 2. There we show the local energy, $`E(𝐑)`$, which is the energy for a given spatial configuration $`𝐑=(𝐫_1,\mathrm{},𝐫_6)`$, binned with its Monte Carlo statistical variance as a function of the sum $`R=_i|𝐫_i|`$ of the particle distances from the center of mass. The average energy is shown by the solid line, while the dashed line gives the experimental binding. At the bottom of the figure is a histogram of the number of samples to show their distribution with $`R`$. We observe that the energy in compact configurations is clearly too small, but for $`R>18`$ fm the energies scatter about the experimental binding.
The asymptotic two-cluster behavior of our six-body wave function can be studied by computing the two-cluster $`\alpha d`$ distribution function, $`𝒜\psi _\alpha \psi _d^{m_d},𝐫_{\alpha d}\psi _{\mathrm{Li}}^{m_6}`$, described in Ref. . It can be expressed in terms of Clebsch-Gordan factors, $`Y_{LM}(\mathrm{\Omega }_{\alpha d})`$, and two radial functions $`R_L(r_{\alpha d})`$ for the $`S`$\- and $`D`$-waves, which are plotted in Fig. 3. The $`R_L`$ can be used to extract the asymptotic normalization constants $`C_0`$ and $`C_2`$, by taking their ratio with the corresponding Whittaker functions, and the asymptotic $`D/S`$ ratio $`\eta `$. We find that $`C_0`$ does not become asymptotic until $`r_{\alpha d}8`$ fm, where we obtain $`C_0=2.28\pm 0.02`$. However, $`\eta `$ is already asymptotic for $`r_{\alpha d}4`$ fm, with the value $`0.026\pm 0.001`$.
This value of $`C_0`$ is at the lower end of the range of values (2.3–2.9) used in a number of other studies . However, the value of $`\eta `$ is consistent with the recent analysis of $`{}_{}{}^{6}\mathrm{Li}`$ \+ $`{}_{}{}^{4}\mathrm{He}`$ elastic scattering , which obtains $`0.025\pm 0.006\pm 0.010`$. This is a big improvement over our earlier value of $`0.07\pm 0.02`$ reported in Ref. , which was obtained with the shell-model-like wave function. However, our quadrupole moment $`Q=0.71\pm 0.08`$ fm<sup>2</sup> remains an order of magnitude too large compared to the experimental value of $`0.08`$ fm<sup>2</sup>. The integrated $`D_2^{\alpha d}`$ parameter value is $`0.27\pm 0.01`$ fm<sup>2</sup>.
The spectroscopic factor is 0.86, in good agreement with the Robertson value of $`0.85\pm 0.04`$ . The spectroscopic factor is not very sensitive to the asymptotic behavior of the wave function, and differs little from our earlier value of 0.84 obtained with the shell-model-like wave function. We note that our shell-model-like wave functions for <sup>7</sup>Li and <sup>6</sup>He provided an excellent prediction for the spectroscopic factors observed in $`{}_{}{}^{7}\mathrm{Li}(e,e^{}p)`$ scattering .
### B $`\alpha `$d initial state
The initial alpha-deuteron scattering wave function $`|\psi _{\alpha d};LSJM`$ having orbital angular momentum $`L`$ and channel spin $`S`$ (=1) coupled to total $`JM`$, is expressed as
$$|\psi _{\alpha d};LSJM=𝒜\left\{\varphi _{\alpha d}(r_{\alpha d})Y_{LM_L}(\widehat{𝐫}_{\alpha d})\underset{ij}{}G_{ij}|\psi _\alpha \psi _d^{m_d}\right\}_{LSJM},$$
(19)
where curly braces indicate angular momentum coupling, $`𝒜`$ antisymmetrizes between clusters, and $`𝐫_{\alpha d}`$ denotes the separation between the centers of mass of the $`\alpha `$ and $`d`$ clusters.
Our scattering wave function is built from the alpha and deuteron ground-state wave fuctions, $`\psi _\alpha `$ and $`\psi _d^{m_d}`$ (where $`m_d`$ denotes the deuteron spin projection), from a nucleon pair correlation operator $`G_{ij}`$, and from an intercluster scalar correlation $`\varphi _{\alpha d}(r_{\alpha d})`$. $`G_{ij}`$ is the identity operator if particles $`i`$ and $`j`$ belong to the same cluster (alpha or deuteron). Otherwise, it is a pair correlation operator derived to describe nuclear matter , and it contains both central and non-central terms. It takes into account, in an approximate way, the distortions introduced in each cluster by the interactions with the particles in the other cluster, and becomes the identity operator as particle separations become large.
The intercluster correlation $`\varphi _{\alpha d}(r_{\alpha d})`$ is a two-body radial wave function describing alpha-deuteron scattering as the scattering of two point particles. It is generated from a Schrödinger equation, with the potential chosen to reproduce the results of phase-shift analyses of alpha-deuteron scattering data in the two-body model. Several such potentials exist in the literature, and we saw no need to re-fit the data. The potentials which we have applied to this study are the potentials of Kukulin and Pomerantsev, Refs. (hereafter, “KP I” and “KP II”, respectively), and the potential of Langanke, Ref. . Because the potential of Langanke was produced for the slightly different problem of describing the $`\alpha d`$ system in the resonating group method, it required modification to match our purposes. This consisted of a small adjustment to the Woods-Saxon radius to reproduce the energy of the $`3^+`$ resonance derived from electron scattering, 711 keV. The most important difference between these potentials, which give very similar cross sections in our calculations, is that the KP potentials were fitted to the phase shifts in the odd and even partial waves separately, to produce the expected parity-dependent potential. The work by Langanke was concerned only with even partial waves, and did not fit the odd partial waves, although it provides a fair qualitative description of the $`P`$-wave phase shifts at low energy.
The potentials are very similar to each other in form. All of the potentials used contain both central and spin-orbit terms, and the spin-orbit force in the even partial waves is well-constrained by the spacing of the $`D`$-wave scattering resonances. While a tensor component could in principle be derived to reproduce the ground-state quadrupole moment (assuming total alpha-deuteron cluster parentage of the ground state), its effect on the scattering data is too small to support deriving it from the scattering phase shifts. (A more ambitious approach to derive a tensor interaction would involve computing energies at fixed separation between the $`\alpha `$ and $`d`$ clusters and varying the deuteron orientation; however, we have not pursued such an approach here.) Coupling to other (e.g., cluster breakup) channels has also been neglected; this is well-justified below the $`{}_{}{}^{5}\mathrm{He}+p`$ and $`{}_{}{}^{5}\mathrm{Li}+n`$ thresholds at 3.12 and 4.20 MeV, respectively. We note that although the KP potentials and the Langanke potential all reproduce the scattering phase shifts very well, and they produce very similar ground-state wave functions at small cluster separation, the ground-state binding of the KP II potential is too large by almost a factor of two. All of the potentials we used to generate $`\varphi _{\alpha d}`$ are deep potentials with a forbidden zero-node state. They produce the expected one-node structure of the ground-state alpha-deuteron wave function required by the Pauli principle, as illustrated by our $`R_L`$ functions in Fig.3, and enforce (by orthogonality to the forbidden state) the most important consequences of the Pauli principle for $`\varphi _{\alpha d}`$.
A complication arises from the phenomenological treatment of the initial state. In reality, the ground-state and $`S`$-wave scattering-state wave functions are orthogonal, because they are different eigenstates of the same ($`J^\pi =1^+`$) potential. Because our calculation does not generate them explicitly as such, our ground state and $`S`$-wave scattering states are not necessarily orthogonal. For purposes of our capture calculation, this has the effect of generating spurious contributions to the $`M1`$ transition operator. For $`S`$-wave captures, we have modified the initial state to ensure orthogonality with the ground state. This is done by subtracting from the initial state, Eq.(19), its projection onto a complete set of ground states,
$$|\psi _{\alpha d};LSJM^{}=|\psi _{\alpha d};LSJM\underset{m_6}{}|\psi _{\mathrm{Li}}^{m_6}\psi _{\mathrm{Li}}^{m_6}|\psi _{\alpha d};LSJM,$$
(20)
and computing $`M1`$ transitions from the adjusted state. This procedure reduced the $`M1`$ contribution to our computed cross sections by more than a factor of 10, roughly to the level of Monte Carlo sampling noise. Further evidence that the remaining contribution is spurious is its strong dependence on the phenomenological radial correlation in the scattering state, discussed below.
## III Cross Section and Transition Operators
The cross section for $`\alpha `$$`d`$ radiative capture at c.m. energy $`E_{c.m.}`$ can be written as
$$\sigma (E_{c.m.})=\frac{8\pi }{3}\frac{\alpha }{v_{\mathrm{rel}}}\frac{q}{1+q/m_{\mathrm{Li}}}\underset{LSJ\mathrm{}}{}\left[\left|E_{\mathrm{}}^{LSJ}(q)\right|^2+\left|M_{\mathrm{}}^{LSJ}(q)\right|^2\right],$$
(21)
where $`\alpha `$ is the fine structure constant ($`\alpha =e^2`$; we take $`\mathrm{}=c=1`$ in this Section), $`v_{\mathrm{rel}}`$ is the $`\alpha `$$`d`$ relative velocity, and $`E_{\mathrm{}}^{LSJ}(q)`$ and $`M_{\mathrm{}}^{LSJ}(q)`$ are the reduced matrix elements (RMEs) of the electric and magnetic multipole operators connecting the $`\alpha `$$`d`$ scattering state in channel $`LSJ`$ to the <sup>6</sup>Li ground state having $`J^\pi ,T=1^+`$,0. The c.m. energy of the emitted photon is given by
$`q`$ $`=`$ $`m_{\mathrm{Li}}\left[1+\sqrt{1+{\displaystyle \frac{2}{m_{\mathrm{Li}}}}(m_d+m_\alpha m_{\mathrm{Li}}+E_{c.m.})}\right]`$ (22)
$``$ $`m_d+m_\alpha m_{\mathrm{Li}}+E_{c.m.},`$ (23)
where $`m_d`$, $`m_\alpha `$, and $`m_{\mathrm{Li}}`$ are the rest masses of deuteron, <sup>4</sup>He, and <sup>6</sup>Li, respectively. The astrophysical $`S`$-factor is then related to the cross section via
$$S(E_{c.m.})=E_{c.m.}\sigma (E_{c.m.})\mathrm{exp}(4\pi \alpha /v_{\mathrm{rel}}).$$
(24)
The $`M1`$, $`M2`$, $`E1`$, $`E2`$, and $`E3`$ transitions involving $`\alpha `$$`d`$ scattering states with relative orbital angular momentum up to $`L=2`$ have been retained in the evaluation of the cross section. Initial states with $`L=3`$ and $`L=4`$ were examined in the early phases of this work, but proved to be very small, and were not retained in the final calculation.
Since the energies of interest in the present study are below 4 MeV, and consequently $`qR_{\mathrm{Li}}0.05`$ ($`R_{\mathrm{Li}}`$ is the <sup>6</sup>Li point rms radius), the long-wavelength aproximation (LWA) of the multipole operators would naively be expected to be adequate to compute the associated RMEs. While the LWA is indeed sufficient for $`E2`$ transitions, which dominate the cross section at energies $`E_{c.m.}>400`$ keV, it is however inaccurate to compute the RMEs of $`M1`$ and, particularly, $`E1`$ transitions in the LWA since these are suppressed. (This fact has already been alluded to in the introduction). It is therefore necessary to include higher order terms, beyond the leading one which normally is taken to define the LWA, in the expansion in powers of $`q`$, so-called “retardation terms”. It is useful to review how these corrections arise for $`E1`$ transitions. We should point out that some of these same issues have recently been discussed in Ref. , in the context of a calculation of $`E1`$ strength in $`d`$$`p`$ radiative capture at energies below 100 keV.
### A The $`E_1`$ operators
In principle, the $`E1`$ RMEs are obtained from
$$E_1^{LSJ}(q)=\frac{\sqrt{3}}{JM,1\lambda |1m_6}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|E_{1\lambda }(q)|\psi _{\alpha d};LSJM,$$
(25)
$$E_{1\lambda }(q)=\frac{1}{q}𝑑𝐱𝐣(𝐱)\times j_1(qx)𝐘_{1\lambda }^{11}(\widehat{𝐱}),$$
(26)
where $`𝐣(𝐱)`$ is the nuclear current density operator, $`j_1(qx)`$ is the spherical Bessel function of order one, and $`𝐘_{1\lambda }^{11}(\widehat{𝐱})`$ are vector spherical harmonic functions. For ease of presentation, the factor $`\sqrt{3}/JM,1\lambda |1m_6`$ occurring in the definition of the RMEs will be suppressed in the following discussion. By expanding the Bessel function in powers of $`q`$ and after standard manipulations , the $`E_1`$ operator can be expressed as
$$E_{1\lambda }(q)E_{1\lambda }(q;\mathrm{LWA1})+E_{1\lambda }(q;\mathrm{LWA2})+E_{1\lambda }(q;\mathrm{LWA3}),$$
(27)
where
$$E_{1\lambda }(q;\mathrm{LWA1})=\frac{\sqrt{2}}{3}[H,𝑑𝐱xY_{1\lambda }(\widehat{𝐱})\rho (𝐱)],$$
(28)
$$E_{1\lambda }(q;\mathrm{LWA2})=\frac{\mathrm{i}q^2}{3\sqrt{2}}𝑑𝐱xY_{1\lambda }(\widehat{𝐱})𝐱𝐣(𝐱),$$
(29)
$$E_{1\lambda }(q;\mathrm{LWA3})=\frac{\sqrt{2}q^2}{15}[H,𝑑𝐱x^3Y_{1\lambda }(\widehat{𝐱})\rho (𝐱)].$$
(30)
Here the continuity equation has been used to relate $`𝐣(𝐱)`$ occurring in $`E_{1\lambda }(q;\mathrm{LWA1})`$ and $`E_{1\lambda }(q;\mathrm{LWA3})`$ to the commutator $`\mathrm{i}[H,\rho (𝐱)]`$, where $`\rho (𝐱)`$ is the charge density operator. Evaluating the RMEs of these operators leads to
$$E_1^{LSJ}(q)E_1^{LSJ}(q;\mathrm{LWA1})+E_1^{LSJ}(q;\mathrm{LWA2})+E_1^{LSJ}(q;\mathrm{LWA3}),$$
(31)
with
$$E_1^{LSJ}(q;\mathrm{LWA1})=\frac{\sqrt{2}q}{3}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|𝑑𝐱xY_{1\lambda }(\widehat{𝐱})\rho (𝐱)|\psi _{\alpha d};LSJM,$$
(32)
$$E_1^{LSJ}(q;\mathrm{LWA2})=\frac{\mathrm{i}q^2}{3\sqrt{2}}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|𝑑𝐱xY_{1\lambda }(\widehat{𝐱})𝐱𝐣(𝐱)|\psi _{\alpha d};LSJM,$$
(33)
$$E_1^{LSJ}(q;\mathrm{LWA3})=\frac{\sqrt{2}q^3}{15}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|𝑑𝐱x^3Y_{1\lambda }(\widehat{𝐱})\rho (𝐱)|\psi _{\alpha d};LSJM,$$
(34)
where terms up to order $`q^3`$ have been retained, and it has been assumed that the initial and final VMC wave functions are exact eigenfunctions of the Hamiltonian – which, incidentally, is clearly not the case, see the previous Section – so that $`[H,\rho (𝐱)]q\rho (𝐱)`$ in the matrix element, ignoring the kinetic energy of the recoiling <sup>6</sup>Li.
The nuclear electromagnetic charge and current density operators have one- and two-body components.
$`\rho (𝐱)`$ $`=`$ $`{\displaystyle \underset{i}{}}\rho _i^{(1)}(𝐱)+{\displaystyle \underset{i<j}{}}\rho _{ij}^{(2)}(𝐱);`$ (35)
$`𝐣(𝐱)`$ $`=`$ $`{\displaystyle \underset{i}{}}𝐣_i^{(1)}(𝐱)+{\displaystyle \underset{i<j}{}}𝐣_{ij}^{(2)}(𝐱).`$ (36)
The one-body terms $`\rho _i^{(1)}`$ and $`𝐣_i^{(1)}`$ have the standard expressions obtained from a non-relativistic reduction of the covariant single-nucleon current. The charge density is written as
$$\rho _i^{(1)}(𝐱)=\rho _{i,\mathrm{NR}}^{(1)}(𝐱)+\rho _{i,\mathrm{RC}}^{(1)}(𝐱),$$
(37)
with
$$\rho _{i,\mathrm{NR}}^{(1)}(𝐱)=ϵ_i\delta (𝐱𝐫_i),$$
(38)
$$\rho _{i,\mathrm{RC}}^{(1)}(𝐱)=\frac{1}{8m^2}\left(2\mu _iϵ_i\right)_x[𝝈_i\times 𝐩_i,\delta (𝐱𝐫_i)]_+,$$
(39)
where the Darwin-Foldy relativistic correction has been neglected in the expression for $`\rho _{i,\mathrm{RC}}^{(1)}`$, since it gives no contribution to Eq. (32) (see below). Hereafter, $`[\mathrm{},\mathrm{}]_+`$ denotes the anticommutator. The one-body current density is expressed as
$$𝐣_i^{(1)}(𝐱)=\frac{1}{2m}ϵ_i[𝐩_i,\delta (𝐱𝐫_i)]_+\frac{1}{2m}\mu _i𝝈_i\times _x\delta (𝐱𝐫_i).$$
(40)
The following definitions have been introduced:
$`ϵ_i`$ $``$ $`{\displaystyle \frac{1+\tau _{i,z}}{2}},`$ (41)
$`\mu _i`$ $``$ $`{\displaystyle \frac{\mu ^S+\mu ^V\tau _{i,z}}{2}},`$ (42)
where $`\mu ^S`$ ($`\mu ^S`$=0.88 n.m.) and $`\mu ^V`$ ($`\mu ^V`$=4.706 n.m.) are the isoscalar and isovector combinations of the proton and neutron magnetic moments.
The two-body charge and current density operators have pion range terms and additional, short-range terms due to heavy meson exchanges . However, the pion-exchange current density is an isovector operator, and its matrix element vanishes identically in the radiative $`\alpha `$$`d`$ capture, while the heavy-meson exchange (charge and current) operators are expected to give negligible contributions. Hence, in the present study we only retain the charge density operator associated with pion exchange. It is explicitly given by
$$\rho _{ij,\pi }^{(2)}(𝐱)=\frac{f_\pi ^2}{2mm_\pi ^2}𝝉_i𝝉_jI_\pi ^{}(r_{ij})[𝝈_j\widehat{𝐫}_{ij}𝝈_i_x\delta (𝐱𝐫_i)+ij],$$
(43)
where $`m_\pi `$ is the charged pion mass, $`f_\pi `$ the pion-nucleon coupling constant ($`f_\pi ^2/4\pi =0.075`$), $`𝐫_{ij}=𝐫_i𝐫_j`$ is the separation of nucleons $`i`$ and $`j`$, and the function $`I_\pi (r)`$ is defined as
$$I_\pi (r)=\frac{1}{4\pi r}\left[\mathrm{e}^{m_\pi r}\mathrm{e}^{\mathrm{\Lambda }_\pi r}\frac{1}{2}\left(1\frac{m_\pi ^2}{\mathrm{\Lambda }_\pi ^2}\right)\mathrm{\Lambda }_\pi r\mathrm{e}^{\mathrm{\Lambda }_\pi r}\right].$$
(44)
The prime denotes differentiation with respect to its argument. For the parameter $`\mathrm{\Lambda }_\pi `$, we have taken the value 1.05 GeV.
We now consider, in turn, the various contributions LWA1, LWA2, and LWA3 to the $`E_1`$ RME. Insertion of the NR, RC, and $`\pi `$-exchange charge density operators in Eq. (32) allows us to write correspondingly
$$E_1^{LSJ}(q;\mathrm{LWA1})E_1^{LSJ}(q;\mathrm{LWAc})+E_1^{LSJ}(q;\mathrm{LWAb})+E_1^{LSJ}(q;\mathrm{LWA}\pi )$$
(45)
where
$$E_1^{LSJ}(q;\mathrm{LWAc})=\frac{\sqrt{2}q}{3}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|\underset{i}{}ϵ_ir_iY_{1\lambda }(\widehat{𝐫}_i)|\psi _{\alpha d};LSJM,$$
(46)
$$E_1^{LSJ}(q;\mathrm{LWAb})=\frac{\sqrt{2}q}{3}\sqrt{\frac{3}{4\pi }}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|\underset{i}{}\frac{2\mu _iϵ_i}{4m^2}(𝝈_i\times 𝐩_i)_\lambda |\psi _{\alpha d};LSJM,$$
(47)
$`E_1^{LSJ}(q;\mathrm{LWA}\pi )={\displaystyle \frac{\sqrt{2}q}{3}}\sqrt{{\displaystyle \frac{3}{4\pi }}}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|{\displaystyle \underset{i<j}{}}`$ $`{\displaystyle \frac{f_\pi ^2}{2mm_\pi ^2}}𝝉_i𝝉_jI_\pi ^{}(r_{ij})[\sigma _{i,\lambda }𝝈_j\widehat{𝐫}_{ij}`$ (49)
$`+ij]|\psi _{\alpha d};LSJM,`$
where $`(\sigma \times 𝐩)_\lambda `$ and $`\sigma _{i,\lambda }`$ denote spherical components.
Next, for the terms of order $`q^2`$ and $`q^3`$, Eqs. (33) and (34), we find:
$`E_1^{LSJ}(q;\mathrm{LWA2})={\displaystyle \frac{\mathrm{i}q^2}{3\sqrt{2}}}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|{\displaystyle \underset{i}{}}`$ $`{\displaystyle \frac{ϵ_i}{2m}}[𝐩_i,𝐫_ir_iY_{1\lambda }(\widehat{𝐫}_i)]_+`$ (50)
$``$ $`{\displaystyle \frac{\mu _i}{2m}}\sqrt{{\displaystyle \frac{3}{4\pi }}}(𝐫_i\times 𝝈_i)_\lambda |\psi _{\alpha d};LSJM,`$ (51)
$$E_1^{LSJ}(q;\mathrm{LWA3})=\frac{\sqrt{2}q^3}{15}\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|\underset{i}{}ϵ_ir_i^3Y_{1\lambda }(\widehat{𝐫}_i)|\psi _{\alpha d};LSJM,$$
(52)
where in the last equation the contributions from relativistic corrections and pion-exchange have been neglected.
The leading contribution $`E_1^{LSJ}(q;\mathrm{LWAc})`$ to the $`E_1`$ RME vanishes in $`\alpha `$$`d`$ capture, since the initial and final states have zero isospin, and the corresponding wave functions do not depend on the c.m. position:
$$E_1^{LSJ}(\mathrm{LWAc})\mathrm{\Psi }_{\mathrm{Li}}^{m_6}|\underset{i}{}𝐫_{i,\lambda }|\psi _{\alpha d};LSJM=0,$$
(53)
where the coordinates $`𝐫_i`$ are defined relative to that of the c.m..
There are also more subtle corrections, which can potentially lead to additional, (relatively) significant $`E_1`$ contributions. One such correction is the possibility of isospin admixtures with $`T>0`$ in the <sup>6</sup>Li and $`\alpha `$$`d`$ states, originating from the isospin-symmetry-breaking interactions present in the AV18/UIX model. These $`T>0`$ components are ignored in the VMC wave functions used here.
An additional correction – which we take into account – arises from a relativistic correction to the center of mass, and makes the LWAc contribution nonzero. This correction arises because translation-invariant wave functions require not center-of-mass but center-of-energy coordinates. Such coordinates constitute the relativistic analog of the center-of-mass coordinates appropriate to nonrelativistic problems. The distinction is usually not crucial at low energies, but it matters here because the LWAc term would otherwise be zero. Since the coordinates in which we compute wave functions are not center-of-energy coordinates, the dipole moment
$$𝐝=\underset{i}{}ϵ_i𝐫_i,$$
(54)
which enters Eq. (53), should be replaced by the expression
$`𝐝`$ $`=`$ $`{\displaystyle \underset{i}{}}ϵ_i(𝐫_i𝐫_{CE})`$ (55)
$`=`$ $`{\displaystyle \underset{i}{}}ϵ_i𝐫_{CE},`$ (56)
where
$$𝐫_{CE}=\frac{_iE_i𝐫_i}{_iE_i},$$
(57)
and
$$E_i=m+\frac{p_i^2}{2m}+\frac{1}{2}\underset{ji}{}v_{ij}+\frac{1}{3}\underset{jki}{}V_{ijk}.$$
(58)
Here, $`v_{ij}`$ and $`V_{ijk}`$ are the two- and three-nucleon potentials. We have thus replaced the particle masses in the expression for center of mass by their total energies (rest energy plus kinetic and potential energies). Since the nucleon masses contribute most of the nucleons’ relativistic energies, the center of mass is only slightly different from the center of energy. (The rms radius of the center of energy for our <sup>6</sup>Li variational ground state is $`(9.1\pm 2.7)\times 10^3`$ fm.) We note that this correction has been taken into account in the past (in two-body models) by putting the measured alpha and deuteron masses into calculations, rather than $`4m`$ and $`2m`$ . These differ by the mass defects of the alpha particle and deuteron.
The $`E1`$ cross sections arising from the center-of-energy correction are larger than the $`E1`$ cross sections measured in Ref. by about a factor of six. There is a significant ($`30\%`$) LWAb spin-orbit contribution which becomes more important with increasing energy; the next-largest contributions are from the LWA3 and LWA2 operators. All of these terms except LWAb are too small in our calculation to account for the observed transition by more than a factor of ten, but LWA3 is responsible for about 20% of the total $`S`$ factor at zero energy. We have also included the contribution labelled LWA$`\pi `$ in the present study, and we find it to be much smaller than LWAb. Note that although the LWAb term is large enough to account for the magnitude of the $`E1`$ data alone, it is of the wrong sign to account for the asymmetry from which the $`E1`$ strength is extracted.
### B The Monte Carlo Calculation
We calculated matrix elements by modifying the variational Monte Carlo code described in Ref. to perform Monte Carlo integrations of matrix elements between the scattering and ground states described above. The Monte Carlo algorithm used was the Metropolis algorithm, and we tried several weighting functions, all based on the $`{}_{}{}^{6}\mathrm{Li}`$ ground state. The final calculation used the weighting function
$$W(𝐑)=\sqrt{\psi _{\mathrm{Li}}^{m_6}(𝐑)|\psi _{\mathrm{Li}}^{m_6}(𝐑)}$$
(59)
(where the bra-ket product here indicates summation over spin and isospin degrees of freedom for a given spatial configuration $`𝐑=(𝐫_1,𝐫_2,\mathrm{},𝐫_6)`$ of the particles; in all other cases, it also indicates integration over space coordinates). This weighting function was chosen to reflect approximately the expected behavior of the matrix element, and to obtain good sampling both at small cluster separation and out to 40 fm cluster separation. In the low-energy calculation, 3% of the $`E1`$ matrix element and 15% of the $`E2`$ matrix element come from the region beyond 30 fm cluster separation.
Because all of the energy dependence is contained in the relative wave function $`\varphi _{\alpha d}(r_{\alpha d})`$ and the transition operators, the matrix element for a given scattering partial wave can be re-written as
$$T_{\mathrm{}}^{LSJ}(q)=_0^{\mathrm{}}dxx^2\varphi _{\alpha d}(x)\psi _{\mathrm{Li}}^{m_6}|T_\mathrm{}\lambda (q)𝒜\{\delta (xr_{\alpha d})Y_L^{M_L}(\widehat{𝐫}_{\alpha d}))\underset{ij}{}G_{ij}|\psi _\alpha \psi _d^{m_d}\}_{LSJM},$$
(60)
using a technique from Ref. . Here $`T_\mathrm{}\lambda `$ denotes any of the $`E_{\mathrm{}}`$ and $`M_{\mathrm{}}`$ operators. The partial-wave expansion of the current operator contains the photon energy only in multiplicative factors on each term of the expansion. The integration over all coordinates except $`x`$ can therefore be calculated once for each partial wave by the Monte Carlo code, and the result can then be used to compute the full integral for as many energies as desired by recomputing $`\varphi _{\alpha d}`$ only, greatly reducing the amount of computation.
The calculation sampled 2 000 000 Monte Carlo configurations, summing over all 15 partitions of the six particles into clusters at each step. Typical results for the integrand in Eq.(60) at low energy are shown for two transitions, at two different energies, in Fig. 4. Calculations at energies above about 1 MeV had integrands that were essentially zero beyond 30 fm cluster separation, and therefore suffered little from limited sampling.
We also performed a second set of calculations, this time computing functions which had the expected asymptotic forms of the operator densities as a function of $`x`$ at large cluster separations, and normalizing them to the Monte-Carlo sampled operator densities between 12 and 22 fm. These asymptotic forms, given by multiplying Eq. (12) by powers of $`r_{\alpha d}`$ appropriate for each operator, were then used instead of the Monte Carlo operator densities at cluster separations greater than 12 fm, in order to check that our results were not affected by the effective cutoff in sampling beyond 45 fm (due to very small $`W(𝐑)`$) or by poor sampling elsewhere in the tails of the wave function. For all of the dominant reaction mechanisms, this gave essentially the same result as the explicit calculation. The smooth curves of Fig. 4 were computed in this way.
## IV Results
Our main result is the total capture cross section as a function of energy, shown in comparison with laboratory data in Fig. 5 as an astrophysical $`S`$-factor. Note the good agreement with the capture data below 3 MeV, reflecting a reasonable value of the alpha-deuteron asymptotic normalization coefficient of the <sup>6</sup>Li ground state, $`C_0`$, and accurate reproduction of the scattering phase shifts. The disagreement with the data above 3 MeV is a generic feature of direct-capture models, and is usually attributed to neglected couplings to breakup channels, and to channels with nonzero isospin in the scattering state (such as the $`J^\pi ,T=0^+,1`$ state at 2.09 MeV). As shown in Fig. 6, the behavior at the $`3^+`$ resonance is nicely reproduced in location, width, and amplitude, simply by constraining the cluster potentials in the $`D`$ waves to reproduce the experimental phase shifts. The failure to match the energy dependence of the cross sections inferred below 500 keV from Coulomb-breakup experiments is shared by all other direct capture calculations, and could reflect either something that is not understood about the experiments, or something that is missing from the models. Our result with more detailed wave functions and operators than previously considered may be taken as a further indication that previous models were correct in this region.
In Figs. 7 and 8, we show the breakdown of the cross section into contributions from different multipolarities and incoming partial waves, again as $`S`$ factors. Fig. 7 shows results of two calculations, one from the Langanke potential and one from the KP II potential. The results are dominated by $`E2`$ transitions arising from $`D`$-wave scattering and by $`E1`$ transitions arising from $`P`$-wave scattering. Results from the two sets of scattering states are almost indistinguishable for these transitions, especially at low energies. In Fig. 8, the two sets of curves were both computed from the KP II potential. One set of curves was computed directly from the Monte Carlo calculation, while the other was computed from the expected asymptotic behavior of the integrand of Eq. 60 beyond 12 fm, as described above. Fig. 9 shows a further breakdown of the $`E1`$ operator into its various terms, along with the measured $`E1`$ contributions . It is seen that:
1.) The $`E1`$ cross section is larger than the data of Robertson et al. by a factor of about 7 at 2 MeV. Most other models have also arrived at $`E1`$ cross sections that are too large relative to these data. A new aspect of our calculation is that the largest contribution is calculated from the nucleon-nucleon potentials as a relativistic center-of-mass correction. This corresponds closely to inserting measured cluster masses by hand in simpler models, as indicated by the fact that virtually identical (within 50%) results for the LWAc term are obtained from our six-body model and from a simple capture model using the same cluster potential, but with laboratory alpha and deuteron masses introduced. In Fig. 9, the top curve is the total $`E1`$ $`S`$ factor (arising mainly from the center-of-energy correction), and the other curves indicate the $`S`$ factors that would result from various small terms in the E1 operator individually. (These differ from their actual contributions to the $`S`$ factor, since these contributions add coherently in the matrix element.) The center-of-energy and LWAb terms are obviously much more important than the other terms in our calculation, and the smaller terms do not affect the total significantly.
We note also that Robertson et al. report the opposite sign from what is expected for the $`E1`$ transition in their experiment. None of the effects in our calculation can produce an effect of the correct magnitude with the observed “wrong” sign, or reproduce the sharp drop in $`E1`$ amplitude from 1.63 to 1.33 MeV seen in the data.
2.) There is no significant $`M1`$ transition originating from the $`S`$-wave scattering state at lowest order in LWA, and we estimate the next-to-lowest term to be much smaller than what we obtain for the lowest-order term. The $`M1`$ strength is expected to be generally very small, since it is zero by orthogonality arguments in a two-body model. However, at very low energy, any $`S`$-wave contribution present may be expected to dominate the total cross section, and many-body wave functions can (in principle) provide small pieces of the final-state wave function which can be reached in such a transition. Such an effect is present in neutron captures on the deuteron and on <sup>3</sup>He. The presence of a nonzero $`M1`$ contribution in our calculation raises two questions: to what extent is it a product of our choice of wave functions, and how will the presence of meson exchange currents (MEC), which is a factor-of-ten correction for the $`M1`$ cross section in $`{}_{}{}^{3}\mathrm{He}(n,\gamma )^4\mathrm{He}`$, affect our result? The answer to the first question lies in the discussion of orthogonalization at the end of Sec. II and in Fig. 7; orthogonal ground and scattering states produce much smaller $`M1`$ contributions than non-orthogonalized states, and the remaining $`M1`$ contribution is very sensitive to the choice of phenomenological scattering potentials, suggesting a near-cancellation that ought to be exact. The second question is less important than it may appear at first glance. The large relative MEC contributions to neutron captures on light nuclei typically arise from the isovector part of the $`M1`$ operator. The $`\alpha d`$ capture process is isoscalar; the isoscalar MEC operator is both significantly smaller than the corresponding isovector operator, and quantitatively less certain in magnitude. We have therefore neglected MEC contributions to the M1 operator. Only electric multipoles contribute to Coulomb breakup, so the discrepancy between our results and the indirect data (circles in Fig. 5) is not the result of omitting MEC.
To compare our results with those obtained in a simpler model, we have also computed $`\alpha d`$ capture by treating the two clusters as point particles, and the ground state as an $`S`$-wave bound state of these two clusters. We used the modified Langanke potential, described above, to compute both states. With this potential, the $`{}_{}{}^{6}\mathrm{Li}`$ ground state energy is 0.07 MeV too high, and this affects the energy dependence of the cross section. The asymptotic normalization of the wave function, which sets the scale for the nonresonant cross section, depends in any model on the detailed short-range behavior of the ground state, and is larger for this model than for our six-body $`{}_{}{}^{6}\mathrm{Li}`$ ground state. While the 711 keV resonance is equally well-reproduced in strength and width by both two-body and six-body models, the total cross sections away from resonance are as much as 16% smaller in the six-body model. The simple model also yields an $`E1`$ strength which is greater by about 30% at 3 MeV. The source of the greater $`E1`$ strength is an interplay between the differing asymptotic normalizations, the treatment of the center-of-energy correction (treated in the simple model by using laboratory values of alpha and deuteron masses), and differences in the sizes of LWA2 and LWAb terms.
We have also calculated thermally-averaged reaction rates suitable for use in astrophysical calculations. We integrate from 0.1 keV to 10 MeV at temperatures from $`10^7`$ K to $`3\times 10^{10}`$ K to produce the quantity $`N_A\sigma v`$ customarily used in reaction network calculations. Our result, shown in Fig. 10, is as much as 40% larger than previous estimates over a wide range of temperatures. The sources of these differences are not clear. They probably arise from several specific details of the capture models used, which have been normalized to the data using assumptions about the relative roles of the $`E1`$ and $`E2`$ components to produce the rates recommended in compilations.
## V Discussion
The total cross sections we have computed compare favorably with the data in the region from the $`3^+`$ resonance to 3 MeV. This reflects mainly a combination of reasonable fits of the scattering states to experimental phase shifts, and reasonable reproduction of the asymptotic behavior of the <sup>6</sup>Li ground state in the $`S`$-wave $`\alpha d`$ channel. Our calculation of the $`E1`$ cross section from the relativistic center-of-energy correction, and examination of several smaller corrections, is new; however, like all other models not actually scaled to the laboratory data, it is higher than the data. In our calculation, the $`E1`$ transition contributes just under 70% of the total cross section at 50 keV.
A very large cross section for alpha-deuteron capture would be of some interest for cosmology, since it would allow significant production of $`{}_{}{}^{6}\mathrm{Li}`$ in the big bang. However, we find that big-bang nucleosynthesis calculations using our new cross sections can produce $`{}_{}{}^{6}\mathrm{Li}`$ at a maximum level of 0.000 60 relative to <sup>7</sup>Li, or $`{}_{}{}^{6}\mathrm{Li}`$/H=$`5\times 10^{14}`$. The maximum occurs at a baryon/photon ratio of $`2\times 10^{10}`$, at the low end of the range of possible values suggested by other light-element abundances. At the higher baryon/photon ratio of $`5\times 10^{10}`$ suggested by recent extragalactic deuterium abundance measurements, there is a factor of three less $`{}_{}{}^{6}\mathrm{Li}`$, but because of the increasing <sup>7</sup>Li production, $`{}_{}{}^{6}\mathrm{Li}`$/<sup>7</sup>Li = 0.000 07. Since the abundance ratio found in low-metallicity halo stars so far is in the vicinity of $`{}_{}{}^{6}\mathrm{Li}`$/<sup>7</sup>Li = 0.05, it seems very unlikely that standard big-bang nucleosynthesis could account for observed abundances. It should be kept in mind that non-standard models not relying solely on the alpha-deuteron capture mechanism could, in principle, produce interesting amounts of $`{}_{}{}^{6}\mathrm{Li}`$ .
In the future we expect to make several improvements on the present work. While the VMC wave functions for the s-shell nuclei give energies within 2% of the experimental value, the trial functions for p-shell nuclei are not as good. It should be worthwhile to try the more accurate GFMC ground states for the p-shell nuclei and calculate mixed estimates of the transition operators . Eventually one would also like to construct the scattering states with the GFMC method, but this will be considerably more difficult. Further, new three-body potentials that give a better fit to the energies of the light p-shell nuclei are currently under development and should improve both the VMC and GFMC results. An important aspect of these studies will continue to be the construction of ground states that have the proper asymptotic cluster behavior. This must be imposed in the variational trial function, because the GFMC algorithm is not sensitive to the tail of the wave function and will not be able to correct for any failures.
The present calculation should serve as a useful starting point for studies of several other reactions of astrophysical interest involving p-shell nuclei, including $`{}_{}{}^{4}\mathrm{He}(t,\gamma )^7\mathrm{Li}`$, $`{}_{}{}^{4}\mathrm{He}(^3\mathrm{He},\gamma )^7\mathrm{Be}`$, and $`{}_{}{}^{7}\mathrm{Be}(p,\gamma )^8\mathrm{B}`$. Work on these reactions is now in progress.
###### Acknowledgements.
The authors wish to thank V.R. Pandharipande and S.C. Pieper for many useful comments. Computations were performed on the IBM SP of the Mathematics and Computer Science Division, Argonne National Laboratory. The work of KMN and RBW is supported by the U. S. Department of Energy, Nuclear Physics Division, under contract No. W-31-109-ENG-38, and that of RS by the U. S. Department of Energy under contract No. DE-AC05-84ER40150.
|
warning/0006/astro-ph0006367.html
|
ar5iv
|
text
|
# Reionization of the Universe and the Photoevaporation of Cosmological Minihalos
## 0.1 Ionization Fronts in the IGM
The neutral, opaque IGM out of which the first bound objects condensed was dramatically reheated and reionized at some time between a redshift $`z50`$ and $`z5`$ by the radiation released by some of these objects. When the first sources turned on, they ionized their surroundings by propagating weak, R-type ionization fronts which moved outward supersonically with respect to both the neutral gas ahead of and the ionized gas behind the front, racing ahead of the hydrodynamical response of the IGM, as first described by Shapiro (1986) and Shapiro & Giroux (1987). These authors solved the problem of the time-varying radius of a spherical I-front which surrounds isolated sources in a cosmologically-expanding IGM analytically, taking proper account of the I-front jump condition generalized to cosmological conditions. They applied these solutions to determine when the I-fronts surrounding isolated sources would grow to overlap and, thereby, complete the reionization of the universe. The effect of density inhomogeneity on the rate of I-front propagation was described by a mean “clumping factor” $`c_l>1`$, which slowed the I-fronts by increasing the average recombination rate per H atom inside clumps. This suffices to describe the rate of I-front propagation as long as the clumps are either not self-shielding or, if so, only absorb a fraction of the ionizing photons emitted by the central source. Numerical radiative transfer methods are currently under development to solve this problem in 3D for the inhomogeneous density distribution which arises as cosmic structure forms, so far limited to a fixed density field without gas dynamics (e.g. Abel, Norman, & Madau 1999; Razoumov & Scott 1999; Ciardi et al. 2000). A different, more approximate approach which is intended to mimic the average rate at which I-fronts expanded and overlapped during reionization within the context of cosmological gas dynamics simulation has also been developed (Gnedin 2000). The question of what dynamical effect the I-front had on the density inhomogeneities it encountered, however, requires further analysis. Here we shall briefly summarize our results of the first radiation-hydrodynamical simulations of the back-reaction of a cosmological I-front on a gravitationally-bound density inhomogeneity it encounters – a dwarf galaxy minihalo – during reionization.
## 0.2 The Photoevaporation of Dwarf Galaxy Minihalos Overtaken by a Cosmological Ionization Front
We have performed radiation-hydrodynamical simulations of the photoevaporation of a cosmological minihalo overrun by a weak, R-type I-front in the surrounding IGM, created by an external source of ionizing radiation (Shapiro and Raga 2000a,b). Our simulations in 2D, axisymmetry used an Eulerian hydro code with Adaptive Mesh Refinement and the Van Leer flux-splitting algorithm, which solved nonequilibrium ionization rate equations (for H, He, C, N, O, Ne, and S) and included an explicit treatment of radiative transfer by taking into account the bound-free opacity of H and He. A possible heavy element abundance of $`10^3`$ times solar was assumed, as well.
Here we compare some of those results for two different sources: a quasar-like source with emission spectrum $`F_\nu \nu ^{\mathit{1.8}}`$ ($`\nu >\nu _H`$) and a stellar source with a 50,000 K blackbody spectrum, with luminosity and distance adjusted to keep the ionizing photon fluxes the same in the two cases. In particular, if $`r_{\mathrm{Mpc}}`$ is the distance (in Mpc) between source and minihalo and $`N_{\mathrm{ph},56}`$ is the H-ionizing photon luminosity (in units of $`10^{56}s^1`$), then the flux at the location of the minihalo would, if unattenuated, correspond to $`N_{\mathrm{ph},56}/r_{\mathrm{Mpc}}^2=1`$.
Our initial condition before ionization, shown in Figure 1, is that of a $`10^7M_{}`$ minihalo in an Einstein-de Sitter universe ($`\mathrm{\Omega }_{\mathrm{CDM}}=1\mathrm{\Omega }_{\mathrm{bary}}`$; $`\mathrm{\Omega }_{\mathrm{bary}}h^2=0.02`$; $`h=0.7`$) which collapses out and virializes at $`z_{\mathrm{coll}}=9`$, yielding a truncated, nonsingular isothermal sphere of radius $`R_c=0.5\mathrm{kpc}`$ in hydrostatic equilibrium with virial temperature $`T_{\mathrm{vir}}=5900\mathrm{K}`$ and dark-matter velocity dispersion $`\sigma _V=6.3\mathrm{km}\mathrm{s}^1`$, according to the solution of Shapiro, Iliev, & Raga (1999), for which the finite central density inside a radius about 1/30 of the total size of the sphere is 514 times the surface density. This hydrostatic core of radius $`R_c`$ is embedded in a self-similar, spherical, cosmological infall according to Bertschinger (1985).
The results of our simulations on an $`(r,x)`$-grid with $`256\times 512`$ cells (fully refined), summarized in Figures 2–6, include the following points:
* The background IGM and infalling gas outside the minihalo \[centered at $`(r,x)=(0,7.125\times 10^{21}\mathrm{cm})`$\] are quickly ionized, and the resulting pressure gradient in the infall region converts the infall into an outflow.
* As expected, the hydrostatic core of the minihalo shields itself against ionizing photons, trapping the I-front which enters the halo, causing it to decelerate inside the halo to the sound speed of the ionized gas before it can exit the other side, thereby transforming itself into a weak, D-type front preceded by a shock.
* The side facing the source expels a supersonic wind backwards towards the source, which shocks the IGM outside the minihalo, while the remaining neutral halo material is accelerated away from the source by the so-called “rocket effect” as the halo photoevaporates (cf. Spitzer 1978). Since this is a case of gas bound to a dark halo with $`\sigma <10\mathrm{km}\mathrm{s}^1`$, this photoevaporation proceeds unimpeded by gravity.
* Figures 2(a) and (b) show the position of the I-front inside the minihalo as it slows from weak, R-type to weak D-type led by a shock as it advances across the original hydrostatic core, for the two cases. Figures 2(a) and (b) also show the mass of the neutral zone within the original hydrostatic core shrinking as the minihalo photoevaporates within about 100 Myrs. The photoevaporation time is 50 % larger for the stellar source than for the quasar source.
* Figures 3 and 4 show the structure of the photoevaporative flow 50 Myrs after the global I-front first overtakes the minihalo, with key features of the flow indicated by the labels on the temperature plot in Figure 4. For the stellar case, a strong shock labelled “4S” clearly leads the D-type I-front (labelled “4I”) as it advances through the minihalo core, by contrast with the quasar case in which hard photons penetrate deeper into the neutral gas and preheat it, thereby weakening the shock which leads the I-front. The softer stellar spectrum also explains why helium on the ionized side of the I-front is mostly He II, rather than He III as in the quasar case, while the neutral side is completely He I, rather than a mix of He I and II as in that case.
* Figure 5 shows the spatial variation of the relative abundances of C, N, O ions along the symmetry axis after 50 Myrs. While the quasar case shows the presence at 50 Myrs of low as well as high ionization stages for the metals, the softer spectrum of the stellar case yields less highly ionized gas on the ionized side of the I-front (e.g. mostly C III, N III, O III) and the neutral side as well (e.g. C II, N I, O I and II).
* The column densities of H I, He I and II, and C IV for minihalo gas of different velocities as seen along the symmetry axis at different times are shown in Figure 6. At early times, the minihalo gas resembles a weak Damped Lyman $`\alpha `$ (“DLA”) absorber with small velocity width ($`10\mathrm{km}\mathrm{s}^1`$) and $`N_{\mathrm{H}\mathrm{I}}10^{20}\mathrm{cm}^2`$, with a Lyman-$`\alpha `$-Forest(“LF”)-like red wing ($`\text{velocity width}10\mathrm{km}\mathrm{s}^1`$) with $`N_{\mathrm{H}\mathrm{I}}10^{16}\mathrm{cm}^2`$ on the side moving toward the source, with a He I profile which mimics that of H I but with $`N_{\mathrm{He}\mathrm{I}}/N_{\mathrm{H}\mathrm{I}}[\mathrm{He}]/[\mathrm{H}]`$, and with a weak C IV feature with $`N_{\mathrm{C}\mathrm{IV}}10^{11}(10^{12})\mathrm{cm}^2`$ for the stellar (quasar) cases, respectively, displaced in this same asymmetric way from the velocity of peak H I column density. For He II at early times, the stellar case has $`N_{\mathrm{He}\mathrm{II}}10^{18}\mathrm{cm}^2`$ shifted by 10’s of $`\mathrm{km}/\mathrm{sec}`$ to the red of the H I peak, while for the quasar case, He II simply follows the H I profile, except that $`N_{\mathrm{He}\mathrm{II}}/N_{\mathrm{H}\mathrm{I}}10`$ in the red wing but $`N_{\mathrm{He}\mathrm{II}}/N_{\mathrm{H}\mathrm{I}}10^2`$ in the central H I feature. After 160 Myr, however, only a narrow H I feature with LF-like column density $`N_{\mathrm{H}\mathrm{I}}10^{13}(10^{14})\mathrm{cm}^2`$ remains, with $`N_{\mathrm{He}\mathrm{I}}/N_{\mathrm{H}\mathrm{I}}1/4(<1/10)`$, $`N_{\mathrm{He}\mathrm{II}}/N_{\mathrm{H}\mathrm{I}}10^3(10^2)`$, and $`N_{\mathrm{C}\mathrm{IV}}/N_{\mathrm{H}\mathrm{I}}3[\mathrm{C}]/[\mathrm{C}]_{}([\mathrm{C}]/[\mathrm{C}]_{})`$ for the stellar (quasar) cases, respectively.
* Observations of the absorption spectra of high redshift sources like those which reionized the universe may reveal the presence of photoevaporative flows like these and provide a useful diagnostic of the reionization process.
* Future work will extend this study to minihalos of higher virial temperatures, for which gravity competes more effectively with photoevaporation.
This work was supported by grants NASA NAG5-2785, NAG5-7363, and NAG5-7821, NSF ASC-9504046, and Texas ARP 3658-0624-1999, and a 1997 CONACyT National Chair of Excellence at UNAM for PRS.
|
warning/0006/hep-th0006109.html
|
ar5iv
|
text
|
# Gauge transformations as local space-time transformations
### Gauge transformations as local space-time transformations
The geometrical framework of General Relativity grounds on two basic assumption inferred from experience: Lorentz Invariance, the physical equivalence of all space-time reference frames related by Lorentz transformations, and the Principle of Equivalence, asserting the possibility of eliminating the effects of gravitation in an infinitesimal neighborhood of every point by an appropriate choice of coordinates. The former fixes the rigid properties of space-time in the absence of gravitational phenomena; it determines the local gauge group $`SO(1,3)`$ or, equivalently, the Minkowskian metric $`\eta _{\alpha \beta }`$. The latter identifies gravitational phenomena with a non trivial parallel transport of space-time reference frames forcing us to relax the rigidity hypothesis and fixing the local space-time structure in terms of $`SO(1,3)`$ pseudo-Riemannian geometry. Technically one proceeds by promoting the Minkowskian metric to a point dependent metric $`\eta _{\alpha \beta }g_{\mu \nu }(x)`$ satisfying the maximal compatibility condition $`_\kappa g_{\mu \nu }=0`$. Lorentz Invariance and the Principle of Equivalence allows us to understand gravitational phenomena in pure terms of space-time geometry. How to generalize these basic assumptions to take into account non-gravitational interactions? If we insist with a geometrical picture we should focus on geometrical properties. It is well known that the only geometrical property shared by all fundamental interactions is gauge invariance. This basic fact is at the heart of most attempts of breaking through toward a unified picture. However, the emphasis always has been on gauge invariance alone or geometry alone rather than on gauge invariance as a local geometric property of space-time. There is a deep asymmetry in nowadays comprehension of gauge invariance associated to gravitational and non-gravitational interactions. On one side we have a clear understanding of the local $`SO(1,3)`$ gravitational gauge invariance in terms of observer’s freedom of choosing at will an orthonormal reference frame in every space-time point. To the side of non-gravitational interactions, gauge invariance enters in a rather technical way, guarantying renormalizability and hence perturbative consistence of the quantum theory. Nonetheless, the physical meaning of the ‘internal’ $`U(1)\times SU(2)\times SU(3)`$ gauge invariance associated to electromagnetic, weak and strong forces remains obscure. Attempts of shedding light on this point date back to the work of Weyl, Kaluza and Klein in the twenties reaching present day speculations on supergravity and strings. It is curious to note that even though most of these theories introduce extra dimensions besides the four familiar ones, no serious attempts have been made to assimilate ‘internal’ gauge invariance with the local freedom associated to the choice of an orthonormal reference frame in extra dimensions –simply in line with gravity. While gravitational gauge invariance is a robust (insensitive to continuous deformations) local property of space-time, ‘internal’ gauge invariance is in general assimilated with fragile global properties of particular highly symmetrical solutions. We do find such a deep asymmetry in the comprehension of the only geometrical property surviving in our effective description of fundamental interactions rather unattractive. If we do accept as a working hypothesis the introduction of extra dimensions we should keep in mind that we are already adding to the theory an extra gauge freedom completely analogous to the gravitational gauge freedom. Why not identifying this freedom with non-gravitational gauge freedom?
In support of this elementary geometrical argument there is a well known physical fact. Besides the generic consideration that all fundamental interactions share gauge invariance, it is a very remarkable experimental fact that gauge groups realized in nature are either pseudo-orthogonal or unitary. Orthogonal and unitary groups are intimately connected, playing the identical role of tangent space structural groups in real and complex geometry respectively. This fact indicates us clearly the way to take. The ‘internal’ gauge group $`U(1)\times SU(2)\times SU(3)`$ should simply appear on the side of the gravitational gauge group $`SO(1,3)`$ as –possibly part of– a larger space-time structural group<sup>1</sup><sup>1</sup>1The possibility of constructing differential geometries with generic structural groups was considered by S. Weinberg in a Kaluza-Klein context . The geometry we consider in this paper is a particular case of Weinberg’s generalization. However, it will be excessive to consider our geometrical structure as a generalized differential geometry; it is rather an “interpolation” between real and complex classical differential geometries.. In different words, we surrender to the experimental evidence that the local gauge group associated to fundamental interactions is $`SO(1,3)\times U(1)\times SU(2)\times SU(3)`$ and we enforce this at a space-time level by extending the request of space-time Lorentz Invariance to the one of space-time General Gauge Invariance. This amounts to introduce ten extra space-time dimensions of a somehow ‘complex’ nature besides the four familiar real ones. In the absence of interactions the real and complex nature of different space-time directions is characterized by an appropriate rigid metric structure generalizing the Minkowskian structure. By assuming the validity of a General Principle of Equivalence for all fundamental interactions we are then led to relax the rigidity conditions identifying all fundamental interactions with a non-trivial parallel transport of space-time reference frames. Technically this is obtained again by promoting the rigid space-time metric to a point dependent metric satisfying a maximal compatibility condition. General Gauge Invariance and the General Principle of Equivalence bring us to a unified description of fundamental interactions. This is not only in the mere technical sense that we succeed in recasting all fundamental force fields in a single geometrical object but on the deeper physical and philosophical ground that all interactions appear as identical operative consequences of a non trivial parallel transport of space-time reference frames. In the first part of this paper we reconstruct the local geometrical structure of space-time out of these two basic assumption.
### On the four dimensional nature of space-time and extra dimensions
The generalization of a fundamental concept like the one of space-time is clearly an extremely delicate task that we have the right to perform only if bringing a concrete improvement in our understanding of Nature. Therefore, in introducing extra space-time dimensions and a particular hypothesis on their nature, we have to immediately face two urgent questions. First of all, we do have to explain why our direct perception of space-time is limited to four dimensions. Secondly, but of a major importance, we do have to display an unmistakable trace left by extra dimensions in the effective four dimensional physics. The traditional approach to the first point is that of introducing additional hypothesis on the topological nature of extra dimensions –hence on the global structure of space-time– assuming them to dynamically compactify on extremely small length scales. Basic experimental evidence on chiral fermions is apparently against this hypothesis in pure metrical theories . In any case there is no answer to the second point in the traditional approach. It is instead an intrinsic property of the local geometrical structure we have been lead to –given essentially by the fact that the complex part of the space-time connection couples to matter like a magnetic field– that the free motion of matter in a background gets automatically squeezed on a real $`1+3`$ effective space-time. This gives a natural answer to the question of dimensional reduction without the introduction of any subsidiary hypothesis. In addition the trace left by extra dimensions in the effective four dimensional physics appears unmistakably in the structure of elementary matter fields.
### The fermionic sector of the Standard Model of Elementary Particles
The structure of elementary matter fields before electro-weak symmetry breaking is a basic building block of the Standard Model of Elementary Particles. As extracted from experience the fundamental fermionic representation appears complicated and redundant. Complicated because the basic pattern –a family– consist of five irreducible representations of $`U(1)\times SU(2)\times SU(3)`$ carrying a definite chirality: a right-handed isospin singlet $`𝖨=\mathrm{𝟏}`$, color singlet $`𝖢=\mathrm{𝟏}`$, of hypercharge $`𝖸=1`$ identified with the right-handed electron $`e_R`$; a left-handed isospin doublet $`𝖨=\mathrm{𝟐}`$, color singlet $`𝖢=\mathrm{𝟏}`$, of hypercharge $`𝖸=1/2`$ which components correspond to the left-handed electronic neutrino $`\nu _{eL}`$ and electron $`e_L`$; two right-handed isospin singlet $`𝖨=\mathrm{𝟏}`$, color triplets $`𝖢=\mathrm{𝟑}`$, of hypercharge $`𝖸=1/3`$ and $`𝖸=2/3`$ identified with the right-handed quarks down $`d_R`$ and up $`u_R`$ respectively; and a left-handed isospin doublet $`𝖨=\mathrm{𝟐}`$, color triplets $`𝖢=\mathrm{𝟑}`$, of hypercharge $`𝖸=1/6`$ corresponding to the left-handed components of the quarks down $`d_L`$ and up $`u_L`$. In a more synthetic notation we can write these fifteen elementary matter fields as
$$\begin{array}{cccc}e_R=(\mathrm{𝟏},\mathrm{𝟏})_{+1}^+\hfill & \left(\begin{array}{c}\hfill \nu _{eL}\\ \hfill e_L\end{array}\right)=(\mathrm{𝟐},\mathrm{𝟏})_{+1/2}^{}\hfill & \begin{array}{c}u_R=(\mathrm{𝟏},\mathrm{𝟑})_{2/3}^+\hfill \\ d_R=(\mathrm{𝟏},\mathrm{𝟑})_{+1/3}^+\hfill \end{array}\hfill & \left(\begin{array}{c}\hfill u_L\\ \hfill d_L\end{array}\right)=(\mathrm{𝟐},\mathrm{𝟑})_{1/6}^{}\hfill \end{array}$$
The fundamental representation is redundant because this basic pattern is repeated –at least– three times in the generations corresponding to the leptons $`\mu `$, $`\nu _\mu `$ and the quarks charm $`c`$ and strange $`s`$
$$\begin{array}{cccc}\mu _R=(\mathrm{𝟏},\mathrm{𝟏})_{+1}^+\hfill & \left(\begin{array}{c}\hfill \nu _{\mu L}\\ \hfill \mu _L\end{array}\right)=(\mathrm{𝟐},\mathrm{𝟏})_{+1/2}^{}\hfill & \begin{array}{c}c_R=(\mathrm{𝟏},\mathrm{𝟑})_{2/3}^+\hfill \\ s_R=(\mathrm{𝟏},\mathrm{𝟑})_{+1/3}^+\hfill \end{array}\hfill & \left(\begin{array}{c}\hfill c_L\\ \hfill s_L\end{array}\right)=(\mathrm{𝟐},\mathrm{𝟑})_{1/6}^{}\hfill \end{array}$$
and to the leptons $`\tau `$, $`\nu _\tau `$ and the quarks top $`t`$ and bottom $`b`$
$$\begin{array}{cccc}\tau _R=(\mathrm{𝟏},\mathrm{𝟏})_{+1}^+\hfill & \left(\begin{array}{c}\hfill \nu _{\tau L}\\ \hfill \tau _L\end{array}\right)=(\mathrm{𝟐},\mathrm{𝟏})_{+1/2}^{}\hfill & \begin{array}{c}t_R=(\mathrm{𝟏},\mathrm{𝟑})_{2/3}^+\hfill \\ b_R=(\mathrm{𝟏},\mathrm{𝟑})_{+1/3}^+\hfill \end{array}\hfill & \left(\begin{array}{c}\hfill t_L\\ \hfill b_L\end{array}\right)=(\mathrm{𝟐},\mathrm{𝟑})_{1/6}^{}\hfill \end{array}$$
As a further assumption the Standard Model of Elementary Particles prescribes elementary matter fields to minimally interact with the gauge potentials $`𝖠_\mu ^y`$, $`𝖠_\mu ^{\iota 𝗂}`$, $`𝗂=1,2,3`$, and $`𝖠_\mu ^{\kappa 𝖼}`$, $`𝖼=1,2,\mathrm{},8`$, describing hyper $`U(1)`$, weak $`SU(2)`$ and strong $`SU(3)`$ interactions. In a standard notation matter field equations read
$$\gamma ^\mu (i_\mu 𝖸_𝗉𝖠_\mu ^y𝖠^{\iota 𝗂}𝖨_𝗉^𝗂𝖠^{\kappa 𝖼}𝖢_𝗉^𝖼)\psi _𝗉=0$$
where $`\psi _𝗉`$ is any of the particle listed above, $`𝖸_𝗉`$ the relative hypercharge and $`𝖨_𝗉^𝗂`$, $`𝖢_𝗉^𝖼`$ the relative isospin and color representations (multiplet and spin indices are understood).
Some light on the single family structure is shed by the observation that –while $`U(1)\times SU(2)\times SU(3)`$ naturally embeds in $`SU(5)`$ and this is $`SO(10)`$– the left-handed fermionic representation $`(\mathrm{𝟏},\mathrm{𝟏})_1^{}(\mathrm{𝟐},\mathrm{𝟏})_{+1/2}^{}(\mathrm{𝟏},\overline{\mathrm{𝟑}})_{+2/3}^{}(\mathrm{𝟏},\overline{\mathrm{𝟑}})_{1/3}^{}(\mathrm{𝟐},\mathrm{𝟑})_{1/6}^{}`$ matches the simpler $`SU(5)`$ representation $`\overline{\mathrm{𝟓}}^{}+\mathrm{𝟏}0^{}`$; if we then accept to include as a further fundamental fermion a highly non-interactive right-handed neutrino $`\nu _R=(\mathrm{𝟏},\mathrm{𝟏})_{0}^+`$ the total left-handed fermionic representation recasts in to the single $`SO(10)`$ representation $`\mathrm{𝟏}6^{}`$. Standard gauge theories based on $`SU(5)`$ or $`SO(10)`$ –not including gravity– pay the simplification of gauge group and fundamental fermionic representation at the very high price of introducing several unobserved vector mesons other than a whole army of Higgs bosons necessary to break down the symmetry to the observed one. In addition they leave unsolved the problem of family replication and give no contribution to our understanding of ‘internal’ gauge invariance. From our viewpoint, we observe that the natural embedding of gauge group and matter representation in $`SO(10)`$ is perfectly compatible with –and perhaps a further indication of– the introduction of ten extra dimensions.
In the second and third part of this paper we analyze the free motion of a fermion in a given background. The fermionic sector of the Standard Model of Elementary Particles emerges as the low energy limit of the theory.
## 1 Space-time structure
The geometrical structure of General Relativity is that of $`SO(1,3)`$ pseudo-Riemannian geometry. Space-time is described by a real manifold. Gravitational forces by a metric connection with tangent space structure
$$\eta _{\alpha \beta }=\left(\begin{array}{cccc}1& & & \\ & 1& & \\ & & 1& \\ & & & 1\end{array}\right)$$
The gravitational field can be identified with different geometrical objects: the metric tensor $`g_{\mu \nu }`$, the affine connection $`\mathrm{\Gamma }_{\mu \nu }^\rho `$ or the spin connection $`\mathrm{\Omega }_{\kappa ,\alpha \beta }`$. The vanishing of space-time torsion –as required by the Principle of Equivalence– makes the three structure to carry the same amount of geometrical information. The gauge character of gravitational interactions is most easily displayed in terms of spin connection. A fundamental notion in the theory is that of reference frame. A reference frame at a point $`x`$ is a set of three orthonormal space-like vectors $`e_1^\mu |_x,e_2^\mu |_x,e_3^\mu |_x`$ plus a clock. Geometrically the clock is interpreted as a time-like unit vector $`e_0^\mu |_x`$ which is conveniently chosen orthogonal to the first three. An assignation of a reference frame in every space-time point is therefore a set of four vector fields satisfying
$$e_\alpha ^\mu e_\beta ^\nu g_{\mu \nu }=\eta _{\alpha \beta }$$
We see that a reference frame is determined up to a point dependent Lorentz pseudo-rotation,
$$e_\alpha ^\mu \mathrm{\Lambda }_\alpha ^\beta (x)e_\beta ^\mu $$
where in every point $`\mathrm{\Lambda }_\alpha ^\gamma \mathrm{\Lambda }_\beta ^\delta \eta _{\gamma \delta }=\eta _{\alpha \beta }`$. In order to compare physical phenomena in different space-time points we are lead to construct the (spin) connection
$$\mathrm{\Omega }_{\kappa ,\alpha \beta }=e_\alpha ^\mu \left(_\kappa e_\beta ^\nu \right)g_{\mu \nu }$$
where $``$ denotes the covariant derivative induced by the metric $`g_{\mu \nu }`$ (on curved indices only). The connection is clearly anti-symmetric in the indices $`\alpha `$ and $`\beta `$, belonging therefore to the Lie algebra $`so(1,3)`$. Under a local $`SO(1,3)`$ transformation of reference frames
$$\mathrm{\Omega }_{\kappa ,\alpha \beta }\mathrm{\Lambda }_\alpha ^\gamma \mathrm{\Lambda }_\beta ^\delta \mathrm{\Omega }_{\kappa ,\gamma \delta }+\mathrm{\Lambda }_\alpha ^\gamma \left(_\kappa \mathrm{\Lambda }_\beta ^\delta \right)\eta _{\gamma \delta }$$
which clearly display the gauge nature of the gravitational field. The freedom of choosing at will a reference frame in every space-time point is the deep physical meaning of the $`SO(1,3)`$ gauge invariance of gravitational forces.
We are willing to enforce the whole gauge group $`SO(1,3)\times U(1)\times SU(2)\times SU(3)`$ associated to fundamental interactions at a space-time level. This requires the introduction of ten complex dimensions besides the four familiar real ones. Therefore, it is convenient to start by briefly recalling the main feature of complex geometry.
There are essentially two equivalent approaches to the definition of a complex manifold. The first one is repeating the definition of real manifold in terms of local coordinates and transformation functions, with real vector spaces replaced by complex ones and real smooth functions by homomorphic functions. The second one –that better adapts to our task– is starting from a real manifold and introducing a complex structure on it. A complex structure $`I_i^j`$ is a tensor field acting on the tangent space at every point of the manifold as the multiplication by the imaginary unit $`i`$. More explicitly $`I_i^kI_k^j=\delta _i^j`$ plus an integrability condition necessary to guarantee that imaginary units defined in different points fit together consistently. A complex metric connection on a complex manifold is introduced through a Riemannian metric $`g_{ij}`$ preserving the complex structure, $`g_{ij}=I_i^kI_j^lg_{kl}`$. To $`g_{ij}`$ is then associated an antisymmetric two-form $`\omega _{ij}=I_i^kg_{kj}`$. Metric and antisymmetric two-form completely characterize the geometry and are simply recasted in real and imaginary part of a single Hermitian metric $`h_{ij}=g_{ij}+i\omega _{ij}`$. Observe that to such a geometry is in general associated a non-vanishing torsion proportional to the exterior derivative of $`\omega _{ij}`$. Further requiring the vanishing of torsion makes metric and skew-symmetric two-form maximally compatible, $`_k\omega _{ij}0`$. Such a geometry plays the role of Riemannian geometry in the complex realm and is known as Kählerian geometry. On the tangent space at every point we always can choose coordinates in such a way that $`g_{ij}`$ and $`\omega _{ij}`$ take the standard form
$$\delta _{ab}=\left(\begin{array}{ccc}1& 0& \\ 0& 1& \\ & & \mathrm{}\end{array}\right)\epsilon _{ab}=\left(\begin{array}{ccc}0& 1& \\ 1& 0& \\ & & \mathrm{}\end{array}\right)$$
where dots indicate a certain number of blocks of the same type. These matrices specify the tangent space structure of the geometry, playing the very same role of the Minkowskian metric in pseudo-Riemannian geometry. We are now ready to go back to the notion of reference frame. An assignation of a reference frame in every point of a Kählerian manifold is a maximal set of independent vector fields $`e_a^i`$ fulfilling the conditions
$$e_a^ie_b^jg_{ij}=\delta _{ab}e_a^ie_b^j\omega _{ij}=\epsilon _{ab}$$
Again, we see that a reference frame is determined up to transformations
$$e_a^iU_a^b(x)e_b^i$$
that satisfies in every point $`U_a^cU_b^d\delta _{cd}=\delta _{ab}`$ and $`U_a^cU_b^d\epsilon _{cd}=\epsilon _{ab}`$. The first condition ensures $`U_a^b`$ to be an orthogonal transformation, the second a symplectic one. The intersection of orthogonal and symplectic groups is the unitary group $`U(n)=SO(2n)Sp(2n)`$. On a Kählerian manifold a reference frame is therefore determined up to an unitary transformation. Comparing informations in different space-time points involves again the connection
$$\mathrm{\Omega }_{k,ab}=e_a^i\left(_ke_b^j\right)g_{ij}$$
The requirement of a vanishing torsion makes $`\mathrm{\Omega }_{k,ab}`$ to belong to the Lie algebra $`u(n)`$. Under a local unitary transformation $`\mathrm{\Omega }_{k,ab}`$ transform as a $`U(n)`$ gauge connection
$$\mathrm{\Omega }_{k,ab}U_a^cU_b^d\mathrm{\Omega }_{k,cd}+U_a^c\left(_kU_b^d\right)\delta _{cd}$$
However, $`\mathrm{\Omega }_{k,ab}`$ can not be associated to a non-gravitational gauge field. The indices $`k`$, $`a`$ and $`b`$ have the same transformation properties. Somehow we need a structure supporting objects having both real space-time and complex gauge indices.
We have all the necessary ingredients to build up such a generalization. In analogy with real Riemannian and complex Kählerian geometries we start with a space-time described by a real manifold. Enforcing General Gauge Invariance we introduce a tangent space structure having as isometries the gauge group of fundamental interactions. In accordance with the General Principle of Equivalence we then construct the local geometrical framework based on this structure. Fundamental interactions will be identified with appropriate components of the total connection.
### General Gauge Invariance and tangent space structure
We may assume a generic tangent space structure to be given by a Hermitian matrix $`\widehat{\eta }_{AB}+i\widehat{\epsilon }_{AB}`$. The real symmetric part $`\widehat{\eta }_{AB}`$ always can be chosen diagonal with all eigenvalues $`\pm 1`$. The imaginary anti-symmetric part $`\widehat{\epsilon }_{AB}`$ always can be chosen two by two block diagonal. For its eigenvalues we have many different options. When all eigenvalues equal zero we obtain a (pseudo-)Riemannian real structure. When all eigenvalues equal $`\pm i`$ we obtain a (pseudo-)Kählerian complex structure. In the general case we have a group or real (pseudo-)Riemannian directions corresponding to null eigenvalues, plus one or more groups of equivalent complex (pseudo-)Kählerian directions corresponding to possibly degenerate complex eigenvalues. The case that better fits to the group $`SO(1,3)\times U(1)\times SU(2)\times SU(3)`$ is obviously the latter.
General Gauge Invariance brings us to infer that space-time has $`4+10`$ dimensions and –in the absence of fundamental interactions– a rigid metric structure given by $`\widehat{\eta }_{AB}+i\widehat{\epsilon }_{AB}`$, where the real part $`\widehat{\eta }_{AB}`$ is the Minkowskian metric
$$\widehat{\eta }_{AB}=\text{diag}(c^2,1,1,1,1,1,1,1,1,1,1,1,1,1)$$
and the imaginary part $`\widehat{\epsilon }_{AB}`$ is the antisymmetric two-form
$$\widehat{\epsilon }_{AB}=\left(\begin{array}{cccccccccccccc}0& 0& & & & & & & & & & & & \\ 0& 0& & & & & & & & & & & & \\ & & 0& \hfill 0& & & & & & & & & & \\ & & 0& \hfill 0& & & & & & & & & & \\ & & & & & & & & & & & & & \\ & & & & 0& w& & & & & & & & \\ & & & & w& 0& & & & & & & & \\ & & & & & & 0& w& & & & & & \\ & & & & & & w& 0& & & & & & \\ & & & & & & & & & & & & & \\ & & & & & & & & 0\hfill & s& & & & \\ & & & & & & & & s\hfill & 0& & & & \\ & & & & & & & & & & 0& s& & \\ & & & & & & & & & & s& 0& & \\ & & & & & & & & & & & & 0& s\\ & & & & & & & & & & & & s& 0\end{array}\right)$$
$`c`$ is at usual the speed of light and $`w`$, $`s`$ two real positive parameters not fixed by present considerations<sup>2</sup><sup>2</sup>2A word about notation. We have quite a large number of objects and different indices. We refer as ordinary space-time directions the four real directions. All other directions are collectively addressed as extra space-time directions. The ones corresponding to eigenvalues $`\pm iw`$ are called weak directions while the ones corresponding to $`\pm is`$ strong directions. As a rule we put a hat $`\widehat{}`$ over all $`14`$-dimensional objects. This allows us to use the same symbol to denote the same geometrical object in different contexts. Capital Latin letters form $`A`$ run from $`0`$ to $`13`$ and denote space-time flat-indices. Greek letters from $`\alpha `$ denote ordinary space-time flat-indices and take values $`0`$, $`1`$, $`2`$, $`3`$. Latin letters from $`a`$ take values $`4,\mathrm{},13`$ and denote flat-indices in weak and strong directions. We also split this set of indices in $`a_w`$ going from $`4`$ to $`7`$ and $`a_s`$ going from $`8`$ to $`13`$. Furthermore, we find it convenient to introduce the notation $`1_w=4`$, $`2_w=5`$, $`3_w=6`$, $`4_w=7`$ and $`1_s=8`$, …, $`6_s=13`$. The same type of letters from the mid of the alphabet will denote the corresponding curved-indices.. Observe that while real directions may be rescaled by setting –as we do– $`c=1`$, the same is not possible for single complex directions. As an example, rescaling the fourth and fifth coordinates (we count from zero) in such a way that $`\widehat{\epsilon }_{45}=\widehat{\epsilon }_{54}=1`$ will introduce $`w`$ in $`\widehat{\eta }_{44}`$ and $`\widehat{\eta }_{55}`$. In the following we assume $`ws`$, <sup>3</sup><sup>3</sup>3We may equally be tempted to assume all the five complex directions as equivalent, obtaining an attractive grand unified $`U(1)\times SU(5)`$ gauge group. However, we find more economic to break space-time symmetry –it is already broken– at a fundamental level by assuming the existence of different groups of inequivalent complex directions. At the end of the story this will avoid the introduction a large number of unobserved vector mesons and corresponding Higgs scalars necessary to break the group down to the desired symmetry. that is, two groups of inequivalent complex directions.
The role of Lorentz transformations is now taken by the isometries of the local Hermitian structure $`\widehat{\eta }_{AB}+i\widehat{\epsilon }_{AB}`$. In different words we claim the physical equivalence of all space-time reference frames related by transformations $`\widehat{\mathrm{\Lambda }}_A^B`$ fulfilling conditions
$$\widehat{\mathrm{\Lambda }}_A^C\widehat{\mathrm{\Lambda }}_B^D\widehat{\eta }_{CD}=\widehat{\eta }_{AB}\text{and}\widehat{\mathrm{\Lambda }}_A^C\widehat{\mathrm{\Lambda }}_B^D\widehat{\epsilon }_{CD}=\widehat{\epsilon }_{AB}$$
By dividing $`\widehat{\mathrm{\Lambda }}_A^B`$ in appropriate blocks we realize that all non diagonal blocks have to equal zero
$$\widehat{\mathrm{\Lambda }}_A^B=\left(\begin{array}{ccc}\hfill \mathrm{\Lambda }_\alpha ^\beta & & \\ & & \\ & & \\ & W_{a_w}^{b_w}& \\ & & \\ & & \\ & & S_{a_s}^{b_s}\hfill \end{array}\right)$$
while $`\mathrm{\Lambda }_\alpha ^\beta SO(1,3)`$, $`W_{a_w}^{b_w}U(2)`$ and $`S_{a_s}^{b_s}U(3)`$. The isometry group is therefore the direct product $`SO(1,3)\times U(1)\times U(1)\times SU(2)\times SU(3)`$. The extension includes naturally the pattern of local symmetries of fundamental interactions. The only extracharge we have to pay is an additional $`U(1)`$ gauge symmetry.
For later use and also to fix notation it convenient to explicitly write down the conditions fulfilled by infinitesimal transformations. Rewriting $`\widehat{\mathrm{\Lambda }}_A^B=\delta _A^B+\widehat{\lambda }_A^B+\mathrm{}`$, with $`\widehat{\lambda }_A^B`$ taken infinitesimal and using the conditions above we obtain
$$\widehat{\lambda }_A^C\widehat{\eta }_{CB}+\widehat{\lambda }_B^C\widehat{\eta }_{CA}=0\text{and}\widehat{\lambda }_A^C\widehat{\epsilon }_{CB}\widehat{\lambda }_B^C\widehat{\epsilon }_{CA}=0$$
Dividing ordinary space-time, weak and strong indices we find that:
$$\widehat{\lambda }_{\alpha \beta }=\widehat{\lambda }_{\beta \alpha }$$
are the usual infinitesimal generators of ordinary Lorentz transformations;
$$\widehat{\lambda }^x=2w(\widehat{\lambda }_{1_w2_w}+\widehat{\lambda }_{3_w4_w})+2s(\widehat{\lambda }_{1_s2_s}+\widehat{\lambda }_{3_s4_s}+\widehat{\lambda }_{5_s6_s})$$
and
$$\widehat{\lambda }^y=2w(\widehat{\lambda }_{1_w2_w}+\widehat{\lambda }_{3_w4_w})2s(\widehat{\lambda }_{1_s2_s}+\widehat{\lambda }_{3_s4_s}+\widehat{\lambda }_{5_s6_s})$$
generate respectively a $`U(1)`$ rotation of all complex directions and a relative $`U(1)`$ rotation of the two blocks of inequivalent complex directions. These generators of space-time transformations will later be identified with the additional extracharge and with hypercharge;
$$\begin{array}{c}\widehat{\lambda }^{\iota _1}=\frac{1}{2}\left(\widehat{\lambda }_{1_w4_w}\widehat{\lambda }_{2_w3_w}\right)\hfill \\ \widehat{\lambda }^{\iota _2}=\frac{1}{2}\left(\widehat{\lambda }_{1_w3_w}+\widehat{\lambda }_{2_w4_w}\right)\hfill \\ \widehat{\lambda }^{\iota _3}=\frac{1}{2}\left(\widehat{\lambda }_{1_w2_w}\widehat{\lambda }_{3_w4_w}\right)\hfill \end{array}$$
generates space-time $`SU(2)`$ transformations that will appear in the effective four dimensional physics as isospin transformations;
$$\begin{array}{c}\widehat{\lambda }^{\kappa _1}=\frac{1}{2}\left(\widehat{\lambda }_{1_s4_s}\widehat{\lambda }_{2_s3_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _2}=\frac{1}{2}\left(\widehat{\lambda }_{1_s3_s}+\widehat{\lambda }_{2_s4_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _3}=\frac{1}{2}\left(\widehat{\lambda }_{1_s2_s}\widehat{\lambda }_{3_s4_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _4}=\frac{1}{2}\left(\widehat{\lambda }_{1_s6_s}\widehat{\lambda }_{2_s5_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _5}=\frac{1}{2}\left(\widehat{\lambda }_{1_s5_s}+\widehat{\lambda }_{2_s6_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _6}=\frac{1}{2}\left(\widehat{\lambda }_{3_s6_s}\widehat{\lambda }_{4_s5_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _7}=\frac{1}{2}\left(\widehat{\lambda }_{3_s5_s}+\widehat{\lambda }_{4_s6_s}\right)\hfill \\ \widehat{\lambda }^{\kappa _8}=\frac{1}{2\sqrt{3}}\left(\widehat{\lambda }_{1_s2_s}+\widehat{\lambda }_{3_s4_s}2\widehat{\lambda }_{5_s6_s}\right)\hfill \end{array}$$
generates space-time $`SU(3)`$ transformations appearing in four dimensions as color transformations. All remaining independent combinations equal zero.
### The General Principle of Equivalence and local geometry
In complete analogy with the general relativistic picture of gravitational interactions we assume all fundamental interactions to appear as a consequence of a non trivial parallel transport of space-time reference frames. We enforce the General Principle of Equivalence by promoting the metric of the rigid space-time to a point dependent metric $`\widehat{\eta }_{AB}+i\widehat{\epsilon }_{AB}\widehat{g}_{IJ}(\widehat{x})+i\widehat{\omega }_{IJ}(\widehat{x})`$ satisfying the maximal compatibility condition $`\widehat{}_K\widehat{g}_{IJ}=0`$ and $`\widehat{}_K\widehat{\omega }_{IJ}=0`$. An assignation of reference frames in every space-time point is again a set of maximally independent vector fields $`\widehat{e}_A^I`$ fulfilling
$$\widehat{e}_A^I\widehat{e}_B^J\widehat{g}_{IJ}=\widehat{\eta }_{AB}\widehat{e}_A^I\widehat{e}_B^J\widehat{\omega }_{IJ}=\widehat{\epsilon }_{AB}$$
Reference frames are so determined up to a point dependent $`\widehat{\mathrm{\Lambda }}(\widehat{x})`$ transformation
$$\widehat{e}_A^I\widehat{\mathrm{\Lambda }}_A^B(\widehat{x})\widehat{e}_B^I$$
Comparing physical informations in different space-time points involves as usual the connection
$$\widehat{\mathrm{\Omega }}_{K,AB}=\widehat{e}_A^I\left(\widehat{}_K\widehat{e}_B^J\right)\widehat{g}_{IJ}$$
belonging this time to the lie algebra $`so(1,3)u(1)u(1)su(2)su(3)`$ in virtue of the maximal compatibility condition. Under a point dependent redefinition of reference frames $`\widehat{\mathrm{\Omega }}_{K,AB}`$ transforms like an $`SO(1,3)\times U(1)\times U(1)\times SU(2)\times SU(3)`$ gauge potential
$$\widehat{\mathrm{\Omega }}_{K,AB}\widehat{\mathrm{\Lambda }}_A^C\widehat{\mathrm{\Lambda }}_B^D\widehat{\mathrm{\Omega }}_{K,CD}+\widehat{\mathrm{\Lambda }}_A^C\left(_K\widehat{\mathrm{\Lambda }}_B^D\right)\widehat{\eta }_{CD}$$
The components of $`\widehat{\mathrm{\Omega }}_{K,AB}`$ having $`K`$ as an ordinary space-time index and $`A`$, $`B`$ both ranging over ordinary or extra space-time have the right transformation properties to represent gravitational and non-gravitational fundamental gauge fields respectively. To fully appreciate the geometrical meaning of gravitational and non-gravitational interactions we develop a little further the geometrical structure.
Consider first the case in which a gauge exist such that $`\widehat{\mathrm{\Omega }}_{K,AB}0`$. In the absence of forces space-time is flat. It is then possible to find global coordinate frames $`\widehat{x}^A=(x^\alpha ,\xi ^a)`$ such that the metric structure $`\widehat{g}_{IJ}+i\widehat{\omega }_{IJ}`$ takes everywhere the standard form $`\widehat{\eta }_{AB}+i\widehat{\epsilon }_{AB}`$. It is worth stressing that this rigid geometry is very different from a fourteen dimensional Minkowskian geometry. We are not allowed to mix real and complex directions together, nor complex directions of a different nature. As a flat geometry our generalization is a rather trivial one: the total space-time is the direct product of a four dimensional Minkowskian space-time times a complex affine-space of real dimension four times a second complex affine-space of real dimension six. However –as we show in next sections– this rigid structure already has very important phenomenological consequences. It is responsible for the effective four dimensional structure of elementary matter fields. In physical terms it is better to think of space-time in a different way. Picture out the fourteen dimensional flat space-time as foliated in null four dimensional hyper-surfaces of the degenerate skew-symmetric two-form $`\widehat{\epsilon }_{AB}`$ (space-time looks like the ordinary three dimensional Euclidean space foliated in one dimensional straight field lines of an homogeneous magnetic field). On every leave of such a foliation the symmetric two-form $`\widehat{\eta }_{AB}`$ induces an ordinary Minkowskian structure. Leaves are trivially embedded in space-time. Every $`1+3`$ null hyper-surface is a good candidate for ordinary space-time.
When fundamental interactions are turned on it is no longer possible to globally bring the metric structure in the standard form. However, we can still picture out the fourteen dimensional space-time as foliated in null four dimensional hyper-surfaces of the degenerate two-form $`\widehat{\omega }_{IJ}(\widehat{x})`$ (we can maintain the three dimensional analogy by replacing the homogeneous magnetic field with an inhomogeneous one). On every leave of the foliation the symmetric two-form $`\widehat{g}_{IJ}(\widehat{x})`$ induces now an $`SO(1,3)`$ pseudo-Riemannian connection. In the effective four dimensional physics this is identified with gravitation. Every leave is also non trivially embedded in the total space-time. The twist in space-time of $`\widehat{\omega }_{IJ}`$ null hyper-surfaces is the ingredient making the difference between a product manifold and our geometrical structure. The effect of such a twist on the effective four dimensional physics is that of non-gravitational fundamental interactions. To make the content of the theory more transparent we specialize to coordinates adapted to the foliation. Among the various coordinate frames doing the job one choice is particularly convenient. As a consequence of the maximal compatibility condition the two-form $`\widehat{\omega }_{IJ}`$ is closed (enforcing the analogy between $`\widehat{\omega }_{IJ}`$ and a three dimensional magnetic field). Therefore, a classical theorem of Darboux ensures the possibility of globally<sup>4</sup><sup>4</sup>4We assume space-time topology to be the one of $`IR^{14}`$. We also assume the foliation of space-time to be regular ignoring problems connected to global integrability. In a general topology our consideration are valid in a finite neighborhood of a point. finding coordinates in such a way that $`\widehat{\omega }_{IJ}\widehat{\epsilon }_{IJ}`$. Denoting by $`\widehat{x}^I=(x^\mu ,\xi ^i)`$ one of such Darboux coordinates frames we immediately realize the $`x^\mu `$ to parameterize foliation leaves and the $`\xi ^{i_w}`$, $`\xi ^{i_s}`$ weak and strong directions respectively. Every point $`\xi `$ in extra space-time labels a leave of the foliation. In a Darboux coordinate frame we introduce the following parameterization of the Hermitian metric
$$\begin{array}{ccc}\widehat{g}_{IJ}=\left(\begin{array}{cc}g_{\mu \nu }& g_{\mu \lambda }a_j^\lambda \\ a_i^\kappa g_{\kappa \nu }& g_{ij}+a_i^\kappa a_j^\lambda g_{\kappa \lambda }\end{array}\right)& \text{and}& \widehat{\omega }_{IJ}=\left(\begin{array}{cc}0& 0\\ 0& \epsilon _{ij}\end{array}\right)\end{array}$$
$`g_{\mu \nu }(x,\xi )`$ is the pseudo-Riemannian metric induced on the leave labeled by $`\xi `$; $`a_i^\mu (x,\xi )`$ and $`g_{ij}(x,\xi )`$ encode informations on its extrinsic geometry. This parameterization allows to directly relate space-time reference frames $`\widehat{e}_A^I`$ with ordinary $`e_\alpha ^\mu `$ and extra $`e_a^i`$ reference frames
$$\widehat{e}_A^I=\left(\begin{array}{cc}e_\alpha ^\mu & 0\\ e_a^ka_k^\mu & e_a^i\end{array}\right)$$
Substituting in the general conditions we obtain indeed
$$\begin{array}{ccc}e_\alpha ^\mu e_\beta ^\nu g_{\mu \nu }=\eta _{\alpha \beta }\hfill & & \\ e_a^ie_b^jg_{ij}=\delta _{ij}\hfill & \text{and}& e_a^ie_b^j\epsilon _{ij}=\epsilon _{ab}\hfill \end{array}$$
The splitting of ordinary and extra space-time coordinates divides connection components in six groups: $`\widehat{\mathrm{\Omega }}_{\mu ,\alpha \beta }`$, $`\widehat{\mathrm{\Omega }}_{\mu ,\alpha b}`$, $`\widehat{\mathrm{\Omega }}_{\mu ,ab}`$ and $`\widehat{\mathrm{\Omega }}_{i,\alpha \beta }`$, $`\widehat{\mathrm{\Omega }}_{i,\alpha b}`$, $`\widehat{\mathrm{\Omega }}_{i,ab}`$. The first three describe the properties of the parallel transport of space-time reference frames in ordinary directions; they completely characterize the intrinsic and extrinsic geometry of every leave of the foliation as a submanifold of space-time. The last three describe the properties of parallel transport of space-time reference frames in extra directions; they measure the amount of change in intrinsic and extrinsic geometry when moving from a leave to the next. (In the three dimensional analogy the former are the equivalent of intrinsic -trivial- geometry, curvature and torsion of every field line; the latter are the equivalent of vorticity, expansion, shear and similar geometrical properties of the foliation of space in terms of field lines.) If we assume to be constrained to live on a single four dimensional leave of the foliation –and in the next sections we show that this is indeed the case– we are only allowed to transport reference frames in ordinary directions. Therefore, the geometrical space-time properties we can directly perceive in our effective four dimensional physics are the ones described by $`\widehat{\mathrm{\Omega }}_{\mu ,\alpha \beta }`$, $`\widehat{\mathrm{\Omega }}_{\mu ,\alpha b}`$ and $`\widehat{\mathrm{\Omega }}_{\mu ,ab}`$.
$`\widehat{\mathrm{\Omega }}_{\mu ,\alpha \beta }`$ equals the induced connection on every leave
$$\widehat{\mathrm{\Omega }}_{\mu ,\alpha \beta }=\mathrm{\Omega }_{\mu ,\alpha \beta }$$
giving a complete characterization of its intrinsic geometry. The induced connection belongs to $`so(1,3)`$, transforms like a Lorentz gauge potential and is identified with Einstein’s gravitational field.
$`\widehat{\mathrm{\Omega }}_{\mu ,\alpha b}`$ generalizes the notion of curvature of a line embedded in the three dimensional Euclidean space. In a higher dimensional context it is costume to rewrite it in terms of the extrinsic curvature (or second fundamental form) $`\kappa _{\mu \nu a}=e_\mu ^\beta \widehat{\mathrm{\Omega }}_{\nu ,\beta a}`$. In our geometry $`\widehat{\mathrm{\Omega }}_{\mu ,\alpha b}`$ is identically equal to zero as a consequence of the connection transformation properties
$$\widehat{\mathrm{\Omega }}_{\mu ,\alpha b}=e_\alpha ^\nu \kappa _{\mu \nu b}=0$$
Every leave of the foliation is minimally embedded in the total space-time.
$`\widehat{\mathrm{\Omega }}_{\mu ,ab}`$ generalizes the notion of torsion of a line embedded in the three dimensional Euclidean space. This is perhaps the best way of visualizing its geometrical meaning in a higher dimensional context. We address these components of the total connection as the extrinsic torsion of every leave of the foliation. We also introduce the special symbol $`𝒜_{\mu ,ab}\widehat{\mathrm{\Omega }}_{\mu ,ab}`$ to denote it. The extrinsic torsion transforms like a one-form under reparameterization of the leave. It fulfills the conditions $`𝒜_{\mu ,ab}+𝒜_{\mu ,ba}=0`$ and $`𝒜_{\mu ,a}^c\epsilon _{cb}𝒜_{\mu ,b}^c\epsilon _{ca}=0`$ belonging to the Lie algebra $`u(1)u(1)su(2)su(3)`$. Under a point dependent redefinition of extra reference frames $`e_a^iU_a^b(x,\xi )e_b^i`$, $`𝒜_{\mu ,ab}`$ transforms like a $`U(1)\times U(1)\times SU(2)\times SU(3)`$ gauge connection. In the next section we show that $`𝒜_{\mu ,ab}`$ effectively couples to matter field as a gauge connection. Therefore, we identify the extrinsic torsion with non-gravitational fundamental interactions. In terms of the parameterization introduced above
$$𝒜_{\mu ,ab}=\frac{1}{2}\left(e_a^i(_\mu e_b^j)(_\mu e_a^i)e_b^j\right)g_{ij}$$
For later use it is convenient to rearrange the various components of $`𝒜_{\mu ,ab}`$ as follows: the gauge potential associated to extracharge and hypercharge are chosen as $`𝒜_\mu ^x=\frac{1}{2w}(𝒜_{\mu ,1_w2_w}+𝒜_{\mu ,3_w4_w})+\frac{1}{2s}(𝒜_{\mu ,1_s2_s}+𝒜_{\mu ,3_s4_s}+𝒜_{\mu ,5_s6_s})`$ and $`𝒜_\mu ^y=\frac{1}{2w}(𝒜_{\mu ,1_w2_w}+𝒜_{\mu ,3_w4_w})\frac{1}{2s}(𝒜_{\mu ,1_s2_s}+𝒜_{\mu ,3_s4_s}+𝒜_{\mu ,5_s6_s})`$; the ones corresponding to isospin $`𝒜_\mu ^{\iota _1}`$, $`𝒜_\mu ^{\iota _2}`$, $`𝒜_\mu ^{\iota _3}`$ and to color $`𝒜_\mu ^{\kappa _1}`$,$`𝒜_\mu ^{\kappa _2}`$,…, $`𝒜_\mu ^{\kappa _8}`$ are defined according to the notation for infinitesimal generators.
Of the remaining three groups of connection components, the vanishing on symmetry grounds of $`\widehat{\mathrm{\Omega }}_{i,\alpha b}0`$ guarantees the integrability of directions everywhere orthogonal to ordinary space-time hyper-surfaces. Therefore, we can imagine space-time as further foliated in ten dimensional extra hyper-surfaces. $`\widehat{\mathrm{\Omega }}_{i,ab}`$ is then interpreted as the connection induced on every such leave and $`\widehat{\mathrm{\Omega }}_{i,\alpha \beta }`$ as its extrinsic torsion.
The triviality of the parallel transport of space-time reference frames is measured, as usual, by the gauge covariant curvature field strength
$$\widehat{R}_{IJAB}=_I\widehat{\mathrm{\Omega }}_{J,AB}_J\widehat{\mathrm{\Omega }}_{I,AB}+\widehat{\mathrm{\Omega }}_{I,A}^C\widehat{\mathrm{\Omega }}_{J,CB}\widehat{\mathrm{\Omega }}_{I,B}^C\widehat{\mathrm{\Omega }}_{J,CA}$$
The splitting of ordinary and extra space-time coordinates divides field strength components in nine groups. On symmetry grounds only four of these are non vanishing: $`\widehat{R}_{\mu \nu \alpha \beta }`$, $`\widehat{R}_{\mu \nu ab}`$ and $`\widehat{R}_{ij\alpha \beta }`$ $`\widehat{R}_{ijab}`$. The ones measuring the triviality of the parallel transport of reference frames in ordinary directions are $`\widehat{R}_{\mu \nu \alpha \beta }`$ and $`\widehat{R}_{\mu \nu ab}`$
$`\widehat{R}_{\mu \nu \alpha \beta }`$ equals the induced curvature on every leave, that is the gravitational field strength
$$\widehat{R}_{\mu \nu \alpha \beta }=R_{\mu \nu \alpha \beta }=_\mu \mathrm{\Omega }_{\nu ,\alpha \beta }_\nu \mathrm{\Omega }_{\mu ,\alpha \beta }+\mathrm{\Omega }_{\mu ,\alpha }^\gamma \mathrm{\Omega }_{\nu ,\gamma \beta }\mathrm{\Omega }_{\nu ,\alpha }^\gamma \mathrm{\Omega }_{\mu ,\gamma \beta }$$
These equations coincides with the Gauss equations for the embedding of ordinary space-time leaves in the total manifold. $`R_{\mu \nu \rho \sigma }=e_\rho ^\alpha e_\sigma ^\beta R_{\mu \nu \alpha \beta }`$ is the usual Riemann tensor which contraction is used in writing down Einstein’s equations.
$`\widehat{R}_{\mu \nu ab}`$ equals the non-gravitational field strength associated to the gauge potential $`𝒜_{\mu ,ab}`$
$$\widehat{R}_{\mu \nu ab}=_{\mu \nu ab}=_\mu 𝒜_{\nu ,ab}_\nu 𝒜_{\mu ,ab}+𝒜_{\mu ,a}^c𝒜_{\nu ,cb}𝒜_{\mu ,b}^c𝒜_{\nu ,ca}$$
Its definition correspond to the Ricci equations for the embedding of leaves in the total space-time. The field strength of hyper, weak and strong forces as well as the one associated to the additional extracharge are components of the total space-time curvature.
The vanishing on symmetry grounds of $`\widehat{R}_{\mu \nu \alpha b}0`$ is perfectly compatible with the Codazzi-Mainardi equations once we remember that the extrinsic curvature is identically equal to zero. The remaining two groups of non vanishing components of the total curvature tensor express analogue properties for the orthogonal foliation: $`\widehat{R}_{ijab}`$ equals the curvature field strength induced on every ten dimensional extra leave while $`\widehat{R}_{ij\alpha \beta }`$ the curvature associated to the relative extrinsic torsion.
### An extra coupling
As a consequence of the maximal compatibility condition the skew-symmetric two-form $`\widehat{\omega }_{IJ}`$ is closed. This induces a twist on the bundle of complex fields defined on the manifold. In more familiar words the principle that all possible coupling have to be included in field equations makes $`\widehat{\omega }_{IJ}`$ to couple to matter like a magnetic field (further enforcing the three dimensional analogy). Therefore, in all covariant differentiation operators ordinary derivatives $`_I`$ have to be replaced by $`_Iil^2\widehat{\theta }_I`$; the vector potential $`\widehat{\theta }_I`$ is associated to the two-form $`\widehat{\omega }_{IJ}`$ in the usual way: $`\widehat{\omega }_{IJ}=_I\widehat{\theta }_J_J\widehat{\theta }_I`$. On dimensional grounds we are forced to introduce a fundamental length $`l`$ in the theory. In the absence of other scales we are naturally lead to identify $`l`$ with the Planck length $`l10^{33}cm`$. However, all basic feature of elementary matter fields discussed in this paper only depend on the assumption that $`l`$ is small (i.e. we are considering the theory at energy scales much smaller than the one associated to $`l`$) and not on the particular value it takes. It is worth remarking that $`\widehat{\theta }_I`$ is not an external gauge structure superimposed on the manifold. It is an intrinsic geometrical object that naturally comes in requiring the structural group of the geometry to be –in part– unitary.
In order to write down field equations for test matter fields it is convenient to choose a Darboux coordinate frame $`\widehat{x}^I=(x^\mu ,\xi ^i)`$ and partially fix the gauge freedom associated the one-form $`\widehat{\theta }_I`$ by choosing $`\widehat{\theta }_I=(0,\theta _i)`$. Differentiation in ordinary space-time directions is then realized by the usual momenta $`i_\mu `$. Together with the $`x^\mu `$ these operators close the ordinary space-time Heisenberg algebra generating $`SO(1,3)`$ in the standard way. Differentiation in weak and strong directions is instead realized by the operators
$$\mathrm{\Pi }_i=il_il^1\theta _i$$
These ones are no longer a set of mutually commuting operators. They fulfill commutation relations
$$[\mathrm{\Pi }_i,\mathrm{\Pi }_j]=i\epsilon _{ij}$$
where $`\epsilon _{ij}\widehat{\epsilon }_{ij}`$ is the non-degenerate extra part of $`\widehat{\epsilon }_{IJ}`$. The $`\mathrm{\Pi }_i`$ behaves as five pairs of canonically conjugate operators. For this reason, it is no longer convenient to consider the multiplication by $`\xi ^i`$ as conjugate to $`\mathrm{\Pi }_i`$. We rearrange canonical variables by introducing the independent set of operators
$$\mathrm{\Xi }^i=\xi ^i+l\epsilon ^{ij}\mathrm{\Pi }_j$$
where $`\epsilon _{ik}\epsilon ^{jk}=\delta _i^j`$. We can check out immediately that the $`\mathrm{\Xi }^i`$ commute with the $`\mathrm{\Pi }_i`$ and fulfill commutation relations
$$[\mathrm{\Xi }^i,\mathrm{\Xi }^j]=il^2\epsilon ^{ji}$$
The $`\mathrm{\Xi }^i`$ are five more pairs of canonically conjugate operators. These complete the representation of Heisenberg algebra in extra directions. Infinitesimal generators of the two $`U(1)`$, of $`SU(2)`$ and of $`SU(3)`$ –somehow the equivalent of the orbital angular momentum in ordinary directions– are constructed in terms of $`\mathrm{\Pi }_i`$ as
$$L^{ij}=\frac{1}{4}(\delta ^{ik}\epsilon ^{jl}\delta ^{jk}\epsilon ^{il})\{\mathrm{\Pi }_k,\mathrm{\Pi }_l\}$$
The extra orbital angular momentum $`L^{ij}`$ fulfills the conditions $`L_{ij}+L_{ij}=0`$ and $`L_i^k\epsilon _{kj}L_j^k\epsilon _{ki}=0`$. Following the notation previously introduced we group its components in the generators $`L^x`$, $`L^y`$ of the two $`U(1)`$ groups; the generators $`L^{\iota _1}`$, $`L^{\iota _2}`$, $`L^{\iota _3}`$ of $`SU(2)`$; and the generators $`L^{\kappa _1}`$, $`L^{\kappa _2}`$, …, $`L^{\kappa _8}`$ of $`SU(3)`$.
## 2 Spinor in a background
Unification of fundamental forces in space-time connection should be mirrored by unification of elementary matter in a fermionic field. The ability of including force and matter in a single geometrical frame is a crucial test for the theory. In addition we have to justify why our direct perception of the fourteen dimensional space-time continuous is reduced to the familiar four real dimensions. To these tasks we consider the free propagation of a test spinor in a given fourteen dimensional background.
We start or analysis by drawing an analogy with the three dimensional motion of a charged particle in a magnetic field . The analogy is between the fourteen dimensional space-time and the three dimensional space. The background geometry is identified with the Euclidean metric plus the magnetic field. The field is required to have a constant magnitude $`1/l`$ but is otherwise allowed to vary arbitrarily in direction. The test spinor is identified with the charged particle. The magnetic field null direction spans magnetic lines. Therefore, these have to be pictured out as ordinary space-time leaves. The two directions orthogonal to the magnetic field are instead identified with extra directions. Observe that the analogy is not complete in that the magnetic field is not required to be parallelly transported by the Euclidean structure and the spin freedom of the three dimensional particle is disregarded. Nonetheless, the two systems share the same qualitative dynamical behavior. The three dimensional model is meant as a tool to concretely picture out the behavior of the fourteen dimensional spinor in space-time.
Consider first the case of an homogeneous magnetic field (Figure 3a). A charged particle moving in such a background performs a rapid rotation of radius $`l`$ in the plane orthogonal to the field while freely propagating along a straight magnetic line. The stronger the field magnitude the smaller the radius of the circle. For very large values of $`1/l`$ the trajectory of the particle is indistinguishable from a straight line. A conservation law protect the particle from drifting in directions orthogonal to the field so that the motion starting in a neighborhood of a line will never drift away. The motion of the particle is effectively constrained on a magnetic line.
Allow now the field to vary in direction (Figure 3b). Magnetic lines are no longer straight. However, in the limit of very large values of $`1/l`$ the motion of the particle still separate on three different energies scales. A consistent fraction of energy (going as $`1/l`$) is stored in a rapid rotation orthogonal to the field direction. On a lower energy scale (going as $`l^0`$) the particle drifts along the magnetic line. The effective one dimensional motion couples to the magnetic line torsion as to a $`U(1)`$ gauge potential; the effective charge is proportional to the angular momentum stored in the fast rotation . On an even lower energy scale (going as $`l^2`$) the particle drifts in directions normal to the field. In contrast with the homogeneous case, the motion starting in a neighborhood of a generic magnetic line does not remain there forever. However, if the effective Hamiltonian describing the very slow drift in normal directions presents an absolute minimum in correspondence of a given magnetic line, whatever initial conditions will bring the particle in a neighborhood of that line. The motion of the particle gets effectively constrained on a privileged magnetic line.
The picture does not change substantially if we switch from classical to quantum dynamics . For very large values of $`1/l`$ the wavefunction of the system effectively separates in a part taking into account dynamics in normal directions and a part describing the effective motion of the particle along the field line. The normal rotational motion gets quantized and the particle effectively behaves as frozen in a rotational eigenstate –not necessarily the ground state. Normal rotational eigenstates: are confined in a neighborhood of size $`l`$ of the magnetic line; are labeled by a $`U(1)`$ quantum number corresponding to the angular momentum stored in normal directions; are infinitely degenerate as far as the very slow drift in normal directions is negligible. In correspondence to every normal eigenstate the particle effectively propagates along the field line. The effective quantum motion still couples to the field line torsion as to a $`U(1)`$ gauge field; the effective charge is quantized and proportional to the angular momentum stored in normal directions.
Having this in mind we go back to the original problem. By a spinor on a pseudo-Riemannian manifold we mean a field transforming according to a spinorial representation of the structural pseudo-orthogonal group. The natural embedding of the structural group of space-time in $`SO(1,13)`$ allows us a natural definition of space-time spinors. We introduce Dirac’s gamma matrices $`\widehat{\gamma }^A`$ fulfilling standard anti-commutation relations
$$\{\widehat{\gamma }^A,\widehat{\gamma }^B\}=2\widehat{\eta }^{AB}$$
As usual, independent commutators of the gammas generate the spin-$`1/2`$ representation of $`SO(1,13)`$
$$\widehat{\mathrm{\Sigma }}^{AB}=\frac{i}{4}[\widehat{\gamma }^A,\widehat{\gamma }^B]$$
Among the $`\widehat{\mathrm{\Sigma }}^{AB}`$ we identify ordinary and extra spin angular momentum operators as the linear combinations generating the subgroup $`SO(1,3)\times U(1)\times U(1)\times SU(2)\times SU(3)`$. Once more we take advantage of the notation introduced for infinitesimal generators: $`\widehat{\mathrm{\Sigma }}^{\alpha \beta }`$ for ordinary space-time $`SO(1,3)`$; $`\widehat{\mathrm{\Sigma }}^x`$ and $`\widehat{\mathrm{\Sigma }}^y`$ generating the two $`U(1)`$ subgroups; $`\widehat{\mathrm{\Sigma }}^{\iota _1}`$, $`\widehat{\mathrm{\Sigma }}^{\iota _2}`$, $`\widehat{\mathrm{\Sigma }}^{\iota _3}`$ generating $`SU(2)`$; and $`\widehat{\mathrm{\Sigma }}^{\kappa _1}`$, $`\widehat{\mathrm{\Sigma }}^{\kappa _2}`$, …, $`\widehat{\mathrm{\Sigma }}^{\kappa _8}`$ for the generators of $`SU(3)`$. In fourteen dimensions gamma and sigma matrices have size $`2^7`$ by $`2^7`$ and act on Dirac spinors having $`2^7`$ components.
The free propagation of a Dirac spinor field $`\widehat{\psi }(\widehat{x})`$ in a curved background is described by the covariant matter field equations
$$\left(\widehat{\gamma }^A\widehat{e}_A^I\widehat{D}_I\widehat{𝗉}\right)\widehat{\psi }(\widehat{x})=0$$
where $`\widehat{D}_I=i_Il^2\widehat{\theta }_I+\frac{1}{2}\widehat{\mathrm{\Omega }}_{I,AB}\widehat{\mathrm{\Sigma }}^{AB}`$ is the covariant derivative on spinors with extra coupling and $`\widehat{𝗉}`$ is a fourteen dimensional mass term. The sign in front of $`l^2\widehat{\theta }_I`$ is clearly arbitrary. This choice corresponds to standard particle physics notations. The sign in front of the connection term is fixed by general covariance. Observe that since $`\widehat{\mathrm{\Omega }}_{I,AB}`$ belongs to $`so(1,3)u(1)u(1)su(2)su(3)`$ only the linear combinations of $`\widehat{\mathrm{\Sigma }}^{AB}`$ defining ordinary and extra spin angular momentum appear in the equation. This is consistent with the required local covariance. As an additional physical proviso we require the integral of $`\widehat{\psi }^{}\widehat{\psi }`$ over the whole thirteen dimensional space (no time) to be finite. As a consequence not all values of $`\widehat{𝗉}`$ yield acceptable solutions. In the following we identify square integrable solutions of matter field equations with elementary fermions.
Matter field equations are not easy to solve in a generic background. On the other hand, we expect on physical grounds that the fundamental length $`l`$ is extremely small –perhaps of order of the Planck scale $`l10^{33}cm`$. This allows us to proceed perturbatively in $`l`$. The qualitative behavior of the system is analogue to that of a charged particle in a strong magnetic background. The freedoms of the system divide on three different energy scales producing field configurations to squeeze in a neighborhood of a $`1+3`$ hyper-surface. To get quantitative we adopt the following strategy. Introduce Darboux coordinates $`\widehat{x}^I=(x^\mu ,\xi ^i)`$ and partially fix extra gauge freedom by choosing $`\widehat{\theta }_I=(0,\theta _i)`$. Recalling that $`\xi ^i=\mathrm{\Xi }^il\epsilon ^{ij}\mathrm{\Pi }_j`$ rewrite $`\widehat{\gamma }^A\widehat{e}_A^ID_I`$ in terms of the canonical operators $`\mathrm{\Pi }_i`$, $`i_\mu `$, $`x^\mu `$ and $`\mathrm{\Xi }^i`$ plus ordinary and extra spin. This produces a natural expansion of field equations in powers of $`l`$. The solution is then constructed by means of ordinary perturbation theory. An explicit analysis shows that the leading order term goes as $`l^1`$ and represents –up to higher order terms– an harmonic oscillator in the canonical variables $`\mathrm{\Pi }_i`$ and extra spin. These are the only freedoms active on this energy scale. They are associated to a fast rotation of size $`l`$ in extra directions and carry $`U(1)\times U(1)\times SU(2)\times SU(3)`$ quantum numbers. We will see shortly that these freedom are associated to family quantum numbers. The next to leading term is of order $`l^0`$ and resembles a covariant Dirac operator on an effective $`1+3`$ space-time. The freedoms active on this energy scale are the canonical partners $`i_\mu `$, $`x^\mu `$ and ordinary spin. These describe effective ordinary motion. The remaining freedoms $`\mathrm{\Xi }^i`$ become dynamical only at higher order. They are associated to the very slow drift in extra directions. As long as the effects of such a very slow drift are negligible $`U(1)\times U(1)\times SU(2)\times SU(3)`$ multiplets are degenerate. In a while it will become clear that this degeneracy is responsible for the phenomenon of families replication.
We now focus on the analysis of two simple background configurations that allow us to capture the basic feature of the dimensional reduction process in a particularly clean way. The free motion in a flat background –somehow a ‘special relativistic’ problem– displays clearly the emergence of effective four dimensional chiral fermions, the mechanism making them to carry hypercharge, isospin and color quantum numbers and the one producing the replication of a basic pattern in different generations. The free motion on an intrinsically flat ordinary space-time leave non trivially embedded in space-time allows us to check how the effective four dimensional motion couples to non-gravitational interactions.
### Free motion in a flat background
In a flat background it is possible and convenient to globally parallelize reference frames. We take therefore $`e_\alpha ^\mu \delta _\alpha ^\mu `$, $`e_a^i\delta _a^i`$ and $`a_i^\mu 0`$. With this natural choice matter field equations take the simple form
$$\left[\frac{1}{l}\widehat{\gamma }^a\mathrm{\Pi }_ai\widehat{\gamma }^\mu _\mu \right]\widehat{\psi }=\widehat{𝗉}\widehat{\psi }$$
Ordinary and extra freedoms decouples. The operators $`\mathrm{\Xi }^i`$ are constants of motion. The problem is capable of a direct exact solution. However, in view of the more complicated case in a curved background we proceed by constructing solutions by means of perturbation theory. The leading order term $`\widehat{\gamma }^a\mathrm{\Pi }_a`$ is identified with the solvable part of the problem. The remaining terms with the perturbation.
The leading order corresponds to the eigenvalues problem
$$\widehat{\gamma }^a\mathrm{\Pi }_a\widehat{\psi }_{\underset{¯}{𝗉}}=\widehat{𝗉}\widehat{\psi }_{\underset{¯}{𝗉}}$$
where the multi-index $`\underset{¯}{𝗉}`$ denotes the set of quantum numbers necessary to label eigenstates corresponding to the eigenvalue $`\widehat{𝗉}`$. In order to construct solutions it is useful to explicitly separate ordinary four dimensional and extra ten dimensional spinor indices by parameterizing the $`\widehat{\gamma }^A`$ in terms of four and ten dimensional gamma matrices. Denoting by $`\gamma ^\mu `$, $`\mu =0,1,2,3`$, ordinary four by four $`SO(1,3)`$ Dirac matrices, by $`\gamma ^a`$, $`a=4,5,\mathrm{},13`$, extra thirty-two by thirty-two $`SO(10)`$ gamma matrices and by $`\gamma _{}=i\gamma ^0\gamma ^1\gamma ^2\gamma ^3`$ and $`\gamma _{}=i\gamma ^{1_w}\gamma ^{2_w}\mathrm{}\gamma ^{6_s}`$ the four and ten dimensional chiral matrices, we construct the representation
$$\begin{array}{cc}\widehat{\gamma }^\alpha =\gamma ^\alpha 11_{32}\hfill & \alpha =0,1,2,3\hfill \\ \widehat{\gamma }^a=\gamma _{}\gamma ^a\hfill & a=4,5,\mathrm{},13\hfill \end{array}$$
It is readily check that these matrices fulfill the correct anti-commutation relations. In correspondence we rewrite the spinor field as the tensor product of an ordinary four dimensional spinor times an extra ten dimensional spinor: $`\widehat{\psi }_{\underset{¯}{𝗉}}(\widehat{x})=\psi _{\underset{¯}{𝗉}}(x)\varphi _{\underset{¯}{𝗉}}(\xi )`$. Four and ten dimensional components may be chosen as functions of ordinary and extra coordinates respectively. The leading order eigenvalue problem decouples as
$$\{\begin{array}{c}\gamma _{}\psi _{\underset{¯}{𝗉}}(x)=\pm \psi _{\underset{¯}{𝗉}}(x)\hfill \\ 𝒫\varphi _{\underset{¯}{𝗉}}(\xi )=𝗉\varphi _{\underset{¯}{𝗉}}(\xi )\hfill \end{array}$$
where we have introduced the particle operator $`𝒫=\gamma ^a\mathrm{\Pi }_a`$. For very small values of $`l`$ extra freedoms are frozen in $`𝒫`$ eigenstates. In the next section we show that particle operator eigenstates: are squeezed in a neighborhood of size $`l`$ of an ordinary $`1+3`$ real hyper-surfaces; are organized in $`U(1)\times U(1)\times SU(2)\times SU(3)`$ multiplets; every multiplet is repeated in the spectrum as a consequence of $`\mathrm{\Xi }^i`$ conservation. In correspondence to every particle operator eigenstate the four dimensional spinor $`\psi _{\underset{¯}{𝗉}}(x)`$ –which inherit the multiplet structure and relative degeneracy from the corresponding $`𝒫`$ eigenstate– describes the effective four dimensional motion of the system. Leading order matter field equations require four dimensional spinors to be chiral. This is a first very important phenomenological prediction of the theory.
Recall now that the whole theory is $`C\widehat{T}P`$ invariant. Since leading order matter field equations do not depend on bosonic ordinary space-time variables, $`C\widehat{T}P`$ decomposes in the product of two discrete symmetries which are independently conserved
$$C\widehat{T}P=(\gamma _{}CTP_{}11_{32})(\gamma _{}CP_{})$$
where $`CTP_{}`$ and $`CP_{}`$ are the four and ten dimensional part of the total $`C\widehat{T}P`$ operator
$$\begin{array}{c}CTP_{}\psi (x)=\gamma _{}\psi ^{}(x)\hfill \\ CP_{}\varphi (\xi )=\gamma _{}\varphi ^{}(\xi )\hfill \end{array}$$
The first term in the $`C\widehat{T}P`$ decomposition acts as the identity operator on spinor indices. Its role is that of enforcing $`CTP_{}`$ on four dimensional chiral spinors. The second term relates ordinary and extra spinor indices in a non trivial manner. We introduce the special symbol $`\widehat{\mathrm{\Gamma }}\gamma _{}CP_{}`$ to denote it. Clearly $`\widehat{\mathrm{\Gamma }}^2=1`$. $`\widehat{\mathrm{\Gamma }}`$ is a discrete symmetry dividing the spectrum in two sectors. Assuming the system in one of these sectors the chirality of every four dimensional spinor $`\psi _{\underset{¯}{𝗉}}(x)`$ is fixed by the $`CP_{}`$ value of the corresponding $`\varphi _{\underset{¯}{𝗉}}(\xi )`$.
We evaluate next to leading corrections by means of ordinary perturbation theory. Eigenvalues corrections $`\mathrm{\Delta }\widehat{𝗉}`$ and leading order eigenfunctions left unspecified by degeneracies are determined by diagonalizing the perturbation over degenerate leading order eigenstates. This amounts to solve the ordinary space-time eigenvalue problem $`i\gamma ^\mu _\mu \psi _{\underset{¯}{𝗉}}=\mathrm{\Delta }\widehat{𝗉}\psi _{\underset{¯}{𝗉}}`$. Eigenvalues corrections $`\mathrm{\Delta }\widehat{𝗉}`$ are readily determined by the chirality constraint on $`\psi _{\underset{¯}{𝗉}}(x)`$. They all equal zero: $`\mathrm{\Delta }\widehat{𝗉}=0`$. While degeneracy is not lifted, zero order eigenfunctions are partially determined. In the absence of interactions, four dimensional chiral spinors $`\psi _{\underset{¯}{𝗉}}(x)`$ describing the effective motion in ordinary space-time obey the free massless Dirac equation
$$i\gamma ^\mu _\mu \psi _{\underset{¯}{𝗉}}=0$$
This is consistent with the prescriptions of the Standard Model.
It is readily checked that $`\widehat{\psi }_{\underset{¯}{𝗉}}(\widehat{x})=\psi _{\underset{¯}{𝗉}}(x)\times \varphi _{\underset{¯}{𝗉}}(\xi )`$ are exact solutions of our original problem. No further corrections have to be evaluated.
We have all the necessary ingredients to identify space-time spinors with elementary fermions. For very small values of $`l`$ the fourteen dimensional matter field effectively decouples in the tensor product of a four dimensional chiral spinor $`\psi _{\underset{¯}{𝗉}}(x)`$ times a ten dimensional particle operator eigenspinor $`\varphi _{\underset{¯}{𝗉}}(\xi )`$.
Particle operators eigenspinors are squeezed in a neighborhood of an ordinary space-time hyper-surface. Particle operators eigenspinors are organized in $`U(1)\times U(1)\times SU(2)\times SU(3)`$ multiplets. The whole pattern of multiplets is repeated many times in the spectrum. Every particle operator eigenfunction carries a definite value of $`CP_{}`$. Given the very large amount of energy necessary to excite transitions between different particle operator eigenstates the system behaves as frozen in extra eigenstates.
In correspondence to every particle operator eigenstate the four dimensional chiral spinor $`\psi _{\underset{¯}{𝗉}}(x)`$ propagates freely in ordinary spacetime. Four dimensional spinors are organized in $`U(1)\times U(1)\times SU(2)\times SU(3)`$ multiplets in accordance with the structure of their extra partners. The pattern of multiplets is repeated a certain number of times. The chirality of every four dimensional spinor is fixed by the $`CP_{}`$ value of its extra partner.
The effective four dimensional qualitative structure is identical to the one of the Standard Model of Elementary Particles. Four dimensional fermions are chiral. Multiplet quantum numbers corresponds family quantum numbers. The repetition of a basic pattern of multiplets to the phenomenon of families replication. Four dimensional matter field equations in absence of interactions are correct.
In order to make the correspondence quantitative we have to check two more things. First of all, we should investigate how the effective four dimensional dynamics couples to a curved background. This allows us to double check the correct identification of $`U(1)\times SU(2)\times SU(3)`$ quantum numbers with hypercharge, isospin and color of elementary matter fields and the correct identification of extrinsic torsion with hyper, weak and strong interactions. Second, we have to explicitly evaluate the particle operator spectrum checking whether the basic pattern of multiplets corresponds to the familiar family structure.
### Free motion in a minimal curved background
In a generic background it is not possible to globally parallelize reference frames. Space-time leaves are no longer flat. This yields an effective force on particles freely propagating in the background. In order to explore how four dimensional fermions couple to non-gravitational interactions we assume the existence of a privileged intrinsically flat but extrinsically curved leave to be identified with ordinary space-time. To ensure field configurations to remain in a neighborhood of the privileged leave we further assume it to correspond to a stationary point of reference frames as functions of extra coordinates. Expanding in extra directions we then obtain the following ansatz for reference frames in a neighborhood of the privileged leave: $`e_\alpha ^\mu =\delta _\alpha ^\mu +𝒪(l)`$, $`e_a^i=e_a^i(x)+𝒪(l^2)`$ and $`a_i^\mu =𝒪(l)`$, where all the functions are evaluated on the leave. This approach has the disadvantage that the extra freedoms $`\mathrm{\Xi }^i`$ are assumed from the very beginning to be non-dynamical. The method can not be used to evaluate higher order corrections. It is nevertheless sufficient for our present task. By explicitly computing connection components we obtain matter field equations up to higher order in $`l`$ as
$$\left[\frac{1}{l}\widehat{\gamma }^ae_a^i(x)\mathrm{\Pi }_i+\widehat{\gamma }^\mu (i_\mu +\frac{1}{2}𝒜_{\mu ,ab}\widehat{\mathrm{\Sigma }}^{ab})+𝒪(l)\right]\widehat{\psi }=\widehat{𝗉}\widehat{\psi }$$
Ordinary and extra freedoms no longer decouples. We can not proceed to a direct solution of the problem. On the other hand, we can still proceed perturbatively.
The leading order term of the operator depends now explicitly on ordinary space-time bosonic variables. In order to apply our previous analysis a little preparatory work is necessary. As a first thing we observe that –since extra reference frames fulfill the identity $`e_a^i(x)e_b^j(x)\epsilon _{ij}=\epsilon _{ab}`$– the operators $`e_a^i(x)\mathrm{\Pi }_i`$ close canonical commutation relations $`[e_a^i\mathrm{\Pi }_i,e_b^i\mathrm{\Pi }_j]=i\epsilon _{ab}`$. Therefore, the $`x`$ dependence of the leading order term is fictitious and can be removed by an appropriate redefinition of bosonic canonical variables. To this task we look for a purely bosonic unitary transformation $`U`$ such that
$$Ue_a^i(x)\mathrm{\Pi }_iU^{}=\mathrm{\Pi }_a$$
where the $`\mathrm{\Pi }_a`$ are new canonical operators fulfilling $`[\mathrm{\Pi }_a,\mathrm{\Pi }_b]=i\epsilon _{ab}`$. Operating $`U`$ on field equations we bring the leading order term back to the simple form $`\widehat{\gamma }^a\mathrm{\Pi }_a/l`$. At the same time we generate a new term $`i\widehat{\gamma }^\mu U(_\mu U^{})`$ in the perturbation. In the general case the transformation $`U`$ brings the $`\mathrm{\Xi }^i`$ –not dynamical here– in a new set of canonical variables. However, we can insist the new operators to commute with ordinary space-time canonical operators; we require $`[i_\mu ,\mathrm{\Pi }_a]=0`$. This condition yields the equation
$$[U(_\mu U^{}),\mathrm{\Pi }_a]=𝒜_{\mu ,a}^b\mathrm{\Pi }_b$$
allowing us the determination of the quantity $`U(_\mu U^{})`$ without even knowing the explicit form of $`U`$. We redly obtain the solution
$$U(_\mu U^{})=\frac{i}{2}𝒜_{\mu ,ab}L^{ab}$$
where $`L^{ab}`$ are the extra orbital angular momentum operators introduced in the previous section. This allows us to operate the unitary transformation. Denoting by $`\widehat{J}^{ab}=L^{ab}\widehat{\mathrm{\Sigma }}^{ab}`$ the extra total angular momentum operators we obtain the matter field equations in the form
$$\left[\frac{1}{l}\widehat{\gamma }^a\mathrm{\Pi }_a+\widehat{\gamma }^\mu \left(i_\mu \frac{1}{2}𝒜_{\mu ,ab}(x)\widehat{J}^{ab}\right)+𝒪(l)\right]\widehat{\psi }=\widehat{𝗉}\widehat{\psi }$$
Up to extrinsic torsion and higher order terms these resemble matter field equations in a flat background. Observe that $`𝒜_{\mu ,ab}`$ couples to ordinary variables like a non-gravitational gauge connection. Extra indices play the role of gauge indices. In order to bring matter field equations in an even more suggestive form we explicitly introduce the two $`U(1)`$, $`SU(2)`$ and $`SU(3)`$ infinitesimal generators. Without further delay we address these extra space-time symmetries as extracharge, hypercharge, isospin and color. Decompose $`\widehat{J}^{ab}`$ in $`\widehat{J}^x=L^x\widehat{\mathrm{\Sigma }}^x`$, $`\widehat{J}^y=L^y\widehat{\mathrm{\Sigma }}^y`$, … etc., according to the notation introduced in the previously section. Recall the similar decomposition for the extrinsic torsion $`𝒜_{\mu ,ab}`$ in $`𝒜_\mu ^x`$, $`𝒜_\mu ^y`$, … etc.. Denote by $`𝖷\widehat{J}^x`$ the $`U(1)`$ extracharge generator, by $`𝖸\widehat{J}^y`$ the $`U(1)`$ hypercharge generator, by $`𝖨^𝗂\widehat{J}^{\iota 𝗂}`$, $`𝗂=1,2,3`$, the $`SU(2)`$ isospin generators, and by $`𝖢^𝖼\widehat{J}^{\kappa 𝖼}`$, $`𝖼=1,2,\mathrm{},8`$, the $`SU(3)`$ color generators. Matter field equations take then the form
$$\left[\frac{1}{l}\widehat{\gamma }^a\mathrm{\Pi }_a+\widehat{\gamma }^\mu \left(i_\mu 𝖷𝒜_\mu ^x𝖸𝒜_\mu ^y𝒜_\mu ^{\iota 𝗂}𝖨^𝗂𝒜_\mu ^{\kappa 𝖼}𝖢^𝖼\right)+𝒪(l)\right]\widehat{\psi }=\widehat{𝗉}\widehat{\psi }$$
Ordinary and extra freedom are still coupled. In order to effectively decouple them we precede by means of perturbation theory.
The discussion of the leading order goes exactly as in a flat background.
The analysis gets a little more subtle when next to leading corrections are evaluated. As before the chirality constraint on $`\psi _{\underset{¯}{𝗉}}(x)`$ forces eigenvalues corrections $`\mathrm{\Delta }\widehat{𝗉}`$ to vanish identically. On the other hand, the presence of isospin $`𝖨^𝗂`$ and color $`𝖢^𝖼`$ operators in the perturbation produces now a coupling among extra eigenstates belonging to a $`U(1)\times U(1)\times SU(2)\times SU(3)`$ multiplet. As a consequence the effective equations of motion for four dimensional spinors belonging to a multiplet no longer decouples. In order to make the analysis transparent it is useful to explicitly introduce indices labeling different states of $`U(1)\times U(1)\times SU(2)\times SU(3)`$ multiplets. We denote by $`\varphi _{\underset{¯}{𝗉},\mathrm{a}}|\underset{¯}{𝗉},\mathrm{a}`$, $`\mathrm{a}=1,\mathrm{},n`$, particle operator eigenstates belonging to a given multiplet of dimension $`n`$. Introduce the matrices
$$\begin{array}{cc}𝖨_{𝗉,\mathrm{ab}}^𝗂=\underset{¯}{𝗉},\mathrm{a}|𝖨^𝗂|\underset{¯}{𝗉},\mathrm{b}\hfill & 𝗂=1,2,3\hfill \\ 𝖢_{𝗉,\mathrm{ab}}^𝖼=\underset{¯}{𝗉},\mathrm{a}|𝖢^𝖼|\underset{¯}{𝗉},\mathrm{b}\hfill & 𝖼=1,2,\mathrm{},8\hfill \end{array}$$
where Roman indices $`\mathrm{a}`$ and $`\mathrm{b}`$ run over the degeneracy dimension $`n`$. It is immediate to check that $`𝖨_{𝗉,\mathrm{ab}}^𝗂`$ and $`𝖢_{𝗉,\mathrm{ab}}^𝖼`$ furnish an $`n`$ by $`n`$ representation of $`SU(2)\times SU(3)`$. Since $`𝖷`$ and $`𝖸`$ commute with isospin and color generators, the representation carries definite extracharge and hypercharge values: $`\underset{¯}{𝗉},\mathrm{a}|𝖷|\underset{¯}{𝗉},\mathrm{b}=𝖷_𝗉11_{\mathrm{ab}}`$, $`\underset{¯}{𝗉},\mathrm{a}|𝖸|\underset{¯}{𝗉},\mathrm{b}=𝖸_𝗉11_{\mathrm{ab}}`$. Diagonalizing the perturbation over leading order eigenstates amounts to solve the massless Dirac problems for chiral spinors coupled to a $`U(1)\times U(1)\times SU(2)\times SU(3)`$ gauge field
$$\underset{\mathrm{b}}{}\gamma ^\mu (i11_{\mathrm{ab}}_\mu 𝖷_𝗉11_{\mathrm{ab}}𝒜_\mu ^x𝖸_𝗉11_{\mathrm{ab}}𝒜_\mu ^y𝒜^{\iota 𝗂}𝖨_{𝗉,\mathrm{ab}}^𝗂𝒜^{\kappa 𝖼}𝖢_{𝗉,\mathrm{ab}}^𝖼)\psi _{\underset{¯}{𝗉},\mathrm{b}}=0$$
By recompressing degeneracy indices inside the multi-index $`\underset{¯}{𝗉}`$ we bring the equation in the standard form. Effective four dimensional matter field equations reads
$$\gamma ^\mu (i_\mu 𝖷_𝗉𝒜_\mu ^x𝖸_𝗉𝒜_\mu ^y𝒜^{\iota 𝗂}𝖨_𝗉^𝗂𝒜^{\kappa 𝖼}𝖢_𝗉^𝖼)\psi _{\underset{¯}{𝗉}}=0$$
where multiplet and spinor indices are understood. Up to extracharge coupling these equations reproduce correctly the prescriptions of the Standard Model of Elementary Particle. The identification of $`U(1)\times SU(2)\times SU(3)`$ quantum numbers with hypercharge, isospin and color as well as the identification of extrinsic torsion with hyper, weak and strong interactions are therefore consistent.
## 3 Fermion quantum numbers
A catalog of elementary particles is obtained by explicitly analyzing the spectrum of the particle operator
$$𝒫=\gamma ^a\mathrm{\Pi }_a$$
### Symmetries
The first thing striking us is the simplicity in form of $`𝒫`$. The particle operator is obtained as the contraction of ten extra bosonic operators $`\mathrm{\Pi }_a`$ fulfilling commutation relations $`[\mathrm{\Pi }_a,\mathrm{\Pi }_b]=i\epsilon _{ab}`$ with ten extra fermionic operators $`\gamma ^a`$ fulfilling anti-commutation relations $`\{\gamma ^a,\gamma ^b\}=\delta ^{ab}`$. The form is invariant under the exchange of bosonic and fermionic freedoms. Therefore, we expect the presence of a supersymmetry. In analogy with $`𝒫=\gamma ^a\delta _a^b\mathrm{\Pi }_b`$ we introduce a second operator $`\overline{𝒫}=\gamma ^a\epsilon _a^b\mathrm{\Pi }_b`$ where $`\epsilon _a^b`$ –somehow the anti-symmetric analogue of $`\delta _a^b`$– has the same form of $`\epsilon _{ab}`$ with $`w`$ and $`s`$ set equal to one<sup>5</sup><sup>5</sup>5It is worth remarking that $`\epsilon _a^b`$ as well as the previously introduced $`\epsilon ^{ab}`$ are not obtained from $`\epsilon _{ab}`$ by highering indices with the flat metric.. We can readily check that $`𝒫`$ and $`\overline{𝒫}`$ anticommute while their squares both equal the infinitesimal generator of extracharge transformations $`𝖷=\delta ^{ab}\mathrm{\Pi }_a\mathrm{\Pi }_b\epsilon _{ab}\mathrm{\Sigma }^{ab}`$. This is actually the reason for choosing the generators of the two $`U(1)`$ groups as we did. Moreover, the particle operator and its supersymmetric partner anticommute with the extra chiral gamma $`\gamma _{}`$. The operators $`𝒫`$, $`\overline{𝒫}`$, $`𝖷`$ and $`\gamma _{}`$ close the $`N=2`$ superalgebra
$$\begin{array}{c}\{𝒫,𝒫\}=\{\overline{𝒫},\overline{𝒫}\}=2𝖷\\ \{𝒫,\overline{𝒫}\}=\{𝒫,\gamma _{}\}=\{\overline{𝒫},\gamma _{}\}=0\end{array}$$
As a consequence we can define an operator $`S=i\gamma _{}\overline{𝒫}`$ commuting with $`𝒫`$ that can be used to label particle operator eigenstates. The square of $`S`$ equals the extracharge $`S^2=𝖷`$. The relevant information is therefore a sign. We define the operator $`\varsigma =\text{sign}(CP_{}S)`$ dividing the spectrum in the sectors $`\varsigma =\pm 1`$.
A direct computation shows that the particle operator commutes with the generators of extracharge, hypercharge, isospin and color transformations. All together the operators $`𝒫`$, $`S`$, $`𝖷`$, $`𝖸`$, $`𝖨^𝗂`$, $`𝗂=1,2,3`$, $`𝖢^𝖼`$, $`𝖼=1,\mathrm{},8`$ close the commutation relations
$$\begin{array}{c}[𝒫,S]=[𝒫,𝖷]=[𝒫,𝖸]=[𝒫,𝖨^𝗂]=[𝒫,𝖢^𝖼]=0\\ [S,𝖷]=[S,𝖸]=[S,𝖨^𝗂]=[S,𝖢^𝖼]=0\\ [𝖷,𝖸]=[𝖷,𝖨^𝗂]=[𝖷,𝖢^𝖼]=0\\ [𝖸,𝖨^𝗂]=[𝖸,𝖢^𝖼]=0\\ [𝖨^𝗂,𝖢^𝖼]=0\end{array}$$
and
$$[𝖨^𝗂,𝖨^𝗃]=iϵ^{\mathrm{𝗂𝗃𝗄}}𝖨^𝗄[𝖢^𝖼,𝖢^𝖽]=i𝖿^{\mathrm{𝖼𝖽𝖾}}𝖢^𝖾$$
where $`ϵ^{\mathrm{𝗂𝗃𝗄}}`$ and $`𝖿^{\mathrm{𝖼𝖽𝖾}}`$ are standard $`su(2)`$ and $`su(3)`$ structure constants. Extracharge, hypercharge, isospin and color quantum number can be used to label particle operator eigenstates. Eigenvalues of $`𝒫`$ equal the square root of extracharge up to a sign. Choosing to explicitly indicate the value of $`𝖷`$ among eigenstates labels, only the sign $`\pi =\text{sign}(𝒫)`$ of the particle operator matters. For every eigenstate we indicate: the value of $`\pi `$; the value of $`\varsigma `$; the value of $`𝖷`$; the value of $`𝖸`$; the isospin representation $`𝖨=\mathrm{𝟏},\mathrm{𝟐},\mathrm{𝟑},\mathrm{}`$ and the value of its third generator $`𝖨_3`$; the color representation $`𝖢=\mathrm{𝟏},\mathrm{𝟑},\overline{\mathrm{𝟑}},\mathrm{}`$ and the values of its third and eighth generators $`𝖢_3`$ and $`𝖢_8`$. Representations of $`su(2)`$ and $`su(3)`$ are indicated by their dimension in boldface characters as long as such a notation is unambiguous. Isospin and color quantum numbers will also be represented collectively by the sole representations $`𝖨`$ and $`𝖢`$ when the focus is on the multiplet and not on the particular state.
Finally, we should remember that $`𝒫`$, $`S`$, $`𝖷`$, $`𝖸`$, $`𝖨^𝗂`$, $`𝗂=1,2,3`$, $`𝖢^𝖼`$, $`𝖼=1,\mathrm{},8`$, commute with the ten canonical operators controlling the very slow drift in extra directions
$$[𝒫,\mathrm{\Xi }^a]=[S,\mathrm{\Xi }^a]=[𝖷,\mathrm{\Xi }^a]=[𝖸,\mathrm{\Xi }^a]=[𝖨^𝗂,\mathrm{\Xi }^a]=[𝖢^𝖼,\mathrm{\Xi }^a]=0$$
where $`a=1_w,\mathrm{},4_w,1_s,\mathrm{},6_s`$. With the $`\mathrm{\Xi }^a`$ we can construct up to five mutually commuting operators labeling different generations of particles. The form of these operators and hence of the quantum numbers $`\underset{¯}{𝐆}`$ labeling generations are presumably determined by higher order corrections in the perturbative expansion.
Particle operator eigenstates $`\varphi _{\underset{¯}{𝗉}}(\xi )|\underset{¯}{𝗉}`$ are completely specified by
$$|\underset{¯}{𝗉}=|\pm ;\pm ;𝖷;𝖸;𝖨,𝖨_3;𝖢,𝖢_3,𝖢_8;\underset{¯}{𝐆}$$
where the first sign is refers to $`\pi `$ and the second to $`\varsigma `$. In this paper we do not consider effects of symmetry breaking. As a consequence we omit generations quantum numbers among eigenstates labels in the sequel.
### Creation and annihilation operators
In order to explicitly construct eigenvalues and eigenstates of the particle operator it is useful to introduce creation and annihilation operators corresponding to bosonic and fermionic extra freedoms. In weak directions we define
$$\begin{array}{cc}\{\begin{array}{c}a_{1w}=\frac{1}{\sqrt{2w}}\left(\mathrm{\Pi }_{1_w}+i\mathrm{\Pi }_{2_w}\right)\hfill \\ a_{1w}^{}{}_{}{}^{}=\frac{1}{\sqrt{2w}}\left(\mathrm{\Pi }_{1_w}i\mathrm{\Pi }_{2_w}\right)\hfill \end{array}\hfill & \{\begin{array}{c}a_{2w}=\frac{1}{\sqrt{2w}}\left(\mathrm{\Pi }_{3_w}+i\mathrm{\Pi }_{4_w}\right)\hfill \\ a_{2w}^{}{}_{}{}^{}=\frac{1}{\sqrt{2w}}\left(\mathrm{\Pi }_{3_w}i\mathrm{\Pi }_{4_w}\right)\hfill \end{array}\hfill \\ \{\begin{array}{c}b_{1w}=\frac{1}{2}\left(\gamma ^{1_w}+i\gamma ^{2_w}\right)\hfill \\ b_{1w}^{}{}_{}{}^{}=\frac{1}{2}\left(\gamma ^{1_w}i\gamma ^{2_w}\right)\hfill \end{array}\hfill & \{\begin{array}{c}b_{2w}=\frac{1}{2}\left(\gamma ^{3_w}+i\gamma ^{4_w}\right)\hfill \\ b_{2w}^{}{}_{}{}^{}=\frac{1}{2}\left(\gamma ^{3_w}i\gamma ^{4_w}\right)\hfill \end{array}\hfill \end{array}$$
while in strong directions
$$\begin{array}{ccc}\{\begin{array}{c}a_{1s}=\frac{1}{\sqrt{2s}}\left(\mathrm{\Pi }_{1_s}+i\mathrm{\Pi }_{2_s}\right)\hfill \\ a_{1s}^{}{}_{}{}^{}=\frac{1}{\sqrt{2s}}\left(\mathrm{\Pi }_{1_s}i\mathrm{\Pi }_{2_s}\right)\hfill \end{array}\hfill & \{\begin{array}{c}a_{2s}=\frac{1}{\sqrt{2s}}\left(\mathrm{\Pi }_{3_s}+i\mathrm{\Pi }_{4_s}\right)\hfill \\ a_{2s}^{}{}_{}{}^{}=\frac{1}{\sqrt{2s}}\left(\mathrm{\Pi }_{3_s}i\mathrm{\Pi }_{4_s}\right)\hfill \end{array}\hfill & \{\begin{array}{c}a_{3s}=\frac{1}{\sqrt{2s}}\left(\mathrm{\Pi }_{5_s}+i\mathrm{\Pi }_{6_s}\right)\hfill \\ a_{3s}^{}{}_{}{}^{}=\frac{1}{\sqrt{2s}}\left(\mathrm{\Pi }_{5_s}i\mathrm{\Pi }_{6_s}\right)\hfill \end{array}\hfill \\ \{\begin{array}{c}b_{1s}=\frac{1}{2}\left(\gamma ^{1_s}+i\gamma ^{2_s}\right)\hfill \\ b_{1s}^{}{}_{}{}^{}=\frac{1}{2}\left(\gamma ^{1_s}i\gamma ^{2_s}\right)\hfill \end{array}\hfill & \{\begin{array}{c}b_{2s}=\frac{1}{2}\left(\gamma ^{3_s}+i\gamma ^{4_s}\right)\hfill \\ b_{2s}^{}{}_{}{}^{}=\frac{1}{2}\left(\gamma ^{3_s}i\gamma ^{4_s}\right)\hfill \end{array}\hfill & \{\begin{array}{c}b_{3s}=\frac{1}{2}\left(\gamma ^{5_s}+i\gamma ^{6_s}\right)\hfill \\ b_{3s}^{}{}_{}{}^{}=\frac{1}{2}\left(\gamma ^{5_s}i\gamma ^{6_s}\right)\hfill \end{array}\hfill \end{array}$$
It is readily checked that $`a_\mathrm{i}`$, $`a_{\mathrm{i}}^{}{}_{}{}^{}`$, $`b_\mathrm{i}`$ and $`b_{\mathrm{i}}^{}{}_{}{}^{}`$, $`\mathrm{i}=1w,2w,1s,2s,3s`$ fulfill standard commutation, anti-commutation relations $`[a_\mathrm{i},a_{\mathrm{j}}^{}{}_{}{}^{}]=\delta _{\mathrm{ij}}`$ and $`\{b_\mathrm{i},b_{\mathrm{j}}^{}{}_{}{}^{}\}=\delta _{\mathrm{ij}}`$. We rewrite the whole algebra of operators in terms of creators and annihilators.
The particle operator $`𝒫`$ takes the form
$$\begin{array}{c}𝒫=\sqrt{2w}\left(a_{1w}b_{1w}^{}{}_{}{}^{}+a_{1w}^{}{}_{}{}^{}b_{1w}+a_{2w}b_{2w}^{}{}_{}{}^{}+a_{2w}^{}{}_{}{}^{}b_{2w}\right)+\hfill \\ +\sqrt{2s}\left(a_{1s}b_{1s}^{}{}_{}{}^{}+a_{1s}^{}{}_{}{}^{}b_{1s}+a_{2s}b_{2s}^{}{}_{}{}^{}+a_{2s}^{}{}_{}{}^{}b_{2s}+a_{3s}b_{3s}^{}{}_{}{}^{}+a_{3s}^{}{}_{}{}^{}b_{3s}\right)\hfill \end{array}$$
its supersymmetric partner $`\overline{𝒫}`$
$$\begin{array}{c}\overline{𝒫}=i\sqrt{2w}\left(a_{1w}b_{1w}^{}{}_{}{}^{}a_{1w}^{}{}_{}{}^{}b_{1w}+a_{2w}b_{2w}^{}{}_{}{}^{}a_{2w}^{}{}_{}{}^{}b_{2w}\right)+\hfill \\ i\sqrt{2s}\left(a_{1s}b_{1s}^{}{}_{}{}^{}a_{1s}^{}{}_{}{}^{}b_{1s}+a_{2s}b_{2s}^{}{}_{}{}^{}a_{2s}^{}{}_{}{}^{}b_{2s}+a_{3s}b_{3s}^{}{}_{}{}^{}a_{3s}^{}{}_{}{}^{}b_{3s}\right)\hfill \end{array}$$
Even though $`𝒫`$ and $`\overline{𝒫}`$ are not diagonal is the standard occupation number basis associated to $`a_\mathrm{i}`$ and $`b_\mathrm{i}`$, creators and annihilators remains extremely useful tools in constructing the spectrum. Indeed, the extracharge operator $`𝖷=𝒫^2=\overline{𝒫}^2`$ as well as the hypercharge $`𝖸`$, isospin $`𝖨_3`$ and color $`𝖢_3`$, $`𝖢_8`$ operators labeling eigenstates are diagonal in this basis. For extracharge we obtain
$$\begin{array}{c}𝖷=2w\left(a_{1w}^{}{}_{}{}^{}a_{1w}+a_{2w}^{}{}_{}{}^{}a_{2w}+b_{1w}^{}{}_{}{}^{}b_{1w}+b_{2w}^{}{}_{}{}^{}b_{2w}\right)\hfill \\ +2s\left(a_{1s}^{}{}_{}{}^{}a_{1s}+a_{2s}^{}{}_{}{}^{}a_{2s}+a_{3s}^{}{}_{}{}^{}a_{3s}+b_{1s}^{}{}_{}{}^{}b_{1s}+b_{2s}^{}{}_{}{}^{}b_{2s}+b_{3s}^{}{}_{}{}^{}b_{3s}\right)\hfill \end{array}$$
while for hypercharge
$$\begin{array}{c}𝖸=2w\left(a_{1w}^{}{}_{}{}^{}a_{1w}+a_{2w}^{}{}_{}{}^{}a_{2w}+b_{1w}^{}{}_{}{}^{}b_{1w}+b_{2w}^{}{}_{}{}^{}b_{2w}\right)\hfill \\ 2s\left(a_{1s}^{}{}_{}{}^{}a_{1s}+a_{2s}^{}{}_{}{}^{}a_{2s}+a_{3s}^{}{}_{}{}^{}a_{3s}+b_{1s}^{}{}_{}{}^{}b_{1s}+b_{2s}^{}{}_{}{}^{}b_{2s}+b_{3s}^{}{}_{}{}^{}b_{3s}\right)\hfill \end{array}$$
In order to construct eigenstates it is useful to have an explicit expression of isospin highering and lowering operators $`𝖨_\pm =𝖨_1\pm i𝖨_2`$ other that $`𝖨_3`$. We obtain the isospin algebra in the form
$$\begin{array}{c}𝖨_+=a_{2w}^{}{}_{}{}^{}a_{1w}+b_{2w}^{}{}_{}{}^{}b_{1w}\hfill \\ 𝖨_{}=a_{1w}^{}{}_{}{}^{}a_{2w}+b_{1w}^{}{}_{}{}^{}b_{2w}\hfill \\ 𝖨_3=\frac{1}{2}\left(a_{2w}^{}{}_{}{}^{}a_{2w}a_{1w}^{}{}_{}{}^{}a_{1w}+b_{2w}^{}{}_{}{}^{}b_{2w}b_{1w}^{}{}_{}{}^{}b_{1w}\right)\hfill \end{array}$$
Highering and lowering operators for the color algebra are also necessary. We introduce the mnemonic notation $`𝖢_{}=𝖢_1\pm i𝖢_2`$, $`𝖢_{}=𝖢_4\pm i𝖢_5`$ and $`𝖢_{}=𝖢_6\pm i𝖢_7`$ for the operators moving states in the $`su(3)`$ weights plane. Arrows indicate directions in which states are moved. We obtain the color algebra in the form
$$\begin{array}{c}𝖢_{}=a_{2s}^{}{}_{}{}^{}a_{1s}+b_{2s}^{}{}_{}{}^{}b_{1s}\hfill \\ 𝖢_{}=a_{1s}^{}{}_{}{}^{}a_{2s}+b_{1s}^{}{}_{}{}^{}b_{2s}\hfill \\ 𝖢_{}=a_{3s}^{}{}_{}{}^{}a_{2s}+b_{3s}^{}{}_{}{}^{}b_{2s}\hfill \\ 𝖢_{}=a_{2s}^{}{}_{}{}^{}a_{3s}+b_{2s}^{}{}_{}{}^{}b_{3s}\hfill \\ 𝖢_{}=a_{3s}^{}{}_{}{}^{}a_{1s}+b_{3s}^{}{}_{}{}^{}b_{1s}\hfill \\ 𝖢_{}=a_{1s}^{}{}_{}{}^{}a_{3s}+b_{1s}^{}{}_{}{}^{}b_{3s}\hfill \\ 𝖢_3=\frac{1}{2}\left(a_{2s}^{}{}_{}{}^{}a_{2s}a_{1s}^{}{}_{}{}^{}a_{1s}+b_{2s}^{}{}_{}{}^{}b_{2s}b_{1s}^{}{}_{}{}^{}b_{1s}\right)\hfill \\ 𝖢_8=\frac{1}{2\sqrt{3}}\left(2a_{3s}^{}{}_{}{}^{}a_{3s}a_{2s}^{}{}_{}{}^{}a_{2s}a_{1s}^{}{}_{}{}^{}a_{1s}+2b_{3s}^{}{}_{}{}^{}b_{3s}b_{2s}^{}{}_{}{}^{}b_{2s}b_{1s}^{}{}_{}{}^{}b_{1s}\right)\hfill \end{array}$$
Incidentally we observe that the expression of highering and lowering operators in terms of creators and annihilators is a particularly happy one. Indeed, we only have to remember the fundamental $`su(2)`$ diagram $`_{{}_{1}{}^{}_2}`$ to reconstruct $`𝖨_+`$ as the operator annihilating in $`1`$ and re-creating in $`2`$ and $`𝖨_{}`$ as the one annihilating in $`2`$ and re-creating in $`1`$. The same holds for $`su(3)`$. Given our choice of conventions the realization of the color algebra transforms according the $`\overline{\mathrm{𝟑}}`$ representation, so that the fundamental diagram we have to remember is $`_{{}_{1}{}^{}_2}^^^3`$. As an example, it is clear that the operator $`𝖢_{}`$ corresponds to annihilating in $`1`$ and re-creating in $`3`$, $`𝖢_{}`$ to annihilating in $`3`$ and re-creating in $`2`$, etc.. A different choice of sign in front of the term $`l^2\widehat{\theta }_I`$ in matter field equations would have produced the representation $`\mathrm{𝟑}`$ corresponding to the fundamental diagram $`_____3^{{}_{}{}^{2}^1}`$.
We have all the necessary ingredients to explicitly construct particle operator eigenstates in terms of occupation number states. As usual, these are obtained from the vacuum $`|\mathrm{𝟎}`$ by acting with creation operators
$`|n_{1w},n_{2w},\sigma _{1w},\sigma _{2w};n_{1s},n_{2s},n_{3s},\sigma _{1s},\sigma _{2s},\sigma _{3s}=`$
$`{\displaystyle \frac{a_{1w}^{}{}_{}{}^{}{}_{}{}^{n_{1w}}}{\sqrt{n_{1w}!}}}{\displaystyle \frac{a_{2w}^{}{}_{}{}^{}{}_{}{}^{n_{2w}}}{\sqrt{n_{2w}!}}}b_{1w}^{}{}_{}{}^{}{}_{}{}^{\sigma _{1w}}b_{2w}^{}{}_{}{}^{}{}_{}{}^{\sigma _{2w}}{\displaystyle \frac{a_{1s}^{}{}_{}{}^{}{}_{}{}^{n_{1s}}}{\sqrt{n_{1s}!}}}{\displaystyle \frac{a_{2s}^{}{}_{}{}^{}{}_{}{}^{n_{2s}}}{\sqrt{n_{2s}!}}}{\displaystyle \frac{a_{3s}^{}{}_{}{}^{}{}_{}{}^{n_{3s}}}{\sqrt{n_{3s}!}}}b_{1s}^{}{}_{}{}^{}{}_{}{}^{\sigma _{1s}}b_{2s}^{}{}_{}{}^{}{}_{}{}^{\sigma _{2s}}b_{3s}^{}{}_{}{}^{}{}_{}{}^{\sigma _{3s}}|\mathrm{𝟎}`$
the integer numbers $`n_\mathrm{i}=0,1,2,\mathrm{}`$ denote eigenvalues of bosonic number operators $`a_{\mathrm{i}}^{}{}_{}{}^{}a_\mathrm{i}`$; the binary numbers $`\sigma _\mathrm{i}=0,1`$ denote eigenvalues of fermionic number operators $`b_{\mathrm{i}}^{}{}_{}{}^{}b_\mathrm{i}`$ (clearly $`\mathrm{i}=1w,2w,1s,2s,3s`$ and here we are not summing over repeated indices). Observe that this is more than a formal expression. Number operators are explicitly realized as differential operators on square integrable functions of extra coordinates and spin. The vacuum $`|\mathrm{𝟎}`$ as well as all excited states correspond to real extra wavefunctions having –as harmonic oscillator eigenstates– a typical size $`l`$. This proves the effective confinement of the system in a neighborhood of size $`l`$ of an ordinary $`1+3`$ hyper-surface. In order to simplify notation we group weak bosonic occupation number $`n_{w1},n_{w2}`$ in $`\underset{¯}{n_w}`$, weak fermionic occupation numbers $`\sigma _{1w},\sigma _{2w}`$ in $`\underset{¯}{\sigma _w}`$ … etc. . We also write things like $`|\underset{¯}{2_1},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}`$ to denote the state $`|2,0,0,1;0,0,0,0,0,0`$.
### Spectrum and particle states
The particle operator spectrum is readily determined by recalling that $`𝒫^2=𝖷`$. The extracharge $`𝖷`$ consist in the sum of two harmonic oscillator corresponding to the frequency $`2w`$ plus three harmonic oscillators corresponding to the frequency $`2s`$. Therefore, the particle operator spectrum is
$$𝗉=\pm \sqrt{2w𝗇_𝗐+2s𝗇_𝗌}$$
with $`𝗇_𝗐,𝗇_𝗌=0,1,2\mathrm{}`$ . The spectrum is discrete. In order to explicitly construct eigenstates corresponding to the eigenvalues $`\pm |𝗉|`$ we proceed in the following standard manner. Select all occupation number states corresponding to the extracharge value $`𝗉^2`$. With these states construct all the linear combinations that are simultaneously annihilated by $`𝖨_+`$, $`𝖢_{}`$, $`𝖢_{}`$ and $`𝖢_{}`$. Among these restricted number of states select new linear combinations having a definite value of $`𝒫`$ and $`S`$. These are the highest weight states of the multiplets corresponding to the given eigenvalue. Act now on these states with the lowering operators $`𝖨_{}`$, $`𝖢_{}`$, $`𝖢_{}`$ and $`𝖢_{}`$ to construct the whole representations. As a check control that the total number of $`𝒫`$ eigenstates coincides with the original number of $`𝖷`$ eigenstates.
Let us now proceed to the explicit construction of the first few particle operator eigenstates and to the identification of the corresponding elementary particles:
$`𝗉=0`$
To $`𝖷=0`$ corresponds the vacuum state $`|\underset{¯}{0},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}|\mathrm{𝟎}`$. This is the only particle operator eigenstate with $`𝗉=0`$
$$|;;0;0;\mathrm{𝟏},0;\mathrm{𝟏},0,0=|\underset{¯}{0},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}$$
the signs $`\pi `$ and $`\varsigma `$ are clearly not defined. The state has hypercharge $`𝖸=0`$, is an isospin singlet $`𝖨=\mathrm{𝟏}`$ and a color singlet $`𝖢=\mathrm{𝟏}`$. Its $`CP_{}`$ value is $`+1`$. The corresponding particle is a very non-interactive one. It has the right quantum numbers to be identified with a right-handed neutrino
$$\psi _{0;0;\mathrm{𝟏};\mathrm{𝟏}}\nu _{eR}$$
This identification correspond to assume the system in the sector $`\widehat{\mathrm{\Gamma }}=+1`$ fixing the chirality of all particles.
$`𝗉=\pm \sqrt{2w}`$
To $`𝖷=2w`$ correspond four occupation number states: $`|\underset{¯}{1_1},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}`$, $`|\underset{¯}{0},\underset{¯}{1_1};\underset{¯}{0},\underset{¯}{0}`$, $`|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}`$ and $`|\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}`$. The last two are the only ones simultaneously annihilated by $`𝖨_+`$, $`𝖢_{}`$, $`𝖢_{}`$ and $`𝖢_{}`$. Explicitly analyzing the action of $`𝒫`$ on them we easily construct the particle operator eigenstates
$$\begin{array}{cc}|\pm ;\pm ;2w;2w;\mathrm{𝟐},\frac{1}{2};\mathrm{𝟏},0,0\hfill & =\frac{1}{\sqrt{2}}\left(|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}\right)\hfill \\ |\pm ;\pm ;2w;2w;\mathrm{𝟐},\frac{1}{2};\mathrm{𝟏},0,0\hfill & =\frac{1}{\sqrt{2}}\left(|\underset{¯}{1_1},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{1_1};\underset{¯}{0},\underset{¯}{0}\right)\hfill \end{array}$$
These carry hypercharge $`𝖸=2w`$, are isospin doublets $`𝖨=\mathrm{𝟐}`$ and color singlets $`𝖢=\mathrm{𝟏}`$. Their $`CP_{}`$ value is $`1`$. The corresponding particle is therefore left-handed and has the right isospin and color quantum numbers to be identified with a leptonic doublet
$$\psi _{2w;2w;\mathrm{𝟐};\mathrm{𝟏}}\left(\begin{array}{c}\nu _{eL}\\ e_L\end{array}\right)$$
Requiring the doublet hypercharge to be $`1/2`$ we fix the value of the weak parameter $`w`$ to $`1/4`$. This is geometrically very satisfactory being the inverse of the number of weak directions. Choosing the value of the weak parameter $`w`$ as $`1/4`$ fixes the weak part of hypercharge of all the remaining particles of the spectrum.
$`𝗉=\pm \sqrt{2s}`$
To $`𝖷=2s`$ correspond six occupation number states out of which we construct the particle operator eigenstates
$$\begin{array}{cc}|\pm ;\pm ;2s;2s;\mathrm{𝟏},0;\overline{\mathrm{𝟑}},0,\frac{1}{\sqrt{3}}\hfill & =\frac{1}{\sqrt{2}}\left(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_3},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_3}\right)\hfill \\ |\pm ;\pm ;2s;2s;\mathrm{𝟏},0;\overline{\mathrm{𝟑}},\frac{1}{2},\frac{1}{2\sqrt{3}}\hfill & =\frac{1}{\sqrt{2}}\left(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_2},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_2}\right)\hfill \\ |\pm ;\pm ;2s;2s;\mathrm{𝟏},0;\overline{\mathrm{𝟑}},\frac{1}{2},\frac{1}{2\sqrt{3}}\hfill & =\frac{1}{\sqrt{2}}\left(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_1},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_1}\right)\hfill \end{array}$$
These carry hypercharge $`𝖸=2s`$, are isospin singlets $`𝖨=\mathrm{𝟏}`$ and color anti-triplets $`𝖢=\overline{\mathrm{𝟑}}`$. Their $`CP_{}`$ value is again $`1`$. The corresponding particle has therefore the right quantum numbers to be identified with an anti right-handed quark (which is a left-handed particle). The negative value of the hypercharge indicates that the quark is of type down<sup>6</sup><sup>6</sup>6We should remark that $`s`$ is positive. A negative value of $`s`$ will simply interchange the role of extracharge and hypercharge.
$$\psi _{2s;2s;\mathrm{𝟏};\overline{\mathrm{𝟑}}}d_R^{}$$
The identification requires the particle hypercharge to be $`1/3`$ fixing the value of the strong parameter $`s`$ to $`1/6`$. Again, this is geometrically very satisfactory. As for the weak parameter, it is the inverse of the number of strong directions. Choosing $`s`$ to be $`1/6`$ fixes the strong part of hypercharge of all the remaining particles.
The identification of particle operator eigenstates corresponding to $`𝗉=\pm \sqrt{2w}`$ and $`𝗉=\pm \sqrt{2s}`$ with a neutrino-electron type isospin doublet and with a quark down type color anti-triplets fix the only two parameters left unspecified by the gauge group<sup>7</sup><sup>7</sup>7Clearly, the parameters $`w`$ and $`s`$ can always be rescaled through a redefinition of $`l`$. What really matters is fixing $`w/s`$ to 3/2 –the ratio of numbers of strong and weak directions.. Extracharge and hypercharge of all remaining elementary fermions are determined by the choice
$$w=\frac{1}{4}s=\frac{1}{6}$$
The correct chirality assignment for the first three states is obtained if the fourteen dimensional system is in the sector $`\widehat{\mathrm{\Gamma }}=+1`$. The remaining discrete symmetries $`\pi `$ and $`\varsigma `$ further divide the spectrum in sectors. It is natural to assume the system in a single one of these sectors. We proceed by analyzing a few more particle states:
$`𝗉=\pm \sqrt{2w+2s}`$
To $`𝖷=2w+2s`$ correspond $`24`$ occupation number states. Out of these only $`|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{1_3},\underset{¯}{0}`$, $`|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_3}`$, $`|\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{1_3},\underset{¯}{0}`$ and $`|\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{1_3}`$ are simultaneously annihilated by the highering operators $`𝖨_+`$, $`𝖢_{}`$ $`𝖢_{}`$ and $`𝖢_{}`$. Imposing that linear combinations of these are simultaneous eigenstates of $`𝒫`$ and $`S`$ we obtain a linear system of eight equations for the four coefficients. This is immediately solved yielding the highest weight states of four isospin doublets $`𝖨=\mathrm{𝟐}`$, color anti-triplets $`𝖢=\overline{\mathrm{𝟑}}`$
$$\begin{array}{c}|\pm ;\pm ;2w+2s;2w2s;\mathrm{𝟐},\frac{1}{2};\overline{\mathrm{𝟑}},0,\frac{1}{\sqrt{3}}=\hfill \\ \frac{1}{\sqrt{4w+4s}}\left(\sqrt{2w+2s}|\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{1_3}\pm \sqrt{2w}|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_3}\sqrt{2s}|\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{1_3},\underset{¯}{0}\right)\hfill \\ \text{and}\hfill \\ |\pm ;;2w+2s;2w2s;\mathrm{𝟐},\frac{1}{2};\overline{\mathrm{𝟑}},0,\frac{1}{\sqrt{3}}=\hfill \\ \frac{1}{\sqrt{4w+4s}}\left(\sqrt{2s}|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_3}+\sqrt{2w}|\underset{¯}{0},\underset{¯}{1_2};\underset{¯}{1_3},\underset{¯}{0}\pm \sqrt{2w+2s}|\underset{¯}{1_2},\underset{¯}{0};\underset{¯}{1_3},\underset{¯}{0}\right)\hfill \end{array}$$
The remaining multiplets states are easily constructed by means lowering operators. The multiplets carry hypercharge $`𝖸=1/6`$. Their $`CP_{}`$ value is $`+1`$, corresponding therefore to right-handed particles. These are the correct quantum numbers for anti left-handed quark doublets
$$\psi _{2w+2s;2w2s;\mathrm{𝟐};\overline{\mathrm{𝟑}}}\left(\begin{array}{c}u_L\\ d_L\end{array}\right)^{}$$
Observe that the four multiplet belongs to different sector of the theory.
$`𝗉=\pm \sqrt{4w}`$
Out of the eight occupation number states corresponding to $`𝖷=4w`$ we construct two $`𝒫`$ eigenstates transforming like isospin singlets $`𝖨=\mathrm{𝟏}`$, color singlets $`𝖢=\mathrm{𝟏}`$ and two eigenstates transforming like isospin triplets $`𝖨=\mathrm{𝟑}`$, color singlet $`𝖢=\mathrm{𝟏}`$. Maximal weight states are obtained as
$$\begin{array}{c}|\pm ;\pm ;4w;4w;\mathrm{𝟏},0;\mathrm{𝟏},0,0=\hfill \\ =\frac{1}{2}\left(|\underset{¯}{1_1},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}|\underset{¯}{1_2},\underset{¯}{1_1};\underset{¯}{0},\underset{¯}{0}\pm \sqrt{2}|\underset{¯}{0},\underset{¯}{1_11_2};\underset{¯}{0},\underset{¯}{0}\right)\hfill \\ \text{and}\hfill \\ |\pm ;;4w;4w;\mathrm{𝟑},1;\mathrm{𝟏},0,0=\frac{1}{\sqrt{2}}\left(|\underset{¯}{2_2},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}\pm |\underset{¯}{1_2},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}\right)\hfill \end{array}$$
These carry hypercharge $`𝖸=1`$ and correspond to $`CP_{}=+1`$. The singlet states have the correct quantum numbers to be identified with right-handed leptons of electronic type
$$\psi _{4w;4w;\mathrm{𝟏};\mathrm{𝟏}}e_R$$
The identification of isospin triplets is more puzzling. However, we should remark that the four multiplets belongs to different sectors of the theory.
$`𝗉=\pm \sqrt{4s}`$
To $`𝖷=4s`$ correspond eighteen occupation number states out of which we construct two $`𝒫`$ eigenstates transforming like isospin singlets $`𝖨=\mathrm{𝟏}`$, color triplet $`𝖢=\mathrm{𝟑}`$ and two eigenstates transforming like isospin singlets $`𝖨=\mathrm{𝟏}`$ and according the $`𝖢=\overline{\mathrm{𝟔}}`$ color representation. Once more we display the explicit form of highest weight eigenstates
$$\begin{array}{c}|\pm ;\pm ;4s;4s;\mathrm{𝟏},0;\mathrm{𝟑},\frac{1}{2},\frac{1}{2\sqrt{3}}=\hfill \\ =\frac{1}{2}\left(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_2},\underset{¯}{1_3}|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_3},\underset{¯}{1_2}\pm \sqrt{2}|\underset{¯}{0},\underset{¯}{0};\underset{¯}{0},\underset{¯}{1_21_3}\right)\hfill \\ \text{and}\hfill \\ |\pm ;;4s;4s;\mathrm{𝟏},0;\overline{\mathrm{𝟔}},0,\frac{2}{\sqrt{3}}=\frac{1}{\sqrt{2}}\left(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{2_3},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_3},\underset{¯}{1_3}\right)\hfill \end{array}$$
All these multiplets carry hypercharge $`𝖸=2/3`$ and correspond to the $`CP_{}`$ value $`+1`$. The color triplets have therefore the correct quantum numbers to be identified with right-handed quarks of type up
$$\psi _{4s;4s;\mathrm{𝟏};\mathrm{𝟑}}u_R$$
Puzzling is the identification of the multiplets transforming according the $`\overline{\mathrm{𝟔}}`$ color representation. Again, we remark that the four multiplets belong to different sectors of the theory.
As a consequence of charge conjugation symmetry the sectors $`\pi =+`$ and $`\pi =`$ have exactly the same structure. Therefore, we can restrict attention to one of the two. In Table 1 we summarize the content in multiplet of the sector $`\pi =+`$ of the particle states investigated so far. The sector $`\varsigma =+`$ reproduces correctly the structure of a family.
Beyond the family structure, the theory predicts the existence of infinite many other elementary fermions corresponding to higher excited states. Moreover, even if we labeled $`𝒫`$ eigenstates with the names of first generation particles $`e`$, $`\nu _e`$, $`d`$, $`u`$, we should keep in mind that because of $`\mathrm{\Xi }^a`$ degeneracy this pattern is repeated in the second generation $`\mu `$, $`\nu _\mu `$, $`s`$, $`c`$, in the third generation $`\tau `$, $`\nu _\tau `$, $`b`$, $`t`$, and possibly in many others<sup>8</sup><sup>8</sup>8The number of generations is proportional to the extra-volume and is not necessarily infinite.. There is no way of distinguishing an electron from a muon or a tau before electro-weak symmetry breaking. It is not unreasonable to expect that symmetry breaking is produced by the coupling with a fourteen dimensional scalar field or simply by higher order terms in the expansion of matter field equations in a curved background. In any case –whatever produces the symmetry breaking– the comprehension of the correct mechanism of mass generation will make the prediction of new particles and generations in a severe test for the theory.
Let us now focus on Table 1. Is it really natural to identify its content with a family of elementary fermions? As a matter of fact the assignment of hypercharge, isospin, color and chirality of every particle is correct. A definitive answer to the question can only be given through the exhibition of a natural symmetry breaking mechanism. However, it is not unreasonable to assume that a particle could possibly acquire a mass in reason of the average value of the extracharge –and perhaps hypercharge and complexity of isospin and color representations– of its right and left components. If this were the case, the quark down would be lighter than the quark up and –even worst– than the electron. Moreover, the fact that the right-handed components of the quark down and the left-handed components of the quark doublet appear in the spectrum through their anti-particles is not particularly appealing. To have a deeper insight into the problem we analyze the content of higher particle operator eigenstates. This can be done in very general terms. In the discussion of generic particle operator eigenstates it is better to indicate isospin and color representations by the total $`su(2)`$ isospin and by the two non negative integer labeling $`su(3)`$ representations. We first consider states in which only weak modes are excited. These correspond to eigenvalues $`𝗉=\pm \sqrt{2w𝗇_𝗐}`$ where $`𝗇_𝗐`$ is any non negative integer
$`𝗉=\pm \sqrt{2w𝗇_𝗐}`$
By counting the possible way of distributing $`𝗇_𝗐`$ quanta in two bosonic plus two fermionic harmonic oscillators, we realize that to the extracharge value $`𝖷=2w𝗇_𝗐`$ correspond $`4𝗇_𝗐`$ occupation number states. Out of these we easily select the ones simultaneously annihilated by $`𝖨_+`$, $`𝖢_{}`$, $`𝖢_{}`$, $`𝖢_{}`$ and rearrange them in $`𝒫`$, $`S`$ eigenstates. We obtain the maximal weight states
$$\begin{array}{c}|\pm ;()^{𝗇_𝗐};2w𝗇_𝗐;2w𝗇_𝗐;\frac{𝗇_𝗐}{2},\frac{𝗇_𝗐}{2};(0,0),0,0=\hfill \\ \frac{1}{\sqrt{2}}\left(|\underset{¯}{𝗇_{𝗐}^{}{}_{2}{}^{}},\underset{¯}{0};\underset{¯}{0},\underset{¯}{0}\pm |\underset{¯}{(𝗇_𝗐1)_2},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}\right)\hfill \\ \text{and}\hfill \\ |\pm ;\pm ()^{𝗇_𝗐};2w𝗇_𝗐;2w𝗇_𝗐;\frac{𝗇_𝗐2}{2},\frac{𝗇_𝗐2}{2};(0,0),0,0=\frac{1}{\sqrt{2𝗇_𝗐}}(|\underset{¯}{1_1(𝗇_𝗐2)_2},\underset{¯}{1_2};\underset{¯}{0},\underset{¯}{0}+\hfill \\ \sqrt{𝗇_𝗐1}|\underset{¯}{(𝗇_𝗐1)_2},\underset{¯}{1_1};\underset{¯}{0},\underset{¯}{0}\sqrt{𝗇_𝗐}|\underset{¯}{(𝗇_𝗐2)_2},\underset{¯}{1_11_2};\underset{¯}{0},\underset{¯}{0})\hfill \end{array}$$
corresponding to two representation of isospin $`𝖨=\frac{𝗇_𝗐}{2}`$ and two representation of isospin $`𝖨=\frac{𝗇_𝗐2}{2}`$; color quantum numbers are singlets $`𝖢=(0,0)`$. The total number of states contained in the four representations is correctly of $`4𝗇_𝗐`$. The particle operator level $`𝗉=\sqrt{2w𝗇_𝗐}`$ contains therefore the multiplets
$$(\frac{𝗇_𝗐}{2},(0,0))(\frac{𝗇_𝗐2}{2},(0,0))$$
where the second irreducible representation appears only for $`𝗇_𝗐2`$. All states carry hypercharge $`𝖸=2w𝗇_𝗐`$ and correspond to the $`CP_{}`$ value $`(1)^{𝗇_𝗐}`$.
Next, we consider particle operator eigenstates in which only strong modes are excited. These correspond to eigenvalues $`𝗉=\pm \sqrt{2s𝗇_𝗌}`$ with $`𝗇_𝗌`$ any non negative integer
$`𝗉=\pm \sqrt{2s𝗇_𝗌}`$
We count again all possible ways of distributing $`𝗇_𝗌`$ quanta in three bosonic plus three fermionic harmonic oscillators. In this way we find $`4𝗇_{𝗌}^{}{}_{}{}^{2}+2`$ occupation number states corresponding to the extracharge value $`𝖷=2s𝗇_𝗌`$. Maximal weight states are constructed in analogy with the with the previous case. We obtain
$$\begin{array}{c}|\pm ;()^{𝗇_𝗌};2s𝗇_𝗌;2s𝗇_𝗌;0,0;(0,𝗇_𝗌),0,\frac{𝗇_𝗌}{\sqrt{3}}=\hfill \\ \frac{1}{\sqrt{2}}\left(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{𝗇_{𝗌}^{}{}_{3}{}^{}},\underset{¯}{0}\pm |\underset{¯}{0},\underset{¯}{0};\underset{¯}{(𝗇_𝗌1)_3},\underset{¯}{1_3}\right)\hfill \\ |\pm ;\pm ()^{𝗇_𝗌};2s𝗇_𝗌;2s𝗇_𝗌;0,0;(1,𝗇_𝗌2),\frac{1}{2},\frac{2𝗇_𝗌3}{2\sqrt{3}}=\frac{1}{\sqrt{2𝗇_𝗌}}(|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_2(𝗇_𝗌2)_3},\underset{¯}{1_3}+\hfill \\ \sqrt{𝗇_𝗌1}|\underset{¯}{0},\underset{¯}{0};\underset{¯}{(𝗇_𝗌1)_3},\underset{¯}{1_2}\pm \sqrt{𝗇_𝗌}|\underset{¯}{0},\underset{¯}{0};\underset{¯}{(𝗇_𝗐2)_3},\underset{¯}{1_21_3})\hfill \\ \text{and}\hfill \\ |\pm ;()^{𝗇_𝗌};2s𝗇_𝗌;2s𝗇_𝗌;0,0;(0,𝗇_𝗌3),0,\frac{𝗇_𝗌3}{\sqrt{3}}=\hfill \\ \frac{1}{\sqrt{2𝗇_𝗌}}\left(\right|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_1(𝗇_𝗌3)_3},\underset{¯}{1_21_3}+|\underset{¯}{0},\underset{¯}{0};\underset{¯}{1_2(𝗇_𝗌3)_3},\underset{¯}{1_11_3}+\hfill \\ +\sqrt{𝗇_𝗌2}|\underset{¯}{0},\underset{¯}{0};\underset{¯}{(𝗇_𝗌2)_3},\underset{¯}{1_11_2}\pm \sqrt{𝗇_𝗌}|\underset{¯}{0},\underset{¯}{0};\underset{¯}{(𝗇_𝗌3)_3},\underset{¯}{1_11_21_3})\hfill \end{array}$$
corresponding to isospin singlets $`𝖨=0`$ and to the color representations $`𝖢=(0,𝗇_𝗌)`$, $`𝖢=(1,𝗇_𝗌2)`$ and $`𝖢=(0,𝗇_𝗌3)`$ respectively. The total number of states contained in the six irreducible representations is correctly of $`4𝗇_{𝗌}^{}{}_{}{}^{2}+2`$. The content in multiplets of the level $`𝗉=\sqrt{2s𝗇_𝗌}`$ is therefore
$$(0,(0,𝗇_𝗌))(0,(1,𝗇_𝗌2))(0,(0,𝗇_𝗌3))$$
where the second and third irreducible representations appear only for $`𝗇_𝗌2`$ and $`𝗇_𝗌3`$ respectively. All states carry hypercharge $`𝖸=2s𝗇_𝗌`$ and correspond to the $`CP_{}`$ value $`(1)^{𝗇_𝗌}`$.
Out of these two particular cases, we can reconstruct the contents in multiplet of a generic particle operator eigenstate
$`𝗉=\pm \sqrt{2w𝗇_𝗐+2s𝗇_𝗌}`$
The representations content of the particle operator state containing $`𝗇_𝗐`$ weak oscillation quanta and $`𝗇_𝗌`$ strong oscillation quanta coincides with the direct product of the representations contents of the particle operator states corresponding to $`𝗉=\pm \sqrt{2w𝗇_𝗐}`$ and $`𝗉=\pm \sqrt{2s𝗇_𝗌}`$. In the level $`𝗉=\sqrt{2w𝗇_𝗐+2s𝗇_𝗌}`$ we find twice the representation
$`({\displaystyle \frac{𝗇_𝗐}{2}},(0,𝗇_𝗌))({\displaystyle \frac{𝗇_𝗐}{2}},(1,𝗇_𝗌2))({\displaystyle \frac{𝗇_𝗐}{2}},(0,𝗇_𝗌3))`$
$`({\displaystyle \frac{𝗇_𝗐2}{2}},(0,𝗇_𝗌))({\displaystyle \frac{𝗇_𝗐2}{2}},(1,𝗇_𝗌2))({\displaystyle \frac{𝗇_𝗐2}{2}},(0,𝗇_𝗌3))`$
where the second irreducible representation in the direct sum appears only for $`𝗇_𝗌2`$, the third one for $`𝗇_𝗌3`$, the fourth one for $`𝗇_𝗐2`$, the fifth one for $`𝗇_𝗐2`$ and $`𝗇_𝗌2`$ and the sixth one only for $`𝗇_𝗐2`$ and $`𝗇_𝗌3`$. One copy of the total representation correspond to the sector $`\varsigma =+`$ the other one to the sector $`\varsigma =`$. Maximal weight states corresponding to all irreducible representations are easily constructed in terms of the states given above. All eigenstates carry hypercharge $`𝖸=2w𝗇_𝗐2s𝗇_𝗌`$ and correspond to the $`CP_{}`$ value $`(1)^{𝗇_𝗐+𝗇_𝗌}`$.
To large extracharge values correspond higher and higher –and presumably heavier– $`su(2)`$ and $`su(3)`$ representations. Simple isospin and color multiplets appear only in the very bottom of the spectrum. In particular, particle states carrying the correct hypercharge, isospin, color and chirality to be identified with known elementary matter fields all carry extracharge $`𝖷2`$. In Table 2 we report the lowest weight representations of the lower corner of the particle spectrum in the $`𝖷/𝖸`$ plane. Particles carrying quantum numbers of known elementary fermions are indicated by the name of the corresponding first generation particles. They organizes in a diamond at the very bottom of the spectrum. Unknown elementary fermions are indicated by a question mark. If we accept the idea that particles acquire mass in reason of the values of their extracharge, hypercharge and the complexity of their isospin and color representations, the diamond correspond to the lightest states of the spectrum.
The content of the first six rows of Table 2 coincides with the sector $`\varsigma =+`$ of Table 1 (plus an anti right-handed electron coming from the states $`𝗇_𝗐=0`$, $`𝗇_𝗌=3`$ that also have extracharge $`𝖷=1`$). Table 2 allows us to fully appreciate the fundamental symmetries of the particle operator. The spectrum contains exactly a family and an anti-family of known elementary fermions. These are organized inside the diamond in rows going from bottom-left to top-right as: leptons $`\nu _{eR}(\nu _{eL},e_L)e_R`$, anti-quarks $`d_R^{}(d_L,u_L)^{}u_R^{}`$, quarks $`u_R(u_L,d_L)d_R`$ and anti-leptons $`e_R^{}(e_L,\nu _{eL})^{}\nu _{eR}^{}`$.
Moving along bottom-left/top-right diagonals amounts to add a weak quanta to the state, $`(𝗇_𝗐,𝗇_𝗌)(𝗇_𝗐+1,𝗇_𝗌)`$ (eg$`\nu _{eR}`$ corresponds to $`𝗇_𝗐=0`$, $`𝗇_𝗌=0`$, $`(\nu _{eL},e_L)`$ to $`𝗇_𝗐=1`$, $`𝗇_𝗌=0`$ and $`e_R`$ to $`𝗇_𝗐=2`$, $`𝗇_𝗌=0`$; $`u_R`$ corresponds to $`𝗇_𝗐=0`$, $`𝗇_𝗌=2`$, $`(u_L,d_L)`$ to $`𝗇_𝗐=1`$, $`𝗇_𝗌=2`$ and $`d_R`$ to $`𝗇_𝗐=2`$, $`𝗇_𝗌=2`$; etc.). The relative infinitesimal generator is $`(𝖷+𝖸)/2`$. At a space-time level it produces a $`U(1)`$ rotation of all weak directions. Therefore, it is natural to expect that the effective breakdown of weak symmetry couples states aligned on bottom-left/top-right diagonals.
Moving in orthogonal directions –from bottom-right to top-left– amounts instead to add a strong quanta to the state, $`(𝗇_𝗐,𝗇_𝗌)(𝗇_𝗐,𝗇_𝗌+1)`$. The corresponding infinitesimal generator $`(𝖷𝖸)/2`$ produces a $`U(1)`$ rotation of all strong directions. If the color symmetry remains a good one, there is no reason to expect a mixing between states belonging to different bottom-left/top-right diagonals. Moreover, the exact conservations of the strong part of $`CP_{}=CP_{weak}CP_{strong}`$ further divides the spectrum in two sub-sectors corresponding to $`CP_{strong}=\pm 1`$. The lowest energy sector, $`CP_{strong}=+1`$, contains the first and third bottom-left/top-right diagonals; these correspond exactly to a fundamental family of elementary fermions –repeated in many generations. Observe that the qualitative expectation for masses goes in the correct order of neutrino, electron, quark up and quark down. The other sector, $`CP_{strong}=1`$, contains an anti-family.
### Higher particle states
New particles are presumably aligned along diagonals with a fixed value of $`𝗇_𝗌`$. The lightest new states correspond to $`𝗇_𝗌=0`$, $`2`$ and $`4`$. The simplest case is clearly the first, corresponding to isospin multiplets, color singlets. We report the content in multiplets of the first six states belonging to the diagonal $`𝗇_𝗌=0`$ in the plane $`𝖨/𝖸`$
Unknown particles are indicated by their electric charge after symmetry breaking. Solid lines indicate a possible pattern of symmetry breaking. Observe that the highest state with hypercharge $`𝖸=1`$, the lowest with $`𝖸=3/2`$ as well as the highest with $`𝖸=2`$ and the lowest with $`𝖸=5/2`$ are in the sector $`\varsigma =`$. If $`\varsigma `$ is a relevant selection rule the first new observable fermions are presumably a mixing of the higher $`𝖸=3/2`$ with the lower $`𝖸=2`$ states.
## 4 Particle-Energy relation, extraforce and $`\nu \nu _R`$
All what has been discussed in this paper is a direct consequence of a single basic assumption: local gauge invariance associated to all fundamental interactions reflects observer’s freedom of choosing a space-time reference frame in every point.
Given this, the gauge group associated to gravitational, electromagnetic, weak and strong interactions determines the local geometry of space-time –including its dimension. The geometrical framework naturally embodies degrees of freedom corresponding to gravitational and non-gravitational gauge interactions. In order to make a first connection between theory and observation, in this paper we considered the free motion of a test spinor in a given background. As a consequence of the local geometrical structure the fermion is reduced on a $`1+3`$ effective space-time; the effective four dimensional dynamics is described by chiral fermions carrying quantum numbers of known elementary particles –including generations; matter couples to gauge forces in the standard way. The fermionic sector of the Standard Model of Elementary Particle emerges as the low energy limit of the theory.
Of course, before claiming that we have a unified comprehension of fundamental interactions and elementary matter fields, there are two major issues that have to be investigated: the fundamental equations relating space-time geometry and matter distribution –the extension of Einstein equations– and the analysis of a natural mechanism producing electro-weak symmetry breaking. In the low energy limit, this should bring us to an understanding of the splitting of coupling constants associated to fundamental interactions as well as to a first principle prediction of masses and mixing angles. Both subjects are currently under investigation. For now, we observe that already at the stage considered in this paper the theory leads to important phenomenological consequences (other than the predictions of new particles and generations):
### Particle-Energy relation
In the very fast rotation of the fourteen dimensional spinor around ordinary space-time is stored an amazing amount of energy. This appears as a fourteen dimensional mass term in matter field equations. Therefore, in close analogy to the special relativistic picture associating to every body of mass $`m`$ an intrinsic energy $`E_m=mc^2`$, our theory predicts an even more fundamental relation: to every elementary particle $`\psi _𝗉`$ is associated an intrinsic energy
$$E_𝗉=\frac{\mathrm{}c}{l}𝗉$$
A fundamental length of the order of the Planck scale $`l10^{33}cm`$ associates to the electron an intrinsic energy of about $`10^{15}TeV`$. This enormous amount of energy stored in every elementary particle does not affect directly the effective four dimensional dynamics<sup>9</sup><sup>9</sup>9In many respects the situation is analogous to that of many adiabatic systems. To fix ideas we may think of the separation of vibrational and rotational nuclear freedoms in a molecule (with no symmetries). Most of the energy is stored in a vibrational state. In correspondence the molecule performs a free rotation. On rotational energy scales there is no direct perception of the amount of energy stored in vibrations. In addition, vibrational transitions are too rare to be appreciated on rotational energy scales.. However, it affects it indirectly. The ability of a particle of carrying extracharge, hypercharge, isospin and color is proportional to its intrinsic energy. A more striking consequence is that every elementary particle but $`\nu _R`$ is unstable. In the effective four dimensional theory, the conservation of extracharge, hypercharge, isospin and color prevent the system from collapsing down into the ground state. However, on very long time scales –or very high energies– we should expect these transitions to become relevant.
### Extraforce
Enforcing the gauge group of fundamental interactions at a space-time level indicates the existence of a further $`U(1)`$ gauge symmetry. This is presumably associated to an additional force mediated by an additional gauge boson. Predictions on the behavior of the extraforce and on the mass of the corresponding vector boson go necessarily through the extension of Einstein equation to the whole fourteen dimensional space-time and through the analysis of symmetry breaking.
### $`\nu _R`$
The theory predicts the existence of a particle carrying the right quantum numbers to be identified with the right-handed neutrino $`\nu _R`$. It is the ground state of the square of the particle operator $`𝒫`$; the only elementary particle carrying a vanishing intrinsic energy. In some sense, therefore, it represents the most elementary particle we can think of in four dimensions. The right-handed neutrino carries zero extracharge, zero hypercharge, no isospin and no color. It has no electromagnetic, weak, strong nor extra interactions. It only interacts through the gravitational field. It is natural to expect that after electro-weak symmetry breaking $`\nu _R`$ couples to $`\nu _L`$ originating the lightest massive particle. In addition it is natural to expect that almost the totality of the matter in the universe is in the ground state $`\nu _R`$. This could possibly give an answer to the dark matter problem.
## Considerations and Acknowledgments
The idea that the fermionic sector of the Standard Model could emerge as a low energy limit from an extension of space-time relativity came first to me when I realized that the effective motion of a generic dynamical system constrained on a sub-space by a scalar potential gets naturally coupled to an $`SO(n_1)\times SO(n_2)\times \mathrm{}`$ gauge potential. Dimension and number of $`SO(n)`$ factors depend on subspace codimension and symmetries of the confining potential (a beautiful example is found in the separation of vibrational and rotational freedoms in polyatomic molecules). A gauge group given by the direct product of orthogonal groups is not that different from a gauge group given by the direct product of unitary groups. The possibility that space-time dimensional reduction is produced by a scalar potential is unappealing. However, the property of magnetic fields of effectively reducing the motion along field lines seemed to match perfectly the unitary structure of non-gravitational gauge groups. I started working on the idea of a mixed real-complex geometry when I was ‘Bruno Rossi’ Post Doc fellow at MIT. There I first presented the idea of “Dimensional Reduction without Compactification” in a seminar in April 1998. I kept on working on the problem at KFKI-RMKI in Budapest –I take the chance to thank Prof. K. Tóth for that position– where I first obtained the fermionic spectrum summarized in Table 1. The work was reorganized and completed at the Physics Department of the University of Parma where I spent the winter 2000 as an external collaborator.
It is a genuine pleasure to thank my friends Roberto De Pietri, László Szabados and Péter Vecsernyés for all the help they gave me. Discussing with them the many aspects of the problem was precious to me.
|
warning/0006/math0006211.html
|
ar5iv
|
text
|
# Classification of Left-Covariant Differential Calculi on the Quantum Group SL_𝑞(2)This paper was supported by the Deutsche Forschungsgemeinschaft
## 1 Introduction
The theory of bicovariant differential calculus over Hopf algebras is one of the commonly used and best understood theories related to quantum groups. Its origin was a paper of S. L. Woronowicz where also left-covariant differential calculi were considered.
It is well known that bicovariant differential calculi on quantum groups often (but not always, see ,) have the unpleasant property (besides non-uniqueness) that their dimensions do not coincide with the dimensions of the canonical differential calculi of the corresponding Lie groups. There were made some attempts using a generalized adjoint action in order to circumvent this defect ,. Another way is to look for left-covariant differential calculi on the quantum group. Then the corresponding quantum tangent spaces are not invariant under the adjoint action in general. The first such example (the legendary 3D-calculus) was developed by S. L. Woronowicz for the quantum group $`\mathrm{SU}_q(2)`$. Among others it was shown therein that the cohomology spaces of the differential complex are the same as in the classical situation. Further examples of such kind were given by K. Schmüdgen and A. Schüler . The paper contains also a first classification of left-covariant differential calculi on the quantum group $`\mathrm{SL}_q(2)`$ (under very restrictive conditions). A method for the construction of left-covariant differential calculi on a quantum linear group was initiated by K. Schmüdgen . Since the pioneering work of Woronowicz the general theory of differential calculi on quantum groups was refined and developed further. An extensive overview can be found in Chapter 14 of the monograph .
In contrast to the classical situation there is no distinguished differential calculus on a quantum group and the non-commutative geometry of a quantum group depends on the differential calculus in general. Thus it seems to be natural to ask how many such calculi (satisfying other additional natural conditions) do really exist. This problem is studied in the present paper. Our aim is to give a step-by-step classification of left-covariant differential calculi $`(\mathrm{\Gamma },\mathrm{d})`$ on the quantum group $`\mathrm{SL}_q(2)`$ under rather general assumptions. In order to motivate the addition of further assumptions we investigate the outcome of the classification after each step. Although the methods we use can be easily described, the computations are very boring. The computer algebra program FELIX by J. Apel and U. Klaus was very helpful to carry out long computations. The main result of the present paper is Theorem 7. Suppose that $`q`$ is a nonzero complex number and not a root of unity. Then the assertion of Theorem 7 states that there are exactly 11 (single) left-covariant first order differential calculi over the Hopf algebra $`𝒪(\mathrm{SL}_q(2))`$ having the following properties: The quantum tangent space is a subspace of the algebra $`𝒰`$ (see Section 3), the one-forms $`\omega (u_2^1)`$, $`\omega (u_1^2)`$ and $`\omega (u_1^1u_2^2)`$ form a basis of the (necessarily 3-dimensional) $`𝒪(\mathrm{SL}_q(2))`$-bimodule $`\mathrm{\Gamma }`$, the dimension of the space of left-invariant differential 2-forms in the universal higher order differential calculus is at least 3 and the first order calculus is invariant with respect to all Hopf algebra automorphisms of $`𝒪(\mathrm{SL}_q(2))`$. The list of these calculi is given in Corollary 5.3. Using the method of Woronowicz it is proved in Theorem 8 that the dimensions of the cohomology spaces of these 11 differential complexes are the same as in the classical situation.
This paper is organized as follows. In Section 2 we recall some basic notions and facts about the general theory of left-covariant differential calculus on quantum groups. If not stated otherwise we follow the definitions and notations of Woronowicz and of the monograph . In Section 3 the structure of the dual Hopf algebra $`𝒰`$ of $`𝒪(\mathrm{SL}_q(2))`$ is described. In Section 4 we determine all 4-dimensional unital right coideals of $`𝒰`$. In Section 5 further restrictions on the calculus are added. In Section 6 we investigate additional structures such as $``$-structures and braidings. In Section 8 the cohomology spaces of the most important differential complexes are studied. Section 7 contains the main theorem (Theorem 7) of the present paper. The outcoming calculi are then studied in detail.
Throughout Sweedler’s notation for coproducts and coactions and Einsteins convention of summing over repeated indices are used. The symbols $``$ and $`_𝒜`$ denote tensor products over the complex numbers and over an algebra $`𝒜`$, respectively. All algebras are complex and unital.
## 2 Left-covariant differential calculi on quantum groups
First let us recall some facts of the general theory (see , ). Let $`𝒜`$ be a Hopf algebra with coproduct $`\mathrm{\Delta }`$, counit $`\epsilon `$, and invertible antipode $`S`$. An $`𝒜`$-bimodule $`\mathrm{\Gamma }`$ is called first order differential calculus (FODC for short) over $`𝒜`$, if there is a linear mapping $`\mathrm{d}:𝒜\mathrm{\Gamma }`$ such that
* $`\mathrm{d}`$ satisfies the Leibniz rule: $`\mathrm{d}(ab)=(\mathrm{d}a)b+a\mathrm{d}b`$ for any $`a,b𝒜`$,
* $`\mathrm{\Gamma }=\mathrm{Lin}\{a\mathrm{d}b|a,b𝒜\}`$.
An FODC $`\mathrm{\Gamma }`$ is called left-covariant if there is a linear mapping $`\mathrm{\Delta }_\mathrm{L}:\mathrm{\Gamma }𝒜\mathrm{\Gamma }`$ such that $`\mathrm{\Delta }_\mathrm{L}(a(\mathrm{d}b)c)=\mathrm{\Delta }(a)(\mathrm{id}\mathrm{d})\mathrm{\Delta }(b)\mathrm{\Delta }(c)`$, where $`(ab)(c\rho )=acb\rho `$ and $`(a\rho )(bc)=ab\rho c`$ for any $`a,b,c𝒜`$ and $`\rho \mathrm{\Gamma }`$. Elements $`\rho \mathrm{\Gamma }`$ for which $`\mathrm{\Delta }_\mathrm{L}(\rho )=1\rho `$ are called left-invariant. Because of Theorem 2.1 in any left-covariant $`𝒜`$-bimodule is a free left module and any basis of the vector space $`\mathrm{\Gamma }_\mathrm{L}`$ of left-invariant 1-forms is a free basis of the left (right) $`𝒜`$-module $`\mathrm{\Gamma }`$. The dimension of $`\mathrm{\Gamma }_\mathrm{L}`$ is called the dimension of the FODC $`\mathrm{\Gamma }`$. In this paper we are dealing only with finite dimensional FODC.
Suppose that $`\mathrm{\Gamma }`$ is an $`n`$-dimensional first order differential calculus over $`𝒜`$. Let us fix a basis $`\{\omega _i|i=1,\mathrm{},n\}`$ of $`\mathrm{\Gamma }_\mathrm{L}`$. Then there are functionals $`X_i`$, $`i=1,\mathrm{},n`$ in the dual Hopf algebra $`𝒜^{}`$ such that the differential $`\mathrm{d}`$ can be written in the form
$`\mathrm{d}a={\displaystyle \underset{i=1}{\overset{n}{}}}a_{(1)}X_i(a_{(2)})\omega _i,a𝒜.`$ (1)
Recall that $`𝒜^{}`$ is the set of all linear functionals $`f`$ on $`𝒜`$ for which there exist functionals $`f_1,\mathrm{},f_N,g_1,\mathrm{},g_N`$ on $`𝒜`$ such that $`f(ab)=_{i=1}^Nf_i(a)g_i(b)`$ for all $`a,b𝒜`$.
The vector space $`𝒳_\mathrm{\Gamma }:=\mathrm{Lin}\{X_i|i=1,\mathrm{},n\}`$ is called the quantum tangent space of the left-covariant FODC $`\mathrm{\Gamma }`$. We define a mapping $`\omega :𝒜\mathrm{\Gamma }_\mathrm{L}`$ by $`\omega (a)=S(a_{(1)})\mathrm{d}a_{(2)}`$. Then by (1) the equation
$`\omega (a)={\displaystyle \underset{i=1}{\overset{n}{}}}X_i(a)\omega _i,a𝒜`$ (2)
holds. Since $`\mathrm{d1}=0`$, we have $`\omega (1)=0`$ and $`X_i(1)=0`$ for any $`i`$. The following lemma is the starting point for the first part of our classification.
Lemma 1. If $`𝒳`$ is the quantum tangent space of a FODC $`\mathrm{\Gamma }`$, then $`\overline{𝒳}=𝒳\epsilon `$ is a unital right coideal of $`𝒜^{}`$ (i. e. $`\mathrm{\Delta }(\overline{𝒳})\overline{𝒳}𝒜^{}`$).
Conversely, any unital right coideal $`\overline{𝒳}`$ of $`𝒜^{}`$ determines a unique FODC with quantum tangent space $`𝒳^+:=\{X\overline{𝒳}|X(1)=0\}`$.
Proof. See and .
In particular, the coproduct of elements of the quantum tangent space takes the form
$`\mathrm{\Delta }X_i`$ $`=1X_i+X_jf_i^j,`$ (3)
where the functionals $`f_i^j𝒜^{}`$ describe the bimodule structure of $`\mathrm{\Gamma }`$:
$`\omega _ia`$ $`=a_{(1)}f_j^i(a_{(2)})\omega _j\text{for any }a𝒜\text{.}`$ (4)
Left-covariant first order differential calculi $`\mathrm{\Gamma }`$ over $`𝒜`$ are also characterized by the right ideal
$$_\mathrm{\Gamma }:=\{a\mathrm{ker}\epsilon 𝒜|\omega (a)=0\}=\{a𝒜|X(a)=0X\overline{𝒳}_\mathrm{\Gamma }\}$$
(5)
of $`𝒜`$. Two FODC $`(\mathrm{\Gamma }_1,\mathrm{d}_1)`$ and $`(\mathrm{\Gamma }_2,\mathrm{d}_2)`$ over $`𝒜`$ are called isomorphic, if $`_{\mathrm{\Gamma }_1}=_{\mathrm{\Gamma }_2}`$ or equivalently if $`𝒳_{\mathrm{\Gamma }_1}=𝒳_{\mathrm{\Gamma }_2}`$.
Let $`\mathrm{\Gamma }^k`$ denote the $`k`$-fold tensor product $`\mathrm{\Gamma }_𝒜\mathrm{}_𝒜\mathrm{\Gamma }`$ of the $`𝒜`$-bimodule $`\mathrm{\Gamma }`$, $`\mathrm{\Gamma }^0:=𝒜`$, $`\mathrm{\Gamma }^1:=\mathrm{\Gamma }`$ and $`\mathrm{\Gamma }^{}:=_{k=0}^{\mathrm{}}\mathrm{\Gamma }^k`$. Then $`\mathrm{\Gamma }^{}`$ becomes an algebra with multiplication $`_𝒜`$. Let $`𝒮`$ be a graded two-sided ideal in $`\mathrm{\Gamma }^{}`$, $`𝒮_{k=2}^{\mathrm{}}\mathrm{\Gamma }^k`$, $`𝒮=_{k=2}^{\mathrm{}}𝒮\mathrm{\Gamma }^k`$. Then the $`𝒜`$-bimodule $`\mathrm{\Gamma }^{}:=\mathrm{\Gamma }^{}/𝒮`$ as well as $`\mathrm{\Gamma }^k:=\mathrm{\Gamma }^k/(𝒮\mathrm{\Gamma }^k)`$ are well defined. The bimodule $`\mathrm{\Gamma }^{}`$ is called a differential calculus over the Hopf algebra $`𝒜`$ if there is a linear mapping $`\mathrm{d}:\mathrm{\Gamma }^{}\mathrm{\Gamma }^{}`$ of grade one (i. e. $`\mathrm{d}:\mathrm{\Gamma }^k\mathrm{\Gamma }^{k+1}`$) such that
* $`\mathrm{d}`$ satisfies the graded Leibniz rule $`\mathrm{d}(\rho \rho ^{})=\mathrm{d}\rho \rho ^{}+(1)^m\rho \mathrm{d}\rho ^{}`$ for $`\rho \mathrm{\Gamma }^m,\rho ^{}\mathrm{\Gamma }^{}`$,
* $`\mathrm{d}^2=0`$,
* $`\mathrm{\Gamma }=\mathrm{Lin}\{a\mathrm{d}b|a,b𝒜\}`$.
If $`\mathrm{\Gamma }`$ is left-covariant, then $`\mathrm{\Gamma }^{}`$ is also left-covariant with $`\mathrm{\Delta }_\mathrm{L}(\rho _𝒜\rho ^{})=\rho _{(1)}\rho _{(1)}^{}\rho _{(0)}_𝒜\rho _{(0)}^{}`$. Suppose that $`𝒮`$ is an invariant subspace of the left coaction, i. e. $`\mathrm{\Delta }_\mathrm{L}(𝒮)𝒜𝒮`$. Then $`\mathrm{\Gamma }^{}`$ inherits the left coaction of $`\mathrm{\Gamma }^{}`$ and $`\mathrm{\Gamma }^{}`$ is called a left-covariant differential calculus over $`𝒜`$.
Suppose that $`\mathrm{\Gamma }^{}`$ is a left-covariant differential calculus over $`𝒜`$. Then the Maurer-Cartan formula
$`\mathrm{d}\omega (a)=\omega (a_{(1)})\omega (a_{(2)}),a𝒜`$ (6)
is always fulfilled. Moreover, for any given left-covariant FODC $`\mathrm{\Gamma }`$ over $`𝒜`$ there exists a universal differential calculus $`{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{}`$. This means that any left-covariant differential calculus $`\stackrel{~}{\mathrm{\Gamma }}^{}`$ over $`𝒜`$ with $`\stackrel{~}{\mathrm{\Gamma }}^1=\mathrm{\Gamma }`$ is isomorphic to $`{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{}/\stackrel{~}{𝒮}`$, where $`\stackrel{~}{𝒮}`$ is a two-sided ideal in $`{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{}`$. The differential calculus $`{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{}`$ can be given by the two-sided ideal $`𝒮`$ generated by the elements of the vector space
$$𝒮_{}^{2}{}_{\mathrm{L}}{}^{}:=\mathrm{Lin}\{\omega (a_{(1)})_𝒜\omega (a_{(2)})|a\}.$$
(7)
Lemma 2. The following equation holds for any left-covariant differential calculus $`\mathrm{\Gamma }`$ over $`𝒜`$ with quantum tangent space $`𝒳`$:
$`dim({}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2})_\mathrm{L}=dim\{T\overline{𝒳}\overline{𝒳}|\mathrm{m}T=0\}dim𝒳,`$ (8)
where $`\mathrm{m}`$ denotes the multiplication map $`\mathrm{m}:\overline{𝒳}\overline{𝒳}𝒜^{}𝒜^{}𝒜^{}`$.
Proof. It was proved in that in the present situation the formula
$`dim({}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2})_\mathrm{L}=dim\{T𝒳𝒳|\mathrm{m}T𝒳\}`$ (9)
is valid. Let $`V,V^{}`$, and $`W`$ denote the vector spaces
$$V:=\{T𝒳𝒳|\mathrm{m}T𝒳\},V^{}:=\{T\overline{𝒳}𝒳|\mathrm{m}T=0\},\text{and}$$
$$W:=\{T\overline{𝒳}\overline{𝒳}|\mathrm{m}T=0\},$$
respectively. Then the mapping $`\phi :VV^{}`$, $`\phi (T)=T1\mathrm{m}T`$, and its inverse $`\psi :V^{}V`$, $`\psi (T)=T(\epsilon \mathrm{id})T`$ give an isomorphism between $`V`$ and $`V^{}`$. Obviously we have $`\overline{𝒳}\overline{𝒳}=(\overline{𝒳}𝒳)(\overline{𝒳}1)`$ and hence
$$dimW=dimV^{}+dimW^{},$$
where $`W^{}:=\{T^{}\overline{𝒳}|T\overline{𝒳}𝒳,\mathrm{m}(T^{}1T)=0\}`$. But the vector space $`W^{}`$ is isomorphic to $`𝒳`$. Indeed, $`𝒳W^{}`$ since $`\mathrm{m}(X11X)=0`$ but $`1W^{}`$. The latter follows from the fact that $`\epsilon (\mathrm{m}(11))=1`$ and $`\epsilon (\mathrm{m}(T))=0`$ for any $`T\overline{𝒳}𝒳`$.
## 3 The Hopf dual of $`𝓞\mathbf{(}\mathrm{𝐒𝐋}_𝒒\mathbf{(}\mathrm{𝟐}\mathbf{)}\mathbf{)}`$
In what follows we assume that $`q`$ is a transcendental complex number. The structure of the coordinate Hopf algebra $`𝒪(\mathrm{SL}_q(2))`$ (with generators $`u_j^i`$, $`i,j=1,2`$) of the quantum group $`\mathrm{SL}_q(2)`$ is well known. We now restate the description of the Hopf dual $`𝒰=𝒪(\mathrm{SL}_q(2))^{}`$ obtained in the monograph . Let $`𝒰`$ denote the unital algebra generated by the elements $`E,F,G`$ and $`f_\mu `$ ($`\mu ^\times =\{0\}`$) and by the relations
$$\begin{array}{c}f_\mu f_\nu =f_{\mu \nu },f_\mu E=\mu ^2Ef_\mu ,f_\mu F=\mu ^2Ff_\mu ,f_\mu G=Gf_\mu ,\\ GE=E(G+2),GF=F(G2),EFFE=\frac{f_qf_{q^1}}{qq^1}.\end{array}$$
(10)
The element $`f_1`$ is the unit in the algebra $`𝒰`$. The element $`f_1`$ is also denoted by $`\epsilon _{}`$.
Let us fix one square root $`q^{1/2}`$ of $`q`$ and define $`K=f_{q^{1/2}}`$. Then there is a Hopf algebra structure on $`𝒰`$ such that
$`\mathrm{\Delta }(E)`$ $`=EK+K^1E,`$ $`\epsilon (E)`$ $`=0,`$ $`S(E)`$ $`=qE,`$ (11)
$`\mathrm{\Delta }(F)`$ $`=FK+K^1F,`$ $`\epsilon (F)`$ $`=0,`$ $`S(F)`$ $`=q^1F,`$
$`\mathrm{\Delta }(G)`$ $`=1G+G1,`$ $`\epsilon (G)`$ $`=0,`$ $`S(G)`$ $`=G,`$
$`\mathrm{\Delta }(f_\mu )`$ $`=f_\mu f_\mu ,`$ $`\epsilon (f_\mu )`$ $`=1,`$ $`S(f_\mu )`$ $`=f_{\mu ^1}.`$
To make calculations easier we use the notation $`F^{(k)}:=F^kK^k/[k]!`$, $`E^{(k)}:=K^kE^k/[k]!`$ and $`G^{(k)}=G^k/k!`$ for any $`k_0`$, where $`[k]!=[k][k1]\mathrm{}[1]`$, $`[0]!=1`$ and $`[k]=(q^kq^k)/(qq^1)`$. The coproducts of these elements are
$`\mathrm{\Delta }(F^{(k)})`$ $`={\displaystyle \underset{r=0}{\overset{k}{}}}F^{(kr)}K^{2r}F^{(r)},`$ (12)
$`\mathrm{\Delta }(E^{(k)})`$ $`={\displaystyle \underset{s=0}{\overset{k}{}}}K^{2s}E^{(ks)}E^{(s)},`$
$`\mathrm{\Delta }(G^{(k)})`$ $`={\displaystyle \underset{t=0}{\overset{k}{}}}G^{(kt)}G^{(t)}.`$
The dual pairing $`,`$ of the Hopf algebras $`𝒰=𝒪(\mathrm{SL}_q(2))^{}`$ and $`𝒪(\mathrm{SL}_q(2))`$ is given by the matrices $`,u_j^i`$, where
$$E=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right),F=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right),G=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),f_\mu =\left(\begin{array}{cc}\mu ^1& 0\\ 0& \mu \end{array}\right).$$
(13)
For the algebra $`𝒰`$ a PBW-like theorem holds: the elements of the set $`\{F^{(i)}f_\mu E^{(j)}G^{(k)}|i,j,k_0,\mu ^\times \}`$ form a vector space basis of the algebra $`𝒰`$.
Remark. That the Hopf algebra $`𝒰`$ defined above is the full Hopf dual of the Hopf algebra $`𝒪(\mathrm{SL}_q(2))`$ is proved in the monograph for transcendental $`q`$. The assumption that $`q`$ is transcendental is only necessary in order to apply this theorem. If this fact holds already for $`q`$ not a root of unity, then all results of the present paper remain valid under this assumption. In what follows, we classify the corresponding left-covariant differential calculi whose quantum tangent space are contained in $`𝒰`$. All these considerations hold if $`q0`$ and $`q`$ is not a root of unity.
## 4 Unital right coideals of $`𝓤`$
By Lemma 2 left-covariant first order differential calculi over the Hopf algebra $`𝒜:=𝒪(\mathrm{SL}_q(2))`$ and unital right coideals of $`𝒜^{}`$ are in one-to-one correspondence. Now we determine all unital right coideals of $`𝒰`$ of dimension $`4`$. In this way we decribe all left-covariant first order differential calculi over $`𝒜`$ of dimension less than four.
For $`X=F^{(i)}f_\mu E^{(j)}G^{(k)}`$, $`i,j,k_0`$, $`\mu ^\times `$ we set
$`_1(X):=i,_2(X):=\mu ,_3(X):=j,_4(X):=k,_{13}(X):=i+j.`$
For a finite linear combination $`X=_{i=1}^na_iX_i`$ of such elements we define $`_m(X):=\mathrm{max}\{_m(X_i)|i=1,\mathrm{},n\}`$ for $`m\{1,3,4,13\}`$. Then from the PBW-theorem we conclude that $`𝒰=_{\mu ^\times }[𝒰]_\mu `$, where
$$[𝒰]_\mu :=\mathrm{Lin}\{F^{(i)}f_\mu E^{(j)}G^{(k)}|i,j,k_0\}.$$
Proposition 3. Let $`𝒳`$ be a right coideal of $`𝒰`$ and let $`X𝒳`$. Then there exist complex numbers $`\alpha _{i,\nu ,j,k}`$ ($`\nu ^\times `$, $`i,j,k_0`$) such that $`X=_{i,\nu ,j,k}\alpha _{i,\nu ,j,k}F^{(i)}f_\nu E^{(j)}G^{(k)}`$. Let us fix $`\mu ^\times `$ and $`r,s,t_0`$. Consider the element
$`X_{r,\mu ,s,t}`$ $`:={\displaystyle \underset{i,j,k}{}}\alpha _{i,\mu ,j,k}F^{(ir)}f_{q^{rs}\mu }E^{(js)}G^{(kt)}`$ (14)
of $`𝒰`$, where the sum is running over all $`i,j,k_0`$ with $`ir`$, $`js`$ and $`kt`$. Then the vector space $`\mathrm{Lin}\{X_{r\mu st}|r,s,t_0,\mu ^\times \}`$ is the smallest (with respect to inclusion) right coideal of $`𝒰`$ containing $`X`$.
Proof. Using (12) we compute the coproduct of $`X`$ and obtain
$`\mathrm{\Delta }\left({\displaystyle \underset{i\nu jk}{}}\alpha _{i\nu jk}F^{(i)}f_\nu E^{(j)}G^{(k)}\right)=`$
$`={\displaystyle \underset{i\nu jk}{}}\alpha _{i\nu jk}{\displaystyle \underset{rst}{}}F^{(ir)}K^{2r}f_\nu K^{2s}E^{(js)}G^{(kt)}F^{(r)}f_\nu E^{(s)}G^{(t)}`$
$`={\displaystyle \underset{r\nu st}{}}{\displaystyle \underset{ijk}{}}\alpha _{i\nu jk}F^{(ir)}f_{q^{rs}\nu }E^{(js)}G^{(kt)}F^{(r)}f_\nu E^{(s)}G^{(t)}`$
$`={\displaystyle \underset{r\nu st}{}}X_{r\nu st}F^{(r)}f_\nu E^{(s)}G^{(t)}.`$
Because of the PBW-theorem the elements of the right hand side of the tensor product are linearly independent. Hence the elements $`X_{r\nu st}`$ belong to any right coideal containing $`X`$ (i. e. they belong to $`𝒳`$ too). Observe that $`X=_{\mu ^\times }X_{0\mu 00}`$. Taking in account that $`(\mathrm{\Delta }\mathrm{id})\mathrm{\Delta }(X)=(\mathrm{id}\mathrm{\Delta })\mathrm{\Delta }(X)`$ it follows that the vector space $`\mathrm{Lin}\{X_{r\mu st}|r,s,t_0,\mu ^\times \}`$ is a right coideal of $`𝒰`$.
Since $`X=_{\mu ^\times }X_{0\mu 00}`$ we obtain the following.
Corollary 4. Any right coideal $`𝒳`$ of $`𝒰`$ is isomorphic to the direct sum $`_{\mu ^\times }[𝒳]_\mu `$ of its homogeneous components $`[𝒳]_\mu :=𝒳[𝒰]_\mu `$.
What can be said about the dimension of a right coideal $`𝒳`$ of $`𝒰`$? Let $`X`$ be a nonzero element of $`𝒳`$. By Corollary 4 we may assume without loss of generality that there is a $`\mu ^\times `$ such that $`X[𝒳]_\mu `$. Due to the PBW-theorem there is for any $`t_0`$, $`t_4(X)`$ a unique $`X_t𝒰`$ such that $`X=_{t=0}^{_4(X)}X_tG^{(t)}`$ and $`X_{_4(X)}0`$. Let now $`p_0`$, $`p_{13}(X)`$. We define
$`t_{13}(p)=\mathrm{max}\{m|X_m0,_{13}(X_m)p\}.`$
Then the coefficients of $`X_{r\mu st}`$ in Proposition 4 indicate that for any $`t_0`$, $`tt_{13}(p)`$ there is at least one number $`r(t,p)_0`$, $`r(t,p)p`$ such that $`X_{r(t,p),\mu ,pr(t,p),t}0`$. Since $`_2(X_{r(t,p),\mu ,pr(t,p),t})=\mu q^p`$, $`q`$ is not a root of unity and $`_4(X_{r(t,p),\mu ,pr(t,p),t})=t_{13}(p)t`$ we conclude that the elements $`X_{r(t,p),\mu ,pr(t,p),t}`$, $`0p_{13}(X)`$, $`0tt_{13}(p)`$ are linearly independent. Hence the number
$`{\displaystyle \underset{p=0}{\overset{_{13}(X)}{}}}(t_{13}(p)+1)={\displaystyle \underset{t=0}{\overset{_4(X)}{}}}(s_{13}(t)+1)`$ (15)
with $`s_{13}(t)=\mathrm{max}\{_{13}(X_m)|mt,X_m0\}`$ for $`t_0`$, $`t_4(X)`$, is a lower bound for the dimension of the right coideal $`𝒳`$.
Example 1. Suppose that $`X=K^6G^{(5)}+(F^{(1)}K^6K^6)G^{(4)}+(F^{(2)}K^6E^{(1)}+K^6E^{(3)})G^{(2)}+F^{(3)}K^6`$ is an element of a right coideal $`𝒳`$ of $`𝒰`$. We have $`X[𝒰]_{q^3}`$, $`_4(X)=5`$ and $`_{13}(X)=3`$. The coefficients $`X_k`$ of $`G^{(k)}`$ in $`X`$ are $`X_0=F^{(3)}K^6`$, $`X_2=F^{(2)}K^6E^{(1)}+K^6E^{(3)}`$, $`X_4=F^{(1)}K^6K^6`$, $`X_5=K^6`$ and $`X_t=0`$ otherwise. The values of the functions $`t_{13}(n)`$ ($`s_{13}(n)`$) are $`5,4,2,2`$ for $`n=0,1,2,3`$ ($`3,3,3,1,1,0`$ for $`n=0,1,2,3,4,5`$). Hence the dimension of each right coideal $`𝒳`$ containing $`X`$ is at least 17.
In the remaining part of this section we determine all unital right coideals of $`𝒰`$ of dimension $`4`$.
### 4.1 $`𝐝𝐢𝐦𝓧\mathbf{}\mathrm{𝟐}`$
The following list contains all possibilities (if not otherwise stated, parameters are arbitrary complex numbers):
* $`𝒳_1^1=1`$.
* $`𝒳_1^2=\mathrm{Lin}\{G,1\}`$,
* $`𝒳_2^2=\mathrm{Lin}\{F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$,
* $`𝒳_3^2=\mathrm{Lin}\{K^2E^{(1)}+\alpha K^2,1\}`$,
* $`𝒳_4^2=\mathrm{Lin}\{f_\mu ,1\}`$, $`\mu ^\times `$, $`\mu 1`$.
Obviously $`dim𝒳=1`$ and $`1𝒳`$ imply that the only 1-dimensional unital right coideal of $`𝒰`$ is $`𝒳_1^1`$.
Suppose that $`dim𝒳=2`$. By (15) we have $`_4(X)<dim𝒳=2`$ for any $`X𝒳`$. If there is an $`X[𝒳]_\mu `$, $`\mu ^\times `$ with $`_4(X)=1`$ then we must have $`s_{13}(0)=s_{13}(1)=0`$ by (15) and we obtain $`X=\alpha f_\mu G+\beta f_\mu `$, $`\alpha 0`$. The only nonzero elements $`X_{r\nu st}`$ are $`X`$ and $`X_{0\mu 01}=\alpha f_\mu `$. Since $`dim𝒳=2`$ and $`1𝒳`$, $`1`$ must be one of those two elements. Therefore, $`f_\mu =1`$, i. e. $`\mu =1`$. Hence $`𝒳`$ is isomorphic to $`𝒳_1^2`$.
If $`_4(X)=0`$ for any $`X𝒳`$ then by (15) $`s_{13}(0)`$ can take the values $`1`$ and $`0`$. In the first case we have $`1=s_{13}(0)=_{13}(X)`$ from which $`X=\alpha F^{(1)}f_\mu +\beta E^{(1)}f_\mu +\gamma f_\mu `$ ($`\alpha 0`$ or $`\beta 0`$) follows. Then $`X_{1\mu 00}=\alpha K^2f_\mu `$, $`X_{0\mu 01}=\beta K^2f_\mu `$ and therefore $`K^2f_\mu 𝒳`$. Since $`1𝒳`$, we must have $`f_\mu =K^2`$. This gives the coideals $`𝒳_2^2`$ and $`𝒳_3^2`$. In the remaining case we have $`_{13}(X)=0`$ and we get $`𝒳_4^2`$.
### 4.2 $`𝐝𝐢𝐦𝓧\mathbf{=}\mathrm{𝟑}`$
By (15), $`_4(X)2`$ for any $`X𝒳`$. If $`_4(X)=2`$ then $`s_{13}(i)=0`$ for any $`i=0,1,2`$. If $`_4(X)=1`$ then $`s_{13}(1)=0`$ and $`s_{13}(0)`$ can take the values $`0`$ and $`1`$. Finally, if $`_4(X)=0`$ then $`s_{13}(0)`$ can be $`2,1`$ or $`0`$. We obtain the following list of unital right coideals:
* $`𝒳_1^3=\mathrm{Lin}\{G^2,G,1\}`$,
* $`𝒳_2^3=\mathrm{Lin}\{K^2G+\alpha F^{(1)}K^2+\beta K^2E^{(1)},K^2,1\}`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_3^3=\mathrm{Lin}\{G+\alpha F^{(1)}+\beta E^{(1)},K^2,1\}`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_4^3=\mathrm{Lin}\{f_\mu G,f_\mu ,1\}`$, $`\mu ^\times `$, $`\mu 1`$,
* $`𝒳_5^3=\mathrm{Lin}\{G,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$,
* $`𝒳_6^3=\mathrm{Lin}\{G,K^2E^{(1)}+\alpha K^2,1\}`$,
* $`𝒳_7^3=\mathrm{Lin}\{G,f_\mu ,1\}`$, $`\mu ^\times `$, $`\mu 1`$,
* $`𝒳_8^3=\mathrm{Lin}\{F^{(2)}K^4+\alpha F^{(1)}K^4E^{(1)}+\alpha ^2K^4E^{(2)}+\beta F^{(1)}K^4+\alpha \beta K^4E^{(1)}+\gamma K^4,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$,
* $`𝒳_9^3=\mathrm{Lin}\{K^4E^{(2)}+\alpha K^4E^{(1)}+\beta K^4,K^2E^{(1)}+\alpha K^2,1\}`$,
* $`𝒳_{10}^3=\mathrm{Lin}\{F^{(1)}f_\mu +\alpha f_\mu E^{(1)}+\beta f_\mu ,f_{\mu q^1},1\}`$, $`\mu q`$,
* $`𝒳_{11}^3=\mathrm{Lin}\{f_\mu E^{(1)}+\alpha f_\mu ,f_{\mu q^1},1\}`$, $`\mu q`$,
* $`𝒳_{12}^3=\mathrm{Lin}\{F^{(1)}K^2+\alpha K^2,K^2E^{(1)}+\beta K^2,1\}`$,
* $`𝒳_{13}^3=\mathrm{Lin}\{F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{14}^3=\mathrm{Lin}\{K^2E^{(1)}+\alpha K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{15}^3=\mathrm{Lin}\{f_\mu ,f_\nu ,1\}`$, $`\mu \nu `$, $`\mu 1`$, $`\nu 1`$.
### 4.3 $`𝐝𝐢𝐦𝓧\mathbf{=}\mathrm{𝟒}`$
Finally we give the complete list of 4-dimensional unital right coideals of $`𝒰`$. For the proof of the completeness of the list the method explained above is used.
* $`𝒳_1^4=\mathrm{Lin}\{G^{(3)},G^{(2)},G,1\}`$,
* $`𝒳_2^4=\mathrm{Lin}\{K^2G^{(2)}+\alpha F^{(1)}K^2+\beta K^2E^{(1)},K^2G,K^2,1\}`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_3^4=\mathrm{Lin}\{G^{(2)}+\alpha F^{(1)}+\beta E^{(1)},G,K^2,1\}`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_4^4=\mathrm{Lin}\{f_\mu G^{(2)},f_\mu G,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_5^4=\mathrm{Lin}\{G^{(2)},G,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$,
* $`𝒳_6^4=\mathrm{Lin}\{G^{(2)},G,K^2E^{(1)}+\alpha K^2,1\}`$,
* $`𝒳_7^4=\mathrm{Lin}\{G^{(2)},G,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_8^4=\mathrm{Lin}\{F^{(1)}K^2G+\alpha K^2E^{(1)}G+\beta K^2G+\gamma K^2E^{(1)}+\delta K^2,G,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$,
(LI) $`\gamma 2\alpha `$
* $`𝒳_9^4=\mathrm{Lin}\{K^2E^{(1)}G+\alpha K^2G+\beta F^{(1)}K^2+\gamma K^2,G,K^2E^{(1)}+\alpha K^2,1\}`$,
(LI) $`\beta 0`$
* $`𝒳_{10}^4=\mathrm{Lin}\{K^4G+\alpha F^{(2)}K^4+\alpha \beta F^{(1)}K^4E^{(1)}+\alpha \beta ^2K^4E^{(2)}+\alpha \gamma F^{(1)}K^4+\alpha \beta \gamma K^4E^{(1)},F^{(1)}K^2+\beta K^2E^{(1)}+\gamma K^2,K^4,1\}`$, $`\alpha 0`$,
* $`𝒳_{11}^4=\mathrm{Lin}\{K^4G+\alpha K^4E^{(2)}+\alpha \beta K^4E^{(1)},K^2E^{(1)}+\beta K^2,K^4,1\}`$, $`\alpha 0`$,
* $`𝒳_{12}^4=\mathrm{Lin}\{G+\alpha F^{(2)}+\alpha \beta F^{(1)}E^{(1)}+\alpha \beta ^2E^{(2)}+\alpha \gamma F^{(1)}+\alpha \beta \gamma E^{(1)},F^{(1)}K^2+\beta K^2E^{(1)}+\gamma K^2,K^4,1\}`$, $`\alpha 0`$,
* $`𝒳_{13}^4=\mathrm{Lin}\{G+\alpha E^{(2)}+\alpha \beta E^{(1)},K^2E^{(1)}+\beta K^2,K^4,1\}`$, $`\alpha 0`$,
* $`𝒳_{14}^4=\mathrm{Lin}\{f_\mu G+\alpha F^{(1)}f_\mu +\beta f_\mu E^{(1)},f_\mu ,f_{\mu q^1},1\}`$,
$`\mu 1`$, $`\mu q`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_{15}^4=\mathrm{Lin}\{K^2G+\alpha F^{(1)}K^2+\beta K^2E^{(1)},G,K^2,1\}`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_{16}^4=\mathrm{Lin}\{K^2G+\alpha F^{(1)}K^2+\beta K^2E^{(1)},F^{(1)}K^4+\gamma K^4E^{(1)}+\delta K^4,K^2,1\}`$, $`|\alpha |+|\beta |0`$,
(LI) $`\beta \alpha \gamma `$,
* $`𝒳_{17}^4=\mathrm{Lin}\{K^2G+\alpha F^{(1)}K^2+\beta K^2E^{(1)},K^4E^{(1)}+\gamma K^4,K^2,1\}`$,
$`|\alpha |+|\beta |0`$,
(LI) $`\alpha 0`$,
* $`𝒳_{18}^4=\mathrm{Lin}\{K^2G+\alpha K^2E^{(1)},F^{(1)}K^2+\beta K^2E^{(1)},K^2,1\}`$, $`\alpha 0`$,
(LI)
* $`𝒳_{19}^4=\mathrm{Lin}\{K^2G+\alpha F^{(1)}K^2,K^2E^{(1)},K^2,1\}`$, $`\alpha 0`$,
(LI)
* $`𝒳_{20}^4=\mathrm{Lin}\{K^2G+\alpha F^{(1)}K^2+\beta K^2E^{(1)},f_\mu ,K^2,1\}`$, $`|\alpha |+|\beta |0`$, $`\mu 1`$, $`\mu q`$,
* $`𝒳_{21}^4=\mathrm{Lin}\{G+\alpha F^{(1)}+\beta E^{(1)},K^2G,K^2,1\}`$, $`|\alpha |+|\beta |0`$,
* $`𝒳_{22}^4=\mathrm{Lin}\{G+\alpha E^{(1)},F^{(1)}+\beta E^{(1)},K^2,1\}`$, $`\alpha 0`$,
(LI)
* $`𝒳_{23}^4=\mathrm{Lin}\{G+\alpha F^{(1)},E^{(1)},K^2,1\}`$, $`\alpha 0`$,
(LI)
* $`𝒳_{24}^4=\mathrm{Lin}\{G+\alpha F^{(1)}+\beta E^{(1)},F^{(1)}K^2+\gamma K^2E^{(1)}+\delta K^2,K^2,1\}`$,
$`|\alpha |+|\beta |0`$,
(LI) $`\beta \alpha \gamma `$
* $`𝒳_{25}^4=\mathrm{Lin}\{G+\alpha F^{(1)}+\beta E^{(1)},K^2E^{(1)}+\gamma K^2,K^2,1\}`$, $`|\alpha |+|\beta |0`$,
(LI) $`\alpha 0`$
* $`𝒳_{26}^4=\mathrm{Lin}\{G+\alpha F^{(1)}+\beta E^{(1)},f_\mu ,K^2,1\}`$, $`\mu 1`$, $`\mu q^1`$,
* $`𝒳_{27}^4=\mathrm{Lin}\{f_\mu G,G,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{28}^4=\mathrm{Lin}\{f_\mu G,F^{(1)}f_{\mu q}+\alpha f_{\mu q}E^{(1)}+\beta f_{\mu q},f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{29}^4=\mathrm{Lin}\{f_\mu G,f_{\mu q}E^{(1)}+\alpha f_{\mu q},f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{30}^4=\mathrm{Lin}\{f_\mu G,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{31}^4=\mathrm{Lin}\{f_\mu G,K^2E^{(1)}+\alpha K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{32}^4=\mathrm{Lin}\{f_\mu G,f_\mu ,f_\nu ,1\}`$, $`\mu \nu `$, $`\mu 1`$, $`\nu 1`$,
* $`𝒳_{33}^4=\mathrm{Lin}\{G,F^{(2)}K^4+\alpha F^{(1)}K^4E^{(1)}+\alpha ^2K^4E^{(2)}+\beta F^{(1)}K^4+\alpha \beta K^4E^{(1)}+\gamma K^4,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$, $`\alpha 0`$,
* $`𝒳_{34}^4=\mathrm{Lin}\{G,K^4E^{(2)}+\alpha K^4E^{(1)}+\beta K^4,K^2E^{(1)}+\alpha K^2,1\}`$,
* $`𝒳_{35}^4=\mathrm{Lin}\{G,F^{(1)}f_\mu +\alpha f_\mu E^{(1)}+\beta f_\mu ,f_{\mu q^1},1\}`$, $`\mu q`$,
* $`𝒳_{36}^4=\mathrm{Lin}\{G,f_\mu E^{(1)}+\alpha f_\mu ,f_{\mu q^1},1\}`$, $`\mu q`$,
* $`𝒳_{37}^4=\mathrm{Lin}\{G,F^{(1)}K^2+\alpha K^2,K^2E^{(1)}+\beta K^2,1\}`$,
(LI)
* $`𝒳_{38}^4=\mathrm{Lin}\{G,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{39}^4=\mathrm{Lin}\{G,K^2E^{(1)}+\alpha K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{40}^4=\mathrm{Lin}\{G,f_\mu ,f_\nu ,1\}`$, $`\mu \nu `$, $`\mu 1`$, $`\nu 1`$,
* $`𝒳_{41}^4=\mathrm{Lin}\{F^{(3)}K^6+\alpha F^{(2)}K^6E^{(1)}+\alpha ^2F^{(1)}K^6E^{(2)}+\alpha ^3K^6E^{(3)}+\beta F^{(2)}K^6+\alpha \beta F^{(1)}K^6E^{(1)}+\alpha ^2\beta K^6E^{(2)}+\gamma F^{(1)}K^6+\alpha \gamma K^6E^{(1)}+\delta K^6,F^{(2)}K^4+\alpha F^{(1)}K^4E^{(1)}+\alpha ^2K^4E^{(2)}+\beta F^{(1)}K^4+\alpha \beta K^4E^{(1)}+\gamma K^4,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,1\}`$,
* $`𝒳_{42}^4=\mathrm{Lin}\{K^6E^{(3)}+\alpha K^6E^{(2)}+\beta K^6E^{(1)}+\gamma K^6,K^4E^{(2)}+\alpha K^4E^{(1)}+\beta K^4,K^2E^{(1)}+\alpha K^2,1\}`$,
* $`𝒳_{43}^4=\mathrm{Lin}\{F^{(2)}f_\mu +\alpha F^{(1)}f_\mu E^{(1)}+\alpha ^2f_\mu E^{(2)}+\beta F^{(1)}f_\mu +\alpha \beta f_\mu E^{(1)}+\gamma f_\mu ,F^{(1)}f_{\mu q^1}+\alpha f_{\mu q^1}E^{(1)}+\beta f_{\mu q^1},f_{\mu q^2},1\}`$, $`\mu q^2`$,
* $`𝒳_{44}^4=\mathrm{Lin}\{f_\mu E^{(2)}+\alpha f_\mu E^{(1)}+\beta f_\mu ,f_{\mu q^1}E^{(1)}+\alpha f_{\mu q^1},f_{\mu q^2},1\}`$, $`\mu q^2`$,
* $`𝒳_{45}^4=\mathrm{Lin}\{F^{(2)}K^4+\alpha _1F^{(1)}K^4E^{(1)}+\alpha _2K^4E^{(2)}+(\beta _1+\alpha _1\beta _2)F^{(1)}K^4+(\alpha _1\beta _1+\alpha _2\beta _2)K^4E^{(1)}+\gamma K^4,F^{(1)}K^2+\beta _1K^2,K^2E^{(1)}+\beta _2K^2,1\}`$, $`\alpha _2\alpha _1^2`$,
(LI) $`(q^2q^2)\gamma q\alpha _1+(q^21)(\beta _1^2+2\alpha _1\beta _1\beta _2+\alpha _2\beta _2^2)`$
* $`𝒳_{46}^4=\mathrm{Lin}\{F^{(2)}K^4+\alpha F^{(1)}K^4E^{(1)}+\alpha ^2K^4E^{(2)}+\beta F^{(1)}K^4+\gamma K^4E^{(1)}+\delta K^4,F^{(1)}K^2+\alpha K^2E^{(1)},K^2,1\}`$, $`\gamma \alpha \beta `$,
(LI)
* $`𝒳_{47}^4=\mathrm{Lin}\{F^{(1)}K^4E^{(1)}+\alpha K^4E^{(2)}+\beta F^{(1)}K^4+(\gamma +\alpha \beta )K^4E^{(1)}+\delta K^4,F^{(1)}K^2+\gamma K^2,K^2E^{(1)}+\beta K^2,1\}`$,
(LI) $`(q^2q^2)\delta q+(q^21)\beta (2\gamma +\alpha \beta )`$
* $`𝒳_{48}^4=\mathrm{Lin}\{K^4E^{(2)}+\alpha F^{(1)}K^4+\beta K^4E^{(1)}+\gamma K^4,K^2E^{(1)},K^2,1\}`$, $`\alpha 0`$,
(LI)
* $`𝒳_{49}^4=\mathrm{Lin}\{F^{(2)}K^4+\alpha F^{(1)}K^4E^{(1)}+\alpha ^2K^4E^{(2)}+\beta F^{(1)}K^4+\alpha \beta K^4E^{(1)}+\gamma K^4,F^{(1)}K^2+(\beta \alpha \delta )K^2,K^2E^{(1)}+\delta K^2,1\}`$,
(LI) $`\alpha (qq^3)\gamma (qq^1)\beta ^2`$
* $`𝒳_{50}^4=\mathrm{Lin}\{F^{(2)}K^4+\alpha F^{(1)}K^4E^{(1)}+\alpha ^2K^4E^{(2)}+\beta F^{(1)}K^4+\alpha \beta K^4E^{(1)}+\gamma K^4,F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{51}^4=\mathrm{Lin}\{K^4E^{(2)}+\alpha K^4E^{(1)}+\beta K^4,F^{(1)}K^2+\gamma K^2,K^2E^{(1)}+\alpha K^2,1\}`$,
(LI) $`(1+q^2)\beta \alpha ^2`$
* $`𝒳_{52}^4=\mathrm{Lin}\{K^4E^{(2)}+\alpha K^4E^{(1)}+\beta K^4,K^2E^{(1)}+\alpha K^2,f_\mu ,1\}`$, $`\mu 1`$,
* $`𝒳_{53}^4=\mathrm{Lin}\{F^{(1)}f_\mu +\alpha f_\mu E^{(1)}+\beta f_\mu ,F^{(1)}K^2+\gamma K^2E^{(1)}+\delta K^2,f_{\mu q^1},1\}`$, $`\mu q`$,
(LI) $`\gamma \alpha `$, $`\mu ^2q^2`$
* $`𝒳_{54}^4=\mathrm{Lin}\{F^{(1)}f_\mu +\alpha f_\mu E^{(1)}+\beta f_\mu ,K^2E^{(1)}+\gamma K^2,f_{\mu q^1},1\}`$, $`\mu q`$,
(LI) $`\mu ^2q^2`$
* $`𝒳_{55}^4=\mathrm{Lin}\{F^{(1)}f_\mu +\alpha f_\mu ,f_\mu E^{(1)}+\beta f_\mu ,f_{\mu q^1},1\}`$, $`\mu q`$,
(LI) $`\mu ^2q^2`$
* $`𝒳_{56}^4=\mathrm{Lin}\{F^{(1)}f_\mu +\alpha f_\mu E^{(1)}+\beta f_\mu ,f_{\mu q^1},f_\nu ,1\}`$, $`\mu q`$, $`\mu q\nu `$, $`\nu 1`$,
* $`𝒳_{57}^4=\mathrm{Lin}\{F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,f_\mu E^{(1)}+\gamma f_\mu ,f_{\mu q^1},1\}`$, $`\mu q`$,
(LI) $`\mu ^2q^2`$
* $`𝒳_{58}^4=\mathrm{Lin}\{F^{(1)}K^2+\alpha K^2,K^2E^{(1)}+\beta K^2,f_\mu ,1\}`$, $`\mu 1`$,
(LI) $`\mu ^21`$
* $`𝒳_{59}^4=\mathrm{Lin}\{F^{(1)}K^2+\alpha K^2E^{(1)}+\beta K^2,f_\mu ,f_\nu ,1\}`$, $`\mu \nu `$, $`\mu 1`$, $`\nu 1`$,
* $`𝒳_{60}^4=\mathrm{Lin}\{K^2E^{(1)}+\alpha K^2,f_\mu ,f_\nu ,1\}`$, $`\mu \nu `$, $`\mu 1`$, $`\nu 1`$,
* $`𝒳_{61}^4=\mathrm{Lin}\{f_\mu E^{(1)}+\alpha f_\mu ,K^2E^{(1)}+\beta K^2,f_{\mu q^1},1\}`$, $`\mu q`$,
* $`𝒳_{62}^4=\mathrm{Lin}\{f_\mu E^{(1)}+\alpha f_\mu ,f_{\mu q^1},f_\nu ,1\}`$, $`\mu q`$, $`\mu q\nu `$, $`\nu 1`$,
* $`𝒳_{63}^4=\mathrm{Lin}\{f_{\mu _1},f_{\mu _2},f_{\mu _3},1\}`$, $`\mu _i\mu _j`$, $`\mu _i1`$ for any $`ij`$.
## 5 Further restrictions on the calculus
### 5.1 Generators of the FODC
Let $`\mathrm{\Gamma }`$ be a 3-dimensional left-covariant differential calculus on $`\mathrm{SL}_q(2)`$. We shall look for calculi $`\mathrm{\Gamma }`$ satisfying the following additional condition:
* The left-invariant one-forms $`\omega (u_2^1)`$, $`\omega (u_1^2)`$ and $`\omega (u_1^1u_2^2)`$ generate $`\mathrm{\Gamma }`$ as a left $`𝒜`$-module.
It is simple to check that this condition is fulfilled if and only if the subspace $`\overline{𝒳}`$ of the linear functionals on the matrix elements of the fundamental corepresentation $`u`$ is four dimensional. Note that the classical differential calculus on $`\mathrm{SL}(2)`$ obviously has this property. All coideals from Subsection 4.3 satisfying condition (LI) are marked with the label (LI). The necessary and sufficient conditions for the parameter values are indicated after the label.
### 5.2 Universal (higher order) differential calculi
In the previous section we have seen that there is a very large number of left-covariant first order differential calculi on the quantum group $`\mathrm{SL}_q(2)`$. Now the corresponding left-covariant universal (higher order) differential calculi will be considered, too. We require that
* the dimension of the space of left-invariant differential 2-forms is at least 3.
Recall that any differential calculus is a quotient of the universal differential calculus. Hence if there is a calculus such that the dimension of the space of its left-invariant differential 2-forms is equal to 3 then the universal calculus satisfies the above condition.
In order to study this condition we use Lemma 2 and the list in Subsection 4.3. As a sample let us consider $`𝒳=𝒳_8^4`$, $`\gamma 2\alpha `$. We then have
$`\overline{𝒳}`$ $`=\{X_1,X_2,X_3,X_4\}`$
$`=\{FKG+\alpha KEG+\beta K^2G+\gamma KE+\delta K^2,G,FK+\alpha KE+\beta K^2,1\}.`$
For the elements $`\mathrm{m}(X_iX_j)`$ we obtain
$`X_1X_1=`$ $`q^1F^2K^2G^2+(1+q^2)\alpha FK^2EG^2+q^1\alpha ^2K^2E^2G^2`$ (16)
$`+\text{terms of lower degree}`$
$`X_1X_2=`$ $`FKG^2+\alpha KEG^2+\beta K^2G^2+\gamma KEG+\delta K^2G`$ (17)
$`X_1X_3=`$ $`q^1F^2K^2G+(1+q^2)\alpha FK^2EG+q^1\alpha ^2K^2E^2G`$ (18)
$`+\text{terms of lower degree}`$
$`X_1X_4=`$ $`FKG+\alpha KEG+\beta K^2G+\gamma KE+\delta K^2`$ (19)
$`X_2X_1`$ $`X_1X_2+2X_1X_4=4\alpha KEG+2\beta K^2G+4\gamma KE+2\delta K^2`$ (20)
$`X_2X_2=`$ $`G^2`$ (21)
$`X_2X_3`$ $`X_1X_4+2X_3X_4=(\gamma +4\alpha )KE+(\delta +2\beta )K^2`$ (22)
$`X_2X_4=`$ $`G`$ (23)
$`X_3X_1`$ $`X_1X_32X_3^2=((1q^2)\gamma 4\alpha )FK^2E4q^1\alpha ^2K^2E^2`$ (24)
$`+\text{terms of lower degree}`$
$`X_3X_2`$ $`X_1X_4=\gamma KE\delta K^2`$ (25)
$`X_3X_3=`$ $`q^1F^2K^2+(1+q^2)\alpha FK^2E+q^1\alpha ^2K^2E^2+(1+q^2)\beta FK^3`$
$`+(1+q^2)\alpha \beta K^3E+(\beta ^2+\alpha /(qq^1))K^4\alpha /(qq^1)`$ (26)
$`X_3X_4=`$ $`FK+\alpha KE+\beta K^2`$ (27)
$`X_4X_1`$ $`X_1X_4=0`$ (28)
$`X_4X_2`$ $`X_2X_4=0`$ (29)
$`X_4X_3`$ $`X_3X_4=0`$ (30)
$`X_4X_4=`$ $`1`$ (31)
From Lemma 2 we conclude that there are at least $`3+dim𝒳=6`$ linearly independent relations in $`\overline{𝒳}^2`$. Hence there are at most 10 linearly independent elements in $`\overline{𝒳}^2`$. Because of the PBW-theorem the 9 elements in (16)–(19), (21), (23), (5.2), (27) and (31) are linearly independent. Suppose that $`\alpha `$ is nonzero. Then (20) and the difference (22)$``$(25) are further linearly independent elements which is a contradiction. If $`\alpha =0`$ then $`\gamma 0`$ because of (LI). Therefore, (22) and (24) are two additional linearly independent elements in $`\overline{𝒳}^2`$ and we obtain again a contradiction.
The same procedure can be applied to all right coideals of the list in Subsection 4.3. In order to carry out these computations we used the computer algebra program FELIX . The result is the following (the number of the general coideal and the corresponding parameter values are given in parenthesis):
* (47; $`\alpha =\beta =\gamma =0`$, $`\delta =q^5/(q^21)^2`$)
$`\overline{𝒳}=\{FK^2E+q^5/(q^21)^2K^4,FK,KE,1\}`$
* (53; $`\alpha =0`$, $`\beta 0`$, $`\gamma =(qq^1)^2\beta ^2`$, $`\delta =(1+q)\beta `$, $`\mu =q^{1/2}`$)
$`\overline{𝒳}=\{F+\beta K,FK(qq^1)^2\beta ^2KE+(1+q)\beta K^2,K^1,1\}`$
* (53; $`\alpha =0`$, $`\beta 0`$, $`\gamma =(qq^1)^2\beta ^2`$, $`\delta =(1+q)\beta `$, $`\mu =q^{1/2}`$)
$`\overline{𝒳}=\{F\epsilon _{}+\beta \epsilon _{}K,FK(qq^1)^2\beta ^2KE+(1+q)\beta K^2,\epsilon _{}K^1,1\}`$
* (53; $`\alpha =0`$, $`\beta 0`$, $`\gamma =(qq^1)^2\beta ^2`$, $`\delta =(1q)\beta `$, $`\mu =\mathrm{i}q^{1/2}`$)
$`\overline{𝒳}=\{Ff_\mathrm{i}+\beta f_\mathrm{i}K,FK+(qq^1)^2\beta ^2KE+(1q)\beta K^2,f_\mathrm{i}K^1,1\}`$
* (53; $`\alpha =0`$, $`\beta 0`$, $`\gamma =(qq^1)^2\beta ^2`$, $`\delta =(1q)\beta `$, $`\mu =\mathrm{i}q^{1/2}`$)
$`\overline{𝒳}=\{Ff_\mathrm{i}+\beta f_\mathrm{i}K,FK+(qq^1)^2\beta ^2KE+(1q)\beta K^2,f_\mathrm{i}K^1,1\}`$
* (53; $`\alpha 0`$, $`\gamma =\alpha `$, $`\beta =\delta =0`$, $`\mu =\mathrm{i}q`$) $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2}=4`$
$`\overline{𝒳}=\{Ff_\mathrm{i}K+\alpha f_\mathrm{i}KE,FK\alpha KE,f_\mathrm{i},1\}`$
* (53; $`\alpha 0`$, $`\gamma =\alpha `$, $`\beta =\delta =0`$, $`\mu =\mathrm{i}q`$) $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2}=4`$
$`\overline{𝒳}=\{Ff_\mathrm{i}K+\alpha f_\mathrm{i}KE,FK\alpha KE,f_\mathrm{i},1\}`$
* (54; $`\alpha =\beta =\gamma =0`$, $`\mu =q^3`$)
$`\overline{𝒳}=\{FK^5,KE,K^4,1\}`$
* (54; $`\alpha =\beta =\gamma =0`$, $`\mu =q^1`$)
$`\overline{𝒳}=\{FK^3,KE,K^4,1\}`$
* (55; $`\alpha =\beta =0`$, $`\mu =1`$)
$`\overline{𝒳}=\{FK^1,K^1E,K^2,1\}`$
* (55; $`\alpha =\beta =0`$, $`\mu =1`$)
$`\overline{𝒳}=\{F\epsilon _{}K^1,\epsilon _{}K^1E,\epsilon _{}K^2,1\}`$
* (55; $`\alpha =\beta =0`$, $`\mu =q^1`$)
$`\overline{𝒳}=\{FK^3,K^3E,K^4,1\}`$
* (57; $`\alpha =\beta =\gamma =0`$, $`\mu =q^3`$)
$`\overline{𝒳}=\{FK,K^5E,K^4,1\}`$
* (57; $`\alpha =\beta =\gamma =0`$, $`\mu =q^1`$)
$`\overline{𝒳}=\{FK,K^3E,K^4,1\}`$
* (57; $`\gamma 0`$, $`\alpha =1/((qq^1)^2\gamma ^2)`$, $`\beta =1/((q1)(q^21)\gamma )`$, $`\mu =q^{1/2}`$)
$`\overline{𝒳}=\{FK+\alpha KE+\beta K^2,E+\gamma K,K^1,1\}`$
* (57; $`\gamma 0`$, $`\alpha =1/((qq^1)^2\gamma ^2)`$, $`\beta =1/((q1)(q^21)\gamma )`$, $`\mu =q^{1/2}`$)
$`\overline{𝒳}=\{FK+\alpha KE+\beta K^2,\epsilon _{}E+\gamma \epsilon _{}K,\epsilon _{}K^1,1\}`$
* (58; $`\alpha =\beta =0`$, $`\mu =q^2`$)
$`\overline{𝒳}=\{FK,KE,K^4,1\}`$
* (58; $`\alpha =\beta =0`$, $`\mu =q`$)
$`\overline{𝒳}=\{FK,KE,K^2,1\}`$
* (58; $`\alpha 0`$, $`\beta =q^3/((q^21)^2\alpha )`$, $`\mu =q^1`$)
$`\overline{𝒳}=\{FK+\alpha K^2,KEq^3/((q^21)^2\alpha )K^2,K^2,1\}`$
* (58; $`\alpha =\beta =0`$, $`\mu =q`$)
$`\overline{𝒳}=\{FK,KE,\epsilon _{}K^2,1\}`$
Remark. The solutions 6 and 7 have the property $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2}=4`$. In all other cases we have $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2}=3`$.
### 5.3 Hopf algebra automorphisms
Let $`𝒜`$ be a Hopf algebra and $`\phi `$ an automorphism of $`𝒜`$. Let $`\mathrm{\Gamma }`$ be an $`𝒜`$-bimodule. Then the actions $`.:𝒜\times \mathrm{\Gamma }\mathrm{\Gamma }`$, $`a.\rho :=\phi (a)\rho `$ and $`.:\mathrm{\Gamma }\times 𝒜\mathrm{\Gamma }`$, $`\rho .a:=\rho \phi (a)`$ determine another $`𝒜`$-bimodule structure on $`\mathrm{\Gamma }`$, since $`\phi `$ is an algebra homomorphism. Let $`\mathrm{\Gamma }`$ be a left $`𝒜`$-module. Then the mapping $`\mathrm{\Delta }_\mathrm{L}^{}:\mathrm{\Gamma }𝒜\mathrm{\Gamma }`$, $`\mathrm{\Delta }_\mathrm{L}^{}(\rho )=\phi (\rho _{(1)})\rho _{(0)}`$ determines a second left $`𝒜`$-comodule structure on $`\mathrm{\Gamma }`$, since $`\phi `$ is a comodule homomorphism.
If $`(\mathrm{\Gamma },\mathrm{d})`$ is a left-covariant FODC over $`𝒜`$, set $`\mathrm{\Gamma }_\phi :=\mathrm{\Gamma }`$ with module action . and comodule mapping $`\mathrm{\Delta }_\mathrm{L}^{}`$. Further define $`\mathrm{d}_\phi :𝒜\mathrm{\Gamma }`$, $`\mathrm{d}_\phi a:=\mathrm{d}\phi (a)`$. Then $`(\mathrm{\Gamma }_\phi ,\mathrm{d}_\phi )`$ is a left-covariant FODC over $`𝒜`$. Indeed, $`\mathrm{d}_\phi `$ satisfies the Leibniz rule and
$$\mathrm{Lin}\{a.\mathrm{d}_\phi b|a,b𝒜\}=\mathrm{Lin}\{\phi (a)\mathrm{d}\phi (b)|a,b𝒜\}=\mathrm{Lin}\{a\mathrm{d}b|a,b𝒜\}$$
since $`\phi `$ is invertible.
Definition 1. Let $`(\mathrm{\Gamma },\mathrm{d})`$ be a left-covariant FODC over the Hopf algebra $`𝒜`$. We call $`(\mathrm{\Gamma },\mathrm{d})`$ Hopf-invariant if $`(\mathrm{\Gamma }_\phi ,\mathrm{d}_\phi )`$ is isomorphic to $`(\mathrm{\Gamma },\mathrm{d})`$ for all Hopf algebra automorphism $`\phi `$ of $`𝒜`$.
Proposition 5. Let $`(\mathrm{\Gamma },\mathrm{d})`$ be a left-covariant FODC over $`𝒜`$. The following statements are equivalent:
(i) $`(\mathrm{\Gamma },\mathrm{d})`$ is Hopf-invariant.
(ii) $`\phi (_\mathrm{\Gamma })=_\mathrm{\Gamma }`$ for any Hopf algebra automorphism $`\phi `$ of $`𝒜`$.
(iii) $`\phi (𝒳_\mathrm{\Gamma })=𝒳_\mathrm{\Gamma }`$ for any Hopf algebra automorphism $`\phi `$ of $`𝒜^{}`$ which is implemented by $`𝒜`$.
Proof. It suffices to show that for any Hopf algebra automorphism $`\phi `$ of $`𝒜`$ we have $`_{\mathrm{\Gamma }_\phi }=\phi ^1(_\mathrm{\Gamma })`$ and $`𝒳_{\mathrm{\Gamma }_\phi }=\phi (𝒳_\mathrm{\Gamma })`$. For this we compute
$`\omega _\phi (a):=S(a_{(1)}).\mathrm{d}_\phi a_{(2)}=\phi (S(a_{(1)}))\mathrm{d}\phi (a_{(2)})=S(\phi (a)_{(1)})\mathrm{d}\phi (a)_{(2)}=\omega (\phi (a))`$
for any $`a𝒜`$. Hence (with $`𝒜^+=𝒜\mathrm{ker}\epsilon `$)
$`_{\mathrm{\Gamma }_\phi }`$ $`=\{a𝒜^+|\omega _\phi (a)=0\}`$
$`=\{a𝒜^+|\omega (\phi (a))=0\}=\{\phi ^1(b)𝒜^+|\omega (b)=0\}=\phi ^1(_\mathrm{\Gamma })`$
and (with $`𝒜^{}{}_{}{}^{+}:=𝒜^{}\mathrm{ker}\epsilon `$)
$`𝒳_{\mathrm{\Gamma }_\phi }`$ $`=\{X𝒜^{}{}_{}{}^{+}|X(_{\mathrm{\Gamma }_\phi })=0\}=\{X𝒜^{}{}_{}{}^{+}|X(\phi ^1(_\mathrm{\Gamma }))=0\}`$
$`=\{Y\phi 𝒜^{}{}_{}{}^{+}|Y(_\mathrm{\Gamma })=0\}=\phi (𝒳_\mathrm{\Gamma }).`$
It is easy to prove that any Hopf algebra automorphism $`\phi `$ of $`𝒪(\mathrm{SL}_q(2))`$ is given by $`\phi (u_1^1)=u_1^1`$, $`\phi (u_2^2)=u_2^2`$, $`\phi (u_2^1)=\alpha u_2^1`$, $`\phi (u_1^2)=\alpha ^1u_1^2`$, where $`\alpha ^\times `$ (see also \[7, Section 4.1.2\]). For this $`\phi =:\phi _\alpha `$ we obtain from (13) the formulas
$`\phi _\alpha (E)=\alpha ^1E,\phi _\alpha (F)=\alpha F,\phi _\alpha (f_\mu )=f_\mu ,\mu ^\times .`$ (32)
Proposition 6. A left-covariant FODC $`\mathrm{\Gamma }`$ over $`𝒪(\mathrm{SL}_q(2))`$ is Hopf-invariant if and only if its quantum tangent space $`𝒳_\mathrm{\Gamma }`$ is generated by elements which are homogeneous with respect to the $``$-grading of $`𝒰`$.
Proof. Clearly $`\phi _\alpha (X)=\alpha ^nX`$ for any $`X𝒳`$, $`\mathrm{deg}X=n`$. Therefore the condition of the Proposition is sufficient.
Conversely, suppose that $`X=_{i=1}^k\lambda _iX_i𝒳`$, $`k>1`$, where $`\lambda _i^\times `$, $`X_i𝒳`$, $`\mathrm{deg}X_i=n_i`$ and $`n_in_j`$ for any $`ij`$. If $`(\mathrm{\Gamma },\mathrm{d})`$ is Hopf-invariant then we have $`𝒳\phi _q(X)q^{n_k}X=_{i=1}^{k1}(q^{n_i}q^{n_k})X_i`$. Since $`q`$ is not a root of unity, we obtain that $`_{i=1}^{k1}\mu _iX_i𝒳`$, $`\mu _i^\times `$. By induction on $`k`$ one easily proves that $`X_i𝒳`$ for any $`i`$.
Corollary 7. The items 1, 8-14, 17, 18 and 20 are precisely the quantum tangent spaces of the list in Subsection 5.2 which are Hopf-invariant.
## 6 More structures on left-covariant differential calculi on $`\mathrm{SL}_q(2)`$
The Hopf algebra $`𝒪(\mathrm{SL}_q(2))`$ admits 3 non-equivalent real forms. Namely,
* $`q`$; $`(u_1^1)^{}=u_2^2`$, $`(u_2^1)^{}=qu_1^2`$, $`(u_1^2)^{}=q^1u_2^1`$, $`(u_2^2)^{}=u_1^1`$;
* $`q`$; $`(u_1^1)^{}=u_2^2`$, $`(u_2^1)^{}=qu_1^2`$, $`(u_1^2)^{}=q^1u_2^1`$, $`(u_2^2)^{}=u_1^1`$;
* $`|q|=1`$; $`(u_j^i)^{}=u_j^i`$ for any $`i,j=1,2`$.
The Hopf $``$-algebras corresponding to them are $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$ and $`𝒪(\mathrm{SL}_q(2,))`$. Let us introduce the dual involution on $`𝒰`$ in the way $`f^{}(a):=\overline{f(S(a)^{})}`$. Then the corresponding $``$-structures on $`𝒰`$ are given by
* $`E^{}=F`$, $`F^{}=E`$, $`G^{}=G`$, $`f_\mu ^{}=f_{\overline{\mu }}`$,
* $`E^{}=F`$, $`F^{}=E`$, $`G^{}=G`$, $`f_\mu ^{}=f_{\overline{\mu }}`$,
* $`E^{}=qE`$, $`F^{}=q^1F`$, $`G^{}=G`$, $`f_\mu ^{}=f_{\overline{\mu }^1}`$,
respectively. A FODC $`(\mathrm{\Gamma },\mathrm{d})`$ over a Hopf $``$-algebra $`𝒜`$ is called a $``$-calculus if there exists an involution $`:\mathrm{\Gamma }\mathrm{\Gamma }`$ such that $`(a(\mathrm{d}b)c)^{}=c^{}(\mathrm{d}b^{})a^{}`$ for any $`a,b,c𝒜`$.
Proposition 8. Let $`(\mathrm{\Gamma },\mathrm{d})`$ be a finite-dimensional left-covariant FODC over a Hopf $``$-algebra $`𝒜`$. Then $`(\mathrm{\Gamma },\mathrm{d})`$ is a $``$-calculus if and only if its quantum tangent space $`𝒳_\mathrm{\Gamma }`$ is $``$-invariant.
Let $`\mathrm{\Gamma }`$ be a left-covariant bimodule over a Hopf algebra $`𝒜`$. An invertible linear mapping $`\sigma :\mathrm{\Gamma }_𝒜\mathrm{\Gamma }\mathrm{\Gamma }_𝒜\mathrm{\Gamma }`$ is called a braiding of $`\mathrm{\Gamma }`$ if $`\sigma `$ is a homomorphism of $`𝒜`$-bimodules, commutes with the left coaction on $`\mathrm{\Gamma }`$ and satisfies the braid relation
$$(\sigma \mathrm{id})(\mathrm{id}\sigma )(\sigma \mathrm{id})=(\mathrm{id}\sigma )(\sigma \mathrm{id})(\mathrm{id}\sigma )$$
on $`\mathrm{\Gamma }^3`$. If $`(\mathrm{\Gamma },\mathrm{d})`$ is a left-covariant differential calculus over $`𝒜`$ then we require that $`(\mathrm{id}\sigma )(𝒮\mathrm{\Gamma }^2)=0`$. Such a braiding neither needs to exist nor it is unique for a given left-covariant differential calculus over $`𝒜`$.
Let $`(\mathrm{\Gamma },\mathrm{d})`$ be a left-covariant FODC over $`𝒜`$. Fix a basis $`\{X_i|i=1,\mathrm{},n\}`$ of its quantum tangent space $`𝒳_\mathrm{\Gamma }`$ and let $`\{\omega _i|i=1,\mathrm{},n\}`$ be the dual basis of $`\mathrm{\Gamma }_\mathrm{L}`$. If $`\sigma `$ is a braiding of $`\mathrm{\Gamma }`$ with $`\sigma (\omega _i\omega _j)=\sigma _{ij}^{kl}\omega _k\omega _l`$ then we can define a bilinear mapping $`[,]:𝒳_\mathrm{\Gamma }\times 𝒳_\mathrm{\Gamma }𝒳_\mathrm{\Gamma }`$ by
$`[X_i,X_j]:=X_iX_j\sigma _{kl}^{ij}X_kX_l.`$ (33)
This mapping can be viewed as a generalization of the Lie bracket of the left-invariant vector fields if it also satisfies a generalized Jacobi identity. For the calculi and braidings described below the mapping $`\beta :=[,]`$ fulfills the equation
$`(\beta (\beta \mathrm{id})\beta (\mathrm{id}\beta ))A_3^\mathrm{t}=0`$ (34)
where $`A_3^\mathrm{t}(X_iX_jX_k)=(\mathrm{id}\sigma _{12})(\mathrm{id}\sigma _{23}+\sigma _{23}\sigma _{12})_{rst}^{ijk}X_rX_sX_t`$. Unfortunately, such an equation does not hold for bicovariant differential calculi in general.
## 7 The calculi in detail
The main result of the considerations of the preceding sections is the following.
Theorem 9. Suppose that $`q`$ is a transcendental complex number. Let $`(\mathrm{\Gamma },\mathrm{d})`$ be a left-covariant FODC over $`𝒪(\mathrm{SL}_q(2))`$ having the following properties:
* $`\mathrm{\Gamma }`$ is a 3-dimensional left-covariant bimodule
* the left-invariant one-forms $`\omega (u_2^1),\omega (u_1^2)`$ and $`\omega (u_1^1u_2^2)`$ form a basis of the left module $`\mathrm{\Gamma }`$
* the universal differential calculus $`{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{}`$ associated to $`\mathrm{\Gamma }`$ satisfies the inequality $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{2}3`$
* $`\mathrm{\Gamma }`$ is Hopf-invariant.
Then $`\mathrm{\Gamma }`$ is isomorphic to one of the 11 left-covariant FODC 1,8-14,17,18,20 of the list in Subsection 5.2.
Remarks. 1. The calculus 17 is isomorphic to the 3D-calculus of Woronowicz . Two other 3-dimensional calculi appear in . They are isomorphic to the calculus 13 (for $`r=2`$) and 8 (for $`r=3`$), respectively. The calculi 10 and 11 are subcalculi of the bicovariant $`4\mathrm{D}_+`$\- and $`4\mathrm{D}_{}`$-calculus, respectively (see also \[7, Section 14.2.4\]). One can easily check all these isomorphisms by comparing the corresponding right ideals $`_\mathrm{\Gamma }`$.
2. In \[8, Section 7\] the notion of elementary left-covariant FODC was introduced. It is an easy computation to show that the six calculi 10–12,17,18 and 20 in Theorem 7 are elementary, while the others are not.
After the last section we list some important facts about these 11 calculi. First we fix the basis $`\omega _H:=\omega ((u_1^1u_2^2)/2),\omega _X:=\omega (u_2^1),\omega _Y:=\omega (u_1^2)`$ in $`\mathrm{\Gamma }_\mathrm{L}`$ and determine the dual basis $`\{H,X,Y\}`$ (i. e. $`\omega (a)=H(a)\omega _H+X(a)\omega _X+Y(a)\omega _Y`$ for any $`a𝒪(\mathrm{SL}_q(2))`$). We examine whether or not the calculus is a $``$-calculus with respect to the three given involutions on $`𝒪(\mathrm{SL}_q(2))`$. We compute the pairing between the quantum tangent space and linear (fundamental representation) and quadratic elements of $`𝒪(\mathrm{SL}_q(2))`$. Using this the generators of the right ideal $`_\mathrm{\Gamma }`$ are computed.
The quadratic-linear relations between generators of $`𝒳_\mathrm{\Gamma }`$ are given. Because of their simple form one can easily show that the algebra without unit generated by the elements $`H,X`$, and $`Y`$ and these relations has the PBW-basis $`\{H^{n_1}X^{n_2}Y^{n_3}|n_1,n_2,n_3_0,n_1+n_2+n_3>0\}`$. This proves also that $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{3}=1`$ and $`dim{}_{\mathrm{u}}{}^{}\mathrm{\Gamma }_{}^{k}=0`$ for any $`k>3`$ and for each of the 11 calculi.
The matrix $`(f_j^i)`$ describes the commutation rules between one-forms and functions (see equation (4)), where $`i,j\{H,X,Y\}`$. Finally, the generators of the right ideal and equation (7) determine the vector space of left-invariant symmetric 2-forms. If there exists a braiding $`\sigma `$ of the left-covariant bimodule $`\mathrm{\Gamma }`$ such that $`(\mathrm{id}\sigma )(𝒮\mathrm{\Gamma }_\mathrm{L}^2)=0`$ then we give one of them by its eigenvalues and eigenspaces. It was determined by means of the computer algebra program FELIX .
## 8 Cohomology
Let us fix one of the differential calculi of Theorem 7. The differential mapping $`\mathrm{d}:\mathrm{\Gamma }^{}\mathrm{\Gamma }^{}`$ satisfies the equation $`\mathrm{d}^2=0`$. Hence it defines a complex
$$\begin{array}{ccccccccccc}\{0\}& \stackrel{\mathrm{d}}{}& 𝒜& \stackrel{\mathrm{d}}{}& \mathrm{\Gamma }& \stackrel{\mathrm{d}}{}& \mathrm{\Gamma }^2& \stackrel{\mathrm{d}}{}& \mathrm{\Gamma }^3& \stackrel{\mathrm{d}}{}& \mathrm{\Gamma }^4=\{0\}.\end{array}$$
(35)
Now the corresponding cohomology spaces will be determined. In Woronowicz introduced a method to do this. Let us recall his strategy.
The first observation is that the algebra $`𝒜=𝒪(\mathrm{SL}_q(2))`$ has a vector space basis
$$\{v_{ij}^\lambda |\lambda \frac{1}{2}_0;\lambda i,j\lambda ;\lambda i,\lambda j_0\}$$
(36)
such that $`\mathrm{\Delta }(v_{ij}^\lambda )=v_{ik}^\lambda v_{kj}^\lambda `$ for any $`\lambda ,i,j`$. Let us fix such a basis. Then for $`\rho _j\mathrm{\Gamma }_\mathrm{L}^{}`$ we obtain
$$\mathrm{d}(v_{ij}^\lambda \rho _j)=v_{ik}^\lambda X_r(v_{kj}^\lambda )\omega _r\rho _j+v_{ij}^\lambda \mathrm{d}\rho _j.$$
(37)
The elements $`_{r,j}X_r(v_{kj}^\lambda )\omega _r\rho _j+\mathrm{d}\rho _k`$ are left-invariant. Hence the differential mapping $`\mathrm{d}`$ preserves the direct sum decomposition
$$\mathrm{\Gamma }^{}=\underset{\lambda ,i}{}C(v_i^\lambda )\mathrm{\Gamma }_\mathrm{L}^{},$$
(38)
where $`C(v_i^\lambda )=\mathrm{Lin}\{v_{kl}^\mu |\mu =\lambda ,k=i\}`$. Moreover, formula (37) for the differential on $`C(v_i^\lambda )\mathrm{\Gamma }_\mathrm{L}^{}`$ does not depend on $`i`$. Therefore, for any $`\lambda \frac{1}{2}_0`$ we obtain $`2\lambda `$ isomorphic differential complexes.
Let $`V^\lambda =\mathrm{Lin}\{e_\mu ^\lambda |\lambda \mu \lambda ;\lambda \mu _0\}`$ be the representation of $`𝒜^{}`$ defined by $`f.e_\mu ^\lambda =e_\nu ^\lambda f(v_{\nu \mu }^\lambda )`$ for any $`f𝒜^{}`$. It is isomorphic to the representation given by
$`E.e_\nu ^\lambda `$ $`=[\lambda +\nu ]e_{\nu 1}^\lambda ,`$ $`F.e_\nu ^\lambda `$ $`=[\lambda \nu ]e_{\nu +1}^\lambda ,`$
$`f_\mu .e_\nu ^\lambda `$ $`=\mu ^{2\nu }e_\nu ^\lambda ,`$ $`G.e_\nu ^\lambda `$ $`=2\nu e_\nu ^\lambda .`$ (39)
Because of the left-covariance of the differential calculus, $`V^\lambda `$ induces a representation of the quantum tangent space $`𝒳_\mathrm{\Gamma }`$ of the differential calculus. We obtain a new complex
$$\begin{array}{ccccccc}\{0\}& \stackrel{\mathrm{d}}{}& V^\lambda & \stackrel{\mathrm{d}}{}& V^\lambda \mathrm{\Gamma }_\mathrm{L}& \stackrel{\mathrm{d}}{}& V^\lambda \mathrm{\Gamma }_\mathrm{L}^2\\ & & & & & \stackrel{\mathrm{d}}{}& V^\lambda \mathrm{\Gamma }_\mathrm{L}^3& \stackrel{\mathrm{d}}{}& \{0\},\end{array}$$
(40)
where
$$\mathrm{d}(e_\mu ^\lambda \rho _\mu )=e_\nu ^\lambda X_r(v_{\nu \mu }^\lambda )\omega _r\rho _\mu +e_\mu ^\lambda \mathrm{d}\rho _\mu .$$
(41)
Lemma 10. The cohomology spaces of the complex (41) are isomorphic to $`H_\lambda ^0=H_\lambda ^3=^p`$, $`H_\lambda ^1=H_\lambda ^2=\{0\}`$, where $`p=1`$ for $`\lambda =0`$ and $`p=0`$ otherwise.
Theorem 11. Let $`\mathrm{\Gamma }`$ be one of the differential calculi of Theorem 7. Then the cohomology spaces of the differential complex (35) are isomorphic to
$`H^0`$ $`=H^3=^p,`$ $`H^1`$ $`=H^2=\{0\},`$ (42)
where $`p=1`$ for $`\lambda =0`$ and $`p=0`$ otherwise.
Proof. The assertion follows from Lemma 8 and the preceding considerations.
Proof of the Lemma. For $`\lambda =0`$ we have $`\mathrm{d}e_0^0=0`$ and $`\mathrm{d}(e_0^0\omega )0`$ for any $`\omega \mathrm{\Gamma }_\mathrm{L}`$. Since $`dim\mathrm{\Gamma }_\mathrm{L}^2=dim\mathrm{\Gamma }_\mathrm{L}(=3)<\mathrm{}`$, the mapping $`\mathrm{d}:\mathrm{\Gamma }_\mathrm{L}\mathrm{\Gamma }_\mathrm{L}^2`$ is also surjective. Therefore, for any $`\xi \mathrm{\Gamma }_\mathrm{L}^2`$ we obtain $`\mathrm{d}\xi (=\mathrm{d}^2\omega )=0`$. Hence the assertion of the lemma for $`\lambda =0`$ is valid.
Let now $`\lambda 0`$. We set
$$(e_\mu ^\lambda \underset{j=1}{\overset{k}{}}\omega _{i_j}):=\mu +\underset{j=1}{\overset{k}{}}(\omega _{i_j}),\text{ where }(\omega _H)=0,(\omega _Y)=(\omega _X)=1.$$
(43)
Then $``$ defines a direct sum decomposition of $`V^\lambda \mathrm{\Gamma }_\mathrm{L}^{}`$ and $`\mathrm{d}`$ preserves this decomposition: $`(\mathrm{d}\rho )=(\rho )`$ for any homogeneous element $`\rho `$ of $`V^\lambda \mathrm{\Gamma }_\mathrm{L}^{}`$. Hence it suffices to check that the sequence (40) restricted to homogeneous elements of the same degree $`\mu `$ is exact for any $`\lambda \frac{1}{2}`$ and any $`\mu `$ with $`\lambda \mu `$.
For $`\rho =e_\mu ^\lambda `$ the equation $`\mathrm{d}\rho (=X_i.e_\mu ^\lambda \omega _i)=0`$ implies that $`X.e_\mu ^\lambda =Y.e_\mu ^\lambda =0`$. Hence $`\mu =\lambda `$ and $`\mu =\lambda `$. This is a contradiction to $`\lambda 0`$, so we have $`H_\lambda ^0=\{0\}`$.
Since $`\mathrm{d}\rho =0`$ for any $`\rho V^\lambda \mathrm{\Gamma }_\mathrm{L}^3`$ we have to show that $`\rho =\mathrm{d}\xi `$ for some $`\xi V^\lambda \mathrm{\Gamma }_\mathrm{L}^2`$. If $`(\rho )=\mu `$ then $`\lambda \mu \lambda `$ and $`\rho `$ is a constant multiple of $`e_\mu ^\lambda \omega _H\omega _X\omega _Y`$. One can take $`\xi _1:=c^1e_{\mu +1}^\lambda \omega _H\omega _X`$ or $`\xi _2:=c^1e_{\mu 1}^\lambda \omega _H\omega _Y`$ with some $`c^\times `$. Indeed, because of $`\omega _H\omega _H\omega _X=\omega _X\omega _H\omega _X=\mathrm{d}(\omega _H\omega _X)=0`$ the element $`\mathrm{d}(e_{\mu +1}^\lambda \omega _H\omega _X)=Y.e_{\mu +1}^\lambda \omega _Y\omega _H\omega _X`$ is a nonzero multiple of $`\rho `$ for $`\mu \lambda `$. Similarly, $`\mathrm{d}(c\xi _2)`$ is a nonzero multiple of $`\rho `$ for $`\mu \lambda `$. Therefore, $`H_\lambda ^3=\{0\}`$.
Now we prove that $`H_\lambda ^1=\{0\}`$. It is easy to see that if $`\omega =\mu `$ for a nonzero $`\omega V^\lambda \mathrm{\Gamma }_\mathrm{L}`$ then $`|\mu |\lambda +1`$. Hence we have to show that $`\mathrm{d}\omega 0`$ if $`|\mu |=\lambda +1`$ and that the vector space $`\{\omega V^\lambda \mathrm{\Gamma }_\mathrm{L}|\omega =\mu ,\mathrm{d}\omega =0\}`$ is one-dimensional for $`|\mu |\lambda `$. We define $`V_1^{\lambda \mu }:=\{\omega V^\lambda \mathrm{\Gamma }_\mathrm{L}|\omega =\mu \}`$. Since $`e_\lambda ^\lambda \omega _Y`$ generates the vector space $`V_1^{\lambda ,\lambda +1}`$ we have to show that $`H.e_\lambda ^\lambda \omega _H\omega _Y+e_\lambda ^\lambda \mathrm{d}\omega _Y0`$. Similarly, $`\mathrm{d}(e_\lambda ^\lambda \omega _X)=H.e_\lambda ^\lambda \omega _H\omega _X+e_\lambda ^\lambda \mathrm{d}\omega _X`$ must be nonzero. Using the explicit formulas for the quantum tangent space, for the differentials $`\mathrm{d}\omega _X`$ and $`\mathrm{d}\omega _Y`$ and (8) this is easily done.
Secondly, the elements $`e_\mu ^\lambda \omega _H`$, $`e_{\mu +1}^\lambda \omega _X`$ and $`e_{\mu 1}^\lambda \omega _Y`$ generate the vector space $`V_1^{\lambda \mu }`$. We have $`dimV_1^{\lambda \mu }=2`$ for $`|\mu |=\lambda `$ and $`dimV_1^{\lambda \mu }=3`$ for $`|\mu |<\lambda `$. Moreover,
$`\mathrm{d}(e_\mu ^\lambda \omega _H)=`$ $`X.e_\mu ^\lambda \omega _X\omega _H+Y.e_\mu ^\lambda \omega _Y\omega _H`$
$`+(H.e_\mu ^\lambda \omega _H\omega _H+e_\mu ^\lambda \mathrm{d}\omega _H)`$ (44)
$`\mathrm{d}(e_{\mu +1}^\lambda \omega _X)=`$ $`Y.e_{\mu +1}^\lambda \omega _Y\omega _X+(H.e_{\mu +1}^\lambda \omega _H\omega _X+e_{\mu +1}^\lambda \mathrm{d}\omega _X)`$ (45)
$`\mathrm{d}(e_{\mu 1}^\lambda \omega _Y)=`$ $`X.e_{\mu 1}^\lambda \omega _X\omega _Y+(H.e_{\mu 1}^\lambda \omega _H\omega _Y+e_{\mu 1}^\lambda \mathrm{d}\omega _Y)`$ (46)
Observe that $`\xi =c_{HX}(\xi )\omega _H\omega _X+c_{HY}(\xi )\omega _H\omega _Y+c_{XY}(\xi )\omega _X\omega _Y`$ for any $`\xi V^\lambda \mathrm{\Gamma }_\mathrm{L}^2`$, where $`c_{HX}(\xi ),c_{HY}(\xi ),c_{XY}(\xi )V^\lambda `$. Now if $`\mu =\lambda `$ then by (46) $`c_{XY}(\mathrm{d}(e_{\lambda 1}^\lambda \omega _Y))=X.e_{\lambda 1}^\lambda 0`$ and therefore the range of $`\mathrm{d}`$ is at least one-dimensional. Similarly, by (45) $`c_{XY}(\mathrm{d}(e_{\mu +1}^\lambda \omega _X))0`$ for $`\mu =\lambda `$. Hence $`dim\mathrm{ker}\mathrm{d}V_1^{\lambda \mu }1`$ for $`|\mu |=\lambda `$. If $`|\mu |<\lambda `$ then $`c_{HY}(\mathrm{d}(e_\mu ^\lambda \omega _H))0`$, $`c_{HY}(\mathrm{d}(e_{\mu +1}^\lambda \omega _X))=0`$ and $`c_{XY}(\mathrm{d}(e_{\mu +1}^\lambda \omega _X))0`$. Therefore, $`dim\mathrm{d}(V_1^{\lambda \mu })2`$ and so $`dim\mathrm{ker}\mathrm{d}V_1^{\lambda \mu }1`$. Together, $`dim\mathrm{ker}\mathrm{d}V_1^{\lambda \mu }1`$ for $`|\mu |\lambda `$ and $`\mathrm{ker}\mathrm{d}V_1^{\lambda \mu }=\{0\}`$ for $`|\mu |=\lambda +1`$. Because of $`\mathrm{d}e_\mu ^\lambda 0`$, $`(\mathrm{d}e_\mu ^\lambda )=\mu `$ and $`\mathrm{d}(\mathrm{d}e_\mu ^\lambda )=0`$ we also have $`dim\mathrm{ker}\mathrm{d}V_1^{\lambda \mu }1`$ for $`|\mu |\lambda `$. This means that $`H_\lambda ^1=\{0\}`$.
The last assertion, $`H_\lambda ^2=\{0\}`$, follows from dimension computations.
1. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2(q^1q)}{q^4+1}}\left(FK^2E+{\displaystyle \frac{q^5(K^41)}{(q^21)^2}}\right),`$ $`X`$ $`:=q^{1/2}FK,`$ $`Y`$ $`:=q^{1/2}KE`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q^2+q^2}}\left(\begin{array}{cc}q^2& 0\\ 0& q^2\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`qHXq^1XH`$ $`={\displaystyle \frac{2(q+q^1)}{q^2+q^2}}X`$
$`q^1HYqYH`$ $`={\displaystyle \frac{2(q+q^1)}{q^2+q^2}}Y`$
$`q^2XYq^2YX`$ $`={\displaystyle \frac{q^2+q^2}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& 0& 0\\ \frac{2q^{3/2}(q^1q)}{q^2+q^2}K^3E& K^2& 0\\ \frac{2q^{3/2}(q^1q)}{q^2+q^2}FK^3& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{q^4+1}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& \frac{2(q^2+1)}{q^2+q^2}\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& \frac{2(q^2+1)}{q^2+q^2}\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& \frac{2q^2+2q^4}{q^2+q^2}& 0& 0& 0& \frac{2(q^1q)}{q^2+q^2}& 0& 0& 0& \frac{2q^42q^2}{q^2+q^2}\\ \hfill X& 0& 1& 0& 0& 0& q& 0& 0& 0\\ \hfill Y& 0& 0& 1& 0& 0& 0& 0& q& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^4u_2^2(1+q^4)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2+(q^3q)u_1^1`$, $`(u_1^11)u_2^1`$, $`(u_1^11)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H`$ $`\omega _X\omega _X`$ $`q^1\omega _H\omega _X+q\omega _X\omega _H`$
$`q^2\omega _X\omega _Y+q^2\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q\omega _H\omega _Y+q^1\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`q^2\omega _H\omega _Xq^2\omega _X\omega _H`$, $`q^2\omega _H\omega _Yq^2\omega _Y\omega _H`$,
$`\mathrm{ker}(q^2+\sigma ):`$ $`q^3\omega _X\omega _Yq^3\omega _Y\omega _X\frac{4(qq^1)}{(q^2+q^2)^2}\omega _H\omega _H`$.
2. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^2q^2}}(K^41),`$ $`X`$ $`:=q^{5/2}FK^5,`$ $`Y`$ $`:=q^{1/2}KE`$
Real forms: $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+q^1}}\left(\begin{array}{cc}q^1& 0\\ 0& q\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^2HXq^2XH`$ $`=2X`$
$`q^2HYq^2YH`$ $`=2Y`$
$`q^3XYq^3YX+{\displaystyle \frac{(q^2+1)^2(q^21)}{4q^3}}H^2`$ $`={\displaystyle \frac{q+q^1}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& \frac{q^2q^2}{2}X& 0\\ 0& K^6& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{q^4q^2}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^2\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^2\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^2& 0& 0& 0& 0& 0& 0& 0& 2q^2\\ \hfill X& 0& q^2& 0& 0& 0& q^3& 0& 0& 0\\ \hfill Y& 0& 0& 1& 0& 0& 0& 0& q& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^2u_2^2(1+q^2)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^1q^2)u_2^1`$, $`(u_1^11)u_1^2`$
Left-invariant symmetric 2-forms: $`p=(q^2+1)^2(q^21)/4`$
$`\omega _H\omega _Hp\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q^2\omega _H\omega _X+q^2\omega _X\omega _H`$
$`q^3\omega _X\omega _Y+q^3\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^2\omega _H\omega _Y+q^2\omega _Y\omega _H`$
Braiding:
3. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^2q^2}}(K^41),`$ $`X`$ $`:=q^{3/2}FK^3,`$ $`Y`$ $`:=q^{1/2}KE`$
Real forms: $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+q^1}}\left(\begin{array}{cc}q& 0\\ 0& q^1\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^2HXq^2XH`$ $`=2X`$
$`q^2HYq^2YH`$ $`=2Y`$
$`qXYq^1YX`$ $`={\displaystyle \frac{q+q^1}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& \frac{q^2q^2}{2}X& 0\\ 0& K^2& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{1q^2}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^2\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^2\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^2& 0& 0& 0& 0& 0& 0& 0& 2q^2\\ \hfill X& 0& q^2& 0& 0& 0& q^1& 0& 0& 0\\ \hfill Y& 0& 0& 1& 0& 0& 0& 0& q& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^2u_2^2(1+q^2)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^1q^2)u_2^1`$, $`(u_1^11)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H`$ $`\omega _X\omega _X`$ $`q^2\omega _H\omega _X+q^2\omega _X\omega _H`$
$`q^1\omega _X\omega _Y+q\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^2\omega _H\omega _Y+q^2\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`q\omega _H\omega _Xq^1\omega _X\omega _H`$, $`q^1\omega _H\omega _Yq\omega _Y\omega _H`$,
$`\mathrm{ker}(q^2+\sigma ):`$ $`\omega _X\omega _Y\omega _Y\omega _X`$.
4. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{qq^1}}(K^21),`$ $`X`$ $`:=q^{1/2}FK^1,`$ $`Y`$ $`:=q^{1/2}K^1E`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+1}}\left(\begin{array}{cc}q& 0\\ 0& 1\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^1HXqXH`$ $`=2X`$
$`qHYq^1YH`$ $`=2Y`$
$`qXYq^1YX{\displaystyle \frac{qq^1}{4}}H^2`$ $`=H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^2& \frac{qq^1}{2}X& \frac{qq^1}{2}Y\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& q^1\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^1\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q& 0& 0& 0& 0& 0& 0& 0& 2q^1\\ \hfill X& 0& q& 0& 0& 0& 1& 0& 0& 0\\ \hfill Y& 0& 0& q& 0& 0& 0& 0& 1& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+qu_2^2(1+q)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^1q)u_2^1`$, $`(u_1^1q)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H+{\displaystyle \frac{1q^2}{4}}\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q\omega _H\omega _X+q^1\omega _X\omega _H`$
$`q^1\omega _X\omega _Y+q\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^1\omega _H\omega _Y+q\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`\omega _H\omega _X\omega _X\omega _H`$, $`\omega _H\omega _Y\omega _Y\omega _H`$, $`\omega _X\omega _Y\omega _Y\omega _X`$.
5. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^1q}}(\epsilon _{}K^21),`$ $`X`$ $`:=q^{1/2}F\epsilon _{}K^1,`$ $`Y`$ $`:=q^{1/2}\epsilon _{}K^1E`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q1}}\left(\begin{array}{cc}q& 0\\ 0& 1\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^1HXqXH`$ $`=2X`$
$`qHYq^1YH`$ $`=2Y`$
$`qXYq^1YX{\displaystyle \frac{qq^1}{4}}H^2`$ $`=H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}\epsilon _{}K^2& \frac{q^1q}{2}X& \frac{q^1q}{2}Y\\ 0& \epsilon _{}& 0\\ 0& 0& \epsilon _{}\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& q^1\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^1\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q& 0& 0& 0& 0& 0& 0& 0& 2q^1\\ \hfill X& 0& q& 0& 0& 0& 1& 0& 0& 0\\ \hfill Y& 0& 0& q& 0& 0& 0& 0& 1& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1qu_2^2(1q)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^1+q)u_2^1`$, $`(u_1^1+q)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H+{\displaystyle \frac{1q^2}{4}}\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q\omega _H\omega _X+q^1\omega _X\omega _H`$
$`q^1\omega _X\omega _Y+q\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^1\omega _H\omega _Y+q\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`\omega _H\omega _X\omega _X\omega _H`$, $`\omega _H\omega _Y\omega _Y\omega _H`$, $`\omega _X\omega _Y\omega _Y\omega _X`$.
6. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^2q^2}}(K^41),`$ $`X`$ $`:=q^{3/2}FK^3,`$ $`Y`$ $`:=q^{3/2}K^3E`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+q^1}}\left(\begin{array}{cc}q& 0\\ 0& q^1\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^2HXq^2XH`$ $`=2X`$
$`q^2HYq^2YH`$ $`=2Y`$
$`q^3XYq^3YX{\displaystyle \frac{(q^2+1)^2(q^21)}{4q^3}}H^2`$ $`={\displaystyle \frac{q+q^1}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& \frac{q^2q^2}{2}X& \frac{q^2q^2}{2}Y\\ 0& K^2& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{q^2q^4}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^2\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^2\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^2& 0& 0& 0& 0& 0& 0& 0& 2q^2\\ \hfill X& 0& q^2& 0& 0& 0& q^1& 0& 0& 0\\ \hfill Y& 0& 0& q^2& 0& 0& 0& 0& q^1& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^2u_2^2(1+q^2)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^1q^2)u_2^1`$, $`(u_1^1q^2)u_1^2`$
Left-invariant symmetric 2-forms: $`p=(1+q^2)^2(1q^2)/4`$
$`\omega _H\omega _H+p\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q^2\omega _H\omega _X+q^2\omega _X\omega _H`$
$`q^3\omega _X\omega _Y+q^3\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^2\omega _H\omega _Y+q^2\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`q\omega _H\omega _Xq^1\omega _X\omega _H`$, $`q^1\omega _H\omega _Yq\omega _Y\omega _H`$,
$`\mathrm{ker}(q^2+\sigma ):`$ $`q^2\omega _X\omega _Yq^2\omega _Y\omega _X`$.
7. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^2q^2}}(K^41),`$ $`X`$ $`:=q^{1/2}FK,`$ $`Y`$ $`:=q^{5/2}K^5E`$
Real forms: $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+q^1}}\left(\begin{array}{cc}q^1& 0\\ 0& q\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^2HXq^2XH`$ $`=2X`$
$`q^2HYq^2YH`$ $`=2Y`$
$`q^3XYq^3YX+{\displaystyle \frac{(q^2+1)^2(q^21)}{4q^3}}H^2`$ $`={\displaystyle \frac{q+q^1}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& 0& \frac{q^2q^2}{2}Y\\ 0& K^2& 0\\ 0& 0& K^6\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{q^2q^4}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^2\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^2\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^2& 0& 0& 0& 0& 0& 0& 0& 2q^2\\ \hfill X& 0& 1& 0& 0& 0& q& 0& 0& 0\\ \hfill Y& 0& 0& q^2& 0& 0& 0& 0& q^3& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^2u_2^2(1+q^2)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^11)u_2^1`$, $`(u_1^1q^2)u_1^2`$
Left-invariant symmetric 2-forms: $`p=(q^2+1)^2(q^21)/4`$
$`\omega _H\omega _Hp\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q^2\omega _H\omega _X+q^2\omega _X\omega _H`$
$`q^3\omega _X\omega _Y+q^3\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^2\omega _H\omega _Y+q^2\omega _Y\omega _H`$
Braiding:
8. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^2q^2}}(K^41),`$ $`X`$ $`:=q^{1/2}FK,`$ $`Y`$ $`:=q^{3/2}K^3E`$
Real forms: $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+q^1}}\left(\begin{array}{cc}q& 0\\ 0& q^1\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^2HXq^2XH`$ $`=2X`$
$`q^2HYq^2YH`$ $`=2Y`$
$`qXYq^1YX`$ $`={\displaystyle \frac{q+q^1}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& 0& \frac{q^2q^2}{2}Y\\ 0& K^2& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{1q^2}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^2\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^2\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^2& 0& 0& 0& 0& 0& 0& 0& 2q^2\\ \hfill X& 0& 1& 0& 0& 0& q& 0& 0& 0\\ \hfill Y& 0& 0& q^2& 0& 0& 0& 0& q^1& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^2u_2^2(1+q^2)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^11)u_2^1`$, $`(u_1^1q^2)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H`$ $`\omega _X\omega _X`$ $`q^2\omega _H\omega _X+q^2\omega _X\omega _H`$
$`q^1\omega _X\omega _Y+q\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^2\omega _H\omega _Y+q^2\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`q\omega _H\omega _Xq^1\omega _X\omega _H`$, $`q^1\omega _H\omega _Yq\omega _Y\omega _H`$,
$`\mathrm{ker}(q^2+\sigma ):`$ $`\omega _X\omega _Y\omega _Y\omega _X`$.
9. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^2q^2}}(K^41),`$ $`X`$ $`:=q^{1/2}FK,`$ $`Y`$ $`:=q^{1/2}KE`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+q^1}}\left(\begin{array}{cc}q^1& 0\\ 0& q\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`q^2HXq^2XH`$ $`=2X`$
$`q^2HYq^2YH`$ $`=2Y`$
$`q^1XYqYX`$ $`={\displaystyle \frac{q+q^1}{2}}H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^4& 0& 0\\ 0& K^2& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& \frac{1q^2}{2}\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^2\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q^2\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^2& 0& 0& 0& 0& 0& 0& 0& 2q^2\\ \hfill X& 0& 1& 0& 0& 0& q& 0& 0& 0\\ \hfill Y& 0& 0& 1& 0& 0& 0& 0& q& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^2u_2^2(1+q^2)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^11)u_2^1`$, $`(u_1^11)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H`$ $`\omega _X\omega _X`$ $`q^2\omega _H\omega _X+q^2\omega _X\omega _H`$
$`q\omega _X\omega _Y+q^1\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q^2\omega _H\omega _Y+q^2\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`q^3\omega _H\omega _Xq^3\omega _X\omega _H`$, $`q^3\omega _H\omega _Yq^3\omega _Y\omega _H`$,
$`\mathrm{ker}(q^2+\sigma ):`$ $`q^2\omega _X\omega _Yq^2\omega _Y\omega _X`$.
10. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{q^1q}}(K^21),`$ $`X`$ $`:=q^{1/2}FK,`$ $`Y`$ $`:=q^{1/2}KE`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{q+1}}\left(\begin{array}{cc}1& 0\\ 0& q\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`qHXq^1XH`$ $`=2X`$
$`q^1HYqYH`$ $`=2Y`$
$`q^1XYqYX+{\displaystyle \frac{qq^1}{4}}H^2`$ $`=H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}K^2& 0& 0\\ 0& K^2& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& q\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^1\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^1& 0& 0& 0& 0& 0& 0& 0& 2q\\ \hfill X& 0& 1& 0& 0& 0& q& 0& 0& 0\\ \hfill Y& 0& 0& 1& 0& 0& 0& 0& q& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1+q^1u_2^2(1+q^1)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^11)u_2^1`$, $`(u_1^11)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H+{\displaystyle \frac{1q^2}{4}}\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q^1\omega _H\omega _X+q\omega _X\omega _H`$
$`q\omega _X\omega _Y+q^1\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q\omega _H\omega _Y+q^1\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`\omega _H\omega _X\omega _X\omega _H`$, $`\omega _H\omega _Y\omega _Y\omega _H`$, $`\omega _X\omega _Y\omega _Y\omega _X`$.
11. Quantum tangent space $`𝒳_𝚪`$:
$`H`$ $`:={\displaystyle \frac{2}{qq^1}}(\epsilon _{}K^21),`$ $`X`$ $`:=q^{1/2}FK,`$ $`Y`$ $`:=q^{1/2}KE`$
Real forms: $`𝒪(\mathrm{SU}_q(2))`$, $`𝒪(\mathrm{SU}_q(1,1))`$, $`𝒪(\mathrm{SL}_q(2,))`$
Fund. repr.: $`H={\displaystyle \frac{2}{1q}}\left(\begin{array}{cc}1& 0\\ 0& q\end{array}\right)`$, $`X=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)`$, $`Y=\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)`$
Relations:
$`qHXq^1XH`$ $`=2X`$
$`q^1HYqYH`$ $`=2Y`$
$`q^1XYqYX+{\displaystyle \frac{qq^1}{4}}H^2`$ $`=H`$
Module structure, differentials:
$`(f_j^i)=\left(\begin{array}{ccc}\epsilon _{}K^2& 0& 0\\ 0& K^2& 0\\ 0& 0& K^2\end{array}\right)`$ $`\begin{array}{cc}\hfill \mathrm{d}\omega _H=& q\omega _X\omega _Y\hfill \\ \hfill \mathrm{d}\omega _X=& 2q^1\omega _H\omega _X\hfill \\ \hfill \mathrm{d}\omega _Y=& 2q\omega _H\omega _Y\hfill \end{array}`$
Pairing:
$$\begin{array}{cccccccccc}& (u_1^1)^2& u_1^1u_2^1& u_1^1u_1^2& (u_2^1)^2& u_2^1u_1^2& u_2^1u_2^2& (u_1^2)^2& u_1^2u_2^2& (u_2^2)^2\\ & & & & & & & & & \\ \hfill H& 2q^1& 0& 0& 0& 0& 0& 0& 0& 2q\\ \hfill X& 0& 1& 0& 0& 0& q& 0& 0& 0\\ \hfill Y& 0& 0& 1& 0& 0& 0& 0& q& 0\end{array}$$
Right ideal $`𝓡_𝚪`$: $`u_1^1q^1u_2^2(1q^1)`$, $`(u_2^1)^2`$, $`(u_1^2)^2`$, $`u_2^1u_1^2`$, $`(u_1^11)u_2^1`$, $`(u_1^11)u_1^2`$
Left-invariant symmetric 2-forms:
$`\omega _H\omega _H+{\displaystyle \frac{1q^2}{4}}\omega _X\omega _Y`$ $`\omega _X\omega _X`$ $`q^1\omega _H\omega _X+q\omega _X\omega _H`$
$`q\omega _X\omega _Y+q^1\omega _Y\omega _X`$ $`\omega _Y\omega _Y`$ $`q\omega _H\omega _Y+q^1\omega _Y\omega _H`$
Braiding: $`(1\sigma )(q^2+\sigma )=0`$
$`\mathrm{ker}(q^2+\sigma ):`$ $`\omega _H\omega _X\omega _X\omega _H`$, $`\omega _H\omega _Y\omega _Y\omega _H`$, $`\omega _X\omega _Y\omega _Y\omega _X`$.
Acknowledgement: The author would like to thank A. Schüler for stimulating discussions.
|
warning/0006/cond-mat0006448.html
|
ar5iv
|
text
|
# Surface scaling behavior of isotropic Heisenberg systems: Critical exponents, structure factor, and profiles
## I Introduction
Classical spin systems with $`O(N)`$ symmetry display long - range magnetic order at low temperatures, if the spatial dimensionality is sufficiently high or the interaction range is sufficiently large. Long - range order sets in spontaneously below a certain critical temperature $`T=T_c`$ at zero magnetic field $`𝐇=\mathrm{𝟎}`$, where the field $`𝐇`$ has $`N`$ components. In the vicinity of the critical point $`(T,𝐇)=(T_c,\mathrm{𝟎})`$ these systems are characterized by an $`N`$ component order parameter given by the magnetization. Both from the theoretical and the experimental point of view these systems provide important realizations of the well known $`O(N)`$ universality class for critical behavior. In particular, the cases $`N=1`$, $`N=2`$, and $`N=3`$ are most relevant for experiments, because they correspond to the Ising universality class (magnets with uniaxial anisotropies, simple and binary fluids), the XY universality class (magnets with planar anisotropies, <sup>4</sup>He near the normal - superfluid transition), and the Heisenberg universality class (isotropic magnets, e.g. Ni, EuO, EuS), respectively. Real samples of any material always have surfaces which display a distinct critical behavior in the vicinity of the critical point of the bulk material. In turn, the penetration depth of surface effects is set by the bulk correlation length $`\xi =\xi _\pm ^0|\tau |^\nu `$ which diverges at the critical point $`\tau =(TT_c)/T_c0`$, where $`𝐇=\mathrm{𝟎}`$ is assumed. The exponent $`\nu `$ is the universal correlation length exponent and $`\xi _\pm ^0`$ denotes the nonuniversal correlation length amplitudes above $`(+)`$ and below $`()`$ $`T_c`$ ($`N=1`$ only). It is now well established that surface critical behavior obeys the priciple of universality, i.e., critical exponents and scaling functions which characterize surface critical behavior do not depend on microscopic details of the material . General field - theoretic considerations based on the $`\varphi ^4`$ Ginzburg - Landau model for semiinfinite systems show that for $`O(N)`$ symmetric systems there are three surface universality classes which are denoted as ordinary $`(O)`$, surface - bulk or special $`(SB)`$, and extraordinary $`(E)`$. These three universality classes can be briefly characterized by the absence or presence of surface order at the critical point of the bulk material. In particular, $`O`$ stands for a disordered surface and $`E`$ symbolizes an ordered surface, whereas $`SB`$ indicates that surface and bulk are critical simultaneously. The physical nature of the $`SB`$ surface universality class is that of a multicritical point . Surface order can arise spontaneously $`(E)`$ or can be imposed from outside by a surface field $`𝐇_1`$. The latter case is quite common in fluid systems with confining walls (Ising universality class, $`N=1`$) , where the bulk transition is then denoted as normal transition. It has been shown by rigorous arguments, that apart from corrections to scaling the normal and the extraordinary transition are equivalent . This is also quite important for the theory, because the normal transition gives access to the extraordinary surface universality class even when the surface does not order spontaneously, i.e., the $`SB`$ multicritical point does not exist.
Within the framework of $`O(N)`$ symmetric spin models on lattices we will only consider the case of short - range interactions which is characterized by a nearest - neighbor coupling constant $`J`$ in the bulk and by another nearest - neighbor coupling constant $`J_1`$ in the surface. In this case the absence or presence of the $`SB`$ transition crucially depends on the spatial dimension $`d`$ and the value of $`N`$. In $`d=2`$ dimensions the $`SB`$ transition does not exist, because the ”surfaces” in this case are only one - dimensional. Note that for $`N2`$ a $`O(N)`$ spin system does not show long - range order in $`d=2`$ . In $`d=3`$ the value of $`N`$ becomes crucial. For $`N=1`$ the $`SB`$ transition exists, because the surface, i.e., the Ising model in $`d=2`$, exhibits long - range order below a finite critical temperature. In the phase diagram of a semi - infinite Ising ferromagnet the $`SB`$ multicritical point is located at $`T=T_c`$, $`𝐇=𝐇_1=0`$ ($`N=1`$ here), and the critical value $`J_1^c/J`$ of the surface - to - bulk coupling ratio $`J_1/J`$. Note that besides the lines of the $`O`$ and $`E`$ transitions a line of surface transitions terminates in the $`SB`$ multicritical point . For the Ising ferromagnet one finds $`J_1^c/J=3/2`$ on a simple cubic lattice to a remarkably high accuracy . For $`N=2`$ the surface of the spin system is a two - dimensional XY model which does no longer show long - range order . Instead, a line of Kosterlitz - Thouless surface transitions occurs in the $`(T,J_1/J)`$ plane of the phase diagram which terminates in a $`SB`$ multicritical point at $`T=T_c`$ and $`J_1/J=J_1^c/J1.5`$ for a XY ferromagnet on a simple cubic lattice. For $`N=3`$ the $`SB`$ transition does not exist, because the two - dimensional Heisenberg model has no phase transition . In the absence of symmetry breaking surface interactions the isotropic Heisenberg ferromagnet in $`d=3`$ should therefore always display ordinary surface critical behavior, regardless of the value of $`J_1/J`$. In $`d4`$ dimensions, i.e., within mean - field theory, the $`SB`$ transition exists for any value of $`N`$.
The dimensional crossover behavior of the $`SB`$ multicritical point outlined above greatly complicates the field - theoretic description of the $`SB`$ transition for general $`N`$, apart from additional Goldstone problems below $`T_c`$ for $`N>1`$ which can be accounted for in finite systems with periodic boundary conditions . Therefore, most of the field - theoretic work is devoted to the Ising universality class and to the ordinary transition for general $`N1`$ . In the following we will focus on the ordinary transition in isotropic Heisenberg magnets and we therefore only summarize some aspects of the ordinary surface universality class for later reference. For a recent general survey of surface critical behavior the reader is referred to Ref..
Apart from the standard bulk critical exponents additional surface critical exponents must be introduced in order to describe critical behavior at the ordinary transition. Let $`m_1`$ denote the modulus of the surface magnetization per spin $`𝐌_1/\text{Area}`$ and let $`H`$ and $`H_1`$ denote the moduli of the bulk and surface field $`𝐇`$ and $`𝐇_1`$, respectively. Then the critical behavior of $`m_1`$ and of the susceptibilities $`\chi _1`$ and $`\chi _{11}`$ is given by
$`m_1`$ $``$ $`(\tau )^{\beta _1}`$ (1)
$`\chi _1={\displaystyle \frac{m_1}{H}}`$ $``$ $`|\tau |^{\gamma _1}`$ (2)
$`\chi _{11}={\displaystyle \frac{m_1}{H_1}}`$ $``$ $`|\tau |^{\gamma _{11}}`$ (3)
for $`H=H_1=0`$ and sufficiently small $`\tau `$. The suceptibilities $`\chi _1`$ and $`\chi _{11}`$ are known as the layer susceptibility and the surface or local susceptibility. In terms of the coordinates $`𝐱_{||}`$ parallel to the surface and $`z`$ perpendicular to the surface the correlation decay exponents $`\eta _{||}`$ and $`\eta _{}`$ are defined by
$`G(|𝐱_{||}𝐱_{}^{}{}_{||}{}^{}|,z,z)`$ $``$ $`|𝐱_{||}𝐱_{}^{}{}_{||}{}^{}|^{(d2+\eta _{||})}`$ (4)
$`G(0,z,z^{})`$ $``$ $`|zz^{}|^{(d2+\eta _{})}`$ (5)
in the limit of large distances, where $`T=T_c`$ and $`H=H_1=0`$ is assumed. The surface exponents defined in Eqs.(1) and (4) are not independent. By virtue of the scaling relations
$`\beta _1`$ $`=`$ $`\nu (d2+\eta _{||})/2`$ (6)
$`\gamma _1`$ $`=`$ $`\nu (2\eta _{})`$ (7)
$`\gamma _{11}`$ $`=`$ $`\nu (1\eta _{||})`$ (8)
$`\eta _{}`$ $`=`$ $`(\eta +\eta _{||})/2`$ (9)
only one of the surface exponents defined above, say, $`\eta _{||}`$ is independent for given $`\nu `$ and $`\eta `$, where $`\eta `$ is the decay exponent of the bulk correlation function according to $`G(|𝐱^{(1)}𝐱^{(2)}|)|𝐱^{(1)}𝐱^{(2)}|^{(d2+\eta )}`$ at $`T=T_c`$. The remaining magnetic exponents $`\delta _1`$ and $`\delta _{11}`$ defined by
$`m_1(T_c,H,H_1=0)`$ $``$ $`H^{1/\delta _1}`$ (10)
$`m_1(T_c,H=0,H_1)`$ $``$ $`H_1^{1/\delta _{11}}`$ (11)
are related to the bulk exponents $`\nu `$ and $`\eta `$ and the surface exponent $`\eta _{||}`$ according to $`\delta _1=\nu (d+2\eta )/(2\beta _1)`$ and $`\delta _{11}=\nu (d\eta _{||})/(2\beta _1)`$, respectively . In analogy with Eq.(1) the layer specific heat $`C_1`$ and the surface or local specific heat $`C_{11}`$ display the critical singularites
$`C_1={\displaystyle \frac{e_1}{t}}`$ $``$ $`|\tau |^{\alpha _1}`$ (12)
$`C_{11}={\displaystyle \frac{e_1}{c}}`$ $``$ $`|\tau |^{\alpha _{11}},`$ (13)
where $`e_1`$ is the surface energy density per spin and $`c=(J_1^cJ_1)/J`$ is the surface enhancement. Within the framework of the field - theoretic renormalization group it has been shown by an explicit calculation of the energy density profile and from the short - distance expansion of the energy density operator near the surface that $`\alpha _1=\alpha 1`$ and $`\alpha _{11}=\alpha 2\nu `$ at the ordinary transition. The critical singularities of all bulk and surface quantities can therefore be expressed by the bulk exponents $`\nu `$ and $`\eta `$ and the surface exponent $`\eta _{||}`$, i.e., there is only a single independent surface exponent at the ordinary transition . Within the field - theoretic renormalization group in $`d=4\epsilon `$ all surface exponents are known to two - loop order for general $`N`$ . The extrapolation of the $`\epsilon `$ \- expansion to $`d=3`$ $`(\epsilon =1)`$ already yields reasonable estimates for the surface exponents, however, their numerical accuracy remains onknown due to the lack of higher order results which are needed to apply resummation techniques. Therefore, alternative analytic approaches have been pursued in order to obtain improved estimates. In particular for the Heisenberg model $`\beta _1=0.81\pm 0.04`$ has been obtained which agrees with the extrapolated value for $`\beta _1`$ . The result of a perturbative calculation for the nonlinear sigma model in $`d=2+\epsilon `$ has been combined with the results in $`d=4\epsilon `$ by means of a Padé approximant resulting in the numerical estimates $`\beta _1=0.84\pm 0.01`$ and $`\eta _{||}=1.39\pm 0.02`$. A massive field-theoretic approach recently provided the new estimates $`\beta _1=0.880`$, 0.862, 0.889 from a \[2/0\], \[0/2\], and \[1/1\] Padé approximant in $`d=3`$, respectively . From an earlier experiment on Ni(100) and Ni(110) surfaces the average estimates $`\beta _10.81`$ and $`\beta _10.79`$ have been obtained, respectively.
For the experimental verification of surface critical behavior the struture factor in surfaces and thin films is of particular interest . From Eq.(4) one concludes that for small parallel momentum transfer $`p`$ the surface correlation function displays the algebraic $`p^{1+\eta _{||}}`$ singularity, which gives access to the exponent $`\eta _{||}`$ in a surface scattering experiment. Consequently, substantial theoretical effort has been spent on the theoretical understanding of correlation functions in surfaces and films and the crossover behavior between them . A thorough survey of the properties of the static structure factor can be found in Ref.. Scattering experiments also give access to the order parameter profile which is governed by universal shape functions. In $`d=2`$ and at $`T=T_c`$ the shape functions of the order parameter and the energy densitiy profile for Ising and Potts models can be obtained from conformal invariance considerations . Generalizations to $`O(N)`$ symmetric models in $`d=2`$ are also possible . In higher dimensions one has to resort to field - theoretic methods , where only the energy density profile is nonzero at the ordinary transition . The identification of the surface universality class for a specific system is a rather delicate problem. For example, the presence of weak surface fields leads to a crossover from ordinary to extraordinary behavior as the surface is approached from the interior . For multi - component systems, e.g., binary alloys the surface universality classes may even be different for different orientations of the surface .
Whereas most of the theoretical results quoted above can be applied to the case $`d=3`$, $`N=3`$ under consideration, numerical investigations have primarily focussed on the Ising universality class $`(N=1)`$ in $`d=3`$. From a Monte - Carlo simulation of the isotropic Heisenberg ferromagnet on a simple cubic lattice with an open surface on one side and a self - consistently determined surface field on the opposite side an estimate $`\beta _1=0.75\pm 0.10`$ has been found for the surface exponent $`\beta _1`$ (ordinary transition) . The field dependence of the surface magnetization $`m_1`$ at $`T=T_c`$ (see Eq.(10)) was determined later and the estimate $`\delta _1=2.3\pm 0.1`$ was found. Transfer matrix Monte - Carlo calculations for the Heisenberg ferromagnet on small lattices provided the estimate $`\beta _1=0.80\pm 0.03`$ . Most of the numerical effort has been spent on the surface critical behavior of the Ising model and a few detailed studies also exist for the XY model . The most accurate numerical estimates of Ising surface exponents and amplitude ratios at the ordinary and the SB transition can be found in Ref.. We close this brief overview with the remark that the surface exponent $`\beta _1`$ for the ordinary transition is not affected by surface bond disorder or the presence of steps on the surface . A theoretical explanation for this behavior can be found from the construction of upper and lower bounds on $`\beta _1`$ .
Although a substantial wealth of information is already available for the surface critrical behavior of the isotropic ferromagnet with $`O(N=3)`$ symmetry at the surface, the resulting estimates for the surface critical exponents are too disparate to provide a reliable basis for further investigations, e.g, the surface contribution to the dynamic structure factor. Furthermore, a systematic scaling analysis of static surface correlations is still missing, which provides essential information for the numerical analysis of experimental scattering data and Monte - Carlo data of the dynamic structure factor. Moreover, previous investigations were focussed on the asymptotic scaling regime of the ordinary surface universality class. In real systems or computer simulations of isotropic Heisenberg magnets, however, the surface properties at hand may be such that asymptotic scaling is only obtained after a crossover regime of a a certain width has been traversed. This also means that some sort of nonasymptotic surface behavior must occur before the crossover regime is reached even if the bulk is already critical, i.e., displays asymptotic scaling behavior governed by bulk exponents. It is important for the analysis of numerical data to localize these regimes and to describe their properties.
It is the purpose of this investigation to fill at least some of the aforementioned gaps. In particular new independent estimates for the surface critical exponents of the isotropic Heisenberg ferromagnet at the ordinary transition are provided. With Refs. and as guidelines for the universal form of the shape functions the surface structure factor and the order parameter and energy density profiles are analyzed at $`T=T_c`$ for various values of the surface - to - bulk coupling ratio $`J_1/J`$. Particular attention will be paid to the crossover regime, which is best described by the shape crossover of order parameter and energy density profiles. Outside the crossover regime a nonasymptotic scaling regime is investigated and characterized by nonuniversal $`J_1/J`$ dependent exponents. This work is mainly focussed on scaling properties and therefore the simulations are restricted to simple cubic lattices with (100) surfaces.
## II Model and simulation method
The model Hamiltonian describes an isotropic Heisenberg ferro - or antiferromagnet on a simple cubic lattice with two (100) surfaces in a cubic geometry. It is given by (see also Ref.)
$$=J\underset{i,jV\backslash A}{}𝐒_i𝐒_jJ_1\underset{i,jA}{}𝐒_i𝐒_j,$$
(14)
where $`V`$ denotes the set of all lattice sites (volume) and $`A`$ denotes the set of surface sites. The coupling constants $`J`$ and $`J_1`$ are assumed to have the same sign and the spins $`𝐒_i=(S_i^x,S_i^y,S_i^z)`$ with $`|𝐒_i|=1`$ only interact with nearest neighbors. A nearest neigbor pair $`i,j`$ of spins is part of the interior $`V\backslash A`$ of the system, if at least one of the spins is not part of the surface $`A`$. A nearest neighbor pair $`i,j`$ is part of the surface if both spins belong to $`A`$. The surfaces are defined by the lattice planes $`(x,y,z=1)`$ and $`(x,y,z=L)`$, where $`1x,yL^{}`$. Along the $`x`$ and $`y`$ directions periodic boundary conditions are applied. In order to avoid unwelcomed finite - size effects the geometry of the lattice is chosen as cubic, i.e., $`L^{}=L`$. The cubic shape turns out to be a reasonable compromise between achievable system sizes $`L`$ and the sensitivity of the system to the surface - to - bulk coupling ratio $`J_1/J`$.
The Monte - Carlo algorithm is chosen as a hybrid scheme which consists of Metropolis sweeps, Wolff single cluster updates , and overrelaxation sweeps . Typically, one hybrid Monte - Carlo step consists of 10 individual steps each of which can be one of the updates listed above. The Metropolis and the Wolff algorithm work the standard way, where the reduced coordination number of the lattice at the surfaces and the modified surface coupling $`J_1`$ must be taken into account. The acceptance probability $`p`$ of a proposed spin flip in the Metropolis algorithm is defined by $`p(\mathrm{\Delta }E)=1/[\mathrm{exp}(\mathrm{\Delta }E/k_BT)+1]`$, where $`k_B`$ is the Boltzmann constant and $`\mathrm{\Delta }E`$ is the change in configurational energy of the proposed move. The overrelaxation part of the algorithm performs a microcanonical update of the configuration by sequentially rotating each spin in the lattice such that its energy contribution to the energy of the whole configuration remains constant. The implementation of this update scheme is straightforward, because according to Eq.(14) the energy of a spin with respect to its neighborhood has the functional form of a scalar product. The angle of rotation can be chosen randomly for each spin, however, it turns out that in view of minimal autocorrelation times a reflection, i.e., a rotation of all spins by 180 degrees is the most efficient overrelaxation move. One hybrid Monte - Carlo step consists of two Metropolis (M) for single cluster Wolff (C) and four overrelaxation (O) updates. The individual updates are mixed automatically in the program to generate the update sequence M O C O C M O C O C . The random number generator is the shift register generator R1279 defined by the recursion relation $`X_n=X_{np}X_{nq}`$ for $`(p,q)=(1279,1063)`$. Generators like these are known to cause systematic errors in combination with the Wolff algorithm . However, for lags $`(p,q)`$ used here these errors are far smaller than typical statistical errors. They are further reduced by the hybrid nature of the algorithm .
The Monte - Carlo scheme described above is employed for lattice sizes $`L`$ between $`L=12`$ and $`L=72`$. For each choice of parameters we perform at least 20 blocks of $`10^3`$ hybid steps for equilibration followed by $`10^4`$ hybrid steps for measurements. Each measurement block yields an estimate for all quantities of interest and from these we obtain our final estimates and estimates of their statistical error following standard procedures. The integrated autocorrelation time of the hybrid algorithm is determined by the autocorrelation function of the energy or, equivalently, the modulus of the magnetization, which yield the slowest modes for the Wolff algorithm. The autocorrelation times do not exceed 10 hybrid Monte - Carlo steps for the largest lattice size $`(T=T_c)`$, so the equilibration and measurement periods quoted above translate to roughly 100 and 1000 autocorrelation times, respectively. In order to obtain the best statistics for all magnetic quantities a measurement is made after every hybrid Monte - Carlo step. All error bars quoted in the following correspond to one standard deviation. The hybrid scheme samples the surfaces of the system reasonably often, so a preferential sampling of surface configurations is not required. The simulations have been performed on DEC alpha and AXP workstations at the Physics department of the BUGH Wuppertal.
## III Surface scaling exponents
The simulations presented here have been performed at $`T=T_c`$ for several values of the surface - to - bulk coupling ratio $`J_1/J`$. The estimate for $`T_c`$ used here is taken from Ref., where the critical coupling $`K_c`$ has been determined as $`K_cJ/(k_BT_c)=0.693035(37)`$. In view of the limited system size $`L72`$ the relative accuracy of $`10^4`$ in $`T_c`$ is sufficient in order to perform a standard finite - size scaling analysis in terms of the system size $`L`$ up to the usual corrections to scaling. As our main reference for bulk critical exponents we choose the work of Guida and Zinn - Justin , where the bulk critical exponents of the $`O(N)`$ universality class have been obtained from high order Borel resummed perturbation theory for the Ginzburg - Landau model.
In order to access surface critical exponents the surface magnetization $`m_1`$, its second moment, and the surface energy density $`e_1`$ has been measured. The layer and surface specific heats $`C_1`$ and $`C_{11}`$ (see Eq.(12)) have not been investgated, because $`\alpha _1`$ and $`\alpha _{11}`$ can be expressed by bulk exponents only. The layer magnetization $`𝐦(z)`$, $`1zL`$ and the total magnetization $`𝐦_{tot}`$ are defined by
$`𝐦(z)`$ $``$ $`{\displaystyle \underset{x,y=1}{\overset{L}{}}}𝐒_{xyz}/L^2`$ (15)
$`𝐦_{tot}`$ $``$ $`{\displaystyle \underset{z=1}{\overset{L}{}}}𝐦(z)/L,`$ (17)
respectively, and $`m_{tot}|𝐦_{tot}|`$ denotes the modulus of the total magnetization. The magnetization profile $`m(z)`$ is then defined as the projection of $`𝐦(z)`$ onto the total magnetization $`𝐦_{tot}`$, i.e.,
$$m(z)𝐦(z)𝐦_{tot}/m_{tot}$$
(18)
and $`m_1(m(1)+m(L))/2`$ defines the surface magnetization. Note that the surfaces at $`z=1`$ and $`z=L`$ are identical (see Eq.(14)). In terms of $`m_{tot}`$ and $`m_1`$ the layer and surface susceptibilities (i.e., their longitudinal components) $`\chi _1`$ and $`\chi _{11}`$ for a completely finite system are defined as
$`\chi _1`$ $`=`$ $`L^2(m_{tot}m_1m_{tot}m_1)/(k_BT)`$ (19)
$`\chi _{11}`$ $`=`$ $`L^2(m_1^2m_1^2)/(k_BT),`$ (20)
where $`\mathrm{}`$ denotes the thermal average. The energy profile $`e(z)`$ is defined accordingly, where apart from the exchange energy between the spins within layer $`z`$ half the interaction energy to the layers $`z1`$ and $`z+1`$ also contributes to $`e(z)`$, so $`e_{tot}_{z=1}^Le(z)/L`$ is the total energy density. In the following all energies will be given in units of $`k_BT_c`$, i.e., extra factors $`k_BT`$ (see, e.g, Eq.(19)) are unity at $`T=T_c`$.
The surface magnetization $`m_1`$ at $`T=T_c`$ as function of the system size $`L`$ is shown in Fig.1 for $`J_1/J=0.3`$, 0.5, 1.0, 1.5, and 2.0. For $`J_1/J1`$ the functional form of $`m_1(L)m_1`$ is accurately captured by
$$m_1(L)=B_{m_1}L^{\beta _1/\nu }\left(1+C_{m_1}L^\omega \right),$$
(21)
where $`B_{m_1}`$ is the magnetization amplitude and $`C_{m_1}`$ is the amplitude of the leading correction to scaling. The associated Wegner exponent $`\omega =0.78`$ is taken from Ref.. A least square fit of Eq.(21) to the data for $`J_1/J1`$ displayed in Fig.1 yields the estimates $`\beta _1/\nu =1.185(6)`$, 1.175(13), and 1.171(7) for $`J_1/J=0.3`$, 0.5, and 1.0, respectively. The error indicated in parenthesis corresponds to one standard deviation. From these estimates one obtaines the weighted average $`\beta _1/\nu =1.179(6)`$, where the smallest of the individual errors is taken as the final error estimate. Two of the three individual estimates are included in the error interval of the final estimate. From the literature value $`\nu =0.7073(35)`$ one obtains the estimate
$$\beta _1=0.834(6)$$
(22)
for the surface exponent of the magnetization. For $`J_1/J=1.5`$ Eq.(21) does not capture the functional form of $`m_1(L)`$, because within the available range of lattice sizes the system undergoes a crossover towards the asymptotic ordinary surface critical behavior. A more detailed discussion of this crossover is postponed to Sec. 5, where the order parameter and energy density profiles are presented. If the decay of $`m_1(L)`$ for $`J_1/J=1.5`$ is described by an effective exponent according to Eq.(21), one finds a value around 0.6 for $`L20`$ and a value around 0.9 for $`L48`$. This indicates that only a part of the full crossover process is captured by the simulation. This leads to the conclusion that the data for $`J_1/J=2.0`$ have not yet even entered the crossover regime to ordinary surface critical behavior. It is insructive to compare these data with corresponding data for the Ising model. The surface - to - bulk coupling ratio $`J_1/J=2.0`$ already belongs to the extraordinary regime of the Ising model , where the surface exhibits long - range order at $`T=T_c`$. The comparison is shown in Fig.2, where the data for an Ising model according to Eq.(14) have been obtained from a hybrid algortihm which corresponds to the one described above, except that overrelaxation moves cannot be performed in this case . The surface magnetization decays with an effective exponent of about 0.16 (see Fig.2(a), solid line), whereas for the Ising model (see Fig.2(b), solid line) $`m_1(L)`$ approaches the spontaneous surface magnetization $`m_{10}`$ according to
$$m_1(L)=m_{10}B_{m_1}^IL^{\beta /\nu }$$
(23)
up to corrections to scaling, where $`\beta /\nu 0.517`$ is the scaling dimension of the order parameter in the Ising universality class and $`B_{m_1}^I`$ is a nonuniversal amplitude. Fig.2 illustrates how the presence of real long - range surface order (b) can be distinguished from spurious long - range surface order (a) which only appears as a nonasymptotic finite - size effect. However, within a typical range of numerically accessible system sizes the crossover to the asymptotic ordinary surface behavior cannot be observed for $`J_1/J2.0`$ and therefore a data analysis within the framework of finite - size scaling only yields an effective exponent $`\beta _{1,eff}=\beta _{1,eff}(J_1/J)`$. For $`J_1/J=2.0`$ one has $`\beta _{1,eff}(2.0)/\nu 0.16`$ (see Fig.2(a)) and for $`J_1/J=3.0`$ an effective exponent $`\beta _{1,eff}(3.0)/\nu 0.08`$ is found (not shown). We do not quote error bars here, because the estimates for $`\beta _{1,eff}(J_1/J)`$ are presumably affected by systematic errors which are larger than the statistical ones. In the interior of the system the ’bulk’ magnetization $`m_bm(L/2)`$ obeys standard critical finite size scaling with the exponent $`\beta /\nu 0.518`$ (see Eqs.(37) and (38) and Fig.10 in Sec. 5).
The layer susceptibility $`\chi _1`$ is analyzed in the same way as the surface magnetization $`m_1(L)`$. The data are displayed in Fig.3. For $`J_1/J1`$ the data can be interpreted according to
$$\chi _1(L)=B_{\chi _1}L^{\gamma _1/\nu }\left(1+C_{\chi _1}L^\omega \right),$$
(24)
which is the exact analog of Eq.(21). The term in parenthesis captures the leading correction to scaling, $`\gamma _1/\nu `$ is the corresponding surface exponent for finite - size scaling, and $`B_{\chi _1}`$ is a nonuniversal overall amplitude. With $`\omega 0.78`$ as above a least square fit of Eq.(24) to the data shown in Fig.3 yields the estimates $`\gamma _1/\nu =1.314(23)`$, 1.305(12), and 1.308(25) for $`J_1/J=0.3`$, 0.5, and 1.0, respectively. As before we adopt the weighted average $`\gamma _1/\nu =1.307(12)`$ as our final estimate, where the smallest of the individual errors is taken as the error estimate. From the literature value $`\nu =0.7073(35)`$ the estimate
$$\gamma _1=0.924(10)$$
(25)
is found. For $`J_1/J1.5`$ the data qualitatively behave as in Fig.1: For $`J_1/J=1.5`$ the system undergoes a crossover to ordinary surface behavior and for $`J_1/J=2.0`$ the asymptotic behavior is out of reach for the simulation. Nonetheless, the increase of $`\chi _1`$ in the latter case can be described by an effective exponent $`\gamma _{1,eff}(J_1/J=2.0)/\nu 2.31`$. Likewise, $`\gamma _{1,eff}(J_1/J=3.0)/\nu 2.44`$ is obtained (not shown). Error bars are not quoted for the reasons indicated above.
The exponets $`\beta _1`$ and $`\gamma _1`$ are not independent. From the scaling relations given by Eq.(6) and bulk scaling relations one can infer the simple rule $`\beta _1+\gamma _1=\beta +\gamma `$. From the literature one obtains $`\beta +\gamma =1.7557(56)`$ and Eqs.(22) and (25) yield $`\beta _1+\gamma _1=1.758(11)`$ which verifies the above scaling law. The effective exponents $`\beta _{1,eff}(2.0)`$ and $`\gamma _{1,eff}(2.0)`$ for $`J_1/J=2.0`$ (see Figs.1 and 3) yield $`\beta _{1,eff}(2.0)+\gamma _{1,eff}(2.0)1.75`$ which is remarkably close to the value of $`\beta +\gamma `$ quoted above. Likewise, $`\beta _{1,eff}(3.0)+\gamma _{1,eff}(3.0)1.78`$ is found for $`J_1/J=3.0`$. From Eq.(6) one furthermore obtains the correlation exponents
$$\eta _{||}=1.358(12),\eta _{}=0.697(6)$$
(26)
and the exponent
$$\gamma _{11}=0.253(9)$$
(27)
of the surface susceptibility $`\chi _{11}`$. At $`T=T_c`$ the surface susceptibility therefore behaves according to
$$\chi _{11}(L)=\chi _{110}B_{\chi _{11}}L^{1\eta _{||}}\left(1+C_{\chi _{11}}L^\omega \right),$$
(28)
where the term in parenthesis captures the leading correction to scaling. Fits of Eq.(28) to the data for $`J_1/J1`$ are shown in Fig.4, where $`1\eta _{||}=0.358`$ and $`\omega =0.78`$ are kept fixed and the amplitudes $`\chi _{110}`$, $`B_{\chi _{11}}`$ and $`C_{\chi _{11}}`$ are used as fit parameters. For $`J_1/J1.5`$ Eq.(28) does no longer describe $`\chi _{11}`$ due to the strong nonasymptotic effects described above (for more details see Sec. 4). A comparison of different estimates for $`\beta _1`$ from various sources is shown in Table I. The estimates are rather consistent within their mutual errors, but the best agreement is found with the field theoretic estimate given in Ref..
We close this section with an investigation of the finite-size behavior of the surface energy density $`e_1`$. The data for the dimensionless surface energy density $`\epsilon _1=e_1/(k_BT_c)`$ are shown in Fig.5 for $`0.5J_1/J2.0`$. The $`L`$ dependence of $`e_1`$ for $`J_1/J=0.3`$ (not shown), 0.5 (a), and 1.0 (b) is qualitatively different from the $`L`$ dependence for $`J_1/J=1.5`$ (c), 2.0 (d), and 3.0 (not shown). The leading $`L`$ dependence of $`e_1`$ is written as
$$\epsilon _1(L)=\epsilon _{10}+B_{e_1}^{(2)}L^2+B_{e_1}^{(3)}L^3,$$
(29)
where the amplitudes $`\epsilon _{10}`$, $`B_{e_1}^{(2)}`$, and $`B_{e_1}^{(3)}`$ are taken as fit parameters. For $`J_1/J1`$ the coefficient $`B_{e_1}^{(2)}`$ turns out to be two orders of magnitude smaller than the coefficient $`B_{e_1}^{(3)}`$ which suggests that $`L^3`$ is the leading finite - size correction in this case. According to naive finite - size scaling one expects $`x_{e_1}=(1\alpha _1)/\nu `$ as the leading finite - size exponent for the surface energy density. In fact, due to the scaling relation $`\alpha _1=\alpha 1`$ for the ordinary transition one finds $`x_{e_1}=(2\alpha )/\nu =d=3`$ from hyperscaling in agreement with the simulation data in Figs.5 (a) and (b). In contrast, $`B_{e_1}^{(2)}`$ and $`B_{e_1}^{(3)}`$ are of roughly equal magnitude (and of opposite sign) for the case $`J_1/J1.5`$ displayed in Figs.5 (c) and (d). Therefore, $`L^2`$ rather than $`L^3`$ is the leading finite - size correction here. The discrepancy between Figs.5 (a), (b) and Figs.5 (c), (d) is due to strong nonasymptotic contributions to the finite size behavior for $`J_1/J1.5`$. One possible source of the $`L^2`$ contribution is the scaling dimension $`x_{f_s}=(2\alpha _s)/\nu =(2\alpha \nu )/\nu =d1=2`$ of the surface free energy $`f_s`$ (surface tension) which prevails in Eq.(29) due to the nonscaling dependence of the surface tension on $`J_1/J`$ . Another source of the leading $`L^2`$ dependence is the regular (noncritical) finite - size behavior of the surface tension for periodic boundary conditions . Other finite - size corrections such as $`L^1`$ substantially reduce the goodness - of - fit and can therefore be ruled out as leading terms. Possible logarithmic corrections of the form $`L^2\mathrm{log}L`$ cannot be identified unambigiously.
## IV Surface structure factor
The surface structure factor is given by the discrete Fourier transform of the surface spin - spin correlation function
$`G_1^{\alpha \alpha }(𝐱𝐱^{})`$ $`=`$ $`(S_{x,y,1}^\alpha S_{x^{},y^{},1}^\alpha S_{x,y,1}^\alpha ^2`$ (30)
$`+`$ $`S_{x,y,L}^\alpha S_{x^{},y^{},L}^\alpha S_{x,y,L}^\alpha ^2)/2,`$ (31)
where the Fourier transform is taken with respect to $`𝐱𝐱^{}`$, $`𝐱=(x,y)`$, and the upper index $`\alpha `$ refers to the spin component. Note that off - diagonal components of $`G_1`$ vanish identically. The longitudinal and transverse components of $`G_1^{\alpha \alpha }`$ with respect to the total magnetization are very similar and we therefore restrict the following discussion to the transverse component only. The momentum transfer $`𝐩`$ in the (100) direction, i.e., parallel to the surface, in units of the inverse lattice constant is given by $`𝐩=(p,0,0)=(2\pi n/L,0,0)`$, $`n=0,1,\mathrm{},L/2`$. Numerical data of the surface structure factor for $`J_1/J=0.5`$ in the (100) direction for three different lattice sizes are displayed in Fig.6. The data essentially collapse onto a single curve for $`p>0`$, finite - size effects are only visible at $`p=0`$, where the structure factor shown in Fig.6 reduces to the transverse component of the surface susceptibility $`\chi _{11}`$ (see Eq.(28)). For $`J_1/J=0.3`$ identical properties are obtained (not shown). It turns out that lattice effects in the structure factor near the Brillouin zone boundary can be captured by the replacement $`p2\mathrm{sin}(p/2)`$ to a remarkable accuracy . The shape of the surface structure factor is reasonably well captured by the mean - field type expression
$`S_1^\alpha (p,L)`$ $`=`$ $`\chi _{110}^\alpha B_{\chi _{11}}^\alpha {\displaystyle \frac{c^\alpha 2\mathrm{sin}(p/2)}{c^\alpha +2\mathrm{sin}(p/2)}}`$ (32)
$`\times `$ $`\left[{\displaystyle \frac{2\mathrm{\Gamma }^\alpha \mathrm{sin}(p/2)}{\mathrm{tanh}\left(2\mathrm{\Gamma }^\alpha \mathrm{sin}(p/2)L\right)}}\right]^{\eta _{||}1},`$ (33)
where the upper index $`\alpha `$ indicates the component (longitudinal $`l`$ or transverse $`t`$ with respect to $`𝐦_{tot}`$) of $`S_1^\alpha (p,L)`$. The coefficients $`\chi _{110}^\alpha `$ and $`B_{\chi _{11}}^\alpha `$ are the coefficients of the surface susceptibility $`\chi _{11}^\alpha `$ (for $`\alpha =l`$ see Eq.(28)). The width parameter $`\mathrm{\Gamma }^\alpha `$ and the surface enhancement parameter $`c^\alpha `$ are used to fit the momentum dependence of the surface structure factor and the exponent $`\eta _{||}`$ is taken from Eq.(26). The fit is performed in two stages. First, $`\chi _{110}^\alpha `$ and $`B_{\chi _{11}}^\alpha `$ are determined from a least square fit of Eq.(32) for $`p=0`$ to the surface susceptibility. Second, the width parameter $`\mathrm{\Gamma }^\alpha `$ and, if needed (see below), the surface enhancement parameter $`c^\alpha `$ are adjusted to obtain a least square fit of the momentum dependence to the data. In practice, this fit procedure has only been performed for the largest lattice size $`L=72`$. For smaller systems the fit parameters are taken from $`L=72`$ in order to test the accuracy of the shape predicted by Eq.(32). The result for $`J_1/J=0.5`$ is displayed in Fig.6 for the transverse component $`S_1^t(p,L)`$ of the surface structure factor. For the largest system $`L=72()`$ the short dashed line shows the fit, whereas for other lattice sizes such as $`L=24(\times )`$ and $`48(+)`$ the evaluation of Eq.(32) for $`L=24`$ (solid line) and $`L=48`$ (long dashed line) confirms the predicted shape to a remarkable accuracy. For $`J_1/J=0.3`$ and 0.5 the parameter $`c^\alpha `$ is very large and therefore the surface enhancement prefactor $`(c^\alpha 2\mathrm{sin}(p/2))/(c^\alpha +2\mathrm{sin}(p/2))`$ can be omitted. Noticeable deviations between the data and Eq.(32) for $`\alpha =t`$ only occur for small $`p`$ on the smaller lattices, where additional lattice corrections to Eq.(32) may become important. However, for any $`p>0`$ the data for all system sizes collapse onto a single curve, which can be obtained from Eq.(32) by performing the limits $`c^\alpha \mathrm{}`$ and $`L\mathrm{}`$ at finite $`p`$.
For $`J_1/J=1.0`$ the surface enhancement parameter $`c^\alpha `$ becomes important. The data are shown in Fig.7, where the surface enhancement prefactor provides an important correction to the momentum dependence of $`S_1^\alpha (p,L)`$. The fit procedure works as described above, but deviations from the assumed shape for small $`p`$ also occur for $`L=72`$. The surface structure factor is more sensitive to crossover phenomena occurring for larger values of $`J_1/J`$ than the surface quantities discussed in the previous section. Corrections to scaling, which are not included in Eq.(32), may also account for part of the deviations between the data and the simple model for the shape function. Note that the analytic results obtained in Ref. do not hold for the cubic geometry used here. Furthermore, nonasymptotic surface enhancement corrections, which have not been considered in Ref., become essential for the data analysis. For $`p>0`$ the data still collapse onto a single curve, which is given by Eq.(32) in the limit $`L\mathrm{}`$ at finite $`p`$ and finite $`c^\alpha 15`$.
The behavior of $`S_1^t(p,L)`$ for $`J_1/J=2.0`$ is shown in Fig.8, for $`J_1/J=3.0`$ similar results have been obtained (not shown). Although scaling appears to be valid to a very high degree of accuracy, the behavior of $`S_1^t(p,L)`$ is very different from Figs.6 and 7. First, the surface susceptibility grows according to the power law
$$\chi _{11}^\alpha (L)=B_{\chi _{11}}^\alpha L^{1\eta _{||,eff}(J_1/J)}$$
(34)
rather than approaching a finite limit as in Eq.(28). From a least square fit of Eq.(34) to $`\chi _{11}^t`$ one obtains the effective exponent $`\eta _{||,eff}(J_1/J=2.0)0.64`$, where deviations from the pure power law given by Eq.(34) are very small. For $`J_1/J=3.0`$ one obtains $`\eta _{||,eff}(3.0)0.82`$. We refrain from quoting error bars here, because the values for $`\eta _{||,eff}(J_1/J)`$ may be affected by systematic errors due to corrections to Eq.(34) of unknown form. Mutual interaction between the surfaces mediated by the bulk may also cause systematic errors. With $`B_{110}^t`$ and $`\eta _{||,eff}(J_1/J)`$ taken from Eq.(34) the shape function
$`S_1^\alpha (p,L)`$ $`=`$ $`B_{\chi _{11}}^\alpha {\displaystyle \frac{c^\alpha 2\mathrm{sin}(p/2)}{c^\alpha +2\mathrm{sin}(p/2)}}`$ (35)
$`\times `$ $`\left[{\displaystyle \frac{2\mathrm{\Gamma }^\alpha \mathrm{sin}(p/2)}{\mathrm{tanh}\left(2\mathrm{\Gamma }^\alpha \mathrm{sin}(p/2)L\right)}}\right]^{\eta _{||,eff}(J_1/J)1},`$ (36)
is fitted to $`S_1^t`$, where the remaining two parameters $`c^\alpha `$ and $`\mathrm{\Gamma }^\alpha `$ are used. As described above the fit is only performed for $`L=72`$ (short dashed line in Fig.8). For $`L=24(\times )`$ and $`L=48(+)`$ Eq.(35) ($`\alpha =t`$) is shown for $`L=24`$ (solid line) and $`L=48`$ (long dashed line), where all fit parameters are kept fixed. The shape function given by Eq.(35) is remarkably accurate. However, the observed scaling is completely different from the asymptotic scaling shown in Fig.6 and, apart from surface enhancement corrections, in Fig.7. The effective exponent $`\eta _{||,eff}(J_1/J)`$ is not related to the surface exponent $`\eta _{||}`$ given by Eq.(26), because it is nonuniversal, i.e., it depends on $`J_1/J`$. From general considerations it is tempting to pose the (effective) surface scaling law $`\beta _{1,eff}(J_1/J)=\nu [d2+\eta _{||,eff}(J_1/J)]/2`$ (see Eq.(6)). The direct determination of $`\beta _{1,eff}`$ from $`m_1(L)`$ for $`J_1/J2.0`$ is plagued with considerable uncertainties, because the decay of $`m_1(L)`$ with $`L`$ becomes quite slow for $`J_1/J2.0`$. Systematic errors may exceed the formal statistical error of a fit in this case and therefore the aforementioned effective scaling law cannot be confirmed unambigiously. Note that the crossover regime to the asymptotic scaling remains unaccessible within the range of system sizes used here. Nonetheless, the accuracy of the scaling law for $`S_1(p,L)`$ for the nonasymptotic regime still lacks theoretical understanding.
For $`J_1/J=1.5`$ none of the above descriptions applies to the data and scaling appears to be violated. This is in accordance with the findings of Sec. 3, where $`m_1(L)`$ and $`\chi _1(L)`$ display sizable deviations from simple scaling laws. The effects of the crossover to the ordinary surface universality class are particularly pronounced in the shape function of the order parameter profile to which we turn in the following section.
## V Profiles
The order parameter profile $`m(z)`$ (see Eq.(18)) provides local information about the order in the system. Furthermore, $`m(z)`$ is by construction very sensitive to the system size at $`T=T_c`$ and should therefore be a valuable probe for scaling behavior which is easier to interpret than the structure factor. For $`J_1/J1`$ all previous investigations have shown that the system essentially displays the asymptotic critical behavior of the ordinary surface universality class. The magnetization profile confirms this again, so the scaling analysis can be restricted to the case $`J_1/J=0.5`$. The scaled magnetization profile $`M(z/L)`$ is defined by
$$M(z/L)m(z)/m_b,$$
(37)
where
$$m_bm(L/2)=B_{m_b}L^{\beta /\nu }\left(1+C_{m_b}L^\omega \right).$$
(38)
The factor in parenthesis captures corrections to scaling and the exponents $`\beta `$, $`\nu `$, and $`\omega `$ are taken from Ref.. The coefficient $`C_{m_b}`$ is determined by a least square fit of Eq.(38) to $`m_b`$. For numerical convenience $`z1/2`$, $`z=1,\mathrm{},L`$ is chosen as the position coordinate for the profile. The scaling plot for $`J_1/J=0.5`$ is shown in Fig.9, where $`z`$ now refers to the shifted layer index. The data scale very accurately and the shape of the profile is in accordance with the expectation for the ordinary transition. The scaling function $`M(z/L)`$ can be represented by the simple fit formula
$$M(\zeta )=B_M\left[(\zeta +\zeta _0)(1\zeta +\zeta _0)\right]^{(\beta _1\beta )/\nu },$$
(39)
where $`\zeta _0=z_0/L`$ and $`z_0`$ is the extrapolation length, $`\beta `$ and $`\nu `$ are taken from Ref., and $`\beta _1`$ is taken from Eq.(22). From a least square fit of Eq.(39) to the data for $`L=72`$ one finds $`z_00.26`$ in units of the lattice constant. For $`J_1/J=0.3`$ and 1.0 one finds $`z_00.34`$ and $`z_00.46`$, respectively. The choice $`0.3J_1/J1`$ therefore yields quite accurate realizations of Dirichlet boundary conditions, which characterize the ordinary surface universality class.
For $`J_1/J=2.0`$ the application of Eqs.(37) and (38) to the data yields the result shown in Fig.10. Except at the surface layers scaling is fulfilled very accurately, however, the shape of $`m(z)`$ does not show the expected fixed point form. Instead, a strong enhancement of the surface magnetization over the magnetization in the interior is obtained. A behavior like this is typical for the extraordinary transition which does not occur for the Heisenberg model defined by Eq.(14) in $`d=3`$, i.e., the system displays spurious long-range surface order. If the lattice size $`L`$ could be increased further one should therefore find a crossover to the ordinary behavior displayed in Fig.9 (see Fig.2). Nonetheless, the shape of the scaling function $`M(\zeta )`$ is again captured by Eq.(39), if the exponent is used as a third fit parameter. It is tempting to write this exponent in the form $`(\beta _{1,eff}(J_1/J)\beta )/\nu `$, however, the result for $`\beta _{1,eff}(J_1/J)`$ obtained this way is not compatible with the estimate obtained from $`m_1(L)`$, which may be due to systematic errors of various kinds discussed above.
The question how the system actually performs the crossover is answered in Fig.11, where $`M(z/L)`$ is shown for $`J_1/J=1.5`$. For smaller systems the surface magnetization is still enhanced over $`m_b`$, but as $`L`$ is increased, the maximum of the profile at $`z=L/2`$ finally exceeds $`m_1`$ and the profile shape approaches the fixed point form shown in Fig.9. The influence of the surface coupling on $`m(z)`$ is confined to the two outermost lattice layers on either side of the cube, whereas the curvature of the remainder of the profile already has the “correct” sign for all lattice sizes shown in Fig.11. For $`L<24`$ the profile becomes flatter in the middle and the profile shape approaches the one displayed in Fig.10. Note that even at $`z=L/2`$ the data fail to collapse according to Eqs.(37) and (38). Although the isotropic Heisenberg model does not display long - range surface order in the thermodynamic limit, on a finite lattice a strong enhancement of the surface magnetization $`m_1`$ over the bulk magnetization $`m_b`$ does occur for sufficiently strong surface couplings as a finite - size effect for a certain range of system sizes.
The scaling properties of the energy density profile are a little more delicate, because a background energy density must be subtracted from the profile in order to obtain scaling. One finds the scaling form
$$E(z/L)(\epsilon (z)\epsilon _0)/L^{(1\alpha )/\nu },$$
(40)
where $`E(\zeta )`$ is the scaling function and corrections to scaling have been disregarded. For numerical convenience $`z1/2`$, $`z=1,\mathrm{},L`$ is again chosen as the position coordinate for the profile. The scaling plots for $`0.3J_1/J1.0`$ are well represented by the scaling plot for $`J_1/J=0.5`$ which shown in Fig.12. As in Figs.9 \- 11 $`z`$ refers to the shifted layer index. The shape of the scaling function is as expected for the ordinary universality class . For $`L>24`$ the data collapse reasonably well which confirms scaling according to Eq.(40), where $`(1\alpha )/\nu 1.586`$ according to Ref.. The shape of the scaling function $`E(\zeta )`$ can be approximated by the fit formula (see also Ref.)
$$E(\zeta )=B_E\left[\pi /\mathrm{sin}\left(\pi \frac{\zeta +\zeta _0}{1+2\zeta _0}\right)\right]^{(1\alpha )/\nu },$$
(41)
where $`\zeta _0=z_0/L`$ is the scaled extrapolation length of the profile (see Eq.(39)). For small arguments $`\zeta =z/L`$ Eq.(41) captures the algebraic increase of $`\epsilon (z)`$ correctly and the extrapolation length $`z_0`$ becomes negligibly small. However, away from the surface Eq.(41) captures the shape function of the energy density profile only in a qualitative sense.
For $`J_1/J=2.0`$ scaling of the data according to Eq.(40) works even better as shown in Fig.13. Note that the value of the reference energy density $`\epsilon _0`$ does not depend on $`J_1/J`$. The data for system sizes $`L24`$ collapse onto a single curve, which is represented by Eq.(41) with a much better accuracy than for $`J_1/J1.0`$. The extrapolation length $`z_00.12`$ (in units of the lattice constant) is still very small. The behavior the energy density profile displayed in Fig.13 is typical for a system at the extrordinary transition which does not exist for the Heisenberg model in $`d=3`$ with nearest neighbor interactions. According to the above discussion the shape of the energy density profile given by Fig.13 is governed by the presence of spurious long - range order in the surface. The crossover to the aymptotic shape (see Fig.12) will occur if $`L`$ is increased further. For $`J_1/J=2.0`$ the crossover regime is out of reach, but for $`J_1/J=1.5`$ this crossover takes place within the range of accessible system sizes as shown in Fig.14. As in Fig.11 scaling is violated. The curvature of the profile near $`z=L/2`$ changes sign between $`L=36`$ and $`L=60`$ and for $`L60`$ the profile approaches its asymptotic shape. The nonasymptotic surface effects are more pronounced here than for the magnetization profile and penetrate deeper into the system, but the magnitude of the surface induced enhancement of the energy density at the surface decays quickly with increasing $`L`$. The same crossover behavior can be observed for the critical Ising model slightly below the $`SB`$ multicritical point for, e.g., $`J_1/J=1.45`$ .
## VI Summary and outlook
In the absence of symmetry breaking fields the asymptotic critical scaling behavior of surfaces of a critical $`d=3`$ dimensional Heisenberg magnet with short - range interactions is always governed by the ordinary surface universality class. For finite systems, however, a crucial interplay between the available system size and the value of the surface - to - bulk coupling ratio $`J_1/J`$ determines whether or not the asymptotic surface scaling behavior can actually be observed. Within the range of system sizes $`12L72`$ for the $`L\times L\times L`$ geometry used here critical behavior in the ordinary surface universality class can be observed for $`J_1/J1.0`$. The scaling exponents found numerically in this case are consistent with rigorous scaling laws and estimates for theses exponents found previously by various analytical and numerical methods. The shape of the surface structure factor $`S_1(p,L)`$ is captured by very simple mean - field like expressions, in which two amplitudes are fixed by a fit to the surface susceptibility for $`p=0`$. A width parameter and, if needed, a surface enhancement parameter then determine the shape of the momentum dependence. It turns out, that finite - size and lattice effects are very accurately described by the pseudo scaling argument $`2\mathrm{sin}(p/2)L`$ which replaces the true scaling argument $`pL`$. These properties of $`S_1(p,L)`$ are essential for the data interpretation of the dynamic surface structure factor which is the key quantity for the interpretaion of neutron scattering data on magnetic surfaces and will therefore be the main focus of ensuing work. The order parameter and energy density profiles are less relevant for experiments, but they are easier to interpret and yield valuable insight into the scaling behavior of the system. Either profile is found to scale in accordance with the ordinary surface universality class.
For $`J_1/J2.0`$ scaling is still found for all quantities under investigation, however, the scaling exponents are replaced by effective ones and their values depend on $`J_1/J`$. The scaling relations cannot be verified unambigiously, because some of the numerical estimates for the effective exponents are presumably affected by systematic errors of unknown magnitude. Such errors may ensue due to unknown corrections to the effective scaling laws or due to an effective interaction between the two surfaces mediated by the bulk system in between. Nontheless, the effective scaling properties of the surface structure factor $`S_1(p,L)`$ provide valuable information for the analysis of its dynamic counterpart. The striking scaling properties found here still await theoretical explanation.
Coupling ratios $`J_1/J2.0`$ are too large to access the crossover regime from the state of enhanced surface magnetization (spurious long \- range surface order) to the asymptotic (ordinary) surface scaling. For $`J_1/J=1.5`$, however, this crossover becomes the dominating feature in the finite - size behavior of all quantities under investigation, at least within the range of system sizes used here. The value $`J_1/J=1.5`$ only marks the location of the crossover regime for the system sizes at hand rather than a sharp transition in the surface behavior of the Heisenberg model. In the crossover regime scaling is violated and further theoretical insight is needed for a purposeful data analysis. Scaling of bulk quantities is also affected by particularly large correction terms. The crossover process itself is best visualized in the shape crossover of the magnetization and the energy density profiles, which occurs as $`L`$ is increased for $`J_1/J=1.5`$ and $`T=T_c`$.
###### Acknowledgements.
The author gratefully acknowledges financial support of the major part of this work through the Heisenberg program of the Deutsche Forschungsgemeinschaft under grant # Kr 1322/2-1.
|
warning/0006/hep-ph0006316.html
|
ar5iv
|
text
|
# Interference pattern of the Coulomb and the strong Van der Waals forces in p-p scattering11footnote 1NUP-A-2000-13
## 1 Introduction
In the composite model of hadron, in which the fundamental constructive force is a strong or super-strong Coulombic force, the quantum fluctuation of the composite states causes the strong Van der Waals interaction between hadrons. Historically before 1960’s, hadrons were regarded as elementary particles and the interactions arose from the exchange of mesons with masses, therefore the forces were inevitably short range. After the introduction of the composite model of hadron, the idea of the short range force is taken over to the new hadron physics, because when the momentum transfer is small and we do not explore the inside of hadrons, there must be no differences whether the hadrons are composite or elementary. Although the short range nature of the interactions between hadrons are widely believed, it cannot be true when the strong Van der Waals forces are acting between hadrons. The purpose of the present paper is to propose an experiment to confirm the existence of the strong Van der Waals interaction between nucleons by observing the characteristic interference in the low energy (20 $``$ 30 MeV.) proton-proton scattering. The interference pattern arises from the destructive interference between the repulsive Coulomb and the attractive strong Van der Waals forces, and the cross section $`\mathrm{\Delta }\sigma /\sigma `$ has a narrow dip in the neighborhood of $`\theta _{c.m.}=14^{}`$ with the depth around one per cent.
In the confirmation of the strong Van der Waals interaction, it is desirable to have recourse to the difference of analytic structure of the scattering amplitude $`A(s,t)`$ in the neighborhood of $`t=0`$. For the case of the short range force $`A(s,t)`$ is regular at $`t=0`$ and the nearest singularity occurs at $`t=t_{min}`$ , where $`(t_{min})^{1/2}`$ is the smallest mass exchanged in the t-channel. On the other hand when the long range force is acting and the asymptotic form of the potential is $`V(r)C/r^\alpha `$ , an extra singularity occurs at $`t=0`$ in $`A(s,t)`$ and whose spectral function behaves as $`A_t(s,t)=C^{}t^\gamma +\mathrm{}`$ in the threshold region. In the next section, we shall see that the powers $`\alpha `$ and $`\gamma `$ are related by $`\alpha =2\gamma +3`$. Therefore to observe the extra singularity at $`t=0`$ implies the confirmation of the long range force. Moreover since $`t=0`$ is the end point of the physical region $`4\nu t0`$, we can recognize the extra singularity, when the sufficiently precise data are available, without making any analytic continuation.
In particular if the long range interaction is the Van der Waals potential of the London type ($`\alpha =6`$), the amplitude $`A(s,t)`$ must have a singular term $`C^{}(t)^{3/2}`$. Since $`t=2\nu (1z)`$, there are two ways to observe the extra singularity of the amplitude $`(2\nu )^{3/2}(1z)^{3/2}`$. The first one is to make the partial wave projection and to observe the singular threshold behavior $`\nu ^{3/2}`$ in the partial wave amplitudes $`a_{\mathrm{}}(\nu )`$. The second one is to fix $`\nu `$ and to observe the anomalous angular dependency $`(1z)^{3/2}`$, which has a singularity at $`z=1`$. By analysing the once subtracted S-wave amplitude of the p-p scattering $`(a_0(\nu )a_0(0))/\nu `$, the long range force was searched in the previous paper, and we observed a cusp of the form $`(c_0c_1\sqrt{\nu })`$ at $`\nu =0`$ which was characteristic to the Van der Waals interaction of the London type. In section 2, we shall briefly review the previous search of the strong Van der Waals force using the once subtracted partial wave amplitude, and the parameters of the long range interaction will be given. In section 3, by using the parameters of the long range force, the anomalous angular distritution of the amplitude of the p-p scattering is computed, and the characteristic interference pattern of the cross section is predicted. Section 4 will be used for remarks and comments.
## 2 Search for the strong Van der Waals force in the S-wave amplitude
When the scattering amplitude $`A(s,t)`$ has the extra singularity $`C^{}(t)^\gamma `$, the partial wave amplitudes $`a_{\mathrm{}}(\nu )`$ also have extra singularities at $`\nu =0`$ and the threshold behaviors become $`C_{\mathrm{}}^{\prime \prime }\nu ^\gamma `$, where $`C_{\mathrm{}}^{\prime \prime }`$ are proportional to $`C^{}`$. Since in the hadron physics the data of the S-wave phase shift of the p-p scattering are prominent in their accuracy and the scattering length is determined very precisely, we shall analyse the once subtracted S-wave amplitude $`(a_0(\nu )a_0(0))/\nu `$ in search for the extra singularity $`C_0^{\prime \prime }\nu ^{\gamma 1}`$. Our normalization of the partial wave amplitude $`a_0(\nu )`$ is
$$a_0(\nu )=\frac{\sqrt{m^2+\nu }}{X_0(\nu )i\sqrt{\nu }}$$
(1)
for the scattering of neutral particles, and $`X_0(\nu )`$ is the effective range function $`\sqrt{\nu }\mathrm{cot}\delta _0(\nu )`$ which is regular at $`\nu =0`$ and accepts the Taylor expansion of $`\nu `$. In order to observe the singular behavior at $`\nu =0`$, we must remove the known nearby singularities. First of all we have to remove the unitarity cut by constructing a function
$$\frac{K_0(\nu )}{\nu }\frac{a_0(\nu )a_0(0))}{\nu }\frac{1}{\pi }_0^{\mathrm{}}\frac{\mathrm{Im}a_0(\nu ^{})}{\nu ^{}(\nu ^{}\nu )}𝑑\nu ^{}$$
(2)
, which is free from the right hand cut and is called the Kantor amplitude. To facilitate the search of the extra singularity it is desirable to remove also the cut of the one-pion exchange (OPE). The procedure is the same as the case of the unitarity cut, namely first compute an integration
$$\frac{K_0^{1\pi }(\nu )}{\nu }\frac{1}{\pi }_{\mathrm{}}^{1/4}\frac{\mathrm{Im}a_0^{1\pi }(\nu ^{})}{\nu ^{}(\nu ^{}\nu )}𝑑\nu ^{}=\frac{1}{4}\frac{g^2}{4\pi }\{\frac{1}{4\nu }\mathrm{log}(1+4\nu )1\}$$
(3)
, and then subtract it from the amplitude. In the calculation, the neutral pion mass is set equal to 1, and throughout this paper we shall use the neutral pion mass as the unit of the energy and the momentum. Since the two-pion exchange spectrum starts slowly at $`\nu =1`$, the function $`(K_0(\nu )K_0^{1\pi }(\nu ))/\nu `$ must be almost constant and have small slope in the neighborhood of $`\nu =0`$ when the long range interactions are absent. On the other hand, for the Van der Waals interaction of the London type ( $`\alpha =6`$ ) $`\gamma =3/2`$ and $`(K_0(\nu )K_0^{1\pi }(\nu ))/\nu `$ must have a singular term $`C^{\prime \prime }\nu ^{1/2}`$, therefore we can observe a cusp at $`\nu =0`$ as long as the coefficient $`C^{\prime \prime }`$ is not very small. In this way we can examine the long range force if the $`\pi `$-N coupling constant $`g^2/4\pi `$ and the S-wave phase shifts are given. Because the effects of the Coulomb and the vacuum polarization potentials are not considered, this method is applicable to the case where the Coulomb interaction is not important such as to the neutron-neutron scattering.
Let us turn to the proton-proton scattering, where the vacuum polarization as well as the Coulombic interactions are important. The Kantor amplitude introduced in Eq.(2) still has the left hand cuts, namely the Coulombic cut in $`\mathrm{}<\nu 0`$ and the cut of the vaccum polarization in $`\mathrm{}<\nu m_e^2`$. The difficulties are by-passed if we use the modified effective range function $`X_0(\nu )`$ of the proton-proton scattering, which is regular at $`\nu =0`$ and accepts the effective range expansion, when all the forces are short range except for the terms of the Coulomb and of the vacuum polarization. The modified effective range function $`X_0(\nu )`$ for the phase shift $`\delta _0^E(\nu )`$ is
$$X_0(\nu )=\frac{C_0^2\sqrt{\nu }}{1\varphi _0}\{(1+\chi _0)\mathrm{cot}\delta _0^E\mathrm{tan}\tau _0\}+me^2h(\eta )+me^2\mathrm{}_0(\eta ).$$
(4)
In Eq.(3), two well-known functions with the Coulombic order of magnitudes appear, they are expressed using a new variable $`\eta =me^2/(2\sqrt{\nu })`$ :
$$C_0^2=\frac{2\pi \eta }{e^{2\pi \eta }1}andh(\eta )=\eta ^2\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}\frac{1}{\mathrm{}(\mathrm{}^2+\eta ^2)}\mathrm{log}\eta 0.57722\mathrm{}.$$
(5)
In Eq.(4) $`\tau _0`$ is the phase shift due to the vacuum polarization potential
$$V^{vac}(r)=\lambda \frac{e^2}{r}_{4m_e^2}^{\mathrm{}}𝑑t\frac{e^{r\sqrt{t}}}{2t}(1+\frac{2m_e^2}{t})\sqrt{1\frac{4m_e^2}{t}}\lambda \frac{e^2}{r}I(r),$$
(6)
where $`m_e`$ is the mass of the electron and $`\lambda =2e^2/3\pi =1.549\times 10^3`$. Functions $`\tau _0`$, $`\chi _0`$, $`\varphi _0`$ and $`\mathrm{}_0(\eta )`$ have the order of magnitudes of the vaccum polarization, and introduced in the previous paper.
By using the modified effective range function $`X_0(\nu )`$, we define the S-wave amplitude $`a_0(\nu )`$ of the p-p scattering by
$$a_0(\nu )=\frac{\sqrt{m^2+\nu }}{X_0(\nu )me^2h(\eta )i\sqrt{\nu }C_0^2}.$$
(7)
The relation between $`a_0(\nu )`$ and the phase shift $`\delta _0^E`$ is obtained if we substitute $`X_0(\nu )`$ of Eq.(4) into Eq.(7), and which reduces to the well-known form
$$a_0(\nu )=\frac{1}{C_0^2}\frac{\sqrt{m^2+\nu }}{\sqrt{\nu }}e^{i\delta _0^E(\nu )}\mathrm{sin}\delta _0^E(\nu ),$$
(8)
if the functions related to the vacuum polarization are neglected. The form of $`a_0(\nu )`$ of Eq.(7) is the same as that of the neutron-neutron scattering $`(\sqrt{m^2+\nu }/\sqrt{\nu })e^{i\delta }\mathrm{sin}\delta `$ except for the factor $`C_0^2`$ given in Eq.(4), which is the penetration factor. If we compare the S-wave amplitude of the p-p scattering of Eq.(7) with that of the n-n scattering Eq.(1), a combination of functions $`(me^2h(\eta )i\sqrt{\nu }C_0^2)`$ appears in place of $`i\sqrt{\nu }`$. In order to investigate the analytic structure of $`a_0(\nu )`$, it is convenient to rewrite the combination as
$$me^2h(\eta )i\sqrt{\nu }C_0^2=i\sqrt{\nu }+me^2\{\mathrm{log}(i\eta )\psi (1+i\eta )\}.$$
(9)
Since the digamma function $`\psi (z)`$ has poles at non-positive integers, the poles on the $`\eta `$-plane appear on the positive imaginary axis. In terms of $`\sqrt{\nu }`$, which is $`me^2/(2\eta )`$, the series of poles appear on the negative imaginary axisis and converge to $`\sqrt{\nu }=0`$. It is the smallness of the fine structure constant $`e^2`$ and therefore of the residues of such poles that zeros of the denominator of Eq.(7) occur at points very close to the locations of the poles of $`(me^2h(\eta )i\sqrt{\nu }C_0^2)`$. Therefore the partial wave amplitude $`a_0(\nu )`$ of the p-p scattering has a series of poles on the second sheet of $`\nu `$, namely on the lower half plane of $`\sqrt{\nu }`$, whereas on the first sheet of $`\nu `$ the analytic structure of $`a_0(\nu )`$ does not change compared to the case of the n-n scattering. This fact implies that the same definition of the Kantor amplitude $`K_0(\nu )`$ introduced for the neutron-neutron scattering, which is given in Eq.(2), is valid also for the proton-proton scattering, as long as we evaluate Im$`a_0(\nu ^{})`$ of Eq.(2) from Eqs.(7) and (9). Therefore the Kantor amplitude of the p-p scattering $`K_0(\nu )`$ constructed in this way is free from the singularities in the neighborhood of $`\nu =0`$, and so does not have the cut of the vacuum polarization as well as that of the Coulomb interaction.
We can now compute the once subtracted S-wave Kantor amplitude minus the contribution from the one-pion exchange of the proton-proton scattering:
$$\stackrel{~}{K}_0^{once}(\nu )\frac{K_0(\nu )}{\nu }\frac{K_0^{1\pi }(\nu )}{\nu }.$$
(10)
In figure 1 and figure 2, $`\stackrel{~}{K}_0^{once}(\nu )`$ is plotted against $`T_{lab}`$. The graphs exhibit a cusp at $`\nu =0`$, which is characteristic to the attractive long range force. The cusp is fitted by a spectral function of three parameters:
$$A_t^{extra}(s,t)=\pi C^{}t^\gamma e^{\beta t}$$
(11)
, and the results of the chi-square fit in $`0.6MeV.<T_{lab}<125MeV.`$ are
$$\gamma =1.543,\beta =0.06264andC^{}=0.1762$$
(12)
in the unit of the neutral pion Compton wave length. The curve in fig.1 and $`(L)_3`$ curve in fig.2 are the fits by the spectral function of the long range force $`A_t^{extra}(s,t)`$ given in Eqs.(11) and (12), and the $`\chi `$-value per data point is 0.441. On the other hand, other curves in fig.2 are the three parameter fits by the spectral function of the short range forces:
$$A_t(s,t)=\underset{i=1}{\overset{3}{}}c_i\delta (tt_i),$$
(13)
where $`c_i`$ are free parameters and three $`t_i`$ are 4, 9 and 16 for the curve $`(sa)_3`$ (dotted curve), whereas $`t_i`$ are 9, 16 and 25 for the curve $`(sb)_3`$ (dashed curve), and the $`\chi `$-values of the fits per data points are 1.82 and 3.11 respectively. The curves indicate that the short range spectra, which mimic the spectrum of the two-pion exchange, cannot reproduce the cusp of $`\stackrel{~}{K}_0^{once}(\nu )`$ shown in figures 1 and 2. Details of the fits are found in the previous paper.
## 3 Interference pattern of the cross section of p-p scattering
In the previous section, we observed a cusp at $`\nu =0`$ in $`\stackrel{~}{K}_0^{once}(\nu )`$ which arises from the partial wave projection of the singular term $`C^{}(2\nu )^\gamma (1z)^\gamma `$. However the same singularity can also be observed in the angular distribution for fixed $`\nu `$. The aim of this paper is to propose to observe the singularity at $`z=\pm 1`$ and to confirm the existence of the strong Van der Waals force in the p-p scattering. Since the Coulomb potential also gives rise to the poles at $`z=\pm 1`$, we can expect to observe the characteristic interference pattern of the singular behaviors in the cross section. In this section, we shall compute such a pattern by using the parameters of the long range force determined in the previous section. If we consider that the energy dependent phase shift data are extracted from the measurements of different laboratories, the observation of the singular behavior in the angular distribution is the more direct way to confirm the existence of the long range force.
The amplitude due to the extra spectrum of Eq.(11) is
$`{\displaystyle \frac{m}{\sqrt{\nu }}}W(\nu ,z)={\displaystyle \frac{\pi C^{}}{\pi }}{\displaystyle _0^{\mathrm{}}}𝑑t^{}{\displaystyle \frac{t^\gamma e^{\beta t^{}}}{t^{}+2(1z)}}`$ (14)
$`=`$ $`C^{}(2\nu )^\gamma (1z)^\gamma \mathrm{\Gamma }(\gamma +1)\mathrm{\Gamma }(\gamma ,2\nu \beta (1z),\mathrm{})\mathrm{exp}[2\nu \beta (1z)]`$
, where $`\mathrm{\Gamma }(x,a,b)`$ is the incomplete gamma function defined by $`_a^bt^{x1}e^t𝑑t`$. The cross section $`\sigma (\theta )`$ of the proton-proton scattering in the low energy region is
$`\nu \sigma (\theta )`$ $`=`$ $`{\displaystyle \frac{1}{4}}|\stackrel{~}{f}^s(\theta )+2\mathrm{exp}[i\delta _0^E]\mathrm{sin}\delta _0^E+\alpha _e(z)|^2+`$ (15)
$`+`$ $`{\displaystyle \frac{3}{4}}|\stackrel{~}{f}^t(\theta )+6\delta _{1,C}z+\alpha _o(z)|^2+\mathrm{\Delta }.`$
$`\stackrel{~}{f}^s(\theta )`$ and $`\stackrel{~}{f}^t(\theta )`$ are the Coulomb amplitude plus the one-pion exchange contribution of the spin-singlet and the spin-triplet states respectively, and they are
$`\stackrel{~}{f}^s(\theta )`$ $`=`$ $`{\displaystyle \frac{\eta }{2}}\{{\displaystyle \frac{1}{S^2}}\mathrm{exp}[i\eta \mathrm{log}S^2]+{\displaystyle \frac{1}{C^2}}\mathrm{exp}[i\eta \mathrm{log}C^2]\}+`$ (16)
$`+`$ $`f^2m\sqrt{\nu }\{({\displaystyle \frac{1}{1t}}+{\displaystyle \frac{1}{1u}}){\displaystyle \frac{2}{2\nu }}Q_0(1+{\displaystyle \frac{1}{2\nu }})\}`$
and
$`\stackrel{~}{f}^t(\theta )`$ $`=`$ $`{\displaystyle \frac{\eta }{2}}\{{\displaystyle \frac{1}{S^2}}\mathrm{exp}[i\eta \mathrm{log}S^2]{\displaystyle \frac{1}{C^2}}\mathrm{exp}[i\eta \mathrm{log}C^2]\}`$ (17)
$``$ $`{\displaystyle \frac{f^2}{3}}m\sqrt{\nu }\{({\displaystyle \frac{1}{1t}}+{\displaystyle \frac{1}{1u}}){\displaystyle \frac{6}{2\nu }}Q_1(1+{\displaystyle \frac{1}{2\nu }})z\},`$
, in which
$$S=\mathrm{sin}\frac{\theta }{2},C=\mathrm{cos}\frac{\theta }{2}and\eta =\frac{e^2}{\mathrm{}c\beta _L},$$
(18)
where $`\beta _L`$ is the velocity in the laboratory system and it is written in terms of $`\nu `$
$$\beta _L=\frac{2\sqrt{\nu }\sqrt{1+\frac{\nu }{m^2}}}{m(1+\frac{2\nu }{m^2})}.$$
(19)
In Eq.(15) $`\alpha _e(z)`$ and $`\alpha _e(z)`$ are the even and odd functions of $`z`$ which take care of the contributions from the higher partial waves and for very small $`\nu `$ they are negligible. When the hadron interaction is short range, since the poles of the one-pion exchanges are already separated, the nearest singulary of $`\alpha (z)`$’s on the $`z`$-plane occurs at
$$z=\pm (1+\frac{t_{min}}{2\nu })$$
(20)
with $`t_{min}=4`$, the threshold of the spectrum of the two-pion exchange. As an example, we shall fix the incident energy at $`T_{lab}=`$20 MeV., namely at $`\nu =0.515`$. In such a case, the analytic domain on the $`z`$-plane is the Lehmann ellipse whose semi-major axis is around 5 and foci are at $`z=\pm 1`$. However if the two-pion exchange spectrum is very small in the threshold region and the main cotribution comes from the $`\sigma `$-meson (560 MeV.), then $`t_{min}=16`$ and the semi-major axis is around 17. For $`\alpha _e(z)`$ it is sufficient to retain only the S and D waves to reproduce the amplitude within the error $`10^3`$ in the physical region $`1z1`$. For the odd function $`\alpha _o(z)`$ we shall consider two cases, the first one is to retain only the P wave, whereas the second one is to retain the P and F waves.
To observe the extra singularity at $`z=\pm 1`$, we firstly determine the coefficients of the polynomials $`\alpha _e(z)`$ and $`\alpha _o(z)`$ using the cross section in the off-forward region $`z_1zz_1`$. Notation $`\sigma _{smooth}(z)`$ will be used for the cross section which is obtained by using the polynomials thus determined, because this is the smooth continuation of the cross section from the off-forward region to whole the physical region $`1z1`$. When the nuclear forces are short range, the observed cross section must coincide with $`\sigma _{smooth}(z)`$ in the forward region $`z_1z1`$ within the error. On the other hand, if the long range force exists and the extra singularity is of the form $`(1z)^\gamma `$, then the observed cross section deviates from $`\sigma _{smooth}(z)`$ in the forward region. We can evaluate such a deviation if the spectral functions of the long range force Eqs.(11) and (12) are given.
By using the amplitude of the long range force $`W(\nu ,z)`$ introduced in Eq.(14), $`\alpha (z)`$’s are written as
$$\alpha _e(z)=(W(\nu ,z)+W(\nu ,z)2w_02w_2z^2)+c_0^{}+c_2^{}z^2$$
(21)
and
$$\alpha _o(z)=(W(\nu ,z)W(\nu ,z)2w_1z)+c_1^{}z,$$
(22)
in which $`w_i`$’s are chosen in such a way that $`(w_0+w_1z+w_2z^2)`$ becomes the best fit to $`W(\nu ,z)`$ in the given off-forward region $`z_1zz_1`$. In particular for $`z_1=0`$, $`w_0`$, $`w_1`$ and $`w_2`$ are the value, slope and curvature at $`z=0`$ respectively. With these $`\alpha (z)`$’s we can compute the cross section $`\sigma _{long}(z)`$ from Eq.(15). The relative deviation of the cross section $`\mathrm{\Delta }\sigma /\sigma `$ is defined by
$$\mathrm{\Delta }\sigma /\sigma (\theta )=\frac{\sigma _{long}(z)\sigma _{smooth}(z)}{\sigma _{smooth}(z)}$$
(23)
and if we retain only the first order of the small quantities $`\alpha _e(z)`$, $`\alpha _o(z)`$ and $`\mathrm{\Delta }`$ in Eq.(15), the relative deviation reduces to
$`\mathrm{\Delta }\sigma /\sigma (\theta )`$ $`=`$ $`\stackrel{~}{g}^s(\theta )(W(\nu ,z)+W(\nu ,z)2w_02w_2z^2)+`$ (24)
$`+`$ $`\stackrel{~}{g}^t(\theta )(W(\nu ,z)W(\nu ,z)2w_1z),`$
where
$$\stackrel{~}{g}^s(\theta )=\frac{1}{2\sigma _0(\theta )}\mathrm{Re}(\stackrel{~}{f}^s(\theta )+2\mathrm{exp}[i\delta _0^E]\mathrm{sin}\delta _0^E)$$
(25)
and
$$\stackrel{~}{g}^t(\theta )=\frac{3}{2\sigma _0(\theta )}\mathrm{Re}(\stackrel{~}{f}^t(\theta )+6\delta _{1c}z).$$
(26)
In Eqs.(25) and (26), $`\sigma _0(\theta )`$ is the zeroth order of the cross section, and obtained by setting the samall quantities $`\alpha _e(z)`$, $`\alpha _o(z)`$ and $`\mathrm{\Delta }`$ in Eq.(15) equal to zero.
In figure 3 and 4, the coefficient functions $`\stackrel{~}{g}^s(\theta )`$ and $`\stackrel{~}{g}^t(\theta )`$ are displayed, in which we use the phase shifts $`\delta _0^E=50.96^{}`$ and $`\delta _{1c}=0.516^{}`$ at $`T_{lab}=`$ 20 MeV. , and also the $`\pi `$-N coupling constant $`g^2/4\pi =14.4`$ where $`g^2/4\pi =4m^2f^2`$.
In figure 5, the relative deviation $`\mathrm{\Delta }\sigma /\sigma (\theta )`$ is plotted against $`\theta `$, the scattering angle in the center of mass system. The numbers attached to the curves are $`\theta _1`$ the boundary of the off-forward region $`\theta _1\theta 180^{}\theta _1`$, in which the parameters in $`\alpha (z)`$’s are determined. The curves show narrow dips at $`\theta =15^{}`$ with the depth around 1 per cent. This is our prediction of the interference pattern derived from the parameters of the long range force. If the long range force does not exist, the curves of the relative deviation $`\mathrm{\Delta }\sigma /\sigma (\theta )`$ must be zero within the error.
In figure 6, the curves $`\mathrm{\Delta }\sigma /\sigma (\theta )`$, in which the F-wave as well as the P-wave term in $`\alpha _o(z)`$ are retained, are shown. The dips appear at $`\theta =13^{}`$.
## 4 Remarks and Comments
In this paper we propose to measure precisely the angular distribution of the cross section of the low energy proton-proton scattering, in order to confirm the long range interaction in the nuclear force. By using the parameters of the spectrum of the long range force obtained from the analysis of the S-wave phase shift of the p-p scattering, we predict the characteristic interference pattern in the angular distribution of the p-p cross section, which has a dip at $`\theta _{c.m.}=14^{}`$ with the depth around 1 %. The observation of such a pattern is the more direct way to confirm the long range force, because the energy dependent phase shifts are obtained from the data of different laboratories by constructing a consistent curve, in which the correction factor of the incident beam is assigned to each experiment. If we consider the precision of the measurement the low energy proton-proton is an ideal place to observe the strong Van der Waals force.
Another good place to observe the Van der Waals force is the low energy ($`T_{lab}`$ 1 MeV.) neutron-Pb scattering. This is because the strength of the long range potential is magnified by a factor A, the mass number, and it is relatively easy to observe the anomaly of the angular distribution. Since the Van der Waals force is universal, we can expect to observe such a force also in other processes such as in the $`\pi `$-$`\pi `$ scattering. Although the precisions of the $`\pi `$-$`\pi `$ data are not very high, the P-wave amplithde of the $`\pi `$-$`\pi `$ is another place easy to observe the strong Van der Waals force, because we can compute the spectral function of the two-pion exchange from the $`\pi `$-$`\pi `$ data, and by removing the spectrum from the amplitude we can prepare the wide domain of analyticity. If all the forces are short range, the Kantor amplitude must be almost constant.
In figure 7, the once subtracted Kantor amplitude $`K_1(\nu )/\nu `$ of the P-wave pi-pi scattering is shown along with the contribution from the two-pion exchange spectrum. Here again the cusp of the attractive sign appears. By making the chi-square fit, the parameters are determined.
$$\gamma =1.95,C^{}=0.0161and\beta =0.144$$
(27)
The long range force in $`\pi `$-$`\pi `$ is close to the Van der Waals force of the Casimir-Polder type rather than the London type.
|
warning/0006/math0006164.html
|
ar5iv
|
text
|
# Avoiding maximal parabolic subgroups of 𝑆_𝑘
## 1. Introduction and Main Result
Let $`[p]=\{1,\mathrm{},p\}`$ denote a totally ordered alphabet on $`p`$ letters, and let $`\alpha =(\alpha _1,\mathrm{},\alpha _m)[p_1]^m`$, $`\beta =(\beta _1,\mathrm{},\beta _m)[p_2]^m`$. We say that $`\alpha `$ is order-isomorphic to $`\beta `$ if for all $`1i<jm`$ one has $`\alpha _i<\alpha _j`$ if and only if $`\beta _i<\beta _j`$. For two permutations $`\pi S_n`$ and $`\tau S_k`$, an occurrence of $`\tau `$ in $`\pi `$ is a subsequence $`1i_1<i_2<\mathrm{}<i_kn`$ such that $`(\pi _{i_1},\mathrm{},\pi _{i_k})`$ is order-isomorphic to $`\tau `$; in such a context $`\tau `$ is usually called the pattern. We say that $`\pi `$ avoids $`\tau `$, or is $`\tau `$-avoiding, if there is no occurrence of $`\tau `$ in $`\pi `$. Pattern avoidance proved to be a useful language in a variety of seemingly unrelated problems, from stack sorting \[Kn, Ch.~2.2.1\] to singularities of Schubert varieties \[LS\]. A natural generalization of single pattern avoidance is subset avoidance; that is, we say that $`\pi S_n`$ avoids a subset $`TS_k`$ if $`\pi `$ avoids any $`\tau T`$. A complete study of subset avoidance for the case $`k=3`$ is carried out in \[SS\]. For $`k>3`$ situation becomes more complicated, as the number of possible cases grows rapidly. Recently, several authors have considered the case of general $`k`$ when $`T`$ has some nice algebraic properties. Paper \[BDPP\] treats the case when $`T`$ is the centralizer of $`k1`$ and $`k`$ under the natural action of $`S_k`$ on $`[k]`$ (see also Sec. 3 for more detail). In \[AR\], $`T`$ is a Kazhdan–Lusztig cell of $`S_k`$, or, equivalently, the Knuth equivalence class (see \[St, vol.~2, Ch.~A1\]). In this paper we consider the case when $`T`$ is a maximal parabolic subgroup of $`S_k`$.
Let $`s_i`$ denote the simple transposition interchanging $`i`$ and $`i+1`$. Recall that a subgroup of $`S_k`$ is called parabolic if it is generated by $`s_{i_1},\mathrm{},s_{i_r}`$. A parabolic subgroup of $`S_k`$ is called maximal if the number of its generators equals $`k2`$. We denote by $`P_{l,m}`$ the (maximal) parabolic subgroup of $`S_{l+m}`$ generated by $`s_1,\mathrm{},s_{l1},s_{l+1},\mathrm{},s_{l+m1}`$, and by $`f_{l,m}(n)`$ the number of permutations in $`S_n`$ avoiding all the patterns in $`P_{l,m}`$. In this note we find an explicit expression for the generating function of the sequence $`\{f_{l,m}(n)\}`$.
To be more precise, we prove the following more general result. Let us denote $`\sigma =s_1s_2\mathrm{}s_{k1}`$, that is, $`\sigma =(2,3,\mathrm{},k,1)`$, and let $`a`$ be an integer, $`0ak1`$ (here and in what follows $`k=l+m`$). We denote by $`f_{l,m}^a(n)`$ the number of permutations in $`S_n`$ avoiding the left coset $`\sigma ^aP_{l,m}`$; in particular, $`f_{l,m}^0(n)`$ coincides with $`f_{l,m}(n)`$. Let $`F_{l,m}^a(x)`$ denote the generating function of $`\{f_{l,m}^a(n)\}`$,
$$F_{l,m}^a(x)=\underset{n0}{}f_{l,m}^a(n)x^n.$$
Recall that the Laguerre polynomial $`L_n^\alpha (x)`$ is given by
$$L_n^\alpha (x)=\frac{1}{n!}e^xx^\alpha \frac{d^n}{dx^n}\left(e^xx^{n+\alpha }\right),$$
and the rook polynomial of the rectangular $`s\times t`$ board is given by
$$R_{s,t}(x)=s!x^sL_s^{ts}(x^1)$$
for $`st`$ and by $`R_{s,t}(x)=R_{t,s}(x)`$ otherwise (see \[Ri, Ch.~7.4\]).
###### Main Theorem
Let $`\lambda =\mathrm{min}\{l,m\}`$, $`\mu =\mathrm{max}\{l,m\}`$, then
$$F_{l,m}^a(x)R_{l,m}(x)=\underset{r=0}{\overset{\lambda 1}{}}x^rr!\underset{j=0}{\overset{r}{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{l}{j}\right)\left(\genfrac{}{}{0pt}{}{m}{j}\right)}{\left(\genfrac{}{}{0pt}{}{r}{j}\right)}+(1)^\lambda x^\lambda \lambda !\underset{r=0}{\overset{\mu \lambda 1}{}}x^rr!\left(\genfrac{}{}{0pt}{}{\mu r1}{\lambda }\right),$$
or, equivalently,
$$F_{l,m}^a(x)=\underset{r=0}{\overset{k1}{}}x^rr!\frac{(1)^\lambda x^\mu }{\lambda !L_\lambda ^{\mu \lambda }(x^1)}\underset{r=0}{\overset{\lambda 1}{}}(k+r)!x^r\underset{j=r+1}{\overset{\lambda }{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{l}{j}\right)\left(\genfrac{}{}{0pt}{}{m}{j}\right)}{\left(\genfrac{}{}{0pt}{}{k+r}{j}\right)},$$
where $`k=l+m=\lambda +\mu `$.
The proof of the Main Theorem is presented in the next section.
As a corollary we immediately get the following result (see \[Ma, Theorem 1\]).
###### Corollary 1.1
Let $`0ak1`$, then
$$f_{1,k1}^a(n)=\{\begin{array}{cccc}\hfill (& k2)!(k1)^{n+2k}\hfill & & \text{for }nk\hfill \\ & n!\hfill & & \text{for }n<k.\hfill \end{array}$$
###### Demonstration Proof
Since $`R_{1,k1}=1+(k1)x`$, the Main Theorem implies
$$F_{1,k1}^a(x)=\frac{1_{r=0}^{k3}x^r(kr2)r!}{1(k1)x}=\frac{x^{k2}(k2)!}{1(k1)x}+\underset{r=0}{\overset{k3}{}}x^rr!,$$
and the result follows. ∎
Another immediate corollary of the Main Theorem gives the asymptotics for $`f_{l,m}^a(n)`$ as $`n\mathrm{}`$.
###### Corollary 1.2
$`f_{l,m}^a(n)c\gamma ^n`$, where $`c`$ is a constant depending on $`l`$ and $`m`$, and $`\gamma `$ is the maximal root of $`L_\lambda ^{\mu \lambda }`$; in particular, $`\gamma k2+\sqrt{1+4(l1)(m1)}`$.
###### Demonstration Proof
Follows from standard results in the theory of rational generating functions (see e.g. \[St, vol.~1, Ch.~4\]) and the fact that all the roots of Laguerre polynomials are simple (see \[Sz, Ch.~3.3\]). The upper bound on $`\gamma `$ is obtained in \[IL\]. ∎
## 2. Proofs
First of all, we make the following simple, though useful observation.
###### Lemma 2.1
For any natural $`a`$, $`l`$, $`m`$, $`n`$ such that $`1al+m1`$ one has $`f_{l,m}^a(n)=f_{m,l}^{m+la}(n)`$.
###### Demonstration Proof
Denote by $`\rho _n`$ and $`\varkappa _n`$ the involutions $`S_nS_n`$ that take $`(i_1,i_2,\mathrm{},i_k)`$ to $`(i_k,\mathrm{},i_2,i_1)`$ (reversal) to $`(n+1i_1,n+1i_2,\mathrm{},n+1i_n)`$ (complement), respectively. It is easy to see that for any $`TS_k`$, the involutions $`\rho _n`$ and $`\varkappa _n`$ provide natural bijections between the sets $`S_n(T)`$ and $`S_n(\rho _kT)`$, and between $`S_n(T)`$ and $`S_n(\varkappa _kT)`$, respectively. It remains to note that $`\rho _k\varkappa _k\sigma ^aP_{l,m}=\sigma ^{l+ma}P_{l,m}`$. ∎
From now on we assume that $`a0`$, $`l1`$, $`m1`$ are fixed, and denote $`b=nm+a`$. It follows from Lemma 2.1 that we may assume that $`am`$, and hence $`bn`$. This means, in other words, that $`\tau S_k`$ belongs to $`\sigma ^aP_{l,m}`$ if and only if $`(\tau _1,\mathrm{},\tau _l)`$ is a permutation of the numbers $`a+1,\mathrm{},a+l`$. In what follows we usually omit the indices $`a`$, $`l`$, $`m`$ whenever appropriate; for example, instead of $`f_{l,m}^a(n)`$ we write just $`f(n)`$.
For any $`nk`$ and any $`d`$ such that $`1dn`$, we denote by $`g_n(i_1,\mathrm{},i_d)=g_{n;l,m}^a(i_1,\mathrm{},i_d)`$ the number of permutations $`\pi S_n(\sigma ^aP_{l,m})`$ such that $`\pi _j=i_j`$ for $`j=1,\mathrm{},d`$. It is natural to extend $`g_n`$ to the case $`d=0`$ by setting $`g_n(\mathrm{})=f(n)`$.
The following properties of the numbers $`g_n(i_1,\mathrm{},i_d)`$ can be deduced easily from the definitions.
###### Lemma 2.2
(i) Let $`nk`$ and $`1in`$, then
$$g_n(\mathrm{},i,\mathrm{},i,\mathrm{})=0.$$
(ii) Let $`nk`$ and $`a+1i_jb`$ for $`j=1,\mathrm{},l`$, then
$$g_n(i_1,\mathrm{},i_l)=0.$$
(iii) Let $`nk`$, $`1rdl`$, and $`a+1i_jb`$ for $`j=1,\mathrm{},d`$, $`jr`$, then
$$g_n(i_1,\mathrm{},i_d)=\{\begin{array}{cccc}& g_{n1}(i_11,\mathrm{},i_{r1}1,i_{r+1}1,\mathrm{},i_d1)\hfill & & \text{if }1i_ra\hfill \\ & g_{n1}(i_1,\mathrm{},i_{r1},i_{r+1},\mathrm{},i_d)\hfill & & \text{if }b+1i_rn.\hfill \end{array}$$
###### Demonstration Proof
Property (i) is evident. Let us prove (ii). By (i), we may assume that the numbers $`i_1,\mathrm{},i_l`$ are distinct. Take an arbitrary $`\pi S_n`$ such that $`\pi _j=i_j`$ for $`j=1,\mathrm{},l`$. Evidently, for any $`ra`$ there exists a position $`j_r>l`$ such that $`\pi _{j_r}=r`$; the same is true for any $`rb+1`$. Therefore, the restriction of $`\pi `$ to the positions $`1,2,\mathrm{},l,j_1,j_2,\mathrm{},j_a,j_{b+1},j_{b+2},\mathrm{},j_n`$ (in the proper order) gives an occurrence of $`\tau \sigma ^aP_{l,m}`$ in $`\pi `$. Hence, $`\pi S_n(\sigma ^aP_{l,m})`$, which means that $`g_n(i_1,\mathrm{},i_l)=0`$.
To prove (iii), assume first that $`1ra`$. Let $`\pi S_n`$ and $`\pi _j=i_j`$ for $`j=1,\mathrm{},d`$. We define $`\pi ^{}S_{n1}`$ by
$$\pi _j^{}=\{\begin{array}{cccc}& \pi _j1\hfill & & \text{for }1jr1,\hfill \\ & \pi _{j+1}1\hfill & & \text{for }jr\text{ and }\pi _{j+1}>i_r,\hfill \\ & \pi _{j+1}\hfill & & \text{for }jr\text{ and }\pi _{j+1}<i_r.\hfill \end{array}$$
$`1`$
We claim that $`\pi S_n(\sigma ^aP_{l,m})`$ if and only if $`\pi ^{}S_{n1}(\sigma ^aP_{l,m})`$. Indeed, the only if part is trivial, since any occurrence of $`\tau \sigma ^aP_{l,m}`$ in $`\pi ^{}`$ immediately gives rise to an occurrence of $`\tau `$ in $`\pi `$. Conversely, any occurrence of $`\tau `$ in $`\pi `$ that does not include $`i_r`$ gives rise to an occurrence of $`\tau `$ in $`\pi ^{}`$. Assume that there exists an occurrence of $`\tau `$ in $`\pi `$ that includes $`i_r`$. Since $`rdl`$, this occurrence of $`\tau `$ contains $`a`$ entries that are situated to the right of $`i_r`$ and are strictly less than $`i_r`$. However, the whole $`\pi `$ contains only $`a1`$ such entries, a contradiction. It now follows from (1) that property (iii) holds for $`1i_ra`$. The case $`b+1i_rn`$ is treated similarly with the help of transformation $`(\pi S_n)(\pi ^{}S_{n1})`$ given by
$$\pi _j^{}=\{\begin{array}{cccc}& \pi _j\hfill & & \text{for }1jr1,\hfill \\ & \pi _{j+1}1\hfill & & \text{for }jr\text{ and }\pi _{j+1}>i_r,\hfill \\ & \pi _{j+1}\hfill & & \text{for }jr\text{ and }\pi _{j+1}<i_r.\hfill \end{array}$$
Now we introduce the quantity that plays the crucial role in the proof of the Main Theorem. For $`nk`$ and $`1dl`$ we put
$$A(n,d)=A_{l,m}^a(n,d)=\underset{i_1,\mathrm{},i_d=a+1}{\overset{b}{}}g_n(i_1,\mathrm{},i_d).$$
As before, this definition is extended to the case $`d=0`$ by setting
$$A(n,0)=g_n(\mathrm{})=f(n).$$
###### Theorem 2.3
Let $`nk+1`$ and $`1dl1`$, then
$$A(n,d+1)=a(n,d)(md)A(n1,d)dA(n1,d1).$$
$`2`$
###### Demonstration Proof
First of all, we introduce two auxiliary sums:
$`B(n,d)=B_{l,m}^a(n,d)`$ $`={\displaystyle \underset{i_1,\mathrm{},i_d=a+1}{\overset{b+1}{}}}g_n(i_1,\mathrm{},i_d),`$
$`C(n,d)=C_{l,m}^a(n,d)`$ $`={\displaystyle \underset{i_1,\mathrm{},i_d=a}{\overset{b}{}}}g_n(i_1,\mathrm{},i_d);`$
once again, $`B(n,0)=C(n,0)=f(n)`$.
Let us prove three simple identities relating together the sequences $`\{A(n,d)\}`$, $`\{B(n,d)\}`$, $`\{C(n,d)\}`$.
###### Lemma 2.4
Let $`nk`$ and $`1dl`$, then:
$$\begin{array}{c}A(n,d)=A(n,d1)(ma)B(n1,d1)aC(n1,d1),\\ (ma)A(n,d)=(ma)B(n,d)(ma)dB(n1,d1),\\ aA(n,d)=aC(n,d)adC(n1,d1).\end{array}$$
###### Demonstration Proof
To prove the first identity, observe that by definitions and Lemma 2.2(iii) for the case $`r=d`$, one has
$$\begin{array}{c}A(n,d1)A(n,d)=\underset{i_1,\mathrm{},i_{d1}=a+1}{\overset{b}{}}\underset{i_d=1}{\overset{n}{}}g_n(i_1,\mathrm{},i_d)A(n,d)\hfill \\ \hfill =\underset{i_1,\mathrm{},i_{d1}=a+1}{\overset{b}{}}\left(\underset{i_d=1}{\overset{a}{}}g_n(i_1,\mathrm{},i_d)+\underset{i_d=b+1}{\overset{n}{}}g_n(i_1,\mathrm{},i_d)\right)\\ \hfill =\underset{i_1,\mathrm{},i_{d1}=a+1}{\overset{b}{}}\left(ag_{n1}(i_11,\mathrm{},i_{d1}1)+(ma)g_{n1}(i_1,\mathrm{},i_{d1})\right)\\ \hfill =aC(n1,d1)+(ma)B(n1,d1),\end{array}$$
and the result follows.
The second identity is trivial for $`a=m`$, so assume that $`0am1`$ and observe that by definitions and Lemma 2.2(ii) and (iii), one has
$$\begin{array}{c}B(n,d)=\underset{i_1,\mathrm{},i_d=a+1}{\overset{b}{}}g_n(i_1,\mathrm{},i_d)\hfill \\ \hfill +\underset{j=1}{\overset{d}{}}\underset{i_1,\mathrm{},\widehat{ı}_j,\mathrm{},i_d=a+1}{\overset{b}{}}g_n(i_1,\mathrm{},i_{j1},b+1,i_{j+1}\mathrm{},i_d)\\ \hfill =A(n,d)+\underset{j=1}{\overset{d}{}}\underset{i_1,\mathrm{},\widehat{ı}_j,\mathrm{},i_d=a+1}{\overset{b}{}}g_{n1}(i_1,\mathrm{},i_{j1},i_{j+1},\mathrm{},i_d)\\ \hfill =A(n,d)+dB(n1,d1),\end{array}$$
and the result follows.
Finally, the third identity is trivial for $`a=0`$, so assume that $`1am`$ and observe that by definitions and Lemma 2.2(ii) and (iii), one has
$$\begin{array}{c}C(n,d)=\underset{i_1,\mathrm{},i_d=a+1}{\overset{b}{}}g_n(i_1,\mathrm{},i_d)\hfill \\ \hfill +\underset{j=1}{\overset{d}{}}\underset{i_1,\mathrm{},\widehat{ı}_j,\mathrm{},i_d=a+1}{\overset{b}{}}g_n(i_1,\mathrm{},i_{j1},a,i_{j+1}\mathrm{},i_d)=A(n,d)\\ \hfill +\underset{j=1}{\overset{d}{}}\underset{i_1,\mathrm{},\widehat{ı}_j,\mathrm{},i_d=a+1}{\overset{b}{}}g_{n1}(i_11,\mathrm{},i_{j1}1,i_{j+1}1,\mathrm{},i_d1)\\ \hfill =A(n,d)+dC(n1,d1),\end{array}$$
and the result follows. ∎
Now we can complete the proof of Theorem 2.3. Indeed, using twice the first identity of Lemma 2.4, one gets
$`A(n,d+1)`$ $`=A(n,d)(ma)B(n,d)aC(n1,d1),`$
$`dA(n1,d)`$ $`=dA(n1,d1)d(ma)B(n2,d1)daC(n2,d1).`$
Next, the other two identities of Lemma 2.4 imply
$$\begin{array}{c}A(n,d+1)dA(n1,d)=A(n,d)dA(n1,d1)\hfill \\ \hfill \left((ma)B(n1,d)(ma)dB(n2,d1)\right)\left(aC(n1,d)adC(n2,d1)\right)\\ \hfill =A(n,d)dA(n1,d1)(ma)A(n1,d)aA(n1,d),\end{array}$$
and the result follows. ∎
The next result relates the sequence $`\{A(n,d)\}`$ to the sequence $`\{f(n)\}`$.
###### Theorem 2.5
Let $`nk`$ and $`1dl`$, then
$$A(n,d)=\underset{j=0}{\overset{d}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d}{j}\right)f(nj).$$
###### Demonstration Proof
Let $`D(n,d)=D_{l,m}^a(n,d)`$ denote the right hand side of the above identity. We claim that for $`nk+1`$ and $`1dl1`$, $`D(n,d)`$ satisfies the same relation (2) as $`A(n,d)`$ does. Indeed,
$$\begin{array}{c}D(n1,d)=\underset{j=0}{\overset{d}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d}{j}\right)f(n1j)\hfill \\ \hfill =\underset{j=1}{\overset{d+1}{}}(1)^j(j1)!\left(\genfrac{}{}{0pt}{}{m}{j1}\right)\left(\genfrac{}{}{0pt}{}{d}{j1}\right)f(nj)\\ \hfill +(md)(1)^{d+1}d!\left(\genfrac{}{}{0pt}{}{m}{d}\right)f(nd1),\end{array}$$
and
$$\begin{array}{c}D(n1,d1)=\underset{j=0}{\overset{d1}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d1}{j}\right)f(n1j)\hfill \\ \hfill =\underset{j=1}{\overset{d}{}}(1)^j(j1)!\left(\genfrac{}{}{0pt}{}{m}{j1}\right)\left(\genfrac{}{}{0pt}{}{d1}{j1}\right)f(nj),\end{array}$$
and hence
$$\begin{array}{c}D(n,d)(md)D(n1,d)dD(n1,d1)\hfill \\ \hfill =f(n)+(md)(1)^{d+1}d!\left(\genfrac{}{}{0pt}{}{m}{d}\right)f(nd1)\\ \hfill +\underset{j=1}{\overset{d}{}}(1)^jj!\left(\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d}{j}\right)+\frac{md}{j}\left(\genfrac{}{}{0pt}{}{m}{j1}\right)\left(\genfrac{}{}{0pt}{}{d}{j1}\right)+\frac{d}{j}\left(\genfrac{}{}{0pt}{}{m}{j1}\right)\left(\genfrac{}{}{0pt}{}{d1}{j1}\right)\right)f(nj)\\ \hfill =f(n)+\underset{j=1}{\overset{d}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d+1}{j}\right)f(nj)+(1)^{d+1}(d+1)!\left(\genfrac{}{}{0pt}{}{m}{d+1}\right)f(nd1)\\ \hfill =D(n,d+1).\end{array}$$
It follows that $`D(n,d)`$ (as well as $`A(n,d)`$) are defined uniquely for $`nk`$ and $`1ld`$ by initial values $`D(k,d)`$, $`D(n,0)`$, and $`D(n,1)`$ ($`A(k,d)`$, $`A(n,0)`$, and $`A(n,1)`$, respectively). It is easy to see that for $`nk`$ one has $`A(n,0)=D(n,0)=f(n)`$. Next, the first identity of Lemma 2.4 for $`d=1`$ gives
$$A(n,1)=A(n,0)(ma)B(n1,0)aC(n1,0)=f(n)mf(n1)\text{for }nk.$$
On the other hand, by definition,
$$D(n,1)=f(n)mf(n1)\text{for }nk,$$
and hence $`A(n,1)=D(n,1)`$. Finally, a simple combinatorial argument shows that
$$A(k,d)=d!\left(\genfrac{}{}{0pt}{}{l}{d}\right)(kd)!l!m!\text{for }1dl.$$
On the other hand,
$$D(k,d)=\underset{j=0}{\overset{d}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d}{j}\right)(kj)!l!m!,$$
since $`f(r)=r!`$ for $`1rk1`$ and $`f(k)=k!l!m!`$. To prove $`A(k,d)=D(k,d)`$ it remains to check that
$$\underset{j=0}{\overset{d}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{d}{j}\right)(kj)!=d!\left(\genfrac{}{}{0pt}{}{l}{d}\right)(kd)!,$$
which follows from Lemma 2.6 below. ∎
Finally, we are ready to prove the Main Theorem stated in Sec. 1. First of all, by Lemma 2.2(ii), $`A(n,l)=0`$ for $`nk`$. Hence, by Theorem 2.5,
$$\underset{j=0}{\overset{l}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{l}{j}\right)f(nj)=0\text{for }nk,$$
or, equivalently,
$$\underset{j=0}{\overset{l}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{l}{j}\right)x^jf(nj)x^{nj}=0\text{for }nk.$$
As it was already mentioned, $`f(r)=r!`$ for $`1rk1`$, therefore
$$\underset{j=0}{\overset{l}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{l}{j}\right)x^j\left(F_{l,m}^a(x)\underset{i=0}{\overset{kj1}{}}x^ii!\right)=0.$$
$`3`$
Recall that the rook polynomial of the rectangular $`s\times t`$ board, $`st`$, satisfies relation
$$R_{s,t}(x)=\underset{j=0}{\overset{s}{}}j!\left(\genfrac{}{}{0pt}{}{t}{j}\right)\left(\genfrac{}{}{0pt}{}{s}{j}\right)x^j$$
(see \[Ri, Ch.~7.4\]). Hence, (3) is equivalent to
$$\begin{array}{c}F_{l,m}^a(x)R_{\lambda ,\mu }(x)=\underset{j=0}{\overset{l}{}}(1)^jj!\left(\genfrac{}{}{0pt}{}{m}{j}\right)\left(\genfrac{}{}{0pt}{}{l}{j}\right)x^j\underset{i=0}{\overset{kj1}{}}x^ii!\\ =\underset{r=0}{\overset{k1}{}}x^rr!\underset{j=0}{\overset{r}{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{l}{j}\right)\left(\genfrac{}{}{0pt}{}{m}{j}\right)}{\left(\genfrac{}{}{0pt}{}{r}{j}\right)}\\ =\underset{r=0}{\overset{\lambda 1}{}}x^rr!\underset{j=0}{\overset{r}{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{l}{j}\right)\left(\genfrac{}{}{0pt}{}{m}{j}\right)}{\left(\genfrac{}{}{0pt}{}{r}{j}\right)}+\underset{r=\lambda }{\overset{\mu 1}{}}x^rr!\underset{j=0}{\overset{\lambda }{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{l}{j}\right)\left(\genfrac{}{}{0pt}{}{m}{j}\right)}{\left(\genfrac{}{}{0pt}{}{r}{j}\right)}+\underset{r=\mu }{\overset{k1}{}}x^rr!\underset{j=0}{\overset{\lambda }{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{l}{j}\right)\left(\genfrac{}{}{0pt}{}{m}{j}\right)}{\left(\genfrac{}{}{0pt}{}{r}{j}\right)}.\end{array}$$
By Lemma 2.6 below, the third term of the above expression vanishes, while the second term is equal to
$$\underset{r=\lambda }{\overset{\mu 1}{}}x^rr!(1)^\lambda \frac{(r\lambda )!(kr1)!}{\mu r1)!r!},$$
and the first expression of the Main Theorem follows. The second expression is obtained easily from (3) and relation between rook polynomials and Laguerre polynomials given in Sec. 1. ∎
It remains to prove the following technical result, which is apparently known; however, we failed to find a reference to its proof, and decided to present a short proof inspired by the brilliant book \[PWZ\].
###### Lemma~2.6
Let $`1st`$ and let
$$M(s,t)=\underset{i=0}{\overset{s}{}}(1)^j\frac{\left(\genfrac{}{}{0pt}{}{s}{i}\right)\left(\genfrac{}{}{0pt}{}{t}{i}\right)}{\left(\genfrac{}{}{0pt}{}{n}{i}\right)}.$$
Then:
$$M(s,t)=\{\begin{array}{ccc}& \frac{\left(\genfrac{}{}{0pt}{}{nt}{s}\right)}{\left(\genfrac{}{}{0pt}{}{n}{s}\right)}\hfill & \hfill \text{if }ns+t,\\ & 0\hfill & \hfill \text{if }tns+t1,\\ & (1)^s\frac{\left(\genfrac{}{}{0pt}{}{s+tn1}{s}\right)}{\left(\genfrac{}{}{0pt}{}{n}{s}\right)}\hfill & \hfill \text{if }snt1.\end{array}$$
###### Demonstration Proof
Direct check reveals that $`M(s,t)`$ is a hypergeometric series; to be more precise,
$$M(s,t)={}_{2}{}^{}F_{1}^{}\left[{}_{n}{}^{t,s};1\right].$$
Since $`s`$ is a nonpositive integer, the Gauss formula applies (see \[PWZ, Ch.~3.5\]), and we get
$$M(s,t)=\underset{zn}{lim}\frac{\mathrm{\Gamma }(z+t+s)\mathrm{\Gamma }(z)}{\mathrm{\Gamma }(tz)\mathrm{\Gamma }(sz)}.$$
Recall that
$$\mathrm{\Gamma }(x)\mathrm{\Gamma }(1x)=\frac{\pi }{\mathrm{sin}\pi x}.$$
$`4`$
If $`ns+t`$, we apply (4) for $`x=z+t+s`$, $`x=tz`$, $`x=sz`$, $`x=z`$, and get
$$M(s,t)=\frac{\mathrm{\Gamma }(nt+1)\mathrm{\Gamma }(ns+1)}{\mathrm{\Gamma }(nts+1)\mathrm{\Gamma }(n+1)}\underset{zn}{lim}\frac{\mathrm{sin}\pi (tz)\mathrm{sin}\pi (sz)}{\mathrm{sin}\pi z\mathrm{sin}\pi (t+sz)}=\frac{\left(\genfrac{}{}{0pt}{}{nt}{s}\right)}{\left(\genfrac{}{}{0pt}{}{n}{s}\right)}.$$
If $`tns+t1`$, we apply (4) for $`x=tz`$, $`x=sz`$, $`x=z`$, and get
$$M(s,t)=\frac{\mathrm{\Gamma }(nt+1)\mathrm{\Gamma }(ns+1)\mathrm{\Gamma }(s+tn)}{\mathrm{\Gamma }(n+1)}\underset{zn}{lim}\frac{\mathrm{sin}\pi (tz)\mathrm{sin}\pi (sz)}{\mathrm{sin}\pi z}=0.$$
Finally, if $`snt1`$, we apply (4) for $`x=sz`$, $`x=z`$, and get
$$M(s,t)=\frac{\mathrm{\Gamma }(s+tn)\mathrm{\Gamma }(ns+1)}{\mathrm{\Gamma }(tn)\mathrm{\Gamma }(n+1)}\underset{zn}{lim}\frac{\mathrm{sin}\pi (sz)}{\mathrm{sin}\pi z}=(1)^s\frac{\left(\genfrac{}{}{0pt}{}{s+tn1}{s}\right)}{\left(\genfrac{}{}{0pt}{}{n}{s}\right)}.$$
## 3. Concluding remarks
Observe first, that according to the Main Theorem, $`F_{l,m}^a(x)`$ does not depend on $`a`$; in other words, $`|S_n(P_{l,m})|=|S_n(\sigma ^aP_{l,m})|`$ for any $`a`$. We obtained this fact as a consequence of lengthy computations. A natural question would be to find a bijection between $`S_n(P_{l,m})`$ and $`S_n(\sigma ^aP_{l,m})`$ that explains this phenomenon.
Second, it is well known that rook polynomials (or the corresponding Laguerre polynomials) are related to permutations with restricted positions, see \[Ri, Ch.7, 8\]. Laguerre polynomials also arise in a natural way in the study of generalized derangements (see \[FZ\] and references therein). It is tempting to find a combinatorial relation between permutations with restricted positions and permutations avoiding maximal parabolic subgroups, which could explain the occurrence of Laguerre polynomials in the latter context.
Finally, one can consider permutations avoiding nonmaximal parabolic subgroups of $`S_k`$. The first natural step would be to treat the case of subgroups generated by $`k3`$ simple transpositions. It is convenient to denote by $`P_{l_1,l_2,l_3}`$ (with $`l_1+l_2+l_3=k`$) the subgroup of $`S_k`$ generated by all the simple transpositions except for $`s_{l_1}`$ and $`s_{l_1+l_2}`$; further on, we set $`f_{l_1,l_2,l_3}(n)=|S_n(P_{l_1,l_2,l_3})|`$, and $`F_{l_1,l_2,l_3}(x)=_{n0}f_{l_1,l_2,l_3}(n)x^n`$. It is easy to see that $`F_{l_1,l_2,l_3}(x)=F_{l_3,l_2,l_1}(x)`$, so one can assume that $`l_1l_3`$. This said, the main result of \[BDPP\] can be formulated as follows: let $`k3`$, then
$$F_{1,1,k2}(x)=\underset{r=1}{\overset{k2}{}}x^rr!+\frac{(k3)!x^{k4}}{2}\left(1(k1)x\sqrt{12(k1)x+(k3)^2x^2}\right).$$
To the best of our knowledge, this is the only known instance of $`F_{l_1,l_2,l_3}(x)`$. It is worth to note that even in this, simplest, case of nonmaximal parabolic subgroup, the generating function is no more rational.
|
warning/0006/hep-ph0006195.html
|
ar5iv
|
text
|
# RESULTS ON DIFFRACTION AT HERA AND TEVATRON
## 1 Introduction
The understanding of diffractive interactions is of fundamental importance as they govern the high energy behaviour of the elastic cross sections and thus of the total cross sections (via the optical theorem).
In the 70’s, the diffractive processes were intensively studied in hadron–hadron interactions and were well described by the Regge phenomenology. In this framework, the elastic scattering is attributed at high energy to the exchange between the incoming hadrons of a colourless object, the pomeron. The energy dependence of the elastic cross section is parameterised as $`ds/dts^{2(\alpha _{\mathrm{l}\mathrm{P}}(t)1)}`$, and depends on the trajectory of the pomeron $`\alpha _{\mathrm{l}\mathrm{P}}(t)`$: $`\alpha _{\mathrm{l}\mathrm{P}}(t)=1.08+0.25t`$, $`t`$ being the square of the four-momentum transfer. The total, elastic, and diffractive cross sections, thus exhibit a “soft” energy dependence, at high energy.
An important result of HERA studies is that, in contrast to the slow increase with energy of hadron-hadron cross sections (“soft” behaviour), the total $`\gamma ^{}p`$ cross section has a strong (“hard”) energy dependence in the deep inelastic scattering (DIS) domain, which is attributed to a fast rise with energy of the gluon density in the proton. The QCD pomeron being described as a gluonic system, a “hard” behaviour is thus also expected in diffractive interactions.
The interest is now to understand the diffractive interaction in the framework of the QCD theory, and in particular, to study the partonic structure of the pomeron.
## 2 Diffractive structure function at HERA
Diffractive interactions at HERA account for above 10 % of the deep inelastic scattering (DIS) events. The diffractive cross section is measured by selecting the events: $`e+pe+X+Y`$, where the two hadronic systems $`X`$ and $`Y`$ are separated by a large rapidity gap, devoid of particles, $`Y`$ being the system closest to the outgoing proton beam direction. The topology of diffractive events at HERA is presented in Fig. 1b, in contrast to the topology of events in the DIS regime (Fig. 1a). The measurement of the cross section is given in term of the diffractive structure function $`F_2^{D(3)}`$($`Q^2`$, $`x_{\mathrm{l}\mathrm{P}}`$, $`\beta `$). The variable $`Q^2`$ is the negative of the square of q, the four–momentum carried by the virtual photon, and
$$x_{\mathrm{l}\mathrm{P}}=q.(k)/(q.p)\frac{Q^2+M_X^2}{Q^2+W^2};\beta =Q^2/(2q.(k))\frac{Q^2}{Q^2+M_X^2},$$
where k ($`p`$) is the four–momentum carried by the pomeron (incident proton), $`M_X`$ is the invariant mass of the dissociative photon system, and $`W`$ is the energy in the $`\gamma ^{}p`$ center of mass system. For diffractive interactions, the kinematical variable $`t`$ = $`k^2`$ is expected to be small. $`F_2^{D(3)}`$($`Q^2`$, $`x_{\mathrm{l}\mathrm{P}}`$, $`\beta `$) is integrated over $`t`$.
In a picture of diffractive interactions where a pomeron is emitted from the proton, the pomeron having a partonic structure, $`x_{\mathrm{l}\mathrm{P}}`$ is the fraction of the proton momentum carried by the pomeron, and $`\beta `$ is the fraction of the pomeron momentum carried by the quark interacting with the virtual photon. The product $`x_{\mathrm{l}\mathrm{P}}\beta `$ = $`x`$, the usual Bjorken scaling variable.
It has been shown that the amplitudes for diffractive deep inelastic scattering factorise out into a part which depends on $`x_{\mathrm{l}\mathrm{P}}`$ (a ’pomeron flux factor’) and a structure function $`F_2^D(\beta ,Q^2)`$ corresponding to a universal partonic structure of diffraction: $`F_2^{D(3)}(x_{IP},\beta ,Q^2)f(x_{\mathrm{l}\mathrm{P}})F_2^D(\beta ,Q^2)`$.
In a Regge approach, the pomeron flux factor follows a power law: $`f(x_{\mathrm{l}\mathrm{P}})(1/x_{\mathrm{l}\mathrm{P}})^{2\alpha _{IP}1}`$. From the measurement of the $`x_{\mathrm{l}\mathrm{P}}`$ dependence of the diffractive structure function, the pomeron intercept $`\alpha _{IP}(0)`$ can be extracted for different $`Q^2`$ values . As is presented in Fig. 2, the pomeron intercept for $`Q^2`$ $`>`$ 0 has a higher value than 1.08, which is typical of hadron–hadron interactions. The transition from a soft to hard behaviour happens at low $`Q^2`$ values.
The structure function $`F_2(\beta ,Q^2)`$, multiplied by $`x_{\mathrm{l}\mathrm{P}}`$ for clarity, is presented in Fig. 3 for a fixed value of $`x_{\mathrm{l}\mathrm{P}}`$ =0.003, as a function of $`Q^2`$, for different $`\beta `$ bins . It differs markedly from the proton structure function: $`F_2^{D(2)}`$($`Q^2`$, $`\beta `$) is still large at large $`\beta `$, and the rise with $`Q^2`$ persists for values of $`\beta `$ much larger than $``$ 0.15. This is an indication for a dominant gluonic component in the pomeron. A QCD fit was performed on the data, using the Altarelli–Parisi (DGLAP) evolution equations applied to the parton distributions defined, as a function of beta, for a starting scale $`Q_0`$. Two different initial parton distributions were considered for $`Q_0^2`$ = 3.0 $`\mathrm{GeV}^2`$: only quarks, and both quarks and gluons. In the first case (see Fig. 3a), the DGLAP evolution fails to reproduce rising scaling violations at large $`\beta `$, whereas the mixed quark and gluon case (see Fig. 3b) can describe this rise. For the latter parameterisation, the parton distributions are dominated, throughout the $`Q^2`$ range, by gluons, which carry a large fraction of the pomeron momentum; this fraction decreases only slowly at large $`\beta `$ as $`Q^2`$ increases (see Fig. 4).
To complete the understanding of diffraction, studies of the diffractive final state: jet production, hadron transverse momentum distribution, hadron energy flow, charmed particle production… are performed at HERA. The pomeron structure functions extracted from QCD fits to inclusive diffractive DIS are convoluted with scattering amplitudes, to describe the specific final states. The analysis of these different final states is consistent with the picture of a leading gluonic component in the pomeron. This supports the idea of universality of parton distributions in the pomeron. As an example, the differential cross section measurement for diffractive dijet electroproduction (4 $`<`$ $`Q^2`$ $`<`$ 80 $`\mathrm{GeV}^2`$) with $`p_T^{jet}>4`$ GeV, is presented in Fig. 5b. This process is a direct probe of the gluon in the pomeron, via the boson gluon fusion mechanism (see Fig 5a). Figure 5b shows the distribution of the variable $`z_{IP}=(M_{JJ}^2+Q^2)/(M_X^2+Q^2)`$ (where $`M_{JJ}`$ is the invariant mass of the two jet system), which represents the fraction of the pomeron momentum carried by the partons entering the hard process. The bulk of events have $`z_{IP}`$ $`<`$ 1, i.e. $`M_{JJ}`$ $`<`$ $`M_X`$, indicating that pomeron remnants carry a significant fraction of the pomeron momentum. The predictions using the QCD fit to the diffractive structure function, are in good agreement, in shape and in normalisation, with the data.
## 3 Hard diffraction at Tevatron
Hard diffractive interactions are studied at the Tevatron by the CDF and D0 Collaboration, through three different topologies: single diffraction, double diffraction and double pomeron exchange, see Fig 6. The single diffraction samples are selected by requiring a hard signature ($`P_T`$ jets, $`W`$ boson, $`J/\psi `$ meson or a tagged $`b`$ particle), together with the detection of the diffractively scattered $`\overline{p}`$ in the proton spectrometer, or by the presence of a gap without activity in the tracker and calorimeter detectors. The production rate for single diffraction is at the 1 % level, compared to the corresponding non-diffractive process. Double diffraction process is studied through the production of two jets separated by a rapidity gap. The rate for this process was measured at centre of mass energies $`\sqrt{s}`$ = 630 GeV and $`\sqrt{s}`$ = 1800 GeV. The ratio $`R_{630/1800}`$ is 3.4 $`\pm `$ 1.2 (2.4 $`\pm `$ 0.9) for the D0 (CDF ) Collaboration. The decrease of the double diffraction process with increasing energy could be explained by the concept of survival probability: when energy increases, underlying interactions between the beam particle remnants are stronger and can destroy the rapidity gap. Finally, double pomeron exchange is studied by requiring dijet production in the central detector and a rapidity gap on both sides of the detector, or a rapidity gap on one side together with the detection of the diffractively scattered $`\overline{p}`$ in the proton spectrometer. The rate for this process is at the level of $`10^4`$ of the corresponding non-diffractive interactions.
The CDF Collaboration has determined the partonic content of the pomeron, taking advantage of the different sensibilities of the various processes (dijet, $`W`$ and $`b`$ production) to the quark and gluon densities . The production rates are compared to predictions where a hard partonic content of the pomeron was assumed (see Fig. 7a). The gluon density is measured to be 0.55 $`\pm `$ 0.15, in agreement with the ZEUS result, but the measured rate at Tevatron is significantly lower than expected. The $`\beta `$ dependence of the dijet production cross section, measured by the CDF Collaboration , is presented in Fig. 7b, compared to the expected production rate when using the H1 parton densities for the pomeron. The disagreement between the rate of the Tevatron measurement and the expectation from HERA indicates a breakdown of factorisation, which could be understood in terms of survival probability.
## 4 Exclusive vector particle production at HERA
Another way to study diffractive interaction is to analyse the exclusive production of vector meson: $`e+pe+p+V`$. For these clean reactions, quantitative predictions in perturbative QCD are indeed possible when a hard scale is present. This scale can be given by $`Q^2`$, $`|t|`$ or $`m_q`$, the quark mass. Most models rely on the fact that, at high energy, in the proton rest frame, the photon fluctuates into a $`q\overline{q}`$ pair a long time before the interaction, and recombines into a vector meson (VM) a long time after the interaction. The amplitude $``$ then factories into three terms: $`\psi _{\lambda _V}^VT_{\lambda _V\lambda _\gamma }\psi _{\lambda _\gamma }^\gamma `$ where $`T_{\lambda _V\lambda _\gamma }`$ represents the interaction helicity amplitudes ($`\lambda _\gamma `$ and $`\lambda _V`$ being the helicities of the photon and the VM respectively) and $`\psi `$ represents the wave functions. In most models, the photon and vector meson $`q\overline{q}p`$ interaction is described by 2 gluon exchange. The cross section is then proportional to the square of the gluon density in the proton: $`\sigma _{\gamma p}\alpha _s^2(Q^2)/Q^6\left|xg(x,Q^2)\right|^2`$. The main uncertainties of the models come from the choice of scale, of the gluon distribution parameterisation and of the VM wave function (Fermi motion), and from the neglect of non-diagonal gluon distributions and of higher order corrections.
Vector meson production has been intensively studied at HERA, both in photo- ($`Q^2`$ $``$ 0) and electroproduction, for $`\varrho ,\omega ,\varphi ,J/\psi ,\psi ^{}\mathrm{and}\mathrm{{\rm Y}}`$ mesons. At high energy, the $`\varrho `$, $`\omega `$ and $`\varphi `$ photoproduction cross sections measured by fixed target experiment and at HERA present a soft energy dependence, parameterised as $`\sigma W^\delta `$, with $`\delta =0.22`$. In contrast, the $`J/\psi `$ photoproduction cross section, where the mass of the c quark provides a hard scale in the interaction, presents a much stronger energy dependence (“hard” behaviour) , in agreement with the rise of the gluon density at low $`x`$ ($`x`$ $``$ $`Q^2`$/ $`W^2`$). Figure 8a presents the HERA measurement together with predictions of a perturbative QCD model using three parameterisations for the gluon density: GRVHO, MRSR2 and CTEQ4M. The full line corresponds to a fit to the data using the parameterisation $`\sigma W^\delta `$ with $`\delta =0.83\pm 0.07`$, which is in contrast with the value $`\delta `$ = 0.22 for light vector meson photoproduction.
Another way to look at the hard behaviour is to study light vector meson production at high $`Q^2`$, $`Q^2`$ giving here the scale. Measurements of the cross section $`\sigma (\gamma ^{}p\rho p)`$ show an indication for an increasingly stronger energy dependence when $`Q^2`$ increases . With the parameterisation $`\sigma W^\delta `$, the value of $`\delta `$ for $`\varrho `$ meson production appears to reach at high $`Q^2`$ that of $`J/\psi `$ photoproduction (see Fig. 8b).
Signals for $`\mathrm{{\rm Y}}`$ production have been observed recently at HERA (see Fig. 9a). The cross section $`\sigma (\gamma p\mathrm{{\rm Y}}p)\mathrm{BR}(\mathrm{{\rm Y}}\mu ^+\mu ^{}`$) for $`Q^2`$ $``$ 0 is measured to be, respectively, $`19.2\pm 11.0`$ and $`13.0\pm 6.6`$ pb by the H1 and ZEUS Collaborations; due to limited statistics the data cannot distinguish between 1S, 2S and 3S states of the $`\mathrm{{\rm Y}}`$ meson. In order to extract the cross section for the production of $`\mathrm{{\rm Y}}`$(1S), a production ratio of 70 % is used. The result is shown in Fig. 9b, together with recent pQCD calculations . These calculations describe well the data after consideration of two effects: the non vanishing of the real part of the scattering amplitude and the effect of the non-diagonal parton distributions in the proton, which leads to an enhancement of a factor $``$ 5 of the cross section normalisation. Such effects are found to be more important for the production of the $`\mathrm{{\rm Y}}`$ than for the $`J/\psi `$ meson, due to the larger $`b`$ quark mass.
The deep virtual compton scattering (DVCS): $`e+pe+p+\gamma `$ (see Fig. 10a) is a gold-plated process to study pQCD in diffraction. The reaction is perturbatively calculable at high $`Q^2`$, as the incoming and the outcoming photon wave functions and the couplings are known, and no complication from strong interactions between particle in the final state appears. It is then an ideal place to study non-diagonal parton distributions, thus correlations between gluons in the proton. The main background to the DVCS process is the Bethe-Heitler (QED Compton) process, which has the same final state but different phase space. To extract the DVCS cross section, the interference between the two processes has to be taken into account. Events have been selected by the ZEUS Collaboration for $`Q^2`$ $`>`$ 6 $`\mathrm{GeV}^2`$, by requesting the presence of two electromagnetic clusters, corresponding respectively two the scattered electron and the photon. Fig. 10b presents the distribution of the polar angle of the photon cluster. A clear excess in the data (full points) is observed above the Bethe-Heitler background (open triangles). The signal is consistent, in shape and in normalisation, with the prediction of a perturbative QCD calculation computing the DVCS and the Bethe-Heitler processes, taken into account the interference term (open circles).
|
warning/0006/hep-ph0006336.html
|
ar5iv
|
text
|
# Time Evolution of Decay Spectrum in {𝐾⁰,(𝐾⁰)̄}→𝜋⁺𝜋⁻𝑒⁺𝑒⁻
## 1 Introduction
The KTeV experiment has measured a large $`CP`$-violating, $`T`$-odd asymmetry in the decay $`K_L\pi ^+\pi ^{}e^+e^{}`$, in quantitative agreement with a prediction made some years ago . The origin of this effect lies in the amplitude of the radiative decay $`K_L\pi ^+\pi ^{}\gamma `$, which contains a bremsstrahlung term proportional to $`\eta _+`$, as well as a $`CP`$-conserving direct emission term of magnetic dipole character . The interference of the odd electric multipoles $`E1,E3,E5\mathrm{}`$ present in the bremsstrahlung amplitude, which all have $`CP=+1`$,with the magnetic $`M1`$ multipole of $`CP=1`$, produces $`CP`$-violating components in the polarization state (Stokes vector) of the photon . The Dalitz pair process $`K_L\pi ^+\pi ^{}e^+e^{}`$ acts as an analyser of the photon polarization, exposing the $`CP`$-odd, $`T`$-odd component of the Stokes vector. The specific distribution that reveals the $`CP`$-violation is
$$\frac{d\mathrm{\Gamma }}{d\varphi }1\left(\mathrm{\Sigma }_3\mathrm{cos}2\varphi +\mathrm{\Sigma }_1\mathrm{sin}2\varphi \right)$$
(1)
where $`\varphi `$ is the angle between the $`\pi ^+\pi ^{}`$ and $`e^+e^{}`$ planes. The last term in Eq. (1) is $`CP`$-odd and $`T`$-odd, and produces an asymmetry
$$𝒜_\varphi =\frac{\left(_0^{\pi /2}_{\pi /2}^\pi +_\pi ^{3\pi /2}_{3\pi /2}^{2\pi }\right)\frac{d\mathrm{\Gamma }}{d\varphi }d\varphi }{\left(_0^{\pi /2}+_{\pi /2}^\pi +_\pi ^{3\pi /2}+_{3\pi /2}^{2\pi }\right)\frac{d\mathrm{\Gamma }}{d\varphi }d\varphi }=\frac{2}{\pi }\mathrm{\Sigma }_1.$$
(2)
The measured value $`\left|𝒜_\varphi \right|=(13.6\pm 2.5\pm 1.2)\%`$ is in excellent agreement with the prediction of $`14\%`$.
In a recent report , the NA48 collaboration, while confirming the large $`CP`$-violating effect in $`K_L\pi ^+\pi ^{}e^+e^{}`$, has also studied the decay $`K_S\pi ^+\pi ^{}e^+e^{}`$, finding no asymmetry in this case. This is entirely consistent with the fact that the amplitude $`K_S\pi ^+\pi ^{}\gamma `$ is accurately reproduced by bremsstrahlung alone, so that no electric-magnetic interference is expected.
An interesting question raised by the above observations is the following: How does the asymmetry $`𝒜_\varphi `$ evolve with time if the neutral $`K`$ meson is prepared in an initial state $`K^0`$ or $`\overline{K^0}`$ of definite strangeness? How does the value of $`𝒜_\varphi `$ evolve from zero at short times (when the state decays like $`K_S`$) to the value $`14\%`$ at large times (when the state is essentially $`K_L`$)? Finally, what type of evolution is expected in the decay of an incoherent $`K^0\overline{K^0}`$ mixture, such as an untagged neutral kaon beam originating in the decay $`\varphi K^0\overline{K^0}`$?
This paper answers these questions by analysing in detail the time dependence of the decays $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$. We study the evolution of the full decay spectrum, especially the terms that are odd under $`CP`$. A comparison is made between the behaviour of beams that are initially $`K^0`$, $`\overline{K^0}`$ or an incoherent (untagged) mixture.
As a prelude to the discussion of the channel $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$, we analyse in Section 2 the time-dependence of the photon polarization in $`K^0,\overline{K^0}\pi ^+\pi ^{}\gamma `$. This will reveal the behaviour of the $`CP`$-odd components of the photon Stokes vector, one of which is reflected in the asymmetry $`𝒜_\varphi `$ in $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$, that we discuss in Section 3.
## 2 Time Evolution of Photon Polarization in $`K^0,\overline{K^0}\pi ^+\pi ^{}\gamma `$
The measured branching ratios and photon energy spectra in the decays $`K_L\pi ^+\pi ^{}\gamma `$ and $`K_S\pi ^+\pi ^{}\gamma `$ can be well-described by the matrix elements
$`(K_S\pi ^+\pi ^{}\gamma )`$ $`=`$ $`ef_S\left[{\displaystyle \frac{ϵp_+}{kp_+}}{\displaystyle \frac{ϵp_{}}{kp_{}}}\right]`$
$`(K_L\pi ^+\pi ^{}\gamma )`$ $`=`$ $`ef_L\left[{\displaystyle \frac{ϵp_+}{kp_+}}{\displaystyle \frac{ϵp_{}}{kp_{}}}\right]+e{\displaystyle \frac{f_{DE}}{M_{K}^{}{}_{}{}^{4}}}ϵ_{\mu \nu \rho \sigma }ϵ^\mu k^\nu p_{+}^{}{}_{}{}^{\rho }p_{}^{}{}_{}{}^{\sigma }`$ (3)
where
$`f_S=|f_S|e^{i\delta _0(s=M_{K}^{}{}_{}{}^{2})},`$ $`f_L=\eta _+f_S,`$
$`f_{DE}=|f_S|g_{M1},`$ $`g_{M1}=i(0.76)e^{i\delta _1(s)}.`$
Here $`\delta _{0,1}`$ are the $`\pi ^+\pi ^{}`$ phase shifts in the $`I=0`$ $`s`$-wave and $`I=1`$ $`p`$-wave channel, respectively. Introducing the notation
$`(K_{S,L}\pi ^+\pi ^{}\gamma )`$ $`=`$ $`{\displaystyle \frac{e|f_S|}{M_{K}^{}{}_{}{}^{4}}}\{E_{S,L}(\omega ,\mathrm{cos}\theta )[ϵp_+kp_{}ϵp_{}kp_+]`$ (4)
$`+M_{S,L}(\omega ,\mathrm{cos}\theta )ϵ_{\mu \nu \rho \sigma }ϵ^\mu k^\nu p_{+}^{}{}_{}{}^{\rho }p_{}^{}{}_{}{}^{\sigma }\}`$
we have for the electric and magnetic amplitudes
$`E_S=\left({\displaystyle \frac{2M_K}{\omega }}\right)^2{\displaystyle \frac{e^{i\delta _0(s=M_{K}^{}{}_{}{}^{2})}}{1\beta ^2\mathrm{cos}^2\theta }},`$ $`M_S`$ $`=0`$
$`E_L=\left({\displaystyle \frac{2M_K}{\omega }}\right)^2{\displaystyle \frac{\eta _+e^{i\delta _0(s=M_{K}^{}{}_{}{}^{2})}}{1\beta ^2\mathrm{cos}^2\theta }},`$ $`M_L`$ $`=i(0.76)e^{i\delta _1(s)}.`$ (5)
As in , $`\omega `$ is the photon energy in the kaon rest frame, and $`\theta `$ is the angle of the $`\pi ^+`$ relative to the photon in the $`\pi ^+\pi ^{}`$ c.m. frame. Noting that the strangeness eigenstates $`K^0`$ and $`\overline{K^0}`$ can be written as
$`K^0`$ $`=`$ $`\left(K_S+K_L\right)/𝒩`$
$`\overline{K^0}`$ $`=`$ $`\left(K_SK_L\right)/\overline{𝒩},`$ (6)
the decay amplitudes for $`K^0`$ and $`\overline{K^0}`$ at time $`t`$ are
$`(K^0(t)\pi ^+\pi ^{}\gamma )`$ $``$ $`\{E(t,\omega ,\mathrm{cos}\theta )[ϵp_+kp_{}ϵp_{}kp_+]`$
$`+M(t,\omega ,\mathrm{cos}\theta )ϵ_{\mu \nu \rho \sigma }ϵ^\mu k^\nu p_{+}^{}{}_{}{}^{\rho }p_{}^{}{}_{}{}^{\sigma }\}`$
$`(\overline{K^0}(t)\pi ^+\pi ^{}\gamma )`$ $``$ $`\{\overline{E}(t,\omega ,\mathrm{cos}\theta )[ϵp_+kp_{}ϵp_{}kp_+]`$ (7)
$`+\overline{M}(t,\omega ,\mathrm{cos}\theta )ϵ_{\mu \nu \rho \sigma }ϵ^\mu k^\nu p_{+}^{}{}_{}{}^{\rho }p_{}^{}{}_{}{}^{\sigma }\}`$
where
$`E`$ $`=`$ $`e^{i\lambda _St}E_S(\omega ,\mathrm{cos}\theta )+e^{i\lambda _Lt}E_L(\omega ,\mathrm{cos}\theta )`$
$`M`$ $`=`$ $`e^{i\lambda _Lt}M_L(\omega ,\mathrm{cos}\theta )`$
$`\overline{E}`$ $`=`$ $`e^{i\lambda _St}E_S(\omega ,\mathrm{cos}\theta )e^{i\lambda _Lt}E_L(\omega ,\mathrm{cos}\theta )`$
$`\overline{M}`$ $`=`$ $`e^{i\lambda _Lt}M_L(\omega ,\mathrm{cos}\theta )`$ (8)
with $`\lambda _{S,L}=m_{S,L}\frac{i}{2}\mathrm{\Gamma }_{S,L}`$, $`t`$ being the proper time.
The amplitudes (8) determine the Stokes vector of the photon in $`K^0,\overline{K^0}\pi ^+\pi ^{}\gamma `$. For $`K^0`$ decay, the components of the Stokes vector $`\stackrel{}{S}=(S_1,S_2,S_3)`$ at a time $`t`$ are
$`S_1(t)`$ $`=`$ $`{\displaystyle \frac{2\mathrm{Re}\left[E^{}(t)M(t)\right]}{|E(t)|^2+|M(t)|^2}}`$
$`S_2(t)`$ $`=`$ $`{\displaystyle \frac{2\mathrm{Im}\left[E^{}(t)M(t)\right]}{|E(t)|^2+|M(t)|^2}}`$ (9)
$`S_3(t)`$ $`=`$ $`{\displaystyle \frac{|E(t)|^2|M(t)|^2}{|E(t)|^2+|M(t)|^2}}.`$
The corresponding components for $`\overline{K^0}`$ decay are obtained by replacing $`E(t)\overline{E}(t)`$, $`M(t)\overline{M}(t)`$.
The physical significance of $`S_2(t)`$ is that it represents the net circular polarization of the photon in $`K^0\pi ^+\pi ^{}\gamma `$ at time $`t`$:
$$S_2(t)=\frac{d\mathrm{\Gamma }(\lambda _\gamma =1,t)d\mathrm{\Gamma }(\lambda _\gamma =+1,t)}{d\mathrm{\Gamma }(\lambda _\gamma =1,t)+d\mathrm{\Gamma }(\lambda _\gamma =+1,t)}.$$
(10)
The parameters $`S_1(t)`$ and $`S_3(t)`$, on the other hand describe the dependence of the decay rate on the orientation of the polarization vector $`\stackrel{}{ϵ}`$ relative to $`\stackrel{}{n}_\pi `$, the unit vector normal to the $`\pi ^+\pi ^{}`$ plane:
$$\frac{d\mathrm{\Gamma }(t)}{d\varphi }1\left(S_3(t)\mathrm{cos}2\varphi +S_1(t)\mathrm{sin}2\varphi \right)$$
(11)
where the coordinates are chosen such that
$$\stackrel{}{k}=(0,0,k),\stackrel{}{ϵ}=(\mathrm{cos}\varphi ,\mathrm{sin}\varphi ,0),\stackrel{}{n}_\pi =\frac{\stackrel{}{p}_+\times \stackrel{}{p}_{}}{|\stackrel{}{p}_+\times \stackrel{}{p}_{}|}=(1,0,0).$$
(12)
In Figs. 1a and 2a, we show the components $`S_1(t)`$ and $`S_2(t)`$ as functions of photon energy for the decay of an initial $`K^0`$. The corresponding Stokes vector components for an initial $`\overline{K^0}`$ are shown in Figs. 1b and 2b. Notice the intricate interference effect in the time-dependence, particularly in the region $`t10\tau _S`$. Of special interest is the limiting case $`t\mathrm{}`$, when the beam is essentially pure $`K_L`$. The Stokes vector components $`S_1`$ and $`S_2`$ then reduce to those shown in for the case $`K_L\pi ^+\pi ^{}\gamma `$. For comparison with the $`K^0`$ and $`\overline{K^0}`$ cases, we have also considered the case of an untagged initial beam, consisting of an incoherent equal mixture of $`K^0`$ and $`\overline{K^0}`$. The Stokes vector is then
$`S_1(t)`$ $`=`$ $`{\displaystyle \frac{2\mathrm{Re}\left[E^{}(t)M(t)+\overline{E}^{}(t)\overline{M}(t)\right]}{|E(t)|^2+|M(t)|^2+|\overline{E}(t)|^2+|\overline{M}(t)|^2}}`$
$`S_2(t)`$ $`=`$ $`{\displaystyle \frac{2\mathrm{Im}\left[E^{}(t)M(t)+\overline{E}^{}(t)\overline{M}(t)\right]}{|E(t)|^2+|M(t)|^2+|\overline{E}(t)|^2+|\overline{M}(t)|^2}}`$ (13)
$`S_3(t)`$ $`=`$ $`{\displaystyle \frac{|E(t)|^2|M(t)|^2+|\overline{E}(t)|^2|\overline{M}(t)|^2}{|E(t)|^2+|M(t)|^2+|\overline{E}(t)|^2+|\overline{M}(t)|^2}}.`$
The time-dependent parameters $`S_1(t)`$ and $`S_2(t)`$ are plotted in Figs. 1c and 2c. Notice that the strong fluctuations in $`S_1(t)`$ and $`S_2(t)`$ for the $`K^0`$ and $`\overline{K^0}`$ cases are now smoothed out. For $`t\tau _S`$, of course, all three cases give the same Stokes vector, namely that corresponding to $`K_L\pi ^+\pi ^{}\gamma `$.
The polarization components $`S_1(t)`$ and $`S_2(t)`$ in the incoherent case have a special significance: they represent $`CP`$-violating observables at any time $`t`$. In the model under discussion, they vanish when $`\eta _+=0`$, which is not the case for the components $`S_1(t)`$ and $`S_2(t)`$ originating from $`K^0`$ or $`\overline{K^0}`$. Furthermore, in the “hermitian limit” (i.e. $`\delta _0=\delta _1=0`$, $`\mathrm{arg}\eta _+=\frac{\pi }{2}`$) the parameter $`S_1(t)`$ survives, but $`S_2(t)`$ vanishes. This is the hallmark that characterises the observable $`S_1(t)`$ as being $`CP`$-odd, $`T`$-odd, and the observable $`S_2(t)`$ as being $`CP`$-odd, $`T`$-even .
Of particular relevance in our discussion of $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$ is the behaviour of $`S_1(t,\omega )`$ portrayed in Fig. 1 as a function of photon energy $`\omega `$. This will be found to have a resemblance with the asymmetry $`𝒜_\varphi (t,s_\pi )`$ as a function of the $`\pi ^+\pi ^{}`$ invariant mass (recall that for $`K\pi ^+\pi ^{}\gamma `$, $`s_\pi =M_K^22M_K\omega `$). We now proceed to a systematic analysis of time dependence in the Dalitz pair reaction.
## 3 Time Evolution of $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$
We begin by recalling the matrix element of the long-lived kaon decay $`K_L\pi ^+\pi ^{}e^+e^{}`$, treating the $`e^+e^{}`$ system as an internal conversion pair associated with the radiative decay $`K_L\pi ^+\pi ^{}\gamma `$ . Writing the matrix element in the form
$`(K_L\pi ^+\pi ^{}e^+e^{})`$ $`=`$ $`2{\displaystyle \frac{G_F}{\sqrt{2}}}\mathrm{sin}\theta _C\{{\displaystyle \frac{1}{M_K}}[f(p_++p_{})_\lambda +g(p_+p_{})_\lambda ]`$ (14)
$`+i{\displaystyle \frac{h}{M_K^2}}ϵ_{\lambda \mu \nu \sigma }p_K^\mu (p_++p_{})^\nu (p_+p_{})^\sigma \}`$
$`\times \overline{u}(k_{})\gamma ^\lambda v(k_+),`$
the form factors $`f`$, $`g`$ and $`h`$ are related to the parameters $`f_S`$, $`\eta _+`$ and $`g_{M1}`$ of the radiative decay as follows :
$`f`$ $`=`$ $`CM_K^4|\eta _+|e^{i(\delta (M_K^2)+\phi _+)}{\displaystyle \frac{1}{s_l}}{\displaystyle \frac{4\beta \mathrm{cos}\theta _\pi }{s(1\beta ^2\mathrm{cos}^2\theta _\pi )}}`$
$`g`$ $`=`$ $`CM_K^4|\eta _+|e^{i(\delta (M_K^2)+\phi _+)}{\displaystyle \frac{1}{s_l}}{\displaystyle \frac{4}{s(1\beta ^2\mathrm{cos}^2\theta _\pi )}}`$ (15)
$`h`$ $`=`$ $`CM_K^4{\displaystyle \frac{1}{s_l}}(0.76)e^{i\delta _1(s_\pi )}`$
Here $`\theta _\pi `$ is the angle of the $`\pi ^+`$ in the $`\pi ^+\pi ^{}`$ rest frame, relative to the sum of the $`e^+`$ and $`e^{}`$ momenta in that frame; $`s_l`$ is the invariant mass of the lepton pair; $`s_\pi `$ is the $`\pi ^+\pi ^{}`$ invariant mass; and $`s`$ is defined as $`s=\frac{1}{2}(M_K^2s_\pi s_l)`$. The constant $`C`$ is given by
$$CM_K^4=\left(\frac{G_F}{\sqrt{2}}\mathrm{sin}\theta _C\frac{1}{M_K}\right)^1\pi \alpha |f_S|$$
(16)
It is now easy to obtain from the amplitude $`(K_L\pi ^+\pi ^{}e^+e^{})`$ the corresponding time-dependent amplitudes for $`K^0`$ and $`\overline{K^0}`$ decay by making use of the following artifice: To obtain $`(K^0(t)\pi ^+\pi ^{}e^+e^{})`$, we make the replacement
$`f`$ $``$ $`f\left(e^{i\lambda _Lt}+{\displaystyle \frac{1}{\eta _+}}e^{i\lambda _St}\right)/𝒩`$
$`g`$ $``$ $`g\left(e^{i\lambda _Lt}+{\displaystyle \frac{1}{\eta _+}}e^{i\lambda _St}\right)/𝒩`$ (17)
$`h`$ $``$ $`he^{i\lambda _Lt}/𝒩`$
in Eq. ( 14). Similarly, to obtain $`(\overline{K^0}(t)\pi ^+\pi ^{}e^+e^{})`$, we make the replacement
$`f`$ $``$ $`f\left(e^{i\lambda _Lt}+{\displaystyle \frac{1}{\eta _+}}e^{i\lambda _St}\right)/\overline{𝒩}`$
$`g`$ $``$ $`g\left(e^{i\lambda _Lt}+{\displaystyle \frac{1}{\eta _+}}e^{i\lambda _St}\right)/\overline{𝒩}`$ (18)
$`h`$ $``$ $`he^{i\lambda _Lt}/\overline{𝒩}.`$
The normalization factors $`𝒩`$ and $`\overline{𝒩}`$ in Eqs. (17) and (18) are those in the states $`K^0`$ and $`\overline{K^0}`$ (Eq. (6)), where $`𝒩=2p`$, $`\overline{𝒩}=2q`$, with $`|p|^2|q|^2=\frac{2\mathrm{Re}\eta _+}{1+|\eta _+|^2}`$ and $`|p|^2+|q|^2=1`$. With this prescription in mind, we will continue to use the formalism developed in for $`K_L\pi ^+\pi ^{}e^+e^{}`$, with the understanding that in dealing with $`K^0`$ and $`\overline{K^0}`$ decay, the form factors $`f`$, $`g`$ and $`h`$ are to be replaced by the time-dependent combinations given in Eqs. (17) and (18).
Following the procedure of Ref. , we have calculated the angular distribution of the decays $`K^0(t)\pi ^+\pi ^{}e^+e^{}`$ and $`\overline{K^0}(t)\pi ^+\pi ^{}e^+e^{}`$ in the form
$$\frac{d\mathrm{\Gamma }}{d\mathrm{cos}\theta _ld\varphi }=K_1+K_2\mathrm{cos}2\theta _l+K_3\mathrm{sin}^2\theta _l\mathrm{cos}2\varphi +K_9\mathrm{sin}^2\theta _l\mathrm{sin}2\varphi .$$
(19)
Here $`\theta _l`$ is the angle of $`e^+`$ relative to the dipion momentum vector in the $`e^+e^{}`$ frame; and $`\varphi `$ is the angle between the $`\pi ^+\pi ^{}`$ and $`e^+e^{}`$ planes. The last term in Eq. (19) is odd under the $`CP`$-transformation $`\stackrel{}{p}_\pm \stackrel{}{p}_{}`$, $`\stackrel{}{k}_\pm \stackrel{}{k}_{}`$, as well as under the $`T`$-transformation $`\stackrel{}{p}_\pm \stackrel{}{p}_\pm `$, $`\stackrel{}{k}_\pm \stackrel{}{k}_\pm `$. The time-dependent coefficients $`K_2/K_1`$, $`K_3/K_1`$ and $`K_9/K_1`$ are shown in Fig. 3, where we compare the cases $`K^0`$ and $`\overline{K^0}`$, and also show the result for an incoherent $`K^0\overline{K^0}`$ mixture. This figure depicts the manner in which the coefficients of the angular distribution evolve to their asymptotic values appropriate to the decay $`K_L\pi ^+\pi ^{}e^+e^{}`$, namely $`K_2/K_1=0.297`$, $`K_3/K_1=0.180`$, $`K_9/K_1=0.309`$ .
Integrating Eq. (19) over $`\mathrm{cos}\theta _l`$, we obtain the $`\varphi `$-distribution
$$\frac{d\mathrm{\Gamma }}{d\varphi }\left(1\frac{1}{3}\frac{K_2}{K_1}\right)+\frac{2}{3}\left(\frac{K_3}{K_1}\mathrm{cos}2\varphi +\frac{K_9}{K_1}\mathrm{sin}2\varphi \right)$$
(20)
which corresponds to an asymmetry
$$𝒜_\varphi =\frac{2}{\pi }\frac{\frac{2}{3}\frac{K_9}{K_1}}{1\frac{1}{3}\frac{K_2}{K_1}}$$
(21)
The time-dependence of this asymmetry is exhibited in Fig. 4, which is the answer to the question of how this asymmetry evolves from the value zero, appropriate to $`K_S`$ decay, to the value $`14\%`$ observed for $`K_L`$ decay. We have also studied the time-dependent asymmetry $`𝒜_\varphi `$ as a function of the $`\pi ^+\pi ^{}`$ invariant mass. As seen in Fig. 5, the function $`𝒜_\varphi (t,s_\pi )`$ has a similarity to the Stokes vector $`S_1(t,\omega )`$ plotted in Fig. 1, confirming the expectation that the $`CP`$-odd, $`T`$-odd term in the $`\varphi `$-distribution of $`K^0\pi ^+\pi ^{}e^+e^{}`$ is correlated with the $`CP`$-odd, $`T`$-odd component of the Stokes vector in $`K^0\pi ^+\pi ^{}\gamma `$.
Finally, the spectrum-integrated decay rate in $`K^0\pi ^+\pi ^{}e^+e^{}`$ and $`\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$ contains the standard $`K_LK_S`$ interference effect proportional to $`\eta _+`$, and a corresponding asymmetry between $`K^0`$ and $`\overline{K^0}`$, which is shown in Fig. 6.
## 4 Additional Remarks
(i) Our analysis of the time-dependence in $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$ assumed the amplitude to be entirely determined by the radiative decay $`K^0,\overline{K^0}\pi ^+\pi ^{}\gamma `$. A non-radiative contribution to the amplitude $`K_L\pi ^+\pi ^{}e^+e^{}`$ is possible in the form of a “charge-radius” term, with the characteristic feature of producing $`\pi ^+\pi ^{}`$ in an $`s`$-wave. Such a configuration is not possible in the radiative decay $`K_L\pi ^+\pi ^{}\gamma `$. We have investigated the effects of a charge radius term nominally parametrized by the coefficient $`g_P=0.15e^{i\delta _1}`$ defined in . It has the interesting consequence of inducing a small term of the form $`K_4\mathrm{sin}2\theta _l\mathrm{cos}\varphi `$ in the angular distribution $`d\mathrm{\Gamma }/d\mathrm{cos}\theta _ld\varphi `$ given in Eq. (19). Such a term is $`CP`$-odd but $`T`$-even . In Fig. 7, we show the magnitude and time-dependence of the coefficient $`K_4/K_1`$ generated by a charge-radius term $`g_P`$ with the nominal value given above.
(ii) Since the time-dependent interference effects in the decay spectrum of $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$ are strongest in the region $`t10\tau _s`$, one can ask whether similar effects could be observed with a beam of the form $`K_L+\rho K_S`$, where $`\rho `$ is a regeneration amplitude induced by passage of $`K_L`$ through material. We have calculated the decay spectrum of $`(K_L+\rho K_S)\pi ^+\pi ^{}e^+e^{}`$ as a function of time, for some typical regeneration parameters $`\rho =|\rho |e^{i\phi _\rho }`$, such as $`|\rho |=0.002`$, $`0.02`$ and $`0.2`$ with $`\phi _\rho =\pi /4`$. In particular, the time-dependent asymmetry $`𝒜_\varphi `$ for such a beam is shown in Fig. 8. It is clear that a suitable choice of a regenerator would permit a study of the $`CP`$-odd, $`T`$-odd feature in the $`\pi ^+\pi ^{}e^+e^{}`$ decay of a neutral K meson, as well as other aspects of the decay spectrum.
(iii) Several refinements in the matrix element of $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$ can be imagined. There is a small short-distance contribution associated with the interaction $`s\overline{d}e^+e^{}`$ , which, however, is highly suppressed on account of the small quark mixing factor $`V_{ts}V_{td}^{}`$. The magnetic dipole term in the radiative amplitude $`K_L\pi ^+\pi ^{}\gamma `$, represented by the coupling constant $`g_{M1}`$, actually has a certain dependence on $`s_\pi `$ , which can be incorporated into the analysis. Finally, the radiative amplitudes $`K_{L,S}\pi ^+\pi ^{}\gamma `$ could contain additional components such as a direct $`E1`$ multipole in $`K_S`$ decay or a direct $`E2`$ term in $`K_L`$ decay (see, e.g. ). The former would give rise to a departure from pure bremsstrahlung in the photon energy spectrum of $`K_S\pi ^+\pi ^{}\gamma `$, while the latter would produce a $`CP`$-violating charge asymmetry in the Dalitz plot of $`K_L\pi ^+\pi ^{}\gamma `$. There is also the possibility of a direct $`CP`$-violating $`E1`$ multipole in $`K_L`$ decay, that would show up as a difference between $`\eta _{+\gamma }`$ and $`\eta _+`$ in the time-dependence of $`K_L+\rho K_S\pi ^+\pi ^{}\gamma `$.
A specific consequence of introducing direct $`E1`$ multipoles in $`K_L\pi ^+\pi ^{}\gamma `$ is to modify the amplitude $`E_L`$ in Eq.(5) to
$$E_L=\left(\frac{2M_K}{\omega }\right)^2\frac{\eta _+e^{i\delta _0}}{1\beta ^2\mathrm{cos}^2\theta }+g_{E1}e^{i(\phi _++\delta _1)}+i\stackrel{~}{g}_{E1}e^{i\delta _1}$$
(22)
where $`g_{E1}`$ is a measure of direct ($`CP`$-conserving) $`E1`$ emission in $`K_S\pi ^+\pi ^{}\gamma `$, and $`\stackrel{~}{g}_{E1}`$ a measure of a direct $`CP`$-violating $`E1`$ emission in $`K_L\pi ^+\pi ^{}\gamma `$ ($`g_{E1}`$ and $`\stackrel{~}{g}_{E1}`$ being real). The resulting change in the asymmetry $`𝒜_\varphi `$ in $`K_L\pi ^+\pi ^{}e^+e^{}`$ is
$$𝒜_\varphi =15\%\mathrm{sin}\left(\phi _++\delta _0\delta _1\right)+38\%\left[\frac{g_{E1}}{|g_{M1}|}\mathrm{sin}\phi _++\frac{\stackrel{~}{g}_{E1}}{|g_{M1}|}\right].$$
(23)
The observed branching ratio for $`K_S\pi ^+\pi ^{}\gamma `$ limits $`|g_{E1}/g_{M1}|`$ to $`<5\%`$ , while typical estimates of $`|\stackrel{~}{g}_{E1}/g_{M1}|`$ are of order $`10^3`$ .
Needless to say, any incisive study of the channels $`K^0,\overline{K^0}\pi ^+\pi ^{}e^+e^{}`$ and $`K_{L,S}\pi ^+\pi ^{}\gamma `$ should be alert to the possible presence of correction terms of the form described above.
|
warning/0006/hep-ex0006025.html
|
ar5iv
|
text
|
# Measurement of 𝑑𝜎/𝑑𝑦 for High Mass Drell-Yan 𝑒⁺𝑒⁻ Pairs from 𝑝𝑝̄ Collisions at √𝑠=1.8 TeV
## Abstract
We report on the first measurement of the rapidity distribution $`d\sigma /dy`$ over nearly the entire kinematic region of rapidity for $`e^+e^{}`$ pairs in the $`Z`$-boson region of $`66<M_{ee}<116`$ GeV$`/c^2`$ and at higher mass $`M_{ee}>116`$ GeV$`/c^2`$. The data sample consists of 108 pb<sup>-1</sup> of $`p\overline{p}`$ collisions at $`\sqrt{s}=1.8`$ TeV taken by the Collider Detector at Fermilab during 1992–1995. The total cross section in the $`Z`$-boson region is measured to be $`252\pm 11`$ pb. The measured total cross section and $`d\sigma /dy`$ are compared with quantum chromodynamics calculations in leading and higher orders.
PACS numbers: 13.85.Qk, 12.38.Qk
Most measurements at high energy proton-antiproton colliders are performed in the central rapidity production region, $`|y|<1`$. A model dependent extrapolation for $`|y|>1`$ is needed to extract the total cross section for hard processes such as top quark production or $`W`$ and $`Z`$ boson production. This extrapolation is made using Monte Carlo programs (e.g. PYTHIA ), which incorporate quantum chromodynamics (QCD) calculations in leading order (LO) or next to leading order (NLO). A previous measurement of the rapidity distribution, $`d\sigma /dy`$, for dimuon pairs in the $`Z`$-bosons mass region was limited to $`|y|<1`$ . In this communication, we present the first measurement of $`d\sigma /dy`$ for $`e^+e^{}`$ pairs in the $`Z`$-boson mass and high mass region over nearly the entire kinematic region of rapidity. At the Tevatron $`p\overline{p}`$ collider, the kinematic limit at the $`Z`$-boson mass is $`|y|=3.0`$, while we measure $`|y|`$ up to 2.8. The $`d\sigma /dy`$ distributions are compared to the predictions of QCD in LO and NLO. This measurement is also relevant for precision $`W`$ boson mass measurements at hadron colliders, where $`W`$’s are reconstructed using $`e\nu `$ and $`\mu \nu `$ pairs from the Drell-Yan process.
In hadron-hadron collisions at high energies, massive $`e^+e^{}`$ pairs are produced via the Drell-Yan process. In the standard model, quark-antiquark annihilation form an intermediate $`\gamma ^{}`$ or $`Z`$ ($`\gamma ^{}/Z`$) vector boson, which then decays into an $`e^+e^{}`$ pair. In LO, the momentum fraction $`x_1`$ ($`x_2`$) of the partons in the proton (antiproton) are related to the rapidity , $`y`$, of the boson via the equation $`x_{1,2}=(M/\sqrt{s})e^{\pm y}`$. Here $`s`$ is the center of mass energy, and $`M`$ is the mass of the dilepton pair. Therefore, dilepton pairs which are produced at large rapidity originate from events in which one parton is at large $`x`$ and another parton is at very small $`x`$. Since the quark distributions for $`x`$ up to 0.9 are well constrained by the deep-inelastic lepton scattering experiments , comparisons of data and theory for $`d\sigma /dy`$, and the total cross sections provide a test of the theory, e.g. missing NNLO or power correction terms.
The $`e^+e^{}`$ pairs are from 108 pb<sup>-1</sup> of $`p\overline{p}`$ collisions at $`\sqrt{s}=1.8`$ TeV taken by the Collider Detector at Fermilab (CDF) during 1992–1993 ($`18.7\pm 0.7`$ pb<sup>-1</sup>) and 1994–1995 ($`89.1\pm 3.7`$ pb<sup>-1</sup>). CDF is a solenoidal magnetic spectrometer surrounded by projective-tower-geometry calorimeters and outer muon detectors. Only detector components used in this measurement are described here. Charged particle momenta and directions are measured by the spectrometer, which consists of a 1.4 T axial magnetic field, an 84-layer cylindrical drift chamber (CTC), an inner vertex tracking chamber (VTX), and a silicon vertex detector (SVX). The polar coverage of the CTC tracking is $`|\eta |<1.2`$. The $`p\overline{p}`$ collision point along the beam line $`(Z_{\mathrm{vertex}})`$ is determined using tracks in the VTX. The energies and directions of electrons, photons, and jets are measured by three separate calorimeters covering three regions: central $`(|\eta |<1.1)`$, end plug $`(1.1<|\eta |<2.4)`$, and forward $`(2.2<|\eta |<4.2)`$. Each region has an electromagnetic (EM) and hadronic (HAD) calorimeter.
In a previous letter, we presented $`d\sigma /dP_\mathrm{T}`$ of $`Z`$ boson . This analysis is an extension of the $`d\sigma /dP_\mathrm{T}`$ measurement. The $`d\sigma /dP_\mathrm{T}`$ analysis has three categories of $`e^+e^{}`$ pairs: central-central (CC), central-end plug (CP), and central-forward (CF). This analysis extends the sample to the forward rapidity region by including plug-plug (PP) and plug-forward (PF) events. The inclusion of these events increases the event sample by $`20\%`$ and allows for measurement of $`Z`$ bosons with $`|y|`$ up to 2.8. An improvement in this analysis is the additional VTX tracking requirements for plug and forward electrons. The VTX covers the entire rapidity range in this study, and plays an important role in removing background in the high $`\eta `$ region which is not covered by the CTC.
The sample of $`e^+e^{}`$ events was collected by a three-level online trigger that required an electron in either the central or the plug calorimeter. The offline analysis selects events with two or more electron candidates. Since the electrons from the Drell-Yan process are typically isolated, both electrons are required to be isolated from any other activity in the calorimeters. Electrons in the central, end plug, and forward regions are required to be within the fiducial area of the calorimeters and have a minimum $`E_\mathrm{T}`$ of 22, 20, and 15 GeV, respectively. To improve the purity of the sample, electron identification cuts are applied . For CC, CP, and CF events, the central electron (or one of them if there are two) is required to pass strict criteria. The criteria on the other electron are looser. A central electron must have a CTC track that extrapolates to the electron’s shower clusters in the EM calorimeter. These clusters must have EM-like transverse shower profiles. The track momentum and the EM shower energy must be consistent with one another. The track is also used to determine the position and direction of the central electron. The fraction of energy in the HAD calorimeter towers behind the EM shower is required to be consistent with that expected for an EM shower ($`E_{\mathrm{HAD}}/E_{\mathrm{EM}}`$). The plug electrons must also have an EM-like transverse shower profile. The end plug and forward electrons are required to pass the $`E_{\mathrm{HAD}}/E_{\mathrm{EM}}`$ requirement and to have a track in the VTX which originates from the same vertex as the other electron and points to the position of the electromagnetic cluster in the calorimeter. The ratio of found to expected hits in the VTX is required to be greater than $`70\%`$ and $`50\%`$ for plug and forward electrons, respectively. The VTX tracking efficiency is $`(97.8\pm 0.3)\%`$ for plug electrons and $`(97.0\pm 0.9)\%`$ for forward electrons. This requirement on plug and forward electrons reduce the background rates from $`11\%`$ to $`2\%`$ for PP events, and from $`22\%`$ to $`4\%`$ for PF events.
The data sample is divided into two mass bins: the $`Z`$ region ($`66<M_{ee}<116`$ GeV/$`c^2`$) and the high mass region ($`M_{ee}>116`$ GeV/$`c^2`$). After all cuts, the numbers of CC, CP, CF, PP, and PF events in the $`Z`$ mass region are 2894, 3811, 621, 1236, and 589, respectively. The backgrounds are low and are estimated using the data. The backgrounds in the CP, CF, PP, and PF topologies are dominated by jets and are $`28\pm 5`$, $`13\pm 3`$, $`24\pm 2`$, and $`23\pm 3`$ events, respectively. Because of the CTC tracking requirement, the jet background for the CC sample is negligible. The CC background is mainly from $`e^+e^{}`$ pairs from $`W^+W^{}`$, $`\tau ^+\tau ^{}`$, $`c\overline{c}`$, $`b\overline{b}`$, and $`t\overline{t}`$ sources. This background, estimated using $`e^\pm \mu ^{}`$ pairs , is $`3\pm 2`$ events. In the high mass bin, the numbers of CC, CP, CF, PP, and PF events are 61, 59, 9, 18, and 5, respectively. The mean mass of the high mass bin is 152.6 GeV/$`c^2`$.
The acceptance for Drell-Yan $`e^+e^{}`$ pairs is obtained using the Monte Carlo event generator, PYTHIA , and CDF detector simulation programs. PYTHIA generates the LO QCD interaction ($`q+\overline{q}\gamma ^{}/Z`$), simulates initial state QCD radiation via its parton shower algorithms, and generates the decay, $`\gamma ^{}/Ze^+e^{}`$. To approximate higher order QCD corrections to the LO mass distribution, a “$`K`$-factor” is used as an event weight: $`K(M^2)=1+\frac{4}{3}(1+\frac{4}{3}\pi ^2)\alpha _s(M^2)/2\pi `$, where $`\alpha _s`$ is the two loop QCD coupling. This factor improves the agreement between the NLO and LO mass spectra. (For $`M>50`$ GeV/$`c^2`$, $`1.25<K<1.36`$.) The CTEQ3L PDFs are used in the acceptance calculations. Final state QED radiation from the $`\gamma ^{}/Ze^+e^{}`$ vertex is added by the PHOTOS Monte Carlo program. Generated events are processed by CDF detector simulation programs and are reconstructed as data. The calorimetry energy scales and resolutions used in the detector simulation are extracted from the data, as are the cut efficiencies and corresponding errors. Simulated events are accepted if after the reconstruction they pass the $`e^+e^{}`$ pair mass, the detector fiducial, and kinematic ($`E_T`$) cuts.
In the $`d\sigma /dy`$ measurement, various samples are combined and binned in $`y`$. The $`d\sigma /dy`$ is calculated with
$$\frac{d\sigma }{dy}=\frac{\mathrm{\Delta }N}{C\mathrm{\Delta }y_r_rϵA_r}.$$
The $`\mathrm{\Delta }N`$ is the background-subtracted event count in a $`y`$ bin, $`C`$ is a bin centering correction, $`\mathrm{\Delta }y`$ is the bin width, the sum $`r`$ is over the 1992–1993 and 1994–1995 runs, $`_r`$ is the integrated luminosity, and $`ϵA_r`$ is the run’s combined event selection efficiency and acceptance. The backgrounds subtracted from the event count are predicted using the data and background samples. The factor $`C`$ corrects the average value of the cross section in the bin to its bin center value. Acceptances are calculated separately for CC, CP, CF, PP, and PF pairs. They are combined with the corresponding event selection efficiencies to give $`ϵA_r`$. Figure 1 shows the $`ϵA_r`$ for events in the $`Z`$ region as a function of $`y`$ for the CC sample, the CC+CP+CF sample, and the CC+CP+CF+PP+PF sample, respectively. It indicates that the PP+PF events extend the acceptance of the $`y`$ measurement to $`|y|=2.8`$. We significantly increase our statistics by extending the acceptance beyond $`y>1.2`$. The $`d\sigma /dy`$ of $`e^+e^{}`$ events in the $`Z`$ mass region is shown in Table I.
The systematic errors considered are from variations in the background estimates using different methods, the background in the efficiency sample, the uncertainty in energy resolution of the calorimeter, the choice of PDFs, and the distribution of $`Z`$ $`P_\mathrm{T}`$ used in the Monte Carlo event generator. The systematic error from the calorimeter resolution is $`0.2\%`$ in the low $`y`$ region and increases to $`0.7\%`$ at $`|y|>2.5`$. The systematic error from variations in the $`Z`$ $`P_\mathrm{T}`$ is $`0.5\%`$ at $`|y|=0`$ and is $`2.0\%`$ at $`|y|=2.8`$. The systematic error from the choice of PDFs is less than $`1\%`$ in $`|y|<2.0`$ and increases to $`2\%`$ at $`|y|=2.8`$. For the total cross section measurement, the combined systematic error is $`0.6\%`$ (excluding the luminosity uncertainty). The $`p\overline{p}`$ collision luminosity is derived with CDF’s beam-beam cross section, $`\sigma _{\mathrm{BBC}}=51.15\pm 1.60`$ mb . The luminosity error of 3.9$`\%`$ contains the $`\sigma _{\mathrm{BBC}}`$ error and uncertainties specific to running conditions.
Figures 2(a) and 2(b) compare the measured $`d\sigma /dy`$’s to theoretical predictions in the $`Z`$ mass and high mass regions, respectively. The top horizontal axes on these figures are the corresponding values of the $`x_1`$ and $`x_2`$ as a function of $`y`$. The predictions are LO calculations with CTEQ5L PDFs and NLO calculations with MRST99 and CTEQ5M-1 PDFs. The predictions in Figure 2(a) have been normalized by a factor “F”, the ratio of measured total cross section to the prediction (F=1.51, 1.14, and 1.13 for the CTEQ5L, MRST99 and CTEQ5M-1 PDFs, respectively). The predictions in Figure 2(b) are normalized to the data using the factor F from the $`Z`$ mass region. We compare the data to the theory using statistical errors only. As the $`\chi ^2`$ values listed in Figure 2(a) indicate, the LO calculation using recent LO PDFs does not fit the shape as well as the NLO calculation with the most recent NLO PDFs.
Model independent measurements of the total production cross sections for $`e^+e^{}`$ pairs are extracted by integrating the measured values of $`d\sigma /dy`$. Because there are no data for $`|y|>2.8`$, we use a NLO calculation with the CTEQ5M-1 PDFs to estimate $`\gamma ^{}/Z`$ production in that region. The cross section in the region of $`|y|>2.8`$ is about $`0.02\%`$ of the predicted total cross section. The extracted cross section in the $`Z`$ mass region is $`252.1\pm 3.9(\mathrm{stat}.)\pm 1.6(\mathrm{syst}.)\pm 9.8(\mathrm{lum}.)`$ pb. The corresponding $`\sigma (p\overline{p}Z)Br(Zee)`$ is $`253\pm 4(\mathrm{stat}.+\mathrm{syst}.)\pm 10(\mathrm{lum}.)`$ pb. This measurement is in agreement with our previous measurements in the dielectron ($`248\pm 5(\mathrm{stat}.+\mathrm{syst}.)\pm 10(\mathrm{lum}.)`$ pb, using only the CC+CP+CF $`e^+e^{}`$ sample), and dimuon ($`237\pm 9(\mathrm{stat}.+\mathrm{syst}.)\pm 9(\mathrm{lum}.)`$ pb using a CC sample) channels. These previous measurements use a QCD model to correct for the missing events at high rapidity. The combined $`e^+e^{}`$ and $`\mu ^+\mu ^{}`$ cross section is $`250\pm 4`$(stat.+syst.)$`\pm 10`$ pb. Since the $`p\overline{p}`$ inelastic cross section used by CDF in luminosity calculations differs from DO/ ’s by $`+5.9\%`$ , measured cross sections must be renormalized before comparisons to other experiments. The DO/ measurement of $`221\pm 11`$ pb when renormalized to the CDF luminosity measurement is $`234\pm 12`$ pb.
The total cross section measurements can also be compared to QCD calculations. Fixed order QCD calculations have uncertainties from PDF measurements and corrections from higher orders of perturbation theory, i.e., the $`K`$-factor. The NLO-to-LO total cross section correction is significant: $`K1.4`$. In contrast, the NLO total cross section is lower than NNLO prediction by only $`2.3\%`$. The NNLO prediction with the latest NLO MRST99 PDFs is $`227\pm 9`$ pb, where the $`4\%`$ error is mostly from uncertainties in the NLO PDFs. Although a full set of NNLO PDFs is not available, recent estimates of NNLO PDFs indicate that the NNLO PDFs will increase the theoretical cross sections by $`5\%`$. Given these uncertainties, the theoretical expectation is consistent with the $`Z`$ cross section measurements.
The measurement of the Drell-Yan total cross section in the high mass region is $`4.0\pm 0.4(\mathrm{stat}.+\mathrm{syst}.)\pm 0.2(\mathrm{lum}.)`$ pb. The corresponding prediction of the total cross section from the NNLO QCD theory using MRST99 PDFs is $`3.3`$ pb.
In summary, the rapidity distributions of $`e^+e^{}`$ pairs in the $`Z`$ boson mass and high mass region have been measured for the first time over nearly the entire kinematic region. This measurement uses a new tracking technique in the high rapidity region to reduce the background and uncertainties associated with it. In addition, unlike the previous measurement of the total cross section, this measurement is model independent.
The vital contributions of the Fermilab staff and the technical staffs of the participating institutions are gratefully acknowledged. This work is supported by the U.S. Department of Energy and National Science Foundation; the Natural Sciences and Engineering Research Council of Canada; the Istituto Nazionale di Fisica Nucleare of Italy; the Ministry of Education, Science and Culture of Japan; the National Science Council of the Republic of China; the A.P. Sloan Foundation, and the Swiss National Science Foundation.
|
warning/0006/cond-mat0006444.html
|
ar5iv
|
text
|
# Parametric Simultons in Nonlinear Lattices
## I INTRODUCTION
Since the pioneering work of Fermi, Pasta, and Ulam on the nonlinear dynamics in lattices, the understanding of the dynamical localization in ordered, spatially extended discrete systems have experienced considerable progress. In particular, the lattice solitons, which are localized nonlinear excitations due to the balance between nonlinearity and dispersion of the system, are shown to exist, and many important applications are found in transport of energy, proton contactivity, structural phase transition and associated central-peak phenomena, etc. In recent years, the interest in localized excitations in nonlinear lattices has been renewed due to the identification of a new type of anharmonic localized modes\[4-6\]. These modes, called the intrinsic localized modes (ILM’s), or discrete breathers, are the discrete analog of the lattice envelope (or breather) solitons with their spatial extension only a few lattice spacing and the vibrating frequency above the upper cutoff of phonon spectrum band. The ILM’s have been observed in a number of experiments\[7-13\]. Recently, much attention has been paid to the gap solitons in diatomic lattices\[14-25\]. In linear case, a diatomic lattice allows two phonon bands. There is a upper cutoff for phonon frequency and a frequency gap (forbidden band) between acoustic and optical bands, induced by mass and/or force-constant difference of two different types of particles. No interaction occurs between phonons and the phonons can not propagate in the system when their frequencies are in the gap or above the phonon bands. However, these properties of the phonons change drastically when nonlinearity is introduced into the system. New types of localized modes, especially the gap solitons, may appear as nonlinear localized excitations with their vibrating frequencies in the band gap. The gap solitons and ILM’s as well as their chaotic motion have been observed in damped and parametrically excited one-dimensional (1D) diatomic pendulum lattices\[26-28\].
On the other hand, in recent years numerous achievements have been made for optical solitons in nonlinear optical media\[29-31\]. Besides the temporal optical solitons, which is promising for long-distance information transmission in fiber, spatial optical solitons also attract much attention. The spatial optical solitons are believed to be the candidates for all-optical devices, such as optical switches and logic gates, etc. Recently, the study of optical parametric processes, in particular the second harmonic generation (SHG), which marked the birth of nonlinear optics, has generated a great deal of new interest. It was suggested that it is possible to obtain large nonlinear phase shifts by using a cascaded second-order nonlinearity. In 1974, Karamzin and Sukhorukov (KS) recognized that the cascaded second-order parametric processes may support solitons under general phase-matching conditions. They derived two coupled nonlinear equations for the envelopes of the fundamental and second harmonic waves. The difference between the KS equations and the envelope equations for usual SHG is the inclusion of dispersion and/or diffraction, which are necessary for short pulses and/or narrow light beams. Simultaneous solitons (i. e. two components are solitons) are found for the KS equations and these solitons are later termed as the parametric simultons. The concept of the simultons has been generalized to the nonlinear optical media with periodically varying refractive index. Since the eigenspectrum of linear electromagnetic waves consists of many photonic bands and the vibrating frequencies of the simultons may be in the gaps between these bands, the name parametric band-gap simulton is given by Drummond et al\[37-39\]. Different from the self-trapping mechanism for Keer solitons, the formation of the simultons is due to the energy transfer and mutual self-trapping between the fundamental and the second harmonic waves.
On the contrary, the SHG in lattices is much less investigated. Although in the standard textbook of solid state physics there exists a simple experimental description for three phonon processes in solids, it seems that there is no detailed theoretical approach to the SHG in nonlinear lattices until recently. In a recent paper, Konotop considered theoretically the SHG in a nonlinear diatomic lattice and obtained some interesting results.
In many aspects a nonlinear lattice is similar to a nonlinear, periodic optical media. The discreteness of lattice results in the symmetry breaking for continuous translation and makes the property of the system periodic, in particular the frequency spectrum of corresponding linear wave splits into many bands. It should be stressed that the SHG does not occur in 1D monatomic lattices (see the next section). However, a SHG can be realized if we consider nonlinear multi-atomic lattices . The reason is that in the monatomic lattices, an efficient energy transfer (resonance) between any two modes in the system does not occur. But the situation is different for the multi-atomic lattices. A multi-atomic lattice allows many branches of linear dispersion relation, and the dispersion relation is periodic with respect to lattice wave vector. It is just the multiple-value and periodic property of the dispersion relation makes it possible that the phase-matching condition for the SHG, i. e. the condition by which the resonance between the fundamental and second harmonic waves may occur, can be satisfied by selecting the wave vectors and the corresponding frequencies from different spectrum branches.
Motivated by the study of the optical simultons, in this paper we show that lattice simultons are possible in the multiatomic lattices with cubic nonlinearity (different from the case in nonlinear optics, here the order of nonlinearity means the order in the Hamiltonian of the system). The paper is organized as follows. The next section presents our model and an asymptotic expansion based on a quasi-discreteness approach. In section III we solve the KS equations derived in section II and provide some lattice simulton solutions. A discussion and summary is given in the last section.
## II MODEL AND ASYMPTOTIC EXPANSION
### A The model
As mentioned in the last section, the SHG may in princile occur in any multi-atomic lattice, but for definiteness and for the sake of simplicity we consider here a 1D diatomic lattice with a cubic interaction potential. The Hamiltonian of the system is given by
$`H={\displaystyle \underset{n}{}}[{\displaystyle \frac{1}{2}}m\left({\displaystyle \frac{dv_n}{dt}}\right)^2+{\displaystyle \frac{1}{2}}M\left({\displaystyle \frac{dw_n}{dt}}\right)^2+{\displaystyle \frac{1}{2}}K_2(w_nv_n)^2+{\displaystyle \frac{1}{2}}K_2^{}(v_{n+1}w_n)^2`$ (1)
$`+{\displaystyle \frac{1}{3}}K_3(w_nv_n)^3+{\displaystyle \frac{1}{3}}K_3^{}(v_{n+1}w_n)^3+{\displaystyle \frac{1}{3}}V_3v_n^3+{\displaystyle \frac{1}{3}}V_3^{}w_n^3],`$ (2)
where $`v_n`$=$`v_n(t)`$ ($`w_n=w_n(t)`$ ) is the displacement from its equilibrium position of the $`n`$th particle with mass $`m`$ ($`M`$). $`n`$ is the index of the $`n`$th unit cell with a lattice constant $`a=2a_0`$, $`a_0`$ is the equilibrium lattice spacing between two adjacent particles. Here for generality we assume that the nearest-neighbor force constants $`K_j(j=2,3)`$ in the same cells are different from the nearest-neighbor force constants $`K_j^{}(j=2,3)`$ in different cells. $`V_3`$ and $`V_3^{}`$ are the force constants related to the on-site cubic potential for two types of particles. Without loss of generality we assume $`m<M`$, $`K_j^{}K_j(j=2,3)`$, and $`V_3^{}V_3`$. The equations of motion for describing the lattice read
$`{\displaystyle \frac{d^2}{dt^2}}v_n=I_2(w_nv_n)+I_2^{}(w_{n1}v_n)+I_3(w_nv_n)^2I_3^{}(w_{n1}v_n)^2\alpha _mv_n^2,`$ (3)
$`{\displaystyle \frac{d^2}{dt^2}}w_n=J_2(v_nw_n)+J_2^{}(v_{n+1}w_n)J_3(v_nw_n)^2+J_3^{}(v_{n+1}w_n)^2\alpha _Mw_n^2,`$ (4)
where $`I_j=K_j/m,I_j^{}=K_j^{}/m,J_j=K_j/M,J_j^{}=K_j^{}/M(j=2,3),\alpha _m=V_3/m`$ and $`\alpha _M=V_3^{}/M`$. The linear dispersion relation of Eqs. (2) and (3) is given by
$$\omega _\pm (q)=\frac{1}{\sqrt{2}}\left\{(I_2+I_2^{}+J_2+J_2^{})\pm \left[(I_2+I_2^{}+J_2+J_2^{})^216I_2J_2^{}\mathrm{sin}^2(qa/2)\right]^{\frac{1}{2}}\right\}^{\frac{1}{2}},$$
(5)
where the minus (plus) sign corresponds to acoustic (optical) mode. Thus we have two phonon bands $`\omega _\pm (q)`$ and obviously $`\omega _\pm (q+Q)=\omega _\pm (q)`$, here $`Q=2j\pi /a`$, $`j`$ is an integer and $`Q`$ is the reciprocal lattice vector of the system. At the wave number $`q=0`$, the phonon spectrum has a lower cutoff $`\omega _{}(0)=0`$ for the acoustic mode and an upper cutoff $`\omega _+(0)=(I_2+I_2^{}+J_2+J_2^{})^{1/2}`$ for the optical mode. At $`q=\pi /a`$ there exists a frequency gap between the upper cutoff of the acoustic branch $`\omega _{}(\pi /a)`$ and the lower cutoff of the optical branch $`\omega _+(\pi /a)`$, where $`\omega _\pm (\pi /a)=(1/\sqrt{2})\{(I_2+I_2^{}+J_2+J_2^{})\pm [(I_2+I_2^{}+J_2+J_2^{})^216I_2J_2^{}]^{1/2}\}^{1/2}`$. The width of the frequency gap is $`\omega _+(\pi /a)\omega _{}(\pi /a)`$, which approaches zero when $`mM`$ and $`K_2^{}K_2`$. This is just the limit of monatomic lattice with the lattice constant $`a_0=a/2`$. We assume the gap is not small, i. e. we have $`(1m/M)`$ and $`(1K_2/K_2^{})`$ are of order unity.
Because of the periodic property of $`\omega _\pm (q)`$, the condition of a second harmonic resonance in the system (2) and (3) reads
$`q_2=2q_1+Q,`$ (6)
$`\omega _2=2\omega _1,`$ (7)
where $`q_1(q_2)`$ and $`\omega _1(\omega _2)`$ are the wave vector and frequency of the corresponding fundamental (second harmonic) wave, respectively. Eqs.(5) and (6) are also called the phase-matching conditions for the SHG. It is easy to show that in the limit $`mM`$ and $`K_2^{}K_2`$ the condition (5) and (6) can not be satisfied except for zero-frequency mode, i. e. the SHG is impossible in monatomic lattices. For the diatomic lattice, in order to fulfil (5) and (6) we may chose $`\omega _1\omega _{}(q)`$ and $`\omega _2\omega _+(q)`$, then the conditions (5) and (6) give
$`\left[(I_2+I_2^{}+J_2+J_2^{})^24I_2J_2^{}\mathrm{sin}(q_1a)\right]^{1/2}`$ (8)
$`=3(I_2+I_2^{}+J_2+J_2^{})4\left[(I_2+I_2^{}+J_2+J_2^{})^216I_2J_2^{}\mathrm{sin}^2(q_1a/2)\right]^{1/2}.`$ (9)
It is available to solve $`q_1`$ from the above equation. For simplicity we consider cutoff modes of the system. We take $`q_1=\pi /a,q_2=0`$ and $`Q=2\pi /a`$, then the condition (5) is automatically satisfied. The condition (6) (the same as (7) ) now reads
$$I_2+I_2^{}+J_2+J_2^{}=\frac{8}{\sqrt{3}}\sqrt{I_2J_2^{}}.$$
(10)
Eq.(8) also means that $`\omega _1=\omega _{}(\pi /a)=(1/2)(I_2+I_2^{}+J_2+J_2^{})^{1/2}=(4I_2J_2^{}/3)^{1/4}`$ and $`\omega _2=\omega _+(0)=(I_2+I_2^{}+J_2+J_2^{})^{1/2}=2(4I_2J_2^{}/3)^{1/4}`$ If the all harmonic force constants are equal, i. e. $`K_2^{}=K_2`$, Eq.(8) gives $`m=M/3`$. Another particular case is all masses are the same, i. e. $`m=M`$. In this case Eq.(8) requires $`K_2^{}=K_2/3`$. In general, the phase-matching conditions (5) and (6) impose a constraint on masses and harmonic force constants of the lattice.
### B Asymptotic expansion
We employ the quasi-discreteness approach (QDA) developed in Refs. 17 and 24 for diatomic lattices to investigate the SHG in the system (2) and (3). We are interested in the cascading processes of the system in which the width of excitation is narrower than usual SHG case. Thus we use different spatial-temporal scales in deriving the envelope equations for the fundamental and the second harmonic waves. We make the expansion
$$u_n(t)=ϵ\left[u_{n,n}^{(0)}+ϵ^{1/2}u_{n,n}^{(1)}+ϵu_{n,n}^{(2)}+\mathrm{}\right],$$
(11)
where $`u_n(t)`$ represents $`v_n(t)`$ or $`w_n(t)`$. $`ϵ`$ is a smallness and ordering parameter denoting the relative amplitude of the excitation and $`u_{n,n}^{(\nu )}=u^{(\nu )}(\xi _n,\tau ;\varphi _n(t))`$, with
$`\xi _n=ϵ^{1/2}(na\lambda t),`$ (12)
$`\tau =ϵt,`$ (13)
$`\varphi _n=qna\omega (q)t,`$ (14)
where $`\lambda `$ is a parameter to be determined by a solvability condition (see below). Substituting (9)-(12) into Eqs.(2) and (3) and equating the coefficients of the same powers of $`ϵ`$, we obtain
$`{\displaystyle \frac{^2}{t^2}}v_{n,n}^{(j)}I_2(w_{n,n}^{(j)}v_{n,n}^{(j)})I_2^{}(w_{n,n1}^{(j)}v_{n,n}^{(j)})=M_{n,n}^{(j)},`$ (15)
$`M_{n,n}^{(0)}=0,`$ (16)
$`M_{n,n}^{(1)}=2\lambda {\displaystyle \frac{^2}{t\xi _n}}v_{n,n}^{(0)}I_2^{}a{\displaystyle \frac{}{\xi _n}}w_{n,n1}^{(0)},`$ (17)
$`M_{n,n}^{(2)}=2\lambda {\displaystyle \frac{^2}{t\xi _n}}v_{n,n}^{(1)}\left(2{\displaystyle \frac{^2}{t\tau }}+\lambda ^2{\displaystyle \frac{^2}{\xi _n^2}}\right)v_{n,n}^{(0)}+I_2^{}\left(a{\displaystyle \frac{}{\xi _n}}w_{n,n1}^{(1)}+{\displaystyle \frac{a^2}{2!}}{\displaystyle \frac{^2}{\xi _n^2}}w_{n,n1}^{(0)}\right)`$ (18)
$`+I_3(w_{n,n}^{(0)}v_{n,n}^{(0)})^2I_3^{}(w_{n,n1}^{(0)}v_{n,n}^{(0)})^2\alpha _m(v_{n,n}^{(0)})^2,`$ (19)
$`\mathrm{}\mathrm{}`$ (20)
and
$`{\displaystyle \frac{^2}{t^2}}w_{n,n}^{(j)}J_2(v_{n,n}^{(j)}w_{n,n}^{(j)})J_2^{}(v_{n,n+1}^{(j)}w_{n,n}^{(j)})=N_{n,n}^{(j)},`$ (21)
$`N_{n,n}^{(0)}=0,`$ (22)
$`N_{n,n}^{(1)}=2\lambda {\displaystyle \frac{^2}{t\xi _n}}w_{n,n}^{(0)}+J_2^{}a{\displaystyle \frac{}{\xi _n}}v_{n,n+1}^{(0)},`$ (23)
$`N_{n,n}^{(2)}=2\lambda {\displaystyle \frac{^2}{t\xi _n}}w_{n,n}^{(1)}\left(2{\displaystyle \frac{^2}{t\tau }}+\lambda ^2{\displaystyle \frac{^2}{\xi _n^2}}\right)w_{n,n}^{(0)}+J_2^{}\left(a{\displaystyle \frac{}{\xi _n}}v_{n,n+1}^{(1)}+{\displaystyle \frac{a^2}{2!}}{\displaystyle \frac{^2}{\xi _n^2}}v_{n,n+1}^{(0)}\right)`$ (24)
$`J_3(v_{n,n}^{(0)}w_{n,n}^{(0)})^2+J_3^{}(v_{n,n+1}^{(0)}w_{n,n}^{(0)})^2\alpha _M(w_{n,n}^{(0)})^2,`$ (25)
$`\mathrm{}\mathrm{}`$ (26)
with $`j=0,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}2},\mathrm{}`$. Eqs.(13) and (17) can be rewritten in the following form
$`\widehat{L}w_{n,n}^{(j)}=J_2M_{n,n}^{(j)}+J_2^{}M_{n,n+1}^{(j)}+\left({\displaystyle \frac{^2}{t^2}}+I_2+I_2^{}\right)N_{n,n}^{(j)},`$ (27)
$`\left({\displaystyle \frac{^2}{t^2}}+I_2+I_2^{}\right)v_{n,n}^{(j)}=I_2w_{n,n}^{(j)}+I_2^{}w_{n,n1}^{(j)}+M_{n,n}^{(j)},`$ (28)
where the operator $`\widehat{L}`$ is defined by
$`\widehat{L}u_{n,n}^{(j)}=\left({\displaystyle \frac{^2}{t^2}}+I_2+I_2^{}\right)\left({\displaystyle \frac{^2}{t^2}}+J_2+J_2^{}\right)u_{n,n}^{(j)}(I_2J_2+I_2^{}J_2^{})u_{n,n}^{(j)}`$ (29)
$`I_2J_2^{}\left(u_{n,n+1}^{(j)}+u_{n,n1}^{(j)}\right),`$ (30)
where $`u_{n,n}^{(j)}(j=0,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}2},\mathrm{})`$ are a set of arbitrary functions. From Eq.(21) we can solve $`w_{n,n}^{(j)}`$ and obtain a series of solvability conditions (envelope equations) whereas Eq.(22) is used to solve $`v_{n,n}^{(j)}`$.
### C Envelope equations for cascading processes
We now solve Eqs.(22) and (23) order by order. For $`j=0`$ it is easy to get
$`w_{n,n}^{(0)}=A_1(\tau ,\xi _n)\mathrm{exp}(i\varphi _n^{})+A_2(\tau ,\xi _n)\mathrm{exp}(i\varphi _n^+)+\mathrm{c}.\mathrm{c}.,`$ (31)
$`v_{n,n}^{(0)}={\displaystyle \frac{I_2+I_2^{}e^{iqa}}{\omega _{}^2+I_2+I_2^{}}}A_1(\tau ,\xi _n)\mathrm{exp}(i\varphi _n^{})+{\displaystyle \frac{I_2+I_2^{}e^{iqa}}{\omega _+^2+I_2+I_2^{}}}A_2(\tau ,\xi _n)\mathrm{exp}(i\varphi _n^+)+\mathrm{c}.\mathrm{c}.`$ (32)
with $`\varphi _n^\pm =qna\omega _\pm (q)t`$. $`\omega _\pm (q)`$ have been given in Eq.(4). $`A_1`$ and $`A_2`$ are yet to be determined two envelope (or amplitude) functions of the acoustic and the optical excitations, respectively. They are the functions of the slow variables $`\xi _n`$ and $`\tau `$. c. c. denotes the corresponding complex conjugate. For simplicity we specify two modes, i. e. the acoustic upper cutoff mode ($`q_1=\pi /a`$, $`\omega _1=\omega _{}(\pi /a)=(4I_2J_2^{}/3)^{1/4}`$ ) and the optical upper cutoff mode ($`q_2=0`$, $`\omega _2=\omega _+(0)=2\omega _1=2(4I_2J_2^{}/3)^{1/4}`$ ). Thus we have
$`w_{n,n}^{(0)}=A_1(\tau ,\xi _n)(1)^n\mathrm{exp}(i\omega _1t)+A_2(\tau ,\xi _n)\mathrm{exp}(i\omega _2t)+\mathrm{c}.\mathrm{c}.,`$ (33)
$`v_{n,n}^{(0)}={\displaystyle \frac{I_2I_2^{}}{\omega _1^2+I_2+I_2^{}}}A_1(\tau ,\xi _n)(1)^n\mathrm{exp}(i\omega _1t)`$ (34)
$`+{\displaystyle \frac{I_2+I_2^{}}{\omega _2^2+I_2+I_2^{}}}A_2(\tau ,\xi _n)\mathrm{exp}(i\omega _2t)+\mathrm{c}.\mathrm{c}..`$ (35)
From the discussion in subsection II. A, the modes chosed in such way satisfy the phase-matching conditions (5) and (6) for the SHG. Thus in Eqs.(27) and (28) $`A_1`$ ($`A_2`$) represents the amplitude of the fundamental (second harmonic) wave, respectively.
In the next order ($`j`$=1), a solvability condition of Eqs.(21) and (22) requires the parameter $`\lambda =0`$, thus $`\xi _n=na`$. The second-order solution reads
$`w_{n,n}^{(1)}=B_0+[B_1(1)^n\mathrm{exp}(i\omega _1t)+B_2\mathrm{exp}(i\omega _2t)+\mathrm{c}.\mathrm{c}.],`$ (36)
$`v_{n,n}^{(1)}=B_0+\{{\displaystyle \frac{(I_2I_2^{})B_1+I_2^{}aA_1/\xi _n}{\omega _1^2+I_2+I_2^{}}}(1)^n\mathrm{exp}(i\omega _1t)`$ (37)
$`+{\displaystyle \frac{(I_2+I_2^{})B_2I_2^{}aA_2/\xi _n}{\omega _2^2+I_2+I_2^{}}}\mathrm{exp}(i\omega _2t)+\mathrm{c}.\mathrm{c}.\},`$ (38)
where $`B_j`$ ($`j`$=0, 1, 2) are undetermined functions of $`\xi _n`$ and $`\tau `$.
In the order $`j`$=2, we have the third-order approximate equatuion
$$\widehat{L}w_{n,n}^{(2)}=J_2M_{n,n}^{(2)}+J_2^{}M_{n,n+1}^{(2)}+\left(\frac{^2}{t^2}+I_2+I_2^{}\right)N_{n,n}^{(2)}.$$
(39)
Eq.(22) is not necessary since from (30) we can obtain closed equations for $`A_1`$ and $`A_2`$. Using Eqs.(26)-(29) we can get $`M_{n,n}^{(2)}`$, $`M_{n,n+1}^{(2)}`$ and $`N_{n,n}^{(2)}`$. By a detailed calculation we obtain the sovability condition of Eq.(30)
$`i{\displaystyle \frac{A_1}{\tau }}+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_1{\displaystyle \frac{^2A_1}{\xi _n^2}}+\mathrm{\Delta }_1A_1^{}A_2=0,`$ (40)
$`i{\displaystyle \frac{A_1}{\tau }}+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_2{\displaystyle \frac{^2A_2}{\xi _n^2}}+\mathrm{\Delta }_2A_1^2=0,`$ (41)
where the coefficients are expressed as
$`\mathrm{\Gamma }_1={\displaystyle \frac{I_2^{}J_2^{}a^2}{\omega _1[\lambda _1^1+\lambda _1(I_2I_2^{})(J_2J_2^{})]}},`$ (42)
$`\mathrm{\Gamma }_2={\displaystyle \frac{I_2J_2^{}a^2}{\omega _2[\lambda _2^1\lambda _2(I_2+I_2^{})(J_2+J_2^{})]}},`$ (43)
$`\mathrm{\Delta }_1={\displaystyle \frac{[1\lambda _2(I_2+I_2^{})]\lambda _3\lambda _1^1\alpha _M\lambda _1\lambda _2(I_2^2(I_2^{})^2)(J_2J_2^{})\alpha _m}{\omega _1[\lambda _1^1+\lambda _1(I_2I_2^{})(J_2J_2^{})]}},`$ (44)
$`\mathrm{\Delta }_2={\displaystyle \frac{\lambda _4\lambda _2^1\alpha _M\lambda _1^2(I_2I_2^{})^2(J_2+J_2^{})\alpha _M}{2\omega _2[\lambda _2^1+\lambda _2(I_2+I_2^{})(J_2+J_2^{})]}},`$ (45)
with
$`\lambda _j={\displaystyle \frac{1}{\omega _j^2+I_2+I_2^{}}},(j=1,2)`$ (46)
$`\lambda _3=(I_3I_3^{})(J_2J_2^{})\lambda _1^1(J_3J_3^{})+(I_2I_2^{})\left[(J_3+J_3^{})\lambda _1(I_3+I_3^{})(J_2J_2^{})\right],`$ (47)
$`\lambda _4=[1\lambda _1(I_2I_2^{})]^2[J_3\lambda _2^1+I_3(J_2+J_2^{})]+[1+\lambda _1(I_2I_2^{})]^2[J_3^{}\lambda _2^1I_3^{}(J_2+J_2^{})].`$ (48)
Introducing the transformation $`u_j=ϵA_j(j=1,2)`$ and noting that $`\xi _n=ϵ^{1/2}x_n(x_nna)`$ and $`\tau =ϵt`$, Eqs.(31) and (32) can be rewritten into the form
$`i{\displaystyle \frac{u_1}{t}}+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_1{\displaystyle \frac{^2u_1}{x_n^2}}+\mathrm{\Delta }_1u_1^{}u_2=0,`$ (49)
$`i{\displaystyle \frac{u_2}{t}}+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_2{\displaystyle \frac{^2u_2}{x_n^2}}+\mathrm{\Delta }_2u_1^2=0.`$ (50)
We should point out that Eqs.(5) and (6) are perfect phase-matching conditions for the SHG. If we allow a small mismatch for frequency, $`\delta \omega `$, the conditions (5) and (6) become
$$\omega _2=2\omega _1+\delta \omega ,q_2=2q_1+Q.$$
(51)
In this case Eqs.(40) and (41) change into
$`i\left({\displaystyle \frac{u_1}{t}}+v_1{\displaystyle \frac{u_1}{x_n}}\right)+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_1{\displaystyle \frac{^2u_1}{x_n^2}}+\mathrm{\Delta }_1u_1^{}u_2\mathrm{exp}(i\delta \omega t)=0,`$ (52)
$`i\left({\displaystyle \frac{u_2}{t}}+v_2{\displaystyle \frac{u_2}{x_n}}\right)+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_2{\displaystyle \frac{^2u_2}{x_n^2}}+\mathrm{\Delta }_2u_1^2\mathrm{exp}(i\delta \omega t)=0,`$ (53)
where $`v_j(j=1,2)`$ are the group velocities of of the fundamental and the second harmonic waves near at $`q=\pi /a`$ and $`q=0`$, respectively.
Eqs.(43) and (44) are the coupled-mode equations for the fundamental and the second harmonic waves. Such equations have been obtained by Karamzin and Sukhorukov in the context of nonlinear optics. One of important features of the KS equations is the inclusion of dispersion, which is absent in usual SHG envelope equations.
## III LATTICE SIMULTON SOLUTIONS
In this section, we solve the KS equations (43) and (44) derived in our lattice model and thus present some lattice simulton solutions for the system (2) and (3). In general, the property of the solutions of Eqs.(43) and (44) depends strongly on the coefficients appearing in the equations, in particular on their signs. At first we notice that in our system, $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$, which are respectively the group-velocity dispersion of the fundamental and the second harmonic waves, are both negative. But the signs of the nonlinear coefficients, $`\mathrm{\Delta }_1`$ and $`\mathrm{\Delta }_2`$, may be generally of both signs. Thus the situation here is different from the KS equations derived for the cascading process in nonlinear optics, where the nonlinear coefficients have the same sign, while the group-velocity dispersions may have different signs.
To solve Eqs.(43) and (44), we make the transformation
$`u_1(x_n,t)=U_1(\zeta )\mathrm{exp}[i(k_1x_n\mathrm{\Omega }_1t)],`$ (54)
$`u_2(x_n,t)=U_2(\zeta )\mathrm{exp}[i(k_2x_n\mathrm{\Omega }_2t)],`$ (55)
with $`\zeta =kx_n\mathrm{\Omega }t`$. Substituting (45) and (46) into (43) and (44), we obtain
$`{\displaystyle \frac{d^2U_1}{d\zeta ^2}}+\alpha _1U_1U_2\beta _1U_1=0,`$ (56)
$`{\displaystyle \frac{d^2U_2}{d\zeta ^2}}+\alpha _2U_1^2\beta _2U_2=0,`$ (57)
where $`\alpha _1=2\mathrm{\Delta }_1/(\mathrm{\Gamma }_1k^2),\alpha _2=2\mathrm{\Delta }_2/(\mathrm{\Gamma }_2k^2),\beta _1=2(\mathrm{\Omega }_1v_1k\frac{1}{2}\mathrm{\Gamma }_1k_1^2)/(\mathrm{\Gamma }_1k^2),\beta _2=2(\mathrm{\Omega }_2v_2k\frac{1}{2}\mathrm{\Gamma }_2k_2^2)/(\mathrm{\Gamma }_2k^2)`$, $`\mathrm{\Omega }=v_1k+\mathrm{\Gamma }_1kk_1`$, with $`k_2=2k_1,\mathrm{\Omega }_2=2\mathrm{\Omega }_1+\delta \omega `$ and $`k_1=(v_2v_1)/(\mathrm{\Gamma }_12\mathrm{\Gamma }_2)`$. One of the coupled soliton-soliton (i. e. simultaneous solitons for two wave components) solutions of Eqs.(47) and (48) reads
$`U_1={\displaystyle \frac{6}{\sqrt{\alpha _1\alpha _2}}}\left({\displaystyle \frac{2}{3}}\mathrm{sech}^2\zeta \right),`$ (58)
$`U_2={\displaystyle \frac{6}{\alpha _1}}\left({\displaystyle \frac{2}{3}}\mathrm{sech}^2\zeta \right),`$ (59)
where a condition $`\beta _1=\beta _2=4`$ is required. The parameter $`k`$ is given by
$$k=\frac{2(v_2v_1)k_1+(\mathrm{\Gamma }_12\mathrm{\Gamma }_2)k_1^2+\delta \omega }{2(\mathrm{\Gamma }_22\mathrm{\Gamma }_1)}.$$
(60)
The lattice configuration in this case takes the form
$`w_n(t)=(1)^n{\displaystyle \frac{12}{\sqrt{\alpha _1\alpha _2}}}\left[{\displaystyle \frac{2}{3}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\right]\mathrm{cos}[k_1na(\omega _1+\mathrm{\Omega }_1)t]`$ (61)
$`{\displaystyle \frac{12}{\alpha _1}}\left[{\displaystyle \frac{2}{3}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\right]\mathrm{cos}[k_2na(\omega _2+\mathrm{\Omega }_2)t],`$ (62)
$`v_n(t)=(1)^n{\displaystyle \frac{12}{\sqrt{\alpha _1\alpha _2}}}{\displaystyle \frac{I_2I_2^{}}{\omega _1^2+I_2+I_2^{}}}\left[{\displaystyle \frac{2}{3}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\right]\mathrm{cos}[k_1na(\omega _1+\mathrm{\Omega }_1)t]`$ (63)
$`{\displaystyle \frac{12}{\alpha _1}}{\displaystyle \frac{I_2+I_2^{}}{\omega _2^2+I_2+I_2^{}}}\left[{\displaystyle \frac{2}{3}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\right]\mathrm{cos}[k_2na(\omega _2+\mathrm{\Omega }_2)t].`$ (64)
If $`q_1(q_2)`$ is exactly equal to $`\pi /a`$ (zero) but with $`\delta \omega 0`$, one has $`v_1=v_2=0`$. In this case $`k_1=k_2=0`$, $`\mathrm{\Omega }_1=2\mathrm{\Gamma }_1k^2,\mathrm{\Omega }_2=2\mathrm{\Gamma }_2k^2,\mathrm{\Omega }=0`$ and $`k=\{\delta \omega /[2(\mathrm{\Gamma }_22\mathrm{\Gamma }_1)]\}^{1/2}`$. (52) and (53) present a nonpropagating simulton excitation, in which the vibrating frequency of the acoustic- (optical-) mode component being within the acoustic(optical) phonon band. In our model, $`\mathrm{\Gamma }_22\mathrm{\Gamma }_1>0`$ thus $`\delta \omega `$ should be taken positive in this case. In addition, from (52) and (53) we see that the envelopes for both the acoustic and optical components are kinks (or dark solitons). Futhermore, if $`K_2^{}=K_2`$, the displacement of light particles, $`v_n(t)`$, only has an optical-mode component.
The other simulton solution of Eqs.(47) and (48) reads
$`U_1={\displaystyle \frac{6}{\sqrt{\alpha _1\alpha _2}}}\mathrm{sech}^2\zeta ,`$ (65)
$`U_2={\displaystyle \frac{6}{\alpha _1}}\left({\displaystyle \frac{4}{3}}\mathrm{sech}^2\zeta \right),`$ (66)
where we have $`\beta _1=\beta _2=4`$. The parameter $`k`$ now reads
$$k=\frac{2(v_2v_1)k_1+2(\mathrm{\Gamma }_2\mathrm{\Gamma }_1)k_1^2\delta \omega }{2(2\mathrm{\Gamma }_1+\mathrm{\Gamma }_2)}.$$
(67)
The lattice configuration is now given by
$`w_n(t)=(1)^{n+1}{\displaystyle \frac{12}{\sqrt{\alpha _1\alpha _2}}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\mathrm{cos}[k_1na(\omega _1+\mathrm{\Omega }_1)t]`$ (68)
$`{\displaystyle \frac{12}{\alpha _1}}\left({\displaystyle \frac{4}{3}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\right)\mathrm{cos}[k_2na(\omega _2+\mathrm{\Omega }_2)t],`$ (69)
$`v_n(t)=(1)^{n+1}{\displaystyle \frac{12}{\sqrt{\alpha _1\alpha _2}}}{\displaystyle \frac{I_2I_2^{}}{\omega _1^2+I_2+I_2^{}}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\mathrm{cos}[k_1na(\omega _1+\mathrm{\Omega }_1)t]`$ (70)
$`{\displaystyle \frac{12}{\alpha _1}}{\displaystyle \frac{I_2+I_2^{}}{\omega _2^2+I_2+I_2^{}}}\left[{\displaystyle \frac{4}{3}}\mathrm{sech}^2(kna\mathrm{\Omega }t)\right]\mathrm{cos}[k_2na(\omega _2+\mathrm{\Omega }_2)t].`$ (71)
Thus in this case the acoustic-mode component is a staggered envelope soliton but the optical-mode component is still an envelope kink. If $`v_1=v_2=0`$ we have $`k_1=k_2=0,\mathrm{\Omega }_1=2\mathrm{\Gamma }_1k^2,\mathrm{\Omega }_2=2\mathrm{\Gamma }_2k^2,\mathrm{\Omega }=0`$ and $`k=\{\delta \omega /[2(\mathrm{\Gamma }_22\mathrm{\Gamma }_1)]\}^{1/2}`$. In this situation the simulton (57) and (58) is also a nonpropagating excitation with the vibrating frequency of the acoustic- (optical-) mode component within(above) the acoustic(optical) phonon band. In order to make $`k`$ to be real we should take $`\delta \omega <0`$ in this case.
A common requirement for the existence of the simulton solutions (52), (53), (57) and (58) is sgn($`\alpha _1\alpha _2)>0`$, which means sgn($`\mathrm{\Delta }_1\mathrm{\Delta }_2)>0`$ because $`\mathrm{\Gamma }_1\mathrm{\Gamma }_2>0`$ in our model. It can be met by chosing different values of system parameters. For example, in the following two particular cases we have sgn($`\mathrm{\Delta }_1\mathrm{\Delta }_2)>0`$:
1. $`K_2^{}=K_2,K_3^{}=K_3=0`$. In this case $`\mathrm{\Delta }_1=\alpha _M/\omega _1,\mathrm{\Delta }_2=J_2\alpha _M/[2\omega _2(I_2+J_2)]`$.
2. $`K_2^{}=K_2,V_3^{}=V_3=0`$. In this case $`\mathrm{\Delta }_1=(J_3^{}J_3)(1+I_2/J_2)/\omega _1`$, $`\mathrm{\Delta }_2=(I_3^{}I_3+J_3^{}J_3)/[2\omega _2(I_2+J_2)]`$.
## IV DISCUSSION AND SUMMARY
We have analytically shown that the lattice simultons are possible in nonlinear diatomic lattices. Based on the QDA for the nonlinear excitations in diatomic lattices developed before, we have considered the resonant coupling between two phonon modes, one from the acoustic and other one from the optical branch, respectively. The KS equations are derived for the envelopes of the fundamental and second harmonic waves by taking new multiple spatial-temporal scale variables, which are necessary for narrower nonlinear excitations. Exact coupled soliton (simulton) solutions are obtained for the KS equations and the simulton configurations for the lattice displacements are explicitly given.
Similar to the optical simultons in nonlinear optical media, the physical mechanism for the formation of the lattice simultons is due to the cascading effect between two lattice wave components. In this process, the fundamental and the second harmonic waves interact with themselves through repeated wave-wave interactions. For instance the energy of the fundamental wave is first upconverted to the second harmonic wave and then downconverted again, resulting in a mutual self-trapping of each wave thus the formation of two simultaneous solitons.
Mathematically, in addition to the resonance conditions (5) and (6), the formation of a lattice simulton needs a balance between the cubic nonlinearity (in the Hamiltonian) and the dispersion, the latter is provided by the discreteness of the system. Thus for deriving the envelope equations in this case, we must chose the multiple-scale variables different from the ones used in usual SHG. In our derivation for the KS equations based on the QDA, only one small parameter, i. e. the amplitude of the excitation, is used. This method gives a clear, justified and self-consistent hierarchy of scales and thus the corresponding solvability conditions, which are just the envelope equations we need. Thus it is satisfactory according to the point of view of singular perturbation theory.
Cubic nonlinearity exists in most of realistic atomic potentials. Thus it is possible to observe the lattice simultons reported here. It must be emphasized that the multi-value property of the linear dispersion relation is important for generating the simultons in lattices. Thus a diatomic or multi-atomic lattice is necessary for observing such excitations.
The theory given above can be applied to multi-atomic and higher-dimensional lattices, and higher-order nonlinearity can also be included. For instance, if we consider the Hamiltonian with cubic and quartic nonlinearities, Eqs.(31) and (32) sould be generalized to
$`i{\displaystyle \frac{A_1}{\tau }}+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_1{\displaystyle \frac{^2A_1}{\xi _n^2}}+\mathrm{\Delta }_1A_1^{}A_2+(\mathrm{\Lambda }_{11}|A_1|^2+\mathrm{\Lambda }_{12}|A_2|^2)A_1=0,`$ (72)
$`i{\displaystyle \frac{A_1}{\tau }}+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_2{\displaystyle \frac{^2A_2}{\xi _n^2}}+\mathrm{\Delta }_2A_1^2+(\mathrm{\Lambda }_{21}|A_1|^2+\mathrm{\Lambda }_{22}|A_2|^2)A_2=0,`$ (73)
where $`\mathrm{\Lambda }_{ij}(i,j=1,2)`$ are self-phase and cross-phase modulational coefficients contributed by the quartic nonlinearity of the system. Eqs.(59) and (60) can be derived using the multiple-scale variables $`\xi _n=ϵx_n,\tau =ϵ^2t`$ under the assumption $`v_n(t)=O(ϵ),w_n(t)=O(ϵ)`$, $`K_3=O(K_3^{})=O(ϵ)`$, and $`V_3=O(V_3^{})=O(ϵ)`$. A small frequency mismatch can also be included in (59) and (60) and similar equations like (43) and (44) with additional self- and cross-phase modulational terms can also be written down. A detailed study will be presented elsewhere.
## ACKNOWLEDGMENTS
This work is supported in part by grants from the Hong Kong Research Grants Council(RGC), the Hong Kong Baptist University Faculty Research Grant(FRG), the Natural Science Foundation of China, and the Development Foundation for Science and Technology of ECNU.
|
warning/0006/hep-ex0006010.html
|
ar5iv
|
text
|
# New Results on Vector Meson Production at HERA
## 1 Introduction
Production of vector mesons at HERA has become a rich field of experimental and theoretical research. In this report only fairly new results will be shown concentrating on $`J/\psi `$ and $`\varphi `$ meson. For an overview of data on $`\rho `$ mesons see $`^\mathrm{?}`$. Vector mesons can be produced elastically (i.e. exlusively), i.e. in the reaction $`epeVp`$, where $`V`$ denotes the vector meson and the proton remains a proton. The largest background is the proton dissociative process where the proton breaks up into a low mass system. H1 and ZEUS have developed efficient procedures to remove and correct for this and other backgrounds $`^{\mathrm{?},\mathrm{?},\mathrm{?}}`$.
There is by now no doubt that pQCD can describe the elastic production of vector mesons $`^{\mathrm{?},\mathrm{?}}`$ via exchange of a colour neutral system of gluons, i.e. in leading order via exchange of two gluons. The QCD scale can be given by the mass of the vector meson, as in the case of the $`J/\psi `$ or by $`Q^2`$ as for the “light” vector mesons $`\rho ,\omega `$ and $`\varphi `$. Whether the momentum transfer $`t`$ between incoming and outgoing proton (or the outgoing dissociated proton), can also serve as a hard scale is still under experimental investigation $`^\mathrm{?}`$. A signature for the “hard” behaviour is the fast rise of the integrated $`\gamma p`$ cross section for vector meson production with $`W_{\gamma p}`$.
## 2 Total Cross Sections for $`𝑱\mathbf{/}𝝍`$ and $`\mathit{\varphi }`$ Mesons
Photoproduction of $`J/\psi `$ mesons has been measured by the H1 collaboration $`^\mathrm{?}`$ in the range of $`26W_{\gamma p}285\text{GeV}`$. Leptonic decays into $`e^+e^{}`$ or $`\mu ^+\mu ^{}`$ are used depending on the detector region. The data are shown in Fig. 1 and are compared with results from pQCD calculations by Frankfurt et al. $`^\mathrm{?}`$ using various gluon density functions. The main prediction concerns the slope of the data which is seen to be well represented by the calculation using the CTEQ4M or MRSR2 gluon densities. A fit to the data of the form $`W_{\gamma p}^\delta `$ yields $`\delta =0.83\pm 0.07`$ which is much larger than $`\delta 0.3`$ found in soft processes.
New data on the production of $`\varphi `$ mesons at $`Q^2\text{ }>1\text{ }\text{GeV}\text{2}`$ became available from H1 $`^\mathrm{?}`$ and ZEUS $`^{\mathrm{?},\mathrm{?}}`$ both using the decay to $`K^+K^{}`$. ZEUS also shows first results on $`\omega `$ production in a range $`3<Q^2<20\text{ }\text{GeV}\text{2}`$ using the decay $`\omega \pi ^+\pi ^{}\pi ^0`$. The integrated cross sections $`\sigma (\gamma ^{}p\varphi p)`$ and $`\sigma (\gamma ^{}p\omega p)`$ are shown in Fig. 1. Parameterising them as $`W_{\gamma p}^\delta `$ a clear increase of $`\delta `$ with $`Q^2`$ can be seen as has been found for the $`\rho `$ meson $`^\mathrm{?}`$ .
We conclude: if the fast increase of $`\sigma (\gamma pVp)`$ with $`W_{\gamma p}`$ is indeed a signature for a hard process, either $`M_V^2`$ or $`Q^2`$ serve as hard scales.
## 3 Determination of the Regge Trajectory for $`𝑱\mathbf{/}𝝍`$ Photoproduction
In order to determine the properties of the exchange mediating the interaction between the $`J/\psi `$ meson and the proton, Regge language is used. The Regge trajectory is determined by H1 for photoproduction of $`J/\psi `$ mesons assuming a simple linear form $`\alpha (t)=\alpha _0+\alpha ^{}t`$. $`\alpha (t)`$ determines the dependence of the cross sections on the energy $`W_{\gamma p}`$ as $`W_{\gamma p}^{4(\alpha (t)1)}`$. In “hard” interactions $`\alpha ^{}`$ is expected to be small $`^\mathrm{?}`$ while in “soft” reactions $`\alpha ^{}0.25\text{ }\text{GeV}\text{-2}`$ has been found $`^\mathrm{?}`$.
H1 determined the trajectory for $`J/\psi `$ production using data from one experiment only and thus avoiding normalisation problems between data from different experiments. The data in the range $`40W_{\gamma p}150`$ GeV are used. $`d\sigma /dt`$ is determined at 5 values of $`t`$. For these 5 values of $`t`$ the dependence of $`d\sigma /dt`$ on $`W_{\gamma p}`$ is shown in Fig. 3 (left). A fit to the form $`W_{\gamma p}^{4(\alpha (t)1)}`$ is performed and the resulting values for $`\alpha (t)`$ are shown in Fig. 3 (right). A linear fit to $`\alpha (t)`$ yields the parameters of the fitted trajectory which are also given in the figure. The 1$`\sigma `$ contour is shown as are results of various predictions for the pomeron based on non-perturbative models and on a NLO BFKL calculation $`^\mathrm{?}`$. In contrast to the results for the $`\rho `$ and $`\varphi `$ mesons $`\alpha _0`$ for the $`J/\psi `$ is larger and $`\alpha ^{}`$ is compatible with 0.
## 4 The Scale for Production of Vector Mesons
When $`Q^2`$ increases the exchanged photon acquires a longitudinal component. $`R=\sigma _L/\sigma _T`$ has previously been measured for the $`\rho `$ and $`J/\psi `$ meson $`^{\mathrm{?},\mathrm{?}}`$. In Fig. 3c the ratio $`R`$ is shown for the $`\varphi `$ meson as a function of $`Q^2`$ and is seen to increase. The curve based on pQCD and parton hadron duality $`^\mathrm{?}`$ gives a good description of the data.
If in elastic vector meson production the photon couples indeed to quarks (as opposed to coupling to a hadron) and the interaction of the quark pair of the vector meson with the proton is universal the ratio of cross sections should only depend on the quark content of the vector mesons. This is indeed observed in Fig. 3a,b where $`\sigma _\varphi /\sigma _\rho `$ and $`\sigma _{J/\psi }/\sigma _\rho `$ are plotted as functions of $`Q^2`$. The ratio is seen to approach the value of $`2/9`$ for the $`\varphi `$ expected from the simple SU(4) quark counting. The ratio $`\sigma _{J/\psi }/\sigma _\rho `$ increases slowly and is still below the expected value of $`8/9`$ at $`Q^240\text{ }\text{GeV}\text{2}`$, but the errors are still large. A universal behaviour of the cross section for all vector mesons is observed in Fig. 3d, where all available data, for $`\rho ,\omega ,\varphi `$ and $`J/\psi `$ mesons are shown as a function of $`Q^2+M_V^2`$. The data have been scaled to a common $`W_{\gamma p}=75`$ GeV using the measured $`W_{\gamma p}`$ dependence. They have been scaled by the quark charges assuming the SU(4) ratio $`\rho :\omega :\varphi :J/\psi =9:1:2:8`$. The data are seen to agree well with each other and can be described by a function $`(Q^2+M_V^2+a)^b`$ with $`a=0.42\pm 0.09\text{ }\text{GeV}\text{2}`$ and $`b=2.37\pm 0.10`$, which was fitted to the $`\rho `$ data. We can conclude that within present errors $`(Q^2+M_V^2)`$ is a good scale for photoproduction of vector mesons at low values of $`t`$.
## Acknowledgments
I thank the organisers for an interesting meeting and my H1 and ZEUS colleagues for discussions. Skiing after 10 years was fun!
## References
|
warning/0006/astro-ph0006334.html
|
ar5iv
|
text
|
# Sloshing in High Speed Galaxy Interactions
## 1 INTRODUCTION
Several groups have reported observations of lopsided H I (Baldwin, Lynden-Bell, & Sancisi, 1980; Richter & Sancisi, 1994) and stellar distributions (Lauer et al., 1993, 1996; Rix & Zaritsky, 1995; Davidge et al., 1997; Zaritsky & Rix, 1997; Conselice, 1997) in a significant fraction of spiral galaxies. The percentage of galaxies that are measurably lopsided range from 30% of field spirals (Zaritsky & Rix, 1997; Kornreich et al., 1998) to 50% of all spiral galaxies (Swaters et al., 1999). These numbers may be high if, as proposed by Zaritsky & Rix (1997), lopsided galaxies are brighter than symmetric galaxies, thereby biasing the statistics. Nevertheless, a significant number of lopsided galaxies are observed. These findings have prompted a number of papers exploring the origins and dynamics of these asymmetries.
Baldwin, Lynden-Bell, & Sancisi (1980) proposed a simple kinematic model for lopsided galaxies in which different rings of the galaxy, assumed non-interacting, are initially shifted from their centered equilibrium positions. The shifted rings precess in the overall gravitational potential in a direction opposite to that of the mass motion. Because the precession rate decreases in general with radial distance, an initial disturbance tends to “wind up” into a leading spiral arm in a time appreciably less than the Hubble time. Miller & Smith (1988, 1992) have made extensive computer simulation studies of the unstable eccentric motion of matter in the nuclei of galaxies, which they suggest is pertinent to the off-center nuclei observed in a number of galaxies such as M31, M33, and M101.
More recent ideas on the origin of lopsidedness include galaxy mergers (Zaritsky & Rix, 1997), and counter-rotating matter (Sellwood & Merritt, 1994). Schoenmakers, Franx, & de Zeeuw (1997) interpret optical asymmetry as an indicator of asymmetry in the overall galactic potential and therefore an indicator of the spatial distribution of the dark matter in a galaxy, which may have a triaxial distribution. By analyzing the spiral components present in the surface brightness or H I distribution, the velocity gradient, and therefore, the shape of the gravitational potential, may be uncovered. Jog (1997) studied the orbits of stars and gas in lopsided potentials and concluded that a stationary lopsided disk may indicate asymmetry in the halo.
A further possibility is that an optical disk may be in a quasi-stationary lopsided state in a symmetric potential, as discussed by Syer & Tremaine (1996). In this model, gaseous and stellar matter that is not fully relaxed swirl about the minimum of the halo potential. The result is a lopsided flow within a symmetric mass distribution. Numerical simulations of this situation have been done by Levine & Sparke (1998) using a gravitational $`N`$-body tree-code method for disk galaxies shifted from the center of the main halo potential. The results are suggestive of lopsidedness with large lifetimes.
An $`N`$-body simulation study of a rotating spheroidal stellar system including the dynamics of a massive central object by Taga & Iye (1998a) indicates that the central object goes into a long lasting oscillation similar to those found earlier by Miller & Smith (1988, 1992) and which may explain asymmetric structures observed in M31 and NGC 4486B. A linear stability analysis of a self-gravitating fluid disk with a massive central object also by Taga & Iye (1998b) indicates a linear instability (Taga & Iye, 1998b). A study of the eccentric motion of a disk galaxy made up of a large number of self-gravitating rings shows instability in the inner part of the disk ($`<2`$ kpc) and the excitation of long lasting trailing one armed spiral waves (Lovelace et al., 1999).
In this paper we explore two issues surrounding lopsided galaxies. First we examine the gravitational response of an $`N`$-body disk galaxy model embedded in a halo to the high speed passage of a companion galaxy. Then, assuming that a disk galaxy has been perturbed by a passing galaxy or other mechanism, we explore the response of the system to a variety of displacements and relative velocities of a disk of stars offset from their halo centers of mass. In some of these runs, we have both the stars and halo free to move, while the halo is fixed in the remaining runs. We report on runs with different initial conditions and different mass distributions.
We use our two dimensional $`N`$-body GALAXY code (Schroeder & Comins, 1989; Schroeder, 1989; Shorey, 1996), with 100,000 collisionless particles representing stars (or perhaps more properly, galactic clusters) and a gravitating halo that comprises 75% of the system’s total mass. One of our goals is to determine how perturbations, such as the passage of a companion galaxy or offset halo-disk centers of mass, affect the formation and evolution of bars in galaxies. A sufficiently massive halo will quickly dampen almost any bar (Binney & Tremaine, 1987; Sellwood, 1983). We have therefore chosen this $`75\%`$ halo because it has insufficient mass to suppress the development of a bar that may be driven into existence during the run (Comins et al., 1997). The halos we use are sufficiently massive, however, to simulate the gravitational effect of a significant dark matter galactic component. We therefore expect that the suppression of any bar that develops and decays in our runs would be due to the dynamics that the disk undergoes.
Both the creation and the destruction of bars have been discussed by several groups. Noguchi (1996) proposed that bars in late-type galaxies have intrinsic origins, while bars in early-type galaxies have been formed by tidal interactions with other galaxies. Specifically, Miwa & Noguchi (1998) found that tidal effects due to a close encounter between galaxies may trigger the formation of a bar in an otherwise barstable disk. Athanassoula (1996a, b, 1999) performed $`N`$-body simulations of the interactions of a barred disk galaxy with a close companion galaxy. She found that the bar may be destroyed as a consequence of this interaction, or that an offcentered bar may form. The destruction of the bar occurs most likely in the case of a merger, when most of the companion reaches the target’s center.
The initial particle velocity distribution in each run is determined by $`Q_0`$, the initial value of Toomre’s $`Q`$. We chose to set $`Q_0`$ to a marginally stable $`Q_0=1.1`$ over the entire disk. The offset runs lasted just under 10 rotation periods, as measured at the half-mass radius of the disk. The particles initially form a counter-clockwise rotating disk.
The simulation is done on a Cartesian grid with $`256\times 256`$ cells. During the runs with the intruder present, the test galaxy and its halo had a diameter of 128 cells. To maximize use of computer memory and time, the intruder galaxy was a point particle, sent into a coplanar orbit. The perigalactic distance between the two galaxies was chosen so that they would not overlap even if they both had realistic radii. This precludes any dynamical friction between them. Their minimum separation was 42 cell widths and the intruder was on the grid for less than one rotation period of the test galaxy.
During the runs without the intruder, the test galaxy’s initial diameter is enlarged to 192 cells. This enables us to more accurately simulate small-scale features. For a typical disk galaxy of radius 10 kpc (Binney & Tremaine, 1987) this translates into a cell width of roughly 104 pc. Each rotation period of the disk corresponds to roughly 250 Myr.
Runs were made with three different halo mass distributions:
1. The distribution, as described by Sellwood & Carlberg (1984) and Carlberg & Freedman (1985), that creates a rotation curve simulating that of an Sc galaxy.
2. The isothermal distribution described in Binney & Tremaine (1987).
3. An isothermal distribution plus the gravitational potential of a supermassive central black hole.
The rotation curves for these potentials are all presented together in Figure Sloshing in High Speed Galaxy Interactions.
Tables 1, 2, and 3 list the basic properties and results for these runs. Column 1 in each of these tables lists the run numbers referred to throughout the paper. All runs used an Sc distribution for the disk.
In $`\mathrm{\S }2`$ we present the results of 5 runs in which an intruder galaxy passes the target galaxy. In $`\mathrm{\S }3`$ we present the results of 12 runs in which the stars and halo initially have different centers of mass, along with three reference runs with the same mass distributions and no initial offsets. In $`\mathrm{\S }4`$ we compare these runs to each other, to previous simulations, and to observations that they may explain. In $`\mathrm{\S }5`$ we present our conclusions.
## 2 EFFECTS OF A HIGH SPEED GALACTIC COLLISION
Disk galaxies interacting with neighboring galaxies have lopsided structures as seen, for example, in Karachentsev 64 and M101. Unlike simulations of slow-speed interactions that lead to mergers (Hernquist & Mihos, 1995; Zaritsky & Rix, 1997), we send a simulated companion in a high speed, slightly positive-energy, coplanar orbit that takes it past the target galaxy and off the grid. The intruder is modeled as a point mass with 20% the total (halo plus disk) mass of the target galaxy. It is initially sent onto the grid passing the target galaxy in the same direction as the target galaxy’s star particles orbit. The intruder enters at an angle adjusted so that, even if it did have a realistic distribution of stars, it would not collide with the target galaxy. This prevents the merger of the two galaxies, which we do not seek to simulate.
The first two intruder runs had the halo simulated by a second $`N`$-Body component of 10000 particles with a different $`Q`$ than the star particles. Therefore, both the star and halo particles responded to the gravitational influence of the intruder, as well as to each other. (Halos that move we denote as dynamic, while fixed halos are called static.) One of these runs (hereafter Run 1) simulated a dynamic halo of cold matter, with an initial $`Q_{0,halo}=0.3`$. The other run (Run 2) simulated a dynamic halo of hot matter, with $`Q_{0,halo}=5.0`$.
Figure Sloshing in High Speed Galaxy Interactions shows the displacement of the center of mass motion of the stars relative to the center of mass of the halo in Run 1. The halo’s center of mass is always located at (0.0,0.0) in this, and similar, figures. The maximum separation between the halo and star centers of mass is three quarters of a cell width (78 pc). Although “noisy”, Figure Sloshing in High Speed Galaxy Interactions shows distinct periodic, cyclic motion as the two centers of mass waltz around each other. The dynamics of this run are complicated by the fact that the cold halo particles develop a bar, which persists throughout the $`6\frac{1}{4}`$ rotation periods of the run.
Figure Sloshing in High Speed Galaxy Interactions shows the separation in Run 2 between the stellar and halo centers of mass as. This is the hot halo case. Due to the greater kinetic energy in its particles, we expect that the halo center of mass will be displaced less than the stars in this run. This is what we observe. As a result, the maximum separation is nearly two grid cells or 208 pc. No bar develops in this halo.
Next we ran three intruder runs with a hydrodynamic simulation of the halo. These were done to avoid the effects of dynamic friction that were observed to occur between the star and halo particles (each with initially different $`Q_0`$) in Runs 1 and 2. We used an Sc mass distribution of compressible, shockable, gravitating gas. Instead of assigning an initial $`Q`$, we give the gas an initial uniform energy density. This is equivalent to giving the gas a temperature profile.
Figure Sloshing in High Speed Galaxy Interactions shows the center of mass separations between the halo and stars for the case with a low initial energy density gas corresponding to an initial gas temperature of $`T_0=0.5`$ K as measured at the half mass radius. During the run, denoted Run 3, the maximum separation between the halo and star centers of mass reached nearly 3 cell widths or 312 pc. The result for the $`T_0=5.0`$ K run (Run 4) is similar with a maximum separation of nearly 6 cell widths.
The results of the hottest gas run (Run 5), with $`T_0=5\times 10^5`$ K, are especially intriguing. Figures Sloshing in High Speed Galaxy Interactions and Sloshing in High Speed Galaxy Interactions show the gas and star distributions throughout this run. Note that in this and all other intruder runs, the intruder travels counterclockwise, entering the grid at roughly the 5 o’clock position and leaves at roughly the 1 o’clock position at time $`t=0.98`$ rotation periods. After leaving the grid, the intruder’s gravitational potential is ignored. In response to the intruder, a bulge is clearly visible at the three o’clock position in both the gas and stars at time $`0.861`$ rotation periods. Note the similarity between timestep 4.898 in Figures Sloshing in High Speed Galaxy Interactions and Sloshing in High Speed Galaxy Interactions and the optical image of Karachentsev 64 (Figure Sloshing in High Speed Galaxy Interactions). Karachentsev 64 is a pair of interacting spiral galaxies.
The centers of mass of both components drift upward throughout this run. This is due to the asymmetry created in the spiral arms by the intruder’s passage. The more open arm, on the left, carries angular momentum downward and to conserve it, the rest of the system moves upward. The relative displacement between the centers of mass in this run is shown in Figure Sloshing in High Speed Galaxy Interactions. While the dynamics of the system greatly complicate the relative center of mass motion, clearly it still exists.
Because of the differences in behaviors of particles and gas acting as the halo, these three gas halo runs cannot be compared directly with the previous two $`N`$-body halo runs. Nevertheless, the behavior of all five consistently show an offset between the two centers of mass. Furthermore, the “hotter” the halo, the greater the offset.
In all of our runs, the center of mass of the stellar disk moves relative to the center of mass of its halo as a result of the gravitational interaction with an intruder. This provides the motivation to explore the dynamics of visible matter sloshing around in a deep gravitational halo. The difficulties with doing this following the perturbation of an intruder, as in Runs 1 through 5, are two-fold. First, as we saw in Run 5, the transients created by the intruder’s extremely powerful and asymmetric passage make it difficult to see how the star system is responding just to its halo’s potential. Second, there may be other mechanisms that cause a separation of centers of mass that are not so disturbing to the system. Therefore, we now look at systems with initial offsets between the star and halo centers of mass, but without any intruder to mix up the system.
## 3 OFFSET RUNS
### 3.1 Dynamic Halo Runs
We proceed now by displacing a variety of stellar disks from the centers of mass of their halos. Our first four offset runs, labeled Runs 6 – 9, have both stars and halo free to move. Predictably, the two runs with dynamic $`N`$-body halos show significant dynamical friction between the stars and halo particles. These lead to rapid damping of the center of mass motion. In the low angular momentum Run 6, the stars and halo have essentially the same centers of mass and are in equilibrium configurations within less than four rotation periods. The relative positions of the centers of mass show little, if any, periodic motion during the run (Figure Sloshing in High Speed Galaxy Interactions).
Transient structure occurs in the stars in Run 6. They evolve a multiarmed spiral psttern which changes to a two-armed spiral and then to a symmetric disk by the beginning of the fourth rotation period. This is also a typical evolution of a non-offset $`N`$-body system moving in the potential of a fixed halo. In comparison, the transient structures seen in the high angular momentum version, Run 7, shows a significant feathery one-armed spiral structure that evolves to a small bar with a ring surrounding it and then to a symmetric system, after less than four rotation periods (Figure Sloshing in High Speed Galaxy Interactions). Similar one-arm features have been reported by Lotan-Luban (1990), as mentioned in Struck (1999), in head-on, low impact parameter collisions. The center of mass of the stars in our Run 7 follow an arc inward, dampened after 1.8 rotation periods by the dynamical friction, again with little indication of periodic motion relative to the halo’s center of mass.
If, as seems likely, most of the true halo mass is a smoother distribution of matter, less prone to the effects of dynamical friction, it would be more appropriate to simulate it’s dynamic behavior by that of a gravitating fluid. Therefore, we made two runs with our gravitating gas component serving as the halo, as in Runs 3 – 5, above. Run 8 has the same initial conditions as Run 6: hot halo, large initial center of mass offset, and small initial angular momentum for the stars.
Both the halo gas and the stars in Run 8 go through non-axisymmetric transients. After about $`1\frac{3}{4}`$ rotation periods, the gas settles into a two-armed spiral structure. The stars develop a distinct bar with spiral arms winding out from its ends. The spiral structure is much more distinct in the gas than in the star particle component (Figures 11a and 11b).
Unlike the runs with dynamical friction, the center of mass of the stars for the present run show distinct orbital motion around the center of halo mass (Figure 12a). The halo’s center of mass moved less than a grid cell width during any fifty timesteps, during which the star’s center of mass moved up to three cell widths (Figure 12b). This is important in our justifying shortly the value of examining runs with fixed halos.
Run 9 is the high angular momentum version of Run 8. In all other respects the initial conditions of the two runs are the same. The qualitative evolution of the gas is similar in both, although the concentration of gas in the center of the system is lower in the higher angular momentum case. As a result, the halo in Run 9 remains axisymmetric over a larger radius. The initial perturbation on the stars due to the halo is lower in this run than in Run 8, because the stars are not plunging toward the center of the halo’s mass distribution. The stars develop transient arms, but the bar is much less concentrated than in Run 8. Indeed, the central region of the stars remains axisymmetric (Figure 13b). The higher angular momentum of the stars gives the halo more time to move during each timestep. The stellar center of mass shows a more circular orbit around the halo center of mass than in Run 8 (Figure Sloshing in High Speed Galaxy Interactions). This is consistent with the decreased perturbation caused by the stars on the halo.
The orbital motion of the stellar component’s center of mass around the halo seen in the last two runs and the associated structure in the star component suggest that such sloshing will have observable effects on galaxies. This belief is strengthened when we compare these results to a “traditional” run, in which the centers of mass of the disk and halo are initially the same and remain the same throughout the run. Labeled Run 10 in Table 3, the non-offset Sc disk shows the typical evolution seen in $`N`$-body simulations of the disk. It goes through an initial high-m (multiarmed) spiral phase, and then into a distribution with a persistent bar, without spiral arms (Figure Sloshing in High Speed Galaxy Interactions). This bar’s long axis extends one quarter of the diameter of the disk.
Run 10 shows profound differences in mass distribution compared to Run 8, which also develops a bar. While the non-offset star distribution maintains an overall disk-shaped structure, virtually all the star particles in Run 8 are either in the bar or in the two weak spiral arms, while there are many stars in an axisymmetric distribution surround the bar throughout Run 10. Furthermore, the bar in Run 8 is much thicker than the bar in Run 10. Although both runs have bars, clearly, the stars in the run with the dynamic halo (Run 8) have followed a different dynamical evolution than the run with the static halo.
The underlying spiral structure created in the halo of these runs is of concern, as there is no observational evidence yet for such structure. To eliminate any nonaxisymmetric effects from a halo, we have made one final set of runs in which the halo remains fixed and axisymmetric. This construct will affect the timescales for the motion of the stellar center of mass around the halo center of mass. However, as noted above, the relatively small motions of the halo in the low angular momentum situation suggest that this effect will be relatively small for these runs. Indeed, since the halos are expected to be even higher than the 75% of the total galactic mass we have used here, the effect of halo motion will be even less than Runs 8 and 9 indicate. Furthermore, the savings in computer time enabled us to make many more runs with fixed halos than we could have otherwise accomplished.
Runs 11 through 16, 18, and 20 (see Table 3) have displaced stellar centers of mass in fixed halo mass distributions. (Runs 10, 17, and 19 are undisplaced reference runs.) As seen in Table 3, the four parameters we studied can be summarized as follows:
* Initial Offset (Column 2): The initial offset of the disk’s center of mass is either $`2\sqrt{(}2)`$ cells (hereafter, “small offset”) or $`10\sqrt{(}2)`$ cells (“large offset”) from the halo’s (fixed) center of mass. The small offset runs correspond to 3% of the disk’s radius. A grid cell corresponds to roughly 104 pc.
* Angular Momentum (Column 3): The disk’s center of mass initially has either a low angular momentum or a high angular momentum around the halo center of mass. The high angular momentum runs’ angular momenta are 6% higher than the low angular momentum runs.
* Mass Distribution (Column 4): The halo has a variety of radial mass distributions, as described above. The disk is always an Sc distribution.
* Counter-rotating Component (Column 5): In some of the runs we have set 50% of the star particles in counter-rotating motion.
### 3.2 Sc Halo Mass Distribution
Run 11 is a small offset, low center of mass angular momentum run. This angular momentum is in the same direction as the stellar angular momentum. As with all the off-center runs reported in this paper, this run shows an initial transient containing more spiral structure than the traditional run (Run 10). In this offset run, the transient is followed by a period of spiral structure with eight or nine arms that combine and decrease in number. This is followed by the establishment of a bar. The bar becomes robust and permanent after 4 rotation periods. The amounts of $`m=1`$ (one-armed spiral), and $`m=2`$ (bar) motion are comparable to the traditional run. This evolution is summarized in Figures Sloshing in High Speed Galaxy Interactions and Sloshing in High Speed Galaxy Interactions. The center of mass motion of this system spirals around the halo center of mass (Figure Sloshing in High Speed Galaxy Interactions), as also seen in Levine & Sparke (1998).
Run 12 has an initial offset 5 times larger than Run 11, but is otherwise the same. There is a marked transient one-armed feature that dominates the structure for about $`\frac{3}{4}`$ of a rotation period. This is followed by several rotation periods in which multiple, flocculent arms spiral out from the m=1 arm. The m=1 feature persists for the entire run, long after the higher-m structure has dissipated. By the end of 10 rotation periods, there is still no sign of a bar. The center of mass of the disk in this run spirals inward less than is seen in Run 11 (Figure Sloshing in High Speed Galaxy Interactions).
Run 13 is a high angular momentum version of Run 12. The initial one-armed spiral is 50% stronger than in Run 12. Particularly notable about this run is the formation of several very high density concentrations of $`N`$body particles (Figure Sloshing in High Speed Galaxy Interactions). These regions persist for about $`1\frac{1}{2}`$ rotation periods. The end state of the run after 10 rotation periods is qualitatively the same as in Run 12. No bar develops during this run.
The center of mass of Run 13 again spirals in toward the center of the halo mass, but the higher angular momentum of this system results in a longer time during which the two centers of mass are different. The results of this run are consistent with a similar run presented by Levine & Sparke (1998).
Run 14 is the same as Run 13 except that the initial angular momentum of the disk’s center of mass is in the opposite sense to the particles’ orbital motion. The dynamics of the resulting one-armed spiral are significantly different than in Run 13. In the present run, the spiral is stationary in angular location for nearly 1.5 rotation periods, while continually expanding radially (Figures 21a, 21b, and 21c). The arm then transforms into a wide, diffuse, leading arm spiral that persists for the remainder of the run. No bar develops during this run.
Recalling the leading arm spirals we saw develop in runs with counter-rotating disks (Comins et al., 1997), we made two Sc runs with counter-rotating disk particles. Both of these runs had low angular momentum orbits for the disk’s center of mass.
Run 15 has equal numbers of particles rotating in opposite directions, but is otherwise the same as Run 11. Each of the two counter-rotating sets of particles were separately given the same low angular momentum, in opposite directions. As a result, of course, the total angular momentum of the system was zero. The initial offset served as a perturbation that created tightly wound spirals in the inner region of the disk (Figure 22a). These evolved after $`1\frac{1}{2}`$ rotation periods into the same single-armed spiral that we saw in Comins et al. (1997). This spiral expanded radially (Figure 22b) as far out as the inner Lindblad resonance where its outer edge bifurcated into spirals like the tongue of a snake (Figure 22c). At this point the spiral arm is about to change directions. While damping thereafter, the spiral structure persisted for the remainder of the run. The center of mass of the entire disk followed a damped oscillation across the halo center of mass throughout the run.
Run 16 also had half of its particles counter-rotating, but otherwise is the same as Run 12. The center of mass of its disk is displaced 5 times farther from the halo center of mass than in Run 15. Again, the center of mass of the entire disk followed a damped oscillation across the halo center of mass, while the individual counter-rotating halves of the disk separately spiraled inward.
The combined motion created by the initial center of mass offset and the counter-rotating disks led to significantly different behavior than any other run. Within one quarter of a rotation period, the system had developed a reflection-symmetric pattern (Figure Sloshing in High Speed Galaxy Interactions). After three more rotation periods, a small central bar formed, surrounded by a ring, from which a single, tightly-wound arm spiraled. Thereafter, the bar oscillated in and out of existence, due to energy transfer from the bar to the spiral or ring structure. The spiral arm decayed over 6 rotation periods, reversing its direction several times as it decayed.
### 3.3 Isothermal Halo Mass Distribution
In order to learn how robust these results are, we then made a series of runs using different halo mass distributions, while keeping the Sc particle disk. In this section we present the runs using an isothermal halo.
Run 17 is a reference case without initial offset. Comparing the results for this case to the comparable Sc case (Run 10), we note that Run 17 develops a bar more quickly, but that this bar is less elliptical than the one in Run 10, extending only over 15% of the disk’s diameter. This bar persists for the entire run. The spiral structure in the two runs (Run 10 and Run 17) are similar in numbers of arms and evolution. While specific differences between the particle dynamics are perceptible between the two runs, they are subjectively very similar throughout.
Run 18 has a high offset disk and low angular momentum. This isothermal run is analogous to Run 12, above. The global spiral features of these two runs are essentially the same. After the transient spiral stage, however, Run 18 develops a short bar, unlike Run 12. The bar extends over 15% of the disk’s diameter by the fifth rotation period, despite the fact that its $`Q`$ is by then over 4.0, and climbing. The bar’s m=2 Fourier amplitude is 20% less than that of the bar in Run 17.
### 3.4 Isothermal Halo Mass Distribution With Galactic Black Hole
Run 19 adds the gravitational influence of a supermassive black hole at the halo’s center of mass to Run 17. The black hole’s mass is about $`4.2\times 10^8M_{}`$, which corresponds to 0.5% of the mass of our model galaxy. The disk in this run has no initial offset. The black hole has the effect of suppressing the bar, while allowing virtually the same transient spiral structure to unfold. The run ends with a high particle concentration near the center of the disk, consistent with the added gravitational potential of the black hole.
Run 20 adds a large offset and high angular momentum to Run 19. Run 20 is therefore analogous to Run 13 (Sc disk and halo). The addition of the black hole mass is significant, especially during the first three rotation periods. Specifically, we observe richer concentrations of particles in Run 20 than in any other run (Figure Sloshing in High Speed Galaxy Interactions). A bar forms in this run. Its end state is very similar to that of Run 18, differing only in the greater central concentration of particles in the run with the black hole.
## 4 ANALYSIS AND DISCUSSION
### 4.1 Theory
Lovelace et al. (1999) studied the eccentric dynamics of a disk with an exponential surface density distribution \[$`\mathrm{\Sigma }_d\mathrm{exp}(r/r_d)`$\] represented by a large number of rings and a central mass $`M_010^9M_{}`$ embedded in a passive dark matter halo. The inner part of the disk $`r2.5`$ kpc was found to be strongly unstable with $`e`$folding time $`30`$ Myr for the conditions considered. Angular momentum of the rings is transferred outward, and to the central mass if it is present. A trailing one-armed spiral wave is formed in the disk. This differs from the prediction of Baldwin, Lynden-Bell, & Sancisi (1980) of a leading one-armed spiral. The outer part of the disk $`rr_d`$ is stable and in this region the angular momentum is transported by the wave. This instability appears qualitatively similar to that found by Taga & Iye (1998b) for a fluid Kuzmin disk with surface density $`\mathrm{\Sigma }_d1/(1+r^2)^{3/2}`$ with a point mass at the center where unstable trailing one-armed spiral waves are found. The present findings do not give evidence of instability of the inner part of the disk, but they do indicate that a long-lasting trailing spiral wave is generated in the disk as found by Lovelace et al. (1999) and Taga & Iye (1998b).
### 4.2 Effects of Dynamic Halos
Particle and hydrodynamic simulations of a galaxy’s halo lead to significantly different histories for the star particles. This is due in large measure to the dynamical friction that occurs between halo and star particles. Any displacement between the particle halo and stellar centers of mass is quickly damped. The hydrodynamic halo does not damp the stellar motion nearly as much, allowing for prolonged sloshing of the stars and concomitant nonaxisymmetric evolution.
All other things being equal, lower temperature (or lower energy) halo mass leads to more halo displacement as the result of perturbation by a passing galaxy or sloshing stars. These simulations do raise the question as to whether nonaxisymmetric perturbations of the visible matter in galaxies is accompanied by similar changes in the distribution of the underlying halo.
### 4.3 Results of Perturbed Static Halo Simulations
#### 4.3.1 Effects of Stellar Disk Displacement from the Halo Center of Mass and of Disk Angular Momentum on Bar Formation
The runs presented here indicate that bar suppression can be caused by energy redistribution due to the transient motion of the disk center of mass moving toward the halo center of mass. In general, a larger offset of the galactic disk from the halo center is more effective in suppressing the formation of a central bar. The exceptions to this rule are the runs including counter-rotating components (Run 15 and Run 16), where in fact the opposite is observed: the smaller offset produces no bar, but the larger offset does. Comins et al. (1997) showed that the presence of undisplaced 50% counter-rotating components is already an effective suppressor of the formation of a central bar.
The initial offset of the stellar disk’s center of mass was modelled as having occurred in one of two ways: for some runs the stellar disk’s center of mass was given nearly zero initial velocity, causing the center of mass to have a very small angular momentum with respect to the halo’s center. Therefore the stellar disk followed a trajectory that leads nearly through the halo center. In other runs the displaced stellar disk’s center of mass was given a small “push” perpendicular to its offset. This caused the disk to follow a spiral orbit around the halo center. Our results indicate that the amount of the initial displacement of the galactic disk has a much stronger influence on the evolution of the model galaxy than the amount of its angular momentum. In fact, Runs 12 and 13 have identical initial parameters, differing only by their initial angular momentum, and they show no qualitative difference in star distribution by the end of the runs after 10 rotation periods.
#### 4.3.2 Effects of Different Halo Potentials
We found little difference in the effect on bar formation from using different halo mass distributions except for the isothermal potential without a central black hole. Here a bar forms for both a small and a large initial offset. This occurred because the potential of the isothermal mass distribution is about 25% lower in the center of the simulation galaxy, compared to the potential of the other mass distributions at the center of the galaxy. This reduced potential at the galaxy’s center is the result of our effort to fit the isothermal potential closely to the Sc potential over the widest possible range outside the center.
In the case of the isothermal potential with a central black hole, the mass of the black hole was chosen such that its gravitational potential combined with the isothermal potential of the halo produces a potential that is fitted closely to the Sc potential over the entire range of the simulation galaxy, including its center. A mass of $`4.2\times 10^8M_{}`$ for the black hole achieved this.
#### 4.3.3 Effects of Counter-rotating Stars
Comins et al. (1997) showed that the presence of counter-rotating components in the galactic disk can suppress the formation of a central bar. Those runs were performed without any offset of the stellar components, and their results were similar to our Run 15, which had a small offset and exhibited no sign of a bar at any stage of the run. In contrast to this observation, a large offset as in Run 16 produces an alternating sequence of a central bar and an axisymmetric central feature. This is due to the superposition of two motions: the rotation of the bar, and the back and forth motion of the disk’s center of mass as described above. At its greatest length, the bar extends close to the corotation resonance.
T. Z. and N. F. C. wish to thank Sun Microsystems, Inc., and EDS Data Systems for their generous donations of computers on which most of this work was done. The work of R. V. E. L. was supported in part by NSF grant AST 93-20068.
|
warning/0006/cond-mat0006101.html
|
ar5iv
|
text
|
# Abstract
## Abstract
The purpose of this article is to present the thermodynamics of the pseudospin electron (PE) model in the case of the different type interactions between pseudospins. First, we provide an overview of the results of works which deal with the theoretical investigation of the PE model with the inclusion of the direct pseudospin-pseudospin interaction (but without the electron transfer). Second, we present the results of the investigation of the model in the case of the absence of the direct pseudospin-pseudospin interaction and Hubbard correlation, when interaction between pseudospins via conducting electron is done.
## Introduction
Pseudospin-electron (PE) model is one of theoretical models which considers the interaction of electrons with local lattice vibrations where an anharmonic variables are represented by pseudospins. The theoretical investigation of the PE model is an enduring subject of interest at the quantum statistics department.
The model is used to describe the strongly correlated electrons of CuO<sub>2</sub> sheets coupled with the vibrational states of apex oxygen ions O<sub>IV</sub> (which move in the double-well potential) in YBaCuO type high-T<sub>c</sub> superconductors (HTSC) . Recently a similar model has been applied for investigation of the proton-electron interaction in molecular and crystalline systems with hydrogen bonds .
The model Hamiltonian is the following:
$`H={\displaystyle \underset{i}{}}H_i+{\displaystyle \underset{ij\sigma }{}}t_{ij}a_{i\sigma }^+a_{i\sigma }{\displaystyle \frac{1}{2}}{\displaystyle \underset{ij}{}}J_{ij}S_i^zS_j^z,`$ (1)
$`H_i=Un_in_i\mu {\displaystyle \underset{\sigma }{}}n_{i\sigma }+g{\displaystyle \underset{\sigma }{}}n_{i\sigma }S_i^zhS_i^z,`$
where the strong single-site electron correlation $`U`$, interaction with the anharmonic mode ($`g`$-term), and the energy of the anharmonic potential asymmetry ($`h`$-term) are included in the single-site part. Hamiltonian (1) also contains terms, which describe electron transfer $`t_{ij}`$ and direct interaction between pseudospins $`J_{ij}`$. The energy of the electron states at the lattice site is accounted from the level of chemical potential $`\mu `$.
In the case of strong coupling ($`gt`$) and correlation ($`Ut`$) the perturbation theory can not be applied. It is reasonable to include these one in zero order Hamiltonian (single-site Hamiltonian $`H_i`$). Its eigenfunctions are build of the vectors $`|n_i,n_i,S_i^z`$, which form the full unit cell state basis :
$`|1=|0,0,{\displaystyle \frac{1}{2}},|2=|1,1,{\displaystyle \frac{1}{2}},|3=|0,1,{\displaystyle \frac{1}{2}},|4=|1,0,{\displaystyle \frac{1}{2}},`$ (2)
$`|\stackrel{~}{1}=|0,0,{\displaystyle \frac{1}{2}},|\stackrel{~}{2}=|1,1,{\displaystyle \frac{1}{2}},|\stackrel{~}{3}=|0,1,{\displaystyle \frac{1}{2}},|\stackrel{~}{4}=|1,0,{\displaystyle \frac{1}{2}}.`$
In the early works the main attention at the investigation of this model has been paid to the examination of electron states, effective electron-electron interaction, to the elucidation of additional possibilities of occurrence of superconducting pair correlations. On the basis of PE model a possible connection between the superconductivity and lattice instability of the ferroelectric type in HTSC has been discussed . A series of works has been carried out in which the pseudospin $`S^zS^z`$, mixed $`S^zn`$ and charge $`nn`$ correlation functions were calculated. It has been shown with the use of the generalized random phase approximation (GRPA) in the limit $`U\mathrm{}`$ , that there exists a possibility of divergences of these functions at some values of temperature. This effect was interpreted as a manifestation of dielectric instability or ferroelectric type anomaly. The tendency to the spatially modulated charge and pseudospin ordering at the certain model parameter values was found out. The analysis of ferroelectric type instabilities in the two-sublattice model of high temperature superconducting systems has been made . The influence of oxygen nonstoichiometry on localization of apex oxygen in YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7-x</sub> type crystals is studied in the work .
The purpose of this article (lecture) is to present the thermodynamics of the PE model in the case of the different type interactions between pseudospins. First, we provide an overview of the results of works which deal with the theoretical investigation of the PE model with the inclusion of the direct pseudospin-pseudospin interaction (but without the electron transfer ($`t_{ij}=0`$)). Second, we present the results of the investigation of the model in the case of the absence of the direct pseudospin-pseudospin interaction and Hubbard correlation ($`J_{ij}=0`$, $`U=0`$), when interaction between pseudospins via conducting electron is done.
## 1 Direct interaction between pseudospins.
### 1.1 Ferroelectric type interaction.
The work is devoted to the study of the PE model in the case of zero electron transfer $`(t_{ij}=0)`$. The direct interaction between pseudospins is taken into account. It is supposed to be a long-ranged that allows one to use the mean field approximation (MFA). In this approximation, the model Hamiltonian has the following form:
$$H=\underset{i}{}\stackrel{~}{H}_i+\frac{N}{2}J\eta ^2,\stackrel{~}{H}_i=\mu \underset{\sigma }{}n_{i\sigma }+Un_in_i+g\underset{\sigma }{}n_{i\sigma }S_i^z(h+J\eta )S_i^z.$$
The interaction $`(J_{ij}J/N)`$ is taken as the ferroelectric type one; the order parameter $`\eta =S_i^z`$ does not depend on the unit cell index.
Grand canonical potential and partition function of the model, calculated per one lattice site are equal to
$`{\displaystyle \frac{\mathrm{\Omega }}{N}}=T\mathrm{ln}Z_i+{\displaystyle \frac{1}{2}}J\eta ^2,`$ (3)
$`Z_i=2\left[\mathrm{cosh}\beta h_n+\mathrm{e}^{\beta (U2\mu )}\mathrm{cosh}\beta (h_ng)+2\mathrm{e}^{\beta \mu }\mathrm{cosh}\beta \left(h_n{\displaystyle \frac{g}{2}}\right)\right],`$
where $`h_n=h/2+J\eta /2`$.
Then, all thermodynamic properties can be derived from the presented formulae (3).
The average number of electrons is determined as follows :
$$\frac{1}{N}\left(\frac{\mathrm{\Omega }}{\mu }\right)_{T,h,\mu }=\frac{1}{N}\underset{i}{}n_in.$$
(4)
The equation for the order parameter is obtained from the thermodynamical relation :
$$\left(\frac{\mathrm{\Omega }}{\eta }\right)_{T,h,\mu }=0.$$
(5)
For the investigation of equilibrium conditions one should separate two regimes: $`\mu =const`$ and $`n=const`$. We would like to note that hereafter we shall demonstrate the results of numerical investigation which show the main features of the considered model.
#### 1.1.1 $`\mu =const`$ regime.
The $`\mu =const`$ regime corresponds to the case when, for example, charge redistribution between the conducting sheets CuO<sub>2</sub> and other structural elements is allowed. For this regime the equilibrium is defined by the minimum of the grand canonical potential that form an equation for pseudospin mean value (5) and expression for $`n`$ (4).
At some regions of $`h`$ values the field dependencies $`\eta (h)`$ possesses $`S`$ \- like behaviour Fig. 1 (the first order phase transition with the jump of order parameter and electron concentration take place at the change of field $`h`$). The phase transition point is presented by a crossing point on the dependence $`\mathrm{\Omega }(h)`$.
The phase diagram $`T_ch`$ is shown in Fig. 2. One can see that with respect to Ising model the phase coexistence curve is shifted in field and distorted from the vertical line and hence the possibility of the first order phase transition with the temperature change exists in PE model.
The phase diagram $`\mu h`$ (Fig. 2) indicate stability regions for states with $`\eta =\pm 1/2`$ ($`U<g`$ and $`T=0`$). The form of diagram depends on the relation between $`U`$ and $`g`$ parameter values . Transitions between regions $`|r|p`$, $`|\stackrel{~}{r}|\stackrel{~}{p}`$ lead to the change of the average number of electrons only. At transitions $`|r|\stackrel{~}{r}`$ the flipping of pseudospin takes place, and at $`|r|\stackrel{~}{p}`$ $`(rp)`$ both processes occur.
Hence, the possibility of the first order phase transition with the change of field $`h`$ and/or chemical potential $`\mu `$ takes place and is shown by thick line on phase diagram in Fig. 2.
#### 1.1.2 $`n=const`$ regime.
In the regime of the fixed electron concentration value the equilibrium is determined by the minimum of free energy $`F=\mathrm{\Omega }+\mu N`$ and form a set of equations (4), (5) for the chemical potential and order parameter.
Typical example of the $`\eta (h)`$ dependence is shown in Fig. 3 which corresponds to the $`n`$ value in the interval $`0n1`$. Phases $`\eta =\frac{1}{2}`$ (phase 1), $`\eta =\frac{1}{2}n`$ (phase 2), $`\eta =\frac{1}{2}\frac{n}{2}`$ (phase 3), $`\eta =\frac{1}{2}`$ (phase 4) exist between phase transition points (which is determined according to the Maxwell rule from the range of $`S`$-like behaviour (between the spinodal points $`Z_i`$)) and outside of them. At the change of the model parameter values the regions, where metastable phases exist, can overlap, some phase transitions disappear (some intermediate phases can not be realized). In case $`1n2`$ the dependence $`\eta (h)`$ is generally similar. The phase 3 and phase 2’ at $`\eta =\frac{3}{2}n`$, which now appears instead of phase 2, may play the role of the intermediate phases.
On the phase diagram $`nh`$ Fig. 4 the thick solid line indicates the phase coexistence curve and hence the possibility of the first order phase transition with the change of the longitudinal field $`h`$ and/or electron concentration $`n`$ takes place.
More detail analyse of a free energy behaviour shows that the above presented (in this paragraph) results are not realized. The investigation of the equilibrium conditions shows that the first order phase transition transforms into the phase separation. One can see regions where state with homogenous distribution of particles is unstable ($`d\mu /dn0`$), and the phase separation into the regions with different concentrations exists (Fig. 5). The phase diagram $`nh`$ (Fig. 5) illustrates the separation phases.
The phase 3 splits into phase 4 (with concentration $`n=0`$, order parameter $`\eta =\frac{1}{2}`$) and phase 1 ($`n=2`$, $`\eta =\frac{1}{2}`$) with weight coefficients $`1n`$ and $`n`$ respectively (thin dotted lines in Fig. 4 and Fig. 5).
Therefore, more convenient and thermodynamically stable is the phase separated state, which is the mixture of states with different electron concentrations and different values of order parameters.
### 1.2 Antiferroelectric type interaction.
In the case of antiferroelectric type interaction it is convenient to introduce two kinds of sites (A-sites, B-sites). These corresponds to the doubling of the lattice period .
Within the framework of the MFA we shall write:
$$S_A^zS_B^z=\eta _A\eta _B+\eta _AS_B^z+\eta _BS_A^z,$$
(6)
where $`\eta _A=S_A^z`$, $`\eta _B=S_B^z`$.
Then, we obtain the following expression for the model Hamiltonian:
$`H`$ $`=`$ $`{\displaystyle \underset{i}{}}\stackrel{~}{H}_{iA}+{\displaystyle \underset{j}{}}\stackrel{~}{H}_{jB}+N\left\{{\displaystyle \frac{J_1}{2}}\eta _A\eta _B+{\displaystyle \frac{J_2}{4}}(\eta _A^2+\eta _B^2)\right\},`$
$`\stackrel{~}{H}_{iA}`$ $`=`$ $`\mu (n_{iA}+n_{iA})+Un_{iA}n_{iA}+g(n_{iA}+n_{iA})S_{iA}^z`$
$`(h+J_1\eta _B+J_2\eta _A)S_{iA}^z,`$
$`\stackrel{~}{H}_{jB}`$ $`=`$ $`\stackrel{~}{H}_{jA}|_{AB}.`$ (7)
The Hilbert space forms as a direct product of the eigenfunctions (2) for $`\stackrel{~}{H}_A`$ and $`\stackrel{~}{H}_B`$ operators (1.2) $`\left\{|n_{iA}N_{iA},S_{iA}^z\right\}\left\{|n_{iB}N_{iB},S_{iB}^z\right\}`$. The analytical consideration in this case in general is very similar to the previous (ferromagnetic interaction) one, but formulae are more complicated (see in details ).
Grand canonical potential can be written in the form:
$`\mathrm{\Omega }`$ $`=`$ $`J_1\eta _A\eta _B+{\displaystyle \frac{J_2}{2}}(\eta _A^2+\eta _B^2)+T\mathrm{ln}\left\{\left(\eta _A^2{\displaystyle \frac{1}{2}}\right)\left(\eta _B^2{\displaystyle \frac{1}{2}}\right)\right\}`$ (8)
$`T\mathrm{ln}\left(1+\mathrm{e}^{\beta (2\mu +U+g)}+2\mathrm{e}^{\beta \left(\mu +\frac{g}{2}\right)}\right)`$
$`T\mathrm{ln}\left(1+\mathrm{e}^{\beta (2\mu +Ug)}+2\mathrm{e}^{\beta \left(\mu \frac{g}{2}\right)}\right).`$
#### 1.2.1 $`\mu =const`$ regime.
The set of equations for $`\eta _A`$, $`\eta _B`$ is defined by the minimum of the grand canonical potential (8). The expression for the electron mean values is determined from the thermodynamical relation (4).
The form of the grand canonical potential field dependence $`\mathrm{\Omega }(h)`$ (and therefore the type and number of phase transitions) depends on the relation between parameters $`J_1`$ and $`J_2`$ values (Fig. 6). There are no any specific behaviour when $`J_1`$ and $`J_2`$ are placed in the regions 5 and 6. The case when $`J_1`$ and $`J_2`$ are placed in the domains 1, 7, 8 is similar to ferroelectric type interaction.
The location of $`J_1`$ and $`J_2`$ parameters within the area 4 leads to the possibility of the two sequential second order phase transitions: from the ferroelectric phase (FP) with the one pseudospin mean value to the antiferroelectric phase (AP) and then to the FP with the another order parameter value (Fig. 7).
The case when $`J_1`$ and $`J_2`$ belong to the region 2, 3 is shown in Fig. 7 and the corresponding phase diagram $`\mu `$$`h`$ in Fig. 8.
One can see that two first order phase transitions with the change of the field $`h`$ and/or chemical potential $`\mu `$ take place. With respect to ferroelectric type interaction between pseudospins (Fig. 2) the phase coexistence curve is split and one obtains the range (the range width is equal $`J_1`$) where the AP exists (Fig. 8). With the temperature increase the first order phase transitions transform into the second order phase transitions. Hence the possibility of the first order phase transition from FP into AP and then the second order phase transition from AP into FP exist with the temperature increase for the narrow range of $`h`$ values.
#### 1.2.2 $`n=const`$ regime.
As it was mentioned above, in the case of the fixed value of the electron concentration (regime $`n=`$const) the first order phase transition transform into the phase separation.
In Fig. 9 one present the phase diagram $`n`$$`h`$ when $`J_1`$, $`J_2`$ are placed in the region 2. Within the area surrounded by the lines the phase separation into the regions with different concentrations and phases FP (solid lines) and AP (dotted lines) takes place. Outside of these boundaries (which surround the phase separated states) the state with the space homogeneity of electron concentration (FP) is stable. Between the boundaries one have the AP.
We would like to remind that in the $`\mu =`$const regime with the temperature increase the first order phase transitions transform into the second order ($`J_1`$, $`J_2`$ $``$ domain 2,3). On the other hand, in the $`n=`$const regime this correspond to the narrowing of the range of the phase separated states and transform into the second order phase transition curves. Then the phase transition curves approach one to another and, finally, disappear at the certain value of temperature.
The location of $`J_1`$, $`J_2`$ within the area 4 leads to the possibility of the two second order phase transitions with the change of the field (similar to the $`\mu =`$const regime).
## 2 Interaction between pseudospins via conducting electron.
In the $`U=0`$ and $`J_{ij}=0`$ limit operator (1) can be transformed to the Hamiltonian of the electron subsystem of binary alloy in the case of equilibrium disorder. Model (1) is close to the Falicov-Kimball (FK) model but differ in thermodynamic equilibrium conditions, i.e. in a way how self-consistency is achieved ($`S^z=\mathrm{const}`$ for the FK model and $`h=\mathrm{const}`$ for the PE one).
In the present part of work we propose (for the case of the $`U=0`$ and $`J_{ij}=0`$ limit) the self-consistent scheme for calculation of mean values of pseudospin and particle number operators, grand canonical potential as well as correlation functions. The approach is based on the GRPA with the inclusion of the mean field corrections. The possibilities of the phase separation and chess-board phase appearance are investigated .
The calculation is performed in the strong coupling case ($`gt`$) using of single-site states as the basic one. The formalism of electron creation (annihilation) operators $`a_{i\sigma }=b_{i\sigma }P_i^+,`$ $`\stackrel{~}{a}_{i\sigma }=b_{i\sigma }P_i^{}`$ ($`P_i^\pm =\frac{1}{2}\pm S_i^z`$) acting at a site with the certain pseudospin orientation is introduced. Expansion of the calculated quantities in terms of electron transfer leads to the infinite series of terms containing the averages of the $`T`$-products of the $`a_{i\sigma }`$, $`\stackrel{~}{a}_{i\sigma }`$ operators. The evaluation of such averages is made using the corresponding Wick’s theorem. The averages of the products of the projection operators $`P_i^\pm `$ are expanded in semi-invariants .
Nonperturbated electron Green function is equal to
$$g(\omega _n)=g_i(\omega _n);g_i(\omega _n)=\frac{P_i^+}{\mathrm{i}\omega _n\epsilon }+\frac{P_i^{}}{\mathrm{i}\omega _n\stackrel{~}{\epsilon }},$$
(9)
where $`\epsilon =\mu +g/2,\stackrel{~}{\epsilon }=\mu g/2`$ are single-site energies. Single-electron Green function (calculated in Hubbard-I type approximation) is $`=G_𝒌(\omega _n)`$ $`=\left(g^1(\omega _n)t_𝒌\right)^1`$ and its poles determine the electron spectrum
$$\epsilon _{\mathrm{I},\mathrm{II}}(t_𝒌)=\frac{1}{2}(2E_0+t_𝒌)\pm \frac{1}{2}\sqrt{g^2+4t_𝒌S^zg+t_𝒌^2}.$$
(10)
In the adopted approximation the diagrammatic series for the pseudospin mean value can be presented in the form
$$S^z=\text{}$$
(11)
Here we use the following diagrammatic notations: $`\text{}S^z,`$ $`g_i(\omega _n),`$ wavy line is the Fourier transform of hopping $`t_𝒌`$. Semi-invariants are represented by ovals and contain the $`\delta `$-symbols on site indexes. In the spirit of the traditional mean field approach the renormalization of the basic semi-invariant by the insertion of independent loop fragments is taken into account in (11).
The analytical expression for the loop is the following:
$`={\displaystyle \frac{2}{N}}{\displaystyle \underset{n,𝒌}{}}t_𝒌^2{\displaystyle \frac{\left(P_i^+(\mathrm{i}\omega _n\epsilon )^1+P_i^{}(\mathrm{i}\omega _n\stackrel{~}{\epsilon })^1\right)}{g^1(\omega _n)t_𝒌}}`$ (12)
$`=\beta (\alpha _1P_i^++\alpha _2P_i^{}).`$
It should be noted that within the self-consistent scheme of the GRPA, the chain fragments form the single-electron Green function in the Hubbard-I approximation and in the sequences of loop diagrams in the expressions for grand canonical potential $`\mathrm{\Omega }`$ and pair correlation functions ($`S_i^zS_j^z`$, $`S_i^zn_j`$, $`n_in_j`$) the connections between any two loops by more than one semi-invariant are omitted. This procedure includes the renormalization of the higher order semi-invariants, which is similar to the one given by expression (11).
From (11) and (12) follows the equation for pseudospin mean value
$$S^z=\frac{1}{2}\mathrm{tanh}\left\{\frac{\beta }{2}(h+\alpha _2(S^z)\alpha _1(S^z))+\mathrm{ln}\frac{1+\mathrm{e}^{\beta \epsilon }}{1+\mathrm{e}^{\beta \stackrel{~}{\epsilon }}}\right\}.$$
(13)
The grand canonical potential in the considered approximation has the form:
$`\mathrm{\Delta }\mathrm{\Omega }=\mathrm{\Omega }\mathrm{\Omega }|_{t=0}={\displaystyle \frac{2}{N\beta }}{\displaystyle \underset{𝒌}{}}\mathrm{ln}{\displaystyle \frac{(\mathrm{cosh}\frac{\beta }{2}\epsilon _\mathrm{I}(t_𝒌))(\mathrm{cosh}\frac{\beta }{2}\epsilon _{\mathrm{II}}(t_𝒌))}{(\mathrm{cosh}\frac{\beta }{2}\epsilon )(\mathrm{cosh}\frac{\beta }{2}\stackrel{~}{\epsilon })}}+S^z(\alpha _2\alpha _1)`$
$`{\displaystyle \frac{1}{\beta }}\mathrm{ln}\mathrm{cosh}\left\{{\displaystyle \frac{\beta }{2}}(h+\alpha _2\alpha _1)+\mathrm{ln}{\displaystyle \frac{1+\mathrm{e}^{\beta \epsilon }}{1+\mathrm{e}^{\beta \stackrel{~}{\epsilon }}}}\right\}+{\displaystyle \frac{1}{\beta }}\mathrm{ln}\mathrm{cosh}\left\{{\displaystyle \frac{\beta }{2}}h+\mathrm{ln}{\displaystyle \frac{1+\mathrm{e}^{\beta \epsilon }}{1+\mathrm{e}^{\beta \stackrel{~}{\epsilon }}}}\right\}`$
The solution of equation (13) and calculation of potential $`\mathrm{\Omega }`$ were performed numerically.
In the $`\mu =\mathrm{const}`$ regime (in the uniform case) there exists a possibility of the first order phase transition with the jump of the pseudospin mean value and reconstruction of the electron spectrum .
In the $`n=\mathrm{const}`$ regime one can see (Fig. 10) that the regions with $`\mathrm{d}\mu /\mathrm{d}n0`$, where states with a homogeneous distribution of particles are unstable, exist. This corresponds to the phase separation into the states with different electron concentrations and pseudospin mean values. In the phase separated region the free energy as a function of $`n`$ deflects up (Fig. 10) and concentrations of the separated phases are determined by the tangent line touch points (these points are also the points of binodal lines which are determined according to the Maxwell rule from the function $`\mu (n)`$).
The analysis of the $`S^zS^z_𝒒`$ correlator temperature behaviour shows that the high temperature phase become unstable with respect to fluctuations with $`𝒒0`$ for certain values of model parameters. The maximal temperature of instability is achieved for $`𝒒=(\pi ,\pi )`$ (in the case of square lattice with nearest-neighbor hopping) and indicates the possibility of phase transition into a modulated (chess-board) phase.
The analytical consideration of the chess-board phase within the framework of the presented above approximation can be performed in a similar way.
From the comparison of the grand canonical potential $`\mathrm{\Omega }`$ values for uniform and chess-board phases, the ($`\mu ,h`$) phase diagram is obtained (Fig. 11). One can see that chess-board phase exists as an intermediate one between the uniform phases with the different $`S^z`$ and $`n`$ values. The transition between different uniform phases (bistability) is of the first order (Fig. 11, dashed line), while the transition between the uniform and modulated ones is of the first (dotted line) or second (solid line) order.
Appearance of various phases in the considered model remind the situation known for the FK model with a rich phase diagram . However, contrary to this model, an existence of the phase transitions between uniform phases is possible in our case. This results from the another regime of thermodynamic averaging (fixation of $`h`$ field which is an analogous to the chemical potential in the FK model).
Tabunshchyk K.V. e-mail: tkir@icmp.lviv.ua
|
warning/0006/hep-ph0006165.html
|
ar5iv
|
text
|
# Untitled Document
hep-ph/0006165 McGill-00/18
Two-Neutrino Five-Photon Scattering
at Low Energies
Y. Aghababaie and C.P. Burgess
<sup>a</sup> Physics Department, McGill University
3600 University St., Montréal, Québec, Canada, H3A 2T8.
Abstract
We extend earlier constructions of the effective action for neutrino-photon scattering, using the connection between low-energy neutrino-photon and photon-photon scattering together with the known effective Lagrangian describing low-energy photon scattering in QED. We use this effective action to calculate analytic expressions for the low-energy cross section for the (unpolarised) processes $`\nu \overline{\nu }5\gamma `$, $`\nu \gamma \nu +4\gamma `$ and $`\gamma \gamma \nu \overline{\nu }+3\gamma `$. As a byproduct we derive compact expressions for the $`N`$-body phase-space integrals for massless particles, including those having non-trivial tensor-structure.
1. Introduction
Neutrinos and photons may be, with gravitons, the only particles which are massless, or very nearly so. As such, they are the only degrees of freedom which arise at extremely low energies within the vacuum sector (and possibly within other sectors) of the Standard Model (SM). This makes the study of their low-energy interactions a theoretical laboratory for very-low-energy Standard-Model physics. This study may also have practical applications, despite the extremely weak strength of the interactions, because neutrinos play a unique role within the extreme environments found in astrophysics and cosmology. Precisely because of their weak couplings they are often the mediators of dynamically interesting processes, for instance by being responsible for heat and momentum transfer, especially in the late stages of stellar collapse.
$`22`$ processes — like $`\nu \overline{\nu }\gamma \gamma `$ and $`\nu \gamma \nu \gamma `$ — were first studied long ago and were found to be highly suppressed ,. The suppression arises because Yang’s theorem prohibits the coupling of two photons to a state of angular momentum one, and this ensures that the $`O(1/M_W^2)`$ contribution to the amplitude for these $`22`$ process must be zero (but see ). The dominant contribution therefore arises at $`O(1/M_W^4)`$, making it smaller than the nominally negligible processes which arise at higher order in $`\alpha `$, such as $`\nu \overline{\nu }3\gamma `$ and $`\nu \gamma \nu +2\gamma `$ . This observation has stimulated more detailed studies of these reactions (both at low energies and at energies above the electron mass , ), as well as generating searches for practical applications of these $`23`$ processes, such as within stars or in neutrino-photon scattering in the presence of magnetic fields .
As pointed out in ref , the effective action for these $`23`$ processes is related to the known Euler-Heisenberg effective action for four-photon scattering through the replacement of one of the electromagnetic tensors, $`F_{\alpha \beta }`$, by a neutrino ‘field strength’, of the form $`N_{\alpha \beta }=_\alpha \left(\overline{\nu }\gamma _\beta \gamma _L\nu \right)(\alpha \beta )`$. In ref. , this connection was examined in some detail: a particular combination of Feynman diagrams was employed to explicitly show the mapping between the calculation of the Euler-Heisenberg action and that governing the interactions of neutrino-antineutrino pairs with three photons.
In this note we have two goals. Our main new result is to extend this treatment to the next least complicated case: that of 2 neutrinos interacting with 5 photons. We do so by computing the relevant terms of the low-energy effective neutrino-photon lagrangian, and use them to calculate analytic expressions for the low-energy neutrino-photon scattering cross sections. In the centre-of-mass (CM) we find these to be of order $`\sigma \alpha ^3G_F^2E^{18}/(2\pi )^6m_e^{16}`$, where $`E`$ denotes the CM scattering energies, as compared with the $`22`$ result: $`\sigma 200\alpha G_F^4E^6/\pi ^3`$. In principle, the numerical factors are such that the $`25`$ processes can dominate the $`22`$ processes for $`Em_e`$ (which, of course, lies at the limit of validity of the low-energy approximation), although we know of no practical application of this observation.
Our secondary goal is not so much new as it is explanatory. Preparatory to describing the above results we recast the argument for the suppression of the $`22`$ processes into a more modern effective-lagrangian language. We also rederive the connection between the electromagnetic and neutrino scattering processes within this context. Although these are old results, we hope that their recasting in this way may suggest more applications elsewhere.
Our presentation is as follows. In the next section, §2, we review the low-energy limit of neutrino-photon scattering, rederiving both the connection to the Euler-Heisenberg effective lagrangian and the suppression of $`22`$ processes. This is followed in §3 by the derivation of the low-energy effective action for $`25`$ neutrino-photon interactions. §4 then applies this action to compute the three $`25`$ cross sections: $`\nu \overline{\nu }5\gamma `$, $`\nu \gamma \nu +4\gamma `$ and $`\gamma \gamma \nu \overline{\nu }+3\gamma `$. We conclude in the last section, §5, with comments and final remarks. Finally, an appendix describes an efficient method for computing the relevant $`N`$-body phase-space integrals which are encountered.
2. Low-Energy Neutrino-Photon Scattering Revisited
Because the SM contains no direct (tree-level) couplings between neutrinos and photons, the starting point for calculating their very-low-energy <sup>1</sup> <sup>1</sup> We use the name ‘very-low-energy’ to mean energies below the electron mass, in order to distinguish this from other potential notions of ‘low energy’, such as $`EM_W`$. interactions is the SM description of their couplings to other particles. These other particles then generate the effective neutrino-photon interactions once they are integrated out to produce the very-low-energy theory. The effective couplings are nonrenormalizable, in the sense that they are proportional to inverse powers of the masses of the particles which were integrated out to obtain them. Our main interest in what follows is in the dominant interactions at very low energy and so we focus on integrating out the lightest particles. Since the two lightest particles which couple to both neutrinos and photons are electrons and muons, we concentrate our attention on these.
2.1) The Weak-Scale Effective Theory
At energies below the $`W`$-boson mass, the couplings of neutrinos and photons to charged leptons are described by the effective lagrangian obtained from the SM by integrating out the top quark and the electroweak gauge bosons, $`W`$ and $`Z`$. The resulting effective interactions which are of most interest in what follows are those which are suppressed by the fewest powers of $`M_W`$ or $`M_Z`$. Those involving just neutrinos, photons and charged leptons, obtained by matching to the SM at $`\mu =M_W`$, are given by: <sup>2</sup> <sup>2</sup> Like all God-fearing people, our conventions are: $`\eta _{\mu \nu }=(,+,+,+)`$ and $`\overline{\psi }=i\psi ^{}\gamma ^0`$.
$$_{\mathrm{wk}}(\mu =M_W)=eA_\mu J_{\mathrm{em}}^\mu +\frac{G_F}{\sqrt{2}}\underset{klmn}{}\left(i\overline{\nu }_k\gamma _\mu \gamma _L\nu _l\right)L_{klmn}^\mu +O\left(\frac{1}{M_W^4}\right),$$
$`(1)\text{ }`$
where $`k,l,m,n=e,\mu ,\tau `$ run over the three lepton flavours, and the charged-lepton currents are given by
$$\begin{array}{cc}\hfill L_{klmn}^\mu & =i\overline{\mathrm{}}_m\gamma ^\mu \left(v_{klmn}+a_{klmn}\gamma _5\right)\mathrm{}_n,\hfill \\ \hfill J_{\mathrm{em}}^\mu & =\underset{k}{}i\overline{\mathrm{}}_k\gamma ^\mu \mathrm{}_k.\hfill \end{array}$$
$`(2)\text{ }`$
The effective couplings, $`v_{klmn}`$ and $`a_{klmn}`$, as found from tree-level matching are given by:
$$\begin{array}{cc}\hfill v_{klmn}(\mu =M_W)& =\delta _{kn}\delta _{lm}+\delta _{kl}\delta _{mn}\left(\frac{1}{2}+2s_w^2\right)\hfill \\ \hfill \text{and}a_{klmn}(\mu =M_W)& =\delta _{kn}\delta _{lm}\frac{1}{2}\delta _{kl}\delta _{mn},\hfill \end{array}$$
$`(3)\text{ }`$
at tree level, where $`s_w=\mathrm{sin}\theta _w`$ is the sine of the weak mixing angle.
Radiative corrections are easily incorporated into this language. Those loops involving high energy degrees of freedom (those involving particles having masses as large as $`M_W`$ or larger) are included by matching to the SM with higher-loop accuracy. Intermediate-scale loops (involving particles having masses between $`m_e`$ and $`M_W`$) are obtained by running the effective theory down to each new particle threshold, and then matching across this threshold.
One such high-energy loop generates an effective coupling between neutrinos and photons which is proportional to $`1/M_W^4`$ :
$$_{\mathrm{eff}}(\mu )=\frac{4\alpha }{\pi M_W^2}\left(\frac{G_F}{\sqrt{2}}\right)\left[1+\frac{4}{3}\mathrm{ln}\left(\frac{M_W^2}{\mu ^2}\right)\right]\left(i\overline{\nu }\gamma _\alpha \gamma _L\genfrac{}{}{0pt}{}{}{}_\beta \nu \right)F^{\beta \lambda }F_{}^{\alpha }{}_{\lambda }{}^{}.$$
$`(4)\text{ }`$
As we shall see, this particular higher-dimension interaction is not generated when lighter particles are integrated out, and so it is the dominant contribution to low-energy $`22`$ photon-neutrino scattering even though it is suppressed by four powers of $`M_W`$. <sup>3</sup> <sup>3</sup> In very-low-energy scattering applications it is the effective couplings renormalized at the electron mass which are required, so $`\mu =m_e`$ is used in eq. (4).
Imagine now writing down the effective theory at scale $`\mu =m_e`$, just before integrating out the electron. The only particles in this low-energy theory are the electron, photon and neutrinos. Conservation of electric charge and lepton numbers require the lowest-dimension neutrino couplings to electrons in this theory to again have the form of eq. (1), although now restricted to electrons and neutrinos. Furthermore, since all of the neutrino interactions in this effective theory must vanish in the limit where the $`W`$ and $`Z`$ become infinitely massive, they must be proportional to at least one factor of $`G_F`$.
It follows that the dominant low-energy interactions in the effective theory at this scale can differ from the electron terms of eq. (1) only through the values taken by the the coefficients $`v_{klee}(\mu =m_e)`$ and $`a_{klee}(\mu =m_e)`$. Happily enough, it also happens that $`v_{klee}(\mu =m_e)`$ cannot differ from its value, eq. (3), at $`\mu =M_W`$, because the current $`\overline{e}\gamma ^\alpha e`$ is conserved, and so does not get renormalized.
At low energies the sole contribution of physics between $`\mu =M_W`$ and $`\mu =m_e`$ therefore is to the running of the coupling $`a_{klee}`$ (and of the electric charge, $`e`$) between these scales, to all orders in all other SM couplings.
2.2) Matching at $`m_e`$
Next integrate out the electron itself to obtain the effective theory of photons and neutrinos only. At this stage we keep effective interactions having more than the minimal dimension, because these receive their largest coefficients when the lightest possible particle – the electron – is integrated out. We obtain in this way all contributions to low-energy neutrino-photon physics which are $`O(1/M_W^2m_e^p)`$, for all $`p`$.
To this order the new contributions to the effective neutrino-photon interaction lagrangian obtained by matching across the electron mass threshold is therefore given by
$$_{\mathrm{elth}}(\mu =m_e)=\frac{G_F}{\sqrt{2}}\underset{kl}{}\left(i\overline{\nu }_k\gamma _\mu \gamma _L\nu _l\right)\left(v_{klee}\overline{e}\gamma ^\mu e+a_{klee}\overline{e}\gamma ^\mu \gamma _5e\right),$$
$`(5)\text{ }`$
where $`X^\mu `$ represents the expectation of the operator $`X^\mu `$, obtained by integrating out the electrons, weighted by the QED lagrangian: <sup>4</sup> <sup>4</sup> A notational aside is in order here, since eq. (5) gives the impression that $`X^\mu `$ does not involve an integration over the electromagnetic field as well as the electron field. In reality this expectation denotes the usual matching procedure: the difference between the average calculated with electrons and photons in the theory just above $`m_e`$, and the average calculated with photons only in the effective theory just below $`m_e`$. This distinction plays no role in the present discussion.
$$X^\mu =𝒟e𝒟\overline{e}X^\mu (e,\overline{e})\mathrm{exp}\left[id^4x\left(_{\mathrm{kin}}ieA^\mu \overline{e}\gamma ^\mu e\right)\right].$$
$`(6)\text{ }`$
Eqs. (5) and (6) contain the nub of the main results, because it permits the following two conclusions:
$``$ Suppression of $`22`$ Processes:
As is easy to show, all operators involving only $`\overline{\nu }\gamma ^\mu \gamma _L\nu `$ – as opposed to $`\overline{\nu }\gamma _\alpha \gamma _L\genfrac{}{}{0pt}{}{}{}_\beta \nu `$ – and two electromagnetic fields vanish on using the equations of motion for the neutrino and photon fields, and so are redundant in the sense that they may be removed by performing a field redefinition. The only possible lowest dimension operator for $`22`$ processes (dimension 6) turns out to be of the form of eq. (4).
It remains to show that operators of this form are always suppressed by at least two powers of $`G_F`$. We have just argued that the right-hand-side of eq. (5) is explicitly proportional to the neutrino current, $`\overline{\nu }_k\gamma _\mu \gamma _L\nu _l`$, and so cannot contribute to an operator with a derivative embedded within the neutrino bilinear. This is why integrating out the electron does not generate the operator, eq. (4), with a coefficient proportional to $`G_F/m_e^2`$. The same argument also precludes generating such a term when the other charged leptons are integrated out. Charged-current interactions of neutrinos with quarks, on the other hand, can be linear in $`\nu `$, and so need not be proportional to $`\overline{\nu }_k\gamma _\mu \gamma _L\nu _l`$. Nonetheless, conservation of quark flavour only permits these interactions to contribute to neutrino/photon scattering at second order in $`G_F`$.
$``$ Connection with Photon-Photon Scattering:
Since the electromagnetic interactions preserve parity ($`𝒫`$) and charge conjugation ($`𝒞`$), these symmetries may be used to further organize the contributions to $`_{\mathrm{elth}}`$. In particular, these symmetries imply that any term in $`_{\mathrm{elth}}`$ involving an odd power of $`F_{\mu \nu }`$ receives contributions only from the vector current, $`\overline{e}\gamma ^\mu e`$, while those involving even powers of $`F_{\mu \nu }`$ arise purely from the axial current, $`\overline{e}\gamma ^\mu \gamma _5e`$.
Furthermore, since the vector current, $`\overline{e}\gamma ^\mu e`$, is also the electromagnetic current for the electron effective theory, its expectation may be expressed in terms of the Euler-Heisenberg effective lagrangian, $`W_{EH}[A]`$, for photon-photon scattering below $`m_e`$ , :
$$\begin{array}{cc}\hfill \overline{e}\gamma ^\mu e& =\frac{1}{e}\left(\frac{\delta Z}{\delta A_\mu }\right),\hfill \\ \hfill \text{where}Z[A]& =e^{iW_{EH}[A]}=𝒟e𝒟\overline{e}\mathrm{exp}\left[id^4x\left(_{\mathrm{kin}}ieA^\mu \overline{e}\gamma ^\mu e\right)\right].\hfill \end{array}$$
$`(7)\text{ }`$
For instance, since the quartic contribution to the Euler-Heisenberg interaction is given by:
$$_{EH}^{(4)}=\frac{\alpha ^2}{180m_e^4}\left[5(F_{\mu \nu }F^{\mu \nu })^214F_{\mu \nu }F^{\nu \lambda }F_{\lambda \rho }F^{\rho \mu }\right],$$
$`(8)\text{ }`$
it follows that the dominant contribution to $`_{\mathrm{elth}}`$ involving two neutrinos and three electromagnetic fields must be , :
$$_{\mathrm{elth}}^{(3)}=\frac{e(\frac{1}{2}+2s_w^2)\alpha }{90\pi m_e^4}\left(\frac{G_F}{\sqrt{2}}\right)\left[5(N_{\mu \nu }F^{\mu \nu })(F_{\lambda \rho }F^{\lambda \rho })14(N_{\mu \nu }F^{\nu \lambda }F_{\lambda \rho }F^{\rho \mu })\right],$$
$`(9)\text{ }`$
with $`N_{\alpha \beta }=_\alpha \left(\overline{\nu }\gamma _\beta \gamma _L\nu \right)(\alpha \beta )`$.
This method clearly works in general: to obtain any term involving an odd power of $`F_{\mu \nu }`$ in $`_{\mathrm{elth}}`$, replace one power of $`e`$ by $`G_F\left(\frac{1}{2}+2s_w^2\right)/\sqrt{2}`$, and sum all possible ways of replacing one electromagnetic field strength by $`N_{\mu \nu }`$.
2.3) The $`25`$ Effective Lagrangian
The next simplest neutrino-photon interaction which is related in this way to the Euler-Heisenberg action describes $`25`$ processes, like $`\overline{\nu }\nu 5\gamma `$. It is related to the sixth order term of the EH action, which is given by :
$$_{EM}^{(6)}=\frac{\pi \alpha ^3}{315m_e^8}\left[9(F_{\alpha \beta }F^{\alpha \beta })^326F_{\mu \nu }F^{\nu \lambda }F_{\lambda \rho }F^{\rho \mu }(F_{\alpha \beta }F^{\alpha \beta })\right].$$
$`(10)\text{ }`$
Now, given our previous arguments we can read off the effective two-neutrino/five-photon operators directly. We find:
$$\begin{array}{cc}\hfill _{\mathrm{eff}}^{\nu 5\gamma }& =\frac{\pi }{315}\frac{\alpha ^{5/2}}{\sqrt{4\pi }}\frac{G_F}{\sqrt{2}}\frac{\left(\frac{1}{2}+2s_w^2\right)}{m_e^8}[69(F_{\alpha \beta }F^{\alpha \beta })^2F_{\mu \nu }N^{\mu \nu }\hfill \\ & 426F_{\mu \nu }F^{\nu \lambda }F_{\lambda \rho }N^{\rho \mu }(F_{\alpha \beta }F^{\alpha \beta })226F_{\mu \nu }F^{\nu \lambda }F_{\lambda \rho }F^{\rho \mu }(F_{\alpha \beta }N^{\alpha \beta })],\hfill \end{array}$$
$`(11)\text{ }`$
A similar method will also give the even powers of $`F_{\mu \nu }`$ in $`_{\mathrm{elth}}`$ given the expression for the axial-vector/vector current correlation in QED <sup>5</sup> <sup>5</sup> After completing this paper it was brought to our attention that, in ref. , the expression for $`\overline{e}\gamma ^\mu \gamma _5e`$ has been worked out, up to fourth order in the fields. We thank H. Gies for pointing this reference out to us. .
3. Cross sections for $`\overline{\nu }\nu 5\gamma `$ and crossed processes
We next apply the lagrangian, eq. (11), to compute the $`25`$ processes $`\overline{\nu }\nu 5\gamma `$, $`\nu \gamma \nu +4\gamma `$ and $`\gamma \gamma \overline{\nu }\nu +3\gamma `$. This is a straightforward, if tedious, exercise within the effective theory, requiring only the Born approximation using interaction (11). (This should be contrasted with the difficulty of extracting the low-energy limit of the scattering amplitude, computed directly from the higher-loop graphs involving the weak interaction, eq. (1), and QED. As is usually the case with effective lagrangians, the payoff in simplicity is much larger for nonleading contributions.) In performing this calculation we employed the symbolic manipulation program FORM , which reduced the squared amplitude to its final form in under ten minutes on a desktop PC. We briefly sketch the method of computation below.
We require, then, the matrix elements of the effective interaction, eq. (11), which we write as follows:
$$_{\mathrm{eff}}^{\nu 5\gamma }=gN^{\mu \nu }T_{\mu \nu }^{\alpha _1\beta _1,\mathrm{},\alpha _5\beta _5}_{\alpha _1}A_{\beta _1}\mathrm{}_{\alpha _5}A_{\beta _5}$$
where $`g`$ is the factor premultiplying the square bracket in eq. (11) and $`T_{\mu \nu }^{\alpha _1\beta _1,\mathrm{},\alpha _5\beta _5}`$ represents the polynomial of momenta in the effective interaction. In terms of these quantities the matrix element relevant to $`\overline{\nu }\nu 5\gamma `$, for instance, becomes:
$$\gamma _1\mathrm{}\gamma _5|_{\mathrm{eff}}^{\nu 5\gamma }|\overline{\nu }\nu =g\stackrel{~}{N}^{\mu \nu }\stackrel{~}{T}_{\mu \nu }^{\alpha _1\beta _1,\mathrm{},\alpha _5\beta _5}\underset{i=1}{\overset{5}{}}\left(\frac{k_{(i)\alpha _i}ϵ_{\beta _i}(k_i;\lambda _i)}{\sqrt{(2\pi )^3k_i^0}}\right),$$
$`(12)\text{ }`$
where $`\stackrel{~}{N}^{\mu \nu }=0|N^{\mu \nu }|\overline{\nu }\nu `$ and
$$\stackrel{~}{T}_{\mu \nu }^{\alpha _1\beta _1,\mathrm{},\alpha _5\beta _5}=\underset{\pi S_5(1,\mathrm{},5)}{}T_{\mu \nu }^{\alpha _{\pi _1}\beta _{\pi _1},\mathrm{},\alpha _{\pi _5}\beta _{\pi _5}},$$
$`(13)\text{ }`$
is the permutation-summed tensor contracting the fields together. The $`ϵ^\mu `$’s are, as usual, the photon polarisation vectors.
After squaring and doing the spin sums, the following phase-space integral is required to obtain the total cross section:
$$_m^{\alpha _1\mathrm{}\alpha _m;\gamma _1\mathrm{}\gamma _m}(w)=\frac{d^3k_1}{2k_1^0}\mathrm{}\frac{d^3k_m}{2k_m^0}k_1^{\alpha _1}k_1^{\gamma _1}\mathrm{}k_m^{\alpha _m}k_m^{\gamma _m}\delta ^4(\underset{i=1}{\overset{m}{}}k_iw).$$
$`(14)\text{ }`$
A general technique for evaluating integrals of this form is given in the Appendix.
$``$ $`\overline{\nu }\nu 5\gamma `$: Using the integrals of the Appendix gives the final result for $`\overline{\nu }\nu 5\gamma `$:
$$\begin{array}{cc}\hfill \sigma (\nu \overline{\nu }5\gamma )& =\frac{1487}{(2\pi )^62^43^95^47^4}\alpha ^5m_e^2G_F^2\left(\frac{1}{2}+2s_w^2\right)^2\left(\frac{s}{m_e^2}\right)^9\hfill \\ & 6.8710^{38}\left(\frac{s}{m_e^2}\right)^9\mathrm{barn},\hfill \end{array}$$
$`(15)\text{ }`$
where $`s=(p+\overline{p})^2=2p\overline{p}`$ is the usual Mandelstam variable, equal to $`s=4E_\nu ^2`$ in the centre-of-mass frame.
$``$ $`\nu \gamma \nu +4\gamma `$: A similar exercise, after crossing the external lines, yields
$$\begin{array}{cc}\hfill \sigma (\nu \gamma \nu +4\gamma )& =\sigma (\overline{\nu }\gamma \overline{\nu }+4\gamma )\hfill \\ & =\frac{131632339}{(2\pi )^6\mathrm{\hspace{0.33em}2}^83^{10}5^57^411}\alpha ^5m_e^2G_F^2\left(\frac{1}{2}+2s_w^2\right)^2\left(\frac{s}{m_e^2}\right)^9\hfill \\ & 8.6810^{38}\left(\frac{s}{m_e^2}\right)^9\mathrm{barn}.\hfill \end{array}$$
$`(16)\text{ }`$
$``$ $`\gamma \gamma \overline{\nu }\nu +3\gamma `$: Crossing the other neutrino yields,
$$\begin{array}{cc}\hfill \sigma (\gamma \gamma \overline{\nu }\nu +3\gamma )& =\frac{797549}{(2\pi )^6\mathrm{\hspace{0.33em}2}^73^95^57^411}\alpha ^5m_e^2G_F^2\left(\frac{1}{2}+2s_w^2\right)^2\left(\frac{s}{m_e^2}\right)^9\hfill \\ & 8.3810^{38}\left(\frac{s}{m_e^2}\right)^9\mathrm{barn}.\hfill \end{array}$$
$`(17)\text{ }`$
4. Conclusion
We have constructed the effective interaction which governs the interactions of five photons and two neutrinos using the general connection between the effective action for neutrino-photon interactions at lowest order in $`G_F`$ and the known Euler-Heisenberg effective interaction for photon-photon scattering. As an application we have computed the two-body cross sections whose low-energy limits are given in terms of this effective interaction. While these cross sections are likely to be too small to be of any astrophysical relevance, it is interesting that the effective interaction for such a high-order process can be obtained with such minimal effort. It is also noteworthy that a seventh-order process can compete with the two-body scattering $`\overline{\nu }\nu 2\gamma `$. Finally, our expressions were obtained by evaluating multi-body phase space integrals, for which we have presented an efficient method of computation.
5. Acknowledgements
We would like to thank G. Mahlon and J. Matias for helpful conversations. This research was supported by funds from NSERC of Canada and FCAR of Québec.
Appendix I. Phase Space Integrals
Our goal in this appendix is to describe how to evaluate the integral, eq. (14), which we reproduce once more for convenience:
$$_m^{\alpha _1\mathrm{}\alpha _m;\gamma _1\mathrm{}\gamma _m}(w)=\frac{d^3k_1}{2k_1^0}\mathrm{}\frac{d^3k_m}{2k_m^0}k_1^{\alpha _1}k_1^{\gamma _1}\mathrm{}k_m^{\alpha _m}k_m^{\gamma _m}\delta ^4(\underset{i=1}{\overset{m}{}}k_iw).$$
We will find, through the integral representation of the delta function, that we are able to reduce the problem to one of taking derivatives of a suitable (simple) integral. Since the general expressions are extremely lengthy, and particular results are not difficult to obtain with the help of a symbol manipulation program once the prescription is known, we will only provide a detailed recipe for evaluating these integrals.
We proceed by using the Fourier representation of the delta function to factorise this integral into products of terms having the form
$$J^{\alpha \beta }(x)=\frac{d^3k}{2k^0}k^\alpha k^\beta e^{ikx},$$
so that
$$_m^{\alpha _1\mathrm{}\alpha _m;\gamma _1\mathrm{}\gamma _m}(w)=\frac{d^4x}{(2\pi )^4}J^{\alpha _1\gamma _1}(x)\mathrm{}J^{\alpha _m\gamma _m}(x)e^{iwx}.$$
$`(18)\text{ }`$
The integral defining $`J^{\alpha \beta }`$ is easily performed as follows:
$$J^{\alpha \beta }(x)=\frac{1}{(i)^2}\frac{}{x_\alpha }\frac{}{x_\beta }\frac{d^3k}{2k^0}e^{ikx}=\frac{4\pi }{x^6}(\eta ^{\alpha \beta }x^24x^\alpha x^\beta ),$$
$`(19)\text{ }`$
where we use the integral
$$\frac{d^3k}{2k^0}e^{ikx}=\frac{2\pi }{x^2}.$$
$`(20)\text{ }`$
We have ensured the convergence of this integral through the appropriate $`ϵ`$ prescription, taking $`x^2=(x^0)^2+𝐱^2+isgn(x^0)ϵ`$. The $`ϵ`$ term in $`x^2`$ forces the incoming momentum to be future-pointing, and allows the $`x`$ integrals to be done unambiguously. Notice that our result for $`J^{\alpha \beta }`$ is traceless, as is required, since the $`k`$s are null.
With $`J^{\alpha \beta }`$ in hand, expand the integrand of (18) to obtain a sum of integrals of the form
$$d^4xx^{\alpha _1}\mathrm{}x^{\alpha _n}\frac{e^{iwx}}{(x^2)^m},$$
where $`w`$ is a future-pointed, timelike four-vector. These integrals can all be evaluated by differentiating
$$_m(w):=d^4x\frac{e^{iwx}}{\left(x^2\right)^m},$$
$`(21)\text{ }`$
with respect to $`w`$, so that finding this integral reduces the problem of calculating (14) to one of expanding a polynomial and taking derivatives.
Let us evaluate (21). If we explicitly factor the denominator, go to the rest frame of $`w`$, and consider the $`t`$ integral first, we need to consider
$$\begin{array}{cc}\hfill _m(w)& =d^2\mathrm{\Omega }_0^{\mathrm{}}𝑑rr^2_{\mathrm{}}^{\mathrm{}}𝑑t\frac{e^{+i\omega t}}{[(tiϵ)+r]^m[(tiϵ)+r]^m}\hfill \\ & =\frac{1}{2}d^2\mathrm{\Omega }_{\mathrm{}}^{\mathrm{}}𝑑rr^2_{\mathrm{}}^{\mathrm{}}𝑑t\frac{e^{+i\omega t}}{()^m[t(r+iϵ)]^m[t+(riϵ)]^m}\hfill \\ & =2\pi ()^m_{\mathrm{}}^{\mathrm{}}𝑑rr^2_{\mathrm{}}^{\mathrm{}}𝑑t\frac{e^{+i\omega t}}{[t(r+iϵ)]^m[t(r+iϵ)]^m}.\hfill \end{array}$$
The $`t`$ integral in this expression is a contour integral, which is nonzero only for $`\omega >0`$, where $`\omega =w^0`$. For $`\omega >0`$ we close the contour upwards in a semicircle, break it into two, one around each pole, and use the Cauchy integral formula to get
$$\begin{array}{cc}& _m(w)=_{\mathrm{}}^{\mathrm{}}𝑑t\frac{e^{+i\omega t}}{[t(r+iϵ)]^m[t(r+iϵ)]^m}\hfill \\ & =\frac{2\pi i}{(m1)!}\left\{\frac{d^{m1}}{dz^{m1}}\left(\frac{e^{iz\omega }}{[z(iϵr)]^m}\right)_{z=r+iϵ}+\frac{d^{m1}}{dz^{m1}}\left(\frac{e^{iz\omega }}{[z(r+iϵ)]^m}\right)_{z=r+iϵ}\right\}\theta (\omega ).\hfill \end{array}$$
Using the Leibniz rule, $`d^m/dz^m(f(z)g(z))=_{s=0}^n\left(\genfrac{}{}{0pt}{}{m}{s}\right)f^{(ns)}(z)g^{(s)}(z)`$, yields,
$$\begin{array}{cc}\hfill _m(w)=& \frac{2\pi i}{[(m1)!]^2\mathrm{\hspace{0.33em}2}^m}\underset{s=0}{\overset{m1}{}}\left(\genfrac{}{}{0pt}{}{m1}{s}\right)()^s(m+s1)!\hfill \\ & \times \frac{(i\omega )^{m1s}}{2^s}\left[\frac{e^{ir\omega }}{r^{m+s}}+()^{m+s}\frac{e^{ir\omega }}{r^{m+s}}\right]\theta (\omega ).\hfill \end{array}$$
$`(22)\text{ }`$
Now using this in eq. (22) yields, after letting $`rr`$ in the second term of (22),
$$\begin{array}{cc}\hfill _m(w)=& \frac{8\pi ^2i}{[(m1)!]^2\mathrm{\hspace{0.33em}2}^m}\underset{s=0}{\overset{m1}{}}\left(\genfrac{}{}{0pt}{}{m1}{s}\right)()^{ms}(m+s1)!\hfill \\ & \times \theta (\omega )\frac{(i\omega )^{m1s}}{2^s}P_{\mathrm{}}^{\mathrm{}}𝑑r\frac{e^{ir\omega }}{r^{m+s2}}.\hfill \end{array}$$
$`(23)\text{ }`$
Finally, since
$$P_{\mathrm{}}^{\mathrm{}}𝑑r\frac{e^{ir\omega }}{r^n}=\frac{i\pi (i\omega )^{n1}}{(n1)!},$$
substitution into (23) yields
$$\begin{array}{cc}\hfill _m(w)=& ()^{m+1}\theta (\omega )\frac{8\pi ^3(i\omega )^{2m4}}{[(m1)!]^2\mathrm{\hspace{0.33em}2}^m}\underset{s=0}{\overset{m1}{}}\left(\genfrac{}{}{0pt}{}{m1}{s}\right)\frac{()^s}{2^s}(m+s1)(m+s2).\hfill \end{array}$$
The sum is easily recognised as the second derivative of $`2^{m3}[x(1x)]^{m1}`$, evaluated at $`x=\frac{1}{2}`$, and equal to $`\frac{2(m1)}{2^{m1}}`$, so we obtain the general expression,
$$_m(w)=\theta (\omega )\frac{32\pi ^3(m1)}{[(m1)!]^2\mathrm{\hspace{0.33em}4}^m}(w^2)^{m2},\mathrm{for}m3.$$
So, to summarise, by replacing the delta function by an integral, we are able to do each of the $`k`$ integrals separately, obtaining (18). Collecting the terms in the expansion and using (21) and its derivatives to proceed with the $`x`$ integrals yields the final answer to (14).
6. References
M. Gell-Mann, Phys. Rev. Lett. 6 (1961) 70.
D.A. Dicus, C. Kao and W. Repko, Phys. Rev. D48 (1993) 5106–5108 (hep-ph/9305284).
C.N. Yang, Phys. Rev. 77 (1950) 242.
D.A. Dicus, C. Kao and W. Repko, Phys. Rev. Lett. 79 (1997) 569–571 (hep-ph/9703210).
D.A. Dicus, C. Kao and W. Repko, Phys. Rev. D59 (1999) 013012 (hep-ph/9806499); ibid. 013005 (hep-ph/9806499).
A. Abada, J. Matias and R. Pittau, Nucl. Phys. B543 (1999) 255–268, (hep-ph/9808294).
See for example, D.A Dicus and W. Repko, hep-ph/0003305, and references therein.
H. Gies and R. Shaisultanov, Phys. Lett. B480 (2000) 129–134.
A. Abada, J. Matias and R. Pittau, Phys. Rev. D59 (1999) 013008 (hep-ph/9806383).
H. Euler, Ann. Phys. (Leipzig) 26, 398 (1936) ;
W. Heisenberg and H. Euler, Zeit. Phys. 98, 714 (1936).
D.A. Dicus, C. Kao and W. Repko, Phys. Rev. D57 (1998) 2443–2447 (hep-ph/9709415).
C. Schubert, hep-ph/0002276; hep-ph/0001288.
H. Gies and R. Shaisultanov, hep-ph/0003144.
J.A.M. Vermaseren, KEK-Preprint-92-1; ftp://nikhefh.nikhef.nl
|
warning/0006/hep-ph0006225.html
|
ar5iv
|
text
|
# The interplay between Weak and Strong Phases and Direct CP Violation from the Charmless B-meson Decays
## I introduction
Recently the CLEO collaboration has reported measurements on the branching ratios of rare hadronic B decays $`B\pi \pi ,\pi K`$. The data have attracted great interest from both theorists and experimentalists. The study of these channels will provide us important insights on understanding the effects of electroweak penguins (EWP) in B-system, and the final state interactions(FSI), as well as extracting the weak CKM phase $`\gamma =arg(V_{ud}V_{ub}^{}/V_{cd}V_{cb}^{})`$. It may also open a window for probing new physics.
From the current data, it is noticed that the branching ratio for $`B\pi ^+\pi ^{}`$ is relatively small, $`Br(B\pi ^+\pi ^{})4`$ (in units of $`10^6`$). The decays for $`B\pi ^+K^{},\pi ^{}\overline{K^0}`$ have almost equal decay rates, i.e., $`Br(B\pi ^+K^{})Br(B\pi ^{}\overline{K^0})17`$. While an unexpectedly large branching ratio for $`B\pi ^0\overline{K^0}`$ decay was also observed, $`Br(B\pi ^0\overline{K^0})14`$. These measurements seem in conflict with the calculations based on the naive factorization hypotheses. For the first three decays, it was pointed out that the factorization approach may still be valid if one takes the weak phase $`\gamma `$ to be greater than $`90^{}`$. With such a large $`\gamma `$, i.e., $`\mathrm{cos}\gamma <0`$, the interference between tree and penguin diagrams has opposite sign in $`B\pi \pi `$ and $`B\pi K`$ decays. Thus the negative $`\mathrm{cos}\gamma `$ will suppress the decay rate for $`B\pi \pi `$ and enhance that for $`B\pi ^+K^{}`$. As a consequence, the almost equal decay rates for $`B\pi ^+K^{}`$ and $`B\pi ^{}\overline{K^0}`$ decay modes indicate the dominance of strong penguin. However, the large rate for $`B\pi ^0\overline{K^0}`$ is not easily explained. Most recent analysis showed that a large final state interaction (FSI) phase would be helpful to enhance the branching ratio for the $`B\pi ^0\overline{K^0}`$ decay, but only considering the elastic-rescatterings remains insufficient to obtain the large central value of the data.
To understand the measured data, besides some model-dependent calculations, a model-independent approach using a single relative strong phase has also been proposed for the study of $`B\pi \pi ,\pi K`$ decays. Since the ordinary factorization approach suffers from the uncertainties due to hadronic matrix elements, such as the meson decay constants, B-meson form factors and so-called effective color number $`N_c`$. The model-independent analyses may be more useful as more data become available. The approach based on the isospin $`SU(2)`$ and approximate flavor $`SU(3)`$ symmetries of the strong interactions has been proposed to constrain and extract the weak phase $`\gamma `$. It has been noticed that the ratios between the CP-averaged decay rates, such as $`R=Br(B\pi ^\pm K^0)/Br(B\pi ^0K^\pm )`$ may provide us important information on the weak phase $`\gamma `$. Most recently, it has been shown that the weak phase $`\gamma `$ may be determined through three ratios among CP-averaged decay rates of $`B\pi ^+\pi ^{}`$, $`\pi ^+\pi ^0`$, $`\pi ^{}K^+`$ and $`\pi ^{}K^0`$ decays. Where two solutions were obtained at the $`1\sigma `$ level, one with positive $`\mathrm{cos}\delta `$ and negative $`\mathrm{cos}\gamma `$, i.e., relative small strong phase $`\delta `$ and large weak phase $`\gamma `$, and another with negative $`\mathrm{cos}\delta `$ and positive $`\mathrm{cos}\gamma `$. The latter with positive $`\mathrm{cos}\gamma `$ seems to be favored by solutions obtained from other constraints in the standard model but appears not to be as favorable as the one with negative $`\mathrm{cos}\gamma `$ studies of all the existing charmless decays are taken in account. However, there is still no complete analysis in the literature.
In this paper, we shall give a general analysis for all 7 decay modes of $`B\pi \pi `$ and $`\pi K`$. For that purpose, we will start from a general model independent parameterization for all the decay amplitudes by considering both the isospin and simple diagrammatic decompositions. We will show that there are in general 15 independent variables. By assuming the SU(3) relations which appropriately account for SU(3) symmetry breaking effects, it allows us to reduce the 15 independent variables into 9. They consist of 6 isospin amplitudes and 2 relative strong phase as well as one weak phase $`\gamma `$. Note that once the relative strong phase is zero, $`B\pi ^{}\overline{K}^0`$ only receives contributions from penguin-type diagram and the amplitudes of the isospin $`I=1/2`$ and $`I=3/2`$ amplitudes from tree-type diagrams cancel each other, which provides an additional constraint. In fact, one of the isospin amplitudes becomes almost irrelevant due to the suppression factor of the CKM mixing element. With such a consideration, there are only 8 relevant unknown quantities with 5 isospin amplitudes and 3 phases. We show that the current 6 measured decay rates allow us to extract 6 unknown quantities as functions of two variable. The upper bound of the decay rate $`B\pi ^0\pi ^0`$ also provides a bound for the difference between the two strong phases. Once taking the numerical value of the weak phase $`\gamma `$ to be the one obtained from other constraints in the standard model and fixing one of the strong phases, all the other parameters can be determined. With these determined parameters, we are then able to predict the branching ratio of the $`B\pi ^0\pi ^0`$ decay mode which is yet unmeasured due to the difficulty of its identification by the current detector. In addition we also present predictions for direct CP violations in all 7 decay channels of $`B\pi \pi ,\pi K`$. In our numerical fitting, we have adopted the $`\chi ^2`$ analysis for the CLEO data in order to have a systematic treatment on the experimental errors.
In general, according to the Watson theorem, there are two independent relative strong phases associating with the isospin amplitudes. They are often assumed to be equal in the literatures. In this work, we shall make a more general analysis with two relative strong phases. It is shown that the equal phase assumption will result in large enhancement of isospin amplitude $`a_{3/2}^c`$ which will be 5 times larger than the one calculated from the factorization approach. The value of the strong phase is found to be $`\delta \pm 95^{}`$. These large values may imply large inelastic FSI or indicate the possible new physics effects. However, if the two strong phases are different, the value of $`a_{3/2}^c`$ can be lower and is comparable with the usual factorization calculations.
It is remarkable to observe that within $`1\sigma `$ all the 6 decay rates can be consistently fitted for a large range of the weak phase $`0^{}<\gamma <180^{}`$ for the above two cases. It is also of interest to note that one of isospin amplitudes and the strong phases have a weak dependence on the weak phase $`\gamma `$. Three isospin amplitudes show a moderate dependence on the weak phase $`\gamma `$. Only one isospin amplitude is sensitive to $`\gamma `$. In particular, the fitting values for the 4 usual isospin amplitudes considered in most of the literature could still be comparable with the ones obtained by using naive factorization approach. The resulting large strong phases may be regarded as a strong indication of large FSI in $`B\pi \pi ,\pi K`$ decays.
## II The general framework
We begin with writing the decay amplitude of $`B\pi \pi ,\pi K`$ in the following general form:
$$A^{\pi \pi (\pi K)}=\lambda _u^{d(s)}A_u^{\pi \pi (\pi K)}+\lambda _c^{d(s)}A_c^{\pi \pi (\pi K)},$$
(1)
where $`\lambda _u^{d(s)}=V_{ub}V_{ud(s)}^{}`$ and $`\lambda _c^{d(s)}=V_{cb}V_{cd(s)}^{}`$ are the products of CKM matrix elements. The term proportional to $`\lambda _t^{d(s)}=V_{tb}V_{td(s)}^{}`$ has been absorbed into the above two terms by using the unitarity relation, $`V_{ub}V_{ud(s)}^{}+V_{cb}V_{cd(s)}^{}+V_{tb}V_{td(s)}=0`$.
we also find it useful to adopt the isospin decomposition for the decay amplitudes
$`A_{\pi ^{}\pi ^+}^{u,c}`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}a_0^{u,c}e^{i\delta _0}+\sqrt{{\displaystyle \frac{1}{3}}}a_2^{u,c}e^{i\delta _2}`$ (2)
$`A_{\pi ^0\pi ^0}^{u,c}`$ $`=`$ $`\sqrt{{\displaystyle \frac{1}{3}}}a_0^{u,c}e^{i\delta _0}\sqrt{{\displaystyle \frac{2}{3}}}a_2^{u,c}e^{i\delta _2}`$ (3)
$`A_{\pi ^{}\pi ^0}^{u,c}`$ $`=`$ $`\sqrt{{\displaystyle \frac{3}{2}}}a_2^{u,c}e^{i\delta _2}`$ (4)
$`A_{\pi ^+K^{}}^{u,c}`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}a_{1/2}^{u,c}e^{i\delta _{1/2}}+\sqrt{{\displaystyle \frac{1}{3}}}a_{3/2}^{u,c}e^{i\delta _{3/2}}`$ (5)
$`A_{\pi ^0\overline{K^0}}^{u,c}`$ $`=`$ $`\sqrt{{\displaystyle \frac{1}{3}}}a_{1/2}^{u,c}e^{i\delta _{1/2}}\sqrt{{\displaystyle \frac{2}{3}}}a_{3/2}^{u,c}e^{i\delta _{3/2}}`$ (6)
$`A_{\pi ^0K^{}}^{u,c}`$ $`=`$ $`\sqrt{{\displaystyle \frac{3}{2}}}a_{3/2}^{u,c}e^{i\delta _{3/2}}{\displaystyle \frac{1}{\sqrt{2}}}A_{\pi ^{}\overline{K^0}}^{u,c}`$ (7)
with
$$A_{\pi ^{}\overline{K^0}}^{u,c}=\sqrt{\frac{2}{3}}b_{1/2}^{u,c}e^{\delta _{1/2}^{}}\sqrt{\frac{1}{3}}a_{3/2}^{u,c}e^{i\delta _{3/2}}$$
(8)
Where $`a_I^{u,c}`$ and $`b_I^{u,c}`$ are the isospin amplitudes and $`\delta _I`$ and $`\delta _{1/2}^{}`$ are the strong phases due to final state interactions. In some literatures the strong phase of isospin amplitude $`b_{1/2}^{u,c}`$ is assumed to be equal to the one of $`a^{u,c}`$ for simplicity. However, in the most general case, these strong phases are not necessarily the same, since they arise from the effective Hamiltonian with different isospin. The subscript $`I=0,2,1/2,3/2`$ denote the isospins of the amplitudes. The advantage of the isospin decomposition allows one to use $`SU(3)`$ relations including leading order $`SU(3)`$ breaking effects. In other words, the isospin amplitudes are assumed to satisfy the following relations:
$`a_0^{u,c}`$ $``$ $`(f_\pi /f_K)a_{1/2}^{u,c},a_2^{u,c}(f_\pi /f_K)a_{3/2}^{u,c}`$ (9)
$`\delta _0`$ $``$ $`\delta _{1/2},\delta _2\delta _{3/2}`$ (10)
where $`f_\pi `$ and $`f_K`$ are the $`\pi ,K`$ meson decay constants with $`f_\pi /f_K0.8`$. For convenience, we define two phase differences as follows:
$`\delta `$ $`=`$ $`\delta _{3/2}\delta _{1/2}`$ (11)
$`\delta ^{}`$ $`=`$ $`\delta _{1/2}^{}\delta _{1/2}`$ (12)
Practically, the decay amplitudes are evaluated by calculating various Feynman diagrams. In order to see how those isospin amplitudes receive contributions from diagrams, we also present a simple diagrammatic decomposition. The diagrams can in general be classified into 6 types denoted by $`T`$(tree), $`C`$(color suppressed tree) $`P`$(QCD penguin), $`P_{EW}`$(electroweak penguin) and $`P_{EW}^C`$( color suppressed electroweak penguin) :
$`A_{\pi ^{}\pi ^+}`$ $`=`$ $`T+P+{\displaystyle \frac{2}{3}}P_{EW}^C`$ (13)
$`A_{\pi ^0\pi ^0}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(C+PP_{EW}{\displaystyle \frac{1}{3}}P_{EW}^C)`$ (14)
$`A_{\pi ^{}\pi ^0}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(TCP_{EW}P_{EW}^C)`$ (15)
$`A_{\pi ^+K^{}}`$ $`=`$ $`T^{}+P^{}+{\displaystyle \frac{2}{3}}P_{EW}^C`$ (16)
$`A_{\pi ^0\overline{K^0}}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(C^{}+P^{}P_{EW}^{}{\displaystyle \frac{1}{3}}P_{EW}^C)`$ (17)
where the primed and unprimed quantities are the amplitudes in $`B\pi K`$ and $`B\pi \pi `$ decays. They roughly differ by a factor $`f_\pi /f_K0.8`$ when the $`SU(3)`$ flavor symmetry breaking effects are considered.
Combining the two decompositions, it is straight forward to get the following relations:
$`a_{1/2}^{u,c}e^{i\delta _{1/2}}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{6}}}\left(2T^{}C^{}+3P^{}P_{EW}^{}+P_{EW}^C\right)^{u,c},`$ (18)
$`a_{3/2}^{u,c}e^{i\delta _{3/2}}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{3}}}\left(T^{}+C^{}+P_{EW}^{}+P_{EW}^C\right)^{u,c}.`$ (19)
From the above equations, one may easily see the relative magnitudes among those $`SU(2)`$ invariant amplitudes. If the inelastic rescattering effects are small, $`T^{},C^{}`$ will only contribute to the term proportional to $`\lambda _u^s`$. Therefore one may expect that $`T^{}(C^{})^uT^{}(C^{})^c`$. This will lead to $`a_{1/2(3/2)}^ua_{1/2(3/2)}^c`$. Since $`a_{1/2}^c`$ receives contributions from QCD penguins, while $`a_{3/2}^c`$ only gets contributions from EWPs, one may conclude that $`a_{1/2}^ca_{3/2}^c`$.
To obtain relations for the isospin amplitude $`b_{1/2}^{u,c}`$, one needs to be careful in adopting the diagrammatic decomposition implied by the naive factorization ansatz. This is because the resulting relative strong phase is zero in the factorization approach , i.e., $`\delta =\delta _{3/2}\delta _{1/2}=0`$ and $`\delta ^{}=0`$. As a consequence,
$`A_{\pi ^0K^{}}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}(T^{}C^{}P^{}P_{EW}^{}{\displaystyle \frac{2}{3}}P_{EW}^C)`$ (20)
$`A_{\pi ^{}\overline{K^0}}`$ $`=`$ $`P^{}{\displaystyle \frac{1}{3}}P_{EW}^C`$ (21)
The amplitudes with isospin $`I=1/2`$ and $`I=3/2`$ from tree-type graphs cancel each other in Eq.(21). Thus the total amplitude only receives contributions from penguin diagrams in this case, namely
$$A_{\pi ^{}\overline{K^0}}^{u,c}=\sqrt{\frac{2}{3}}b_{1/2}^{u,c}\sqrt{\frac{1}{3}}a_{3/2}^{u,c}=\left(P^{}\frac{1}{3}P_{EW}^C\right)^{u,c}$$
(22)
If assuming the $`t`$-quark dominance in the penguin loops, one finds from eq.(22) that :
$$A_{\pi ^{}\overline{K^0}}^uA_{\pi ^{}\overline{K^0}}^corb_{1/2}^ub_{1/2}^c+\frac{1}{\sqrt{2}}a_{3/2}^u\frac{1}{\sqrt{2}}a_{3/2}^c.$$
(23)
which may be assumed for simplicity to be approximately valid after considering final state interactions with nonzero strong phases. In the numerical calculations, we have check that the amplitude $`b_{1/2}^u`$ is less important due to the strong suppression of the CKM factor (for instance, even if taking $`b_{1/2}^ub_{1/2}^c`$, the results remain almost unchanged).
With the above analyses, let us provide an intuitive discussion on how to yield a large branching ratio for $`B\pi ^0\overline{K^0}`$ decay by appropriately choosing the isospin amplitudes. Note the fact that $`\lambda _u^s\lambda _c^s`$, one may roughly estimate the ratio between $`Br(B\pi ^0\overline{K^0})`$ and $`Br(B\pi ^+K^{})`$ by neglecting the terms containing the CKM factor $`\lambda _u^s`$
$`R={\displaystyle \frac{Br(B\pi ^0\overline{K^0})}{Br(B\pi ^+K^{})}}`$ $``$ $`\left|{\displaystyle \frac{\sqrt{\frac{1}{3}}a_{1/2}^c\sqrt{\frac{2}{3}}a_{3/2}^ce^{i\delta }}{\sqrt{\frac{2}{3}}a_{1/2}^c+\sqrt{\frac{1}{3}}a_{3/2}^ce^{i\delta }}}\right|^2,`$ (24)
with $`\delta =\delta _{3/2}\delta _{1/2}`$. Once neglecting $`a_{3/2}^c`$, the ratio may be simply given by $`R\frac{1}{2}`$, which is much smaller than the central value of the data $`R=0.84`$. It indicates that to enhance the decay rate of $`B\pi ^0\overline{K^0}`$, the isospin amplitude $`a_{3/2}^c`$ should not be neglected. Its small value may provide a sizable contribution for a large value of $`\delta >\pi /2`$. This is because in this case there exists a constructive interference between $`a_{1/2}^c`$ and $`a_{3/2}^c`$ in $`B\pi ^0\overline{K^0}`$ and a destructive interference in $`B\pi ^+K^{}`$. The situation is quite similar to the case for a large $`\gamma >\pi /2`$, which is considered to enhance $`B\pi K`$ and decrease $`B\pi \pi `$ decay rates. From eq.(24), it is easily seen that the value of $`a_{3/2}^c`$ satisfies
$`a_{3/2}^c{\displaystyle \frac{\sqrt{2R}1}{\sqrt{2}+\sqrt{R}}}a_{1/2}^c0.12a_{1/2}^c.`$ (25)
With the above considerations, there are only 8 relevant quantities: $`a_{1/2}^{u,c},a_{3/2}^{u,c},b_{1/2}^c,\delta ,\delta ^{},\gamma `$, which should be constrained by six measured decay rates and one upper bound. When taking the weak phase $`\gamma `$ as a free parameter, the rest 6 variables can be determined from 6 equations of eqs.(2-7). As the errors in the current data remain considerable large, one may not take the central values of the data to be too serious. Thus by only using the central values of the data to determine the 6 variables may not be good enough. To take into account the experimental errors in a systematic way, we shall adopt a global $`\chi ^2`$( least square) analysis for the present data. This treatment allows us to obtain not only the central value but also the errors for the fitting amplitudes. Our fitting will be carried out by using the standard $`\chi ^2`$ analysis program package MINUIT.
## III Results and discussions
In order to compare with the values estimated from the factorization, it is necessary to explicitly see how large of the isospin amplitudes $`a_{1/2,3/2}^{u,c}`$ and $`b_{1/2}`$ from the factorization calculations, we present the relevant formulae for the $`B\pi \pi ,\pi K`$ decay amplitudes with the assumption of factorization:
$`A_{\pi ^+K^{}}`$ $`=`$ $`i{\displaystyle \frac{G_F}{\sqrt{2}}}f_KF_0^{B\pi }(m_K^2)(m_B^2m_\pi ^2)\left(\lambda _u^sa_1\lambda _t^s(a_4+a_{10}+(a_6+a_8)R_4)\right)`$ (26)
$`A_{\pi ^0\overline{K^0}}`$ $`=`$ $`i{\displaystyle \frac{G_F}{2}}f_KF_0^{B\pi }(m_K^2)(m_B^2m_\pi ^2)\lambda _t^s\left(a_4+a_6R_5{\displaystyle \frac{1}{2}}(a_{10}+a_8R_5)\right)`$ (28)
$`i{\displaystyle \frac{G_F}{2}}f_\pi F_0^{BK}(m_\pi ^2)(m_B^2m_K^2)\left(\lambda _u^sa_2\lambda _t^s{\displaystyle \frac{3}{2}}(a_9a_7)\right)`$
$`A_{\pi ^0K^{}}`$ $`=`$ $`i{\displaystyle \frac{G_F}{2}}f_KF_0^{B\pi }(m_K^2)(m_B^2m_\pi ^2)(\lambda _u^sa_1\lambda _t^s(a_4+a_6R_4+(a_{10}+a_8R_4))`$ (30)
$`i{\displaystyle \frac{G_F}{2}}f_\pi F_0^{BK}(m_\pi ^2)(m_B^2m_K^2)\left(\lambda _u^sa_2\lambda _t^s{\displaystyle \frac{3}{2}}(a_9a_7)\right)`$
$`A_{\pi ^{}\overline{K^0}}^c`$ $`=`$ $`i{\displaystyle \frac{G_F}{\sqrt{2}}}f_KF_0^{B\pi }(m_K^2)(m_B^2m_\pi ^2)\lambda _t^s\left(a_4+a_6R_5{\displaystyle \frac{1}{2}}(a_{10}+a_8R_5)\right)`$ (31)
where $`f_{\pi ,K}`$ and $`F^{B\pi ,BK}`$ are the decay constants and $`B`$-meson form factors respectively, $`R_4=2m_K^2/(m_bm_u)(m_u+m_s)`$ and $`R_5=2m_K^2/(m_bm_d)(m_d+m_s)`$. In the flavor $`SU(2)`$ limit, one has $`R_4R_52m_K^2/(m_bm_s)`$.
The expressions of the isospin amplitudes can be rewritten as follows:
$`a_{1/2}^ue^{i\delta _{1/2}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}A_{\pi ^+K^{}}^u+\sqrt{{\displaystyle \frac{1}{3}}}A_{\pi ^0\overline{K^0}}^u`$ (32)
$`=`$ $`r(\sqrt{{\displaystyle \frac{2}{3}}}(a_1+a_4+a_{10}+(a_6+a_8)R_4)+\sqrt{{\displaystyle \frac{1}{6}}}(a_4+a_6R_5{\displaystyle \frac{1}{2}}(a_{10}+a_8R_5))`$ (34)
$`\sqrt{{\displaystyle \frac{1}{6}}}(a_2+{\displaystyle \frac{3}{2}}(a_9a_7)X))`$
$`a_{1/2}^ce^{i\delta _{1/2}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}A_{\pi ^+K^{}}^c+\sqrt{{\displaystyle \frac{1}{3}}}A_{\pi ^0\overline{K^0}}^c`$ (35)
$`=`$ $`r(\sqrt{{\displaystyle \frac{2}{3}}}(a_4+a_{10}+(a_6+a_8)R_4)+\sqrt{{\displaystyle \frac{1}{6}}}(a_4+a_6R_5{\displaystyle \frac{1}{2}}(a_{10}+a_8R_5))`$ (37)
$`\sqrt{{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{3}{2}}}}(a_9a_7)X)`$
$`a_{3/2}^ue^{i\delta _{3/2}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{1}{3}}}A_{\pi ^+K^{}}^u\sqrt{{\displaystyle \frac{2}{3}}}A_{\pi ^0\overline{K^0}}^u`$ (38)
$`=`$ $`r\sqrt{{\displaystyle \frac{1}{3}}}\left(a_1+a_2X+{\displaystyle \frac{3}{2}}(a_9a_7)X+{\displaystyle \frac{3}{2}}(a_{10}+a_8)R_4\right)`$ (39)
$`a_{3/2}^ce^{i\delta _{3/2}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{1}{3}}}A_{\pi ^+K^{}}^c\sqrt{{\displaystyle \frac{2}{3}}}A_{\pi ^0\overline{K^0}}^c`$ (40)
$`=`$ $`r{\displaystyle \frac{\sqrt{3}}{2}}(a_{10}+a_8R_4+(a_9a_7)X)`$ (41)
and
$`\sqrt{{\displaystyle \frac{2}{3}}}b_{1/2}^c{\displaystyle \frac{1}{\sqrt{3}}}a_{3/2}^c`$ $`=`$ $`A_{\pi ^{}\overline{K^0}}^c`$ (42)
$`=`$ $`r\left(a_4+a_6R_5{\displaystyle \frac{1}{2}}(a_{10}+a_8R_5)\right)`$ (43)
where $`r=i\frac{G_F}{\sqrt{2}}f_KF_0^{B\pi }(m_K^2)(m_B^2m_\pi ^2)`$ and $`X=(f_\pi /f_K)(F_0^{BK}/F_0^{B\pi })(m_B^2m_K^2)/(m_B^2m_\pi ^2)`$. In our numerical estimates, we will take $`f_\pi =133`$ MeV, $`f_K=158`$ MeV, $`F_0^{B\pi }=0.36`$ and $`F_0^{BK}=0.41.`$ There remains a large uncertainty in strange quark mass $`m_s`$. For $`m_s=(100200)`$ MeV, we find that the numerical values of those amplitudes are given by
$`a_{1/2}^u`$ $``$ $`818846,a_{3/2}^u709,`$ (44)
$`a_{1/2}^c`$ $``$ $`(103131),a_{3/2}^c7,`$ (45)
$`b_{1/2}^c`$ $``$ $`(72100).`$ (46)
The results from the $`\chi ^2`$-fitting are shown in Fig.1-3, where the six amplitudes as well as their errors at 1$`\sigma `$ level are obtained as functions of the weak phase $`\gamma `$ with $`\delta ^{}`$ fixed at $`0,\pi /6,\pi /3`$ respectively. The relative magnitudes of the amplitudes are consistent with the previous discussions. In our fit, the minimum value of $`\chi ^2`$ is found to be extremly low ( typiclly $`\chi _{min}^20.5\times 10^{12}`$). This means that the $`\chi ^2`$-fits are highly consistent and the 6 amplitudes are actually extracted as the solutions of Eqs.2-8. It can be seen from the figures that the $`\gamma `$ dependence of $`a_{1/2}^u`$ is quite strong and the one of $`a_{1/2}^c`$ and $`a_{3/2}^{u,c}`$ is relatively weak. On the other hand, it may be used to extract the angle $`\gamma `$ once one of those amplitudes can be determined or calculated in other independent ways. It is of interest to see that the $`\gamma `$ dependence of the amplitude $`b_{1/2}^c`$ and the strong phase $`\delta `$ is weak, which shows that these two quantities are approximately fixed. The possibility of large $`\delta `$ was also suggested in Ref. to explain the large branching ratio of the $`B\pi ^0\overline{K^0}`$ decay. Recently, perturbative QCD calculations have also shown a large strong phase.
In Fig.1 where the phase difference $`\delta ^{}`$ is set to be zero as usual, the $`\chi ^2`$-fitting shows that the values of $`a_{1/2}^u`$ and $`a_{3/2}^u`$ may be comparable with the ones from the theoretical estimations. only when the weak phase $`\gamma `$ is large. Especially for $`\gamma >2\pi /3`$, the fitting values could coincide with the ones from factorization except for $`a_{3/2}^c`$. For $`\gamma <2\pi /3`$, the two amplitudes are smaller than the ones from naive factorization calculations. It appears that the factorization approach may become suitable for large weak phase $`\gamma `$. This phenomenon was observed by most of the analyses in the literature which neglects the isospin amplitude $`a_{3/2}^c`$. As a consequence, the resulting large value of $`\gamma `$ seems to be conflict with the one obtained from other constraints in the standard model. Before drawing the final conclusion, one also notice that in the factorization approach, one yields a zero strong phase $`\delta =0^{}`$ which actually contradicts with the general fitting value $`\delta \pm 95^{}`$. Therefore, the results of estimates based on the naive factorization approach should be unreliable, and the isospin amplitudes must receive additional large contributions. A large value for the relative strong phase $`\delta \pm 95^{}`$ implies that the final state interactions or inelastic rescattering effects must be significant.
We would like to stress that the most outstanding feature of the $`\chi ^2`$ analysis with $`\delta ^{}=0`$ is that the isospin amplitude $`a_{3/2}^c`$ is likely to be relatively larger than the one estimated from the naive factorization calculations. The fitting central value of $`a_{3/2}^c`$ may be larger by a factor of $`79`$. To explicitly see how the decay rates depend on the isospin amplitude $`a_{3/2}^c`$, we plot in Fig.4 the 6 branching ratios of $`B\pi \pi ,\pi K`$ decays as functions of $`a_{3/2}^c`$. It can be clearly seen that if $`\delta ^{}=0`$, a small value of $`a_{3/2}^c7`$ is not able to reproduce all the CLEO data within the $`1\sigma `$ level, especially the data for the channels of $`\pi ^0\overline{K^0}`$ and $`\pi ^0K^{}`$. To consistently describe the whole data, we need a relative large value $`a_{3/2}^c75`$ for fitting the central value of the data which is about $`10\%`$ of the largest one $`a_0^u`$.
Within the standard model, it seems difficult to enhance the isospin amplitude $`a_{3/2}^c`$ by an order of magnitude even the inelastic FSI is involved, this is because the main inelastic channels such as $`BDD(DD_s,\eta _cK)\pi \pi (K)`$, only contribute to isospin-$`\frac{1}{2}`$ part of the decay amplitude. In the SM, it is known that the ratio $`a_{3/2}^c/a_{3/2}^u`$ can be determined without the hadronic uncertainties in the flavor $`SU(3)`$ limit, this is because the ratio only depends on the short distance Wilson coefficients. Thus a large value of $`a_{3/2}^c/a_{3/2}^u`$ may indicate the existence of new physics. While all models beyond the standard model must effectively provide large contributions to the electroweak penguins in order to enhance the isospin amplitude $`a_{3/2}^c`$, such models are: SUSY with R parity violation, $`Z^{}`$-model and $`Z`$-mediated FCNC models, et. al.
Let us now consider the case that $`\delta ^{}`$ is non-zero, the situation then becomes quite different. In Fig.2 and Fig.3, it is seen that $`a_{3/2}^c`$ decreases as the value of $`\delta ^{}`$ increases. When $`\delta ^{}`$ reaches $`\pi /3`$ , $`a_{3/2}^c`$ will be consistent with the value yielded from factorization approach. It is also noticed from the figures that a large $`\delta ^{}`$ leads to large values of $`a_{1/2}^u`$ and $`\delta `$. The enhancement of $`a_{1/2}^u`$ may be easily understood than the enhancement of $`a_{3/2}^c`$ due to final state interactions. Nevertheless, as will be discussed below, the values of $`\delta ^{}`$ cannot be too large due to the constraint of the upper bound of the branching ratio of $`B\pi ^0\pi ^0`$.
When all the isospin amplitudes and strong phases are determined, one is able to predict the direct CP asymmetries for all the relevant decay channels. The direct CP asymmetry in $`B\pi \pi ,\pi K`$ decays is defined in the standard way
$$A_{CP}=\frac{\mathrm{\Gamma }(\overline{B}\overline{f})\mathrm{\Gamma }(Bf)}{\mathrm{\Gamma }(\overline{B}\overline{f})+\mathrm{\Gamma }(Bf)}a_{ϵ^{\prime \prime }}^f$$
(47)
where $`f`$ denotes the final state mesons. In Fig.5 we plot several $`|A_{CP}|`$’s as functions of the weak phase $`\gamma `$ with different value of $`\delta ^{}`$. When $`\gamma `$ is near $`90^{}`$ and $`\delta ^{}=0`$ one has $`|A_{CP}(\pi ^+K^{})|0.04`$ which is in a good agreement with the most recent CLEO data $`A_{CP}=0.04\pm 0.16`$. At this point, we have a reliable prediction for the direct CP violations in the following decay modes. The $`A_{CP}`$’s with $`45^{}<\gamma <95^{}`$ read
$`|A_{CP}(\pi ^+K^{})|(2.54)\%,|A_{CP}(\pi ^0\overline{K^0})|(2.55)\%,`$ (48)
$`|A_{CP}(\pi ^0K^{})|(7.510)\%,|A_{CP}(\pi ^{}\overline{K}^0)|(56.5)\%,`$ (49)
$`|A_{CP}(\pi ^+\pi ^{})|(7.512.5)\%,|A_{CP}(\pi ^0\pi ^0)|(7.512)\%.`$ (50)
for $`\delta ^{}=0`$, and
$`|A_{CP}(\pi ^+K^{})|(58)\%,|A_{CP}(\pi ^0\overline{K^0})|(810)\%,`$ (51)
$`|A_{CP}(\pi ^0K^{})|(1218)\%,|A_{CP}(\pi ^{}\overline{K}^0)|(812)\%,`$ (52)
$`|A_{CP}(\pi ^+\pi ^{})|(1624)\%,|A_{CP}(\pi ^0\pi ^0)|(1014)\%.`$ (53)
for $`\delta ^{}=30^{}`$. In spite of the large $`\delta `$, the smallness of CP asymmetries in $`B\pi ^{}\pi ^0`$ decay is due to the absence of the interference between tree and penguin diagrams.
There remains an unobserved decay mode in $`B\pi \pi ,\pi K`$ decays, which is the decay mode $`B\pi ^0\pi ^0`$ ( the CLEO collaboration has already reported the indication of $`Br(B\pi ^0\pi ^{})5.6`$ ). As all the relevant isospin amplitudes have been determined as functions of $`\gamma `$, it allows us to predict $`B\pi ^0\pi ^0`$ as a function of $`\gamma `$. It is interesting to note that our $`\chi ^2`$-analysis shows that the resulting branching ratio $`Br(B\pi ^0\pi ^0)`$ is almost independent of the weak phase $`\gamma `$. Its value at $`\delta ^{}=0`$ is close to the one of $`B\pi ^+\pi ^{}`$ decay
$$Br(B\pi ^0\pi ^0)4.6\times 10^6.$$
(54)
with $`\delta ^{}`$ increasing, the branching ratio becomes larger and can reach $`7(10)\times 10^6`$ when $`\delta ^{}=30^{}(60^{})`$. Such a large branching ratio is about an order of magnitude larger than the prediction based on factorization calculations. Most recently, the CLEO collaboration reported an upper bound of $`Br(B\pi ^0\pi ^0)<5.7\times 10^{(6)}`$, this will impose a strong constraint on the value of $`\delta ^{}`$ (see Fig. 6). It is seen that to be consistent with the data at the $`1\sigma `$ level, the upper bound of the branching ratio $`Br(B\pi ^0\pi ^0)`$ limits $`\delta ^{}`$ to be less than $`50^{}`$.
## IV conclusions
In summary, we have made a general less model-dependent investigation on the charmless B-meson decays by using the $`\chi ^2`$ analysis based on the most recent CLEO data. We have used the most general isospin decomposition with two independent strong phases. All the isospin amplitudes in rare hadronic $`B`$ decays $`B\pi \pi ,\pi K`$ can be determined as functions of the weak phase $`\gamma `$ and one strong phase $`\delta ^{}`$. The effects of two equal and unequal strong phases are studied in detail. It is found that the isospin amplitude $`b_{1/2}^c`$ and the strong phase $`\delta `$ only slightly depend on the phase $`\gamma `$. An important observation under the equal strong phase assumption is the relative large isospin amplitude $`a_{3/2}^c`$, $`(a_{3/2}^c70)`$ where the central value is about 5 times greater than the one obtained from the factorization calculations. When the two strong phases are not equal, the allowed values of $`a_{3/2}^c`$ decrease as their difference, i.e., $`\delta ^{}`$, increases. For the most general case with two rather than one large FSI strong phases, the magnitude of all the isospin amplitudes may be around the one estimated from the factorization approach. Nevertheless, one needs to find out the mechanism of producing large strong phases. This could directly be tested by measuring the branching ratio $`Br(B\pi ^0\pi ^0)`$.
The direct CP asymmetries $`A_{CP}^f`$ for all the relevant decay channels have also been given as functions of $`\gamma `$. The resulting numerical value for $`A_{CP}^{\pi ^+K^{}}`$ is consistent with the most recent data. When taking $`\gamma `$ to be in a reasonable range $`\gamma =45^{}95^{}`$, we are led to the results given in eqs.(35) and (36), which can be directly tested by experiments in the near future. A resulting large branching ratio $`Br(B\pi ^0\pi ^0)`$ which is comparable with the one $`Br(B\pi ^+\pi ^{})`$ will also provide an important and consistent test.
From the most general analysis presented in this paper, the data appears to strongly suggest that final state interactions and inelastic rescattering effects must be significant and play an important rule in the charmless $`B\pi \pi ,\pi K`$ decays. Otherwise, our general analyses may be interpreted as hinting at the existence of new physics. For a more definite conclusion, one needs more precise data. The two B-factories BaBaR and BELLE are expected to provide us with more information from the charmless decays.
Acknowledgments:
The author (C.Q.G.) was supported by the NSC under contract number NSC 89-2112-M-007-013. J.N.N. is partially supported by the Natural Science and Engineering Council of Canada. Two of us (Y.L.W. and Y.F.Z.) were supported in part by NSF of China under grant No. $`\mathrm{\#}`$ 19625514. Y.L.W acknowledges the CTS for the support during his visit and he would also like to thank H.Y. Cheng, X.G. He, W.S. Hou and H.N. Li for useful discussions.
|
warning/0006/cond-mat0006052.html
|
ar5iv
|
text
|
# Atomic scale engines: Cars and wheels
## Abstract
We introduce a new approach to build microscopic engines on the atomic scale that move translationally or rotationally and can perform useful functions such as pulling of a cargo. Characteristic of these engines is the possibility to determine dynamically the directionality of the motion. The approach is based on the transformation of the fed energy to directed motion through a dynamical competition between the intrinsic lengths of the moving object and the supporting carrier.
The handling of single atoms and molecules has become widespread in science , but the challenge still remains to further ‘tame’ them and make single molecules perform useful functions. Nanoscale technology has been predicted almost 40 years ago , but inspite of a growing interest in atomic scale engines, such as biological motors , ratchet systems , molecular rotors , and molecular machinery in general , a real breakthrough concerning the construction of a man-made nanoscale counterpart of the ‘steam engine’ has not occured yet. This has mainly been due to the fact that we still miss the crucial link of how to transform energy to directed motion on this scale.
In this Letter we propose possible basic principles of such an engine. The main advantages of this novel approach are: (a) the same concept applies for both translational and rotational motions, (b) the directionality of motion is determined dynamically and does not require spatial asymmetry of the moving object or of the supporting carrier, (c) the velocity obtained can be varied over a wide range, independent of the direction, and (d) the engine is powerful enough to allow for the transportation of a cargo.
The proposed engine consists in general of two parts: the supporting carrier and the moving object. Achieving motion of the engine is based on dynamical competition between the two intrinsic lengths of the carrier and the object. This competition is used to transform initially fed energy to directed motion. To exemplify the concept, we use below a simple model system of a chain in a periodic potential, namely a Frenkel-Kontorova type model . But we would like to emphasize that this choice as example is solely motivated by the simplicity of the model rather than by experimental requirements. In particular, the particles are not meant to be single atoms and the springs are not meant to be single chemical bonds. The sole purpose of the model system is to address in a simple manner the following questions: (a) What is the minimal size of the engine? (b) How are the direction and velocity of the motion determined? (c) How powerful is the engine? The important questions of possible physical realizations will be addressed towards the end of the Letter.
In the model system, as already mentioned, the supporting carrier is taken as an isotropic surface, and the moving object as a chain of $`N`$ identical particles on the surface. Each particle $`i`$ has a mass $`m`$ and is located at coordinate $`x_i`$. For simplicity, we restrict the first part of the discussion to translational motion in one dimension; Fig. 1(a) displays a sketch of the model geometry for $`N=3`$. The $`N`$ equations of motion read as
$$m\ddot{x}_i+\eta \dot{x}_i+\frac{\mathrm{\Phi }(x_i)}{x_i}+\underset{\delta _i}{}\frac{\mathrm{\Psi }(x_ix_{i+\delta _i})}{x_{i+\delta _i}}=0i=1,\mathrm{},N.$$
(1)
The second term in eq. (1) describes the dissipative interaction \[friction\] between the particles and the surface and is proportional to their relative velocities with proportionality constant $`\eta `$. The static interaction between the particles and the surface is represented by the periodic potential $`\mathrm{\Phi }(x)=\mathrm{\Phi }_0\mathrm{cos}(2\pi x/b)`$ with periodicity $`b`$. Concerning the inter-particle interaction, we take a nearest neighbor harmonic interaction $`\mathrm{\Psi }(x_ix_{i+\delta _i})=(k/2)[|x_ix_{i+\delta _i}|a_{i,i+\delta _i}(t)]^2`$ with free equilibrium rest lengths $`a_{i,i+\delta _i}(t)`$ \[$`i+\delta _i=i\pm 1`$ denotes the nearest neightbors of particle $`i`$\]. The $`N1`$ rest lengths depend both on the bond’s position specified by the indices $`i,i+\delta _i`$ and on time $`t`$. We demonstrate that if energy is pumped into the system in a specific manner that provides spatially and temporally correlated changes of the lengths $`a_{i,i+\delta _i}(t)`$, the dynamical local competition between the periodicity $`b`$ and the rest lengths $`a_{i,i+\delta _i}(t)`$ can induce a directed motion of the chain.
Without specifing the dependence of the rest lengths $`a_{i,i+\delta _i}(t)`$ on the bond’s position and time, the above approach describes a whole family of atomic scale engines. In the following we restrict ourselves to a certain choice for $`a_{i,i+\delta _i}(t)`$, that results in an engine having a minimum size of as small as $`N=3`$ particles. The position and time dependent rest lengths $`a_{i,i+\delta _i}(t)`$ are chosen as $`a_{i,i+\delta _i}(t)=a[1+\alpha (qx_{i,i+\delta _i}+\omega t)]`$, resulting in a certain fixed spatial and temporal correlation between the lengths of different bonds. Here, $`x_{i,i+1}=ib`$ and $`x_{i,i1}=x_{i1,i}`$ are the relative positions of the bonds between particles $`i`$ and $`i\pm 1`$. The length $`a`$, the wave vector $`q`$, and the driving frequeny $`\omega `$ are parameters. The function $`\alpha (s)`$ has a periodicity of $`1`$ such that $`\alpha (s+1)=\alpha (s)`$ for all $`s`$, and is chosen as $`\alpha (s)=c\mathrm{sin}(\pi s/s_0)`$ for $`0ss_0`$ and $`\alpha (s)=0`$ for $`s_0s1`$, with $`s_0`$ and $`c`$ being parameters. In Fig. 1(b) shown are $`10`$ snapshots of the motion for a complete step of length $`b`$ to the right for the engine sketched in Fig. 1(a) \[note that the $`10^{\mathrm{th}}`$ snapshot is equivalent to the $`1^{\mathrm{st}}`$ as the chain moved by a length $`b`$ to the right\]. The direction of the motion is dynamically determined and solely given by the bond whose rest length increases first, starting when the particles are in the minima of the surface potential \[the right one in Fig. 1(b), cf. \]. Therefore, the motion can be easily controlled; in particular the direction can be chosen independent of velocity, and the motion can be stopped and restarted. The latter can be achieved with the same or with the opposite direction. Such a control of motion is not possible with other engines such as the ratchet systems. For the above choice of $`a_{i,i+\delta _i}(t)`$ and for not too high frequencies up to a maximum frequency $`\omega _{\mathrm{max}}\pi /(25b)\sqrt{\mathrm{\Phi }/m}`$ \[keeping the other parameters fixed\], the engine’s velocity $`v`$ is proportional to the driving frequency $`\omega `$ through $`v=b\omega `$, so that the maximum velocity is approximately given by $`v_{\mathrm{max}}\pi /25\sqrt{\mathrm{\Phi }/m}`$. For higher driving frequencies the motion gets first irregular and finally diffusive, loosing its directionality. The wave vector $`q`$ determines the correlation between the different rest lengths. A choice of $`q=1/(5b)`$ turns out to yield a large maximum velocity and a ‘powerful’ engine for $`N=3`$. A measure for how powerful this engine is, is determined by a simple procedure , where the chain moves against a constant force applied at each particle. The chain with the parameters shown in Fig. 1 is able to move against a constant force of up to $`F_{\mathrm{max}}\mathrm{\Phi }_0/b`$, maintaining the velocity $`b\omega `$. For higher forces, the chain first remains in its initial location, and finally moves with the force. It is important to note that the maximum possible force depends sensitively on the parameters, in particular on the constant $`c`$ and the wave vector $`q`$. Another possible way to measure how powerful the engine is, is to let the chain transport a cargo of $`N^{}`$ additional inactive particles attached at its end with the same inter-particle potential $`\mathrm{\Psi }(x_ix_{i+\delta _i})`$, but with a constant rest length $`a`$. It turns out that, depending on the choice of parameters, a chain of $`N`$ particles is able to transport up to a maximum of $`N_{\mathrm{max}}^{}N/2`$ additional inactive particles and hence up to about half its own weight on an isotropic surface.
So far, our discussion has been restricted to a linear one-dimensional system. However, one of the main advantages of the above concept is that it can easily be applied to other types of motion and generalized to higher dimensionality. Let us first discuss the case
of higher dimensionality, and consider a chain moving on a two-dimensional surface by replacing in eq. (1): (a) the 1D coordinates $`x_i`$ by 2D ones $`\stackrel{}{x}_i`$, (b) the 1D partial derivations $`/x_i`$ by 2D gradients $`\stackrel{}{}_{\stackrel{}{x}_i}`$, (c) the 1D surface potential $`\mathrm{\Phi }(x)`$ by a 2D counterpart $`\mathrm{\Phi }(\stackrel{}{x})=\mathrm{\Phi }_0\mathrm{cos}(\pi [\stackrel{}{x}^{(1)}\stackrel{}{x}^{(2)}]/b)\mathrm{cos}(\pi [\stackrel{}{x}^{(1)}+\stackrel{}{x}^{(2)}]/b)`$ \[$`\stackrel{}{x}^{(i)}`$ denotes the $`i`$-th component of the vector $`\stackrel{}{x}`$\], and (d) the 1D inter-particle potential $`\mathrm{\Psi }(x_ix_{i+\delta _i})`$ by the 2D counterpart $`\mathrm{\Psi }(\stackrel{}{x}_i\stackrel{}{x}_{i+\delta _i})=(k/2)[\left|\stackrel{}{x}_i\stackrel{}{x}_{i+\delta _i}\right|a_{i,i+\delta _i}(t)]^2`$ \[$`||`$ denotes vector length\]. Using the same form for $`a_{i,i+\delta _i}(t)`$ as in the case of 1D, the chain can be moved along the 2D surface. This raises the striking possibility of building a complex atomic scale ‘car’ by connecting, for example, six chains in such a way that they constitute an array of $`3\times 3`$ particles with $`12`$ bonds, see Fig. 2(a) for a sketch of the geometry. By exciting three parallel chains coherently, this ‘car’ can be moved both forward and backward as well as left and right \[using the six vertical or the six horizontal bonds in Fig. 2(a)\], so that it
can be driven freely over the surface. As an example, in Fig. 2(b) shown are $`10`$ snapshots of the motion for a complete step of length $`b`$ to the right \[note that the $`10^{\mathrm{th}}`$ snapshot is equivalent to the $`1^{\mathrm{st}}`$ as the ‘car’ moved by a length $`b`$ to the right\].
Concerning the application to other types of motion, an engine that performs rotational motion can be achieved by treating the coordinates $`x_i`$ as angular coordinates $`x_i[0,\mathrm{})`$ on a circle of circumfence of length $`\mathrm{}=nb`$ with $`n`$ integer, and periodic boundary conditions, see Fig. 3(a) for a sketch of a ring of $`N=3`$ particles and $`\mathrm{}=4b`$. This ‘wheel’ can rotate either clockwise or counterclockwise. As an example, in Fig. 3(b) shown are $`10`$ snapshots of the rotation for a complete step of length $`b`$ counterclockwise \[note that the $`10^{\mathrm{th}}`$ snapshot is equivalent to the $`1^{\mathrm{st}}`$ as the ‘wheel’ rotated by a length $`b`$ counterclockwise; for the four-fold symmetry of the example, this is equivalent to a rotation with an angle $`\pi /2`$ counterclockwise\].
To demonstrate the robustness of our example engines and the concept in general, we should stress a few important points. First, it is clear that the motion of the chain is not affected by the integer part of the free rest lengths $`a_{i,i+\delta _i}(t)/b`$ between the particles or ‘feet’ that touch the
surface. In the above examples in Figs. 1 and 2, we have chosen the rest lengths $`a_{i,i+\delta _i}(t)`$ to oscillate between $`11b/10`$ and $`187b/100`$ \[a relative change by $`187/110`$\], but the same motion occurs for $`a_{i,i+\delta _i}(t)`$ oscillating between $`21b/10`$ and $`287b/100`$ \[a relative change by $`287/210`$\], between $`31b/10`$ and $`387b/10`$ \[a relative change by $`387/310`$\], and so on. Hence, the relative change can be made arbitrarily small. A second possibility to reduce the required stretching of the bond needed to cause a directed motion is to divide the $`N1`$ bonds into $`g`$ groups \[$`N`$ bonds in the case of rotational motion\], each containing $`(N1)/g`$ bonds \[$`N/g`$ bonds in the case of rotational motion\], by setting $`x_{i,i+1}=\mathrm{int}(i/g)b`$ \[$`\mathrm{int}(x)`$ denotes the largest integer $`n`$ with $`nx`$\]. For instance, for the parameters used for Figs. 1 and 2, but with $`N=10`$ particles \[i.e. $`9`$ bonds\] and $`g=3`$ groups, each containg $`3`$ bonds, it is sufficient that the free rest lengths $`a_{i,i+\delta _i}(t)`$ oscillate between $`b`$ and $`142b/100`$ \[instead of $`11b/10`$ and $`187b/100`$\] to cause a directed translational motion with the same velocity $`b\omega `$.
One important question is that of the influence of the shape of the ‘excitation’ $`\alpha (s)`$, which is sinusodial in our case. As long as the ‘excitation’ is approximately trapezoidal, a proper choice of driving frequency $`\omega `$ and wave vector $`q`$ results in a directed motion. Since eq. (1) is deterministic, another important question is that of the influence of noise. As long as the thermal energy $`k_\mathrm{B}T`$ is much smaller than the energy scale $`\mathrm{\Phi }_0`$, the engines motion is hardly influenced by noise. If the fluctuations become larger, the motion gets erratic. Finally, for $`k_\mathrm{B}T\mathrm{\Phi }_0`$, the motion looses completly its directionality and becomes diffusive.
In comparison with other existing approaches to translational atomic scale engines, e.g. the ratchet systems mentioned earlier, or other models that use a spring with a time-dependent length in combination with static/dynamic friction or with a spatial asymmetric potential , the current approach has a number of advantages. In particular, our atomic scale engines can operate as microscopic ‘shuttles’ or ‘trucks’ that transport a cargo, a property that is absent in all other proposed schemes. In addition, the engines proposed here are not bound to a one-dimensional track and can be applied to different geometries and higher dimensionality.
As a conclusion, we would like to emphasize that the choice of a chain on a surface as the example engine has been motivated solely by the simplicity of the model rather than by experimental requirements. One important feature of the concept introduced here is that it does not impose any specific length or time scales. This means that it is applicable not only for an engine on an atomic scale, but also for mesoscopic or even macroscopic sizes. Hence, the basic concept of competing lengths is more general than presented here and might be applicable in a wide range of situations. A possible realization on an atomic or mesoscopic scale is by building the moving object using nanosize clusters \[the ‘particles’\] and photochromic molecules \[the ‘bonds’\]. The time dependence of the rest lengths, which is a crucial part of the model, can be provided by individually controlling the ‘bonds’ by light induced conformational changes of the chromophors. Using different chromophores which respond to different wavelengths should allow to specifically excite ‘bonds’ at chosen locations at given times. In this proposed realization the surface corrugation can be prepared by nanolithography. We believe therefore that the current concept is simple and robust enough so that it can be realized in actual experiments using already existing techniques. Such an atomic scale engine, when realized experimentally, will be a breakthrough in the field of nanotechnology in providing new ways to manipulate molecules and clusters.
Financial support from the Israel Science Foundation, the German Israeli Foundation, and DIP and SISITOMAS grants is gratefully acknowledged. M.P. gratefully acknowledges the Alexander von Humboldt Foundation (Feodor Lynen program) for financial support.
|
warning/0006/math0006098.html
|
ar5iv
|
text
|
# Theorem(0.1)
Ergodicity of Mapping Class Group Actions
on Representation Varieties, I. Closed Surfaces
Doug Pickrell and Eugene Z. Xia
Department of Mathematics
University of Arizona
Tucson, AZ 85721
pickrellmath.arizona.edu xiamath.umass.edu
Abstract. We prove that the mapping class group of a closed surface acts ergodically on connected components of the representation variety corresponding to a connected compact Lie group.
§0. Introduction.
Throughout this paper we fix a connected compact Lie group $`K`$, and we let $`dg`$ denote the unique normalized Haar measure on $`K`$.
Let $`\mathrm{\Sigma }`$ denote a closed oriented surface with a fixed basepoint, and let $`\mathrm{\Gamma }_\mathrm{\Sigma }=\pi _0(Aut(\mathrm{\Sigma }))`$, the mapping class group. The representation variety $`Hom(\pi _1\mathrm{\Sigma },K)`$ has a canonical $`\mathrm{\Gamma }`$<sub>Σ</sub>-invariant measure class, the Lebesgue class of the set of nonsingular points, and it is well-known that this class is represented by a $`\mathrm{\Gamma }_\mathrm{\Sigma }`$-invariant measure (\[AB\], \[Go1\]). Our aim is to prove the following
###### Theorem(0.1)
The group $`\mathrm{\Gamma }_\mathrm{\Sigma }`$ acts ergodically on the Lebesgue class of each connected component of $`Hom(\pi _1\mathrm{\Sigma },K)`$.
Let $`H^1(\mathrm{\Sigma },K)`$ denote the moduli space of representations, i.e. the quotient of $`Hom(\pi _1\mathrm{\Sigma },K)`$ by the conjugation action of $`K`$. The following was proven by Goldman for $`K`$ locally isomorphic to $`SU(2)\times T`$, where $`T`$ is a torus, in \[Go2\].
###### Corollary(0.2)
The group $`\mathrm{\Gamma }_\mathrm{\Sigma }`$ acts ergodically on the Lebesgue class of each connected component of $`H^1(\mathrm{\Sigma },K)`$. Thus the Lebesgue class of each component has an essentially unique $`\mathrm{\Gamma }_\mathrm{\Sigma }`$-invariant representative, the canonical symplectic volume element.
As in most problems involving representation varieties, we use sewing techniques. Our basic idea is to prove $`(0.1)`$ in initial cases, the one and two-holed tori, for $`a.e.`$ boundary condition; this is easier than dealing with every boundary condition, because we can use harmonic analysis for $`K`$. When we sew, because we integrate, the measure-theoretic ambiguity is washed away, and we obtain the Theorem.
The space $`H^1(\mathrm{\Sigma },K)`$ is $`\mathrm{\Gamma }_\mathrm{\Sigma }`$-equivariantly filtered, the filter corresponding to a (conjugacy class of a) closed subgroup $`K_1`$ consisting of those representations which have image in $`K_1`$. The filter corresponding to $`K_1`$ is isomorphic to $`H^1(\mathrm{\Sigma },K_1)`$, modulo the action of the normalizer of $`K_1`$ inside $`K`$; it has a canonical $`\mathrm{\Gamma }_\mathrm{\Sigma }`$-invariant measure class, and each such class is represented by an invariant measure. Our result implies that $`\mathrm{\Gamma }_\mathrm{\Sigma }`$ acts ergodically on connected components for classes corresponding to connected subgroups. We have not addressed the case of nonconnected $`K`$; in this case $`\mathrm{\Gamma }_\mathrm{\Sigma }`$ does not generally act trivially on $`\pi _0`$ of the moduli space (consider a finite group), and it seems difficult to make an enlightening statement.
The symplectic volume element on $`H^1(\mathrm{\Sigma },K)`$ is the temperature $`T0`$ limit of the 2-dimensional Yang-Mills measure $`d\nu _{\frac{1}{T}YM_2}`$ on the space of gauge equivalence classes of all (generalized) $`K`$-connections. Elsewhere we will prove that this measure is ergodic with respect to its symmetry group, the group of area-preserving diffeomorphisms.
(0.2) Notation. Given a Lie group $`G`$, we will always use left translation to trivialize the tangent bundle:
$$TGG\times 𝔤:v|_g(L_{g^1})_{}(v|_g).$$
$`0.3`$
In this frame the commutator of two vector fields $`x,y:G𝔤`$ is given by
$$[x,y]|_g=dy(x)|_gdx(y)|_g+[x(g),y(g)].$$
$`0.4`$
The adjoint action $`Ad:G\times 𝔤𝔤`$ is abbreviated to $`Ad_g(x)=x^g`$. If $`G`$ acts on a space $`X`$, then $`X^g`$ denotes the fixed point set.
§1. Basic Notions and Sewing.
For the purposes of this paper, we will need to consider a somewhat nonstandard kind of boundary condition for surfaces with boundary.
Consider a connected compact oriented surface $`\mathrm{\Sigma }`$ equipped with a basepoint, and the following additional structure: each boundary component is linked to the basepoint by a path, and each boundary component $`c`$ is labelled with a $`+`$ or $``$, and a group element $`k_c`$ of $`K`$. We interpret the sign to mean that the boundary component $`c`$ has an intrinsic orientation that agrees, or disagrees, with the induced orientation from $`\mathrm{\Sigma }`$; the intrinsic orientation of the boundary component $`c`$ gives us a preferred generator for $`\pi _1(c)\pi _1(\mathrm{\Sigma })`$, which, by slight abuse of notation, we will also denote by $`c`$. We define
$$Hom(\mathrm{\Sigma },K)=\{gHom(\pi _1(\mathrm{\Sigma }),K):g|_c=k_c,c\pi _0(\mathrm{\Sigma })\}.$$
$`1.1`$
This space only depends upon the basepoint and paths to the boundary components up to homotopy. The pure mapping class group $`\mathrm{\Gamma }_\mathrm{\Sigma }`$ does not in general act on this space; only the subgroup generated by Dehn twists along curves which do not cross the paths from the basepoint to the boundary components will act; we denote this group by $`\pi _0(Aut(\mathrm{\Sigma }))`$.
If $`\mathrm{\Sigma }`$ is a closed surface, then we can form the quotient of $`Hom(\mathrm{\Sigma },K)`$ by the global gauge action of $`K`$ by conjugation; the quotient is denoted by $`H^1(\mathrm{\Sigma },K)`$. In this case, $`\pi _0(Aut(\mathrm{\Sigma }))=\mathrm{\Gamma }_\mathrm{\Sigma }`$, the mapping class group.
Let $`s`$ denote a separating oriented simple closed curve on $`\mathrm{\Sigma }`$. We suppose that the basepoint is on $`s`$, and we suppose also that $`s`$ does not cut any of the paths from the basepoint to the boundary components. Let $`\stackrel{ˇ}{\mathrm{\Sigma }}_k=\mathrm{\Sigma }_k^{}\mathrm{\Sigma }_k^+`$ denote the disconnected object obtained by cutting along $`s`$ and attaching one $``$ and one $`+`$, and same group element $`k`$, to the new boundary components. The Seifert-Van Kampen Theorem implies that the projection $`p:\stackrel{ˇ}{\mathrm{\Sigma }}\mathrm{\Sigma }`$ induces an exact sequence
$$0s\pi _1(\mathrm{\Sigma }^{})\pi _1(\mathrm{\Sigma }^+)\mathrm{@}>p_{}>>\pi _1(\mathrm{\Sigma })0.$$
$`1.2`$
where $`<s>`$ denotes the normal subgroup generated by the element $`s^1s`$. Hence we have the following elementary
###### Sewing Lemma(1.3)
Assume $`\mathrm{\Sigma }`$ has a group element boundary condition. Then there is a bijective correspondence
$$Hom(\mathrm{\Sigma },K)=\underset{kK}{}Hom(\mathrm{\Sigma }_k^{},K)\times Hom(\mathrm{\Sigma }_k^+,K),$$
where $`g(g^{},g^+)`$, $`g^{}(s)=g^+(s)=k`$, $`g^\pm =g|_{\pi _1(\mathrm{\Sigma }^\pm )}`$. This correspondence is equivariant with respect to $`\pi _0(Aut(\stackrel{ˇ}{\mathrm{\Sigma }}))`$, the group generated by Dehn twists along curves which cross neither $`s`$ nor the paths from basepoint to boundary components.
§2. Initial Cases.
The basic insight of this paper is that in all cases involving boundary, $`\mathrm{\Gamma }`$-ergodicity is equivalent to $`𝒢`$-ergodicity, where $`𝒢`$ is a $`continuous`$ group of volume-preserving transformations, for $`a.e.`$ boundary condition. The latter problem reduces to a calculation concerning infinitesimal transitivity.
§2.1. The one-holed torus, with group element boundary condition.
In this subsection we let $`\mathrm{\Sigma }`$ denote the one-holed torus, with boundary component $`c`$, which we view as in Figure 1.
Given $`kK`$, we write $`\mathrm{\Sigma }_k`$ to indicate that we impose the boundary condition $`k`$, so that $`\mathrm{\Sigma }_k`$ is an object of the type considered in §1. We have
$$\begin{array}{cccc}Hom(\pi _1\mathrm{\Sigma },K)& & K\times K& (g(g_\alpha ,g_\beta ))\\ & & p\\ & & K^{}=[K,K]\end{array}$$
$`\mathrm{2.1.2}`$
where $`p`$ is the commutator map, $`p(g,h)=ghg^1h^1`$. With respect to this identification, the fibers of $`p`$ are precisely the representation spaces $`Hom(\mathrm{\Sigma }_k,K)`$. Define $`\mathrm{\Gamma }`$ to be the group generated by the transformations $`T_j:K\times KK\times K`$ given by
$$T_1(g,h)=(gh^1,h),T_2(g,h)=(g,hg^1).$$
$`\mathrm{2.1.3}`$
These transformations arise from twists along the curves $`s_1`$ and $`s_2`$ indicated in Figure 1; they are volume-preserving (with respect to Haar measure), hence naturally induce unitary transformations of $`L^2(K\times K)`$, they commute with conjugation by $`K`$, and they fix the map $`p`$. The action of $`\mathrm{\Gamma }`$ restricts to the action of $`\pi _0(Aut(\mathrm{\Sigma }_k))`$ on the fiber $`p^1(k)=Hom(\mathrm{\Sigma }_k,K)`$.
In this subsection we will prove the following result, which is of independent interest.
###### (2.1.4)Theorem
Suppose that $`FL^2(K\times K)`$ is $`\mathrm{\Gamma }`$-invariant. Then $`F`$ is $`a.e.`$ constant on components of $`p^1(k)`$ for $`a.e.`$ $`k`$ $`[d\rho ]`$, where $`d\rho =p_{}(dg\times dh)`$.
Remarks (2.1.5). (a) The measure $`d\rho `$ is in the Lebesgue class of $`K^{}`$, and
$$d\rho (k)=(\underset{\mu }{}d_\mu ^1\chi _\mu (k))dk,$$
$`\mathrm{2.1.6}`$
where the sum is over all irreducible characters of $`K^{}`$, and $`d_\mu =\chi _\mu (1)`$. To see this, first note that because $`p`$ is a $`conj(K)`$-equivariant map, and $`dg\times dh`$ is conjugation invariant, $`d\rho `$ is conjugation invariant. Secondly, if $`f=c_\mu \chi _\mu `$ is a central function, then
$$f𝑑\rho =_K^{}_K^{}f(ghg^1h^1)𝑑g𝑑h=\underset{\mu }{}c_\mu \{\chi _\mu (ghg^1h^1)𝑑h\}𝑑g$$
$$=c_\mu \frac{|\chi _\mu (g)|^2}{\chi _\mu (1)}𝑑g=(c_\nu \chi _\nu (g))(d_\mu ^1\chi _\mu (g^1))𝑑g,$$
which heuristically explains (2.1.6) (The third equality uses the well-known integration formula $`\chi (xhyh^1)𝑑h=\chi (x)\chi (y)/\chi (1)`$, which follows from observing that the left hand side is a central function for $`(x,y)K\times K`$, and computing the expansion in terms of characters for $`K\times K`$). Because characters are orthogonal,
$$|d_\mu ^1\chi _\mu (g)|^2𝑑g=d_\mu ^2|\chi _\mu (g)|^2𝑑g=d_\mu ^1.$$
The Weyl dimension formula implies that this sum is finite, provided $`𝔨`$ does not have $`su(2)`$ factors (see below); hence in most cases, the density in (2.1.6) represents an $`L^2`$ function on $`K`$. In general, if we fix a maximal torus and positive Weyl chamber, so that we can parameterize the representations by dominant integral functionals $`\mu `$, then the Weyl character formula implies that for $`gT`$,
$$d_\mu ^1\chi _\mu (g)=\mathrm{\Delta }(g)^1\underset{wW}{}(1)^{l(w)}\{e^{i\rho }\underset{\mu }{}d_\mu ^1e^{i\mu }\}^w(g),$$
where $`W`$ is the Weyl group, $`l(w)`$ is the length of $`w`$, and $`\rho `$ is half the sum of the positive roots. If we write $`\mu `$ in terms of the fundamental dominant integral functionals, $`\mu =n_j\mu _j`$, then the Weyl dimension formula implies
$$d_\mu =\underset{\alpha >0}{}\frac{\mu +\rho ,\alpha }{\rho ,\alpha }\underset{j}{}n_j^{\mu _j,2\rho },$$
so that $`d_\mu ^1e^{i\mu }=d_\mu ^1e^{i{\scriptscriptstyle n_j\theta _j}}`$ always represents an $`L^2`$ function with respect to the Haar measure of $`T`$. Since this function is the boundary values of a holomorphic function on $`(_1)^r`$, it cannot vanish on a set of positive measure. This explains the meaning of the density (2.1.6), and shows that $`d\rho `$ is in the Lebesgue class (see Appendix B for a more direct proof).
In the case of $`K=SU(2)`$, the density in (2.1.6), as a function of $`diag(z,z^1)`$, $`z=e^{i\theta }`$, is given by
$$\frac{1}{zz^1}\{z\underset{d1}{}d^1zz^1\underset{d1}{}d^1z^d\}=Im\{\frac{e^{i\theta }ln(1e^{i\theta })}{sin\theta }\}$$
$$=ln(22cos\theta )^{1/2}+cos\theta \frac{arg(1cos\theta isin\theta )}{sin\theta }.$$
(b) Theorem (2.1.4) gives an algebraic characterization of functions $`F(g,h)`$ which have the form $`f(ghg^1h^1)`$, in situations where all the fibers $`p^1(k)`$ are connected, e.g. for $`K`$ simply connected (see Appendix A). It seems to be unknown whether there might be a reasonable characterization for more general groups (e.g. finite groups).
###### (2.1.7)Corollary of (2.1.4)
For $`a.e.`$ boundary condition $`kK`$, the action
$$\pi _0(Aut(\mathrm{\Sigma }_k))\times Hom(\mathrm{\Sigma }_k,K)$$
is ergodic on the Lebesgue class of each component.
To prove (2.1.4), we need to be able to analyze the transformations in (2.1.3). Let $`T=T_2`$ denote the unitary transformation on $`L^2(K_1\times K_2)`$ corresponding to the second of these transformations, where we have introduced copies $`K_1`$ and $`K_2`$ of $`K`$, for notational clarity. Recall that the Peter-Weyl Theorem asserts that there is a $`K\times K`$-equivariant isomorphism
$$\underset{\mu }{}(V_\mu )L^2(K):(L_\mu )f,f(g)=\underset{\mu }{}dimn(\mu )^{1/2}tr_\mu (L_\mu \pi _\mu (g^1)),$$
$`\mathrm{2.1.8}`$
where $`(V_\mu )`$ denotes the space of linear transformations of $`V_\mu `$, the sums are over all irreducible representations, and the linear action of $`(g_l,g_r)K\times K`$ on these respective spaces is given by
$$L_\mu \pi _\mu (g_l)L_\mu \pi _\mu (g_r)^1$$
$$f(g)f(g_l^1gg_r).$$
###### (2.1.9)Lemma
Via the isomorphisms
$$L^2(K_1\times K_2)=L^2(K_1;L^2(K_2))=\underset{\mu }{}L^2(K_1;(V_\mu )),$$
$`T=diag(T_\mu )`$, where $`T_\mu `$ is the multiplication operator
$$T_\mu :L^2(K_1;(V_\mu ))L^2(K_1;(V_\mu )):F_\mu (g)F_\mu (g)\pi _\mu (g)^1.$$
$`\mathrm{2.1.10}`$
In particular
$$L^2(K_1;(V_\mu ))^{T_\mu }=\{F_\mu :F_\mu (g)|_{(V_\mu ^g)^{}}=0,a.e.g\},$$
$`\mathrm{2.1.11}`$
and if $`F_\mu `$ is $`T_\mu `$-invariant, then (viewed now as a function of two variables)
$$F_\mu (g,h)=F_\mu (g,ha(g)^1),$$
$`\mathrm{2.1.12}`$
for any measureable function $`a:KK`$ such that $`[a(g),g]=1`$ $`a.e.`$.
Proof of (2.1.9). The formula for $`T_\mu `$ is a direct consequence of the Peter-Weyl theorem, and the other statements follow directly from the formula for $`T_\mu `$.//
Note that
$$A=\{a:KK:[g,a(g)]=1,g\}$$
$`\mathrm{2.1.13}`$
is an abelian subgroup of the gauge group $`Map(K,K)`$. We will assume that the maps in $`A`$ are smooth, unless noted otherwise. It is probably not the case that $`A`$ is a Lie subgroup, because the family of projections onto the subalgebras $`𝔨^g`$, $`gK`$, is not smooth. Nonetheless we will refer to
$$𝔞=\{x:K𝔨:Ad_g(x(g))=x(g),g\}.$$
$`\mathrm{2.1.14}`$
as the Lie algebra of $`A`$, because it has the crucial property
$$exp(𝔞)A.$$
The group $`A`$ acts on $`K_1\times K_2`$ in two ways, corresponding to the actions (2.1.3), by
$$A_1(a):(g,h)(ga(h)^1,h),A_2(a):(g,h)(g,ha(g)^1),$$
$`\mathrm{2.1.15}`$
respectively. Note that the transformations $`T_i^n`$, $`i=1,2`$, correspond to $`a(k)=k^n`$. Note also that the transformations (2.1.15) are volume-preserving.
We can restate (2.1.9) as
###### Lemma(2.1.16)
The $`L^2`$ function $`F(g,h)`$ is $`T_j`$-invariant if and only if $`F`$ is $`A_j`$-invariant, for $`j=1,2`$ (Here we can require the maps in $`A`$ to be $`C^{\mathrm{}}`$, $`C^0`$, or merely measureable - the basic result is insensitive to this requirement).
Let $`𝒢`$ denote the closure of the group of volume-preserving transformations of $`K\times K`$ generated by $`A_1`$ and $`A_2`$, inside the Lie group of all volume-preserving diffeomorphisms of $`K\times K`$ (it will turn out that, for our purposes, we could just as well consider the closure in the group of all volume-preserving transformations, in the natural strong operator topology). It is unclear whether $`𝒢`$ is a Lie group, but it is useful to think in these terms, as we will now see. The Lie algebra actions corresponding to (2.1.15) are given by the vector fields on $`K_1\times K_2`$
$$dA_1(x)|_{g,h}=(x(h),0),dA_2(x)|_{g,h}=(0,x(g)),$$
$`\mathrm{2.1.17}`$
respectively, for $`x𝔞`$. These actions do not necessarily commute.
###### Definitions(2.1.18)
(a) $`𝔤_0`$ is the Lie algebra of vector fields on $`K\times K`$ given by
$$𝔤_0=\{(x|_h,y|_g):x,y𝔞\};$$
(b) $`𝔤`$ is the Lie algebra of vector fields on $`K\times K`$ generated by the family of Lie algebras
$$\{Ad_\sigma 𝔤_0:\sigma A_1orA_2\}.$$
The bracket for $`𝔤_0`$ is given by
$$[(x_1,y_1),(x_2,y_2)]|_{g,h}=(dx_2(y_1)|_hdx_1(y_2)|_h,dy_2(x_1)|_gdy_1(x_2)|_g).$$
$`\mathrm{2.1.19}`$
(see (0.2)).
Heuristically $`𝔤_0`$ is the Lie algebra corresponding to the group generated by the identity components of $`A_1`$ and $`A_2`$, while heuristically $`𝔤`$ is the Lie algebra corresponding to $`𝒢`$. In practice we will think of $`𝔤`$ as an $`Ad(\mathrm{\Gamma })`$-invariant Lie algebra containing $`𝔤_0`$.
###### Lemma(2.1.20)
Assuming we require maps to be $`C^{\mathrm{}}`$, we have $`exp(𝔤)𝒢`$.
Proof of (2.1.20). Suppose that $`\xi =(y,x)𝔤_0`$. Now $`exp\{t(0,x)\}A_2`$ and $`exp\{t(y,0)\}A_1`$, $`t`$. Thus
$$exp(\xi )=\underset{n\mathrm{}}{lim}(exp((y/n,0)exp(0,x/n))^n𝒢,$$
$`\mathrm{2.1.21}`$
because Trotter’s product formula is valid for vector fields on a compact manifold. Therefore for any $`\sigma A_j`$,
$$exp(Ad_\sigma (\xi ))=\sigma exp(\xi )\sigma ^1𝒢.$$
$`\mathrm{2.1.22}`$
Using Trotter’s product formula (and the analogue for brackets) in the same way, we see that for sums and brackets of such vector fields, we again exponentiate into $`𝒢`$.//
Our goal now is to show that the Lie algebra $`𝔤`$ is infinitesimally transitive off a set of codimension $`>1`$ along a generic fiber of the commutator map $`p`$. We calculate that
$$dp|_{g,h}:𝔨𝔨𝔨^{}:(\xi ,\eta )\xi ^{hgh^1}\xi ^{hg}+\eta ^{hg}\eta ^h,$$
$$=(\xi ^{h^1}\xi +\eta \eta ^{g^1})^{hg}.$$
$`\mathrm{2.1.23}`$
###### Proposition(2.1.24)
For $`g`$ and $`h`$ in the complement of a set of codimension $`>1`$, the evaluation map
$$eval|_{g,h}:𝔤ker(dp|_{(g,h)})$$
is surjective, where $`eval`$ is the evaluation map.
Proof of (2.1.24). If $`K`$ is abelian, then at all points
$$eval|_g:𝔞𝔨^{Ad(g)}$$
$`\mathrm{2.1.25}`$
is surjective, and it follows from this that $`𝔤`$ is transitive.
So suppose that $`K`$ is nonabelian. Recall the set of regular points,
$$K^{reg}=\{kK:dimn(𝔨^g)=r\}$$
$`\mathrm{2.1.26}`$
where $`r=rank(𝔨)`$ is the minimal possible dimension of $`𝔨^g`$. The singular set $`KK^{reg}`$ has codimension 3, because for a nonregular point $`g`$, $`𝔨^g`$ always contains a copy of $`su(2)`$, in addition to a maximal torus. For a regular point $`gK`$, $`eval|_g`$ in (2.1.25) will be surjective (while the image shrinks at nonregular points, e.g. $`eval|_1(𝔞)=\{0\}`$); this follows from the real analyticity of the vector bundle $`g𝔨^g`$ over $`K^{reg}`$. Therefore for $`g,hK^{reg}`$,
$$eval|_{g,h}(𝔤_0)=𝔨^h𝔨^gker(dp|_{g,h})𝔨𝔨.$$
$`\mathrm{2.1.27}`$
This always fills out the central part of $`𝔨`$. For this reason, without loss of generality, we can henceforth assume that $`𝔨`$ is semisimple.
The map $`p`$ is regular at all points $`(g,h)`$ such that $`𝔨^g𝔨^h=\{0\}`$, by (2.1.23). The abstract meaning of this condition is that the representation of $`\pi _1(\mathrm{\Sigma })`$ determined by $`(g,h)`$ is irreducible, in the intrinsic sense that the commutant of the image in $`K`$ is the center of $`K`$. To understand this condition more concretely (from a point of view useful to us), suppose that $`g`$ and $`h`$ are regular. Write $`h=exp(y)`$, so that $`𝔨^h=C_𝔨(y)`$, the centralizer of $`y`$. For $`x𝔨^g`$,
$$[x,y]=\underset{\alpha }{}\alpha (x)y_\alpha $$
$`\mathrm{2.1.28}`$
where the sum is over all roots $`\alpha `$ of $`𝔨^g`$, and $`y_\alpha `$ denotes the component of $`y`$ in the root space of $`\alpha `$. In order for $`x𝔨^h`$, we must have $`\alpha (x)=0`$, whenever $`y_\alpha 0`$. The condition $`y_\alpha =0`$ is two independent real conditions, because the root space has one complex dimension. Thus $`\{(g,h):𝔨^g𝔨^h\{0\}\}`$ has codimension at least 2.
We now know that the dimension of $`ker(dp)`$ is $`dimn(𝔨)`$ off a set of codimension 2. Let $`proj_i`$ denote projection onto the $`i`$th factor. The map $`proj_1`$ induces an exact sequence
$$0\{(0,𝔨^g)\}ker(dp)\mathrm{@}>proj_1>>\{\xi 𝔨:(1Ad(h^1))\xi (𝔨^g)^{}\}0;$$
there is a similar sequence for $`proj_2`$. The evaluation of $`𝔤_0`$ at $`(g,h)K^{reg}\times K^{reg}`$ fills out $`ker(proj_1)`$$`+`$$`ker(proj_2)`$. Since
$$\{\xi 𝔨:(1Ad(h^1))\xi (𝔨^g)^{}\}^{}=(1Ad(h))𝔨^g,$$
to prove that $`eval:𝔤ker(dp)`$ is surjective at a regular point, it suffices to prove that
$$(1Ad(h))𝔨^g+proj_1(eval|_{g,h}(𝔤))=𝔨;$$
$`\mathrm{2.1.29}`$
there is a similar statement for $`proj_2`$.
Now $`𝔤`$ is $`\mathrm{\Gamma }`$-invariant, hence
$$\underset{\mathrm{\Gamma }}{}\gamma _{}(eval|_{\gamma ^1(g,h)}(𝔤_0))eval|_{g,h}(𝔤).$$
$`\mathrm{2.1.30}`$
In geometric terms, the sum is the $`\mathrm{\Gamma }`$-invariant distribution generated by $`𝔤_0`$. We will first consider only a small part of this sum, namely the $`T_2`$-invariant distribution generated by $`𝔤_0`$.
Suppose that $`(y|_h,x|_g)𝔤_0`$. We have
$$(T_2^n)_{}(eval|_{T_2^n(g,h)}(y,x))=$$
$$=(0,x|_g)+\frac{d}{dt}|_{t=0}(ge^{ty(hg^n)},hg^n(ge^{ty(hg^n)})^n)$$
$$=\{\begin{array}{c}(y(hg^n),x(g)_{k=1}^ny(hg^n)^{g^k}),n>0\\ (y(hg^n),x(g)+_{k=n+1}^0y(hg^n)^{g^k}),n<0\end{array}.$$
$`\mathrm{2.1.31}`$
From this we see that
$$\underset{hg^nK^{reg}}{}𝔨^{hg^n}proj_1(eval|_{g,h}(𝔤)).$$
$`\mathrm{2.1.32}`$
Now for $`(g,h)K^{reg}\times K^{reg}`$, if $`\{g^n\}`$ is a dense subgroup of $`T=exp(𝔱)`$, then
$$\underset{\{n:hg^nK^{reg}\}}{}𝔨^{hg^n}=\underset{\{x𝔱:he^xK^{reg}\}}{}𝔨^{he^x}.$$
$`\mathrm{2.1.33}`$
For we clearly have $``$. Conversely given $`x𝔱`$ such that $`he^x`$ is regular, we can find a sequence $`\{n_j\}`$ such that $`g^{n_j}e^x`$ as $`j\mathrm{}`$, hence $`hg^{n_j}`$ will be regular for $`j`$ sufficiently large, and $`𝔨^{hg^{n_j}}𝔨^{he^x}`$, so that the opposite inclusion holds.
###### Lemma(2.1.34)
There is a set $`X_2K`$ of codimension $`2`$ such that for $`(g,h)K\times X_2`$,
$$(1Ad(h))𝔨^g+\underset{\{x𝔨^g:he^xK^{reg}\}}{}𝔨^{he^x}=𝔨.$$
$`\mathrm{2.1.35}`$
Proof of (2.1.34). We write $`h=exp(y)`$. We also write $`𝔱=𝔨^g`$. Since $`h`$ is regular, there are open neighborhoods $`𝔲`$ and $`U`$ of $`y`$ and $`h`$, respectively, such that $`exp:𝔲U`$ is an isomorphism; let $`log`$ denote the inverse. There is a Taylor series expansion of the form
$$log(he^x)=\underset{n0}{}c_n(h,x),$$
$`\mathrm{2.1.36}`$
where $`c_n`$ is homogeneous of degree $`n`$ in $`x`$. If $`|ad(y)|<\pi `$, where $`||`$ denotes the operator norm, then we can also expand each $`c_n`$, and the form of these expansions can be read off from the Baker-Campbell-Hausdorff formula, namely $`c_0=y`$,
$$c_1=x+\frac{1}{2}[y,x]+\frac{1}{12}[y,[y,x]]+..=x+O(|y|),$$
$`\mathrm{2.1.37}`$
and for $`n>1`$,
$$c_n=constantad(x)^n(y)+o(|y|),$$
$`\mathrm{2.1.38}`$
as $`|y|0`$, where the constant depends only upon $`n`$. We also have
$$(1Ad(h))x=[x,y]+o(|y|)as|y|0.$$
$`\mathrm{2.1.39}`$
We now claim that the sum in (2.1.35) equals
$$=(1Ad(h))𝔱+𝔨^h+span\{log(he^x):x𝔱,he^xUK^{reg}\}$$
$$=(1Ad(h))𝔱+𝔨^h+span\{c_n(h,x):n0,x𝔱\}.$$
$`\mathrm{2.1.40}`$
The first equality is immediate; the second follows from the fact that the span of the power series (2.1.36) will contain the span of the coefficients (replace $`x`$ by $`sx`$, note that $`he^{sx}UK^{reg}`$ for small $`s`$, and differentiate with respect to $`s`$ at $`s=0`$).
Now we first show that (2.1.35) holds for $`hK^{reg}`$ where $`|y|`$ is small. As in (2.1.28), $`𝔨^h=C_𝔨(y)`$, and we can write
$$y=y_𝔱+\underset{\alpha }{}y_\alpha ,$$
$`\mathrm{2.1.41}`$
relative to the root decomposition of $`𝔨^{}`$ with respect to $`𝔱`$. If all the $`y_\alpha 0`$, then together $`𝔱`$ and $`\{ad(x)^n(y):x𝔱,n1\}`$ will span $`𝔨`$. But (2.1.37), (2.1.38) and (2.1.39) now imply that $`\{c_n(h,x):n1,x𝔱\}`$ and $`(1Ad(h))𝔱`$ will span $`𝔨`$, provided that $`|y|`$ is small.
Note that the condition $`y_\alpha =0`$ is linear, and of codimension 2, as we pointed out below (2.1.28). Thus for $`h`$ in a subset of codimension $`2`$ in a neighborhood of $`1`$, the equation in (2.1.35) holds. It remains to do a similar analysis for a neighborhood of a point $`e^{y_0}1`$.
Suppose that $`h=e^{y_0+z}`$, where $`log(e^{y_0})=y_0`$ and $`z`$ is small. We have $`c_0(h,x)=y_0+z`$,
$$c_1(h,x)=c_1(e^{y_0},x)+\frac{1}{2}[z,x]+o(|z|)=c_1(e^{y_0},x)+O(|z|)$$
$`\mathrm{2.1.42}`$
$$(1Ad(h))x=(1Ad(e^{y_0}))x+[x,z]+o(|z|)$$
$`\mathrm{2.1.43}`$
$$c_n(h,x)=c_n(e^{y_0},x)+constantad(x)^n(z)+o(|z|)$$
$`\mathrm{2.1.44}`$
as $`|z|0`$, where $`n>1`$. The derivative of $`𝔨^h`$ (as it varies in the Grassmannian of subspaces, $`Gr(r,𝔨)`$) is a linear transformation $`T(z):C(y_0)C(y_0)^{}`$. If $`T(z)(\xi _0)=\xi _1`$, then to first order in $`s`$, $`exp(ad(y_0+sz))(\xi _0+s\xi _1)`$$`=`$$`\xi _0+s\xi _1`$, i.e. $`[y_0,\xi _1]+[z,\xi _0]=0`$; in terms of the root decomposition for $`C(y_0)`$, we have
$$T(z)(\xi )=\underset{\beta }{}\frac{\beta (\xi )}{\beta (y_0)}z_\beta .$$
$`\mathrm{2.1.45}`$
Thus
$$𝔨^h=graph(T(z):C(y_0)C(y_0)^{})+o(|z|)$$
$`\mathrm{2.1.46}`$
as $`|z|0`$.
Now consider the possibility that together $`𝔨^{exp(y_0)}`$, $`(1Ad(e^{y_0}))𝔱`$, and $`\{c_n(e^{y_0},x):n0,x𝔱\}`$ do not span $`𝔨`$. The argument proceeds initially as in the case $`exp(y_0)=1`$. If $`\alpha `$, $`z_\alpha 0`$ (the components with respect to the root decomposition for $`𝔱`$), then $`\{ad(x)^n(z):n1,x𝔱\}`$ will span $`𝔱^{}`$. We now use (2.1.42)-(2.1.44). For the variation of the span of $`𝔨^h`$, $`(1Ad(h))𝔱`$, and $`\{c_n(h,x):n0,x𝔱\}`$ to be all of $`𝔨`$, it is therefore sufficient for the natural map of an $`r+1`$-dimensional space to an $`r`$-dimensional space
$$z+graph(T(z))𝔨/𝔱^{}$$
$`\mathrm{2.1.47}`$
to be surjective (note that the $`z`$ comes from the $`c_0`$ term; see the line preceding $`(\mathrm{2.1.42})`$). Thus if (1) $`z_\alpha 0`$, $`\alpha `$, and (2) (2.1.47) is surjective, then for $`h`$ corresponding to small $`z`$, (2.1.35) will hold.
We have already remarked that the first condition is of codimension 2. From the formula (2.1.45) for $`T(z)`$, we see that $`z`$ is generically independent of $`graph(T(z))`$, and $`graph(T(z))`$ is generically transverse to $`𝔱^{}`$. Therefore the condition (2) also has codimension 2, for $`z`$ in a small neighborhood of $`1`$. This completes the proof.//
We can now continue with the proof of (2.1.24). Let $`Q`$ (for rational) denote the set of points of $`K^{reg}`$ with the property that $`\{g^n\}`$ is not dense in the torus $`exp(𝔨^g)`$.
Now by (2.1.33) and (2.1.34) we know that $`eval|_{(g,h)}`$ maps onto $`ker(dp)`$ provided that $`(g,h)`$ is not in the set $`(Q_1\times K_2)(K_1\times X_2)`$. Now by considering the $`T_1`$-invariant distribution generated by $`𝔤_0`$ (and $`proj_2`$), we can also conclude that $`eval|_{(g,h)}`$ maps onto $`ker(dp)`$ provided that $`(g,h)`$ is not in the set $`(K_1\times Q_2)(X_1\times K_2)`$.
Because the condition that $`eval|_{(g,h)}`$ maps onto $`ker(dp)`$ is a linear independence condition, involving real analytic vector fields, the set of points where $`eval`$ does not map onto $`ker(dp)`$ is generically real analytic, and by the proceeding paragraph of codimension $`1`$. Since $`X_1`$ and $`X_2`$ have codimension $`2`$, the only portion of the singular set identified in the previous paragraph which could a priori support an object of codimension one is $`Q_1\times Q_2`$. But $`Q`$ has Hausdorff dimension $`d1`$, where $`d`$ is the dimension of $`K`$, hence $`Q_1\times Q_2`$ has Hausdorff dimension $`2d2`$. So the singular set must have codimension at least $`2`$. This completes the proof of (2.1.24).//
Proof of Theorem (2.1.4). Suppose that $`FL^2(K\times K)`$ is $`\mathrm{\Gamma }`$-invariant. By (2.1.16) $`F`$ is $`𝒢`$-invariant. Now given a generic point where $`𝔤`$ is infinitesimally transitive along the fiber, the $`𝒢`$-orbit of that point will be open in the fiber. For a generic fiber, the complement of these open sets has codimension $`>1`$, by (2.1.24). Hence for a generic fiber, the $`𝒢`$-orbits necessarily coincide with the components of the fiber. Thus an invariant $`F`$ is locally constant on connected components of $`a.e.`$ fiber. //
§2.2. The $`n`$-holed torus, with group element boundary condition.
Let $`\mathrm{\Sigma }_{(k_1,..,k_n)}`$ denote the $`n`$-holed torus with boundary components $`c_1`$$`,..,c_n`$, as in Figure 2 (where we ignore $`s`$ momentarily), with group element boundary condition $`k_{c_j}=k_j`$, $`1jn`$. Let $`\mathrm{\Sigma }_k^{}`$ denote the one-holed torus with group element boundary condition considered in §2.1, which reappears in Figure 2 with boundary component $`s`$ (and where we have moved the basepoint from the vertex to $`s`$, which we can do without affecting the results of §2.1). Let $`\mathrm{\Sigma }_{(k^{},k_1,..,k_n)}^+`$ denote the $`n+1`$-holed sphere with group element boundary condition pictured in Figure 2, where $`k^{}`$ is the labelling for $`s`$, and $`k_j`$ is the label for $`c_j`$; the corresponding $`Hom`$ space is empty unless $`k^{}=k_j`$, in which case it is a point. The Sewing Lemma (1.3) implies that we have a $`\pi _0(Aut(\mathrm{\Sigma }^{}))`$-equivariant bijection
$$Hom(\mathrm{\Sigma }_{(k_1,..,k_n)},K)Hom(\mathrm{\Sigma }_k^{},K)\times Hom(\mathrm{\Sigma }_{(k,k_1,..,k_n)}^+,K)$$
$`\mathrm{2.2.1}`$
where $`k=k_j`$. Note that $`\pi _0(Aut(\mathrm{\Sigma }^{}))`$$`=`$$`\pi _0(Aut(\mathrm{\Sigma }))`$.
Unfortunately this is not a situation where we can integrate over $`k`$, to obtain a result for every boundary condition, because $`k`$ is fixed by the $`k_j`$. We need to vary one of the boundary conditions, say $`k_n`$. We write $`\mathrm{\Sigma }_{(\stackrel{}{k},)}`$ for the object with boundary $`k_{c_j}=k_j`$, $`1j<n`$, where we allow $`k_{c_n}`$ to vary. We then have a $`\pi _0(Aut(\mathrm{\Sigma }))`$-equivariant bijection
$$Hom(\pi _1\mathrm{\Sigma }^{},K)Hom(\mathrm{\Sigma }_{(\stackrel{}{k},)},K).$$
$`\mathrm{2.2.2}`$
An immediate consequence of Theorem (2.1.4) is the following
###### Corollary(2.2.3)
For $`a.e.`$ $`k_n`$ $`[d\rho ]`$, the action
$$\pi _0(Aut(\mathrm{\Sigma }))\times Hom(\mathrm{\Sigma }_{(\stackrel{}{k},k_n)},K)$$
is ergodic on the Lebesgue class of each connected component.
§3. Proof of Ergodicity.
Let $`\mathrm{\Sigma }`$ denote the one-holed surface of genus $`p`$ with basepoint and link to the boundary $`c`$, as depicted in Figure 3 (ignore the paths $`s`$ and $`\alpha `$ at this point).
###### Theorem(3.1)
If the genus $`p>1`$, then for $`\underset{¯}{every}`$ group element boundary condition $`kK^{}`$, the action
$$\pi _0(Aut(\mathrm{\Sigma }_k))\times Hom(\mathrm{\Sigma }_k,K)Hom(\mathrm{\Sigma }_k,K)$$
is ergodic with respect to the Lebesgue class of each connected component.
Note that Theorem (0.1) (when the $`genus>1`$) is the special case $`k=1`$ of $`(3.1)`$. When the surface in (0.1) has $`genus=1`$, then (0.1) essentially reduces to the abelian case, and this is a standard application of Fourier series.
The basic facts about the connectedness properties of $`Hom(\mathrm{\Sigma },K)`$ which we will require are gathered in Appendix A, for the convenience of the reader. In particular (A.3) asserts that $`\pi _0(Aut(\mathrm{\Sigma }))`$ acts on components, so that the statement of the Theorem makes sense.
Proof of (3.1). Consider the decomposition of $`Hom(\mathrm{\Sigma },K)`$ into connected components described in (A.3). If we prove (3.1) for all groups of the form $`T\times K_1`$, where $`T`$ is a torus and $`K_1`$ is simply connected, then we will be done. So henceforth we assume that $`K^{}`$ is simply connected. In this case all the representation spaces are connected, by (A.2).
We have proven that $`\pi _0(Aut(\mathrm{\Sigma }^{}))`$ acts ergodically on $`Hom(\mathrm{\Sigma }^{},K)`$, for a one-holed torus $`\mathrm{\Sigma }^{}`$ as in §2.1, for $`a.e.`$ group element boundary condition. Similarly we have proven that $`\pi _0(Aut(\mathrm{\Sigma }^+))`$ acts ergodically on $`Hom(\mathrm{\Sigma }^+,K)`$ for a two-holed torus $`\mathrm{\Sigma }^+`$ as in §2.2, for $`a.e.`$ boundary condition on one end, and for every boundary condition on the other end. It therefore suffices to prove the following: suppose that $`s`$ is a separating curve as in the Sewing Lemma (1.3), such that $`\mathrm{\Sigma }^+`$ is a two-holed torus; if $`\pi _0(Aut(\stackrel{ˇ}{\mathrm{\Sigma }}))`$ acts ergodically on components of $`Hom(\stackrel{ˇ}{\mathrm{\Sigma }},K)`$, for $`a.e.`$ boundary condition on $`s`$, then the conclusion of (3.1) holds (see Figure 3; $`\mathrm{\Sigma }^+`$ is to the reader’s right of $`s`$).
Let $`k_c`$ denote the fixed boundary condition for $`\mathrm{\Sigma }`$. The measure classes for possible boundary conditions on $`s`$ are the same for $`\mathrm{\Sigma }^\pm `$, the Lebesgue class on $`K^{}`$ (see (2.1.5)). Let $`F`$ denote a characteristic function on $`Hom(\mathrm{\Sigma },K)`$. If $`F`$ is $`\pi _0(Aut(\stackrel{ˇ}{\mathrm{\Sigma }}))`$-invariant, then by the Sewing Lemma (1.3) and our induction hypothesis, it follows that $`F`$ is constant along $`a.e.`$ fiber; hence $`F`$ is of the form $`f(g|_s)`$, where $`f`$ is a characteristic function on $`K^{}`$.
Now suppose that $`F`$ is $`\pi _0(Aut(\mathrm{\Sigma }))`$-invariant. As in Figure 3, we choose a Dehn twist $`\sigma `$ corresponding to a loop $`\alpha `$ that will cross the curve $`s`$, but will not cross the link to the boundary component, so that $`\sigma \pi _0(Aut(\mathrm{\Sigma }))`$. Now the loop $`\alpha `$ does not pass through the basepoint. There are two elemental ways in which we can use $`s`$ to link $`\alpha `$ to the basepoint; if we go from the basepoint in the negative direction along $`s`$ to $`\alpha `$, around $`\alpha `$, and return to the basepoint, then we denote this based loop by $`\overline{\alpha }`$; if we go from the basepoint in the positive direction along $`s`$ to $`\alpha `$, around $`\alpha `$, and return, then we denote the loop by $`\underset{¯}{\alpha }`$. Using Figure 3 we compute that
$$\overline{\alpha }=[\alpha _p,\beta _p]\alpha _p^1\beta _{p1}\alpha _{p1}\beta _{p1}^1,\underset{¯}{\alpha }=s\beta _{p1}\alpha _{p1}\beta _{p1}^1\alpha _p[\alpha _p,\beta _p]^1s^1,$$
$`3.3`$
$$\sigma \alpha _j=\alpha _j,\sigma \beta _{p1}=\overline{\alpha }\beta _{p1},\sigma \beta _p=\beta _p\overline{\alpha }^1,$$
$`3.4`$
$$\sigma s=\underset{¯}{\alpha }s\overline{\alpha }^1=(\underset{1}{\overset{p2}{}}[\alpha _j,\beta _j])[\alpha _{p1},\overline{\alpha }\beta _{p1}]=$$
$`3.5`$
$$s[\alpha _{p1},\beta _{p1}]^1\alpha _{p1}\alpha _p[\alpha _p,\beta _p]^1\alpha _{p1}^1[\alpha _{p1},\beta _{p1}][\alpha _p,\beta _p]\alpha _p^1.$$
$`3.6`$
It is convenient to streamline our notation. We put
$$g_1=\underset{1}{\overset{p2}{}}[g_{\alpha _j},g_{\beta _j}],g=g_{\alpha _{p1}},h=g_{\beta _{p1}},k=g_{\alpha _p},l=g_{\beta _p}.$$
$`3.7`$
We have $`(\sigma F)(g)=f(g_{\sigma s})`$. Hence the $`\sigma `$-invariance of $`F`$ is equivalent to
$$f(g_1[g,h])=f((g_1g)(kk_c^1)g_1[g,h](g_1g)^1(kk_c^1)^1),$$
$`3.8`$
for $`a.e.`$ $`g_1,g,h,k,l`$, subject to the constraint $`g_1[g,h][k,l]=k_c`$.
Define
$$\varphi :\{(g_1,g,h,k,l):g_1[g,h][k,l]=k_c\}K^{}\times K^{}$$
$`3.9`$
$$\varphi :(g_1,g,h,k,l)(g_1[g,h],(g_1g)(kk_c^1)g_1[g,h](g_1g)^1(kk_c^1)^1);$$
(3.8) is equivalent to $`\varphi _1^{}f=\varphi _2^{}f`$, $`a.e.`$. If this equality held at all points, then to prove that $`f`$ is constant, it would suffice to show that the relation defined by $`Im(\varphi )`$ (or the equivalence relation generated by $`Im(\varphi )`$) is transitive; since the equality holds in an $`a.e.`$ sense, we must consider the relation defined by the interior of $`Im(\varphi )`$. It is plausible that $`\varphi `$ is surjective, but we can only prove the following weaker result.
###### Lemma(3.10)
Let $`pr_1:K^{}\times K^{}K^{}:(m,n)m`$. Then $`K^{}pr_1(Interior(Im(\varphi )))`$ has codimension at least 2 in $`K^{}`$.
This Lemma implies that for each $`mpr_1(Interior(Im(\varphi )))`$, we can find open sets $`U_m`$ and $`V_m`$ in $`K^{}`$ such that $`mU_m`$ and $`U_m\times V_mIm(\varphi )`$. Since $`\varphi _1^{}f=\varphi _2^{}f`$, $`a.e.`$, it follows that $`f`$ is constant on $`U_m`$, $`a.e.`$. Since $`pr_1(Interior(Im(\varphi )))`$ is connected, this constant must be the same for each $`U_m`$. This implies that $`f`$ is constant $`a.e.`$. Thus proving (3.10) will complete the proof of (3.1).
Proof of (3.10). We can suppose that $`p=2`$, which amounts to setting $`g_1=1`$, and that $`K=K^{}`$. For notational simplicity we will abbreviate $`Ad(g)()`$ to $`g()`$.
To prove (3.10), we first claim that it suffices to show that the map
$$\psi _m:\{[g,h]=m\}\times \{[k,l]=m^1k_c\}K:(g,h;k,l)gkk_c^1mg^1(kk_c^1)^1$$
$`3.11`$
is regular at some smooth point, for each $`mKY`$, where $`Y`$ has codimension 2. For if $`\psi _m`$ is regular at the smooth point $`(g,h;k,l)`$, then $`(g,h)`$ is regular for the commutator map, which is the first factor of $`\varphi `$. Thus $`Im(d\varphi |_{(g,h,k,l)})`$ spans both the vertical and horizontal directions, hence $`(g,h;k,l)`$ is regular for $`\varphi `$.
To specify $`Y`$, consider the commutator map $`[,]:K\times KK`$. This map is surjective, the fibers generically have dimension $`d=dimn(K)`$, and the exceptional fibers have dimension exceeding $`d`$ (e.g. $`[,]^1(1)`$ has dimension $`d+r`$, $`r=rank(K)`$). Let $`N`$ denote the set of values $`nK`$ such that there exists $`(g,h)`$ with $`[g,h]=n`$ and (i) $`gK^{reg}`$ and (ii) $`𝔨^g𝔨^h=\{0\}`$ (i.e. $`(g,h)`$ is regular for $`[,]`$). By (B.5) of Appendix B, $`KN`$ has codimension at least 2 in $`K`$. We set
$$KY=\{m:mNandm^1k_cN\}.$$
$`3.12`$
The Zariski tangent space to $`[,]^1(m)`$ at $`(g,h)`$ is given by
$$T|_{(g,h)}=\{(x,y):x^{h^1}x+yy^{g^1}=0\}.$$
$`3.13`$
Projection onto the $`x`$ factor induces the exact sequence
$$0\{(0,y):y𝔨^g\}T|_{(g,h)}\{x:(1h^1)x𝔨^g\}0.$$
$`3.14`$
If $`(g,h)`$ satisfies (i) and (ii), then $`(g,h)`$ is a smooth point, and the spaces in (3.14) have dimensions $`r`$, $`d`$, and $`dr`$, respectively. Note that
$$\{x:(1h^1)x𝔨^g\}=((1h)𝔨^g)^{},$$
$`3.15`$
and this space depends only upon $`g`$: since $`hgh^1=(g^1m)^1`$, $`h`$ is unique up to multiplication on the right by $`\lambda C_K(g)`$, and $`\lambda `$ acts trivially on $`𝔨^g`$.
Fix $`mKY`$. The derivative of the map $`\psi _m`$ is given by
$$d\psi |_{(g,h;k,l)}:(x,y;z,w)(x^{m^1(kk_c^1)^1}x)^{kk_c^1g}+(z^{k_c^1gm^1k_c}z)^k$$
$$=(x^{(kk_c^1m)^1}x+z^{m^1k_c}(z^{m^1k_c})^{g^1m})^{kk_c^1g}.$$
$`3.16`$
Together with (3.14) and (3.15) this means that we must show that for suitable $`g,h,k,l`$, the sum of subspaces
$$(1(kk_c^1m)^1)((1h)𝔨^g)^{}+(1g^1m)m^1k_c((1l)𝔨^k)^{}$$
$`3.17`$
is all of $`𝔨`$.
Now to deal with (3.17), we need some control over solutions to the constraint equations $`[g,h]=m`$, $`[k,l]=m^1k_c`$. For this purpose, consider the equation $`[g_1,h_1]=n`$. In (B.1) of Appendix B, we show that for any maximal torus $`T`$, there exists a solution $`(g_1,h_1)`$ with $`g_1T`$. For $`nN`$, by dimensional considerations, $`g_1`$ is a finite multi-valued function of $`T`$ (see (a) of (B.6) for explicit equations). Apply this to $`n=m`$. Given $`T`$, we obtain solutions $`[g,h]=m`$. We have $`g^1m`$$`=`$$`hg^1h^1hTh^1`$. Therefore we obtain a finite number of tori $`hTh^1`$. We claim that the multi-valued map
$$\varphi _m:\{Tori\}\{Tori\}:ThTh^1C_K(g^1m)$$
$`3.18`$
is surjective. In a loose sense the inverse is $`\varphi _{m^1}`$, because $`[hgh^1,h^1]=m^1`$. More precisely, given a torus $`T_1`$, apply the preceding to $`m^1`$ and $`T_1`$ to obtain $`(g_1,h_1)`$ with $`[g_1,h_1]=m^1`$ and $`g_1T_1`$. Define $`(g,h)`$ and $`T`$ so that $`g_1=hgh^1`$, $`h_1=h^1`$, $`T=h_1T_1h_1^1`$. Then $`[g,h]=m`$, and $`T_1=hTh^1`$. This proves the claim.
Similarly the multi-valued map
$$\mathrm{\Phi }_{m^1k_c}:\{Tori\}\{Tori\}:T(kl)T(kl)^1C_K(kk_c^1m),$$
$`3.19`$
where $`[k,l]=m^1k_c`$, $`kT`$, is surjective, and the inverse, again in a loose sense, is $`\mathrm{\Phi }_{k_c^1m}`$. For given $`T_1`$ we can find $`[k_1,l_1]=k_c^1m`$, $`k_1T_1`$. Define $`k=(k_1l{}_{1}{}^{})k_1(k_1l_1)^1`$, $`l=(k_1l_1)k_1^1(k_1l_1)^2`$, $`T=(k_1l_1)T_1(k_1l_1)^1`$. Then $`[k,l]=m^1k_c`$, $`kT`$, and $`T_1=klT(kl)^1`$.
Choose the pairs $`(g,h)`$ and $`(k,l)`$ such that $`[g,h]=m`$ and $`[k,l]=m^1k_c`$, and such that both pairs satisfy (i) and (ii) above. It may be necessary to consider perturbations of these pairs. We will refer to perturbations which fix the constraints as admissible. The conditions (i) and (ii) are stable under small admissible perturbations. The space $`[,]^1(m)`$ has dimension $`d`$, and for $`g`$ as above the possible $`h`$’s with $`[g,h]=m`$ form an $`r`$ dimensional set. Thus an admissible small perturbation of $`(g,h)`$ gives a smooth $`dr`$ dimensional perturbation of $`g`$, the tangent space of which is described by (3.15). The same comments apply to $`(k,l)[,]^1(m^1k_c)`$.
Now consider the subspace represented by the first term in (3.17). We first fix $`g`$ and $`h`$. We claim that we can choose an arbitrarily small admissible perturbation of $`(k,l)`$ such that
$$(1(kk_c^1m)^1)((1h)𝔨^g)^{}=$$
$$Im(1(kk_c^1m)^1)=(𝔨^{kk_c^1m})^{}.$$
$`3.20`$
This will hold if we can arrange for $`((1h)𝔨^g)^{}`$ to intersect $`ker(1(kk_c^1m)^1)`$ trivially, i.e.
$$𝔨^{kk_c^1m}((1h)𝔨^g)^{}=\{0\}.$$
$`3.21`$
Because $`k`$ is regular, $`kk_c^1m=(kl)k(kl)^1`$ is regular. Thus $`𝔨^{kk_c^1m}`$ has dimension $`r`$, and $`((1h)k^g)^{}`$ has dimension $`dr`$ (because of condition (ii)). By (B.1) and (3.19) we can find an arbitrarily small admissible perturbation of $`(k,l)`$ such that the intersection (3.21) will be zero.
We now fix our choice of $`(k,l)`$. We claim that we can find an arbitrarily small admissible perturbation of $`(g,h)`$ such that
$$(1g^1m)m^1k_c((1l)𝔨^k)^{}=(𝔨^{g^1m})^{}.$$
$`3.23`$
The argument is essentially the same. It suffices to establish
$$𝔨^{g^1m}m^1k_c((1l)𝔨^k)^{}=\{0\}.$$
$`3.24`$
As before, $`g^1m`$ is regular, because $`g`$ is regular. By (3.18) we can arrange this by an arbitrarily small admissible perturbation.
We now have found $`g,h,k,l`$ such that the image of the subspace (3.17) equals
$$(𝔨^{kk_c^1m})^{}+(𝔨^{g^1m})^{}=(𝔨^{kk_c^1m}𝔨^{g^1m})^{},$$
$`3.25`$
and this equality is stable under small admissible perturbations. Again by (3.18) and (3.19) we can find a small admissible perturbation so that (3.25) will be all of $`𝔨`$.
We have now proven that the map $`\psi _m`$ is regular at some smooth point for each $`mKY`$, and as we observed at the beginning of the proof, this implies (3.10). //
Appendix A. Connectedness Properties.
The following results can be deduced from \[BR\] (and perhaps elsewhere). We record them here for the convenience of the reader.
###### Lemma(A.1)
Suppose that $`K`$ is simply connected. Then $`\{(g,h)K\times K:[g,h]=k\}`$ is connected, $`kK`$.
If $`C`$ denotes the conjugacy class containing $`k`$, then there is a surjective map
$$\{[g,h]=k\}\{[g,h]C\}/conj(K)$$
and the fibers are homogeneous spaces for $`K`$. The fibers are connected because $`K`$ is connected, and by \[BR\] the moduli space corresponding to $`C`$ is connected because $`K`$ is simply connected. This establishes (A.1) (It would clearly be desirable to give an elementary direct proof of this).
###### Lemma(A.2)
Suppose that $`K`$ is simply connected. Suppose that $`\mathrm{\Sigma }`$ is an object with group element boundary condition which is obtained by sewing one-holed tori to an $`N`$-holed sphere. Then $`Hom(\mathrm{\Sigma },K)`$ is connected.
Proof. The space $`Hom`$ for an $`N`$-holed sphere is empty or a point. When we sew, we obtain a connected object by (1.3).//
Let $`pr:\stackrel{~}{K}K`$ denote the universal covering of $`K`$.
###### Proposition(A.3)
If $`\mathrm{\Sigma }`$ is a one-holed surface with boundary condition $`lK`$, then we have the decomposition into connected components
$$Hom(\mathrm{\Sigma }_l,K)=\underset{\stackrel{~}{l}\stackrel{~}{K}^{}pr^1(l)}{}pr_{}Hom(\mathrm{\Sigma }_{\stackrel{~}{l}},\stackrel{~}{K}).$$
This decomposition is equivariant with respect to $`\pi _0(Aut(\mathrm{\Sigma }))`$.
This follows from (A.2).
Appendix B. Commutators.
At several points of this paper, we used the fact that the commutator map $`[,]:K\times KK^{}`$ is surjective (and we presented an indirect proof of this in (a) of (2.1.5)). Here we discuss some refinements which we use in the proof of (3.10).
###### Proposition(B.1)
Let $`T`$ denote a maximal torus in $`K`$. The map
$$\psi :T\times KK^{}:(\lambda ,h)[\lambda ,h]$$
is surjective.
Proof of (B.1). To simplify the notation, we will write $`K`$ in place of $`K^{}`$; $`d`$ will denote the dimension, and $`r`$ the rank, of $`𝔨`$.
The derivative of $`\psi `$ at $`(\lambda ,h)`$ is given by
$$𝔱\times 𝔨𝔨:(x,y)(x^{\lambda h^1}x^\lambda +y^\lambda y)^h,$$
$`B.2`$
hence the image of the derivative at $`(\lambda ,h)`$ is
$$Ad(h)(Ad(\lambda )((1Ad(h^1))𝔱)+(𝔨^\lambda )^{})$$
$$=Ad(h\lambda )((1Ad(h^1))𝔱+(𝔨^\lambda )^{}).$$
$`B.3`$
We claim that the point $`(\lambda ,h)`$ is critical for $`\psi `$ if and only if (i) $`\lambda K^{reg}`$ or (ii) $`𝔨^h𝔱\{0\}`$. To see this, suppose that $`\lambda `$ is regular and $`𝔨^h𝔱=\{0\}`$. Then $`(𝔨^\lambda )^{}=𝔱^{}`$ has dimension $`dr`$ and $`(1Ad(h^1))𝔱`$ has dimension $`r`$. If the intersection of these two spaces is nonempty, then there is $`x𝔱`$ such that $`x^{h^1}=x+y^{}`$, where $`y^{}𝔱^{}`$ is not zero; but $`x^{h^1}`$ and $`x`$ have the same length, so that $`xy^{}`$ implies $`y^{}=0`$, which is a contradiction. Thus the dimension of the space (B.3) is $`d`$, and this establishes our claim.
We can factor $`\psi =\stackrel{~}{\psi }p`$, where
$$T\times K\mathrm{@}>p>>T\times K/T\mathrm{@}>\stackrel{~}{\psi }>>K:(\lambda ,h)\mathrm{@}>p>>(\lambda ,hT)\mathrm{@}>\stackrel{~}{\psi }>>[\lambda ,h],$$
$`B.4`$
so that at any regular point, $`\stackrel{~}{\psi }`$ will actually be a local diffeomorphism. We claim that the set of critical values for $`\stackrel{~}{\psi }`$ has codimension at least two. This will imply that $`\stackrel{~}{\psi }`$ is surjective, because a boundary for the image would necessarily have codimension one.
Suppose that (i) holds, i.e. $`\lambda _0T^{reg}`$. In this case, as we vary $`h`$, $`\lambda _0h\lambda _0^1h^1`$ will sweep out the $`\lambda _0`$-translate of a nongeneric conjugacy class, which will have dimension $`dr2`$. Thus the dimension of the set of critical values arising from condition (i) will be $`r1+dr2=d3`$.
Now suppose that (ii) holds. The condition $`𝔨^{h_0}𝔱0`$ has codimension at least 2 in $`K`$: if $`h_0=exp(X)`$, where $`X`$ is regular, then we must have $`X_\alpha =0`$ for some root $`\alpha `$ of $`𝔱`$, where $`X_\alpha `$ denotes the $`\alpha `$-root space component of $`X`$ (see the proof of (2.1.24), especially the paragraph containing (2.1.28)). This is a $`T`$-invariant condition, hence the set of critical points corresponding to (ii) has codimension at least 2 in $`K/T`$. It follows that the corresponding set of critical values has dimension $`r+dr2`$. This completes the proof.//
###### Corollary(B.5)
For the commutator map $`[,]:K\times KK^{}`$, the complement of the subset $`N`$ of $`K^{}`$ defined by
$$\{n:(g,h)[,]^1(n)s.t.(i)gK^{reg},(ii)𝔨^g𝔨^h=\{0\}\}$$
has codimension at least 2.
Proof. Given a maximal torus $`T`$, each regular value for the map $`\psi `$ of (B.1) will belong to $`N`$. In the proof of (B.1) we established that the complement of the set of regular values for $`\psi `$ has codimension at least 2. By varying $`T`$, we obtain (B.5).//
Remarks(B.6). (a) It is of interest to consider the more general question of whether, for given $`gK`$, the map
$$\psi _g:T\times K^{}K^{}:(\lambda ,h)[g\lambda ,h]$$
$`B.7`$
is surjective. This has a factorization
$$T\times K\mathrm{@}>p>>D_g=\{(\lambda ,l)T\times K:g\lambda l\}\mathrm{@}>\stackrel{~}{\psi }_g>>K\mathrm{@}>L_g>>K$$
$$(\lambda ,h)\mathrm{@}>p>>(\lambda ,hg\lambda h^1)=(\lambda ,l)\mathrm{@}>\stackrel{~}{\psi }_g>>\lambda l^1=k\mathrm{@}>L_g>>gk,$$
$`B.8`$
where $`g\lambda l`$ means $`g\lambda `$ and $`l`$ are conjugate. The map $`\stackrel{~}{\psi }_g`$ is the restriction to $`D_g`$ of the natural coset fibration
$$T\times K(T\times K)/\mathrm{\Delta }(T),$$
$`B.9`$
where $`\mathrm{\Delta }(T)`$ is the diagonally embedded copy of $`T`$ in $`T\times K`$, and we identify $`(T\times K)/\mathrm{\Delta }(T)`$ with $`K`$ by $`(\lambda ,l)\mathrm{\Delta }(T)\lambda l^1`$. The map $`\stackrel{~}{\psi }_g`$ is surjective if and only if for each $`kK`$, there exists $`\lambda T`$ such that $`g\lambda l=k^1\lambda `$. This is equivalent to a system of $`r`$ polynomial equations
$$\chi _i(g\lambda )=\chi _i(k^1\lambda ),i=1,..,r$$
$`B.10`$
for $`r`$ unknowns $`\lambda _1,..,\lambda _r𝕋`$, where $`\chi _i`$ is the character corresponding to the $`i^{th}`$ fundamental irreducible representation, and $`\lambda =_1^r\lambda _i^{h_i}`$, where the $`h_i`$ are the coroots (e.g. for $`SU(3)`$, we have 2 equations
$$\underset{1}{\overset{3}{}}A_i\lambda _i=0,\overline{A}_1\lambda _2\lambda _3+\overline{A}_2\lambda _1\lambda _3+\overline{A}_3\lambda _1\lambda _2=0,$$
$`B.11`$
for the $`\lambda _i𝕋`$, subject to the constraint $`\lambda _i=1`$, where $`A_i=g_{ii}(k^1)_{ii}`$). It is trivial to check that for $`SU(2)`$, $`\psi _g`$ is always surjective, but this is not so for $`SU(3)`$. Thus in particular the equations (B.11) do not in general have solutions satisfying the reality condition $`|\lambda _i|=1`$; on the other hand (B.1) asserts that such solutions always exist for $`g=1`$.
This suggests a number of questions, such as how does one describe the set of conjugacy classes which meet $`gT`$, when is $`\psi _g`$ surjective, and so on.
(b) Identify $`SU(2)`$ with $`_1`$, the group of unit quaternions, by $`\left(\begin{array}{cc}a& b\\ \overline{b}& \overline{a}\end{array}\right)q=abj`$, and take $`T=𝕋`$. The conjugacy classes in $`_1`$ are obtained by fixing the real part of $`q`$. Now fix $`g=abj`$. The conjugacy classes which meet $`g𝕋`$ are indexed by $`[|a|,|a|]`$. We have
$$D_g=\{(\lambda ,q)𝕋\times _1:e(q)=e(a\lambda )\},$$
$$\stackrel{~}{\psi }_g:D_g_1:(\lambda ,q)\lambda \overline{q},$$
$$TD_g|_{\lambda ,q}=\{(is,q^{})i\times Im():e(a\lambda is)=e(qq^{})\},$$
$`B.12`$
$$d(\stackrel{~}{\psi }_g):TD_g|_{\lambda ,q}Im():(is,q^{})q(is+\overline{q}^{})\overline{q},$$
$$D_{g,critical}=\{(\lambda ,a\lambda +\sqrt{1|a|^2}zj):\lambda ,z𝕋\}.$$
When $`0<|a|<1`$, the singular set is a 2-torus; at the extreme values $`|a|=0,1`$, the critical set degenerates to a circle. The $`SU(2)`$ miracle is that in all cases, the set of critical values
$$\stackrel{~}{\psi }_g(D_{g,critical})=\{\overline{a}\sqrt{1|a|^2}\lambda zj:\lambda ,z𝕋\}$$
$`B.13`$
is a circle. One can easily visualize how $`\stackrel{~}{\psi }_g`$ covers $`_1`$.
The extreme case $`|a|=0`$, when there is just a single (totally geodesic) conjugacy class, corresponds to the condition that $`g`$ is a so-called principal element (\[K\]).
References
\[AB\] M. Atiyah and R. Bott, The Yang-Mills equations over Riemann surfaces, Phil. Trans. R. Soc. Lond., A308, (1982), 523-615.
\[BR\] U. Bhosle and A. Ramanathan, Moduli spaces of principal bundles with parabolic structure over Riemann surfaces, Math. Z. 202 (1989), 161-180.
\[Go1\] W. Goldman, The symplectic nature of fundamental groups of surfaces, Adv. Math. 54 (1984), 200-225.
\[Go2\] ———-, Ergodic theory on moduli spaces, Ann. Math. 146 (1997), 475-507.
\[K\] B. Kostant, The principal three-dimensional subgroups and the Betti numbers of a complex simple Lie group, Amer. J. Math. 81 (1959), 973-1032.
|
warning/0006/hep-th0006235.html
|
ar5iv
|
text
|
# CERN-TH/2000-186, LFM(SCF)-00-1, UB-ECM-PF-00/08,TOHO-FP-0067hep-th/0006235 Hamiltonian Formalism for Space-time Non-commutative Theories
## I Introduction
Space-time non-commutative field theories have peculiar properties due to their acausal behavior and lack of unitarity . In reference it has been shown that there is a relation between lack of unitarity and the obstruction to finding a decoupling limit of string theory in an electromagnetic background . These theories have an infinite number of temporal and spatial derivatives, and therefore are non-local in time and space. The initial value problem of a non-local theory requires to give a trajectory or a finite piece of it. The Euler-Lagrange (EL) equation is a constraint in the space of trajectories.
The Hamiltonian formalism for non-local theories was presented in . In this paper we improve the formalism by clarifying the relation among the Lagrangian and Hamiltonian structures. We first consider an equivalent theory in a space-time of one dimension higher than that of the original theory. This space has ”two times” and the dynamics is described in such a way that the evolution is local with respect to one of the times. For this equivalent theory one can construct the Hamiltonian. A characteristic feature of the Hamiltonian formalism for non-local theories is that it contains the EL equations as Hamiltonian constraints.
The Hamiltonian path integral for the $`d+1`$ dimensional field theory is constructed. The Lagrangian path integral formalism for the $`d`$ dimensional theory is obtained by integrating out the momenta.
We apply the Hamiltonian formalism to time-like and light-like non-commutative theories. As an example we consider the case of non-commutative $`\varphi ^3`$ theory in $`d`$ dimensions with space-time non-commutativity. The action contains the free Klein-Gordon Lagrangian and the interaction Lagrangian $`L_i=\frac{g}{3!}𝑑\stackrel{}{x}\varphi \varphi \varphi `$ , where $``$ refers to the Moyal product. We construct the Hamiltonian in $`d+1`$ dimensions. In the path integral quantization we get the Feynman rules that coincide with those used in references . The theory is unitary at the classical level (tree level) but is not unitary at one loop . This analysis should shed new light on the structure of these theories. The knowledge of the Hamiltonian of time-like and light-like non-commutative field theories could also be useful to study the energy of their solitons.
## II Euler-Lagrange equations for non-local theories
Unlike standard Lagrangians, that depend on the values of a finite number of derivatives at a given time —$`q(t)`$, $`\dot{q}(t)`$, …$`q^{(n)}(t)`$—, a non-local Lagrangian depends on a whole piece of the trajectory $`q(t+\lambda )`$, for all values of $`\lambda `$, that is, $`L^{\mathrm{non}}(t)=L([q(t+\lambda )])`$. At best it can be written as a function of all time derivatives $`q^{(j)}(t)`$, $`j=0,1,2,\mathrm{}`$ at the same $`t`$. This means that the analog of the tangent bundle for Lagrangians depending on positions and velocities is infinite dimensional. The action is
$$S[q]=𝑑tL^{\mathrm{non}}(t).$$
(1)
The EL equation is obtained as the variation of functional (1) and is given by
$$𝑑tE(t,t^{};[q])=0,$$
(2)
where $`E(t,t^{};[q])={\displaystyle \frac{\delta L^{\mathrm{non}}(t)}{\delta q(t^{})}}`$.
The EL equation must be understood as a functional relation to be fulfilled by physical trajectories. It is not a differential system, as one is used to find for local Lagrangians. In the latter case, the theorems of existence and uniqueness of solutions enable to interpret the EL equation as ruling the time evolution of the system, whose state at every instant of time is represented by a point in the space of initial data, e.g., $`J_L=\{q,\dot{q},\mathrm{},q^{(2n1)}\}`$ for a local Lagrangian of order $`n`$.
In the non-local case, if we denote the space of all possible trajectories as $`J=\{q(\lambda ),\lambda R\}`$, (2) is a Lagrangian constraint defining the subspace $`J_RJ`$ of physical trajectories.
## III $`1+1`$ dimensional field theory description of non-local theories
Nevertheless, if we insist in defining a “time evolution” $`T_t`$ for a given initial trajectory $`q(\lambda )`$, a natural choice is:
$$q(\lambda )\stackrel{T_t}{}q(\lambda +t).$$
(3)
We shall hence introduce new dynamical variables $`Q(t,\lambda )`$ such that
$$Q(t,\lambda )=q(\lambda +t).$$
(4)
Thus, $`t`$ is the “evolution” parameter and $`\lambda `$ is a continuous parameter indexing the degrees of freedom. These new variables follow the evolution (4) above, and $`Q(0,\lambda )`$ can be seen as initial data in the local $`1+1`$ dimensional field theory.
In differential form, condition (4) reads:
$$\dot{Q}(t,\lambda )=Q^{}(t,\lambda ),$$
(5)
where ‘dot’ and ‘prime’ respectively stand for $`_t`$ and $`_\lambda `$.
We then consider the Hamiltonian system for the $`1+1`$ dimensional field $`Q`$ with the Hamiltonian
$$H(t,[Q,P])=𝑑\lambda P(t,\lambda )Q^{}(t,\lambda )\stackrel{~}{L}(t,[Q]),$$
(6)
where $`P`$ is the canonical momentum of $`Q`$. The phase space is thus $`T^{}J`$ with the fundamental Poisson brackets
$$\{Q(t,\lambda ),P(t,\lambda ^{})\}=\delta (\lambda \lambda ^{}).$$
(7)
In the Hamiltonian (6) $`\stackrel{~}{L}(t,[Q])`$ is a functional defined by
$$\stackrel{~}{L}(t,[Q]):=𝑑\lambda \delta (\lambda )(t,\lambda ).$$
(8)
The “density” $`(t,\lambda )`$ is constructed from $`L^{\mathrm{non}}(t)`$ by replacing $`q(t)`$ by $`Q(t,\lambda )`$, the $`t`$-derivatives of $`q(t)`$ by $`\lambda `$-derivatives of $`Q(t,\lambda )`$ and $`q(t+\rho )`$ by $`Q(t,\lambda +\rho )`$. In this construction of the Hamiltonian $`\lambda `$ inherits the signature of the original time $`t`$ and is a time-like coordinate. Furthermore the symmetry of the original Lagrangian is realised canonically in the enlarged space . Note that $`(t,\lambda )`$ is local in $`t`$ and is non-local in $`\lambda `$. $`H`$ depends linearly on $`P(t,\lambda )`$ but does not depend on $`\dot{Q}(t,\lambda )`$.
The relation (5) naturally arises as the first Hamilton equation for (6). However there is no a priori relationship between $`P(t,\lambda )`$ and $`Q(t,\lambda )`$ —unlike it happens in the local case—, the second Hamilton equation:
$$\dot{P}(t,\lambda )=P^{}(t,\lambda )+\frac{\delta \stackrel{~}{L}(t,[Q])}{\delta Q(t,\lambda )}$$
(9)
does not imply any further restriction on $`Q(t,\lambda )`$. Thus, the Hamiltonian system (6) on $`T^{}J`$ is not so far equivalent to the non-local Lagrangian system of $`L^{non}(t)`$.
Now, instead of taking the whole phase space $`T^{}J`$, we shall restrict to the subspace defined by the 1-parameter set of primary constraints :
$$\phi (t,\lambda ,[Q,P])P(t,\lambda )F(t,\lambda ,[Q])0$$
(10)
with
$$F(t,\lambda ,[Q]):=𝑑\sigma \chi (\lambda ,\sigma )(t;\sigma ,\lambda ),$$
(11)
where $`(t;\sigma ,\lambda )`$ and $`\chi (\lambda ,\sigma )`$ are defined by
$`(t;\sigma ,\lambda )`$ $`=`$ $`{\displaystyle \frac{\delta (t,\sigma )}{\delta Q(t,\lambda )}},\chi (\lambda ,\sigma )={\displaystyle \frac{ϵ(\lambda )ϵ(\sigma )}{2}}.`$ (12)
Here $`ϵ(\lambda )`$ is the sign distribution. The symbols ”$``$” and ”$``$” respectively stand for “strong” and “weak” equality.
Further constraints are generated by requiring the stability of the primary ones. In the first step, we obtain:
$$\dot{\phi }(t,\lambda ,[Q,P])\phi ^{}(t,\lambda ,[Q,P])+\delta (\lambda )\psi _0(t,[Q])0$$
where
$$\psi _0(t,[Q]):=𝑑\sigma (t;\sigma ,0)0$$
(13)
is the secondary constraint. Further constraints then follow by successive time differentiations of $`\psi _0`$. They can be written all together in a condensed form as:
$$\psi (t,\lambda ,[Q])𝑑\sigma (t;\sigma ,\lambda )0.$$
(14)
Therefore, the constrained Hamiltonian system defined by the Hamiltonian (6) and the primary constraints (10) lives in a reduced phase space $`\mathrm{\Gamma }T^{}J`$ defined by (10) and (14). Taking into account (4), the constraint (14) reduces to the EL equation (2) obtained from $`L^{\mathrm{non}}(t)`$.
The constraints (10) and (14) belong to the second class in non-singular systems. In the next section we will show explicitly, for (non-singular) higher derivative Lagrangian system of order $`n`$, they are used to reduce the phase space to $`2n`$ dimensions reproducing the canonical Ostrogradski formalism . Our formalism developed here turns out to be a generalization of the Ostrogradski formalism to the case of infinite order derivative theories. The infinite chain of second class constraints has also appeared in the description of boundary conditions as constraints . Summarizing, the equivalence has been built in the $`1+1`$ dimensional Hamiltonian formalism of local field theories through the constraints (10) and (14). This type of equivalence between the Hamiltonian and Lagrangian formalism is different from the one in local theories.
## IV Non-singular Higher Order Derivative theories
Here we would like to derive both the Lagrangian and Hamiltonian formalisms for non-singular higher order derivative theories from the Hamiltonian formalism of non-local theories developed in the last section.
Let us consider a regular higher derivative theory described by the Lagrangian $`L(q,\dot{q},\ddot{q},\mathrm{},q^{(n)})`$ and write the expressions obtained in the previous section for the non-local Lagrangian. As we embed the higher order theory in the non-local setting we start with the infinite dimensional phase space $`T^{}J(t)=\{Q(t,\lambda ),P(t,\lambda )\}`$. They are assumed to be expanded in the Taylor basis as
$`Q(t,\lambda )`$ $``$ $`{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}e_m(\lambda )q^m(t),`$ (15)
$`P(t,\lambda )`$ $``$ $`{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}e^m(\lambda )p_m(t),`$ (16)
where $`e^{\mathrm{}}(\lambda )`$ and $`e_{\mathrm{}}(\lambda )`$ are orthonormal basis
$`e^{\mathrm{}}(\lambda )`$ $`=`$ $`(_\lambda )^{\mathrm{}}\delta (\lambda ),e_{\mathrm{}}(\lambda )={\displaystyle \frac{\lambda ^{\mathrm{}}}{\mathrm{}!}}.`$ (17)
The coefficients in (16) are new canonical variables
$`\{q^m(t),p_n(t)\}`$ $`=`$ $`\delta _{}^{m}{}_{n}{}^{}`$ (18)
and the Hamiltonian (6) is
$`H(t)`$ $`=`$ $`{\displaystyle \underset{m=0}{\overset{\mathrm{}}{}}}p_m(t)q^{m+1}(t)L(q^0,q^1,\mathrm{},q^n).`$ (19)
The momentum constraint (10) becomes an infinite set of constraints
$`\phi _m(t)`$ $`=`$ $`p_m(t){\displaystyle \underset{\mathrm{}=0}{\overset{nm1}{}}}(D_t)^{\mathrm{}}{\displaystyle \frac{L(t)}{q^{\mathrm{}+m+1}(t)}}0,`$ (20)
where
$`D_t={\displaystyle \underset{r=0}{}}q^{r+1}{\displaystyle \frac{}{q^r}}.`$ (21)
On the other hand the constraint $`\psi `$ in (14) in terms of the Taylor basis becomes
$`\psi ^m(t)`$ $``$ $`(D_t)^m[{\displaystyle \underset{\mathrm{}=0}{\overset{n}{}}}(D_t)^{\mathrm{}}{\displaystyle \frac{L(t)}{q^{\mathrm{}}(t)}}]0.`$ (22)
These constraints (20) and (22) are second class and are used to reduce the infinite dimensional phase space to finite one leading to the ordinary Ostrogradski Hamiltonian formalism. The operator $`D_t`$ defined in (21) becomes a time evolution operator for $`q`$’s using the first set of Hamilton equation
$`\dot{q}^r`$ $`=`$ $`q^{r+1}.`$ (23)
Using this in (20) the constraints $`\phi _m,(0mn1)`$ coincide with the definition of the Ostragradsky momenta
$`p_m`$ $``$ $`{\displaystyle \underset{\mathrm{}=0}{\overset{nm1}{}}}(_t)^{\mathrm{}}{\displaystyle \frac{L(t)}{(_t^{\mathrm{}+m+1}q(t))}},`$ (25)
$`(0mn1).`$
Now they can be solved for $`q^{\mathrm{}},(n\mathrm{}2n1)`$ as functions of canonical paires $`\{q^j,p_j\},(0jn1)`$
$`q^{\mathrm{}}`$ $``$ $`q^{\mathrm{}}(q^0,q^1,\mathrm{},q^{n1},p_0,p_1,\mathrm{},p_{n1}),`$ (27)
$`(n\mathrm{}2n1).`$
They are combined with the constraints $`\phi _{\mathrm{}},(n\mathrm{}2n1)`$
$`\phi _{\mathrm{}}=p_{\mathrm{}}0,(n\mathrm{}2n1)`$ (28)
to form a second class set and can be used to eliminate the canonical pairs $`\{q^{\mathrm{}},p_{\mathrm{}}\}(n\mathrm{}2n1)`$.
If we take into account (23) the constraint (22) for $`m=0`$ is the Euler-Lagrange equation for the original higher derivative Lagrangian,
$`\psi ^0`$ $``$ $`{\displaystyle \underset{\mathrm{}=0}{\overset{n}{}}}(_t)^{\mathrm{}}{\displaystyle \frac{L(t)}{(_t^{\mathrm{}}q(t))}}=0.`$ (29)
The constraints (22) for $`m>0`$ are the time derivatives of the Euler-Lagrange equation (29) expressed in terms of $`q`$’s. For a non-singular theory, all the constraints (22) can be rewritten as
$`q^{\mathrm{}}`$ $``$ $`q^{\mathrm{}}(q^0,q^1,\mathrm{},q^{n1},p_0,p_1,\mathrm{},p_{n1})0,(\mathrm{}2n)`$ (30)
and can be paired with the constraints $`\phi _{\mathrm{}},(\mathrm{}2n)`$
$`\phi _{\mathrm{}}=p_{\mathrm{}}0,(\mathrm{}2n)`$ (31)
forming second class constraints. They are used to eliminate canonical paires $`\{q^{\mathrm{}},p_{\mathrm{}}\}(\mathrm{}2n)`$.
In this way the infinite dimensional phase space is reduced to a finite dimensional one. The reduced phase space is coordinated by $`T^{}J^n=\{q^l,p_l\}`$ with $`l=0,1,\mathrm{},n1`$. All the constraints are second class and we use the iterative property of Dirac bracket. The Dirac bracket for these variables have the standard form,
$`\{q^m,p_n\}^{}`$ $`=`$ $`\delta _{}^{m}{}_{n}{}^{},\{q^m,q^n\}^{}=\{p_m,p_n\}^{}=0.`$ (32)
The Hamiltonian (6) in the reduced space is given by
$`H(t)`$ $`=`$ $`{\displaystyle \underset{m=0}{\overset{n1}{}}}p_m(t)q^{m+1}(t)L(q^0,q^1,\mathrm{},q^n)`$ (33)
where $`q^n`$ is expressed using (27) as a function of the reduced variables in $`T^{}J^n`$. Note that if we consider the limit $`n`$ going to infinity the constraints (20) and (22) do not allow, in general, to reduced the dimensionality of the infinite dimensional phase space of the non-local system via Dirac brackets.
## V Symplectic formulation of the Euler-Lagrange equation
The Hamiltonian formalism presented in the last sections can be cast into a symplectic form as follows. The Poisson brackets (7) correspond to the symplectic two form $`\mathrm{\Omega }\mathrm{\Lambda }^2(T^{}J)`$:
$$\mathrm{\Omega }=𝑑\lambda \delta P(t,\lambda )\delta Q(t,\lambda ),$$
(34)
where $`\delta `$ stands for the functional exterior derivative.
In the constrained phase space $`\mathrm{\Gamma }_1T^{}J`$ defined by (10) only, the induced (pre)symplectic form is:
$$\mathrm{\Omega }_1=\frac{1}{2}𝑑\lambda 𝑑\lambda ^{}\omega (t;\lambda ,\lambda ^{},[Q])\delta Q(t,\lambda )\delta Q(t,\lambda ^{})$$
(35)
where
$$\omega (t;\lambda ,\lambda ^{},[Q])=\chi (\lambda ^{},\lambda )𝑑\sigma \frac{\delta (t;\sigma ,\lambda )}{\delta Q(t,\lambda ^{})}.$$
(36)
The induced Hamiltonian is:
$$H_1(t,[Q])=𝑑\lambda F(t,\lambda ,[Q])Q^{}(t,\lambda )\stackrel{~}{L}(t,[Q]).$$
(37)
The generator of the time evolution (4) is the vector field
$$X([Q])=𝑑\lambda \dot{Q}(t,\lambda )\frac{\delta }{\delta Q(t,\lambda )}.$$
(38)
Now, $`i(X)\mathrm{\Omega }_1+\delta H_1=0`$ gives a first order formulation of the EL equation. Indeed a short calculation yields
$`i(X)\mathrm{\Omega }_1+\delta H_1={\displaystyle 𝑑\sigma (t;\sigma ,0)\delta Q(t,0)}`$ (39)
$`+{\displaystyle 𝑑\lambda 𝑑\lambda ^{}\delta Q(t,\lambda ^{})\left[\dot{Q}(t,\lambda )Q^{}(t,\lambda )\right]\omega (t;\lambda ,\lambda ^{})}.`$ (40)
Whence, the evolution (5) and the EL equation (2) follow from it.
## VI Path integral quantization
Let us consider the Hamiltonian path integral quantization of the $`1+1`$ dimensional field theory associated with the Hamiltonian (6) for $`L^{\mathrm{non}}(t)`$. The path integral is given by
$`{\displaystyle [dP(t,\lambda )][dQ(t,\lambda )]\mu }`$ (41)
$`e^{i{\scriptscriptstyle 𝑑t𝑑\lambda (P(t,\lambda )[\dot{Q}(t,\lambda )Q^{}(t,\lambda )]+\stackrel{~}{L}(t)\delta (\lambda ))}}.`$ (42)
The integration is performed over the reduced phase space $`\mathrm{\Gamma }`$ and the measure $`\mu `$ is
$`\mu `$ $`=`$ $`det\left(\begin{array}{cc}\{\phi ,\phi \}& \{\phi ,\psi \}\\ \{\psi ,\phi \}& \{\psi ,\psi \}\end{array}\right)\delta (\phi )\delta (\psi ).`$ (43)
First we consider the non-singular higher derivative Lagrangian system of order $`n`$. From the discussions of section IV the constraints are arranged to a set in which the canonical variables $`(q^j,p_j)`$ for $`jn`$ are expressed in terms of ones for $`0j<n`$. The measure becomes
$`\mu `$ $`=`$ $`{\displaystyle \underset{j=n}{\overset{2n1}{}}}\{\delta (p_j)\delta (q^j\mathrm{})\}{\displaystyle \underset{k=2n}{\overset{\mathrm{}}{}}}\{\delta (p_k)\delta (q^k\mathrm{}))\}`$ (44)
where … terms are given functions of $`(q^i,p_i),(i=0,\mathrm{},n1)`$. Integrating over $`(q^i,p_i),(in)`$ (42) becomes
$`{\displaystyle \underset{i=0}{\overset{n1}{}}dq^idp_ie^{i{\scriptscriptstyle 𝑑t_{i=0}^{n1}p_i(\dot{q}^iq^{i+1})}+L(q^0,\mathrm{},q^n)}}`$ (45)
where $`q^n`$ is given as a function of $`(q^i,p_i),(i=0,\mathrm{},n1)`$. This is the Hamiltonian path integral of the Ostrogradski formalism. If we assume non-local systems can be regarded as infinite $`n`$ limit of higher derivative system (45) becomes, by taking $`n\mathrm{}`$,
$`{\displaystyle [dP(t,\lambda )][dQ(t,\lambda )]}`$ (46)
$`e^{i{\scriptscriptstyle 𝑑t𝑑\lambda (P(t,\lambda )[\dot{Q}(t,\lambda )Q^{}(t,\lambda )]+\stackrel{~}{L}(t)\delta (\lambda ))}},`$ (47)
where $`Q`$ and $`P`$, which are $`n\mathrm{}`$ of $`(q^i,p_i),(i=0,\mathrm{},n1)`$, are not restricted by the constraints in contrast to (42).
If we integrate out the momenta and using $`\delta (\dot{Q}(t,\lambda )Q^{}(t,\lambda ))`$ we get
$$[dq(t))]e^{i{\scriptscriptstyle 𝑑tL^{\mathrm{non}}(t)}},$$
(48)
which is the Lagrangian path integral formulation for the non-local theory.
## VII Application to space-time non-commutative $`\varphi ^3`$ theory
Space-time non-commutative theories have peculiar properties due to their acausal behavior and lack of unitarity. Here we would like to use the previous formalism to study the question of unitarity in these theories.
To fix the ideas we consider a non-commutative $`\varphi ^3`$ theory with arbitrary non-commutativity in $`d`$ dimensions. The Lagrangian density is given by
$`^{\mathrm{non}}(x^\mu )`$ $`=`$ $`{\displaystyle \frac{1}{2}}_\mu \varphi (x)^\mu \varphi (x){\displaystyle \frac{m^2}{2}}\varphi (x)^2`$ (50)
$`{\displaystyle \frac{g}{3!}}\varphi (x)\varphi (x)\varphi (x)`$
where $``$ is the star product defined by using a general anti-symmetric background $`\theta ^{\mu \nu }`$,
$`f(x)g(x)=[e^{i\frac{\theta ^{\mu \nu }}{2}_\mu ^\alpha _\nu ^\beta }f(x+\alpha )g(x+\beta )]_{\alpha =\beta =0}.`$ (51)
The EL equation is
$`(\mathrm{}m^2)\varphi (x){\displaystyle \frac{g}{2!}}\varphi (x)\varphi (x)=0.`$ (52)
$`x^0`$ in (50)-(52) will be denoted as $`t`$ hereafter. We introduce a ”new coordinate $`x^0`$” which plays the role of $`\lambda `$ in the previous discussion and introduce the field $`Q(t,x^\mu )`$ in $`d+1`$ dimensions. Now $`t`$ is regard as “evolution time” and $`x^\mu :=(x^0,\stackrel{}{x})`$ is a continuous Lorentzian index. Our metric conventions are $`\eta _{tt}=\eta _{00}=1,\eta _{ii}=+1`$. The relation (4) in this case is
$$Q(t,x^0,\stackrel{}{x})=\varphi (t+x^0,\stackrel{}{x}).$$
(53)
The Lagrangian density in $`d+1`$ dimensions for $`Q(t,x^\mu )`$, (see eq.(8)), is
$`(t,x^\mu )`$ $`=`$ $`{\displaystyle \frac{1}{2}}_\mu Q(t,x)^\mu Q(t,x){\displaystyle \frac{m^2}{2}}Q(t,x)^2`$ (55)
$`{\displaystyle \frac{g}{3!}}Q(t,x)Q(t,x)Q(t,x),`$
where now the derivatives in $``$ are with respect to $`x^\mu =(x^0,\stackrel{}{x})`$. Note that this Lagrangian density is local in the evolution time $`t`$.
The momentum constraint (10) is given by
$`\phi (t,x^\mu )=P(t,x^\mu )\delta (x^0)Q^{}(t,x)`$ (56)
$`+{\displaystyle \frac{g}{2!}}{\displaystyle 𝑑x^{}\chi (x^0,x^0)𝑑y_1𝑑y_2}`$ (57)
$`K(y_1x^{},y_2x^{},xx^{})Q(t,y_1)Q(t,y_2),`$ (58)
where $`Q^{}(t,x)`$ denotes $`_{x^0}Q(t,x^\mu )`$. $`K`$ is the symmetric kernel of three star products,
$`f(x)g(x)h(x)={\displaystyle 𝑑y_1𝑑y_2𝑑y_3}`$ (59)
$`K(y_1x,y_2x,y_3x)f(y_1)g(y_2)h(y_3).`$ (60)
The Hamiltonian (6) is
$`H(t)={\displaystyle 𝑑x[P(t,x)Q^{}(t,x)(t,x)\delta (x^0)]}`$ (61)
$`={\displaystyle }dx[P(t,x)Q^{}(t,x)`$ (62)
$`+\delta (x^0)\{{\displaystyle \frac{1}{2}}Q^{}(t,x)^2+{\displaystyle \frac{1}{2}}(Q(t,x))^2+{\displaystyle \frac{m^2}{2}}Q(t,x)^2`$ (63)
$`+{\displaystyle \frac{g}{3!}}Q(t,x)Q(t,x)Q(t,x)\}].`$ (64)
The Hamilton equations are
$`\dot{Q}(t,x)=Q^{}(t,x),`$ (65)
$`\dot{P}(t,x)=P^{}(t,x)\delta ^{}(x^0)[Q^{}(t,x)]_{x^0=0}`$ (66)
$`+\delta (x^0)\{^2Q(t,x)m^2Q(t,x)\}`$ (67)
$`{\displaystyle \frac{g}{2!}}{\displaystyle 𝑑x^{}𝑑y_1𝑑y_2\delta (x^0)}`$ (68)
$`K(y_1x^{},y_2x^{},xx^{})Q(t,y_1)Q(t,y_2).`$ (69)
The stability of the constraint implies the new constraints (13)
$`\psi (t,x)`$ $``$ $`(^2_{x^0}^2m^2)Q(t,x){\displaystyle \frac{g}{2!}}Q(t,x)Q(t,x)`$ (70)
$`=`$ $`0,atx^0=0.`$ (71)
By requiring further consistency we have an infinite number of constraints which can be written as (14)
$`\psi (t,x)`$ $`=`$ $`0\mathrm{for}\mathrm{}<x^0<\mathrm{}.`$ (72)
Using Hamilton equation (65) for $`Q`$, (72) becomes the EL equation
$`(^2_t^2m^2)Q(t,x){\displaystyle \frac{g}{2!}}Q(t,x)Q(t,x)=0,`$ (73)
where $`_{x^0}`$ on $`Q`$ is replaced by $`_t`$ both in the first term and in the $``$ product. It is the original non-local EL equation (52).
If we write the symplectic form and the Hamiltonian in terms of $`Q(t,x)`$, eqs. (37) and (35), we have
$`\mathrm{\Omega }`$ $`=`$ $`{\displaystyle 𝑑x\delta (x^0)\delta Q^{}(t,x)}\delta Q(t,x)`$ (74)
$``$ $`{\displaystyle \frac{g}{4}}{\displaystyle 𝑑x\delta (Q(t,x)Q(t,x))ϵ(x^0)}\delta Q(t,x)`$ (75)
and
$`H={\displaystyle 𝑑x\frac{\delta (x^0)}{2}\{Q^{}(t,x)^2+(Q(t,x))^2+m^2Q(t,x)^2\}}`$ (76)
$`{\displaystyle \frac{g}{4}}{\displaystyle 𝑑x(Q(t,x)Q(t,x))ϵ(x^0)Q^{}(t,x)}.`$ (77)
These expressions can be rewritten in terms of $`\varphi (x)`$ using (65) i.e. (53). In particular the interaction Hamiltonian becomes
$$H_i=\frac{g}{4}𝑑x(\varphi (x)\varphi (x))ϵ(x^0)\dot{\varphi }(x).$$
(78)
Note that the occurrence of time derivatives of any order in the interaction Hamiltonian is not forbidden in non-local theories. This property is clearly not fulfilled by local theories.
Now we can perform the path integral quantization using (48) to obtain
$$[d\varphi (x)]e^{{\scriptscriptstyle 𝑑x({\scriptscriptstyle \frac{1}{2}}_\mu \varphi (x)^\mu \varphi (x){\scriptscriptstyle \frac{m^2}{2}}\varphi (x)^2{\scriptscriptstyle \frac{g}{3!}}\varphi (x)\varphi (x)\varphi (x))}}.$$
(79)
From which we read the Lagrangian Feynman rules, . They coincide with the ones used in . Therefore, it follows from and that non-commutative $`\varphi ^3`$ theory with time-like non-commutativity is not unitary while non-commutative $`\varphi ^3`$ theory with light-like non-commutativity is unitary.
Note Added: Recently, reference has also considered the Hamiltonian formalism for non-local theories.
We acknowledge discussions with Luis Alvarez-Gaumé, José Barbón, Jaume Gomis, Esperanza Lopez, Karl Landsteiner and to Luis Alvarez-Gaumé for a careful reading of the manuscript. The work of J.G is partially supported by AEN98-0431, GC 1998SGR (CIRIT), the work of J.Ll. is supported by DIGICyT, c. no. PPB96-0384 and by IEC(SCF), and K.K. is partially supported by the Grant-in-Aid for Scientific Research, No.12640258 (Ministry of Education Japan).
|
warning/0006/physics0006008.html
|
ar5iv
|
text
|
# Successive approximations for charged particle motion
## I Introduction
The well known Lagrange variational principle requires
$$\delta 𝑑t=\delta [\stackrel{}{\stackrel{~}{p}}\dot{\stackrel{}{q}}]𝑑t=0$$
(1)
with the Lagrangian $``$, Hamiltonian $``$, and generalized momenta $`\stackrel{}{\stackrel{~}{p}}`$ and coordinates $`\stackrel{}{q}`$.
In this principle all variations of $`\stackrel{}{q}(t)`$ are allowed and therefore the Euler–Lagrange equations of motion hold,
$$\frac{d}{dt}_{\dot{\stackrel{}{q}}}(\stackrel{}{q},\dot{\stackrel{}{q}},t)=_\stackrel{}{q}(\stackrel{}{q},\dot{\stackrel{}{q}},t).$$
(2)
For relativistic single particle motion the Lagrangian is
$$=mc\sqrt{c^2\dot{\stackrel{}{r}}^2}+e\dot{\stackrel{}{r}}\stackrel{}{A}e\mathrm{\Phi }$$
(3)
where the position $`\stackrel{}{r}(\stackrel{}{q})`$ is a function of the generalized coordinates $`\stackrel{}{q}`$. The Jacobian matrix $`\underset{¯}{r}`$ of this function can be written in the form $`\underset{¯}{r}=(_\stackrel{}{q}\stackrel{}{r}^T)^T`$ and has the elements $`r_{ij}=_{q_j}r_i`$. In this efficient notation $`\stackrel{}{r}^T`$ is the transpose
of the $`3\times 1`$ matrix $`\stackrel{}{r}`$. The Jacobian matrix of the function $`\dot{\stackrel{}{r}}(\dot{\stackrel{}{q}})`$ is also $`\underset{¯}{r}`$ since $`\dot{\stackrel{}{r}}=_{i=1}^3\dot{q_i}_{q_i}\stackrel{}{r}=\underset{¯}{r}\dot{\stackrel{}{q}}`$.
The generalized momentum is $`\stackrel{}{\stackrel{~}{p}}=_{\dot{\stackrel{}{q}}}=\underset{¯}{r}^T(m\gamma \dot{\stackrel{}{r}}+e\stackrel{}{A})`$ and the variational principle can thus be written as
$`\delta {\displaystyle [\stackrel{}{\stackrel{~}{p}}^T\underset{¯}{r}^1\dot{\stackrel{}{r}}]𝑑t}`$ $`=`$ $`\delta {\displaystyle [(m\gamma \dot{\stackrel{}{r}}+e\stackrel{}{A})^T\underset{¯}{r}\underset{¯}{r}^1\dot{\stackrel{}{r}}]𝑑t}`$ (4)
$`=`$ $`\delta {\displaystyle [m\gamma v^2+e\stackrel{}{A}^T\dot{\stackrel{}{r}}]𝑑t}.`$ (5)
If only variations $`\delta _{=E}`$ are considered which keep the total energy $`=E`$ constant, the variational principle becomes
$`\delta _{_{=E}}{\displaystyle [\stackrel{}{\stackrel{~}{p}}\dot{\stackrel{}{q}}]𝑑t}`$ $`=`$ $`\delta _{_{=E}}{\displaystyle \stackrel{}{\stackrel{~}{p}}𝑑\stackrel{}{q}}`$ (6)
$`=\delta _{_{=E}}{\displaystyle [m\gamma v^2+e\stackrel{}{A}^T\dot{\stackrel{}{r}}]𝑑t}`$ $`=`$ $`0.`$ (7)
The variational principle for constant total energy is called the principle of Maupertuis. However, in equation (7) it does not lead to Euler–Lagrange equations of motion, since not all variations are allowed.
A particle optical device usually has an optical axis or some design curve along which a particle beam should travel. This design curve $`\stackrel{}{R}(l)`$ is parameterized by a variable $`l`$ and the position of a particle in the vicinity of the design curve has coordinates $`x`$ and $`y`$ along the unit vectors $`\stackrel{}{e}_x`$ and $`\stackrel{}{e}_y`$ in a plane perpendicular this curve. This coordinate system is shown in figure 1. The third coordinate vector $`\stackrel{}{e}_l=d\stackrel{}{R}/dl`$ is tangential to the design curve and the curvature vector is $`\stackrel{}{\kappa }=d\stackrel{}{e}_l/dl`$.
The unit vectors $`\stackrel{}{e}_x`$ and $`\stackrel{}{e}_y`$ in the usual Frenet–Serret comoving coordinate system rotate with the torsion of the design curve. If this rotation is wound back, the equations of motion do not contain the torsion of the design curve. The position and the velocity are
$$\stackrel{}{r}=x\stackrel{}{e}_x+y\stackrel{}{e}_y+\stackrel{}{R}(l),\dot{\stackrel{}{r}}=\dot{x}\stackrel{}{e}_x+\dot{y}\stackrel{}{e}_y+h\dot{l}\stackrel{}{e}_l,$$
(8)
with $`h=1+x\kappa _x+y\kappa _y`$. This method is described in and and is mentioned here since design curves with torsion are becoming important when considering particle motion in helical wigglers, undulators, and wavelength shifters , and for polarized particle motion in helical dipole Siberian Snakes .
The variational principle (7) for the three generalized coordinates $`x(t)`$, $`y(t)`$, and $`l(t)`$ can now be written for the two generalized coordinates $`x(l)`$ and $`y(l)`$. This has the following two advantages: a) The particle trajectory along the design curve is usually more important than the particle position at a time $`t`$, and b) Whereas $`\delta _{_{=E}}`$ does not allow for all variations of the three coordinates, the total energy can be conserved for all variations of the two coordinates $`x`$ and $`y`$ by choosing for each position $`\stackrel{}{r}`$ the appropriate momentum with $`m\gamma v=\sqrt{(Ee\mathrm{\Phi }(\stackrel{}{r}))^2/c^2(mc)^2}`$. We obtain from equation (7)
$$\delta _{_{=E}}\stackrel{}{\stackrel{~}{p}}𝑑\stackrel{}{q}=\delta [m\gamma v^2\frac{dt}{dl}+ev\stackrel{}{A}\frac{d\stackrel{}{r}}{dl}]𝑑l=0$$
(9)
with $`d\stackrel{}{r}/dl=x^{}\stackrel{}{e}_x+y^{}\stackrel{}{e}_y+h\stackrel{}{e}_l`$ and $`dt/dl=|d\stackrel{}{r}/dl|/v`$. Since all variations are allowed, the integrand is a very simple new Lagrangian
$$\stackrel{~}{L}=m\gamma v\sqrt{x_{}^{}{}_{}{}^{2}+y_{}^{}{}_{}{}^{2}+h^2}+e(x^{}A_x+y^{}A_y+hA_l)$$
(10)
which leads to Euler–Lagrange equations of motion
$`\stackrel{~}{p}_x`$ $`=`$ $`_x^{}\stackrel{~}{L},\stackrel{~}{p}_{x}^{}{}_{}{}^{}=_x\stackrel{~}{L},`$ (11)
$`\stackrel{~}{p}_y`$ $`=`$ $`_y^{}\stackrel{~}{L},\stackrel{~}{p}_{y}^{}{}_{}{}^{}=_y\stackrel{~}{L}.`$ (12)
The integral $`_0^l\stackrel{~}{L}(\stackrel{~}{l})𝑑\stackrel{~}{l}`$ is called the eikonal.
Since the Hamiltonian formulation is very common in the area of accelerator physics, we will show how the eikonal can be derived from a Hamiltonian formulation.
The equations of motion for the three generalized coordinates $`x(t)`$, $`y(t)`$, and $`l(t)`$ can be obtained from the Hamiltonian
$`=e\mathrm{\Phi }+`$ (13)
$`\sqrt{m^2c^2+(\stackrel{~}{p}_xeA_x)^2+(\stackrel{~}{p}_yeA_y)^2+(\stackrel{~}{p}_l/heA_l)^2}.`$ (14)
In the case of time independent fields, $``$ is the conserved total energy $`E`$ and there are only five independent variables, rather than six. Note that the velocity dependent or non holonomic boundary condition $`(\stackrel{}{p}(\stackrel{}{q},\dot{\stackrel{}{q}}),\stackrel{}{q},t)=E`$ cannot be included in the Lagrange formalism directly. But in the Hamilton formalism this can be done. Furthermore, a switch of independent variable from $`t`$ to $`l`$ can easily be done in the Hamiltonian formulation. The Lagrange formulation is therefore abandoned (too easily, as will be shown later). In the variational condition
$$\delta [\dot{x}\stackrel{~}{p}_x+\dot{y}\stackrel{~}{p}_y+\dot{l}\stackrel{~}{p}_l]𝑑t=0$$
(15)
one can change to the independent variable $`l`$ as follows:
$$\delta [x^{}\stackrel{~}{p}_x+y^{}\stackrel{~}{p}_y+(t^{})(\stackrel{~}{p}_l)]𝑑l=0.$$
(16)
The six canonical coordinates are now $`x`$, $`\stackrel{~}{p}_x`$, $`y`$, $`\stackrel{~}{p}_y`$, $`t`$, and $``$, and the new Hamiltonian is given by $`\stackrel{~}{H}=\stackrel{~}{p}_l`$ which has to be expressed as a function of the six coordinates ,
$`\stackrel{~}{H}=h[eA_l+`$ (17)
$`\sqrt{(e\mathrm{\Phi })^2(mc^2)^2(\stackrel{~}{p}_xeA_x)^2(\stackrel{~}{p}_yeA_y)^2}].`$ (18)
In the Hamilton formalism it is simple to take advantage of the fact that the total energy is conserved for time independent fields; $`^{}=_t\stackrel{~}{H}=0`$ leads to $`=E`$. Then from the six coordinates only the first four have to be considered, leading to the Lagrangian
$$\stackrel{~}{L}=x^{}\stackrel{~}{p}_x+y^{}\stackrel{~}{p}_y\stackrel{~}{H}.$$
(19)
From $`x^{}=_{\stackrel{~}{p}_x}\stackrel{~}{H}=\frac{h}{\sqrt{}}(\stackrel{~}{p}_xeA_x)`$, $`y^{}=_{\stackrel{~}{p}_y}\stackrel{~}{H}=\frac{h}{\sqrt{}}(\stackrel{~}{p}_yeA_y)`$ where $`\sqrt{}`$ is the square root in $`\stackrel{~}{H}`$ one obtains
$`\sqrt{}`$ $`=`$ $`h\sqrt{{\displaystyle \frac{(Ee\mathrm{\Phi })^2(mc^2)^2}{x^2+y^2+h^2}}}`$ (20)
$`=`$ $`m\gamma v{\displaystyle \frac{h}{\sqrt{x^2+y^2+h^2}}},`$ (21)
$`\stackrel{~}{p}_x`$ $`=`$ $`{\displaystyle \frac{\sqrt{}}{h}}x^{}+eA_x,\stackrel{~}{p}_y={\displaystyle \frac{\sqrt{}}{h}}y^{}+eA_y,`$ (22)
$$\stackrel{~}{L}=m\gamma v\sqrt{x_{}^{}{}_{}{}^{2}+y_{}^{}{}_{}{}^{2}+h^2}+e(x^{}A_x+y^{}A_y+hA_l)$$
(23)
for $`m\gamma v=\sqrt{(Ee\mathrm{\Phi })^2/c^2(mc)^2}`$. The very simple Lagrangian $`\stackrel{~}{L}`$ agrees with the integrand (10) of the eikonal.
In the following it will be shown how the Hamiltonian and the Lagrangian equations of motion for the particle trajectory $`\stackrel{}{q}(l)`$ can be solved in an iterative way. We write a general equation of motion for a coordinate vector $`\stackrel{}{z}`$ in the form
$$\stackrel{}{z}^{}=\stackrel{}{f}^1(\stackrel{}{z},l)+\stackrel{}{f}^2(\stackrel{}{z},l)$$
(24)
where we assume that $`\stackrel{}{z}=0`$ is a solution of the differential equation. Furthermore, we assume $`\stackrel{}{z}`$ to be small and let $`\stackrel{}{f}^1`$ be linear in the coordinates. We assume that the nonlinear part of the equation of motion can be expanded in a Taylor series $`\stackrel{}{f}^2`$. The linearized equation of motion is solved by a trajectory $`\stackrel{}{z}_1(l)=\underset{¯}{M}(l)\stackrel{}{z}_i`$ which depends linearly on the initial coordinates. For the transport matrix $`\underset{¯}{M}(l)`$ we therefore have
$$\underset{¯}{M}^{}\stackrel{}{z}_i=\underset{¯}{f^1}\underset{¯}{M}\stackrel{}{z}_i$$
(25)
for all coordinate vectors $`\stackrel{}{z}_i`$; $`\underset{¯}{f^1}`$ being the Jacobian matrix of $`\stackrel{}{f}^1`$.
One can write every solution of (24) as $`\stackrel{}{z}(l)=\underset{¯}{M}(l)\stackrel{}{\zeta }(l)`$, leading to the equation of motion
$$\underset{¯}{M}^{}\stackrel{}{\zeta }+\underset{¯}{M}\stackrel{}{\zeta }^{}=\underset{¯}{f^1}\underset{¯}{M}\stackrel{}{\zeta }+\stackrel{}{f}^2(\stackrel{}{z}).$$
(26)
The Taylor coefficients of $`\stackrel{}{\zeta }(\stackrel{}{z}_i,l)`$ with respect to the initial coordinates $`\stackrel{}{z}_i=\stackrel{}{\zeta }(0)`$ are called aberration coefficients. With equation (25) one obtains
$$\stackrel{}{z}(l)=\underset{¯}{M}(l)\{\stackrel{}{z}_i+_0^l\underset{¯}{M}^1(\stackrel{~}{l})\stackrel{}{f}^2(\stackrel{}{z}(\stackrel{~}{l}))\}d\stackrel{~}{l}.$$
(27)
Now we assume that the general solution $`\stackrel{}{z}(\stackrel{}{z}_i,l)`$ can be expanded in a power series with respect to the initial coordinates. Then symbolizing the $`j`$th order Taylor polynomial with $`[\mathrm{}]_j`$, we write the orders up to $`j`$ as $`\stackrel{}{z}_j=[\stackrel{}{z}(\stackrel{}{z}_i,l)]_j`$, i.e. we use lower indices to describe the order of $`\stackrel{}{z}_i`$. The upper index in $`\stackrel{}{f}`$ describes the order in $`\stackrel{}{z}`$, which is in turn a nonlinear function of $`\stackrel{}{z}_i`$. When $`\stackrel{}{z}_{n1}`$ is known, one can iterate the expansion up to order $`n`$ with equation (27), since
$$\stackrel{}{z}_n=\underset{¯}{M}(l)\{\stackrel{}{z}_i+_0^l\underset{¯}{M}^1(\stackrel{~}{l})[\stackrel{}{f}^2(\stackrel{}{z}_{n1})]_n\}d\stackrel{~}{l}.$$
(28)
The zeroth order of the expansion with respect to the coordinates must vanish, which means that the trajectory $`\stackrel{}{q}=0`$ must satisfy the equation of motion for some momentum $`p(l)`$. Additionally we require that the vector potential on the design curve is gauged to zero. This can always be achieved. The canonical momentum $`\stackrel{}{p}`$ then also vanishes for the trajectory $`\stackrel{}{q}=0`$. It then follows that the Hamiltonian and the Lagrangian have no components linear in the coordinates and momenta. When computing trajectories through a particle optical device, it is customary to normalize the momenta to the initial design momentum $`p_0=p(0)`$. The following two dimensional generalized coordinates are therefore used:
$`\stackrel{}{q}`$ $`=`$ $`\left({\displaystyle \genfrac{}{}{0pt}{}{x}{y}}\right),\stackrel{}{p}=\left({\displaystyle \genfrac{}{}{0pt}{}{\stackrel{~}{p}_x/p_0}{\stackrel{~}{p}_y/p_0}}\right),`$ (29)
$`L(\stackrel{}{q},\stackrel{}{q}^{},l)`$ $`=`$ $`\stackrel{~}{L}/p_0,\stackrel{}{p}=_\stackrel{}{q}^{}L,\stackrel{}{p}^{}=_\stackrel{}{q}L,`$ (30)
$`H(\stackrel{}{q},\stackrel{}{p},l)`$ $`=`$ $`\stackrel{~}{H}/p_0,\stackrel{}{q}^{}=_\stackrel{}{p}H,\stackrel{}{p}^{}=_\stackrel{}{q}H.`$ (31)
The Euler–Lagrange equations lead to the second order differential equations $`\frac{d}{dl}_\stackrel{}{q}^{}L=_\stackrel{}{q}L`$ for the two dimensional vector $`\stackrel{}{q}`$.
## II Successive approximation in terms of Hamiltonians
In the Hamilton formalism one obtains first order equations of motion for the four dimensional vector $`\stackrel{}{z}^T=(q_1,q_2,p_1,p_2)`$. With the antisymmetric matrix $`\underset{¯}{J}`$ one can write the equation of motion as
$$\underset{¯}{J}=\left(\begin{array}{cc}\hfill \underset{¯}{0}_2& \hfill \underset{¯}{1}_2\\ \hfill \underset{¯}{1}_2& \hfill \underset{¯}{0}_2\end{array}\right),\stackrel{}{z}^{}=\underset{¯}{J}_\stackrel{}{z}H,$$
(32)
with the $`2\times 2`$ identity and zero matrixes $`\underset{¯}{1}_2`$ and $`\underset{¯}{0}_2`$. This structure implies special symmetries for the transport maps $`\stackrel{}{}`$ of particle optics. These maps describe how the final phase space coordinates $`\stackrel{}{z}_f=\stackrel{}{}(\stackrel{}{z}_i)`$ of a particle, after flying through an optical device, are related to the initial coordinates $`\stackrel{}{z}_i`$. These maps are often weakly nonlinear and can be expanded in a Taylor expansion. The Hamiltonian nature implies that the Jacobian $`\underset{¯}{}=(_\stackrel{}{z}\stackrel{}{}^T)^T`$ of any transport map $`\stackrel{}{}(\stackrel{}{z})`$ is symplectic , meaning that
$$\underset{¯}{}\underset{¯}{J}\underset{¯}{}^T=\underset{¯}{J}.$$
(33)
For the successive approximations we separate the equation of motion into its linear and nonlinear part,
$$\stackrel{}{z}^{}=\underset{¯}{J}_\stackrel{}{z}(H^2+H^3).$$
(34)
After we have solved for the linear transport matrix $`\stackrel{}{z}_1=\underset{¯}{M}\stackrel{}{z}_i`$, we can iterate by equation (28) which takes the form
$$\stackrel{}{z}_n=\underset{¯}{M}\{\stackrel{}{z}_i+_0^l\underset{¯}{M}^1[\underset{¯}{J}_\stackrel{}{z}H^3(\stackrel{}{z}_{n1})]_n𝑑\stackrel{~}{l}\}$$
(35)
With the relation $`\underset{¯}{}^1\underset{¯}{J}=\underset{¯}{J}\underset{¯}{}^T`$ from equation (33) this can be written as
$$\stackrel{}{z}_n=\underset{¯}{M}\{\stackrel{}{z}_i+\underset{¯}{J}_0^l\underset{¯}{M}^T[_\stackrel{}{z}H^3(\stackrel{}{z}_{n1})]_n𝑑\stackrel{~}{l}\}.$$
(36)
The corresponding equation for the aberrations $`\stackrel{}{\zeta }_n=\underset{¯}{M}^1\stackrel{}{z}_n`$ becomes
$$\stackrel{}{\zeta }_n=\stackrel{}{z}_i+\underset{¯}{J}_0^l[_\stackrel{}{\zeta }H^3(\underset{¯}{M}\stackrel{}{\zeta }_{n1})]_n𝑑\stackrel{~}{l}.$$
(37)
This form of the iteration equation is quite simple. However, since the Hamiltonian (17) is a complicated function, the evaluation of the four integrals can become very cumbersome.
## III Successive approximation in terms of Lagrangians
In Rose used a variational principle to derive a successive approximation to nonlinear motion based on the eikonal. This method iterates position $`\stackrel{}{q}`$ and momentum $`\stackrel{}{p}`$ in their nonlinear dependence on the initial position $`\stackrel{}{q}_i`$ and momentum $`\stackrel{}{p}_i`$. Knowing the order $`n1`$ dependence $`\stackrel{}{q}_{n1}`$ and $`\stackrel{}{p}_{n1}`$, one has to compute $`\stackrel{}{q}_{n1}^{}`$ by differentiation of $`\stackrel{}{q}_{n1}`$ or by inversion of $`\stackrel{}{p}=_\stackrel{}{q}^{}L(\stackrel{}{q},\stackrel{}{q}^{},l)`$. Then the eikonal can be evaluated to compute the order $`n`$ dependence $`\stackrel{}{q}_n`$ and $`\stackrel{}{p}_n`$. In general it can be cumbersome to compute $`\stackrel{}{q}_{n1}^{}`$ and therefore here we derive a new version of the eikonal method, which iterates directly $`\stackrel{}{q}_n^{}`$ rather than the momentum.
In deriving the simple form of equation (37), advantage has only been taken of the symplectic first order transfer matrix. We therefore wish to exploit this advantage again by working with new coordinates which are identical with the canonical $`\stackrel{}{q}`$ and $`\stackrel{}{p}`$ up to first order so that the new coordinates lead to the same first order transport matrix $`\underset{¯}{M}`$. To first order one obtains
$$\stackrel{}{p}=_\stackrel{}{q}^{}L=\frac{p(s)}{p_0}\stackrel{}{q}^{}+\frac{e}{p_0}\left(\genfrac{}{}{0pt}{}{A_x^1}{A_y^1}\right)+𝒪^2(\stackrel{}{q},\stackrel{}{q}^{})$$
(38)
where $`p(s)`$ is the momentum of a particle traveling on the design curve $`\stackrel{}{q}=0`$, and the upper index 1 specifies the part of the vector potential linear in $`x`$ and $`y`$. We therefore work with the coordinates
$$\stackrel{}{Q}=\left(\genfrac{}{}{0pt}{}{\stackrel{}{q}}{\stackrel{}{u}}\right)=\left(\genfrac{}{}{0pt}{}{\stackrel{}{q}}{\frac{p(s)}{p_0}\stackrel{}{q}^{}+\frac{e}{p_0}\left(\genfrac{}{}{0pt}{}{A_x^1}{A_y^1}\right)}\right).$$
(39)
Moreover, it can be shown that the contribution from the vector potential can be gauged to vanish whenever there is no longitudinal magnetic field $`B_0\stackrel{}{e}_l`$ on the design curve. Then if one investigates trajectories which start with momentum $`p_0`$ in a region free of such a field, we have the simple relation $`\stackrel{}{u}_i=\stackrel{}{q}_i^{}`$.
By splitting the Lagrangian into its second order and its higher order part, the equation of motion becomes
$`\stackrel{}{Q}^{}`$ $`=`$ $`\left({\displaystyle \genfrac{}{}{0pt}{}{\stackrel{}{q}}{\stackrel{}{u}}}\right)^{}`$ (40)
$`=`$ $`\left({\displaystyle \genfrac{}{}{0pt}{}{\frac{p_0}{p(s)}\stackrel{}{u}\frac{e}{p(s)}\left(\genfrac{}{}{0pt}{}{A_x^1}{A_y^1}\right)}{_\stackrel{}{q}L^2}}\right)\left({\displaystyle \genfrac{}{}{0pt}{}{0}{\frac{d}{d\stackrel{~}{l}}_\stackrel{}{q}^{}L^3_\stackrel{}{q}L^3}}\right).`$ (41)
After having solved the linearized equation of motion, we obtain with equation (28)
$`\stackrel{}{Q}`$ $`=`$ $`\underset{¯}{M}\{\stackrel{}{Q}_i{\displaystyle _0^l}\underset{¯}{M}^1\left({\displaystyle \genfrac{}{}{0pt}{}{0}{\frac{d}{d\stackrel{~}{l}}_\stackrel{}{q}^{}L^3_\stackrel{}{q}L^3}}\right)𝑑\stackrel{~}{l}\}`$ (42)
$`=`$ $`\underset{¯}{M}\{\stackrel{}{Q}_i+\underset{¯}{J}{\displaystyle _0^l}\underset{¯}{M}^T\left({\displaystyle \genfrac{}{}{0pt}{}{\frac{d}{d\stackrel{~}{l}}_\stackrel{}{q}^{}L^3_\stackrel{}{q}L^3}{0}}\right)𝑑\stackrel{~}{l}\}.`$ (43)
An integration by parts leads to
$`\stackrel{}{Q}`$ $`=`$ $`\underset{¯}{M}\{\stackrel{}{Q}_i\underset{¯}{J}{\displaystyle _0^l}[\underset{¯}{M}_{}^{}{}_{}{}^{T}\left({\displaystyle \genfrac{}{}{0pt}{}{_\stackrel{}{q}^{}L^3}{0}}\right)+\underset{¯}{M}^T\left({\displaystyle \genfrac{}{}{0pt}{}{_\stackrel{}{q}L^3}{0}}\right)]d\stackrel{~}{l}`$ (45)
$`+\underset{¯}{J}\left[\underset{¯}{M}^T\left({\displaystyle \genfrac{}{}{0pt}{}{_\stackrel{}{q}^{}L^3}{0}}\right)\right]_0^l\}.`$
Writing the Jacobian as $`\underset{¯}{M}^T=_{\stackrel{}{Q}_i}\stackrel{}{Q}_1^T=_{\stackrel{}{Q}_i}(\stackrel{}{Q}\stackrel{}{Q}_2)^T`$ where $`\stackrel{}{Q}=\stackrel{}{Q}_1+\stackrel{}{Q}_2`$ was split into parts which depend on $`\stackrel{}{Q}_i`$ linearly and nonlinearly, we obtain
$`\underset{¯}{M}^1\stackrel{}{Q}=\stackrel{}{Q}_i\underset{¯}{J}{\displaystyle _0^l}[(_{\stackrel{}{Q}_i}(\stackrel{}{q}^T\stackrel{}{q}_2^T))_\stackrel{}{q}^{}L^3`$ (46)
$`+(_{\stackrel{}{Q}_i}(\stackrel{}{q}^T\stackrel{}{q}_2^T))_\stackrel{}{q}L^3]d\stackrel{~}{l}+\underset{¯}{J}[(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^3]_0^l`$ (47)
$`=\stackrel{}{Q}_i\underset{¯}{J}{\displaystyle _0^l}[_{\stackrel{}{Q}_i}L^3(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}^{}(LL^2)`$ (48)
$`(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}(LL^2)]d\stackrel{~}{l}+\underset{¯}{J}[(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^3]_0^l`$ (49)
$`=\stackrel{}{Q}_i\underset{¯}{J}{\displaystyle _0^l}[_{\stackrel{}{Q}_i}L^3+(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}^{}L^2+(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}L^2`$ (50)
$`(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}^{}L(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T){\displaystyle \frac{d}{d\stackrel{~}{l}}}_\stackrel{}{q}^{}L]d\stackrel{~}{l}`$ (51)
$`+\underset{¯}{J}\left[(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^3\right]_0^l.`$ (52)
Note that the $`A_x^{}`$, $`A_y^{}`$, $`p(s)`$ and $`p_0`$ of equation (39) drop out of the right hand side of equation (52) owing to the multiplications by the zeros in equation (45). The second order $`L^2`$ of the Lagrangian is a quadratic form in which every quadratic combination of the $`\stackrel{}{q}`$ and $`\stackrel{}{q}^{}`$ can occur. It can be written using a matrix $`\underset{¯}{L^2}`$ as $`L^2=\stackrel{}{Q}^T\underset{¯}{L^2}\stackrel{}{Q}`$. Part of the above integrand can be rewritten as
$$\stackrel{}{q}_2^T_\stackrel{}{q}^{}L^2+\stackrel{}{q}_2^T_\stackrel{}{q}L^2=\stackrel{}{Q}_2^T\underset{¯}{L^2}\stackrel{}{Q}+\stackrel{}{Q}^T\underset{¯}{L^2}\stackrel{}{Q}_2.$$
(53)
For convenience we write $`L^2(\stackrel{}{a})=\stackrel{}{a}^T\underset{¯}{L^2}\stackrel{}{a}`$. Another integration by parts in equation (52) leads to
$`\underset{¯}{M}^1\stackrel{}{Q}`$ $`=`$ $`\stackrel{}{Q}_i\underset{¯}{J}{\displaystyle _0^l}[_{\stackrel{}{Q}_i}(L^3+L^2(\stackrel{}{Q}_2))`$ (56)
$`+(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_{\stackrel{}{q}_1^{}}L^2(\stackrel{}{Q}_1)+(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_{\stackrel{}{q}_1}L^2(\stackrel{}{Q}_1)]d\stackrel{~}{l}`$
$`+\underset{¯}{J}\left[(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}^{}L+(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^3\right]_0^l.`$
The first part of the integral contains $`L_E=L^3+L^2(\stackrel{}{Q}_2)`$. The integral $`_0^lL_E𝑑\stackrel{~}{l}`$ is called the perturbation eikonal. This scheme embodies the essential requirement that the $`n+1`$ order dependence of $`L_E`$ on the initial variables $`\stackrel{}{Q}_i`$ can be computed already when $`Q_{n1}`$ is known; $`\stackrel{}{Q}_n`$ does not need to be known. For an iteration of $`\stackrel{}{Q}_n`$, knowledge of $`\stackrel{}{Q}_{n1}`$ is sufficient. Since $`\stackrel{}{q}_1`$ satisfies the first order equation of motion, we can use the relation $`_{\stackrel{}{q}_1}L^2(\stackrel{}{Q}_1)=\frac{d}{dl}_{\stackrel{}{q}_1^{}}L^2(\stackrel{}{Q}_1)`$ to perform another integration by parts,
$`\underset{¯}{M}^1\stackrel{}{Q}`$ $`=`$ $`\stackrel{}{Q}_i\underset{¯}{J}_{\stackrel{}{Q}_i}{\displaystyle _0^l}L_Ed\stackrel{~}{l}+\underset{¯}{J}[(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_\stackrel{}{q}^{}L`$ (58)
$`+(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^3(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_{\stackrel{}{q}_1^{}}L^2(\stackrel{}{Q}_1)]_0^l`$
$`=`$ $`\stackrel{}{Q}_i\underset{¯}{J}_{\stackrel{}{Q}_i}{\displaystyle _0^l}L_Ed\stackrel{~}{l}+\underset{¯}{J}[(_{\stackrel{}{Q}_i}\stackrel{}{q}^T)_\stackrel{}{q}^{}L`$ (60)
$`(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^2(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_{\stackrel{}{q}_1^{}}L^2(\stackrel{}{Q}_1)]_0^l.`$
$`_{\stackrel{}{Q}_i}L_E`$ is the part of $`_{\stackrel{}{Q}_i}L`$ which up to order $`n`$ in $`\stackrel{}{Q}_i`$ does not depend on $`\stackrel{}{Q}_n`$. Similarly the term outside the integral is simply the part of $`(_{\stackrel{}{Q}_i}\stackrel{}{q}^T)_\stackrel{}{q}^{}L`$ which up to order $`n`$ does not depend on $`\stackrel{}{Q}_n`$. We therefore write $`(_{\stackrel{}{Q}_i}\stackrel{}{q}^T)_\stackrel{}{q}^{}L(_{\stackrel{}{Q}_i}\stackrel{}{q}_1^T)_\stackrel{}{q}^{}L^2(_{\stackrel{}{Q}_i}\stackrel{}{q}_2^T)_{\stackrel{}{q}_1^{}}L^2(\stackrel{}{Q}_1)=\{(_{\stackrel{}{Q}_i}\stackrel{}{q}^T)_\stackrel{}{q}^{}L\}_E`$. If we now express the Lagrangian in terms of the aberrations $`\stackrel{}{\xi }`$ with $`\stackrel{}{Q}=\underset{¯}{M}\stackrel{}{\xi }`$, we obtain the iteration equation
$`\stackrel{}{\xi }_n=_n\stackrel{}{Q}_i`$ $``$ $`\underset{¯}{J}_{\stackrel{}{Q}_i}{\displaystyle _0^l}L_E(\underset{¯}{M}\stackrel{}{\xi }_{n1})𝑑\stackrel{~}{l}`$ (61)
$`+`$ $`\underset{¯}{J}\left[\{(_{\stackrel{}{Q}_i}\stackrel{}{q}^T)_\stackrel{}{q}^{}L\}_E\right]_0^l.`$ (62)
When computing $`\stackrel{}{\xi }_n`$ from $`\stackrel{}{\xi }_{n1}`$ with this iteration equation, all parts of the right hand side which contribute to higher orders are neglected, as indicated by $`=_n`$. This iteration equation can have several advantages over the Hamiltonian iteration equation (28):
* The Lagrangian (10) is a much simpler function than the Hamiltonian (17).
* The derivative in equation (61) is performed after the integral has been evaluated. Therefore only one integral has to be computed and it describes all four coordinates of $`\stackrel{}{\xi }_n`$.
* The fact that the various coordinates are the derivatives with respect to initial conditions yields very simple relations between the various expansion coefficients of $`\stackrel{}{\xi }_n`$, which are the so called aberration coefficients of particle optical devices. These relations can be much simpler than relations entailed by the symplectic symmetry implicit in the Hamiltonian formulation.
* The second pair of coordinates in equation (39) can be calculated very easily. With equations (26) and (40), the equation of motion for $`\stackrel{}{\xi }`$ is
$$\underset{¯}{M}\stackrel{}{\xi }^{}=\left(\genfrac{}{}{0pt}{}{0}{\frac{d}{d\stackrel{~}{l}}_\stackrel{}{q}^{}L^3_\stackrel{}{q}L^3}\right).$$
(63)
After having computed $`\stackrel{}{q}_n=\stackrel{}{M}_{2\times 4}\stackrel{}{\xi }_n`$ by iteration, the derivative $`\stackrel{}{q}^{}`$ can then easily be computed as $`\stackrel{}{q}_n^{}=\stackrel{}{M}_{2\times 4}^{}\stackrel{}{\xi }_n`$ using equation (63). One thus only needs to iterate the two dimensional vector $`\stackrel{}{q}_n`$ and not a four dimensional vector $`\stackrel{}{z}_n`$ as in the Hamiltonian iteration procedure.
## IV Successive approximation for spin orbit motion
The time variation of a spin $`\stackrel{}{s}`$ in the rest frame of a particle is described by the so called Thomas–BMT equation $`\dot{\stackrel{}{s}}=\stackrel{}{\mathrm{\Omega }}_{BMT}\times \stackrel{}{s}`$ where
$`\stackrel{}{\mathrm{\Omega }}_{BMT}=`$ $``$ $`{\displaystyle \frac{q}{m\gamma }}\{(a\gamma +1)\stackrel{}{B}_{}+(1+a)\stackrel{}{B}_{}`$ (64)
$``$ $`{\displaystyle \frac{\gamma }{c}}\stackrel{}{\beta }\times \stackrel{}{E}(a+{\displaystyle \frac{1}{1+\gamma }})\}`$ (65)
with the electric field $`\stackrel{}{E}`$, the parts of the magnetic field $`\stackrel{}{B}`$ which are perpendicular ($``$) and parallel ($``$) to the particle’s velocity, and the anomalous gyro-magnetic factor $`a=\frac{g2}{2}`$.
Changing to the comoving coordinate system of figure (1), we obtain $`\stackrel{}{s}=S_x\stackrel{}{e}_x+S_y\stackrel{}{e}_y+S_l\stackrel{}{e}_l`$ and $`\stackrel{}{s}^{}=(S_x^{}S_l\kappa _x)\stackrel{}{e}_x+(S_y^{}S_l\kappa _y)\stackrel{}{e}_y+(S_l^{}+S_x\kappa _x+S_y\kappa _y)\stackrel{}{e}_l`$. The equation of motion for the vector $`\stackrel{}{S}`$ of these spin components is then
$`\stackrel{}{S}^{}`$ $`=`$ $`\stackrel{}{\mathrm{\Omega }}\times \stackrel{}{S},`$ (66)
$`\stackrel{}{\mathrm{\Omega }}`$ $`=`$ $`\stackrel{}{\mathrm{\Omega }}_{BMT}{\displaystyle \frac{h}{v}}\sqrt{x^2+y^2+h^2}\stackrel{}{\kappa }\times \stackrel{}{e}_l.`$ (67)
The equations of motion for the phase space vector $`\stackrel{}{z}`$ and the spin $`\stackrel{}{S}`$ have the form
$$\stackrel{}{z}^{}=\stackrel{}{f}(\stackrel{}{z},l),\stackrel{}{S}^{}=\stackrel{}{\mathrm{\Omega }}(\stackrel{}{z},l)\times \stackrel{}{S}.$$
(68)
The general solutions transporting the coordinates along the optical system, starting at the initial values $`\stackrel{}{z}_i`$, $`\stackrel{}{S}_i`$, is given by the transport map $``$ and the rotation matrix $`\underset{¯}{R}SO(3)`$,
$$\stackrel{}{z}(l)=\stackrel{}{}(\stackrel{}{z}_i,l),\stackrel{}{S}(l)=\underset{¯}{R}(\stackrel{}{z}_i,l)\stackrel{}{S}_i.$$
(69)
In order to find the general solution, one could compute the nine coefficients of the rotation matrix by solving the differential equation
$$R_{ij}(\stackrel{}{z}_i,l)^{}=ϵ_{ilk}\mathrm{\Omega }_lR_{kj}(\stackrel{}{z}_i,l),$$
(70)
where the vector product was expressed by the totally antisymmetric tensor $`ϵ_{ilk}`$. However, computing the nine components of the rotation matrix seems inefficient, since a rotation can be represented by three angles. It has turned out to be most efficient to represent the rotation of spins by the quaternion $`A`$ which gives the rotation transformation in the SU(2) representation as
$$A=a_0\underset{¯}{1}\mathrm{i}\stackrel{}{a}\underset{¯}{\overset{}{\sigma }}.$$
(71)
Here $`\underset{¯}{1}`$ is the $`2\times 2`$ identity matrix and the elements of the vector $`\underset{¯}{\overset{}{\sigma }}`$ are the three two dimensional Pauli matrixes. When a rotation by an angle $`\varphi `$ is performed around the unit vector $`\stackrel{}{e}`$, the quaternion representation of the rotation has $`a_0=\mathrm{cos}(\varphi /2)`$ and $`\stackrel{}{a}=\mathrm{sin}(\varphi /2)\stackrel{}{e}`$. Therefore $`a_0^2+\stackrel{}{a}^2=1`$ and the identity transformation is represented by $`a_0=1`$.
If a particle traverses an optical element which rotates the spin according to the quaternion $`A`$ and then passes through an element which rotates the spin according to the quaternion $`B`$, the total rotation of the spin is given by
$`C`$ $`=`$ $`c_0\underset{¯}{1}\mathrm{i}\stackrel{}{c}\underset{¯}{\overset{}{\sigma }}=(b_0\underset{¯}{1}\mathrm{i}\stackrel{}{b}\underset{¯}{\overset{}{\sigma }})(a_0\underset{¯}{1}\mathrm{i}\stackrel{}{a}\underset{¯}{\overset{}{\sigma }})`$ (72)
$`=`$ $`(b_0a_0\stackrel{}{b}\stackrel{}{a})\underset{¯}{1}\mathrm{i}(b_0\stackrel{}{a}+\stackrel{}{b}a_0+\stackrel{}{b}\times \stackrel{}{a})\underset{¯}{\overset{}{\sigma }}.`$ (73)
The concatenation of quaternions can be written in matrix form as
$`\stackrel{}{C}`$ $`=`$ $`\left({\displaystyle \genfrac{}{}{0pt}{}{c_0}{\stackrel{}{c}}}\right)=\underset{¯}{B}\left({\displaystyle \genfrac{}{}{0pt}{}{a_0}{\stackrel{}{a}}}\right),`$ (74)
$`\underset{¯}{B}`$ $`=`$ $`\left(\begin{array}{cccc}\hfill b_0& \hfill b_1& \hfill b_2& \hfill b_3\\ \hfill b_1& \hfill b_0& \hfill b_3& \hfill b_2\\ \hfill b_2& \hfill b_3& \hfill b_0& \hfill b_1\\ \hfill b_3& \hfill b_2& \hfill b_1& \hfill b_0\end{array}\right).`$ (79)
This concatenation of two quaternions can be used to find a differential equation for the spin rotation.
While propagating along the design curve by a distance $`dl`$, spins are rotated by an angle $`\mathrm{\Omega }dl=|\mathrm{\Omega }|dl`$ around the vector $`\stackrel{}{\mathrm{\Omega }}`$. After having been propagated to $`l`$ by the quaternion $`A`$, a spin gets propagated from $`l`$ to $`l+dl`$ by the quaternion with $`b_0=1`$ and $`\stackrel{}{b}=\frac{1}{2}\stackrel{}{\mathrm{\Omega }}dl`$. The resulting total rotation is given by $`A+A^{}dl`$ and we obtain the differential equation
$$\left(\genfrac{}{}{0pt}{}{a_0^{}}{\stackrel{}{a}^{}}\right)=\frac{1}{2}\left(\begin{array}{cccc}\hfill 0& \hfill \mathrm{\Omega }_1& \hfill \mathrm{\Omega }_2& \hfill \mathrm{\Omega }_3\\ \hfill \mathrm{\Omega }_1& \hfill 0& \hfill \mathrm{\Omega }_3& \hfill \mathrm{\Omega }_2\\ \hfill \mathrm{\Omega }_2& \hfill \mathrm{\Omega }_3& \hfill 0& \hfill \mathrm{\Omega }_1\\ \hfill \mathrm{\Omega }_3& \hfill \mathrm{\Omega }_2& \hfill \mathrm{\Omega }_1& \hfill 0\end{array}\right)\left(\genfrac{}{}{0pt}{}{a_0}{\stackrel{}{a}}\right).$$
(80)
Writing the matrix as $`\underset{¯}{\mathrm{\Omega }}`$ and the vector as $`\stackrel{}{A}`$, the spin orbit equation of motion has the form
$$\stackrel{}{z}^{}=\stackrel{}{f}(\stackrel{}{z},l),\stackrel{}{A}^{}=\underset{¯}{\mathrm{\Omega }}(\stackrel{}{z},l)\stackrel{}{A}.$$
(81)
The starting conditions are $`\stackrel{}{z}(0)=\stackrel{}{z}_0`$, $`a_0=1`$, and $`\stackrel{}{a}=0`$. The quaternion $`A`$ depends on the initial phase space coordinates $`\stackrel{}{z}_i`$ and can be expanded in a Taylor series with respect to these coordinates. In the following we want to devise an iteration method for $`A_n`$, which is the Taylor expansion to order $`n`$ of $`A`$.
The rotation vector $`\stackrel{}{\mathrm{\Omega }}`$ is split into its value on the design curve and its phase space dependent part as $`\stackrel{}{\mathrm{\Omega }}(\stackrel{}{z},l)=\stackrel{}{\mathrm{\Omega }}^0(l)+\stackrel{}{\mathrm{\Omega }}^1(\stackrel{}{z},l)`$. The spin motion on the design curve is given by $`\stackrel{}{A}_0^{}(l)=\underset{¯}{\mathrm{\Omega }}^0\stackrel{}{A}_0(l)`$. Similarly to equation (26), spin aberrations are defined with respect to the leading order motion. Small phase space coordinates will create a rotation which differs little from $`\stackrel{}{A}_0(l)`$ and we write the phase space dependent rotation as a concatenation of $`\stackrel{}{A}_0`$ and the $`\stackrel{}{z}`$ dependent rotation $`(1+\delta ,\stackrel{}{\delta })`$ which reduces to the identity for $`\stackrel{}{z}=0`$ by requiring that the aberrations $`\delta `$ and $`\stackrel{}{\delta }`$ vanish on the design curve. With equation (74) we obtain
$$\stackrel{}{A}=\underset{¯}{A_0}\left(\genfrac{}{}{0pt}{}{1+\delta }{\stackrel{}{\delta }}\right).$$
(82)
The quaternion $`A`$ is now inserted in the differential equation (81) to obtain
$`\underset{¯}{A}_0^{}\left({\displaystyle \genfrac{}{}{0pt}{}{1+\delta }{\stackrel{}{\delta }}}\right)`$ $`+`$ $`\underset{¯}{A}_0\left({\displaystyle \genfrac{}{}{0pt}{}{\delta ^{}}{\stackrel{}{\delta }^{}}}\right)`$ (83)
$`=`$ $`(\underset{¯}{\mathrm{\Omega }}^0+\underset{¯}{\mathrm{\Omega }}^1)\underset{¯}{A}_0\left({\displaystyle \genfrac{}{}{0pt}{}{1+\delta }{\stackrel{}{\delta }}}\right).`$ (84)
Taking into account the equation on the design curve and the fact that $`\underset{¯}{A}_0^T`$ describes the inverse rotation of $`\underset{¯}{A}_0`$, we obtain
$$\left(\genfrac{}{}{0pt}{}{\delta ^{}}{\stackrel{}{\delta }^{}}\right)=(\underset{¯}{A}_0^T\underset{¯}{\mathrm{\Omega }}^1\underset{¯}{A}_0)\left(\genfrac{}{}{0pt}{}{1+\delta }{\stackrel{}{\delta }}\right)=\underset{¯}{\overset{~}{\mathrm{\Omega }}}(\stackrel{}{z},l)\left(\genfrac{}{}{0pt}{}{1+\delta }{\stackrel{}{\delta }}\right).$$
(85)
Writing the Taylor expansion to order $`n`$ in $`\stackrel{}{z}_i`$ one finally obtains the iteration equation
$$\left(\genfrac{}{}{0pt}{}{\delta _n}{\stackrel{}{\delta }_n}\right)=_n_0^l\underset{¯}{\overset{~}{\mathrm{\Omega }}}(\stackrel{}{z}_n)\left(\genfrac{}{}{0pt}{}{1+\delta _{n1}}{\stackrel{}{\delta }_{n1}}\right)𝑑\stackrel{~}{l},\left(\genfrac{}{}{0pt}{}{\delta _0}{\stackrel{}{\delta }_0}\right)=0.$$
(86)
This iteration method was used for the spin transport in the program SPRINT and was evaluated using MATHEMATICA in .
In the case of successive approximation in terms of the Hamiltonian, the various aberration coefficients were related by the symplectic symmetry. With the Lagrange formalism the various aberration coefficients were related by their being derivatives of a common integral with respect to different initial coordinates. In the case of the successive approximation for spin motion, the various aberration coefficients in $`\stackrel{}{\delta }`$ and $`\delta `$ are related by the relation $`(1+\delta )^2+\stackrel{}{\delta }^2=0`$.
### Acknowledgment
I owe thanks to D. Barber, H. Mais, and M. Vogt for thoroughly reading the manuscript and for the resulting improvements.
|
warning/0006/hep-ph0006158.html
|
ar5iv
|
text
|
# Space-time description of the hadron interaction at high energies.
## Introduction
In this lecture we will try to describe electromagnetic and strong interactions of hadrons in the same framework which follows from general quantum field theory considerations without the introduction of quarks or other exotic objects.
We will assume that there exist point-like constituents in the sense of quantum field theory which are, however, strongly interacting. It is convenient to refer to these particles as partons. We will not be interested in the quantum numbers of these partons, or the symmetry properties of their interactions. We will assume that, contrary to the perturbation theory, the integrals over the transverse momenta of virtual particles converge like in the $`\lambda \phi ^3`$ theory. It turns out that within this picture a common cause exists for two seemingly very different phenomena: the Bjorken scaling in deep inelastic scattering, and the recent theoretical observation that all hadronic cross sections should approach the same limit (provided that the Pomeranchuk pole exists). The lecture is organized as follows. In the first part we discuss the propagation of the hadrons in the space as a process of creation and absorption of the virtual particles (partons) and formulate the notion of the parton wave function of the hadron. The second part describes momentum and coordinate parton distributions in hadrons. In the third part we consider the process of deep inelastic scattering. It is shown that from the point of view of our approach the deep inelastic scattering satisfies the Bjorken scaling, and, in contrast to the quark model, the multiplicity of the produced hadrons is of the order of $`\mathrm{ln}\frac{\nu }{\sqrt{q^2}}`$. The fourth part is devoted to the strong interactions of hadrons and it is shown that in the same framework the total hadron cross sections have to approach asymptotically the same limiting value. In the last part of the lecture we discuss the processes of elastic and quasi-elastic scattering at high energies. It is demonstrated that the cross sections of the quasi-elastic scattering processes at zero angle tend to zero at asymptotically high energies.
Let us discuss, how one can think of the space-time propagation of a physical particle in terms of virtual particles which are involved in the interaction with photon and other hadrons. It is well known that the propagation of a real particle is described by its Green function, which corresponds to a series of Feynman diagrams of the type
(for simplicity, we will consider identical scalar particles). The Feynman diagrams, having many remarkable properties, have, nevertheless, a disadvantage compared to the old-fashioned perturbation theory. Indeed, they do not show how a system evolves with time in a given coordinate reference frame. For example, depending on the relations between the time coordinates $`x_{10}`$, $`y_{10}`$, $`x_{20}`$ and $`y_{20}`$, the graph in Fig.1b corresponds to different processes:
Similarly, the diagram Fig.1c corresponds to the processes
In quantum electrodynamics, where explicit calculations can be carried out, this complicated correspondence is of little interest. However, for strong interactions, where explicit calculations are impossible, distinguishing between different space developments will be useful.
Obviously, if the interaction is strong (the coupling constant $`\lambda `$ is large), many diagrams are relevant. The first question which arises is which configurations dominate: the ones which correspond to the subsequent decays of the particles - the diagrams Fig.2a and Fig3.a, or those which correspond to the interaction of the initial particle with virtual ”pairs” created in the vacuum. It is clear that if the coupling constant is large and the momentum of the incoming particle is small (see below), configurations with ”pairs” dominate (at least if the theory does not contain infinities). Indeed, if $`x_{20}x_{10}`$ is small, then in the case of configurations without ”pairs” the integration regions corresponding to each correction will tend to zero with an increase of the number of corrections. At the same time, for the configurations containing ”pairs” the region of integration over time will remain infinite. Hence, if the retarded Green function $`G^r(y_2y_1)`$ does not have a strong singularity at $`x_{20}x_{10}0`$, the contribution of the configurations without ”pairs” will be relatively small if the coupling constant is large. Even the graphs of the type Fig.1d are determined mainly by configurations like
This means that if we observe a low energy particle at any particular moment of time (the cut in the diagram in Fig. 4), we will see few partons which are decay products of the particle, and a large number of virtual ”pairs” which will interact with these partons in the future.
What happens if a particle has a large momentum in our coordinate reference frame? To analyze the space-time evolution of a fast particle we have to consider a space-time interval $`(x_1x_2)^2`$ such that $`(x_1x_2)^2\frac{1}{\mu ^2}`$, and $`t_2t_1E/\mu ^21/\mu `$. Here $`\mu `$ is the mass of the particle, $`E`$ its energy. In this case, $`\stackrel{}{x}_1\stackrel{}{x}_2=\stackrel{}{v}(t_2t_1)`$, $`(x_2x_1)^2=\frac{\mu ^2}{E^2}(t_2t_1)^2\frac{1}{\mu ^2}`$. For such intervals the relation between the configurations with and without ”pairs” changes. Configurations corresponding to a decay of one parton into many others start to dominate, while the role of configurations with ”pairs” decreases.
The physical origin of this phenomenon is evident. A fast parton can decay, for example, into two fast partons which, due to the energy-time uncertainty relation, will exist for a long time (of the order of $`E/\mu ^2`$), since
$$\mathrm{\Delta }E=\sqrt{\mu ^2+\stackrel{}{p}^2}\sqrt{\mu ^2+\stackrel{}{p}_1^2}\sqrt{\mu ^2+(\stackrel{}{p}\stackrel{}{p}_1)^2}\frac{\mu ^2}{2|\stackrel{}{p}|}\frac{\mu ^2}{2|\stackrel{}{p}_1|}\frac{\mu ^2}{2|\stackrel{}{p}\stackrel{}{p}_1|}.$$
Each of these two partons can again decay into two partons and this will continue up to the point when slow particles, living for a time of the order of $`\frac{1}{\mu }`$, are created. After that the fluctuations must evolve in the reverse direction, i.e. the recombination of the particles begins.
On the other hand, due to the same uncertainty relation, creation of virtual ”pairs” with large momenta in vacuum is possible only for short time intervals of the order of $`\frac{1}{p}`$. Hence, it affects only the region of small momentum partons. The way in which this phenomenon manifests itself can be seen using the simplest graph in Fig.5. as an example. We will observe that it is possible to place here many emissions in spite of the fact that the interval $`x_{12}^2`$ is of the order of unity ($`1/\mu ^2`$), and the Green function depends only on the invariants.
For the sake of simplicity, let us verify this for one space dimension ($`y_i=(t_i,z_i)`$). Suppose that $`x_1=(t,z)`$ and $`x_2=(t,z)`$, $`x_{12}^2=(2t)^2(2z)^2`$. Then $`t=z+\frac{x_{12}^2}{8z}`$. Let us choose the variables $`y_i`$, $`y_i^{}`$ in the same way: $`y_i=(t_i,z_i)`$, $`y_i^{}=(t_i^{},z_i^{})`$, and consider the following region of integration in the integral, corresponding to the diagram in Fig.5:
$$1<z_n<z_{n1}\mathrm{}<z_1<t,$$
$$1<z_n^{}<z_{n1}^{}\mathrm{}<z_1^{}<t,$$
$$z_iz_i^{},t_i=z_i+\frac{y_i^2}{2z_i},t_i^{}=z_i^{}+\frac{y_i^{}_{}{}^{}2}{2z_i^{}}.$$
The integrations over $`d^2y_1\mathrm{}d^2y_nd^2y_1^{}\mathrm{}d^2y_n^{}`$ can be substituted by integrations over $`y_i^2`$, $`y_{i}^{}{}_{}{}^{2}`$ and $`z_iy_{iz},z_i^{}y_{iz}^{}`$
$$d^2y_i=\frac{1}{2}dy_i^2\frac{dz_i}{z_i},d^2y_i^{}=\frac{1}{2}dy_i^{}_{}{}^{}2\frac{dz_i^{}}{z_i^{}}$$
It is easy to see that in this region of integration the arguments of all Green functions: $`(y_iy_i^{})^2`$, $`(y_iy_{i+1}^{})^2`$, $`(y_i^{}y_{i+1}^{})^2`$ , are of the order of unity, and the integrals do not contain any small factors. All these conditions for $`y_i`$ can be satisfied simultaneously for a large number of emissions: $`n\mathrm{ln}t`$. Indeed, if we write $`z_n`$ in the form $`z_n=C^n`$, all conditions will be fulfilled for
$$n\frac{\mathrm{ln}t}{\mathrm{ln}C},C1.$$
Obviously, one can consider a more complicated diagram than Fig.5 by including interactions of the virtual particles. On the other hand, configurations containing vacuum ”pairs” play a minor role. Moving backwards in time is possible only for short time intervals (Fig.6):
Hence, we reach the following picture. A real particle with a large momentum $`p`$ can be described as an ensemble of an indefinite number of partons of the order of $`\mathrm{ln}\frac{p}{\mu }`$ with momenta in the range from $`p`$ to zero, and several vacuum pairs with small momenta which in the future can interact with the target.
The observation of a slow particle during an interval of the order of $`1/\mu `$ does not tell us anything about the structure of the particle since we cannot distinguish it from the background of the vacuum fluctuations, and we can speak only about the interaction of particles or about the spectrum of states. On the contrary, in the case of a fast particle we can speak about its structure, i.e. about the fast partons which do not mix with the vacuum fluctuations. As a result, in a certain sense a fast particle becomes an isolated system which is only weakly coupled to the vacuum fluctuations. Hence, it can be described using a quantum mechanical wave function or an ensemble of wave functions, which determine probabilities of finding certain numbers of partons and their momentum distribution. Such a description is not invariant , since the number of partons depends on the momentum of the particle, but it can be considered as covariant. Moreover, it may be even invariant, if the momentum distribution of the partons is homogeneous in the region of momenta much smaller than the maximal one, and much larger than $`\mu `$.
Indeed, under the transformation from one reference frame to another in which the particle has, for example, a larger momentum, a new region emerges in the distribution of partons; in the old region, however, the parton distribution remains unchanged. One usually describes hadrons in terms of the quantum mechanics of partons in the reference frame which moves with an infinite momentum, because in this case all partons corresponding to vacuum fluctuations have zero momenta, and such a description is exact. Such a reference frame is convenient for the description of the deep inelastic scattering. However, it is not as good for describing strong interactions, where the slow partons are important. In any case, it appears useful to preserve the freedom in choosing the reference frame and to use the covariant description. This allows a more effective analysis of the accuracy of the derivations.
## 1 Wave function of the hadron. Orthogonality and normalization
The previous considerations allow us to introduce the hadron wave function in the following way. Let us assume, as usual, that at $`t\mathrm{}`$ the hadron can be represented as a bare particle (the parton). After a sufficiently long time the parton will decay into other partons and form a stationary state which we call a hadron. Diagrams corresponding to this process are shown in Fig.7.
Let us exclude from the Feynman diagrams those configurations (in the sense of integrations over intermediate times) which correspond to vacuum pair creation.
For the theory $`\lambda \phi ^3`$ such a separation of vacuum fluctuations corresponds to decomposing $`\phi `$ into positive and negative frequency parts $`\phi =\phi ^++\phi ^{}`$ and substituting $`\phi ^3=(\phi ^++\phi ^{})^3`$ by $`3(\phi ^2\phi ^++\phi ^{}\phi ^{+2})`$. The previous discussion shows that the ignored term $`\phi ^{+3}+\phi ^3`$ would mix only partons with small momenta.
It is natural to consider the set of all possible diagrams with a given number of partons $`n`$ at the given moment of time as a component of the hadron wave function $`\mathrm{\Psi }_n(t,\stackrel{}{y}_1,\mathrm{},\stackrel{}{y}_n,p)`$. Similarly, we can determine the wave functions of several hadrons with large momenta provided the energy of their relative motion is small compared to their momenta. The latter condition is necessary to ensure that slow partons are not important in the interaction. The Lagrangian of the interaction remains Hermitian even after the terms corresponding to the vacuum fluctuations are omitted. As a result, the wave functions will be orthogonal, and will be normalized in the usual way:
$$\underset{n}{}\mathrm{\Psi }_n^b^{}(\stackrel{}{y}_1\mathrm{},\stackrel{}{y}_n,p_b)i\stackrel{}{}\mathrm{\Psi }_n^a(\stackrel{}{y}_1\mathrm{},\stackrel{}{y}_n,p_a)\frac{d^3y_1\mathrm{}d^3y_n}{n!}=(2\pi )^3\delta (\stackrel{}{p}_a\stackrel{}{p}_b)\delta _{ab},$$
(1)
or similarly in the momentum space, after separating
$$\underset{n}{}\frac{1}{n!}\mathrm{\Psi }_n^b^{}(\stackrel{}{k}_1\mathrm{},\stackrel{}{k}_n,\stackrel{}{p})\mathrm{\Psi }_n^a(\stackrel{}{k}_1\mathrm{},\stackrel{}{k}_n,\stackrel{}{p})\frac{d^3k_1\mathrm{}d^3k_n}{2k_{10}\mathrm{}2k_{n0}}\frac{\delta (pk_i)}{(2\pi )^{3n1}}=\delta _{ab}.$$
(2)
For the momentum range $`k_i\mu `$, the wave functions coincide with those calculated in the infinite momentum frame. In this reference frame they do not depend on the momentum of the system (except for a trivial factor). This can be easily proven by expanding the parton momenta
$$\stackrel{}{k}_i=\beta _i\stackrel{}{p}+k_i,$$
(3)
and writing the parton energy in the form
$$\epsilon _i=\sqrt{\stackrel{}{k}_i^2+m^2}=\beta _ip+\frac{m^2+k_i^2}{2p\beta _i}.$$
(4)
Note now that the integrals which determine $`\mathrm{\Psi }_n`$, corresponding to Fig.7, can be represented in the form of the old-fashioned perturbation theory where only the differences between the energies of the intermediate states and the initial state $`E_kE`$ enter, and the momentum is conserved. Hence, the terms linear in $`p`$ cancel in these differences, and concequently
$$E_kE=\frac{1}{2p}\left(\underset{i}{}\frac{m^2+k_i^2}{\beta _i}m^2\right)$$
(5)
Each consequent intermediate state in Fig.7 in the $`\lambda \phi ^3`$ model differs from the previous one by the appearance or disappearance of one particle. The factor $`\frac{1}{k_i}=\frac{1}{2p}\frac{1}{\beta _i}`$, which comes from the propagator of this particle, cancels $`2p`$ in (5). Hence, there remain only integrals over $`d^2k_i\frac{d\beta _i}{\beta _i}`$, and the resulting expression does not depend on $`p`$.
$$\underset{n}{}\frac{1}{n!}\mathrm{\Psi }_n^b^{}(k_i,\beta _i)\mathrm{\Psi }_n^a(k_i,\beta _i)\frac{d^2k_i}{2(2\pi )^2}\frac{d\beta _i}{\beta _i}(2\pi )^3\delta (1\beta _i)=\delta _{ab}.$$
(6)
For slow partons, where the expansion (4) is not correct, the dependence on momentum $`p`$ does not disappear, and contrary to the case of the system moving with $`p=\mathrm{}`$, this dependence cuts off the sum over the number of partons.
## 2 Distribution of the partons in space and momentum
The distribution of partons in longitudinal momenta can be characterized by the rapidity:
$$\eta _i=\frac{1}{2}\mathrm{ln}\frac{\epsilon _i+k_{iz}}{\epsilon _ik_{iz}},$$
(7)
where $`k_{iz}`$ is the component of the parton momentum along the hadron momentum.
$$\eta _i\mathrm{ln}\frac{2\beta _ip}{\sqrt{m^2+k_i^2}}.$$
(8)
As it is well known, this quantity is convenient since it simply transforms under the Lorentz transformations along the $`z`$ direction: $`\eta _i^{}=\eta _i+\eta _0`$ , where $`\eta _0`$ is the rapidity of the coordinate system.
The determination of the parton distribution over $`\eta `$ is based on the observation that in each decay process $`k_1k_2+k_3`$ shown in Fig.7 the momenta $`\stackrel{}{k}_2`$ and $`\stackrel{}{k}_3`$ are, in the average, of the same order. This means that in the process of subsequent parton emission and absorption the rapidities of the partons change by a factor of the order of unity. At the same time the overall range of parton rapidities is large, of the order of $`\mathrm{ln}\frac{2p}{m}`$. This implies that in the rapidity space we have short range forces.
Let us consider the density of the distribution in rapidity
$`\phi (\eta ,k_{},p)=`$ (9)
$`{\displaystyle \underset{n}{}}{\displaystyle \frac{1}{n!}}{\displaystyle }|\mathrm{\Psi }(k_{},\eta ,k_1,\eta _1,\mathrm{},k_n,\eta _n,)|^2(2\pi )^3\delta (\stackrel{}{p}\stackrel{}{k}{\displaystyle }\stackrel{}{k}_i){\displaystyle }{\displaystyle \frac{dk_id\eta _i}{2(2\pi )^3}}`$
in the interval $`1\eta \eta _p`$ (see Fig.8).
The independence of $`\phi `$ on $`p`$ for these values of $`\eta `$ means that $`\mathrm{\Psi }`$ depends only on the differences $`\eta _i\eta _p`$. If $`\phi =\phi (\eta \eta _p,k_{})`$ decreases with the increase of $`\eta \eta _p`$, this corresponds to a weak coupling, i.e. to a small probability of the decay of the initial parton. If the coupling constant grows, the number of partons increases and at a certain value of the coupling constant an equilibrium is reached, since the probability of recombination also increases. The value of this critical coupling constant has to be such that the recombination probability due to the interaction should be larger than the recombination probability related to the uncertainty principle.
The basic hypothesis is that such an equilibrium does occur and that due to the short-range character of interaction it is local. This is equivalent to the hypothesis of the constant total cross sections of interaction at $`p\mathrm{}`$. Hence we assume that the equilibrium is determined by the vicinity of the point $`\eta `$ of the order of unity and it does not depend on $`\eta _p`$. Obviously, this can be satisfied only if $`\phi (\eta ,\eta _p,k_{})=\phi (k_{})`$ does not depend on $`\eta `$ and $`\eta _p`$ at $`1\eta \eta _p`$. According to the idea of Feynman, this situation resembles the case of a sufficiently long one-dimensional matter in which, due to the homogeneity of the space, far from the boundaries the density is either constant or oscillating (for a crystal). In our case the analogue of the homogeneity of space is the relativistic invariance (the shift in the space of rapidities). For the time being we will not consider the case of the crystal. According to (9), the integral of $`\phi (\eta ,\eta _p,k_{})`$ over $`\eta `$ and $`k_{}`$ has the meaning of the average parton density which is, obviously, of the order of $`\eta _p\mathrm{ln}\frac{2p}{m}`$.
Generally speaking, we cannot say anything about the parton distribution in the transverse momenta except for one statement: it is absolutely crucial for the whole concept that it must be restricted to the region of the order of parton masses, like in the $`\lambda \phi ^3`$ theory.
Consider now the spatial distribution of the partons. First, let us discuss parton distribution in the plane perpendicular to the momentum $`\stackrel{}{p}`$. For that purpose it is convenient to transform from $`\mathrm{\Psi }_n(k_1,\eta _1,k_2,\eta _2,\mathrm{}k_n,\eta _n)`$ to the impact parameter representation $`\mathrm{\Psi }_n(\stackrel{}{\rho }_1,\eta _1,\stackrel{}{\rho }_2,\eta _2,\mathrm{}\stackrel{}{\rho }_n,\eta )`$:
$$\mathrm{\Psi }_n(\stackrel{}{\rho }_n,\eta _n)=e^{i{\scriptscriptstyle k_i\rho _i}}\mathrm{\Psi }(k_i,\eta _i)\delta (k_i)(2\pi )^2\frac{d^2k_i}{(2\pi )^2}.$$
(10)
Let us rank the partons in the order of their decreasing rapidities. Consider a parton with the rapidity $`\eta \eta _p`$ and let us follow its history from the initial parton. Initially, we will assume that it was produced solely via parton emissions (Fig.9).
In this case it is clear that if the transversal momenta of all partons are of the order of $`\mu `$, than each parton emission leads to a change of the impact parameter $`\stackrel{}{\rho }`$ by $`\frac{1}{\mu }`$. If $`n`$ emissions are necessary to reduce the rapidity from $`\eta _p`$ to $`\eta `$, and they are independent and random, $`\overline{(\mathrm{\Delta }\rho )^2}n`$. If every emission changes the rapidity of the parton by about one unit, then
$$\overline{(\mathrm{\Delta }\rho )^2}=\gamma (\eta _p\eta ).$$
(11)
Hence, the process of the subsequent parton emissions results in a kind of diffusion in the impact parameter plane. The parton distribution in $`\rho `$ for the rapidity $`\eta `$ has the Gaussian form:
$$\phi (\rho ,\eta )=\frac{C(\eta )}{\pi \gamma (\eta _p\eta )}e^{\frac{\rho ^2}{\gamma (\eta _p\eta )}},$$
(12)
if the impact parameter of the initial parton is considered as the origin. Consequently, the partons with $`\eta 0`$ have the broadest distribution, and, hence, the fast hadron is of the size
$$R=\sqrt{\gamma \eta _p}\sqrt{\gamma \mathrm{ln}\frac{2p}{m}}.$$
(13)
The account of the recombination and the scattering of the partons affects only densities of partons and fluctuations, but does not change the radius of the distribution which can be viewed as the front of the diffusion wave.
Let us discuss the parton distribution over the longitudinal coordinate. A relativistic particle with a momentum $`p`$ is commonly considered as a disk of the thickness $`1/p`$. In fact, this is true only in the first approximation of the perturbation theory. Really, a hadron is a disk with radius $`\sqrt{\gamma \mathrm{ln}\frac{2p}{m}}`$ and the thickness of the order of $`1/\mu `$. Indeed, each parton with a longitudinal momentum $`k_{iz}`$ is distributed in the longitudinal direction in an interval $`\mathrm{\Delta }z_i\frac{1}{k_{iz}}`$. Since the parton spectrum exists in the range of momenta from $`p`$ down to $`k_i\mu `$, the longitudinal projection of the hadron wave function has the structure depicted in Fig.11.
Finally, let us consider what is the lifetime of a particular parton. As we have discussed in the Introduction, in a theory which is not singular at short distances, the intervals $`y_{12}^2`$ between two events represented by a Feynman diagram are of the order of unity. For a fast particle moving along the $`z`$ axis, $`z_{21}=vt_{21}`$ and $`y_{12}^2=t_{21}^2\frac{m^2}{p^2}`$. Consequently, the lifetime of a fast parton with a momentum $`k_i`$ is of the order of $`\frac{k_i}{\mu ^2}`$. The presented arguments were based on the $`\lambda \phi ^3`$ theory which is the only theory which provides a cutoff in transverse momenta. Still, the argument should hold for other theories and for particles with spins, if one assumes that in these theories the cutoff of transverse momenta occures in some way. On the other hand, the $`\lambda \phi ^3`$ theory cannot be considered as a self-consistent example. Indeed, due to the absence of a vacuum state, the series of perturbation theory do not make sense (series with positive coefficients are increasing as factorials). Hence, the picture we have presented here does not correspond literally to any particular field theory. At the same time, it corresponds fully to the main ideas of the quantum field theory and to its basic space-time relations.
## 3 Deep inelastic scattering
It is convenient to consider the deep inelastic scattering of electrons in the frame where the time component of the virtual photon momentum is $`q_0=0`$. In this reference frame the momentum of the photon is equal to $`q_z`$ ($`q^2=q_z^2`$), while the momentum of the hadron is $`p_z=\omega q_z/2`$ ($`\omega =2pq/q^2`$). Suppose that $`q_z`$ is large and $`\omega 1`$. According to our previous considerations, a fast hadron can be viewed as an ensemble of partons. In this system a photon looks as a static field with the wavelength $`1/q_z`$.
The main question is, with which partons can the photon interact. We can consider the static field of a photon as a packet with a longitudinal size of the order of $`1/q_z`$. The interaction time between a hadron with the size $`1/\mu `$ and such a packet is of the order of $`1/\mu `$. However, due to the big difference between the parton and photon wave lengths, the interaction with a slow parton is small. Hence, the photon interacts with partons which have momenta of the order of $`q_z`$. Partons with such momenta are distributed in the longitudinal direction in the region $`1/q_z`$. Because of this, the time of the hadron-photon interaction is in fact of the order of $`1/q_z`$, i.e. much shorter than the lifetime of a parton. This means that the photon interacts with a parton as with a free particle, and so not only the momentum but also the energy is conserved. As a result, the energy-momentum conservation laws select the parton with momentum $`q_z/2`$, which can absorb a photon
$$k_{iz}q_z=k_{iz}^{},|k_{iz}q_z|=k_{iz}.$$
This gives
$$k_{iz}=\frac{q_z}{2},k_{iz}^{}=\frac{q_z}{2}.$$
The cross section of such a process is, obviously, equal to the cross section $`\sigma _0`$ of the absorption of a photon by a free particle, multiplied by the probability to find a parton with a longitudinal momentum $`q_z/2`$ inside the hadron, i.e. by the value $`\phi (\eta _{q/2},\eta _p)`$ (9), integrated over $`k_{}`$. (The necessary accuracy of fulfilment of the conservation laws allows any $`k_{}q_z`$ ).
As it was already discussed, $`\phi (\eta ,\eta _p)=\phi (\eta \eta _p)\phi (\omega )`$. Hence, using the known cross section for the interaction of the photon with a charged spinless particle, we obtain for the cross section of the deep inelastic scattering
$$\frac{d^2\sigma }{dq^2d\omega }=\frac{4\pi \alpha ^2}{q^4}\left(1\frac{pq}{pp_e}\right)\phi (\omega ),$$
(14)
where $`p_e`$ is the electron momentum. If the partons have spins, the situation becomes more complicated, since the cross sections of the interactions between photons and partons with different spins are different. The parton distributions in rapidities for different spins may also be different, leading to the form:
$$\frac{d^2\sigma }{dq^2d\omega }=\frac{4\pi \alpha ^2}{q^4}\left\{\left(1\frac{(pq)}{(pp_e)}\right)\phi _0(\omega )+\left[1\frac{pq}{pp_e}+\frac{1}{2}\left(\frac{pq}{pp_e}\right)^2\right]\phi _{\frac{1}{2}}(\omega )\right\}.$$
(15)
Let us discuss now a very important question, namely: what physical processes take place in deep inelastic scatterings. To clarify this, we go back to Fig.7 determining the hadron wave function. We will neglect the parton recombinations in the process of their creation from the initial parton, i.e. we consider fluctuations of the type shown in Fig.9. Suppose that the photon was absorbed by a parton with a large momentum $`q_z/2`$. As a result, this parton obtained a momentum $`q_z`$ and moves in the opposite direction with momentum $`q_z/2`$. The process is depicted in Fig.12. What will now happen to this parton and to the remaining partons? Within the framework we are using it is highly unlikely that the parton with the momentum $`q_z/2`$ will have time to interact with the other partons. The probability to interact directly with residual partons will be small, because the relative momentum of the parton with $`q_z/2`$ and the rest of the partons is large. It could interact with other partons after many subsequent decays which, in the end, could create a slow parton. However, the time needed for these decays is large, and during this time the parton and its decay products will move far away from the remaining partons, thus the interaction will not take place.
Hence, we come to the conclusion that one free parton is moving in the direction $`q_z`$. What will we observe experimentally, if we investigate particles moving in this direction? To answer this question, it is sufficient to note that, in average, a hadron with a momentum $`k_z`$ consists of $`n`$ partons; $`n=C\mathrm{ln}\frac{k_z}{\mu }`$ at $`k_z\mu `$.
In a sense there should exist an uncertainty relation between the number of partons in a hadron ($`n`$) and the number of hadrons in a parton ($`n_p`$)
$$n_pn\stackrel{>}{}c\mathrm{ln}\frac{k_z}{\mu },$$
(16)
where $`k_z`$ is the momentum of the state.
We came to the conclusion that the parton decays into a large number of hadrons i.e. in fact the parton is very short-lived, highly virtual. Hence, we have to discuss whether this conclusion is consistent with the assumption that the photon-parton interaction satisfies the energy conservation. To answer this question, let us calculate the mass of a virtual parton with momentum $`k_z`$, decaying into $`n`$ hadrons with momenta $`k_i`$ and masses $`m_i`$.
$$M^2=(\sqrt{m_i^2+k_i^2})^2k_z^2=\left(k_z+\underset{i}{}\frac{m_i^2+k_i^2}{2k_{iz}}\right)^2k_z^2k_z\underset{i}{}\frac{m_i^2+k_i^2}{k_{iz}}.$$
If the hadrons are distributed almost homogeneously in rapidities, their longitudinal momenta decrease exponentially with their number, and in the sum only a few terms, corresponding to slow hadrons, are relevant. As a result, $`M^2k_z\mu `$, i.e. the time of the existence of the parton is of the order of $`1/\mu `$, much larger than the time of interaction with a photon $`1/q_z`$.
Let us discuss now, what happens to the remaining partons. Little can be determined using only the uncertainty relation eq.(16). This is because the number of partons before the photon absorption was $`n`$, after the photon absorption it became $`n1`$ and, consequently, according to the uncertainty relation, the number of hadrons corresponding to this state can range from 1 to $`n`$. Hence, everything depends on the real perturbation of the hadron wave function due to the photon absorption.
Consider now the fluctuation shown in Fig.12. The photon absorption will not have any influence on partons created after the parton ”$`b`$” which absorbed the photon was produced, and and which have momenta smaller than “b”. These fluctuations will continue, and the partons can, in particular, recombine back into the parton ”$`c`$”. The situation is different for partons which occured earlier and have large momenta (”$`c^{}`$”, ”$`c^{\prime \prime }`$”). In this case the fluctuation cannot evolve further the same way, since the parton ”$`b`$” has moved in the opposite direction. As a result, it is highly probable that partons ”$`c^{}`$” and ”$`c^{\prime \prime }`$” will move apart and lose coherence. On the other hand, slow partons which were emitted by ”$`c^{}`$” and ”$`c^{\prime \prime }`$” earlier and which are not connected with the parton ”$`b`$”, will be correlated, as before, with each of them. Thus ”$`c^{}`$” and ”$`c^{\prime \prime }`$” will move in space together with their slow partons, i.e. in the form of hadrons. Hence, it appears that partons flying in the initial direction lead to the production of the order of $`c\mathrm{ln}\frac{\omega q_z}{2}c\mathrm{ln}\frac{q_z}{2}=c\mathrm{ln}\omega `$ hadrons with rapidities ranging from $`\mathrm{ln}\frac{\omega q_z}{\mu }`$ to $`\mathrm{ln}\frac{q_z}{\mu }`$. This answer can be interpreted in the following way. After the photon is absorbed, a hole is created in the distribution of partons moving in the initial direction.
Contrary to the case of rapidities of partons, we will count the rapidity of the hole not from zero rapidity but from the rapidity $`\mathrm{ln}\frac{\omega q_z}{\mu }`$. In this case the rapidity of the hole is $`\mathrm{ln}\omega `$. If we now represent the parton hole with rapidity $`\mathrm{ln}\omega `$ as a superposition of the hadron states, this superposition will contain $`\mathrm{ln}\omega `$ hadron states.
Let us represent the whole process by a diagram describing rapidity distributions of partons and hadrons. Before the photon absorption the partons in the hadrons are distributed at rapidities between zero and $`\mathrm{ln}\frac{\omega q_z}{\mu }`$, while after the photon absorption a parton distribution is produced which is shown in Fig. 13.
This parton distribution leads to the hadron distribution shown in Fig. 14. The total multiplicity corresponding to this distribution is
$$\overline{n}=c\mathrm{ln}\frac{q_z}{\mu }+c\mathrm{ln}\omega =c\mathrm{ln}\frac{\nu }{\mu \sqrt{q^2}}.$$
This hadron distribution in rapidities in the deep inelastic scattering differs qualitatively from those previously discussed in the literature. It corresponds to $`c\mathrm{ln}\frac{\sqrt{q^2}}{\mu }`$ hadrons moving in the photon momentum direction, and $`\mathrm{ln}\omega `$ hadrons are moving in the nucleon momentum direction, with a gap in rapidity between these distributions. The hadron distribution which was obtained in the framework of perturbation theory for superconverging theories like $`\lambda \phi ^3`$ (Drell, Yan) differs qualitatively from the distribution in Fig. 14.
In conclusion of this part, it is necessary to point out that the problem of spin properties of the partons exist in this picture even if the partons do not have quark quantum numbers. If, as the experiment shows, the cross section $`\sigma _T`$ for the interaction of the transversal photons is larger than the cross section for the interaction of the longitudinal photons, $`\sigma _L`$, the charged partons have predominantly spin $`1/2`$. This means that at least one fermion, for example a nucleon, has to move in the direction of the photon momentum. In other words, in deep inelastic scattering the distribution of the created hadrons in quantum numbers as the function of their rapidities differs essentially from what we are used to in the strong interactions. Perhaps this is one of the key prediction of the non-quark parton picture for $`\sigma _t\sigma _l`$.
## 4 Strong interactions of hadrons
Let us discuss now the strong interactions of hadrons. First, we consider a collision of two hadrons in the laboratory frame. Suppose that a hadron ”1” with momentum $`\stackrel{}{p}_1`$ hits hadron ”2” which is at rest. Obviously, the parton wave function makes no sense for the hadron at rest, since for the latter the vacuum fluctuations are absolutely essential. However, the hadron at rest can also be understood as an ensemble of slow partons distributed in a volume of the order of $`1/\mu `$, independent of the origin of the partons. Indeed, it does not matter whether these partons are decay products of the initial parton or the result of the vacuum fluctuations. How can a fast hadron, consisting of partons with rapidities from $`\mathrm{ln}\frac{2p_1}{\mu }`$ to zero, interact with the target which consists of slow partons? Obviously, the cross section of the interaction of two point-like particles with a large relative energy is not larger than $`\pi \lambda ^2\frac{1}{s_{12}}e^{\eta _{12}}`$ (where $`\lambda `$ is the wave length in the c.m. frame, $`\eta _{12}`$ is the relative rapidity). That is why only slow partons of the incident hadron can interact with the target with a cross section which is not too small. This process is shown in Fig. 15.
If the slow parton which initiated the interaction was absorbed in this interaction, the fluctuation which lead to its creation from a fast parton was interrupted. Hence, all partons which were emitted by the fast parton in the process of fluctuation cannot recombine any more. They disperse in space and ultimately decay into hadrons leading to the creation of hadrons with rapidities from zero to $`\mathrm{ln}\frac{2p_1}{\mu }`$. The interaction between the partons is short-range in rapidities. Hence, the hadron distribution in rapidities will reproduce the parton distribution in rapidities. In particular, the inclusive spectrum of hadrons will have the form shown in Fig. 8, with an unknown distribution near the boundaries. The total hadron multiplicity will be of the order of $`\eta _p=\mathrm{ln}\frac{2p_1}{\mu }`$. If the probability of finding a slow parton in the hadron does not depend on the hadron momentum (this would be quite natural, since with the increase of the momentum the life-time of the fluctuation is also growing), the total cross section of the interaction will not depend on the energy at high energies.
Before continuing the analysis of inelastic processes, let us discuss, how to reconcile the energy independence of the total interaction cross section at high energies with the observation discussed above that the transverse hadron sizes increase with the increase of the energy as $`\sqrt{\gamma \mathrm{ln}\frac{2p}{\mu }}`$. The answer is that slow partons are distributed almost homogeneously over the disk of the radius $`\sqrt{\gamma \mathrm{ln}\frac{2p}{\mu }}`$ (Eq.(11)) , while their overall multiplicity during the time of $`1/\mu `$ is of order of unity.
Let us see now how the same process will look, for example, in the c.m. frame. In this reference frame the interaction will have the form as shown in Fig. 16.
Each of the hadrons consists of partons with rapidities ranging from $`\mathrm{ln}\frac{2p_c}{\mu }`$ to zero and from zero to $`\mathrm{ln}\frac{2p_c}{\mu }`$, respectively. The slow partons interact with cross sections which are not small. As a result, the fluctuations will be interrupted in both hadrons, and the partons will fly away in the opposite directions, leading to the creation of hadrons with rapidities from $`\mathrm{ln}\frac{2p_c}{\mu }`$ to $`\mathrm{ln}\frac{2p_c}{\mu }`$. From the point of view of this reference frame the inclusive spectrum must have the form shown in Fig. 17, with unknown distributions not only at the boundaries but also in the centre, since the distribution of the slow partons in the hadrons and in vacuum fluctuations is unknown. The hadron inclusive spectrum, however, should not depend on the reference frame. Thus the inclusive spectrum in Fig. 17 should coincide with the inclusive spectrum in Fig. 8, and they should differ only by a trivial shift along the rapidity axis, i.e. due to relativistic invariance we know something about the spectra of slow partons and vacuum fluctuations.
Let us demonstrate that this comparison of processes in two reference frames leads to a very important statement, namely that at ultra-high energies the total cross sections for the interactions of arbitrary hadrons should be equal. Indeed, we have assumed that the distribution of hadrons reproduces the parton distribution.
From the point of view of the laboratory frame the distribution of partons and, consequently, distribution of hadrons in the central region of the spectrum is completely determined by the properties (quantum numbers, mass, etc.) of particle 1, and does not depend on the properties of particle 2. On the other hand, from the point of view of the antilaboratory frame (where the particle 1 is at rest) everything is determined by the properties of particle 2. This is possible only if the distribution of partons in the hadrons with rapidities $`\eta `$ much smaller than the hadron rapidity $`\eta _p`$ does not depend on the quantum numbers and the mass of the hadron, that is the parton distribution with $`\eta \eta _p`$ should be universal. From the point of view of the c.m. system the same region is determined by slow partons of both hadrons and by vacuum fluctuations (which are universal), and, consequently, the distribution of slow partons is also universal.
It is natural to assume that the probability of finding a hadron in a sterile state without slow partons tends to zero with the increase of its momentum, in other words assume that slow partons are always present in a hadron (compare to the decrease of the cross section of the elastic electron scattering at large $`q^2`$). In this case considering the process in the c.m. system, we see that the total cross section of the hadron interaction is determined by the cross section of the interaction of slow partons and by their transverse distribution which is universal. Consequently, the total hadron interaction cross section is also universal, i.e. equal for any hadrons.
This statement looks rather strange if we regard it, for instance, from the following point of view. Let us consider the scattering of a complicated system with a large radius, for example, deuteron-nucleon scattering. As we know, the cross section of the deuteron-nucleon interaction equals the sum of the nucleon-nucleon cross sections, thus it is twice as large as the nucleon-nucleon cross section. How and at what energies can the deuteron-nucleon cross section become equal to the nucleon-nucleon cross section? How is it possible that the density of slow partons in the deuteron turns out to be equal to the density of slow partons in the nucleon? To answer this question, let us discuss the parton structure of two hadrons which are separated in the plane transverse to their longitudinal momenta by a distance much larger than their Compton wave length $`1/\mu `$. Suppose that at the initial moment they were point-like particles. Next, independently of each other, they begin to emit partons with decreasing longitudinal momenta. At the same time the diffusion takes place in the transverse plane so that the partons will be distributed in a growing region. The basic observation which we shall prove and which answers our question is that if the momenta of the initial partons are sufficiently large, then during one fluctuation the partons coming from different initial partons will inevitably meet in space (Fig. 18) in the region of the order of $`1/\mu `$. They will have similar large rapidities and, hence, will be able to interact with a probability of the order of unity. If such “meetings” take place sufficiently frequently, the probability of the parton interaction will be unity. Consequently, the further evolution and the density of the slow partons which are created after the meeting may not depend on the fact that initially the transverse distance between two partons was large.
In terms of the diffusion in the impact parameter plane this statement corresponds to the following picture. Suppose that initial partons were placed at points $`\rho _1`$ and $`\rho _2`$ in Fig. 19 and that their longitudinal momenta are of the same order of magnitude, i.e. difference of their rapidities is of the order of unity, while each of the rapidities is large. We will follow the parton starting from point $`\rho _1`$, which decelerates via emission of other partons. As we have seen, its propagation in the perpendicular plane corresponds to diffusion. The difference of rapidities $`\eta _p\eta `$ at the initial and the considered moments serves the role of time in this diffusion process.
The diffusion character of the process means that the probability density of finding a parton with rapidity $`\eta `$ at the point $`\rho `$ if it started from the point $`\rho _1`$ with rapidity $`\eta _p`$ is
$$\omega (\stackrel{}{\rho },\stackrel{}{\rho }_1,\eta _p\eta )=\frac{1}{\pi \gamma (\eta _p\eta )}e^{\frac{(\stackrel{}{\rho }\stackrel{}{\rho }_1))^2}{\gamma (\eta _p\eta )}}.$$
(17)
The situation is exactly the same for a decelerating parton which started from the point $`\rho _2`$. Thus, the probability of finding both partons at the same point $`\rho `$ with equal rapidities is proportional to
$`\omega (\rho _{12},\eta _p\eta )=`$ (18)
$`{\displaystyle \omega (\stackrel{}{\rho },\stackrel{}{\rho }_1,\eta _p\eta )\omega (\stackrel{}{\rho },\stackrel{}{\rho }_2,\eta _p\eta )d^2\rho }=`$
$`{\displaystyle \frac{1}{2\pi \gamma (\eta _p\eta )}}\mathrm{exp}\left[{\displaystyle \frac{(\stackrel{}{\rho }_1\stackrel{}{\rho }_2))^2}{2\gamma (\eta _p\eta )}}\right]`$
If we now integrate this expression over $`\eta `$, i.e. estimate the probability for the partons to meet at some rapidities, we obtain
$$_0^{\eta _p}\omega (\rho _{12},\eta _p\eta )𝑑\eta \frac{1}{\pi }\mathrm{log}\frac{2\gamma \eta _p}{\rho _{12}^2}|_{\eta _p\mathrm{}}\mathrm{}.$$
(19)
This means that if $`2\gamma \eta _p\rho _{12}^2`$, the partons will inevitably meet. According to (19) we get a probability much larger than unity. The reason is that under these conditions the meeting of partons at different values of $`\eta `$ are not independent events and therefore it does not make sense to add the probabilities. It is easy to prove this statement directly, for example with the help of the diffusion equation. We will not do this, however. According to a nice analogy suggested by A. Larkin, this theorem is equivalent to the statement that if you are in an infinite forest in which there is a house on a finite distance from you, then, randomly wandering in the forest, you sooner or later arrive at this house. Essentially, the reason is that in the two-dimensional space the region inside of which the diffusion takes place and the length of the path travelled during the diffusion increase with time in the same way. ¿From the point of view of the reference frame in which the deuteron is at rest and is hit by a nucleon in the form of a disk, the radius of which is much larger than that of the deuteron, the statement of the equality of cross sections means that the parton states inside the disk are highly coherent.
It is clear from above that the cross sections of two hadrons can become equal only when the radius of parton distribution $`\sqrt{\gamma \eta _p}`$ which is increasing with the energy becomes much larger than the size of both hadrons. Substituting $`4\frac{0,3}{m^2}`$ for the value of $`\gamma `$ ($`m`$ is the proton mass) <sup>1</sup><sup>1</sup>1It will be demonstrated below that $`\gamma =4\alpha ^{}`$, where $`\alpha ^{}`$ is the slope of the Pomeron trajectory. The current data give $`\alpha ^{}0.25GeV^2`$. we see that the deuteron-nucleon cross section will practically never coincide with the nucleon-nucleon cross section, while the tendency for convergence of cross sections for pion-nucleon, kaon-nucleon and nucleon-nucleon scatterings may be manifested already starting at the incident energies $`10^3`$ GeV.
## 5 Elastic and quasi-elastic processes
So far we focused on the implications of the considered picture for inelastic processes with multiplicities, growing logarithmically with the energy. However, with a certain probability it can happen that slow partons scatter at very small angles and the fluctuations will not be interrupted in either of the hadrons (for example, if we discuss the process in the c.m. frame). In this case small angle elastic or quasi-elastic scattering will take place (Fig. 20).
First, let us calculate the elastic scattering amplitude. It is well known that the imaginary part of the elastic scattering amplitude can be written in the form
$$A_1(s_{12})=s_{12}d^2\rho _{12}e^{i\stackrel{}{q}\stackrel{}{\rho }_12}\sigma (\rho _{12},s_{12}),$$
(20)
where $`s_{12}`$ is the energy squared in the c.m. system, $`\rho _{12}`$ is the relative impact parameter, $`\sigma (\rho _{12},s_{12})d^2\rho _{12}`$ \- the total interaction cross section of particles being at the distance $`\rho _{12}`$ and $`\stackrel{}{q}`$ is the momentum transferred. In order to calculate $`\sigma (\rho _{12},s_{12})`$ it is sufficient to notice that, according to (12), the probability of finding a slow parton with rapidity $`\eta `$ at the impact parameter $`\rho _1^{}`$ which originated from the first hadron with an impact parameter $`\stackrel{}{\rho }_1`$ is
$$\phi _1(\stackrel{}{\rho }_1,\stackrel{}{\rho ^{}}_1,\eta _1,\eta _{pc})\frac{C(\eta _1)}{\pi \gamma \eta _{pc}}\mathrm{exp}\left[\frac{(\stackrel{}{\rho }_1\stackrel{}{\rho ^{}}_1)^2}{\gamma \eta _{pc}}\right].$$
(21)
The probaility a parton originating from the second hadron at impact parameter $`\rho _2^{}`$ is
$$\phi _2(\stackrel{}{\rho }_2,\stackrel{}{\rho }_2^{},\eta _2,\eta _{pc})\frac{C(\eta _2)}{\pi \gamma \eta _{pc}}\mathrm{exp}\left[\frac{(\stackrel{}{\rho }_2\stackrel{}{\rho }_2^{})^2}{\gamma \eta _{pc}}\right].$$
(22)
The total cross section of the hadron interaction which is due to the interaction of slow partons is equal to
$$\sigma (\rho _{12},s_{12})=𝑑\eta _1𝑑\eta _2d^2\rho _{12}^{}C(\eta _1)C(\eta _2)$$
$$\times \frac{d^2\rho }{(\pi \gamma \eta _{pc})^2}\mathrm{exp}[\frac{(\stackrel{}{\rho }\stackrel{}{\rho }_1)^2}{\gamma \eta _{pc}}\frac{(\stackrel{}{\rho }\stackrel{}{\rho }_2)^2}{\gamma \eta _{pc}}].$$
We have taken into account that $`\rho _1^{}=\rho +\frac{\rho _{12}^{}}{2}`$, $`\rho _2^{}=\rho \frac{\rho _{12}^{}}{2}`$, and that the dependence on $`\rho _{12}^{}`$ can be neglected in the exponential factor.
After carrying out the integration over $`\rho `$, we obtain
$$\sigma (\rho _{12},s_{12})=\frac{e^{\frac{(\stackrel{}{\rho }_1\stackrel{}{\rho }_2)^2}{2\gamma \eta _{pc}}}\sigma _0}{2\pi \gamma \eta _{pc}}$$
(23)
Inserting (22) into (20), we get
$$A_1=s_{12}\sigma _0e^{\frac{\gamma }{4}q^2\xi },$$
$$\xi =2\eta _{pc}=\mathrm{log}\frac{s_{12}}{\mu ^2}.$$
(24)
We obtained the scattering amplitude corresponding to the exchange by the Pomeranchuk pole with the slope $`\alpha ^{}=\gamma /4`$, $`\sigma _0=g^2`$ where $`g`$ is the universal coupling constant of the Pomeron and hadron. The amplitude (24) is usually represented by diagram in Fig. 21
where a propagator of the form $`e^{\alpha ^{}q^2\xi }`$ corresponds to the Pomeron. In the impact parameter space this propagator has the form (22).
Let us discuss the physical meaning of $`\sigma _0`$ in more detail. For this purpose, let us calculate the zero angle scattering amplitude at ($`\stackrel{}{q}=0`$), without using the impact parameter representation. The probability of finding a parton with rapidity $`\eta `$ and a transverse momentum $`k_{}`$ is described by (9). This expression at $`\eta \eta _p`$ corresponds to the diagram in Fig.22. The wavy line represents integration over parton rapidities from $`\eta _p`$ to zero.
This figure reflects the hypothesis that the calculation of $`\phi (\eta ,k_t,\eta _p)`$ for suffiently large $`\eta _p`$ and $`\eta \eta _p`$ leads to an expression for $`\phi `$ which is factorized in the same way as the Pomeron contribution to the scattering amplitude. This is because the parton distribution in this region is independent of the properties of the hadron as well as the values of $`\eta ,\eta _p`$. Compared to the diagram in Fig. 7, Fig. 22 indicates that the calculation of $`\phi (\eta ,k_t,\eta _p)`$ is similar to the calculation of the inclusive cross section due to the Pomeron exchange. The only difference is that the coupling of the hadron with the Pomeron should be substituted by unity, since a hadron always exists in a Pomeron state. If $`\eta 1`$, $`\phi (\eta ,k_t)`$ corresponds to the diagram in Fig.22a, which shows that $`\phi (\eta ,k_t)`$ depends on $`\eta `$. Similarly, it is possible to determine the probability of finding several slow partons (Fig. 22b), and even the density matrix of slow partons. In this case the amplitude of elastic hadron-hadron scattering in the center of mass frame is determined by the diagram of fig.23 and the value of $`\sigma _0`$ is determined solely by the interaction of slow partons.
Now let us consider the quasi-elastic scattering, corresponding to the Pomeron exchange (Figs.24,25) at zero transverse momentum. While the probability to find the parton in hadron ”a” is determined in eq.(8) by the integral of the wave function squared, the analogous quantity for the amplitude of the inelastic diffractive process (Fig.25) will lead to the integral of the product of the parton functions of different hadrons. They are orthogonal to each other and it is almost obvious that amplitude for inelastic diffractive process at zero angle should vanish for this reason. Indeed the orthogonality condition of eq.(6) has the same structure as the imaginary part of the amplitude. Thus, if at high energies the amplitude factorizes (as it should be for the Pomeron exchange), than the orthonormality condition should also have factorized form in the sense that the integral over parton rapidities with $`\eta \eta _p`$ factors out, and only constants $`g_{ab}`$ depend on the properties of specific hadrons (see Fig.26).
Orthogonality of the wave functions of different hadrons implies that $`g_{ab}=0`$ at $`ab`$. In fact the reason, why the amplitude of inelastic diffractive process vanishes at zero angle is the same as the reason why all cross sections should approach the same value at high energies. Both phenomena are due to the fact that properties of slow partons do not depend on the properties of hadrons to which they belong. We can illustrate this again using the example of quasi-elastic dissociation of the composite system — e.g. deuteron. Let us consider the interaction of a fast nucleon with a deuteron. As we discussed in the previous section, at very large energies partons from different nucleons will always interact with each other independent of the distance between nucleons. This will lead to production of the spectrum of slow partons which does not depend on the relative distance between nucleons in deuteron. This means that the amplitude of the nucleon-nucleon interaction will not depend on the internucleon distance as well. Thus, if nucleons inside the deuteron will remain intact after the interaction, than the deuteron will not dissociate as well, since if the amplitude does not depend on the internucleon distance, the wave function of the deuteron will not change after the interaction.
|
warning/0006/hep-th0006229.html
|
ar5iv
|
text
|
# Nontopological Finite Temperature Induced Fermion Number
## Abstract
We show that while the zero temperature induced fermion number in a chiral sigma model background depends only on the asymptotic values of the chiral field, at finite temperature the induced fermion number depends also on the detailed shape of the chiral background. We resum the leading low temperature terms to all orders in the derivative expansion, producing a simple result that can be interpreted physically as the different effect of the chiral background on virtual pairs of the Dirac sea and on the real particles of the thermal plasma. By contrast, for a kink background, not of sigma model form, the finite T induced fermion number is temperature dependent but topological.
The phenomenon of induced fermion number due to the interaction of fermions with topological backgrounds (e.g., solitons, vortices, monopoles, skyrmions) has many applications ranging from polymer physics to particle physics jr ; ssh ; gw ; wilczek ; gj ; eric ; diakonov ; niemi . The original fractional fermion number result of Jackiw-Rebbi jr has a deep connection with the existence of spinless charged excitations in polymers ssh . The adiabatic analysis of Goldstone-Wilczek gw in systems without conjugation symmetry has important implications for bag models gj , monopoles, and sigma models, which provide effective field theory descriptions of systems ranging from condensed matter, to AMO, to particle and nuclear physics weinberg . The induced fermion number is related to the spectral asymmetry of the relevant Dirac operator, and mathematical results concerning index theorems niemi relate the fermion number to asymptotic topological properties of the background. At finite temperature, the situation is less clear. In several examples ns ; babu ; cp ; goldhaber ; monopole , the fermion number is known to be temperature dependent, but is still topological in the sense that the only dependence on the background field is through its asymptotic properties. In this Letter, we present a simple physical case for which this is not true : in a $`1+1`$ dim chiral sigma model, the finite temperature induced fermion number depends on the detailed structure of the background. This contradicts a previous analysis midorikawa and claim nnp that the finite T fermion number is in general a topological quantity. We give a simple physical explanation of the origin of the nontopological dependence. Our analysis has been motivated in part by the results of pisarski concerning the T dependence of anomalous amplitudes in nuclear decays.
Consider an abelian model in $`1+1`$ dimensions with fermions interacting via scalar and pseudoscalar couplings to two bosonic fields $`\varphi _1`$ and $`\varphi _2`$. For the purposes of this paper $`\varphi _1`$ and $`\varphi _2`$ can be considered as classical external fields. The Lagrangian is
$$=i\overline{\psi }/\psi \overline{\psi }\left(\varphi _1+i\gamma _5\varphi _2\right)\psi $$
(1)
There are two especially interesting physical cases:
(i) kink case jr :
$$\varphi _1=m\mathrm{and}\varphi _2(\pm \mathrm{})=\pm \widehat{\varphi _2}$$
(2)
(ii) sigma model case gw :
$$\varphi _1^2+\varphi _2^2=m^2$$
(3)
In the sigma model case (3), the interaction term in the Lagrangian (1) can be expressed as
$$m\overline{\psi }e^{i\gamma _5\theta }\psi =m\overline{\psi }\left(\mathrm{cos}\theta +i\gamma _5\mathrm{sin}\theta \right)\psi $$
(4)
At $`T=0`$, both these cases have an induced topological current $`J^\mu <\overline{\psi }\gamma ^\mu \psi >`$ given by gw
$$J^\mu =\frac{1}{2\pi }ϵ^{\mu \nu }_\nu \theta +\mathrm{}$$
(5)
where the angular field $`\theta `$ is defined by $`\theta \mathrm{arctan}(\varphi _2/\varphi _1)`$. The dots in (5) refer to higher derivative terms, which are all of the form of a total derivative of $`\theta `$ and its derivatives wilczek . Thus, in particular, the induced fermion number, $`N𝑑xJ^0`$, is
$$N=\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑x\theta ^{}=\frac{1}{\pi }\widehat{\theta }$$
(6)
where $`\pm \widehat{\theta }`$ are the asymptotic values of $`\theta (x)`$ at $`x=\pm \mathrm{}`$. The fermion number $`N`$ is topological as it depends only on $`\widehat{\theta }`$, not on the detailed shape of $`\theta (x)`$. The conjugation symmetric case of Jackiw and Rebbi jr is obtained by taking $`m0`$ in the kink case (2), in which case $`N\pm \frac{1}{2}`$.
At nonzero temperature, the induced fermion number for a static background is niemi ; monopole
$$N=\frac{1}{2}_𝒞\frac{dz}{2\pi i}\mathrm{tr}\left(\frac{1}{Hz}\right)\mathrm{tanh}\left(\frac{\beta z}{2}\right)$$
(7)
where $`\beta =1/T`$ is the inverse temperature, and $`\mathrm{tr}(\frac{1}{Hz})`$ is the resolvent of the Dirac Hamiltonian $`H`$. The contour $`𝒞`$ is $`(\mathrm{}+iϵ,+\mathrm{}+iϵ)`$ and $`(+\mathrm{}iϵ,\mathrm{}iϵ)`$. By considering static backgrounds we avoid the well-known complications of finite temperature calculations in non-static backgrounds das . The technical part of the calculation of the induced fermion number (7) is the computation of the resolvent of $`H`$. Once this is done, the induced fermion number may be expressed as an integral representation, or as a sum by deforming the contour in (7) around the simple poles of the tanh function. For static backgrounds $`\varphi _1(x)`$ and $`\varphi _2(x)`$ in (1), the Dirac Hamiltonian is
$$H=i\gamma ^0\gamma ^1+\gamma ^0\varphi _1(x)+i\gamma ^0\gamma _5\varphi _2(x)$$
(8)
where $`\frac{d}{dx}`$, and we will work with the Dirac matrices $`\gamma ^0=\sigma _3`$, $`\gamma ^1=i\sigma _2`$, and $`\gamma ^5=\sigma _1`$. Also, note that only the even part (in terms of the argument $`z`$) of the resolvent $`\mathrm{tr}(\frac{1}{Hz})`$ contributes to the induced fermion number $`N`$ in (7). (This is most easily seen by deforming the contour around the poles of the tanh function.)
Consider first the kink case in (2). Then the even part of the resolvent can be computed exactly using a trace identity which is a special case of the Callias index theorem callias ; niemi ; ns (alternatively, it can be derived in a more elementary manner as an exact resummation of a SUSY derivative expansion dunne ) :
$`\left[\mathrm{tr}\left({\displaystyle \frac{1}{Hz}}\right)\right]_{\mathrm{even}}`$ $`=`$ $`\mathrm{tr}\left({\displaystyle \frac{m}{(+\varphi _2)(\varphi _2)+m^2z^2}}\right)\mathrm{tr}\left({\displaystyle \frac{m}{(\varphi _2)(+\varphi _2)+m^2z^2}}\right)`$ (9)
$`=`$ $`{\displaystyle \frac{m\widehat{\varphi }_2}{(m^2z^2)\sqrt{m^2+\widehat{\varphi }_2^2z^2}}}`$
Then the induced fermion number (7) for the kink case (2) is
$$N=\frac{2}{\pi }\left(\frac{m\beta }{\pi }\right)^2\mathrm{sin}\widehat{\theta }\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{((2n+1)^2+(\frac{m\beta }{\pi })^2)\sqrt{(2n+1)^2\mathrm{cos}^2\widehat{\theta }+(\frac{m\beta }{\pi })^2}}$$
(10)
where $`\widehat{\theta }\mathrm{arctan}(\widehat{\varphi _2}/m)`$. This result is consistent with previous analyses ns , although these were much less explicit. The induced fermion number (10) is plotted in Fig. 1 as a function of $`\widehat{\theta }`$ for various values of temperature. As $`T0`$, this result reduces smoothly to the zero temperature result (6). Despite its complicated form, the nonzero temperature result (10) is still topological as it only refers to the background through $`\widehat{\theta }`$.
In the sigma model case (3), the trace identity formulae (9) do not apply. Another approach is needed to evaluate the resolvent. One such approach is the derivative expansion ian , in which we assume that the spatial derivatives of the background fields are small compared to the fermion mass scale $`m`$. In other words, the backgrounds $`\varphi _1(x)`$ and $`\varphi _2(x)`$ are assumed to be slowly varying on the scale of the fermion Compton wavelength. Returning to the general Hamiltonian (8), the derivative expansion can be obtained by separating $`H^2`$ as
$$H^2=\left(\begin{array}{cc}^2+\varphi _1^2+\varphi _2^2& 0\\ 0& ^2+\varphi _1^2+\varphi _2^2\end{array}\right)+\left(\begin{array}{cc}\varphi _2^{}& i\varphi _1^{}\\ i\varphi _1^{}& \varphi _2^{}\end{array}\right)$$
(11)
and then expanding $`\mathrm{tr}\left(\frac{1}{Hz}\right)=\mathrm{tr}\left((H+z)\frac{1}{H^2z^2}\right)`$ in powers of derivatives. A simple calculation to first order yields:
$$\left[\mathrm{tr}\left(\frac{1}{Hz}\right)\right]_{\mathrm{even}}=\frac{1}{2}_{\mathrm{}}^{\mathrm{}}𝑑x\frac{(\varphi _1\varphi _2^{}\varphi _2\varphi _1^{})}{(\varphi _1^2+\varphi _2^2z^2)^{3/2}}+\mathrm{}$$
(12)
where the dots refer to terms with three or more derivatives.
In the kink case (2), where $`\varphi _1=m`$ is constant, this first order calculation actually reproduces the exact trace identity result (9). But in the sigma model case (3), where $`\varphi _1^2+\varphi _2^2=m^2`$ is a constant, the first order derivative expansion result (12) implies that:
$$\left[\mathrm{tr}\left(\frac{1}{Hz}\right)\right]_{\mathrm{even}}=\frac{m^2}{2(m^2z^2)^{3/2}}_{\mathrm{}}^{\mathrm{}}𝑑x\theta ^{}+\mathrm{}$$
(13)
So, to first order in the derivative expansion, the induced fermion number for the sigma model case is
$$N^{(1)}=\frac{1}{\pi }\left(\frac{m\beta }{\pi }\right)^2\left(\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{[(2n+1)^2+(\frac{m\beta }{\pi })^2]^{3/2}}\right)_{\mathrm{}}^{\mathrm{}}𝑑x\theta ^{}$$
(14)
which is simply the zero temperature answer (6) multiplied by a smooth function of $`T`$. As $`T0`$, this prefactor reduces to $`\frac{1}{2\pi }`$, so the full zero temperature result (6) is regained. But at finite temperature, the first order (in the derivative expansion) formula (14) for the sigma model case differs from the kink case formula (10), even though each of (14) and (10) reduces to (6) at $`T=0`$.
This raises the question of the higher order corrections to the derivative expansion (12). In the kink case (2), there are no higher order corrections to the even part of the resolvent in (11). This is due to the special form of the Hamiltonian in the kink background, which leads to the first order formula (12) agreeing with the exact trace identity result (9). There can, of course, be higher order corrections to the induced fermion number density, but these are all total (spatial) derivatives, and do not contribute to the integrated induced fermion number, even at nonzero temperature.
But in the sigma model case (3), where the trace identity does not apply, the situation is very different. Going to the next order in the derivative expansion, we find
$`\left[\mathrm{tr}\left({\displaystyle \frac{1}{Hz}}\right)\right]_{\mathrm{even}}={\displaystyle \frac{m^2}{2(m^2z^2)^{3/2}}}{\displaystyle 𝑑x\theta ^{}}`$ $``$ $`{\displaystyle \frac{m^2}{8(m^2z^2)^{5/2}}}{\displaystyle 𝑑x\theta ^{\prime \prime \prime }}`$ (15)
$``$ $`{\displaystyle \frac{m^2(4z^2+m^2)}{16(m^2z^2)^{7/2}}}{\displaystyle 𝑑x(\theta ^{})^3}+\mathrm{}`$
where the dots refer to terms involving five or more derivatives. For a chiral background with $`\theta (x)`$ approaching its asymptotic values exponentially fast, the term $`𝑑x\theta ^{\prime \prime \prime }`$ vanishes. But $`𝑑x(\theta ^{})^3`$ does not vanish. Thus, the first order induced fermion number (14) acquires a third order correction:
$$N^{(3)}=\frac{\beta ^2}{8\pi ^3}\left(\frac{m\beta }{\pi }\right)^2\left(\underset{n=0}{\overset{\mathrm{}}{}}\frac{[4(2n+1)^2+(\frac{m\beta }{\pi })^2]}{[(2n+1)^2+(\frac{m\beta }{\pi })^2]^{7/2}}\right)𝑑x(\theta ^{})^3$$
(16)
This is not just a function of the asymptotic value $`\widehat{\theta }`$ of the chiral field $`\theta (x)`$; it also depends on the actual shape of $`\theta (x)`$. Thus, the induced fermion number is no longer topological. This contradicts midorikawa , where it is stated that the first order derivative expansion contribution (14) is the full answer. However, the energy trace prefactor in (16) vanishes at $`T=0`$, so the nontopological third order contribution (16) vanishes at $`T=0`$. Thus, the nontopological nature of the finite temperature induced fermion number is still consistent (at this order) with the topological nature of the zero temperature induced fermion number (6).
We now turn to a physical explanation of why, in the sigma model case, the finite temperature induced charge is more sensitive to the background field than at zero temperature. Note first of all that the chiral background acts like a static but spatially inhomogeneous electric field, as can be seen by making a local chiral rotation wilczek : $`\psi \stackrel{~}{\psi }=e^{i\theta \gamma _5/2}\psi `$. In terms of these chirally rotated fields the Lagrangian (1), with interaction (4), becomes
$$=i\overline{\stackrel{~}{\psi }}/\stackrel{~}{\psi }m\overline{\stackrel{~}{\psi }}\stackrel{~}{\psi }\overline{\stackrel{~}{\psi }}\gamma ^0\frac{\theta ^{}}{2}\stackrel{~}{\psi }$$
(17)
(The chiral rotation leads to an anomalous Jacobian in the path integral, but this does not affect the induced fermion number.) Thus, the chiral field acts as an inhomogeneous $`A_0(x)=\frac{1}{2}\theta ^{}(x)`$, leading to an electric field
$$E(x)=\frac{1}{2}\theta ^{\prime \prime }(x)$$
(18)
Given that $`\theta (x)`$ itself has a kink-like spatial profile, the electric field is such that it changes sign as a function of $`x`$, as shown in Fig. 2 (we choose $`\theta ^{}>0`$). This electric field acts on the Dirac sea to polarize the vacuum by aligning the virtual vacuum dipoles of the Dirac sea, producing a localized build-up of charge near the kink center. But at nonzero temperature, the electric field also has an effect on the thermal plasma, as we show below.
First, consider the full derivative expansion (12) of the even part of the resolvent, at low but nonzero temperature. At fifth order, there are three independent terms involving $`\theta ^{\prime \prime \prime \prime \prime }`$, $`\theta ^{\prime \prime \prime }(\theta ^{})^2`$, and $`(\theta ^{})^5`$. The $`\theta ^{\prime \prime \prime \prime \prime }`$ term vanishes when integrated over $`x`$, but the other two terms are generally nonzero. However, as $`T0`$ the $`(\theta ^{})^5`$ term dominates the $`\theta ^{\prime \prime \prime }(\theta ^{})^2`$ term. Indeed, for low temperature, the dominant term with $`(2l1)`$ derivatives in the derivative expansion (12) involves $`(\theta ^{})^{2l1}`$. Then, using the chirally rotated form (17) of the Lagrangian, the dominant term at $`(2l1)^{\mathrm{th}}`$ order is simply:
$$N_{\mathrm{dom}}^{(2l1)}=\left(T\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\frac{dk}{2\pi }\frac{\mathrm{tr}([\gamma ^0(p/+m)]^{2l})}{(p^2+m^2)^{2l}}\right)𝑑x\left(\frac{\theta ^{}}{2}\right)^{2l1}$$
(19)
with Euclidean $`p=(\omega _n,k)`$ and $`\omega _n=(2n+1)\pi T`$ the Matsubara modes.
At zero temperature, all these terms $`N^{(2l1)}`$ vanish, except for $`l=1`$. This fact is not obvious; it involves highly nontrivial cancellations between terms in the expansion of the trace. But at nonzero temperature, all the terms in (19) are non-vanishing. Moreover, they have a remarkably simple low temperature $`(Tm)`$ limit:
$$N^{(2l1)}=\delta _{l,1}𝑑x\frac{\theta ^{}}{2\pi }\sqrt{\frac{2mT}{\pi }}e^{m/T}\frac{1}{(2l1)!}𝑑x\left(\frac{\theta ^{}}{2T}\right)^{2l1}+\mathrm{}$$
(20)
Thus, in the low temperature limit, we can resum the entire derivative expansion, to obtain the induced fermion number in the sigma model case (3,4) :
$$N=𝑑x\frac{\theta ^{}}{2\pi }\sqrt{\frac{2mT}{\pi }}𝑑xe^{m/T}\mathrm{sinh}\left(\frac{\theta ^{}}{2T}\right)+\mathrm{}$$
(21)
where the dots refer to subleading terms for $`Tm`$.
Several features of this result (21) deserve comment. First, note that at zero temperature, only the first term survives, producing the familiar result (6) that the induced fermion number depends on the chiral field $`\theta (x)`$ only through its asymptotic value $`\widehat{\theta }\theta (\mathrm{})`$. At zero temperature, one can invoke Lorentz invariance to constrain the form of higher order corrections to (5) to be total derivatives wilczek , but these arguments do not apply at finite temperature. We see this in (21): the temperature dependent corrections are not total derivatives of terms made from $`\theta `$ and its derivatives. At nonzero temperature this shows clearly that the induced fermion number is nontopological - it depends also on the detailed shape of $`\theta (x)`$. Second, the resummed exponential factors $`e^{(m\theta ^{}/2)/T}`$ in (21) are consistent with the derivative expansion assumption that $`\theta ^{}m`$. Finally, the form of these exponential factors suggests an interpretation of the derivative expansion as an adiabatic change of the local Fermi level with a local chemical potential $`\mu =\theta ^{}/2`$, which once again is only sensible in the derivative expansion regime where $`\theta ^{}m`$.
To make this physical picture more precise, we can interpret the result (21) as follows. The first, topological, term refers to the induced charge coming from the polarization of the Dirac sea. This is temperature independent as the short-lived virtual “electron-positron dipoles” of the Dirac sea do not come to thermal equilibrium. The next term in (21) corresponds to the induced charge arising from the response of the real charges in the thermal plasma to the spatially inhomogeneous electric field (18). Indeed, the linear response lebellac of the plasma at low temperature to such an electric field yields an induced fermion number density
$$\rho (x)=\frac{dk}{2\pi }f(x,k)$$
(22)
where the static distribution function $`f(x,k)`$ satisfies the Boltzmann equation
$$v\frac{}{x}f(x,k)=E(x)\frac{}{k}f(x,k)$$
(23)
where $`v=k/\sqrt{k^2+m^2}`$. Regarding $`\mu =\frac{1}{2}\theta ^{}(x)`$ as a local chemical potential, (23) is satisfied by local Fermi particle and antiparticle distribution functions
$$f_\pm (x,k)=\frac{1}{e^{\beta (\sqrt{k^2+m^2}\mu )}+1}$$
(24)
Inserting $`f=f_+f_{}`$ into (22), we obtain precisely the second, nontopological, term in (21) in the low temperature limit.
At T=0, the fermion number may be defined as a sharp observable gk ; but at $`T>0`$, thermal fluctuations introduce an rms deviation. Thus, the finite T fermion number in (7,21) is a thermal expectation value $`N`$, as in the monopole cases cp ; goldhaber ; monopole . We have estimated $`N^2N^2`$, in the derivative expansion regime, in an analogous manner to the computation presented here for $`N`$. We find that the rms deviation vanishes at T=0, but at nonzero T can be significant compared to the thermal shift in (21). Details of this will be reported elsewhere.
To conclude, we comment briefly on possible implications of these results for models in other dimensions for which there is an induced fermion number due to some nontrivial background. In $`2+1`$ dimensions, fermions in a static magnetic background acquire an induced charge that is topological, expressed in terms of the net magnetic flux of the background. At finite temperature, the induced charge remains topological, but is multiplied by a smooth function of the temperature babu . In $`3+1`$ dimensions, fermions in a static Dirac monopole background acquire an induced charge that is temperature dependent at finite $`T`$, but still only depends on the background through the total magnetic charge and the self-adjoint extension parameter cp ; monopole . A more interesting case is a static ’t Hooft-Polyakov monopole background, which has a characteristic size scale. Consider, for example, the coupling
$$_{\mathrm{int}}=\overline{\psi }(A/+\varphi +i\gamma _5m)\psi $$
(25)
where $`\psi `$ is an isodoublet fermion, $`A_\mu `$ is a static $`SU(2)`$ ’t Hooft-Polyakov monopole, and $`\varphi `$ is the corresponding static Higgs field. We have computed the finite temperature induced fermion number, using the $`3+1`$ trace identity used in the zero temperature case manu , and we find precisely the same expression (10) as in the $`1+1`$ kink case, with the identification $`\widehat{\theta }=\mathrm{arctan}(\widehat{\varphi }/m)`$, where $`\widehat{\varphi }`$ is the asymptotic value of the magnitude $`|\varphi |=\sqrt{\varphi ^a\varphi ^a}`$ of the Higgs field. Given that (10) reduces to (6) at $`T=0`$, this monopole result is consistent with the familiar zero temperature result gw ; manu that the induced fermion number is proportional to $`\widehat{\theta }`$ comment . The $`3+1`$ dimensional analogue of the $`1+1`$ sigma model case (3,4) is the sigma model with coupling
$`_{\mathrm{int}}`$ $`=`$ $`m\overline{\psi }\left(\pi _0+i\gamma _5\stackrel{}{\pi }\stackrel{}{\tau }\right)\psi `$ (26)
$`=`$ $`m\overline{\psi }\left({\displaystyle \frac{1}{2}}(g+g^{})+{\displaystyle \frac{1}{2}}(gg^{})\gamma _5\right)\psi `$
where $`\stackrel{}{\tau }`$ are $`su(2)`$ generators, the fields $`\pi _0`$ and $`\stackrel{}{\pi }`$ are constrained by $`\pi _0^2+\stackrel{}{\pi }^2=1`$, and the fields $`g`$ in the second line are defined by $`g=\pi _0+i\stackrel{}{\pi }\stackrel{}{\tau }`$. At zero temperature, there is an induced topological charge density
$$J^0=\frac{1}{24\pi ^2}ϵ^{ijk}\mathrm{tr}\left(g^1_igg^1_jgg^1_kg\right)$$
(27)
The corresponding zero temperature integrated charge is given by the winding number of the background field $`g`$ at zero temperature. We conjecture that at finite temperature this induced charge will acquire additional nontopological contributions similar to those found here for the $`1+1`$ sigma model case.
Acknowledgements:
This work has been supported in part (GD) by the U.S. Department of Energy grant DE-FG02-92ER40716.00, and by PPARC. GD thanks Balliol College, Oxford, for a Visiting Fellowship, and the Theoretical Physics Department at Oxford, and the CSSM at Adelaide for their hospitality and support.
|
warning/0006/cond-mat0006459.html
|
ar5iv
|
text
|
# Pipe network model for scaling of dynamic interfaces in porous media
## Abstract
We present a numerical study on the dynamics of imbibition fronts in porous media using a pipe network model. This model quantitatively reproduces the anomalous scaling behavior found in imbibition experiments \[Phys. Rev. E 52, 5166 (1995)\]. Using simple scaling arguments, we derive a new identity among the scaling exponents in agreement with the experimental results.
Self-affine interfaces are found in a variety of phenomena such as two-phase flow in porous media, thin film deposition, flame fronts, etc. . In particular, scaling properties of imbibition fronts have been the focus in many experimental studies . Static scaling behavior has been suggested to be described by a directed depinning percolation model . To explain dynamic properties, one need to go beyond local models and properly deal with the effects of long-range coupling through pressure. Much progress has been obtained using theoretical approaches including consideration of capillary waves , a Flory-type scaling argument and very recently a phase-field model . These studies contributed to our better understanding of the scaling behavior of imbibition. However, exponents provided in general do $`not`$ compare satisfactorily with the experimental values. In fact, of the four exponents determined experimentally by Horváth and Stanley with good precision, $`none`$ has previously been explained analytically or numerically, and no existing exponent identity applies. In this work, we successfully reproduce all these exponents using a pipe network model once the relevant model parameters are properly fine tuned. A new exponent identity among these exponents is also presented. In addition, we have found an empirical relation between the spatial and temporal correlations.
Horváth and Stanley studied imbibition fronts climbing up in vertical filter paper sheets which move continuously downward into a water container at constant speed $`v`$ . The interface height at horizontal coordinate $`x`$ and time $`t`$ is denoted by $`h(x,t)`$. The average height $`\overline{h}`$ and width $`w`$ were demonstrated to scale as
$`v\overline{h}^\mathrm{\Omega }\text{ and }wv^\kappa `$ (1)
The temporal height-height correlation function $`C(t)`$=$`<[h(x,t^{}+t)h(x,t^{})]^2>^{1/2}`$ was shown to follow a scaling form,
$$C(t)=v^\kappa f(tv^{(\theta _t+\kappa )/\beta }),$$
(2)
where $`f(u)u^\beta `$ for small $`u`$, and it converges to a constant for large $`u`$. The exponents were found to be
$$\mathrm{\Omega }=1.594\text{ , }\kappa =0.48\text{ , }\theta _t=0.37\text{ , and }\beta =0.56.$$
(3)
The first exponent has particularly strong implications because Darcy’s law, which is well established for many porous media , implies $`\mathrm{\Omega }=1`$. The same result was also found in numerous numerical studies . Physically, the anomalous value $`\mathrm{\Omega }=1.594`$ suggests that as the interface height increases, its propagation speed becomes slower than simple considerations of capillary pressure and viscous drag would suggest. This cannot be explained by gravity, evaporation, deformation of wetted paper , or inertial effects since they involve intrinsic length or time scales and can only introduce transients or cut-offs to Darcy’s prediction.
Our network model consists of cylindrical pipes connecting volumeless nodes on a two-dimensional square lattice with periodic boundary conditions in the horizontal direction. According to Possiuelle’s law , the flow rate in a pipe of radius $`r`$ is $`Q=\pi r^4\mathrm{\Delta }p/8\mu d`$ where $`\mathrm{\Delta }p`$ is the pressure difference between two points separated by $`d`$. The fluid viscosity is denoted by $`\mu `$. The pressure at the water level in the tank is maintained at the atmospheric value. The capillary pressure is $`2\gamma \mathrm{cos}\varphi /r`$ in each partially filled pipe where $`\gamma `$ and $`\varphi `$ are the surface tension and the contact angle respectively. Gravity is neglected. This is because in Ref. , scalings were reported only for $`\overline{h}`$ smaller than 65% of the maximum value limited by gravity at $`v=0`$ . We have checked numerically that gravity is negligible in this range, even though it can lead abruptly to the breakdown of Eq. (1) at larger $`\overline{h}`$ and finally to the pinning transition. We calculate the pressure at every wetted node above the water level of the tank by solving Kirchoff’s equations using a locally adaptive over-relaxation method. The propagation of the interface during the associated adaptive time step is hence computed. For computational convenience, possible deviation from Possiuelle’s law at the junction of the pipes is neglected as usual and menisci cannot retreat after reaching a node. Since in real situation air can escape from either side of the sheet, trapping of air is irrelevant and its flow is not simulated.
The pipe model at this point is similar to those in previous studies most of which focused on flow inside porous rocks. Tenuous percolation type wetting patterns are obtained. To simulate more compact patterns observed in paper, we therefore implement a new wettability rule to impose a strong effective surface tension in our network. Now, water can enter an empty pipe from a wetted node only if the neighborhood is sufficiently wet. Specifically, of the 8 most nearby nodes (the 4 nearest and the 4 next nearest neighbors) forming a loop around the node under consideration, there must exist a connected subset of 5 which are already wet. This is the most stringent criterion for the wettability of pipes connected to a wetted node without halting the imbibition entirely or involving consideration of further neighbors. It leads to a very strong effective surface tension. Different and more detailed wettability rules were discussed in Ref. . In addition, fluid pathways in randomly arranged fibers have complicated geometry and there are abundance of narrow bottlenecks and large pores. We therefore propose another important characteristic of our network, namely large spatial fluctuations in the local properties. Its implementation will be discussed later.
The model was first tested in the presence of small fluctuations. Adopt a unit lattice spacing and specify the fluid properties by $`\mu =\gamma \mathrm{cos}\varphi =1`$ . We assume random pipe radii in the range \[0,0.5\] uniformly distributed . This distribution is broader than those used by others . The Darcy’s prediction $`\mathrm{\Omega }=1`$ is readily verified.
To simulate large fluctuations expected in real paper, we simply replace a fraction $`p_w`$ of pipes by wider ones with radius $`r_w`$. For sufficiently large $`p_w`$ and $`r_w`$, $`\mathrm{\Omega }>1`$ is obtained. Both exponents $`\mathrm{\Omega }`$ and $`\kappa `$ then depend non-trivially but continuously on $`p_w`$ and $`r_w`$. Fortunately, we have found that nice power-laws with the experimental values of $`\mathrm{\Omega }`$ and $`\kappa `$ can be reproduced if we put $`p_w=0.18`$ and $`r_w=2.0`$ . We will adopt these parameters in the remainder of this work. Snapshots of simulations of imbibition in stationary sheets are shown in Fig. 1.
We have simulated imbibition in moving paper sheets considered in Ref. . This involves continuous upward shifting of the level on the lattice which represents the contact line of the paper sheet with the water in the reservoir. The interface height $`\overline{h}`$ and width $`w`$ are measured after steady state has been attained and are averaged over 3 independent runs using lattices of width 200 and height 1000. Figures 2(a) and (b) plot respectively $`v`$ and $`w`$ against $`\overline{h}`$. The linearity observed in log-log plots verifies the scaling relations in Eq. (1). We get $`\mathrm{\Omega }=1.62\pm 0.05`$ and $`\kappa =0.49\pm 0.03`$ close to the experimental values in Eq. (3) due to the fine tuning of the pipe distribution mentioned above. The quoted errors represent only the uncertainties in the linear fits.
The spatial correlation $`C(l)`$=$`<`$$`[h(x`$+$`l,t)`$-$`h(x,t)]^2`$$`>^{1/2}`$ and the temporal counterpart $`C(t)`$ defined earlier are computed and plotted in Figs. 3(a) and (b) respectively. The initial linear regions corresponding to $`C(l)l^\alpha `$ and $`C(t)t^\beta `$ for small $`l`$ and $`t`$ give $`\alpha =0.61\pm 0.01`$ and $`\beta =0.63\pm 0.01`$ respectively. This value of $`\beta `$ is in reasonable agreement with the experimental value 0.56. We have verified that $`C(t)`$ in Fig. 3(b) follows the experimentally motivated scaling forms in Eq. (2) and we obtain $`\theta _t=0.38\pm 0.02`$ in excellent agreement with the experimental value 0.37. Similarly, the spatial correlation follows an analogous scaling form $`C(l)=v^\kappa g(lv^{(\theta _l+\kappa )/\alpha })`$ where $`g(u)u^\alpha `$ for small $`u`$ and it converges to a constant for large $`u`$. A further result is obtained by first reparametrizing $`C(t)`$ by the vertical displacement $`z=vt`$ of the sheet. Equation (2) then becomes $`C(z)=v^\kappa f(zv^{(\theta _z+\kappa )/\beta })`$, where $`\theta _z=\theta _t\beta `$. We note that both $`C(l)`$ and $`C(z)`$ can be collapsed individually using a common set of exponents $`\alpha =\beta =0.62`$ and $`\theta _l=\theta _z=0.24`$. Furthermore, they can all be collapsed together as shown in Fig. 3(c) verifying the empirical identities
$`\alpha =\beta \text{ , }\theta _l=\theta _z\text{ and }f(cu)=g(u)`$ (4)
where $`c=1.6`$.
Next, we simulate imbibition in stationary paper sheets. The interface height $`\overline{h}`$ and width $`w`$ averaged over 7 lattices of width $`1000`$ are found to scale as
$$\overline{h}t^ϵ\text{ },\text{ and }wt^{\beta _w}$$
(5)
where $`ϵ=0.374\pm 0.005`$ and $`\beta _w=0.29\pm 0.01`$. The interface speed $`v(t)=d\overline{h}/dt`$ hence computed from direct numerical differentiation on our data is plotted as a function of $`\overline{h}`$ in Fig. 2(a). The width $`w`$ is also plotted against $`\overline{h}`$ in Fig. 2(b). Direct comparison with the moving sheet results is possible because the paper speed $`v`$ in that case is also the average interface speed in the paper frame. These plots show that the scaling relations in Eq. (1) apply to stationary sheets with the same exponents as well. Therefore, we can derive Eq. (1) from Eq. (5) and vice versa. In the derivation, one obtains
$$ϵ=\frac{1}{\mathrm{\Omega }+1}\text{ and }\beta _w=\frac{\mathrm{\Omega }\kappa }{\mathrm{\Omega }+1},$$
(6)
which describe our exponents accurately.
For a stationary sheet, $`w`$ at a given $`\overline{h}`$ is only about 80% of the corresponding value for a moving sheet as observed from Figs. 2(b). The roughness has therefore not yet fully developed for the given height $`\overline{h}`$. In contrast, for the less noisy case the roughness has saturated completely and can be determined solely from $`\overline{h}`$ . More importantly, our result indicates that the roughness gets neither closer to nor further from saturation as $`\overline{h}`$ increases since $`w`$ is always 80% of the saturated value. This constancy of the degree of saturation implies that the system has a unique dynamic time scale dictating both the growth of the roughness and the steady state dynamics of a roughened interface. The time it takes to develop a width $`w`$ is $`w^{1/\beta _w}`$ from Eq. (5). It is thus proportional to the relaxation time extracted from the temporal correlation function $`v^{(\theta _t+\kappa )/\beta }w^{(\theta _t+\kappa )/(\beta \kappa )}`$ from Eqs. (2) and (1). Therefore, $`1/\beta _w=(\theta _t+\kappa )/(\beta \kappa )`$. Applying Eq. (6), we have
$$\beta =\mathrm{\Omega }(\theta _t+\kappa )/(\mathrm{\Omega }+1).$$
(7)
This identity is particularly important because all values have been measured experimentally in Ref. . Inserting the experimental values into the r.h.s. of Eq. (7), we get $`\beta =0.52`$ in reasonable agreement with the experimental value $`0.56`$. Using our numerical estimates instead, we get $`\beta =0.54`$ which is a little smaller than the numerically found value 0.61. In the later case we relate the discrepancy to errors in the numerical determination of $`\beta `$. This may be due to lattice discretization effects which becomes more important at $`C(t)1`$. The spatial counterpart $`\alpha `$ should hence suffer a similar problem. Furthermore, noticeably different values of $`\alpha `$ and $`\beta `$ are obtained if correlations of higher moments are used .
The good agreement between our numerical results and the experimental ones strongly supports that our model has captured the essential physics in the imbibition process. However, some further points are yet to be considered. First, despite the nice power-law fits and data collapses of the relevant quantities observed, existence of extraordinarily slow crossover effects masking different asymptotic scalings should be cautioned. Second, our model reproduces the experimental results after fine tuning of the radius distribution. We have also found a few other completely different distributions which also lead to similar exponents upon fine tuning. The exponents nevertheless are $`not`$ robust with respect to changes in the details of the models. Using distributions other than the fine tuned ones, the exponents are in general different and sometimes scalings may not even hold. We expect that there is a yet unknown selection mechanism so that models generating the experimental exponents are preferred. This is currently under active investigations. However, one cannot determine $`apriori`$ a correct realistic pipe distribution. This is because all pipe networks are simplified models of fluid pathways which are indeed very different from those of paper. Realistic simulations, for example, using lattice-Boltzmann method with sophisticated boundary conditions in 3 dimensions unfortunately covers only microscopic regions and thus cannot be applied to study the macroscopic scalings.
Finally, local models of imbibition have been classified into isotropic and anisotropic universality classes exemplified respectively by the random field Ising model (RFIM) and the directed percolation depinning (DPD) model . The anisotropy in the DPD model is due to a solid-on-solid condition. In contrast, for our model both the wettability rule and the flow dynamics treat the horizontal and vertical directions equivalently. Concerning the local symmetry, it is more closely related to the isotropic class. It was found that $`\alpha =1`$ and $`\beta _w=3/4`$ for the isotropic class while $`ϵ=\alpha =\beta _w0.633`$ for the anisotropic one at criticality. This is to be compared with $`ϵ0.374`$, $`\alpha 0.61`$ and $`\beta _w0.29`$ for our model. It is easy to see that the much larger values of $`ϵ`$ and $`\beta _w`$ for DPD and also $`\beta _w`$ for RFIM are due to the locality of the interections. We believe that the agreement of $`\alpha `$ between DPD and our model is just a coincidence. The morphologies of the surfaces are indeed visually quite different. Previous works on non-local models including pipe networks and the phase-field model are in a weak fluctuations regime and are all consistent only with Darcy’s law. Our pipe network with strong fluctuations is the only model exhibiting distinctly different scalings in good agreement with the experimental ones in Ref. . We therefore suggest that it belongs to a new universality class.
In conclusion, we have simulated imbibition in paper using a pipe network model and reproduced all scaling behaviors observed experimentally in Ref. . We obtain
$$\mathrm{\Omega }=1.62\text{ , }\kappa =0.49\text{ , }\theta _t=0.38\text{ and }\beta =0.63$$
(8)
which is to be compared with Eq. (3). The model displays rich behaviors, and can be tuned to reproduce the experimentally determined values of $`\mathrm{\Omega }`$ and $`\kappa `$. Then $`\beta `$ and $`\theta _t`$ turn out respectively in reasonable and excellent agreement with the experimental values. Assuming a single time scale in the dynamics, we have presented a new exponent identity (Eq. (7)) which is justified by the experimental values. Two other exponent identities in Eq. (6) relating the moving and stationary paper cases are deduced and verified numerically. Further identities in Eq. (4) for exponents and scaling functions are suggested empirically based on a data collapse between the spatial and temporal correlations.
We have benefitted from helpful communications with L.M. Sander, M Rost, M. Dubè, T. Ala-Nissila and F.G. Shin who are gratefully acknowledged. C.H.L. is supported by project no. B-Q075 from Research Grants Council of Hong Kong SAR. V.K.H is supported by the Hungarian Science Foundation grant OTKA-F17310 and NATO grant DGE-9804461.
|
warning/0006/hep-ph0006352.html
|
ar5iv
|
text
|
# CURRENT ISSUES IN PROMPT PHOTON PRODUCTION
## References
|
warning/0006/quant-ph0006046.html
|
ar5iv
|
text
|
# A counterexample to a conjectured entanglement inequality
## Abstract
We give an explicit counterexample to an entanglement inequality suggested in a recent paper \[quant-ph/0005126\] by Benatti and Narnhofer. The inequality would have had far-reaching consequences, including the additivity of the entanglement of formation.
The problem of additivity of entanglement of formation is one of the fundamental open issues in the theory of entanglement. In a recent paper Fabio Benatti and Heide Narnhofer bring to bear on this problem their intuition gained from a completely different enterprise (the study of quantum dynamical entropy), using the equivalence between the entanglement of formation and a quantity called the “entropy of a subalgebra” with respect to a state (a connection noted earlier by Uhlmann and others). They achieve some interesting partial results supporting the additivity conjecture<sup>*</sup><sup>*</sup>*We only mention in passing an omission in the formulation of Proposition 1 in : the unitary $`U`$ must be assumed to factorize. .
In this brief note we will concentrate on an exciting prospect coming up in their paper as inequality (12) (restated below as equation (3)). Benatti and Narnhofer apparently found it as an inequality one would just love to have for a simple proof of additivity, but seem undecided as to its validity. Given the simplicity and generality of the inequality, probably many more results would follow from it. It therefore seemed necessary to us to decide the issue quickly. Unfortunately, (if one can ever say that of a mathematical statement) the inequality turned out to be false in general.
The inequality in question refers to a vector $`\mathrm{\Psi }`$ in a fourfold tensor product $`\mathrm{\Psi }_1_2_3_4`$, of which the factors $`_1`$ and $`_3`$ are associated with one party (say Alice) and factors $`_2`$ and $`_4`$ are associated with another, Bob. We can write $`\mathrm{\Psi }`$ in Schmidt form as
$$\mathrm{\Psi }=\underset{\alpha }{}\sqrt{\lambda _\alpha }\mathrm{\Phi }_\alpha ^{12}\mathrm{\Phi }_\alpha ^{34}$$
(1)
with respect to another split $`12|34`$, of the system, which is the split according to which additivity would be investigated. Of course, by definition of the Schmidt decomposition, we have $`_\alpha \lambda _\alpha =1`$, and the vectors $`\mathrm{\Phi }_\alpha ^{12}_1_2`$ and $`\mathrm{\Phi }_\alpha ^{34}_3_4`$ each form an orthonormal set. Let us denote by $`S(\rho )`$ the von Neumann entropy of a density operator $`\rho `$, and by $`\mathrm{tr}_i(A)`$ (resp. $`\mathrm{tr}_{ij}(A)`$) the partial trace of the operator $`A`$ with respect to the Hilbert space $`_i`$ (resp. $`_i_j`$), assumed to be a tensor factor of the space on which $`A`$ lives. Then the inequality suggested by Benatti and Narnhofer is
$`S\left(\mathrm{tr}_{24}(|\mathrm{\Psi }\mathrm{\Psi }|)\right)`$ $``$ $`{\displaystyle \underset{\alpha }{}}\lambda _\alpha \{S\left(\mathrm{tr}_2(|\mathrm{\Phi }_\alpha ^{12}\mathrm{\Phi }_\alpha ^{12}|)\right)+`$ (3)
$`+S\left(\mathrm{tr}_4(|\mathrm{\Phi }_\alpha ^{34}\mathrm{\Phi }_\alpha ^{34}|)\right)\}.`$
A superficial random numerical test might come out in favor of (3). A counterexample is found, however, with all Hilbert spaces $`_i=\text{ }\text{C}^d`$, $`d`$ arbitrary, namely
$$\mathrm{\Psi }=\frac{1}{d}\underset{i,k=1}{\overset{d}{}}e_ie_ke_ie_k=\frac{1}{d}\underset{\alpha =1}{\overset{d^2}{}}\mathrm{\Phi }_\alpha \overline{\mathrm{\Phi }_\alpha },$$
(4)
where the vectors $`e_i`$ form an orthonormal basis of $`\text{ }\text{C}^d`$. Thus $`\mathrm{\Psi }`$ is the tensor product of the vector $`d^{1/2}_ie_ie_i`$ for Alice and the same vector for Bob, making the left hand side of inequality (3) zero. $`\mathrm{\Psi }`$ is also maximally entangled for the split $`12|34`$. One Schmidt decomposition is into the vectors $`e_ie_k`$ (which would make the right hand side of (3) zero, too). But in the maximally entangled case the Schmidt decomposition is highly non-unique, and we may also choose another decomposition using entangled vectors $`\mathrm{\Phi }_\alpha `$ and their complex conjugates $`\overline{\mathrm{\Phi }_\alpha }`$ with respect to the basis $`e_i`$. We may even take them maximally entangled , whence the right hand side of (3) becomes $`2\mathrm{log}d`$. By deforming the coefficients in the latter decomposition slightly, we also get examples with unique Schmidt decomposition such that LHS(3)$`0`$ and RHS(3)$`2\mathrm{log}d`$.
|
warning/0006/hep-th0006118.html
|
ar5iv
|
text
|
# Untitled Document
Monopole supersymmetries
and
the Biedenharn operator
P. A. HORVÁTHY (<sup>1</sup>) Permanent address: Laboratoire de Mathématiques et de Physique Théorique, Université de TOURS. F-37 200 TOURS (France). e-mail: horvathy@univ-tours.fr, A. J. MACFARLANE (<sup>2</sup>) Permanent address: Centre for Mathematical Sciences, DAMTP, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK. e-mail: a.j.macfarlane@damtp.cam.ac.uk, J.-W. VAN HOLTEN (<sup>3</sup>) e-mail : t32@nikhef.nl
NIKHEF, Postbus 41882
1009 DB Amsterdam, The Netherlands
Abstract. The hidden supersymmetry of the monopole found by De Jonghe et al. is generalized to a spin $`\frac{1}{2}`$ particle in the combined field of a Dirac monopole plus a $`\lambda ^2/r^2`$ potential \[considered before by D’Hoker and Vinet\], and related to the operator introduced by Biedenharn a long time ago in solving the Dirac-Coulomb problem. Explicit solutions are obtained by diagonalizing the Biedenharn operator.
1. Introduction
In , De Jonghe et al. found that a spin-$`\frac{1}{2}`$ particle of mass, here chosen $`m=\frac{1}{2}`$, and charge $`q`$ in the field of a Dirac monopole of unit strength, described by the Pauli Hamiltonian
$$H_D=\left(\stackrel{}{D}^2q\frac{\stackrel{}{\sigma }\stackrel{}{r}}{r^3}\right),$$
$`(1.1)`$
admitted a ‘hidden’ supercharge $`\stackrel{~}{Q}_D`$ given by
$$\stackrel{~}{Q}_D=\left(\stackrel{}{\sigma }\stackrel{}{\mathrm{}}+1\right),$$
$`(1.2)`$
where $`\stackrel{}{\mathrm{}}`$ is the \[non-conserved\] orbital part of the angular momentum, $`\stackrel{}{\mathrm{}}=\stackrel{}{r}\times \stackrel{}{\pi }`$, $`\stackrel{}{\pi }=\stackrel{}{p}q\stackrel{}{A}`$. Their discussion is based on the study of Killing-Yano tensors. The supercharge $`\stackrel{~}{Q}_D`$ anticommutes with the supercharge $`Q_D`$ found by D’Hoker and Vinet , where
$$Q_D=\stackrel{}{\sigma }\stackrel{}{\pi }.$$
$`(1.3)`$
These supercharges form, together with the \[conserved\] total angular momentum $`\stackrel{}{J}=\stackrel{}{L}+\frac{\stackrel{}{\sigma }}{2},\stackrel{}{L}=\stackrel{}{\mathrm{}}q\stackrel{}{r}/r,`$ a closed, non-linear algebra .
In this Letter, we (i) generalize the result of De Jonghe et al.; (ii) relate it to earlier work of Dirac , Biedenharn , and Berrondo and McIntosh ; (iii) use it to solve the system, and (iv) discuss a full Minkowski space generalisation.
Very recently , Plyushchay discussed related problems, but from a rather different veiwpoint : while our results here are derived from supersymmetric quantum mechanics, he uses pseudoclassical mechanics with anticommuting (Grassmann) variables. See also the comments on in .
2. The generalized monopole system
Let $`\stackrel{}{A}_D`$ denote the vector potential of the Dirac monopole of unit strength. Let $`\lambda `$ be a positive constant, $`q>0`$ a half-integer, and consider, setting $`A_4\lambda /r`$ and $`\stackrel{}{A}q\stackrel{}{A}_D`$, the gauge field $`A_\alpha `$, ($`\alpha =1,\mathrm{},4`$) on $`𝐑^4`$. We use the Dirac matrices
$$\stackrel{}{\gamma }=\left(\begin{array}{cc}& i\stackrel{}{\sigma }\\ i\stackrel{}{\sigma }& \end{array}\right),\gamma ^4=\left(\begin{array}{cc}& \mathrm{𝟏}_2\\ \mathrm{𝟏}_2& \end{array}\right),\gamma ^5=\left(\begin{array}{cc}\mathrm{𝟏}_2& \\ & \mathrm{𝟏}_2\end{array}\right).$$
$`(2.1)`$
When restricted to fields which do not depend on $`x^4`$, the associated Dirac operator $`D\text{/}\gamma ^\mu D_\mu `$, $`D_\mu _\mu iA_\mu `$, reads
$$D\text{/}=\left(\begin{array}{cc}& T^{}\\ T& \end{array}\right)=\left(\begin{array}{cc}& \stackrel{}{\sigma }\stackrel{}{\pi }i\frac{\lambda }{r}\\ \\ \stackrel{}{\sigma }\stackrel{}{\pi }+i\frac{\lambda }{r}& \end{array}\right).$$
$`(2.2)`$
$`D\text{/}`$ anticommutes with the chirality operator $`\gamma ^5`$.
The square of the Dirac operator,
$$H=\left(\begin{array}{cc}H_1& \\ & H_0\end{array}\right)=\left(\begin{array}{cc}T^{}T& \\ & TT^{}\end{array}\right)=\left\{\stackrel{}{\pi }^2q\frac{\stackrel{}{\sigma }\stackrel{}{r}}{r^3}+\frac{\lambda ^2}{r^2}\lambda \gamma ^5\frac{\stackrel{}{\sigma }\stackrel{}{r}}{r^3}\right\},$$
$`(2.3)`$
is a supersymmetric Hamiltonian : the non-zero-energy parts of the chiral sectors, defined as the eigensectors of $`\gamma ^5`$, are intertwined by the unitary transformations
$$U=T\frac{1}{\sqrt{H_1}}\text{and}U^1=U^{}=\frac{1}{\sqrt{H_1}}T^{},UH_1U^{}=H_0.$$
$`(2.4)`$
The partner hamiltonians $`H_1`$ and $`H_0`$, have therefore the same positive spectra, and the spectrum of $`D\text{/}`$ can be obtained from that of $`D\text{/}^2`$. For further analysis, it is convenient to set
$$\sigma _r=\stackrel{}{\sigma }(\stackrel{}{r}/r),z=\stackrel{}{\sigma }\stackrel{}{\mathrm{}}+1.$$
$`(2.5)`$
Note that $`\sigma _r{}_{}{}^{2}=1`$, and that $`z`$ anticommutes with $`\sigma _r`$ and $`\stackrel{}{\sigma }\stackrel{}{\pi }`$, $`\{z,\sigma _r\}=0`$ and $`\{z,\stackrel{}{\sigma }\stackrel{}{\pi }\}=0.`$ Also $`z`$ is equal to the supercharge $`\stackrel{~}{Q}_D`$ of (1.2). For $`\lambda 0`$, $`z`$ satisfies the relations $`zT=T^{}z`$, $`zT^{}=Tz`$, $`zH_1=H_0z`$ and $`zH_0=H_1z.`$ Therefore, the operator
$$𝒦=\left(\begin{array}{cc}& iz\\ iz& \end{array}\right)i\left(\begin{array}{cc}& \stackrel{}{\sigma }\stackrel{}{\mathrm{}}1\\ \stackrel{}{\sigma }\stackrel{}{\mathrm{}}+1& \end{array}\right)$$
$`(2.6)`$
commutes with the Dirac operator $`D\text{/}`$, and hence also with its square. Using $`(\stackrel{}{\sigma }\stackrel{}{L})^2=\stackrel{}{L}^2+i\stackrel{}{\sigma }(\stackrel{}{L}\times \stackrel{}{L})=\stackrel{}{L}^2\stackrel{}{\sigma }\stackrel{}{L}`$, one proves furthermore that $`𝒦^2=z^2=\stackrel{}{J}^2+1/4q^2.`$ Thus, since the eigenvalues of $`\stackrel{}{J}^2`$ are $`j(j+1)`$, $`j=q1/2,q+1/2,\mathrm{}`$, the operators $`z`$ (and $`𝒦`$) have irrational eigenvalues,
$$\kappa =\pm \sqrt{(j+1/2)^2q^2}.$$
$`(2.7)`$
The operator $`𝒦`$ is hermitian because $`jq1/2`$. For the lowest allowed value $`j=q1/2`$, however, $`\kappa `$ vanishes and $`𝒦`$ is not invertible. The operator $`𝒦`$ has been used by Dirac in the study of the relativistic hydrogen atom long time ago; its form adapted to the monopole, (2.7), was found by Berrondo and McIntosh . It is more convenient to use, however, the hermitian Biedenharn operator
$$\mathrm{\Gamma }=\left(\begin{array}{cc}\stackrel{}{\sigma }\stackrel{}{\mathrm{}}+1+\lambda \sigma _r& \\ & \stackrel{}{\sigma }\stackrel{}{\mathrm{}}+1\lambda \sigma _r\end{array}\right)\left(\begin{array}{cc}y& \\ \\ & x\end{array}\right).$$
$`(2.8)`$
Here $`y=z+\lambda \sigma _r`$ and $`x=z\lambda \sigma _r,`$ so that $`x\sigma _r=\sigma _ry`$. The eigenvalues of $`\mathrm{\Gamma }`$,
$$\gamma =\pm \sqrt{(j+1/2)^2+\lambda ^2q^2},(\text{sign}\gamma =\text{sign}\kappa )$$
$`(2.9)`$
are in general still irrational. However, owing to the presence of $`\lambda ^2`$, the operator $`\mathrm{\Gamma }`$ is invertible whenever $`\lambda ^2>0`$. The clue is that, in terms of $`\mathrm{\Gamma }`$, $`D\text{/}^2`$ becomes simply
$$D\text{/}^2\left(\begin{array}{c}H_1\\ & H_0\end{array}\right)=(_r+\frac{1}{r})^2+\frac{\mathrm{\Gamma }(\mathrm{\Gamma }+1)}{r^2}.$$
$`(2.10)`$
Here $`p_r=i(1/r)_rr=i(_r+1/r)`$ is the hermitian operator conjugate to $`r`$. This is conveniently checked by writing, using the radial form $`\stackrel{}{\sigma }\stackrel{}{\pi }=i\sigma _r(_r+1/rz/r)`$. The supercharges $`T`$ and $`T^{}`$ as
$$\begin{array}{cccccc}& T& =& i\sigma _r\left(_r+\frac{1}{r}\frac{y}{r}\right)& =& i\left(_r+\frac{1}{r}+\frac{x}{r}\right)\sigma _r,\\ \\ & T^{}& =& i\sigma _r\left(_r+\frac{1}{r}\frac{x}{r}\right)& =& i\left(_r+\frac{1}{r}+\frac{y}{r}\right)\sigma _r.\end{array}$$
$`(2.11)`$
The self-adjointness of $`D\text{/}^2`$ requires $`|\lambda |3/2`$ . It follows from our previous formulæ that $`\mathrm{\Gamma }`$ anticommutes with $`D\text{/}`$ and commutes therefore with its square. Hence the shifted operators $`x`$ and $`y`$ commute with the partner hamiltonians, $`[x,H_0]=0=[y,H_1].`$ In conclusion, our $`4`$-component operators satisfy the non-linear algebra
$$\begin{array}{cccc}D\text{/}^2=H,& \{\mathrm{\Gamma },D\text{/}\}=0,& [\mathrm{\Gamma },H]=0,& \mathrm{\Gamma }^2=\stackrel{}{J}^2+1/4+\lambda ^2q^2,\\ \\ [\stackrel{}{J},\mathrm{\Gamma }]=0,& [\stackrel{}{J},D\text{/}]=0,& [\stackrel{}{J},H]=0,& [J_i,J_j]=iϵ_{ijk}J_k,\end{array}$$
$`(2.12)`$
which only differs from that found for the $`2`$-components objects of in the appearance of $`\lambda ^2`$ in $`\mathrm{\Gamma }^2`$. Note that it is now $`D\text{/}`$ which plays the rôle of $`Q_D`$, and $`\mathrm{\Gamma }`$ plays that of $`\stackrel{~}{Q}_D`$.
3. Explicit solutions
A look at (2.10) shows that the Biedenharn operator $`\mathrm{\Gamma }`$ plays clearly a rôle analogous to angular momentum. Since $`\mathrm{\Gamma }`$ and $`J`$ commute, they can be simultaneously diagonalized. A convenient basis is found as follows . Let us first assume that $`jq+1/2`$, and let $`L_\pm =j\pm 1/2`$ be the orbital angular momentum quantum number. Consider first the two-component spinors
$$\phi _\pm ^\mu =\sqrt{\frac{L_\pm +1/2\mu }{2L_\pm +1}}Y_{L_\pm }^{\mu 1/2}\left(\begin{array}{c}1\\ \\ 0\end{array}\right)\sqrt{\frac{L_\pm +1/2\pm \mu }{2L_\pm +1}}Y_{L_\pm }^{\mu +1/2}\left(\begin{array}{c}0\\ \\ 1\end{array}\right),$$
$`(3.1)`$
where the $`Y`$’s are the monopole harmonics, defined in . The $`\phi `$’s satisfy $`\stackrel{}{J}^2=j(j+1),J_3=\mu ,(\mu =j,\mathrm{},j),\stackrel{}{L}^2=L_\pm (L_\pm +1),`$ and the action of $`\sigma _r`$ upon the $`\phi `$’s can be obtained from . Then, dropping the upper index $`\mu `$, and using $`\stackrel{}{\sigma }\stackrel{}{\mathrm{}}=\stackrel{}{J}^2\stackrel{}{L}^23/4+q\stackrel{}{\sigma }\stackrel{}{r}/r`$, it may be shown that the $`2`$-spinors
$$\chi _\pm =\frac{q}{j+1/2+\kappa }\phi _\pm \phi _{}$$
$`(3.2)`$
satisfy $`z\chi _\pm =\pm |\kappa |\chi _\pm `$ and $`\sigma _r\chi _\pm =\chi _{},`$ as well as $`\stackrel{}{J}^2=j(j+1)`$, $`J_3=\mu ,(\mu =j,\mathrm{},j)`$. Hence, defining $`\gamma `$ by $`\gamma ^2=\kappa ^2+\lambda ^2`$, we have
$$\begin{array}{cc}\varphi _+=(|\kappa |+|\gamma |)\chi _+\lambda \chi _{},& \varphi _{}=\lambda \chi _++(|\kappa |+|\gamma |)\chi _{}\\ \\ \mathrm{\Phi }_+=(|\kappa |+|\gamma |)\chi _++\lambda \chi _{},& \mathrm{\Phi }_{}=\lambda \chi _++(|\kappa |+|\gamma |)\chi _{}\end{array}$$
$`(3.3)`$
diagonalize $`x`$ and $`y`$, $`x\varphi _\pm =\gamma \varphi _\pm `$ and $`y\mathrm{\Phi }_\pm =\gamma \mathrm{\Phi }_\pm .`$ The operator $`\sigma _r=\stackrel{}{\sigma }\stackrel{}{r}/r`$ interchanges the $`x`$ and $`y`$ eigenspinors, $`\sigma _r\varphi _\pm =\mathrm{\Phi }_{},`$ a result which also follows directly from $`x\sigma _r=\sigma _ry`$.
When $`j=q\frac{1}{2}`$, there are no $`L_{}=q1`$ states, though: we only have $`2(q\frac{1}{2})+1=2q`$ states with $`L=L_+=q`$ , namely
$$\phi _+^0=\sqrt{\frac{q+1/2+\mu }{2q+1}}Y_q^{\mu 1/2}\left(\begin{array}{c}1\\ \\ 0\end{array}\right)+\sqrt{\frac{q+1/2\mu }{2q+1}}Y_q^{\mu +1/2}\left(\begin{array}{c}0\\ \\ 1\end{array}\right),$$
$`(3.4)`$
Thus, for $`j=q\frac{1}{2}`$, no $`\phi _{}`$ is available, and $`\chi _+^0=\chi _{}^0=\phi _+^0`$ is annihilated by $`z`$. Therefore, there are no $`\varphi _{}`$-states in the $`\gamma ^5=1`$ sector, and no $`\mathrm{\Phi }_+`$ states in the $`\gamma ^5=1`$ sector. In each $`\gamma ^5`$ sector, (3.4) yields in turn $`(2q)`$ states, namely
$$\varphi _+^0=\mathrm{\Phi }_{}^0\phi _+^0.$$
$`(3.5)`$
They are eigenvectors of $`x`$ and $`y`$ with eigenvalues $`\pm \lambda `$, respectively, and are still interchanged by $`\sigma _r`$. In conclusion, the eigenfunctions of $`D\text{/}^2`$ are
$$\begin{array}{ccc}& \{\begin{array}{cc}& \mathrm{\Psi }_\pm =u_\pm \left(\begin{array}{c}\mathrm{\Phi }_\pm \\ 0\end{array}\right)\hfill \\ & \\ & \psi _\pm =u_\pm \left(\begin{array}{c}0\\ \varphi _\pm \end{array}\right)\hfill \end{array}\text{for}\{\begin{array}{c}\gamma ^5=1\\ \\ \\ \gamma ^5=1\end{array}& \text{for}jq+1/2,\\ \\ & \{\begin{array}{cc}& \mathrm{\Psi }_{}^0=u_{}^0\left(\begin{array}{c}\mathrm{\Phi }_{}\\ 0\end{array}\right)\hfill \\ & \\ & \psi _+^0=u_+^0\left(\begin{array}{c}0\\ \varphi _+\end{array}\right)\hfill \end{array}\text{for}\{\begin{array}{c}\gamma ^5=1\\ \\ \\ \gamma ^5=1\end{array}& \text{for}j=q1/2.\end{array}$$
$`(3.6)`$
where $`u_\pm `$ solves the radial equation
$$\left[\left(_r+\frac{1}{r}\right)^2+\frac{\gamma (\gamma +1)}{r^2}E\right]u_\pm (r)=0.$$
$`(3.7)`$
There are no bound states; the scattering states involve Bessel functions:
$$u_\pm (r)=\frac{1}{\sqrt{r}}J_{|\gamma +{\scriptscriptstyle \frac{1}{2}}|}\left(\sqrt{E}r\right).$$
$`(3.8)`$
4. Symmetries
A spin $`0`$ particle in the field of a Dirac monopole has an $`\mathrm{o}(2,1)`$ symmetry, generated by $`HH_0=\stackrel{}{\pi }^2`$ and by dilations and expansions ,
$$D=tH1/2\left\{\stackrel{}{\pi }\stackrel{}{r}\right\}K=\frac{1}{2}t^2H+tD+\frac{1}{2}r^2.$$
$`(4.1)`$
This symmetry has been extended to the Pauli Hamiltonian (1.1) with formally the same generators (4.1), with $`HH_D`$ replacing $`H_0`$ . The supercharge $`Q_D`$ is a square-root of $`H_D`$. Commuting $`Q_D`$ with the expansion, $`K`$, yields a new fermionic generator, namely
$$S=i[Q_D,K]=\frac{1}{\sqrt{2}}\stackrel{}{\sigma }(\stackrel{}{r}\stackrel{}{\pi }t),$$
$`(4.2)`$
and it is then readily proved that the bosonic operators $`H_D`$, $`D`$, $`K`$ close, with $`Q`$ and $`S`$ into an $`\mathrm{osp}(1/1)`$ superalgebra. . Now remarkably
$$i[Q_D,S]\frac{1}{2}=z,$$
$`(4.3)`$
and $`z^2`$ is a Casimir operator of this $`\mathrm{osp}(1/1)`$ .
The same bosonic $`\mathrm{o}(2,1)`$ symmetry arises for the generalized monopole system (2.4). The Dirac operator $`QD\text{/}`$ is a square-root of $`H`$ by construction. However,
$$Q^{}=\gamma ^5Q\left(\begin{array}{cc}& \stackrel{}{\sigma }\stackrel{}{\pi }i\frac{\lambda }{r}\\ \\ \stackrel{}{\sigma }\stackrel{}{\pi }i\frac{\lambda }{r}& \end{array}\right).$$
$`(4.4)`$
is a new square-root, $`\{Q^{},Q^{}\}=H`$. Commuting $`K`$ with $`Q`$ and with $`Q^{}`$ yields
$$S=\gamma ^5\stackrel{}{\gamma }\stackrel{}{r}tQ\text{and}S^{}=i\gamma ^5S.$$
In this way, we get two, independent, super-extensions of the bosonic $`\mathrm{o}(2,1)`$. The two $`\mathrm{osp}(1/1)`$’s do not close yet: the “mixed” anticommutators between the $`Q`$-type and $`S`$-type charges yield a new bosonic charge, namely
$$Y=\{Q,S^{}\}=\{Q^{},S\}=\gamma ^5\left(z+\frac{1}{2}\right)\lambda \sigma _r,$$
$`(4.5)`$
that commutes with the other bosonic charges. The four operators $`H,D,K,Y`$ do close finally with the four fermionic charges $`Q,Q^{},S,S^{}`$,
$$\begin{array}{cccccc}[Q,D]& =& iQ,& [Q^{},D]& =& iQ^{},\\ \\ [Q,K]& =& iS,& [Q^{},K]& =& iS^{},\\ \\ [Q,H]& =& 0,& [Q^{},H]& =& 0,\\ \\ [Q,Y]& =& iQ^{},& [Q^{},Y]& =& iQ,\\ \\ [S,D]& =& iS,& [S^{},D]& =& iS^{},\\ \\ [S,K]& =& 0,& [S^{},K]& =& 0,\\ \\ [S,H]& =& 2iQ,& [S^{},H]& =& 2iQ^{},\\ \\ [S,Y]& =& iS^{},& [S^{},Y]& =& iS,\\ \\ \{Q,Q\}& =& H,& \{Q^{},Q^{}\}& =& H,\\ \\ \{S,S\}& =& 2K,& \{S^{},S^{}\}& =& 2K,\\ \\ \{Q,Q^{}\}& =& 0,& \{S,S^{}\}& =& 0,\\ \\ \{Q,S\}& =& D,& \{Q^{},S^{}\}& =& D,\\ \\ \{Q,S^{}\}& =& Y,& \{Q^{},S\}& =& Y.\end{array}$$
$`(4.6)`$
which are the commutation relations of the $`\mathrm{osp}(1/2)`$ superalgebra, to which spin adds an extra $`\mathrm{o}(3)`$ . Now the Casimir of $`\mathrm{osp}(1/2)`$ is the square of
$$i[Q,S]\frac{1}{2}=i[Q^{},S^{}]\frac{1}{2}=\mathrm{\Gamma },$$
$`(4.7)`$
which provides a nice interpretation for the Biedenharn operator $`\mathrm{\Gamma }`$. Similar algebras were studied in .
5. Particular cases
(i) For $`\lambda =0`$ , we have $`Q=Q^{}=Q_D`$, $`H_1=H_0=H_D`$, the Pauli Hamiltonian in a pure monopole field . The $`4`$-component Hamiltonian is simply diag$`(H_D,H_D)`$; the Biedenharn and the Dirac operators are related as $`\mathrm{\Gamma }=i\gamma ^4𝒦`$. In this case, we recover the formulæ in .
(ii) Another particular value is $`\lambda =\pm q`$, when the situation is similar to that in Taub-NUT space : the spin drops out in one of the chiral sectors, while the Pauli term gets doubled in the other. For $`\lambda =q`$, for example, the Hamiltonian (2.4) reduces to
$$H=\left(\begin{array}{cc}H_1& \\ \\ & H_0\end{array}\right)=\left(\begin{array}{cc}\stackrel{}{\pi }^2+\frac{q^2}{r^2}2q\frac{\stackrel{}{\sigma }\stackrel{}{r}}{zr^3}& \\ \\ & \stackrel{}{\pi }^2+\frac{q^2}{r^2}\end{array}\right).$$
$`(5.1)`$
Here $`H_0`$ describes a spin $`0`$ particle in the combined field of a Dirac monopole and of an inverse-square potential, while $`H_1`$ corresponds to a particle with anomalous gyromagnetic ratio $`4`$. The Biedenharn operator has half-integer eigenvalues, $`\gamma =\pm (j+\frac{1}{2})`$. Note that the $`\gamma ^5=1`$ eigenspinors now reduce to those in Eq. (3.1), $`\varphi _\pm \phi _\pm .`$
Assume first that $`jq+1/2`$. Since $`\gamma (\gamma +1)`$ is now the same for $`|\gamma |`$ as for $`|\gamma |1`$, these values lead to identical solutions. Therefore, in each $`\gamma ^5`$ sector, the energy levels are two-fold degenerate. The numerator of the $`r^2`$ term reads
$$\begin{array}{cc}(j+\frac{3}{2})(j+\frac{1}{2})& \\ \\ & (j+\frac{1}{2})(j\frac{1}{2})\end{array}\text{for}\{\begin{array}{c}\gamma >0\\ \\ \\ \gamma <0\end{array}$$
$`(5.2)`$
so that we indeed get the same equation with $`\gamma >0`$ for $`j`$, as with $`\gamma <0`$ for $`(j1)`$, provided that $`(j1)`$ states do exist. In both cases, (3.8) yields $`r^{1/2}J_{j+1}\left(\sqrt{E}r\right).`$
The two-fold degeneracy is hence also explained by an extra $`\mathrm{o}(3)`$ symmetry in addition to the rotational symmetry: spin is trivially conserved for $`H_0`$, and this is exported to $`H_1`$ by supersymmetry. The extra $`\mathrm{o}(3)`$ symmetry is generated hence by the spin vectors
$$\stackrel{}{S}_0=\frac{1}{2}\stackrel{}{\sigma }\text{for}H_0,\stackrel{}{S}_1=U^{}\stackrel{}{S}_0U\text{for}H_1,$$
$`(5.3)`$
where $`U`$ and $`U^1=U^{}`$ are the intertwiners of (2.4). The two-fold degeneracy corresponds precisely to this $`\mathrm{o}(3)`$ symmetry. For $`j=q1/2`$ , half of the states are missing.
The system admits further symmetries. Firstly, $`H_0`$ admits the non-relativistic conformal $`\mathrm{o}(2,1)`$ symmetry ; supersymmetry exports this to the partner Hamiltonian $`H_1`$. The symmetries combine with $`D\text{/}`$ and $`i\gamma ^5D\text{/}`$ into an $`\mathrm{osp}(2,1)`$ superalgebra .
(iii) Replacing the scalar potential $`\lambda /r`$ by $`q(11/r)`$ — which corresponds to the long-range limit of the scalar field of a self-dual “BPS” monopole — we get
$$\left(\begin{array}{cc}H_1& \\ & H_0\end{array}\right)=\left(\begin{array}{cc}\stackrel{}{\pi }^2\frac{2q^2}{r}+\frac{q^2}{r^2}+q^22q\frac{\stackrel{}{\sigma }\stackrel{}{r}}{r^3}& \\ \\ & \stackrel{}{\pi }^2\frac{2q^2}{r}+\frac{q^2}{r^2}+q^2\end{array}\right).$$
$`(5.4)`$
The “lower” Hamiltonian, $`H_0`$, has again gyromagnetic ratio $`0`$ yielding an extra $`\mathrm{o}(3)`$ symmetry. Its properties have been thouroughly studied by McIntosh and Cisneros, and by Zwanziger , who found that it admits a Kepler-type dynamical symmetry. Its superpartner $`H_1`$ describes a spinning particle of anomalous gyromagnetic ratio $`4`$: this is the ‘dyon’ of D’Hoker and Vinet . Then supersymmetry can be used to transfer the symmetries of $`H_0`$ to $`H_1`$ ; the system can be solved using the Biedenharn method .
6. Minkowski space extension
In this section we turn to the Dirac-equation in Minkowski space-time, with a combined Coulomb-monopole and massless scalar background. This relativistic system of equations has properties very similar to the Euclidean Dirac problem considered above, and can be solved by analogous methods.
In Minkowski space-time, with metric $`\eta _{\mu \nu }=`$ diag$`(1,+1,+1,+1)`$, we use Dirac matrices with the properties $`\{\gamma _\mu ,\gamma _\nu \}=2\eta _{\mu \nu }`$ and $`\gamma _\mu {}_{}{}^{}=\gamma _0\gamma _\mu \gamma _0`$.
Our starting point is the Dirac equation
$$(iD\text{/}+m+g\phi )\psi =0.$$
$`(6.1)`$
Here $`\phi =\stackrel{~}{g}/r`$ is a dynamical scalar background and $`m`$ the mass (which can be taken as the vacuum expectation value of the scalar field). Eq. (6.1) describes the motion in the long-range field of a Julia-Zee dyon . The covariant derivatives contain electromagnetic background potentials corresponding to a Coulomb field
$$D_0=_0+iq\varphi ,\varphi =\stackrel{~}{q}/r,$$
and the field of a magnetic monopole taken, as in Sec. 2, to be of unit strength.
Multiplication by $`(iD\text{/}+m+g\phi )`$ gives a Klein-Gordon type equation, which on stationary states $`\psi (𝐫,t)=\mathrm{exp}(iEt)\psi _E(𝐫)`$ takes the form
$$\left[(Eq\varphi )^2(\stackrel{}{}iq\stackrel{}{A})^2+(m+g\phi )^2iq\sigma ^{\mu \nu }F_{\mu \nu }ig\gamma _\mu ^\mu \phi \right]\psi _E=\mathrm{\hspace{0.17em}0}.$$
$`(6.2)`$
The generalized Klein-Gordon operator can be block-diagonalized in $`(2\times 2)`$ blocks, with 2-component eigenspinors $`\psi _\pm `$ satisfying
$$\left[(Eq\varphi )^2(\stackrel{}{}iq\stackrel{}{A})^2+(m+g\phi )^2\stackrel{}{\sigma }(q\stackrel{}{B}\pm \stackrel{}{}\mathrm{\Lambda })\right]\psi _{E,\pm }=\mathrm{\hspace{0.17em}0},$$
$`(6.3)`$
with $`\mathrm{\Lambda }=\sqrt{g^2\phi ^2q^2\varphi ^2}`$. Note that the square root is real for $`g^2\phi ^2q^2\varphi ^2`$, and imaginary for $`g^2\phi ^2<q^2\varphi ^2`$. For the case of the Coulomb and scalar potentials this becomes
$$\mathrm{\Lambda }=\frac{\lambda }{r},\lambda =\sqrt{g^2\stackrel{~}{g}^2q^2\stackrel{~}{q}^2}.$$
$`(6.4)`$
Defining the operators $`\stackrel{}{\mathrm{}}`$ and $`\stackrel{}{J}`$ as in Sec. 2, we may cast (6.3) into the form
$$[(_r+\frac{1}{r})^2+\frac{1}{r^2}(\stackrel{}{J}^2\frac{3}{4}\stackrel{}{\mathrm{}}\stackrel{}{\sigma }q^2\lambda \sigma _r)(E\frac{q\stackrel{~}{q}}{r})^2+(m\frac{g\stackrel{~}{g}}{r})^2]\psi _{E,\pm }=0.$$
$`(6.5)`$
To make contact with the Biedenharn operator and the work of Sec. 2, we note that the operators $`y`$ and $`x`$ of (2.8) occur as blocks in (6.5). Writing here $`\mathrm{\Gamma }_+=y`$ and $`\mathrm{\Gamma }_{}=x`$, we have
$$\mathrm{\Gamma }_\pm \left(\mathrm{\Gamma }_\pm +1\right)=\stackrel{}{\mathrm{}}^2+\lambda ^2\left(q\pm \lambda \right)\sigma _r,$$
$`(6.6)`$
and can hence rewrite (6.5) in the standard form
$$\left[\left(_r+\frac{1}{r}\right)^2+\frac{1}{r^2}\mathrm{\Gamma }_\pm \left(\mathrm{\Gamma }_\pm +1\right)+\frac{2k}{r}+\epsilon \right]\psi _{E,\pm }=\mathrm{\hspace{0.17em}0}.$$
$`(6.7)`$
Here the constants $`k`$ and $`\epsilon `$ are given by $`k=q\stackrel{~}{q}Eg\stackrel{~}{g}m`$ and $`\epsilon =m^2E^2`$. Note the symmetry under the simultaneous exchange of $`(E,q\stackrel{~}{q})i(m,g\stackrel{~}{g})`$. The eigenvalues of $`\mathrm{\Gamma }_\pm `$ are the ones given in (2.10): here
$$\gamma =\pm \sqrt{(j+\frac{1}{2}){}_{}{}^{2}+g^2\stackrel{~}{g}^2q^2(\stackrel{~}{q}^2+1)}.$$
$`(6.8)`$
Introducing the notation $`l_\gamma `$, where $`l_\gamma =\gamma `$ for $`\gamma 0`$ and $`l_\gamma =(1+\gamma )`$ for $`\gamma <0`$, we see that the eigenstates of $`\mathrm{\Gamma }_\pm `$ satisfy the equation
$$\left[\left(_r+\frac{1}{r}\right)^2+\frac{1}{r^2}l_\gamma \left(l_\gamma +1\right)+\frac{2k}{r}+\epsilon \right]\psi _{E,l(\gamma )}=\mathrm{\hspace{0.17em}0}.$$
$`(6.9)`$
The spectrum of eigenvalues for bound states is well-known from atomic physics:
$$\epsilon =k^2/n_\gamma ^2,n_\gamma =1+l_\gamma +N,N=0,1,2,\mathrm{}.$$
$`(6.10)`$
However, in this case the bound-state energy eigenvalues themselves then are given by
$$E(j,N)=m\left(g\stackrel{~}{g}q\stackrel{~}{q}\pm n_\gamma \sqrt{n_\gamma ^2+q^2\stackrel{~}{q}^2g^2\stackrel{~}{g}^2}\right).$$
$`(6.11)`$
For the ground state $`j=0`$, $`N=0`$, the wave equation factorizes as expected on the basis of supersymmetric quantum mechanics:
$$\left(_r+\frac{\gamma _g1}{r}+\frac{k_g}{\gamma _g}\right)\left(_r+\frac{\gamma _g+1}{r}+\frac{k_g}{\gamma _g}\right)\psi _0=\mathrm{\hspace{0.17em}0},$$
$`(6.12)`$
with $`\gamma _g`$ and $`k_g`$ the ground state values of $`\gamma `$ and $`k`$.
Acknowledgements. P. A. H. and A. J. M would like to thank NIKHEF for the hospitality extended, while part of this work was completed. Correspondence with Dr. M. Plyushchay is also acknowledged.
References
F. De Jonghe, A. J. Macfarlane, K. Peeters, J.-W. van Holten, Phys. Lett. B 359, 114 (1995).
E. D’Hoker and L. Vinet, Phys. Lett. B 137, 72 (1984).
P. A. M. Dirac, The Principles of Quantum Mechanics, Oxford : Clarendon (1958).
L. C. Biedenharn, Phys. Rev. 126, 845 (1962); L. C. Biedenharn and N. V. V. Swamy, ibid. 5B, 1353 (1964).
M. Berrondo and H. V. McIntosh, Journ. Math. Phys. 11, 125 (1970).
M. Plyushchay, hep-th/0005122, to be published in Phys. Lett. B.
D. Spector, Phys. Lett. 474B331-335 (2000).
E. D’Hoker and L. Vinet, Comm. Math. Phys. 97, 391-427 (1985).
F. Bloore and P. A. Horváthy, Journ. Math. Phys. 33, 1869 (1992).
T. T. Wu and C. N. Yang, Nucl. Phys. B107 365-380 ( 1976).
Y. Kazama, C. N. Yang and A. S. Goldhaber, Phys. Rev. 15D 2287-2299 (1977).
R. Jackiw, Ann. Phys (N.Y.) 129, 183 (1980).
C. Duval and P. A. Horváthy, Journ. Math. Phys. 35, 2516 (1994).
K. S. Nirov and M. Plyushchay, Phys. Lett. B405, 114 (1997).
M. Visinescu, Phys. Lett. B 339, 28 (1994); J.-W. van Holten, ibid. B 342, 47 (1995); A. Comtet and P. A. Horváthy, ibid. B 349, 49 (1995).
L. Gy. Fehér, P. A. Horváthy and L. O’Raifeartaigh, Int. Journ. Mod. Phys. A4, 5277 (1989).
H. V. McIntosh and A. Cisneros, Journ. Math. Phys. 11, 896 (1970); D. Zwanziger, Phys. Rev. 176, 1480 (1968).
E. D’Hoker and L. Vinet, Phys. Rev. Lett. 55, 1043 (1986).
B. Julia and A. Zee, Phys. Rev. D11 2227 (1975).
|
warning/0006/astro-ph0006073.html
|
ar5iv
|
text
|
# Transport phenomena in stochastic magnetic mirrors
## 1 Introduction
The problem of thermal conduction in a stochastic magnetic field is crucial for our understanding of galaxy cluster formation (Suginohara & Ostriker 1998; Cen & Ostriker 1999) and for the theory of cooling flows (Fabian 1990). It is also of great interest for the solar physics and for various questions of plasma physics. At the same time, the question: “whether electron thermal conduction is so strongly inhibited by a stochastic magnetic field in a galaxy cluster, that it can be neglected”, is a very controversial one (Rosner & Tucker 1989; Tribble 1989; Tao 1995; Pistinner & Shaviv 1996; Chandran & Cowley 1998). It is currently estimated that if the coefficient of thermal conductivity is less than $`1/30`$ of the Spitzer value, then the time scale of the heat conduction in the cluster is more than the Hubble time (Suginohara & Ostriker 1998). Otherwise, thermal conduction is important. <sup>1</sup><sup>1</sup>1This numerical estimate, $`1/30`$ of the Spitzer value, is based on numerical simulations with limited resolution, so it is not the last word on the problem.
The problem of thermal diffusion of heat conducting electrons in a stochastic magnetic field should be divided into two separate parts because there are two separate effects that reduce diffusion in the presence of stochastic magnetic field (Pistinner & Shaviv 1996; Chandran, Cowley, & Ivanushkina 1999). The first effect is that the heat conducting electrons have to travel along tangled magnetic field lines, and as a result, they have to go larger distances between hot and cold regions of space. (In other words, the temperature gradients are weaker along magnetic field lines.) The second effect is that electrons, while they are traveling along the field lines, become trapped and detrapped between magnetic mirrors (which are regions of strong magnetic field). A trapped electron is reflected back and forth between magnetic mirrors until collisions make its pitch angle sufficiently small for the electron to escape the magnetic trap.
In this paper we concentrate on the second effect, and we derive the reduction of the effective electron thermal conduction parallel to the magnetic field lines caused by the presence of stochastic magnetic mirrors.
As is well known, a temperature gradient produces electrical current as well as heat flow. Similarly, an electric field produces heat flow as well as current. The four transport coefficients describing this are given in equation (44) and (45). The transport coefficients were first calculated by Spitzer & H$`\ddot{\mathrm{a}}`$rm for an unmagnetized plasma (Spitzer & H$`\ddot{\mathrm{a}}`$rm 1953; Cohen, Spitzer & Routly 1950). Their coefficients also apply in an uniform magnetic field for transport parallel to the field. In this paper, we show how the parallel transport coefficients can be reduced in the presence of stochastic magnetic mirrors, and we calculate their reduced values by the same kinetic approach as that of Spitzer & H$`\ddot{\mathrm{a}}`$rm. The reduction factors are presented in Figure 5. The reduced effective thermal conductivity (that resulting when the electric field is present to cancel the current) is given in Figure 6. Spatial diffusivity of mono-energetic electrons along the magnetic field lines is also presented in Figure 3.
First, in Section 2, we solve the kinetic equation to find the escape time $`\tau _m`$ for electrons trapped between two equal magnetic mirrors. We assume, that all electrons have a single value of speed, $`V`$, i.e. they are mono-energetic. The exact calculations of the escape time are given in Appendices A and B. In addition, we carry out Monte-Carlo particle simulations to confirm our results.
Second, in Section 3, we apply our results for this escape time to find the reduction of diffusion of mono-energetic electrons in a system of stochastic mirrors. It turns out that in the limit $`l_0\lambda `$, where $`l_0`$ is the magnetic field decorrelation length and $`\lambda `$ is the electron mean free path, the parallel diffusivity is unaffected by magnetic mirrors, and is given by the standard value $`D_0=(1/3)V\lambda `$. In the opposite limit, $`l_0\lambda `$, magnetic mirrors do reduce diffusivity. We find that in this case there is a subset of the mirrors, the principle mirrors, that inhibits diffusion the most. These are mirrors whose separation distances are approximately equal to the electron effective mean free path, $`\lambda _{\mathrm{eff}}`$, the typical distance that electrons travel in the loss cones before they are scattered out of them. In order to estimate the reduction of diffusion in this limit, we need consider only the principle mirrors, neglecting all others. Again, we perform the numerical simulations to support these theoretical results.
Third, in Section 4, in order to carry out a precise kinetic treatment involving all electrons, we consider the diffusion reduction to be equivalent to an enhancement of the pitch angle scattering rate of electrons. In deriving the collision integral, we, therefore, modify the pitch angle scattering term by the inverse of the factor by which the spatial diffusion is reduced. We take into account the full perturbed electron-electron collision integral, as well as the electron-proton collision term. We obtain an integro-differential equation for the perturbed electron distribution function in the presence of stochastic magnetic mirrors. If there is no reduction of electron diffusivity, our equation reduces to the well known result obtained by Spitzer and H$`\ddot{\mathrm{a}}`$rm (Spitzer & H$`\ddot{\mathrm{a}}`$rm 1953; Cohen, Spitzer & Routly 1950; Spitzer 1962).
Fourth, in Section 5, we solve our equation numerically, separately for the Lorentz gas in the presence of magnetic mirrors, neglecting electron-electron collisions (in this case the equation simplifies greatly), and for the Spitzer gas in the presence of magnetic mirrors. We find the reductions of the four plasma transport coefficients and of the effective thermal conductivity as functions of the ratio of the magnetic field decorrelation length $`l_0`$ to the electron mean free path at the thermal speed $`V_T=\sqrt{2kT/m_e}`$ (this mean free path is different for the Lorentz and Spitzer models). We find that the major effect of the magnetic mirrors is the reduction of anisotropy of superthermal electrons (this anisotropy is driven by a temperature gradient or/and by an electric field). Electrical current and heat are mainly transported by these electrons, whose diffusivity is suppressed the most.
Finally, we discuss our results and give the conclusions in Section 6.
## 2 Mono-energetic electrons trapped between two equal magnetic mirrors
In this section we solve the kinetic equation to find the escape time $`\tau _m`$ for electrons trapped between two equal magnetic mirrors. We assume here and in the next section that all electrons have a single value of speed, $`V`$, which is unchanged by collisions, i.e. electrons are mono-energetic. In order to derive an analytical solution, we make several additional simplifying assumptions. Let the two magnetic barriers (mirrors) be both equal to $`B_\mathrm{m}`$, and we assume the magnetic field $`B`$ is constant between them. We introduce the mirror strength $`m\stackrel{\mathrm{def}}{=}B_\mathrm{m}/B`$. The separation of the mirrors is $`l_m`$, and their thicknesses are negligible compared to $`l_m`$. In other words, magnetic mirrors are similar to thin step-functions with heights $`B_\mathrm{m}B`$ and with constant field $`B`$ between them (see Figure 1). This is a reasonable assumption, because as we will see in the next section, electron diffusion is controlled by strong mirrors with mirror strengths $`m4`$, which are separated by distances much larger than the magnetic field decorrelation length (if the spectrum of mirrors falls off with their strength significantly faster than $`1/m`$, the case that we consider in this paper).
Under these assumptions, the kinetic equation for the distribution function $`f(t,x,\mu )`$ of mono-energetic electrons trapped between the two mirrors is (Braginskii 1965)
$`{\displaystyle \frac{f}{t}}+\mu V{\displaystyle \frac{f}{x}}={\displaystyle \frac{\nu }{2}}{\displaystyle \frac{}{\mu }}\left[(1\mu ^2){\displaystyle \frac{f}{\mu }}\right].`$ (1)
Here $`x`$ is one-dimensional space coordinate along a magnetic flux tube, $`t`$ is time, $`\mu =\mathrm{cos}\theta `$ is the cosine of the electron’s pitch angle, and $`\nu =V/\lambda `$ is the collision frequency \[$`\lambda `$ is the mean free path, see equations (55) and (58)\]. The right-hand side of equation (1) represents the pitch angle scattering rate, $`\nu `$, of electrons. The electrons are trapped in the region of space between the mirrors, $`l_m/2<x<l_m/2`$, and they can escape through the two windows: $`x=l_m/2`$, $`\mu >\mu _{\mathrm{crit}}=\sqrt{11/m}`$ and $`x=l_m/2`$, $`\mu <\mu _{\mathrm{crit}}`$, as shown in Figure 1. The mirror strength is $`m=B_\mathrm{m}/B`$, and it is the measure of the relative heights of the magnetic barriers. For simplicity, we assume that the barriers are high, i.e. $`m1`$ and $`\mu _{\mathrm{crit}}11/2m`$. In this case the electron distribution is in quasi-static equilibrium,
$`f(t,x,\mu )=e^{t/\tau _m}F(x,\mu ),\tau _m\nu ^1,`$ (2)
and equation (1) reduces to
$`{\displaystyle \frac{F}{\tau _m}}+\mu V{\displaystyle \frac{F}{x}}={\displaystyle \frac{\nu }{2}}{\displaystyle \frac{}{\mu }}\left[(1\mu ^2){\displaystyle \frac{F}{\mu }}\right].`$ (3)
Let us consider an electron traveling in the loss cone $`\mu >\mu _{\mathrm{crit}}=\sqrt{11/m}11/2m`$ (or $`\mu <\mu _{\mathrm{crit}}`$). The effective electron mean free path, which is the typical distance the electron travels before it is scattered by small angle collisions out of the loss cone, is
$`\lambda _{\mathrm{eff}}\stackrel{\mathrm{def}}{=}\lambda /2m\lambda .`$ (4)
In other words, $`\lambda _{\mathrm{eff}}`$ is a decay distance for a flow of electrons traveling in the loss cones. The solution of equation (3) and, therefore, the escape time $`\tau _m`$, depends on the mirror strength $`m`$ and the ratio $`l_m/\lambda `$. There are three limiting cases for which simple approximate solutions exist: (1) $`l_m\lambda _{\mathrm{eff}}=\lambda /2m`$; (2) $`\lambda _{\mathrm{eff}}l_m\lambda ^2/\lambda _{\mathrm{eff}}=2m\lambda `$; and (3) $`\lambda ^2/\lambda _{\mathrm{eff}}l_m`$. We solve equation (3) for case (1) in Appendix A and for cases (2) and (3) in Appendix B, and we obtain the electron escape times
$`\begin{array}{cccc}\tau _m^{(1)}\hfill & =& \nu ^1\mathrm{ln}m,\hfill & l_m\lambda _{\mathrm{eff}},\hfill \\ \tau _m^{(2)}\hfill & =& \nu ^1(l_m/\lambda _{\mathrm{eff}})=\nu ^1(2ml_m/\lambda ),\hfill & \lambda _{\mathrm{eff}}l_m\lambda ^2/\lambda _{\mathrm{eff}},\hfill \\ \tau _m^{(3)}\hfill & =& \nu ^1(3/\pi ^2)(l_m/\lambda )^2,\hfill & \lambda ^2/\lambda _{\mathrm{eff}}l_m.\hfill \end{array}`$ (8)
The following simple physical arguments help to understand these results in these three limiting cases. The collisional scattering is a two-dimensional random walk of a unit vector (which is the direction of the electron velocity) on a surface of a unit-radius sphere with frequency $`\nu `$ (so, the scattered angle $`\mathrm{\Delta }_\mathrm{s}=\sqrt{2\nu t}`$ after time interval $`t`$). The right hand side of the kinetic equation (1) represents a one-dimensional random walk in $`\mu `$-space that follows from the two-dimensional walk because of symmetry. However, it is convenient for the moment to return to the original two-dimensional scattering because it is isotropic. The angular sizes of the two loss cones on the unit-radius sphere are $`\mathrm{\Delta }_{\mathrm{esc}}1/\sqrt{m}`$. First, in the limit $`l_m\lambda _{\mathrm{eff}}`$, collisions are very weak, and the scattered angle over the travel time between mirrors, $`l_m/V`$, is $`\sqrt{l_m/\lambda }\mathrm{\Delta }_{\mathrm{esc}}`$. Therefore, in this case we can disregard the electron motion in $`x`$-space. We divide the surface of the unit-radius sphere into $`m`$ boxes, each of angular size $`\mathrm{\Delta }_{\mathrm{esc}}1/\sqrt{m}`$. The time it takes for the unit vector to random walk from one box to another is $`\nu ^1/m`$, resulting in the total escape time $`\tau _mm\times (\nu ^1/m)=\nu ^1`$. Because the unit vector can “visit” each box more than once, the exact result contains the logarithm of $`m`$. Second, in the limit $`\lambda _{\mathrm{eff}}l_m\lambda ^2/\lambda _{\mathrm{eff}}`$, we have to consider motion in $`x`$-space as well. In this case the electrons move in three-dimensional phase space, and they escape when they are in the two loss cones within distance $`\lambda _{\mathrm{eff}}`$ from the mirrors, as shown by the shaded regions in Figure 1(b). We divide the three-dimensional phase space into $`(l_m/\lambda _{\mathrm{eff}})(1/\mathrm{\Delta }_{\mathrm{esc}}^2)m^2l_m/\lambda `$ boxes, each of size $`\lambda _{\mathrm{eff}}\mathrm{\Delta }_{\mathrm{esc}}^2\lambda /m^2`$. The time it takes to move from one box to another is $`\nu ^1/m`$, resulting in the total escape time $`\tau _m(m^2l_m/\lambda )\times (\nu ^1/m)=\nu ^1(ml_m/\lambda )`$. Note, that the electron distribution function is almost constant in the phase space in this case (see Appendix B). Third, in the limit $`\lambda ^2/\lambda _{\mathrm{eff}}l_m`$, the escape of electrons is controlled by slow diffusion in $`x`$-space, so the escape time is approximately equal to the time of diffusion between mirrors, $`\tau _m\nu ^1(l_m/\lambda )^2`$ in this case.
In our further calculations we use a simple interpolation formula
$`\tau _m\tau _m^{(1)}+\tau _m^{(2)}+\tau _m^{(3)}=\nu ^1\left[\mathrm{ln}m+(l_m/\lambda _{\mathrm{eff}})+(3/\pi ^2)(l_m/\lambda )^2\right]`$ (9)
for the whole range of parameters $`m`$ and $`l_m/\lambda `$. This formula is suggested by the numerical simulations shown in Figure 2. The dots in this figure show the results of our Monte-Carlo particle simulations for three mirror strengths $`m=2`$, $`m=16`$ and $`m=128`$. To obtain these results we followed $`10^3`$$`10^6`$ electrons trapped between two equal magnetic mirrors separated by distance $`l_m`$ ranging from $`1/1024`$ to $`256`$ in units of the mean free path $`\lambda `$. Independently of the initial distribution of electrons, the number of trapped electrons tends to an exponential dependence on time with the characteristic decay time $`\tau _m`$ in just a few collision times \[see equation (2)\]. The solid lines in the figure represent formula (9) and are in a very good agreement with the simulations even for the smallest mirror strength $`m=2`$.
## 3 Diffusion of mono-energetic electrons in a system of random magnetic mirrors
In this section we continue to assume that electrons have a single value of speed, $`V`$. If there were no magnetic mirrors and the magnetic field had constant strength along the field lines, the parallel diffusion of mono-energetic electrons would be the standard spatial diffusion, $`D_0=(1/3)V\lambda `$. Here, $`\lambda `$ is the electron mean free path at speed $`V`$. However, as we have discussed in the introductory section, diffusing electrons move along flux tubes with random magnetic field strength and become trapped and detrapped between magnetic mirrors. These mirrors are regions of strong field and are separated by a field decorrelation length $`l_0`$. As a result, the diffusion is reduced by a factor that depends on the ratio $`l_0/\lambda `$.
In the main part of this section we derive this diffusion analytically and at the end of the section confirm it with numerical simulations. (In contrast to the previous section, where there were only two equal mirrors, in this section, we consider many mirrors with random spacing and strength.)
Consider the limit $`l_0\lambda `$ first. In this case collisions are strong, and according to the third formula in equation (8), the time it takes for electrons to escape a trap between two magnetic mirrors is independent of the mirror strengths and is entirely controlled by the standard spatial diffusion transport of electrons between the mirrors. As a result, magnetic mirrors can be ignored, and there is no reduction of diffusion; $`D=D_0`$.
In the opposite limit, $`l_0\lambda `$, the collisions are weak, and magnetic mirrors do result in a reduction of diffusion. To find this reduction, we divide all mirrors into equal size bins $`b_m=(m\delta /2,m+\delta /2]`$, where $`m`$ is the bin central mirror strength, and constant $`\delta `$ is the width of the bins (the value of $`\delta `$ will be discussed later).
For the moment we consider the diffusion in the presence of only those mirrors that are in a single bin $`b_m`$. It turns out that one of the bins leads to a smaller diffusion than any other bin, and the net diffusion due to all the mirrors is approximately that due to only mirrors in this bin, provided that the bins are sufficiently wide.
Let the spectrum of magnetic mirror strengths be $`𝒫(m)`$. We assume that strong magnetic mirrors are rare, i.e. the spectrum falls off fast with the mirror strength (we will estimate how fast it should fall off, below). The probability that a mirror belongs to bin $`b_m`$ is
$`p_m={\displaystyle _{m\delta /2}^{m+\delta /2}}𝒫(m^{})𝑑m^{}\delta 𝒫(m)+(\delta ^3/24)𝒫^{\prime \prime }(m).`$ (10)
At each decorrelation length $`l_0`$ the magnetic field changes and becomes decorrelated. Therefore, the mean separation of mirrors that are in bin $`b_m`$ is
$`l_m=l_0{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}kp_m(1p_m)^{k1}=l_0/p_m.`$ (11)
Let us consider an electron trapped between two mirrors of bin $`b_m`$. The time $`\tau _m`$ that it takes for this electron to escape the trap is given by equation (9), where we keep only the first two terms (because $`l_0\lambda `$)
$`\tau _m\tau _m^{(1)}+\tau _m^{(2)}=\nu ^1\mathrm{ln}(mq_m).`$ (12)
Here, we introduce the important parameter
$`q_m\stackrel{\mathrm{def}}{=}\mathrm{exp}(l_m/\lambda _{\mathrm{eff}})=\mathrm{exp}(2ml_0/p_m\lambda ),`$ (13)
where the mean distance $`l_m`$ between the two mirrors is given by equation (11). After the electron escapes, it travels freely in the loss cone in one of the two directions along the magnetic field lines until it is again trapped between another two mirrors of bin $`b_m`$. The freely traveling electron becomes first trapped with probabilities $`1e^{l_m/\lambda _{\mathrm{eff}}}=1q_m^1`$ in $`0x<l_m`$, $`e^{l_m/\lambda _{\mathrm{eff}}}e^{2l_m/\lambda _{\mathrm{eff}}}=q_m^1q_m^2`$ in $`l_mx<2l_m`$, $`e^{2l_m/\lambda _{\mathrm{eff}}}e^{3l_m/\lambda _{\mathrm{eff}}}=q_m^2q_m^3`$ in $`2l_mx<3l_m`$, and so on. Therefore, the mean distance squared $`\mathrm{\Delta }x^2_m`$ that the electron travels in the loss cones before trapping is
$`\mathrm{\Delta }x^2_ml_m^2{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}k^2(q_m^{k+1}q_m^k)=l_m^2{\displaystyle \frac{q_m(q_m+1)}{(q_m1)^2}}.`$ (14)
The processes of trapping and detrapping repeat in time intervals $`\tau _m`$. In other words, electrons random walk along the field lines in a system of mirrors that belong to bin $`b_m`$ with steps $`\mathrm{\Delta }x^2_m`$ in time intervals $`\tau _m`$. As a result, the diffusion coefficient for these electrons is $`D(m)=C[\mathrm{\Delta }x^2_m/2\tau _m]`$, where we introduce a scaling constant $`C`$, which is of the order unity and will be determined by the numerical simulations. The corresponding reduction of diffusion is
$`D(m)/D_0=C{\displaystyle \frac{3}{2}}\left({\displaystyle \frac{l_0}{\lambda }}\right)^2{\displaystyle \frac{q_m(q_m+1)}{(q_m1)^2}}{\displaystyle \frac{1}{p_m^2}}{\displaystyle \frac{1}{\mathrm{ln}(mq_m)}},l_0\lambda ,`$ (15)
where we use $`D_0=(1/3)\nu \lambda ^2`$ and equations (11), (12) and (14); $`p_m`$ and $`q_m`$ are given by equations (10) and (13).
For a given spectrum of mirrors $`𝒫(m)`$ and given constants $`\delta `$ and $`C`$, the diffusion reduction (15) due to mirrors of bin $`b_m`$, is a function of mirror strength $`m`$. Let us analyze this function in two limits: $`\mathrm{ln}q_m1`$ and $`\mathrm{ln}q_m\mathrm{ln}m1`$. If $`\mathrm{ln}q_m1`$, then $`q_m1=2ml_0/p_m\lambda 1`$. Therefore, $`D(m)/D_0C(3/4)(1/m^2\mathrm{ln}m)`$ and $`(d/dm)[D(m)/D_0]<0`$. On the other hand, if $`\mathrm{ln}q_m\mathrm{ln}m`$, then $`D(m)/D_0C(3/4)(l_0/\lambda )(1/mp_m)`$. Therefore, $`(d/dm)[D(m)/D_0]>0`$ if the spectrum of mirrors falls off faster than $`1/m`$ with the mirror strength.<sup>2</sup><sup>2</sup>2This criterion is different from the result of Albright et al. (2000), who found $`1/m^2`$ to be the boundary spectrum for the transition between their diffusive and subdiffusion regimes. We believe that the difference arises because, for flat spectra, our bin width $`\delta `$ starts to depend on $`l_0/\lambda `$ (and our simple diffusion model breaks down). In this paper we make an assumption that the spectrum falls off significantly faster than $`1/m`$.
Therefore, a minimum of $`D(m)/D_0`$ exists. Let this minimum be achieved at $`m=m_p`$. Then $`\mathrm{ln}q_{m_p}=l_{m_p}/\lambda _{\mathrm{eff}}2/\mathrm{ln}m_p1`$, or $`l_{m_p}\lambda _{\mathrm{eff}}`$. The minimum can roughly be estimated as $`D(m_p)/D_0=\mathrm{min}\{D(m)/D_0\}1/m_p^2`$, which is in agreement with the qualitative results of Albright et al. (2000).
In other words, if $`l_0\lambda `$, then there is the bin that inhibits diffusion the most. We call it the principle bin, $`b_\mathrm{p}=(m_\mathrm{p}\delta /2,m_\mathrm{p}+\delta /2]`$. The corresponding mirror strength $`m_\mathrm{p}`$ is the principle mirror strength. The minimum of diffusion $`D(m)`$ due to mirrors of bin $`b_m`$ is achieved at the principle strength, $`m=m_p`$. The spacing of mirrors that are in the principle bin is of the order of the effective mean free path for this bin, $`l_{m_p}\lambda _{\mathrm{eff}}=\lambda /2m_p`$. The main idea is that, in order to estimate the net diffusion due to all mirrors, we need consider only magnetic mirrors that are in the principle bin and we can neglect all other bins. Mirrors that are smaller than the principle mirrors “work” poorly in the inhibition of diffusion because they are weak and are separated by distances less than $`\lambda _{\mathrm{eff}}`$ (which is the distance that electrons travel in the loss cones). Mirrors that are larger than the principle mirrors “work” poorly, because they are very rare and are separated by very large distances (provided the mirror spectrum falls off with the mirror strength significantly faster than $`1/m`$). These assumptions are supported by our numerical simulations (see Figure 3).
As a result of these considerations, we can combine our theoretical results for the reduction of diffusion of mono-energetic electrons, $`R_\mathrm{D}=D/D_0`$, into a single formula valid in the two limits for $`l_0/\lambda `$:
$`R_\mathrm{D}=D/D_0=\{\begin{array}{cc}\underset{m}{\mathrm{min}}\{D(m)/D_0\}=D(m_p)/D_0,\hfill & l_0\lambda ,\hfill \\ 1,\hfill & l_0\lambda ,\hfill \end{array}`$ (18)
where $`D(m)/D_0`$ is given by equation (15), and the minimum is achieved at the principle mirror strength $`m=m_p`$ (note that $`\mathrm{ln}q_{m_p}=l_{m_p}/\lambda _{\mathrm{eff}}1`$).
We show the theoretical mono-energetic diffusion reduction (18) by the solid lines in Figure 3 for two mirror spectra: exponential and Gaussian<sup>3</sup><sup>3</sup>3We find the minimum in equation (18) numerically.,
$`\begin{array}{ccc}𝒫(m)\hfill & =& e^{(m2)}\text{ exponential,}\hfill \\ 𝒫(m)\hfill & =& (2/\pi )^{1/2}e^{(m2)^2/2}\text{ Gaussian}.\hfill \end{array}`$ (21)
The results of our Monte-Carlo particle simulations are shown by dots. The constants $`C`$ and $`\delta `$ (shown at the top) are of the order unity, and we adjust them by matching our theoretical results with the results of simulations in case of each of the two spectra ($`C`$ and $`\delta `$ do not depend on $`l_0/\lambda `$). The simulations are based on $`1`$$`6\times 10^5`$ particles. For each particle we choose a distribution of mirrors $`m2`$, which all are separated by the magnetic field decorrelation length $`l_0`$, and are chosen according to the assumed mirror spectrum (21). We follow the particles during $`300`$ collision times $`\nu ^1`$. Then we average the particle displacements squared $`\mathrm{\Delta }x^2`$ at a given time $`t`$ to obtain the diffusion coefficient $`\mathrm{\Delta }x^2/2t`$ given in Figure 3.
Note that the bin width $`\delta `$ is larger for the exponential spectrum than it is for the Gaussian. This is because the later is steeper at large mirror strengths. Figures 4(a) and 4(b) clearly demonstrate the difference. In these figures we plot the natural logarithm of the diffusion reduction (15) caused by mirrors that are in bin $`b_m`$ versus the mirror strength $`m`$ for $`l_0/\lambda =1/16`$ and for the both spectra (21) of mirror strengths. The principle bins are shown by arrows. In case of each spectrum, the reduction has the minimum at the corresponding principle mirror strength $`m_p`$. We see, that the reduction roughly doubles over its minimal value at the boundaries of the principle bin, $`m=m_p+\delta /2`$ and $`m=m_p\delta /2`$.
## 4 The Fokker-Planck kinetic equation
In this section we use the results found above to obtain the modified kinetic equation for electrons traveling in a system of random magnetic mirrors. The reduction of diffusion of mono-energetic electrons with speed $`V`$, $`R_\mathrm{D}`$, obtained in the previous section can be considered to be equivalent to an enhancement of the pitch angle scattering rate, since the pitch angle scattering is directly related to spatial diffusion. We therefore, in deriving the collision integral, modify the pitch angle scattering term by factor $`R_\mathrm{D}^1`$, where $`R_\mathrm{D}`$ is the factor by which the spatial diffusion is reduced (see the previous section). Hereafter, we do not assume electrons to be mono-energetic. We take into account the full perturbed electron-electron collision integral, as well as the electron-proton collision term. When $`R_\mathrm{D}1`$, our equations reduce to those of Spitzer and H$`\ddot{\mathrm{a}}`$rm in their well known paper (Spitzer & H$`\ddot{\mathrm{a}}`$rm 1953; Cohen, Spitzer & Routly 1950; Spitzer 1962).
The electron distribution function is
$`f(\mu ,V)=f_0(V)+f_1(\mu ,V),`$ (22)
where $`f_0`$ is the zero order isotropic part given by the Maxwellian distribution,
$`f_0=n(x)\left[m_e/2\pi kT(x)\right]^{3/2}e^{m_eV^2/2kT(x)}=n\pi ^{3/2}V_\mathrm{T}^3e^{\upsilon ^2},`$ (23)
and $`f_1\mu `$ is the first order anisotropic perturbation (of order the temperature gradient and electric field)
$`f_1(\mu ,V)=\mu nV_\mathrm{T}^3S(\upsilon ).`$ (24)
Here $`m_e`$ is the electron mass, $`k`$ is the Boltzmann constant, and the electron temperature $`T(x)`$ and concentration $`n(x)`$ slowly change in space. We also introduce the dimensionless electron speed $`\upsilon =V/V_T`$, where the thermal electron speed is $`V_T=\sqrt{2kT/m_e}`$. Thus, the function $`S(\upsilon )`$ in equation (24) is dimensionless.
In a steady state, the kinetic equation for the electrons is obviously
$`V_x(f_0/x)(eE/m_e)(f_0/V_x)=(\delta f/\delta t)_c,`$ (25)
where $`(\delta f/\delta t)_c`$ is the Coulomb collision integral that includes electron-proton and electron-electron collisions, $`V_x=\mu V`$ is the $`x`$-component of the electron velocity (the component along the magnetic field lines), and $`E`$ is the electric field in the $`x`$-direction. The electron pressure should be constant, $`P=kn(x)T(x)=\mathrm{const}`$<sup>4</sup><sup>4</sup>4Because the hydrodynamic time scale is much shorter than the transport, e.g thermal conduction, time scale. As a result, the derivatives of the Maxwellian electron distribution are
$`f_0/x=(\upsilon ^22.5)(f_0/T)(dT/dx),f_0/V_x=(2\mu /V_\mathrm{T})\upsilon f_0.`$ (26)
The collision integral is divided up as
$`(\delta f/\delta t)_c=(\delta f_0/\delta t)_0+(\delta f_1/\delta t)_0+(\delta f_0/\delta t)_1=(\delta f_1/\delta t)_0+(\delta f_0/\delta t)_1,`$ (27)
where $`(\delta f_0/\delta t)_00`$ corresponds to Maxwellian collisions acting on $`f_0`$, $`(\delta f_1/\delta t)_0`$ corresponds to Maxwellian collisions (with enhanced pitch angle scattering) acting on $`f_1`$, and $`(\delta f_0/\delta t)_1`$ corresponds to perturbed collisions acting on $`f_0`$ (since $`f_0`$ is isotropic, there is no pitch angle scattering in this collision term). The collision integral (27) can be best obtained, in the Fokker-Planck form, by using the Rosenbluth potentials $`h(\mu ,V)=h_0(V)+h_1(\mu ,V)`$ and $`g(\mu ,V)=g_0(V)+g_1(\mu ,V)`$ (Rosenbluth, MacDonald, & Judd 1957). Here $`h_0`$ and $`g_0`$ are calculated using the Maxwellian parts of the electron and ion distribution functions (23), while the perturbed potentials, $`h_1=2\mu 𝒜_1(V)`$ and $`g_1=\mu _1(V)`$, are proportional to $`\mu `$, and they are calculated using the perturbed part of the electron distribution function (24).
The Maxwellian potentials $`h_0`$ and $`g_0`$ determine the $`(\delta f_1/\delta t)_0`$ part of the Fokker-Planck collision integral, and the perturbed potentials, $`h_1=2\mu 𝒜_1(V)`$ and $`g_1=\mu _1(V)`$, are used to find the $`(\delta f_0/\delta t)_1`$ part of the Fokker-Planck collision integral \[see the equation (31) of Rosenbluth, MacDonald, & Judd 1957\]
$`(\delta f_1/\delta t)_0`$ $`=`$ $`{\displaystyle \frac{A_D}{2n}}\{{\displaystyle \frac{1}{V^2}}{\displaystyle \frac{}{V}}\left[f_1V^2{\displaystyle \frac{dh_0}{dV}}\right]+{\displaystyle \frac{1}{2V^2}}{\displaystyle \frac{^2}{V^2}}\left[f_1V^2{\displaystyle \frac{d^2g_0}{dV^2}}\right]`$ (28)
$``$ $`{\displaystyle \frac{1}{V^2}}{\displaystyle \frac{}{V}}\left[f_1{\displaystyle \frac{dg_0}{dV}}\right]+R_\mathrm{D}^1{\displaystyle \frac{1}{2V^3}}{\displaystyle \frac{dg_0}{dV}}{\displaystyle \frac{}{\mu }}\left[(1\mu ^2){\displaystyle \frac{f_1}{\mu }}\right]\},`$
$`(\delta f_0/\delta t)_1`$ $`=`$ $`\mu {\displaystyle \frac{A_D}{2n}}\{{\displaystyle \frac{2}{V^2}}{\displaystyle \frac{d}{dV}}\left[f_0V^2{\displaystyle \frac{d𝒜_1}{dV}}\right]+{\displaystyle \frac{4}{V^2}}f_0𝒜_1+{\displaystyle \frac{1}{2V^2}}{\displaystyle \frac{d^2}{dV^2}}\left[f_0V^2{\displaystyle \frac{d^2_1}{dV^2}}\right]`$ (29)
$``$ $`{\displaystyle \frac{3}{V^3}}f_0{\displaystyle \frac{d_1}{dV}}+{\displaystyle \frac{3}{V^4}}f_0_1{\displaystyle \frac{3}{V^2}}{\displaystyle \frac{d}{dV}}\left[f_0{\displaystyle \frac{d_1}{dV}}\right]+{\displaystyle \frac{3}{V^2}}{\displaystyle \frac{d}{dV}}\left[f_0{\displaystyle \frac{_1}{V}}\right]\}.`$
For a hydrogen plasma the “diffusion constant” $`A_D`$ is
$`A_D=8\pi ne^4\mathrm{ln}\mathrm{\Lambda }/m_e^2,`$ (30)
where $`e`$ is the absolute value of the electron charge, and $`\mathrm{ln}\mathrm{\Lambda }`$ is the Coulomb logarithm (Spitzer 1962). Note, that the last term in equation (28) is the pitch angle scattering term, and we multiply it by our enhancement factor $`R_\mathrm{D}^1`$ \[compare this term with the right-hand side of equation (1)\].
Using the equations (17) and (18) of Rosenbluth, MacDonald, & Judd (1957), we express the derivatives of the potentials $`h_0`$ and $`g_0`$ in terms of the three Maxwellian diffusion coefficients $`\mathrm{\Delta }V_{}_0`$, $`(\mathrm{\Delta }V_{})^2_0`$ and $`(\mathrm{\Delta }V_{})^2_0`$, which are further given in terms of error functions \[see the equations (5-15)–(5-20) of Spitzer 1962\]
$`\begin{array}{ccc}dh_0/dV\hfill & =& (2n/A_D)\mathrm{\Delta }V_{}_0=(n/V^2)[1+4\upsilon ^2G(\upsilon )],\hfill \\ dg_0/dV\hfill & =& (n/A_D)V(\mathrm{\Delta }V_{})^2_0=n[1+\mathrm{\Phi }(\upsilon )G(\upsilon )],\hfill \\ d^2g_0/dV^2\hfill & =& (2n/A_D)(\mathrm{\Delta }V_{})^2_0=(2n/V)G(\upsilon ).\hfill \end{array}`$ (34)
Here $`\mathrm{\Phi }`$ is the usual error function, and $`G`$ is expressed in terms of $`\mathrm{\Phi }`$ and its derivative $`\mathrm{\Phi }^{}`$, they are functions of the dimensionless speed $`\upsilon =V/V_T`$ \[$`V_T=\sqrt{2kT/m_e}`$\],
$`\mathrm{\Phi }(\upsilon )=(2/\sqrt{\pi }){\displaystyle _0^\upsilon }e^{x^2}𝑑x,G(\upsilon )={\displaystyle \frac{\mathrm{\Phi }(\upsilon )\upsilon \mathrm{\Phi }^{}(\upsilon )}{2\upsilon ^2}}.`$ (35)
The perturbed potentials, $`h_1=2\mu 𝒜_1(V)`$ and $`g_1=\mu _1(V)`$, are calculated using the perturbed electron distribution function (24) and are given by the following formulas \[see the equations (40), (41), (45) and (46) of Rosenbluth, MacDonald, & Judd 1957\]
$`𝒜_1`$ $`=`$ $`(4\pi /3)(n/V_\mathrm{T})\left[\upsilon ^2\overline{I}_3(S;\upsilon )+\upsilon \underset{¯}{I}_{\mathrm{\hspace{0.17em}0}}(S;\upsilon )\right],`$
$`_1`$ $`=`$ $`(4\pi /3)nV_\mathrm{T}\left[0.2\upsilon ^2\overline{I}_5(S;\upsilon )\overline{I}_3(S;\upsilon )\upsilon \underset{¯}{I}_{\mathrm{\hspace{0.17em}2}}(S;\upsilon )+0.2\upsilon ^3\underset{¯}{I}_{\mathrm{\hspace{0.17em}0}}(S;\upsilon )\right],`$ (36)
where we introduce integrals
$`\overline{I}_m(S;\upsilon )={\displaystyle _0^\upsilon }\upsilon ^mS(\upsilon )𝑑\upsilon ,\underset{¯}{I}_m(S;\upsilon )={\displaystyle _\upsilon ^{\mathrm{}}}\upsilon ^mS(\upsilon )𝑑\upsilon ,`$ (37)
Now, substituting equations (23), (24), (34) and (36) into formulas (28) and (29), and using definitions (35), (37) and equation (27), after considerable algebra, we have for the collision integrals
$`(\delta f_1/\delta t)_0`$ $`=`$ $`(nA_D/2V_\mathrm{T}^6)\mu \upsilon ^2(\widehat{}S2\upsilon ^2\mathrm{\Phi }^{}S),`$
$`(\delta f_0/\delta t)_1`$ $`=`$ $`(nA_D/2V_\mathrm{T}^6)\mu \upsilon ^2(\widehat{}S+2\upsilon ^2\mathrm{\Phi }^{}S),`$
$`(\delta f/\delta t)_c`$ $`=`$ $`(nA_D/2V_\mathrm{T}^6)\mu \upsilon ^2(\widehat{}S+\widehat{}S),`$ (38)
where the differential and the integral operators are defined as
$`\widehat{}S(\upsilon )`$ $`=`$ $`d/d\upsilon \left[\upsilon G(dS/d\upsilon )\right]+2\upsilon ^2G(dS/d\upsilon )[\upsilon ^1R_\mathrm{D}^1(1+\mathrm{\Phi }G)4\upsilon ^2\mathrm{\Phi }^{}]S,`$ (39)
$`\widehat{}S(\upsilon )`$ $`=`$ $`(4/15\sqrt{\pi })e^{\upsilon ^2}\left[12\overline{I}_5(S;\upsilon )10\overline{I}_3(S;\upsilon )+2\upsilon ^3(6\upsilon ^25)\underset{¯}{I}_{\mathrm{\hspace{0.17em}0}}(S;\upsilon )\right].`$ (40)
The enhancement of the Maxwellian pitch angle scattering rate, $`R_\mathrm{D}^1`$, enters into the differential operator (39). $`R_\mathrm{D}`$ depends on the dimensionless speed $`\upsilon =V/V_T`$, we will explicitly give this dependence in equations (54) and (57).
Finally, substituting formulas (38) and (26) into equation (25), we obtain the kinetic equation for the dimensionless perturbed electron distribution function $`S(\upsilon )`$ \[see equation (24)\]
$`\widehat{}S`$ $`=`$ $`\gamma _\text{T}\upsilon ^3(2\upsilon ^25)e^{\upsilon ^2}+\gamma _\text{E}\upsilon ^3e^{\upsilon ^2}\widehat{}S,`$ (41)
$`S(\upsilon )`$ $``$ $`0,\text{as }\upsilon 0\text{ and as }\upsilon \mathrm{},`$ (42)
where constants $`\gamma _\text{T}`$ and $`\gamma _\text{E}`$ are
$`\gamma _\text{T}={\displaystyle \frac{k^2T}{2\pi ^{5/2}ne^4\mathrm{ln}\mathrm{\Lambda }}}{\displaystyle \frac{dT}{dx}},\gamma _\text{E}={\displaystyle \frac{kT}{\pi ^{5/2}ne^3\mathrm{ln}\mathrm{\Lambda }}}E.`$ (43)
We also take the obvious boundary conditions (42) for function $`S`$. Equations (39)–(41) reduce to the Spitzer equations for an ionized hydrogen gas (Spitzer & H$`\ddot{\mathrm{a}}`$rm 1953; Cohen, Spitzer & Routly 1950) if we set $`R_\mathrm{D}1`$ and make a substitution $`S(\upsilon )=\pi ^{3/2}e^{\upsilon ^2}D(\upsilon )`$. However, we prefer to use function $`S`$, because of the simpler boundary conditions (42).
## 5 The reduction of transport coefficients by stochastic magnetic mirrors
In a steady state, an electric field $`E`$ and a temperature gradient $`dT/dx`$ both produce anisotropic perturbations of the electron distribution function, $`f_1(\mu ,\upsilon )=\mu nV_\mathrm{T}^3S(\upsilon )`$, see equations (22) and (24). This anisotropy results in an electron flow and, consequently, in an electric current $`j`$ and in a heat flow $`Q`$ along magnetic field lines (in the $`x`$-direction)
$`j`$ $`=`$ $`e{\displaystyle _0^{\mathrm{}}}{\displaystyle _1^1}\mu Vf_1𝑑\mu \mathrm{\hspace{0.17em}2}\pi V^2𝑑V=\sigma E+\alpha (dT/dx),`$ (44)
$`Q`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle _1^1}\mu V(m_eV^2/2)f_1𝑑\mu \mathrm{\hspace{0.17em}2}\pi V^2𝑑V=\beta E\kappa (dT/dx).`$ (45)
Here $`\sigma `$, $`\alpha `$, $`\beta `$ and $`\kappa `$ are the four transport coefficients to be found ($`\sigma `$ and $`\kappa `$ are the electrical and thermal conductivities).
Before we proceed to the calculation of the transport coefficients, let us first call attention to the electron flow produced by the electric field. The electric field produces two different kinds of the electron flow. The first, the main, flow is due to acceleration of electrons, which is described by the term containing $`E`$ in equation (25), and correspondingly by the term containing $`\gamma _\text{E}`$ in equation (41). The second, an additional, flow arises because the electric field changes the size of the two loss cones of a mirror trap, so in Figure 1(b) $`\mu _{\mathrm{crit}}`$ in the right upper corner is not equal to $`\mu _{\mathrm{crit}}`$ in the left lower corner. As a result, the electrons are more likely to escape the trap in the direction opposite to the electric field. Fortunately, this additional flow, which is rather complicated to find precisely, can be neglected compared to the flow due to acceleration. We give a prove of this in Appendix C<sup>5</sup><sup>5</sup>5The main reason is that the difference in the two loss cones due to electric field is inversely proportional to the electron kinetic energy, so the additional flow has a factor $`1/V^2`$ compared to a factor $`1/V_\mathrm{T}^2`$ that enters the main flow due to acceleration. Because both the current and the heat flow are mainly transported by superthermal electrons $`\upsilon =V/V_\mathrm{T}2`$, the additional flow is approximately $`20\%`$ of the main flow, see Appendix C.
In further calculations, it is convenient to break $`S(\upsilon )`$ into the two separate inhomogeneous solutions of equation (41), which we denote as $`S_T(\upsilon )`$ and $`S_E(\upsilon )`$<sup>6</sup><sup>6</sup>6The two homogeneous solutions of equation (41) must be excluded, because they diverge either at $`\upsilon 0`$ or at $`\upsilon \mathrm{}`$, violating the conditions (42), see more details in Cohen, Spitzer & Routly 1950. The first solution, $`S_T`$, is obtained by setting $`\gamma _\text{T}=1`$ and $`\gamma _\text{E}=0`$, and the second solution, $`S_E`$, is obtained by setting $`\gamma _\text{T}=0`$ and $`\gamma _\text{E}=1`$, i. e.
$`\begin{array}{c}S_T(\upsilon )=S(\upsilon ),\text{when }\gamma _\text{T}=1\text{ and }\gamma _\text{E}=0,\hfill \\ S_E(\upsilon )=S(\upsilon ),\text{when }\gamma _\text{T}=0\text{ and }\gamma _\text{E}=1.\hfill \end{array}`$ (48)
The general solution to equation (41) and the perturbed distribution function (24) are the linear combinations of the two inhomogeneous solutions,
$`S(\upsilon )`$ $`=`$ $`\gamma _\text{T}S_T(\upsilon )+\gamma _\text{E}S_E(\upsilon ),`$
$`f_1(\mu ,\upsilon )`$ $`=`$ $`\mu nV_\mathrm{T}^3\left[\gamma _\text{T}S_T(\upsilon )+\gamma _\text{E}S_E(\upsilon )\right].`$ (49)
In other words, $`S_T`$ and $`S_E`$ correspond to anisotropic perturbations of the electron distribution function, which are driven by the temperature gradient and by the electric field respectively, while $`S=\gamma _\text{T}S_T+\gamma _\text{E}S_E`$ is the total anisotropic perturbation.
We now consider separately two cases: first, the Lorentz gas in a system of random mirrors, and second, the Spitzer gas in a system of random mirrors. For the Lorentz gas, electrons are assumed only to collide with protons, so equations (39)–(41) become greatly simplified. For the Spitzer gas, we consider both the electron-electron the electron-proton collisions, so we solve the full set of our equations.
### 5.1 Lorentz gas in a system of random mirrors
Here we assume the electrons to collide only with protons, so we have for operators (39) and (40)
$`\widehat{}S=S/\upsilon R_\mathrm{D},\widehat{}S=0,`$ (50)
resulting in the two simple inhomogeneous solutions (48) of equation (41),
$`S_T(\upsilon )=\upsilon ^4(2\upsilon ^25)e^{\upsilon ^2}R_\mathrm{D},S_E(\upsilon )=\upsilon ^4e^{\upsilon ^2}R_\mathrm{D}.`$ (51)
If there are no magnetic mirrors, so $`R_\mathrm{D}1`$, we substitute equations (51) into formula (49) and easily carry out the two integrals in equations (44) and (45). Taking into consideration definitions (43), we obtain the well-known Lorentz transport coefficients (Spitzer 1962)
$`\sigma _\text{L}=2\left({\displaystyle \frac{2}{\pi }}\right)^{3/2}{\displaystyle \frac{(kT)^{3/2}}{m_e^{1/2}e^2\mathrm{ln}\mathrm{\Lambda }}},\alpha _\text{L}=3\left({\displaystyle \frac{2}{\pi }}\right)^{3/2}{\displaystyle \frac{k(kT)^{3/2}}{m_e^{1/2}e^3\mathrm{ln}\mathrm{\Lambda }}},`$
$`\beta _\text{L}=8\left({\displaystyle \frac{2}{\pi }}\right)^{3/2}{\displaystyle \frac{(kT)^{5/2}}{m_e^{1/2}e^3\mathrm{ln}\mathrm{\Lambda }}},\kappa _\text{L}=20\left({\displaystyle \frac{2}{\pi }}\right)^{3/2}{\displaystyle \frac{k(kT)^{5/2}}{m_e^{1/2}e^4\mathrm{ln}\mathrm{\Lambda }}}.`$ (52)
If there are magnetic mirrors, it is convenient to normalize the resulting transport coefficients to the corresponding Lorentz coefficients (52). Substituting equation (49) into the two integrals in equations (44) and (45), and again using definitions (43), we have
$`\sigma /\sigma _\text{L}=(1/3)\overline{I}_3(S_E;\mathrm{}),\alpha /\alpha _\text{L}=(1/9)\overline{I}_3(S_T;\mathrm{}),`$
$`\beta /\beta _\text{L}=(1/12)\overline{I}_5(S_E;\mathrm{}),\kappa /\kappa _\text{L}=(1/60)\overline{I}_5(S_T;\mathrm{}),`$ (53)
where the integral moments are defined by equations (37), and $`S_T`$ and $`S_E`$ are given by equations (51).
In order to find explicitly the diffusion reduction factor $`R_\mathrm{D}`$ in equations (51) as a function of $`\upsilon `$, we refer to the results of Section 3. In those section we found the diffusion reduction as a function of the ratio of the magnetic field decorrelation length $`l_0`$ to the electron mean free path $`\lambda `$. For Lorentz electrons the mean free path $`\lambda _\mathrm{L}`$ is proportional to the fourth power of the electron speed, $`\lambda _\mathrm{L}V^4`$, (Spitzer 1962, Braginskii 1965). Thus, we have
$`R_\mathrm{D}=R_\mathrm{D}(l_0/\lambda _\mathrm{L})=R_\mathrm{D}(\upsilon ^4l_0/\lambda _{\mathrm{L},T}),`$ (54)
where $`\lambda _{\mathrm{L},T}`$ is obviously the Lorentz electron mean free path at the thermal speed $`V_T=\sqrt{2kT/m_e}`$
$`\lambda _{\mathrm{L},T}=(kT)^2/\pi ne^4\mathrm{ln}\mathrm{\Lambda }0.1\mathrm{Kpc}(T/10^7\mathrm{K})^2(10^3\mathrm{cm}^3/n).`$ (55)
Here we assume the Coulomb logarithm for a cluster of galaxy to be $`\mathrm{ln}\mathrm{\Lambda }40`$ (Suginohara & Ostriker 1998).
We use our theoretical results given by equation (18) for the mono-energetic diffusion reduction $`R_\mathrm{D}=R_\mathrm{D}(l_0/\lambda _\mathrm{L})=R_\mathrm{D}(\upsilon ^4l_0/\lambda _{\mathrm{L},T})`$ in the limits $`\upsilon ^4l_0/\lambda _{\mathrm{L},T}1`$ and $`\upsilon ^4l_0/\lambda _{\mathrm{L},T}1`$; and we use our numerical simulation results presented in Figure 3 for $`\upsilon ^4l_0/\lambda _{\mathrm{L},T}1`$. \[We carry out the cubic spline interpolation of the simulation results. Note, that $`R_\mathrm{D}`$ is not differentiated in operator (39), so our final results are not sensitive to small noise errors in calculation of $`R_\mathrm{D}`$.\]
Using equations (51) and (54) with $`R_\mathrm{D}`$ given in Section 3, and numerically performing the velocity integrals, we find all four transport coefficients (53) normalized to the standard Lorentz coefficients (52). The dashed lines in Figures 5(a)–(h) show the resulting normalized transport coefficients $`\sigma `$, $`\alpha `$, $`\beta `$ and $`\kappa `$ as functions of $`l_0/\lambda _{\mathrm{L},T}`$ for the two mirror spectra: (a) exponential, and (b) Gaussian \[see equations (21)\]. The asymptotic values of the coefficients at large values of $`l_0/\lambda _{\mathrm{L},T}`$ are given by the numbers on the dashed lines, and they are unity. Thus, there is no reductions of the transport coefficients at $`l_0/\lambda _{\mathrm{L},T}1`$, as one can expect because there is no reduction of electron diffusivity in this limit \[see equation (18)\].
In a steady state, the electrical current $`j`$ in a highly ionized plasma should be zero. Thus, if a temperature gradient is present, the resulting electric field $`E`$ is obtained by setting $`j`$ to zero in equation (44). Substituting this result for $`E`$ into equation for the heat flow (45), we find for the effective thermal conductivity
$`\kappa _{\mathrm{eff}}`$ $`=`$ $`\kappa \alpha \beta /\sigma ,`$
$`\kappa _{\mathrm{eff}}/\kappa _\text{L}`$ $`=`$ $`\kappa /\kappa _\text{L}(3/5)(\alpha /\alpha _\text{L})(\beta /\beta _\text{L})(\sigma _\text{L}/\sigma ),`$ (56)
where we use formulas (52) for the Lorentz transport coefficients in the second line of this equation.
Using the transport coefficients reported in Figures 5 by dashed lines and formula (56), it is easy to find the effective thermal conductivity $`\kappa _{\mathrm{eff}}`$ normalized to the standard Lorentz thermal conductivity $`\kappa _\text{L}`$ \[see equation (52)\]. However, it is more useful to give the ratio of $`\kappa _{\mathrm{eff}}`$ to the Lorentz effective conductivity, $`\kappa _{\text{L},\mathrm{eff}}=0.4\kappa _\text{L}`$. This ratio is the actual suppression of the effective conductivity of the Lorentz gas by magnetic mirrors. The dashed lines in Figures 6 show this suppression, $`\kappa _{\mathrm{eff}}/\kappa _{\text{L},\mathrm{eff}}`$, as functions of $`l_0/\lambda _{\mathrm{L},T}`$ for the two mirror spectra: (a) exponential, and (b) Gaussian \[see equations (21)\]. It has been estimated that the time of heat conduction in clusters of galaxies is possibly larger than the Hubble time if the thermal conductivity is less than $`1/30`$ of the Spitzer value (Suginohara & Ostriker 1998). The horizontal dotted lines indicate this reduction of $`1/30`$.
For comparison, the dotted lines represent the mono-energetic diffusion reduction at the electron thermal speed, $`R_\mathrm{D}(l_0/\lambda _{\mathrm{L},T})=D(l_0/\lambda _{\mathrm{L},T})/D_0`$. We see, that the Lorentz gas effective conductivity is reduced to a value two to three times smaller than that of the diffusion reduction. This is because heat is mainly transported by superthermal electrons. These electrons have long mean free paths, and the magnetic mirrors more strongly inhibit their diffusion.
### 5.2 Spitzer gas in a system of random mirrors
Now consider the full collision integral (38) for the Spitzer gas in a system of random magnetic mirrors. We have numerically solved the full set of our equations (39)–(42). The formulas (53) and (56) remain the same as for the Lorentz gas, but the functions $`S_T(\upsilon )`$ and $`S_E(\upsilon )`$ are different. For Spitzer electrons the mean free path is $`\lambda _\mathrm{S}V^4[1+\mathrm{\Phi }(\upsilon )G(\upsilon )]^1`$ \[Spitzer 1962, the error functions $`\mathrm{\Phi }`$ and $`G`$ are given by (35)\]. Thus, formula (54) for the reduction of spatial diffusivity now becomes
$`R_\mathrm{D}=R_\mathrm{D}\left(l_0/\lambda _\mathrm{S}\right)=R_\mathrm{D}\left(\upsilon ^4{\displaystyle \frac{l_0}{\lambda _{\mathrm{S},T}}}{\displaystyle \frac{1+\mathrm{\Phi }(\upsilon )G(\upsilon )}{1+\mathrm{\Phi }(1)G(1)}}\right),`$ (57)
where the Spitzer electron mean free at the thermal speed $`V_T=\sqrt{2kT/m_e}`$ is
$`\lambda _{\mathrm{S},T}=0.614(kT)^2/\pi ne^4\mathrm{ln}\mathrm{\Lambda }0.06\mathrm{Kpc}(T/10^7\mathrm{K})^2(10^3\mathrm{cm}^3/n).`$ (58)
Functions $`S_T(\upsilon )`$ and $`S_E(\upsilon )`$ are defined by equations (48), and they are the two inhomogeneous solutions of the equations (41), (42). To find these solutions we solved equation (41) numerically by iterations. At each iteration step the integral part of this equation, $`\widehat{}S`$, was calculated using the solution for $`S`$ from the previous step, and the new solution for $`S`$ was calculated by the Gaussian decomposition with backsubstitution (Fedorenko 1994), using the boundary conditions (42). Initially, we started with zero function $`S=0`$. The iterations converged very rapidly, and the Gaussian decomposition method is stable. We believe that our numerical method is much better and faster than the method of Spitzer and H$`\ddot{\mathrm{a}}`$rm (1953) because their method was not stable. It took us less than ten seconds of computer time to calculate all digits of the transport coefficients reported by Spitzer and H$`\ddot{\mathrm{a}}`$rm.
The solid lines in Figures 5(a)–(h) show the resulting transport coefficients $`\sigma `$, $`\alpha `$, $`\beta `$ and $`\kappa `$ normalized to the standard Lorentz coefficients (52) as functions of $`l_0/\lambda _{\mathrm{S},T}`$ for the two mirror spectra: (a) exponential, and (b) Gaussian \[see equations (21); remember that $`l_0`$ is the magnetic field decorrelation length\]. The asymptotic values of the coefficients at large values of $`l_0/\lambda _{\mathrm{S},T}`$ are given by the numbers on the solid lines, and they agree with the results of Spitzer and H$`\ddot{\mathrm{a}}`$rm.
The effective thermal conductivity, $`\kappa _{\mathrm{eff}}`$, normalized to the Spitzer effective conductivity, $`\kappa _{\text{S},\mathrm{eff}}=0.0943\kappa _\text{L}`$, is given in Figures 6 by the solid lines for the two mirror spectra. This normalized conductivity is the actual suppression of the effective thermal conductivity of the Spitzer gas by stochastic magnetic mirrors. It is the result that should be applied in astrophysical problems with random magnetic mirrors.
Finally, it is interesting to see how the mirrors change the Spitzer perturbed electron distribution function. In Figure 7 we plot functions $`\upsilon ^2S_T(\upsilon )`$ and $`\upsilon ^2S_E(\upsilon )`$ for the case when $`l_0=\lambda _{\mathrm{S},T}`$ \[note that $`2\pi V^2f_1`$ is the actual distribution of electrons over speed $`V=\upsilon V_T`$, see equation (49)\]. The solid lines represent these functions for the Spitzer gas in a system of random mirrors with the exponential mirror spectrum (for the Gaussian spectrum the results are similar). The dashed lines show the same functions for the Spitzer gas without magnetic mirrors. We see that $`\upsilon ^2S_T(\upsilon )`$ and $`\upsilon ^2S_E(\upsilon )`$ are reduced at large values of $`\upsilon `$, i.e. magnetic mirrors reduce the anisotropy of the superthermal electrons, which carry the electrical current and heat.
## 6 Conclusions
In this paper we have derived the actual parallel effective thermal conductivity that should be applied to astrophysical systems with random magnetic mirrors, as well as other important transport coefficients.
Now, let us apply our results for the reduction of the Spitzer effective electron thermal conductivity, shown in Figure 6 by the solid lines, to the galaxy cluster formation problem. If the reduction is by more than a factor of thirty (shown by the horizontal dotted lines in Figure 6), then the time of heat transport becomes larger than the Hubble time, and the heat conduction can be neglected (Suginohara & Ostriker 1998). <sup>7</sup><sup>7</sup>7See the footnote on page 1. We see that this is the case if the magnetic field decorrelation length $`l_0`$ is roughly less than $`10^4`$$`\mathrm{\hspace{0.17em}10}^2`$ of the electron mean free path at the thermal speed $`\lambda _T=\lambda (\sqrt{2kT/m_e})`$ (we consider the Spitzer gas). Although there is little observational data about the topology of magnetic fields in clusters of galaxy, the magnetic field scale is probably $`1`$$`\mathrm{\hspace{0.17em}10}\mathrm{Kpc}`$ (Kronberg 1994; Eilek 1999 and references in it). According to equation (58) the characteristic electron mean free path at the thermal speed is $`0.06`$$`\mathrm{\hspace{0.17em}60}\mathrm{Kpc}`$ for temperatures $`T=10^7`$$`\mathrm{\hspace{0.17em}10}^8\mathrm{K}`$ and densities $`n=10^4`$$`\mathrm{\hspace{0.17em}10}^3\mathrm{cm}^3`$. We see, that in general, the effective electron thermal conductivity parallel to the magnetic field lines is not reduced enough by magnetic mirrors to be completely neglected. However, as we pointed out in the introductory section, there is an additional effect that electrons have to travel along tangled magnetic field lines larger distances from hot to cold regions of space, so the thermal conduction is further reduced (this effect will be considered in our future paper). At the moment, “whether electron thermal conductivity in clusters of galaxies is sufficiently inhibited that it can be ignored” is still an open question.
Recently, Cowley, Chandran et al. studied the reduction of the parallel thermal conduction, and they concluded that the thermal conductivity in galaxy clusters is reduced enough to be neglected (Chandran & Cowley 1998; Chandran, Cowley & Ivanushkina 1999; Albright et al. 2000). Their conclusions are different from ours. The reason is that our approach in calculation of the conductivity is very different, and our results are qualitatively different. The main difference is that they took the reduction of thermal conductivity to be equal to the reduction of diffusivity of thermal electrons. In fact, the reduction of diffusivity is due to the enhanced pitch angle scattering by stochastic magnetic mirrors, and to find the reduction of thermal conductivity, the full set of kinetic equations must be derived and solved. This consistent way of solving the problem makes a considerable difference (see Figure 6). On the other hand, Cowley, Chandran and et al. first called attention to the importance of the effective mean free path $`\lambda _{\mathrm{eff}}`$ and found the correct qualitative result, that in the limit $`l_0\lambda `$ the diffusion reduction is controlled by the mirrors whose spacing is of order of the effective mean free.
###### Acknowledgements.
We are happy to acknowledge many useful discussions of this problem with Jeremiah Ostriker, Jeremy Goodman and David Spergel. We would also like to thank Makoto Matsumoto, Takuji Nishimura and Shawn J. Cokus providing us with fast random number generators (which are given at http://www.math.keio.ac.jp/matumoto/emt.html). This work was partially supported by DOE under Contract No. DE-AC 02-76-CHO-3073. Leonid Malyshkin would also like to thank the Department of Astrophysical Sciences at Princeton University for financial support.
## Appendix A Solution of equation (3) in the limit $`𝒍_𝒎\mathbf{}𝝀_{\mathrm{𝐞𝐟𝐟}}`$
Here we solve equation (3) by expansion in the limit $`l_m\lambda _{\mathrm{eff}}`$. This condition means that collisions are too weak to scatter the electron out of the loss cones. Therefore, $`F(x,\mu )0`$ when $`|\mu |>\mu _{\mathrm{crit}}`$.
We make use the fact that $`(V/\nu )(/x)\lambda /l_m1`$. Also we will show that $`1/\nu \tau _m1`$. The validity of this last assumption appears below. To zero order, we have $`F/x=0`$, and $`F(x,\mu )=F_0(\mu )`$. $`F_0(\mu )=F_0(\mu )`$ because of electron reflection at the mirrors and the symmetry of the loss cones. Up to first order, $`F(x,\mu )=F_0(\mu )+F_1(x,\mu )`$, and we have
$`\mu V{\displaystyle \frac{F_1}{x}}={\displaystyle \frac{\nu }{2}}{\displaystyle \frac{}{\mu }}\left[(1\mu ^2){\displaystyle \frac{F_0}{\mu }}\right]+{\displaystyle \frac{F_0}{\tau _m}}.`$ (A1)
We integrate this equation over $`x`$ along a closed back and forth trajectory of a trapped electron shown by the dotted lines in Figure 1(b), to obtain
$`\{\begin{array}{c}/\mu \left[(1\mu ^2)F_0/\mu \right]+2F_0/\nu \tau _m=0,\hfill \\ F_0(\mu )=F_0(\mu ),F_0(\pm \mu _{\mathrm{crit}})=0.\hfill \end{array}`$ (A4)
We solve equation (A4) by a further expansion, $`1/\nu \tau _m1`$. The even solution in the “inside” region $`1|\mu |e^{\nu \tau _m}`$ up to first order is
$`F_0^{(i)}=C^{(i)}\left[1{\displaystyle \frac{1}{\nu \tau _m}}\mathrm{ln}{\displaystyle \frac{1}{1\mu ^2}}\right],1|\mu |e^{\nu \tau _m}.`$ (A5)
On the other hand, the zero order solution in the “boundary” regions $`1|\mu |1`$ is
$`F_0^{(b)}`$ $`=`$ $`C^{(b)}\mathrm{ln}{\displaystyle \frac{1|\mu |}{1\mu _{\mathrm{crit}}}},1\mu _{\mathrm{crit}}1|\mu |1.`$ (A6)
We match solutions (A5) and (A6) together in regions $`e^{\nu \tau _m}1|\mu |1`$ to finally obtain $`\tau _m=\nu ^1\mathrm{ln}m`$, justifying $`1/\nu \tau _m1`$. This is the first result in equation (8).
## Appendix B Solution of equation (3) in the limits $`𝝀_{\mathrm{𝐞𝐟𝐟}}\mathbf{}𝒍_𝒎\mathbf{}𝝀^\mathrm{𝟐}\mathbf{/}𝝀_{\mathrm{𝐞𝐟𝐟}}`$ and $`𝝀^\mathrm{𝟐}\mathbf{/}𝝀_{\mathrm{𝐞𝐟𝐟}}\mathbf{}𝒍_𝒎`$
Let us consider the kinetic equation (3) in the more limited case $`\lambda l_m`$ (note that $`\lambda _{\mathrm{eff}}\lambda `$). This means that in the kinetic equation $`(V/\nu )(/x)\lambda /l_m1`$. We will also show that $`1/\nu \tau _m1`$. The validity of this assumption appears below. To zero order, we have $`F/\mu =0`$, so $`F(x,\mu )=F_0(x)`$. $`F_0(x)=F_0(x)`$ because of symmetry. Up to first order, $`F(x,\mu )=F_0(x)+F_1(x,\mu )`$, and we have
$`{\displaystyle \frac{\nu }{2}}{\displaystyle \frac{}{\mu }}\left[(1\mu ^2){\displaystyle \frac{F_1}{\mu }}\right]=\mu V{\displaystyle \frac{F_0}{x}}{\displaystyle \frac{F_0}{\tau _m}}.`$ (B1)
We integrate the above equation over $`\mu `$, and then set $`\mu =\pm 1`$ to find the constant of integration. As a result, we obtain $`F_0/\tau _mV(F_0/x)`$ \[so, $`1/\nu \tau _m`$ is of second order\], and $`F_1/\mu =(V/\nu )(F_0/x)`$. We integrate this last equation over $`\mu `$ once more, and obtain
$`F_1=(\mu V/\nu )(F_0/x)+C(x),`$ (B2)
where $`C(x)`$ is another integration constant.
We continue the expansion of the kinetic equation (3) to next order. Up to second order, $`F(x,\mu )=F_0(x)+F_1(x,\mu )+F_2(x,\mu )`$. Using equation (B2), we have
$`{\displaystyle \frac{\nu }{2}}{\displaystyle \frac{}{\mu }}\left[(1\mu ^2){\displaystyle \frac{F_2}{\mu }}\right]={\displaystyle \frac{\mu ^2V^2}{\nu }}{\displaystyle \frac{^2F_0}{^2x}}+\mu V{\displaystyle \frac{C}{x}}{\displaystyle \frac{F_0}{\tau _m}}.`$ (B3)
We integrate equation (B3) over $`\mu `$ from $`1`$ to $`1`$ and obtain
$`\{\begin{array}{c}^2F_0/^2x+(3\nu /\tau _mV^2)F_0=0,\hfill \\ F_0(x)=F_0(x).\hfill \end{array}`$ (B6)
Finally, we integrate this equation and obtain zero order solution for the time-dependent distribution function (2)
$`f(t,x)=e^{t/\tau _m}F_0(x)=e^{t/\tau _m}\mathrm{cos}\left(x\sqrt{3\nu /\tau _mV^2}\right),`$ (B7)
where we drop an unnecessary normalization constant of integration.
Now, to find $`\tau _m`$, we calculate the flux of escaping electrons through the two escape windows (see Figure 1)
$`N/t=2{\displaystyle _{\mu _{\mathrm{crit}}}^1}\mu Vf(t,l_m/2)𝑑\mu =(V/m)e^{t/\tau _m}\mathrm{cos}\left[(l_m/2)\sqrt{3\nu /\tau _mV^2}\right],`$ (B8)
where we use equation (B7) for $`f(t,l_m/2)`$. On the other hand, the flux is equal to the change of the total number of electrons
$`N/t={\displaystyle _{l_m/2}^{l_m/2}}{\displaystyle _1^1}(f/t)𝑑\mu 𝑑x=(4/\tau _m)e^{t/\tau _m}\sqrt{\tau _mV^2/3\nu }\mathrm{sin}\left[(l_m/2)\sqrt{3\nu /\tau _mV^2}\right].`$ (B9)
Equating the two formulas for $`N/t`$, we obtain
$`(3/16)(\nu \tau _m/m^2)=\mathrm{tan}^2\left(\sqrt{3\nu l_m^2/4\tau _mV^2}\right).`$ (B10)
In the limit $`\lambda l_m\lambda ^2/\lambda _{\mathrm{eff}}`$ the argument of the tangent above is small, so we expand the tangent and obtain $`\tau _m=\nu ^1(l_m/\lambda _{\mathrm{eff}})`$, while $`F_0\mathrm{const}`$. In the limit $`\lambda ^2/\lambda _{\mathrm{eff}}l_m`$ the left hand side of equation (B10) is large, therefore, the argument of the tangent is $`\pi /2`$, and we have $`\tau _m=\nu ^1(3/\pi ^2)(l_m/\lambda )^2`$ \[the third line in equation (8)\], i.e. the escape time is controlled by diffusion in $`x`$-space. In both limits $`1/\nu \tau _m1`$, as we assumed above, (and of second order).
Now, the limit $`\lambda _{\mathrm{eff}}l_m\lambda `$ is still left. The result in this case is the same as the result in case $`\lambda l_m\lambda ^2/\lambda _{\mathrm{eff}}`$. However, instead of solving the kinetic equation, we give the following qualitative arguments supported by our numerical simulations (see Figure 2). The relaxation time of the electron distribution in $`\mu `$-space can be estimated as $`\mathrm{\Delta }t_\mu \nu ^1`$. The relaxation time in $`x`$-space can be estimated as the crossing time $`\mathrm{\Delta }t_xl_m/V=\nu ^1(l_m/\lambda )`$ in case $`l_m\lambda `$, and as the time of diffusion across $`\mathrm{\Delta }t_x\nu ^1(l_m/\lambda )^2`$ in case $`\lambda l_m`$. All relaxation times are small compared to the escape time $`\tau _m`$, i.e. $`\mathrm{\Delta }t_\mu ,\mathrm{\Delta }t_x\tau _m`$ for the entire range $`\lambda _{\mathrm{eff}}l_m\lambda ^2/\lambda _{\mathrm{eff}}`$. This means that the distribution function is approximately constant in $`x`$ and $`\mu `$, say $`F_01`$, $`fe^{t/\tau _m}`$. We then carry out calculations similar to those we used in formulas (B8) and (B9) to find that $`\tau _m=\nu ^1(l_m/\lambda _{\mathrm{eff}})`$ \[the second line in equation (8)\].
## Appendix C The Additional electron flow produced by electric field
Let us, for simplicity, consider the Lorentz gas. The results for the Spitzer gas are similar.
First, we derive an estimate for the additional flow $`d\stackrel{~}{}`$ of electrons that are in an interval $`V[V,V+dV)`$ of the velocity space, produced by an electric field $`E`$ due to the change of the two loss cones of a mirror trap. Let us consider only the principle mirrors, because they mainly control the diffusion of electrons (see Section 3). In this appendix, we denote their mirror strength (the principle mirror strength) as $`M`$.
The principle mirror strength is of order of five, so in the case when the magnetic field decorrelation length is more than or approximately equal to the electron mean free path, $`l_0\lambda `$, the escape of electrons from the mirror trap is mainly controlled by their spatial diffusion, see Section 2. Thus, in this limit, the electrons “do not care” about the size of the loss cones, and therefore, no additional flow arises.
In the case $`l_0\lambda `$ there is a non-zero additional flow $`d\stackrel{~}{}`$. In Figure 1(b), because of the electric field, the loss cone on the left, $`\mu _{\mathrm{crit},}`$, is not equal to that on the right, $`\mu _{\mathrm{crit},+}`$. The size of the two loss cones is estimated from the conservation of the electron magnetic moment, $`(1\mu ^2)V^2/B=\mathrm{const}`$, and from the conservation of energy, $`m_eV^2/2+eEx=\mathrm{const}`$. We have
$`\mu _{\mathrm{crit},\pm }^211/M\pm (eEl_M/m_eV^2M),`$ (C1)
where $`l_M`$ is the spacing of the principle mirrors.
Let $`d\stackrel{~}{}_+`$ and $`d\stackrel{~}{}_{}`$ be the absolute values of the fluxes of the escaping electrons to the right and to the left respectively. Then, their sum is
$`d\stackrel{~}{}_++d\stackrel{~}{}_{}=(l_M/\tau _M)\mathrm{\hspace{0.25em}2}\pi V^2f_0dV,`$ (C2)
where $`2\pi V^2f_0dV`$ is the number density of electrons expressed in terms of the Maxwellian zero order electron distribution function $`f_0`$, and $`\tau _M`$ is the escape time, see equations (2), (9), (23) and Section 2. The actual electron flow, $`d\stackrel{~}{}`$, is equal to the difference of $`d\stackrel{~}{}_+`$ and $`d\stackrel{~}{}_{}`$, because they are in opposite directions. An estimate for the ratio $`d\stackrel{~}{}/(d\stackrel{~}{}_++d\stackrel{~}{}_{})`$ is
$`{\displaystyle \frac{d\stackrel{~}{}}{d\stackrel{~}{}_++d\stackrel{~}{}_{}}}={\displaystyle \frac{d\stackrel{~}{}_+d\stackrel{~}{}_{}}{d\stackrel{~}{}_++d\stackrel{~}{}_{}}}\left[{\displaystyle _{\mu _{\mathrm{crit},+}}^1}\mu 𝑑\mu {\displaystyle _{\mu _{\mathrm{crit},}}^1}\mu 𝑑\mu \right]/\left[{\displaystyle _{\mu _{\mathrm{crit},+}}^1}\mu 𝑑\mu +{\displaystyle _{\mu _{\mathrm{crit},}}^1}\mu 𝑑\mu \right].`$ (C3)
Now, using equations (C1)–(C3), we obtain the additional electron flow
$`d\stackrel{~}{}2\pi (eE/m_e)f_0(l_M^2/\tau _M)dV.`$ (C4)
The factor $`(l_M^2/\tau _M)`$ in this equation is proportional to the spatial diffusivity of the electrons (provided that the diffusivity is controlled by the principle mirrors, see Section 3). Thus, it is obviously $`l_M^2/\tau _M=R_\mathrm{D}\lambda _\text{L}^2\nu _\text{L}`$, where $`\lambda _\text{L}`$ and $`\nu _\text{L}`$ are the standard Lorentz mean free path and collision frequency, and $`R_\mathrm{D}`$ is the reduction of the spatial diffusivity reported in Section 3. Using that $`\lambda _\text{L}V^4`$, $`\nu _\text{L}V^3`$, and equations (23), (55), $`V_T=\sqrt{2kT/m_e}`$, we finally obtain
$`d\stackrel{~}{}(1/2)(2/\pi )^{3/2}\left(k^{3/2}T^{3/2}E/m_e^{1/2}e^3\mathrm{ln}\mathrm{\Lambda }\right)R_\mathrm{D}\upsilon ^5e^{\upsilon ^2}d\upsilon .`$ (C5)
Now we like to compare this result for the additional flow $`d\stackrel{~}{}`$ with the main flow $`d`$ produced by the electric field due to acceleration of particles. The later is
$`d={\displaystyle _1^1}\mu Vf_1𝑑\mu \mathrm{\hspace{0.17em}2}\pi V^2𝑑V𝑑\mu =(2/3)(2/\pi )^{3/2}\left(k^{3/2}T^{3/2}E/m_e^{1/2}e^3\mathrm{ln}\mathrm{\Lambda }\right)R_\mathrm{D}\upsilon ^7e^{\upsilon ^2}d\upsilon ,`$ (C6)
where we substituted function $`f_1`$ given by (24), and function $`S(\upsilon )=\gamma _\text{E}S_E(\upsilon )`$ given by (43) and (51). As a result,
$`d\stackrel{~}{}/d(3/4)\upsilon ^2.`$ (C7)
Because the electrical current and the heat flow are mainly transported by superthermal electrons $`\upsilon ^24`$, the additional flow produced by electric field due to non-equal loss cones can indeed be neglected in comparison with the main flow due to acceleration of electrons by electric field.
|
warning/0006/cond-mat0006244.html
|
ar5iv
|
text
|
# Motion in a rocked ratchet with spatially periodic friction
## I Introduction
The search for the possibility of unidirectional motion in a periodic system without the application of any obvious bias is of current research interest. Such possibility requires the system to be out of equilibrium in order for the process to be consistent with the second law of thermodynamics. Several physical models have been proposed to obtain such motion. In all the models noise plays the central role. One of the most discussed models is the one in which an asymmetric periodic potential system is adiabatically rocked by applying constant forces $`F`$ and $`F`$ at regular intervals of time. One obtains unidirectional motion because in such a system the current $`j(+F)j(F)`$. Eventhough the time averaged applied force over a period vanishes the averaged current $`<j>=0.5[j(+F)+j(F)]`$ becomes finite in the presence of noise ( thermal fluctuations). Moreover, the average current $`<j>`$ was found to peak at an intermediate noise strength (or temperature). In this model it has been further shown that by suitably choosing the asymmetric periodic potential one may obtain current reversal as a function of temperature provided the rocking frequency is high. Similar results, however, can be obtained in the presence of a unbiased colored noise instead of the oscillating force. There are several other interesting models to obtain unidirectional motion including models where potential barriers themselves are allowed to fluctuate or models wherein symmetric potential system is driven by temporally asymmetric forces, etc.
The result that thermal noise helps to obtain unidirectional current in a periodic system was quite important. But later on it was pointed out that obtaining mere current does not necessarily mean that the system does work efficiently. Doing work involves flow of current against load and hence one must, in the spirit of the model, obtain current up against a tilted (but otherwise periodic) potential system. Analysis shows, however, that the efficiency of an adiabatically rocked system (ratchet) monotonically decreases with temperature. Therefore though such a ratchet system helps extract a large amount of work at an intermediate temperature ( where the current peaks) the work is accomplished at a larger expense of input energy; thermal fluctuation does not facilitate efficient energy conversion in this model ratchet system. In a subsequent work this deficiency was rectified but in a different model wherein the asymmetric potential oscillates in time, instead of its slope being changed (rocked) between $`+F`$ and $`F`$ adiabatically. In both these models the friction coefficient was constant and uniform in space. The present work makes a detailed study of the rocked ratchet system with nonuniform friction coefficient which varies periodically in space.
In this work we take the friction coefficient to vary with the same periodicity as the potential but with a phase difference, $`\varphi `$. The phase difference $`\varphi `$, the amplitude $`\lambda `$ of variation of friction coefficient, the amplitude $`F_0`$ of rocking, the load, etc. affect the functioning of the ratchet in an intricate and nontrivial manner. The two of the important results we obtain are: (1) The efficiency of the adiabatically rocked ratchet shows a peak as a function of temperature, though the peak (which may or may not exist in case of spatially asymmetric potentials) position does not coincide with the temperature at which the current peaks, and (2) the current could be made to reverse its direction as a function of noise strength and the amplitude $`F_0`$ even at low frequencies of rocking. These attributes are solely related to the medium being inhomogeneous with space dependent friction. It is worth noting that the introduction of space dependent friction, though does not affect the equilibrium properties (such as the relative stability of the locally stable states), changes the dynamics of the system in a nontrivial fashion. Recently it has been shown that these systems exhibit noise induced stability, it shows stochastic resonance in washboard potentials without the application of external periodic input signal, and also unidirectional motion in periodic symmetric potential (non ratchet-like) systems.
In the next section we describe our model and obtain an expression for current and efficiency in the quasi-static limit. In Sec. III we present our results.
## II Equation of motion in inhomogeneous systems
The nature of correct Fokker-Planck equation in the presence of space-dependent diffusion coefficient (inhomogeneous medium) was much debated earlier. Later on the correct expression was found from a microscopic treatment of system-bath coupling. The motion of an overdamped particle, in a potential $`V(q)`$ and subject to a space dependent friction coefficient $`\gamma (q)`$ and an external force field $`F(t)`$ at temperature $`T`$ is described by the Langevin equation
$$\frac{dq}{dt}=\frac{(V^{}(q)F(t))}{\gamma (q)}k_BT\frac{\gamma ^{}(q)}{[\gamma (q)]^2}+\sqrt{\frac{k_BT}{\gamma (q)}}\xi (t),$$
(1)
where $`\xi (t)`$ is a randomly fluctuating Gaussian white noise with zero mean and correlation :
$`<\xi (t)\xi (t^{})>=2\delta (tt^{})`$. Here $`<..>`$ denotes an ensemble average over the distribution of the fluctuating noise $`\xi (t)`$. The primes in Eq. (1) denote the derivative with respect to the space variable $`q`$. It should be noted that the above equation involves a multiplicative noise with an additional temperature dependent drift term. The additional term turns out to be essential in order for the system to approach the correct thermal equilibrium state. We take $`V(q)=V_0(q)+qL`$, where $`V_0(q+2n\pi )=V_0(q)=Vsin(q)`$, $`n`$ being any natural integer. $`L`$ is a constant force ( load) representing the slope of the washboard potential against which the work is done. Also, we take the friction coefficient $`\gamma (q)`$ to be periodic :
$`\gamma (q)=\gamma _0(1\lambda \mathrm{sin}(q+\varphi ))`$, where $`\varphi `$ is the phase difference with respect to $`V_0(q)`$. The equation of motion is equivalently given by the Fokker-Planck equation
$$\frac{P(q,t)}{t}=\frac{}{q}\frac{1}{\gamma (q)}[k_BT\frac{P(q,t)}{q}+(V^{}(q)F(t))P(q,t)].$$
(2)
This equation can be solved for the probability current $`j`$ when $`F(t)=F_0`$ = constant, and is given by
$$j=\frac{k_BT(1\mathrm{exp}(2\pi (F_0L)/k_BT))}{_0^{2\pi }\mathrm{exp}(\frac{V_0(y)+(F_0L)y}{k_BT})𝑑y_y^{y+2\pi }\gamma (x)\mathrm{exp}\frac{V_0(x)(F_0L)x)}{k_BT}}dx.$$
(3)
In the presence of space dependent friction and the phase lag $`\varphi 0,\pi `$ and in the absence of load $`j(F_0)j(F_0)`$ even for a spatially periodic symmetric potential. Thus when the system is subjected to an external ac field $`F(t)`$ the unidirectional particle flow (or rectification of the current) takes place. The phase lag $`\varphi `$ brings in the intrinsic asymmetry in the dynamics of the system. When an externally applied ac force changes slowly enough (quasi-static or adiabatic limit) i.e., when the time scale of variation of $`F(t)`$ is much larger compared to any other time scales involved in the system we can readily obtain an expression for the unidirectional current. For a field $`F(t)`$ of a square wave amplitude $`F_0`$, an average current over the period of oscillation is given by, $`<j>=\frac{1}{2}[j(F_0)+j(F_0)]`$. This particle current can even flow against the applied load $`L`$ and thereby store energy in useful form. In the quasi-static limit following the method of stochastic energetics it can be shown that the input energy $`E_{in}`$ ( per unit time) and the work $`W`$ ( per unit time) that the ratchet system extracts from the external noise are given by $`E_{in}=\frac{1}{2}F_0[j(F_0)j(F_0)]`$ and $`W=\frac{1}{2}L[j(F_0)+j(F_0)]`$ respectively. Thus the efficiency ( $`\eta `$) of the system to transform the external fluctuation to useful work is given by
$$\eta =\frac{L[j(F_0)+j(F_0)]}{F_0[j(F_0)j(F_0)]}.$$
(4)
Henceforth all our variables like $`<j>,E_{in},W,F_0,T`$ are made dimensionless. The amplitude of potential $`V`$ is set to unity as all other energy scales are scaled with respect to $`V`$. We evaluate $`<j>,W,E_{in}`$, and $`\eta `$ numerically using Eq. (3).
## III Results and Discussions
First, we present our results for average (net) unidirectional current in a symmetric periodic potential induced by adiabatic rocking and in the absence of load. We emphasize here that to obtain these currents the system must be inhomogeneous. The phase lag $`\varphi `$ (except for $`\varphi =0,\pi `$, for which unidirectional current is not possible) plays an important role in determining the direction and magnitude of $`<j>`$. For $`2\pi >\varphi >\pi `$, in the presence of external quasistatic force $`F(t)`$ and in the absence of load $`L`$, we have a forward moving ratchet (current flowing in the positive direction) and for $`0<\varphi <\pi `$ we have the opposite. For instance, if we examine the effect of friction coefficient close to the minimum of the potential two different situations are encountered depending on the value of $`\varphi `$. When $`\varphi >\pi `$ and $`F_0>0`$ the particle experiences lower friction near the barriers in the direction of acquired velocity. The situation is reverse when $`F_0<0`$. From Eq. (3) it follows that in a static force $`F_0`$, $`j(F_0)j(F_0)`$ , hence rectification of current occurs in the presence of external adiabatic drive. Moreover, it should be emphasized that the magnitude of current or mobility, in the static field $`F_0`$, depends sensitively on the potential and the frictional profile over the entire period. Depending on the system parameters the current or mobility can be much larger or smaller than the current or mobility of a particle moving in a homogeneous medium characterised by the space averaged frictional coefficient. In fact, in the intermediate values of temperature and $`F_0`$ the mobility can be made much larger than their asymptotic limits. This leads to stochastic resonance in a washboard potential in the absence of ac signal .
In Fig. 1 the average unidirectional current $`<j>`$ is presented as a function of $`\varphi `$ and $`T`$. For this figure the value of amplitude of square-wave ac field $`F_0=0.05`$, and the amplitude of frictional modulation $`\lambda =0.9`$. It can be seen that $`<j>`$ changes sign as a function of $`\varphi `$ and current exhibits either a minimum or a maximum as a function of temperature depending on the value of $`\varphi `$. At the two limits of temperature ($`0`$ and $`inf`$) the currents vanish. This is the case for the value of $`F_0`$ less than the critical value $`F_c`$ of $`F_0`$ where the barrier to motion in either direction vanishes. In our case the critical value of $`F_0`$ is equal to 1. This stochastic resonance-like phenomenon has been observed in rocked ratchet systems characterised by asymmetric periodic potentials . From the contours of the plot it is clear that as we move away from phase shift $`\varphi =\pi `$ positive (or negative) direction the temperature at which maxima (minima) occurs shifts to a larger value and the absolute value of the current at the peak decreases. However, the present symmetric potential situation does not lead to multiple current reversals as a function of temperature in the quasi-static limit of an external drive.
The current as a function of $`T`$ and $`F_0`$ is shown in Fig. 2 for $`\varphi =1.3\pi `$ and $`\lambda =0.9`$. For smaller fields $`F_0`$ compared to the critical field $`F_c`$ the current exhibits a maximum as a function of temperature. As we increase $`F_0`$ the temperature at which the peak occurs decreases. For fields larger than $`F_0`$ the current, however, decreases monotonically with temperature because the barrier to motion disappears. In this high field region mobility of a particle decreases with increase of temperature . However, as we increase the field the net current $`<j>`$ monotonically changes and saturates to a finite value. This is in contrast $`<j>`$ vanishes for large $`F_0`$ in the ratchet subjected to a rocking force in the absence of space dependent friction. In Fig. 3 we have plotted $`<j>`$ versus $`<F_0>`$ and $`\varphi `$ for fixed values of $`T=0.5`$ and $`\lambda =0.9`$. It can again be seen clearly that the current monotonically varies and saturates to a value given by $`\frac{\lambda }{2}sin(\varphi )`$ independent of temperature. This result follows from the analysis of Eq.(3). As expected current reversal can be seen as a function of $`\varphi `$ for large $`F_0`$. It can also be verified that the current $`<j>`$ increases monotonically as a function of the amplitude $`\lambda `$ of the friction coefficient and hence we do not present variation of $`<j>`$, etc., with $`\lambda `$.
We now discuss the efficiency of a symmetric periodic potential system in the presence of space dependent friction driven by an adiabatic periodic field. The efficiency of such a system with uniform friction has been studied earlier and it has been shown that temperature does not facilitate the efficiency $`(\eta )`$ of energy conversion in the system . To calculate $`\eta `$ we make use of Eqs. (3) and (4). In our analysis the load $`L`$ is applied against the direction of net current (in the absence of load). In this situation particle current can flow against the applied load $`L`$ less than some critical value $`L_c`$ thereby storing energy in useful form. For $`L>L_c`$ one cannot talk meaningfully the concept of efficiency as the current flows in the direction of the load and hence no storage of useful energy takes place. In Fig. 4 we have plotted efficiency $`\eta `$, input energy $`E_{in}`$ and work done $`W`$ (scaled up by a factor 60 for convenience of comparison) as a function of $`T`$ for the parameter values, $`F_0=0.5`$, $`\varphi =1.3\pi `$, $`\lambda =0.9`$, and the load $`L=0.04`$. The figure shows that the efficiency exhibits a maximum as a function of temperature indicating that thermal fluctuation facilitates energy conversion. This in contrast to the case of uniform friction coefficient where $`\eta `$ decreases monotonically with the increase of temperature in the same adiabatic limit . It is to be mentioned that the temperature corresponding to the maximum efficiency is not the same as the temperature at which the average current $`<j>`$ becomes maximum in the absence of load. The temperature at which the extracted work maximizes is not the same as the temperature at which the efficiency becomes maximum for the same parameter values. The input energy increases with temperature monotonically and saturates at the high temperature limit. $`\eta `$, $`W`$, and $`E_{in}`$ show similar qualitative behaviour for other parameter values. The above important observation of temperature facilitating the energy conversion is applicable for the spatially symmetric potential. In general in adiabatically rocked systems with frictional nonuniformities the increasing thermal noise need not increase the efficiency. The efficiency is sensitive to the qualitative nature of the periodic potential (and also to the nonuniformity of friction). For instance, asymmetric potential exhibits quite complex behaviour of $`\eta `$ and $`<j>`$. To illustrate this we take $`V_0(q)=\mathrm{sin}q\frac{\mu }{4}\mathrm{sin}2q`$ ( where $`\mu `$ lies between -1 and 1, and is the asymmetry parameter). With this potential we discuss three separate cases: Case A - system in a symmetric potential ($`\mu =0`$) in an inhomogeneous medium ($`\lambda 0`$), Case B - system in an asymmetric potential ($`\mu 0`$) in an inhomogeneous medium ($`\lambda 0`$), and Case C - system in an asymmetric potential ($`\mu 0`$) in a homogeneous medium ($`\lambda =0`$).
In Fig.5, we have presented results of $`\eta `$ versus $`T`$ for all the three cases described above. For this we have taken $`F=0.5`$, $`L=0.002`$, and $`\varphi =1.3\pi `$. For case A $`\lambda =0.9`$, for case B $`\mu =0.08`$ and $`\lambda =0.9`$, and for case C $`\mu =0.08`$. As discussed earlier for the case A temperature maximizes the efficiency. Case C, where the medium is homogeneous, efficiency monotonically decreases with temperature. These observations have been emphasized in earlier literature. The case B, where potential asymmetry and frictional inhomogeneity are present, the efficiency decreases monotonically in this parameter regime. In general whether the temperature facilitates the energy conversion in case B depends sensitively on the system parameters. In some limited parameter range the peaking behaviour is seen as in Fig.6. The parameter values are indicated in the caption. The presence of asymmetry in the potential may or may not help in enhancing the efficiency of the system. This can be seen from Figs. 5 and 6. In all these cases the work done $`W`$ and the input energy $`E_{in}`$ show similar qualitative features as shown in the inset of Fig. 1. Now, we discuss the variation of efficiency as a function of load.
On general grounds it is expected that the efficiency too exhibits maximum as a function of load. It is obvious that the efficiency is zero when load is zero. At the critical value $`L_c`$ (beyond which current flows in the direction of the load) the value of current is zero and hence the efficiency vanishes again. In between these two extreme values of load the efficiency exhibits maximum. Beyond $`L=L_c`$ the current flows down the load and therefore the idea of efficiency becomes invalid. In Fig. 7, we have plotted $`\eta `$ versus load for all the three cases for chosen values of parameters as mentioned in the figure. In all these cases current monotonically decreases as a function of load. The work done against load $`W`$ exhibits a maximum as a function of load. The load at which $`W`$ shows maximum does not coincide with the load at which $`\eta `$ becomes maximum. The input energy $`E_{in}`$ as a function of load varies non monotonically exhibiting a minimum. However, depending on the case under consideration the value of the load at which the minimum in the input energies observed may be larger than $`L_c`$ above which efficiency is not defined.
In Fig. 8, we have plotted the efficiency versus the amplitude of the adiabatic forcing $`F_0`$ for all the three cases. It can be seen from the figure that for the system in an inhomogeneous medium, namely for cases A and B, $`\eta `$ exhibits a maxima and saturate to the same value in the large amplitude limit. In contrast, for the case C after exhibiting maximum $`\eta `$ goes to zero. This follows from the simple fact that in the large amplitude limit in the absence of frictional inhomogeneities the net unidirectional current tends to zero. The peculiar feature of saturation of efficiency in inhomogeneous media is related to the fact that the average current saturates to a constant value in the high amplitude limit as discussed earlier. This somewhat counter-intuitive result is typical to inhomogeneous media. Having discussed efficiency of energy conversion we now study the nature of net current $`<j>`$ in the presence of spatially asymmetric potential to examine if current reversals take place in the adiabatic limit in the absence of load. It is known from the earlier literature that in an adiabatically rocked asymmetric potential ratchet system net current does not exhibit reversals as a function of $`T`$. In these systems current reversals are possible when the frequency of the applied ac field is large. We show here that in the presence of frictional inhomogeneities in addition to asymmetry in the potential one can observe current reversal as a function of thermal noise. In Fig. 9, we have plotted the magnitude of net current $`<j>`$ versus $`T`$ for all the three cases A, B, and C. The corresponding parameter values are mentioned in the caption of the figure. The cases A and C do not exhibit current reversal. This is a general result independent of parameter values. However, in case B current reversal is observed. To obtain current reversal both asymmetry in potential and nonuniform friction coefficient are essential, that is, current reversals arise due to the combined effect of $`\varphi `$ and $`\mu `$. Moreover, it should be noted that to observe current reversals the parameter range should be such that the net current in case A is in the opposite direction to that in case C. For the case B for which current reversal is observed, the plot of efficiency separates into two disjoint branches as the load should be reversed keeping the magnitude same when the current reversal takes place. In the presence of $`\mu 0`$ and $`\lambda 0`$, where the current reversals are observed, the efficiency as a function of temperature is less than the maximum value of efficiency in either of the two cases A and C. That is, $`\eta (\mu 0,\lambda 0)<max[\eta (\mu =0,\lambda 0),\eta (\mu 0,\lambda =0)]`$. To further analyze the nature of current reversals, in Fig. 10, we have plotted $`<j>`$ as a function of $`\varphi `$ and $`T`$ for fixed values of $`\mu =1`$, $`\lambda =0.9`$, and $`F_0=0.5`$. From the contour plots it is clear that as a function of $`\varphi `$ the current reverses sign twice in the intermediate temperature range. Thus the current exhibits reentrant behaviour as a function of phase $`\varphi `$ which is special to the case B. As we decrease the asymmetry in the potential the $`<j>=0`$ contour line shifts towards $`T=0`$ thereby enhancing the domain of current reversal to a lower value of temperature as a function of $`\varphi `$. As a function of temperature the current reversals occur in a definite range of phase $`\varphi `$ which , in turn, depends on other material parameters. The qualitative behaviour of $`<j>`$ remains unaltered for different $`F_0`$ as long as $`F_0`$ is less than the critical value.
In Fig. 11, $`<j>`$ is plotted as a function of $`T`$ and $`F_0`$ for $`\mu =1`$, $`\varphi =1.3\pi `$, and $`\lambda =0.9`$. As opposed to the case of symmetric potential (Fig. 2), currents in the small temperature regime do exhibit maxima and then saturate to a constant value as noted earlier. As a function of $`T`$ the current exhibits similar features as in the case A (Fig. 2). Figure 12, shows $`<j>`$ as a function of $`F_0`$ and $`\varphi `$ for $`T=0.1`$, $`\mu =1`$, and $`\lambda =0.9`$. As opposed to case A (see Fig. 3) $`<j>`$ shows current reversal as a function of $`F_0`$ in the range $`0<\varphi <\pi `$. However, in the asymptotic limit of $`F_0`$ current saturates to a value $`\frac{\lambda }{2}sin(\varphi )`$ independent of the value of the asymmetry parameter $`\mu `$. From the contour plot it follows that as a function of phase $`\varphi `$ we observe the reentrant behaviour of current at high values of $`F_0`$. As we decrease $`\mu `$ the $`<j>=0`$ contour shifts towards smaller values of $`F_0`$ thus making it possible to observe the double reversals at even smaller values of $`F_0`$. This reentrant behaviour as a function of $`\varphi `$ and the current reversal as a function of $`F_0`$ is very specific to the case B alone (compare Fig. 3).
Thus we conclude from our studies that the dynamics of a particle in an inhomogeneous medium is rich and complex. In the presence of adiabatic forcing and asymmetry in the potential current reversals can be observed as a function of $`T`$ and $`F_0`$. And depending on the system parameters thermal fluctuations facilitate the energy conversion. The above behaviour cannot be seen in the homogeneous medium in the same adiabatic limit. However, it is possible to observe these in homogeneous media in the presence of finite frequency ac drive (nonadiabatic regime). This seems to suggest that $`\varphi `$ may play the characteristic role of frequency in our model in the absence of nonadiabatic ac drive. This has been noted earlier in the context of observation of stochastic resonance phenomena in inhomogeneous media in the presence of static tilt alone . As a function of phase $`\varphi `$, we observe reentrant behaviour for the current which arises because of interplay between asymmetry, inhomogeneity, thermal noise, and strength of the adiabatic forcing. Some of the phenomena can be understood at best at a qualitative level only. The effect of nonadiabatic forcing may be of further interest. Work on this line is under investigation.
## Acknowledgements
MCM acknowledges partial financial support and hospitality from the Institute of Physics, Bhubaneswar. MCM and AMJ acknowledge partial financial support from the Board of Research in Nuclear Sciences, DAE, India.
REFERENCES
FIGURE CAPTIONS .
Fig. 1. The net current $`<j>`$ as a function of $`\varphi `$ and $`T`$, for parameter values $`F_0=0.5`$, $`\lambda =0.9`$. In the base plane contour of surface plot $`<j>(\varphi ,T)`$ are given, dotted line indicates $`<j>=0`$ contour.
Fig. 2. $`<j>`$ as a function of temperature $`T`$ and the amplitude of the rocking force $`F_0`$, for $`\varphi =1.3\pi `$, $`\lambda =0.9`$.
Fig. 3. $`<j>`$ as a function of the amplitude of the rocking force $`F_0`$, and phase $`\varphi `$, for $`T=0.5`$.In the base plane contour of surface plot is given.
Fig. 4. Efficiency $`\eta ,E_{in}`$ and $`W`$ as a function of $`T`$ for $`\varphi =1.3\pi `$, $`F_0=0.5`$, $`\lambda =0.9`$, and $`L=0.4`$. $`W`$ has been scaled up by a factor $`60`$ to make it comparable with $`\eta `$ and $`E_{in}`$. Y-axis is in dimensionless units.
Fig. 5. Efficiency versus temperature for (i) case A ($`\mu =0`$, and $`\lambda =0.9`$), (ii) case B ($`\mu =0.08`$ , $`\lambda =0.9`$), and (iii) case C ($`\mu =0.08`$ , $`\lambda =0.0`$), for fixed $`F_0=0.5`$, $`\varphi =1.3\pi `$, and $`L=0.002`$.
Fig. 6. Efficiency $`\eta `$ versus temperature $`T`$, for different values of $`\mu `$ and for $`F_0=0.5`$, $`\lambda =0.9`$, $`L=0.1`$, and $`\varphi =1.3\pi `$.
Fig. 7. Efficiency versus load $`L`$ for (i) case A ($`\mu =0`$, and $`\lambda =0.9`$), (ii) case B ($`\mu =1.0`$ , $`\lambda =0.9`$), and (iii) case C ($`\mu =1.0`$ , $`\lambda =0.0`$), for fixed $`F_0=0.5`$, $`\varphi =1.3\pi `$, and $`T=0.1`$.
Fig. 8. Efficiency versus $`F_0`$ for (i) case A ($`\mu =0`$, and $`\lambda =0.9`$), (ii) case B ($`\mu =1.0`$ , $`\lambda =0.9`$), and (iii) case C ($`\mu =1.0`$ , $`\lambda =0.0`$), for fixed $`L=0.02`$, $`\varphi =1.3\pi `$, and $`T=0.1`$.
Fig. 9. Current $`<j>`$ versus temperature $`T`$ for (i) case A ($`\mu =0`$, and $`\lambda =0.9`$), (ii) case B ($`\mu =1.0`$ , $`\lambda =0.9`$), and (iii) case C ($`\mu =1.0`$ , $`\lambda =0.0`$), for fixed $`F_0=0.5`$, $`\varphi =0.3\pi `$, and $`L=0.0`$.
Fig. 10. The net current $`<j>`$ as a function of $`\varphi `$ and $`T`$, for parameter values $`F_0=0.5`$, $`\lambda =0.9`$, and $`\mu =1.0`$. In the base plane contour of surface plot $`<j>(\varphi ,T)`$ are given, dotted line indicates $`<j>=0`$ contour.
Fig. 11. $`<j>`$ as a function of temperature $`T`$ and the amplitude of the rocking force $`F_0`$, for $`\varphi =1.3\pi `$, $`\lambda =0.9`$, and $`\mu =1.0`$.
Fig. 12. $`<j>`$ as a function of the amplitude of the rocking force $`F_0`$, and phase $`\varphi `$, for $`T=0.5`$, and $`\mu =1.0`$. In the base plane contour of surface plot $`<j>(\varphi ,T)`$ are given, dotted line indicates $`<j>=0`$ contour.
|
warning/0006/astro-ph0006247.html
|
ar5iv
|
text
|
# Near-infrared photometry of isolated spirals with and without an AGN. II: Photometric properties of the host galaxies. Based on data obtained at: the European Southern Observatory, La Silla, Chile, the Télescope Bernard Lyot, Calar Alto Observatory, Las Campanas Observatory. Also based on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Institute
## 1 Introduction
Many studies have been devoted to analyze the properties of Seyfert host galaxies in order to understand the fuelling processes taking place in active galactic nuclei (AGN). At large scale, the fuelling of the active nucleus is supposed to be due to the transport of gas to the central region; this mechanism seems to be connected to the presence of a bar, which provides the non axisymmetric potential invoked in theoretical works (Simkin et al. 1980, Shlosman et al. 1989, Barnes & Hernquist 1991). One of the mechanisms proposed to drive the gas to the very central regions of barred galaxies is that of nested bars (Shlosman et al. 1989; Friedli & Martinet 1993; Combes 1994; Heller & Shlosman 1994), which has been recently proved to fuel molecular gas into an intense central starburst in NGC 2782 (Jogee et al. 1999) and in a Seyfert 2 galaxy, Circinus (Maiolino et al. 1999).
Nevertheless, from the observational side, recent studies (Moles et al. 1995; Ho et al. 1997; Hunt et al. 1999a) based on optical data conclude that barred galaxies are equally found among active and non active galaxies. Therefore, large scale bars are not the specific property that identifies the family of active spiral galaxies. NIR imaging is more reliable in determining the overall mass distribution in galaxies and it has been shown to be more efficient to detect bars (Mulchaey et al. 1997; Seygar & James 1998). From their NIR imaging analysis, McLeod & Rieke (1995) and Mulchaey & Regan (1997) find no evidence for a significant excess of bars in Seyfert galaxies. But a debate still exists on this matter: the analysis by Knapen et al. (2000) of high resolution NIR images by Peletier et al. (1999) points to an excess of bars among Seyferts at a 2.5$`\sigma `$ level, attributing their different result to their better spatial resolution and better matching of active and control samples than in previous works.
A different approach to the problem, also adopted here, consists in deriving detailed information on a number of selected objects, instead of performing statistics on large samples. Regan & Mulchaey (1999) analyze the HST high resolution dust morphology of 12 Seyferts, searching for central bars. The non-ubiquity of such bars led them to conclude that strongly barred potentials cannot be the only mechanism for driving gas into the nucleus; they propose central spiral dust lanes as an alternative method. In the same vein, Martini & Pogge (1999) analyze HST images of 24 Seyfert 2s and find nuclear spirals in 20 of them but only 5 with nuclear bars, concluding that nuclear spirals may be the channel to feed gas into the central engines.
Following this approach, we have chosen to analyze what the similarities and/or differences are between active and non active spirals. Since gravitational interaction has been invoked to be very efficient to induce the formation of bars or any other non axisymmetric component, we take the approach of only selecting isolated objects, in order to avoid the bias introduced by not taking into account the environmental characteristics of the considered galaxies. The DEGAS project (Dynamics and Nuclear Engine of Galaxies of Spiral type) aims at extending the analysis by Moles et al. (1995) and addresses the morphology of the galaxies, including optical and IR images, and the kinematics of the stars and gas, through long slit, high resolution spectroscopy. In this paper we present the analysis of the NIR imaging, which is particularly important because it allows: 1) to separate the various components (the bulge, disk, bar(s) and spiral arms) with the smallest contribution of the active nucleus; 2) to detect and characterize the properties of bars and other structures close to the nucleus, such as bars within bars, elongated disks, rings or lenses, traced by the old stellar population. We present the analysis based on the data set presented in Márquez et al. (1999), and compare the properties of active and non active galaxies in our sample, and those of our sample galaxies with those found in similar studies based on other samples.
Our sample is briefly described in Section 2. The results on the photometric decomposition and the bar properties are discussed in Sections 3 and 4. The discussion and conclusions are given in Section 5.
## 2 The sample and data
The infrared imaging data for 18 active objects and 11 non active galaxies are described in Márquez et al. (1999, hereafter Paper I). The active galaxies have been chosen with the following criteria: (a) Seyfert 1 or 2 from the Véron-Cetty & Véron (1993) catalogue; (b) with morphological information in the RC3 Catalogue; (c) isolated, in the sense of not having a companion within 0.4 Mpc (H<sub>0</sub>=75 km/s/Mpc) and cz$`<`$500 km/s, or companions catalogued by Nilson without known redshift; (d) nearby, cz$`<`$6000 km/s; and (e) intermediate inclination (30 to 65). The non active sample galaxies have been selected among spirals verifying the same conditions (b), (c), (d) and (e), and with morphologies (given by the complete de Vaucouleurs coding, not just the Hubble type) similar to those of the active spirals. Thus, all the galaxies in our sample are isolated, in the sense of avoiding possible effects of interactions with luminous nearby galaxies.
In Paper I we already stressed that these non active galaxies are well suited to be used as a control sample. For each object we give in Paper I: the image in the K’ band, the sharp-divided image (obtained by dividing the observed image by a filtered one), the difference image (obtained by subtracting a model to the observed one), the J-K’ color image, the ellipticity and position angle profiles, the surface brightness profiles in J and K’ and their fits by bulge+disk models and the J-K’ color gradient.
The mean resolution of our images is about 1 arcsecond, corresponding to a physical resolution between 100 and 300 parsecs for the closest and the more distant galaxy respectively. This resolution implies that we are able to map the region where the dynamical resonances are expected to occur (see for instance Pérez et al. 2000) and is therefore well suited for our purposes.
## 3 Results of the photometric bulge+disk decomposition
The surface brightness profiles derived for all the galaxies together with the best bulge/disk decomposition (1D exponential disk and de Vaucouleurs bulge) are displayed in Figs. 1-28 f-g, and the corresponding observed parameters are given in Table 3 of Paper I. Note that several galaxies (namely the active galaxies NGC 3660 and NGC 5728, and the non active galaxies NGC 151, NGC 2811 and NGC 3571) extend notably further out than the infrared images, so the resulting bulge+disk decomposition cannot be satisfactory. Full 2D fits have not been considered since, as already noticed by de Jong (1996b), they would not be better than 1D fits for the purposes of the present study.
The average bulge and disk effective surface brightnesses and effective radii in J and K’ as well as J-K’ colors are given in Table LABEL:decomp. Magnitudes have been corrected for galactic extinction, inclination (assuming transparent disks) and redshift (we have applied the K correction following Hunt et al. 1997). We note that, at face values, some differences are found between active and control galaxies. However, the application of non-parametric tests shows that those differences are not significant.
With respect to the bulge component, we recall that the parameters derived from the photometric decomposition are more critically dependent on the fitting procedure (see de Jong 1996a; Moriondo et al. 1996); in addition to this, the contribution of the AGN in active galaxies is not easy to extract, so bulge parameters can be less accurate. With these caveats in mind, the differences in the average values reported in Table LABEL:decomp are not conclusive; the Kolmogorov-Smirnoff (KS) test gives a probability greater than 98% for both samples to have the same magnitude, surface brightness and scale-length.
In Fig. 1 the bulge equivalent surface brightness is plotted as a function of equivalent radius (i.e. the Kormendy 1977 relation for bulges) for active and control galaxies. Different symbols were used for the various morphological types. However, since the number of galaxies in our sample is small, it is not possible to define Kormendy relations for the various morphological types. The best fits are $`\mu _{bulge}=(16.64\pm 0.24)+(3.77\pm 0.33)logr_{bulge}`$ and $`\mu _{bulge}=(16.48\pm 0.30)+(4.79\pm 0.48)logr_{bulge}`$ for control and active galaxies respectively ($`\mu _{bulge}=(16.52\pm 0.21)+(4.29\pm 0.32)logr_{bulge}`$ when the two samples are taken together). The Kormendy relations obtained for similar data in the near infrared by other authors are also drawn in these figures. Our data are in good agreement with these relations, and extend to somewhat smaller bulges.
The Xanthopoulos (1996) relation in the I band for Seyfert 1 vs. Seyfert 2 galaxies is also given for comparison (27 galaxies). Although it is displaced vertically relative to the relations found in K’, as expected, it shows a comparable slope. Therefore, considering that bulge scale-lengths should essentially be the same from I to K’ for a given galaxy (Evans 1994; see also Hunt et al. 1999a and below), this implies that the bulge component of Seyfert spirals would have an essentially uniform I-K’ color ((I-K’) $``$ 2.7).
The region occupied by the bulges obtained by de Jong (1996b) is also plotted in Fig. 1. The bulge parameters for our sample galaxies define a narrower relation than that found by de Jong, probably because we select only the most isolated objects. The same conclusion was found by Márquez & Moles (1999) when comparing the bulge parameters of their isolated spirals (Márquez & Moles 1996) with those of de Jong in the B band and Baggett et al. (1998) in the V band.
The main result is therefore that no difference is found between the bulges of active and control galaxies, in agreement with Hunt et al. (1999a). It appears that the bulges of active and control galaxies have totally similar structural properties.
With respect to the disk component, the disk surface brightness as a function of its equivalent radius is plotted in Fig. 2 for active and control galaxies. The best fits are $`\mu _{disk}=(16.89\pm 0.43)+(2.46\pm 0.60)logr_{disk}`$ and $`\mu _{disk}=(18.44\pm 0.67)+(0.75\pm 0.94)logr_{disk}`$ for active and control galaxies respectively. The slope of the relation is heavily dependent on the very large and uncertain value found for IC 454. If that point is taken out of the relation, we find for control galaxies $`\mu _{disk}=(16.82\pm 0.57)+(3.21\pm 0.86)logr_{disk}`$. Therefore, both control and active galaxies seem to share the same relation ($`\mu _{disk}=(16.98\pm 0.35)+(2.54\pm 0.50)logr_{disk}`$ for both samples simultaneously).
As for bulges, our data on disks are in good agreement with the relations found by Moriondo et al. (1998) and Hunt et al. (1999a) for normal galaxies. The relation is again narrower for our isolated galaxies than for de Jong’s sample, as reported in Márquez & Moles (1999). The relation found by Xanthopoulos (1995) in the I band is parallel to ours; considering the same scale length for the different bands (as for the bulge, also see below), this would imply an essentially uniform value for the disk color of isolated Seyfert spirals ((I-K’) $``$ 1.5).
Average values for the J-K’ colors of bulges and disks are given in Table LABEL:decomp. Bulge colors agree with those by Moriondo et al. (1998) who find a mean (J-K)=1.06 $`\pm `$ 0.3 for a sample of 14 early-type non active spirals. They also agree with the results by Hunt el al. (1999a), who found (J-K)=1.04 and (J-K)=1.07 for the bulges of Seyfert 1 and normal Sa galaxies, respectively. Disk colors occupy a narrower range, in good agreement with previous (J-K) determinations (Hunt et al. 1997; Hunt et al. 1999a). There appears to be a trend, with a large spread of values, for the bulges and disks of active galaxies to have smaller J-K’ values than the control sample, but an application of the KS test results in a $`>`$ 99% probability for both samples to have the same disk color distribution.
Bulge equivalent radii in the J and K’ bands follow the same relation for both active and control galaxies: $`r_{bulge}(K^{})=(0.915\pm 0.029)r_{bulge}(J)`$. With respect to the disk equivalent radii in the J and K’ bands: $`r_{disk}(K^{})=(0.988\pm 0.026)r_{disk}(J)`$. Both bulge and disk scale lengths appear to be essentially the same in J and K’.
A comparison of the bulge and disk equivalent radii in the J and K’ bands indicates that, again, both active and control galaxies show the same trend: $`r_{bulge}(K^{})=(0.12\pm 0.07)r_{disk}(K^{})`$ and $`r_{bulge}(J)=(0.17\pm 0.11)r_{disk}(J)`$. This is in agreement with previous results by de Jong (1996b) (mean $`r_{bulge}/r_{disk}`$ = 0.14 in K’), Courteau et al. (1996) (mean $`r_{bulge}/r_{disk}`$ = 0.13 in R), and Graham & Prieto (1999) (mean $`r_{bulge}/r_{disk}`$ = 0.24 in K’ for early spirals). A weak correlation between $`r_{bulge}`$ and $`r_{disk}`$ is reported by Seygar & James (1998).
In Fig. 3 we show the relation between the bulge and disk absolute magnitudes in K’. Our results are in good agreement with those by de Jong (1996b), considering that we deal with early-type spirals.
Within the spread of values we find no apparent differences in the photometric parameters between the host galaxies of the active and control samples. This is in good agreement with previous results, except for the disk surface brightnesses of Seyfert galaxies. Bulges of Seyferts in our sample cover the same region as those by Bender et al. (1992) and Andreadakis et al. (1995). The disks of both control and active galaxies are similar to those of normal galaxies in Hunt et al. (1999a; taken from de Jong 1996b and Moriondo et al. 1998) as seen in Fig. 2, and also agree with those by Andreadakis et al. (converted to the K band as in Hunt et al. 1999a). Nevertheless, the disk surface brightness of Seyfert galaxies selected from the CfA sample by Hunt et al. (1999a) and from the 12$`\mu `$m sample by Hunt et al. (1999b) are about 1 magnitude brighter in K than those of normal early type spirals. However, it has to be noted that their control samples were not specially designed to match the active sample; in particular, some clearly interacting systems are found among the active sample, so interactions may have an effect on the enhanced surface brightness they obtain.
This raises the question of how well the control sample used for comparison matches the active sample. Very recently, Knapen et al. (2000) have reported a higher bar percentage for active spirals with respect to non active ones. We stress that, unlike other authors, we are avoiding the possible effects of any strong interaction on the disk surface brightness. In particular, if our isolation criteria are applied to select only the isolated objects among those analyzed by Knapen et al., their conclusions change: the (small) resulting samples of active and control galaxies (13 and 11 with morphological information on the presence or absence of a bar) do show the same percentage of barred galaxies (8 and 7 galaxies, respectively). This result reinforces the importance of not including objects that could be suffering from a gravitational interaction (as it is the case, for instance, for NGC 7469, which is included in Knapen’s sample and is known to reside in an isolated pair; see, for instance, Márquez & Moles 1994). Since the sample of active galaxies by Hunt et al. (1999a) also comes from the CfA sample, the same considerations should apply and could explain the higher disk surface brightness they found as a result of including interacting objects. This point should be statistically confirmed by using a larger sample of active and non active isolated spirals.
## 4 Bars and central properties
We discuss here the properties of primary and secondary bars in the studied galaxies. We remind the reader that we refer to secondary bars every time a central elongation is detected, be it a bar, a lens, an inclined disk or a ring. Whether that elongation is actually a bar is a question that will be addressed using kinematic data.
Observed bar lengths and ellipticities are given in Paper I. They have been obtained from the ellipse fitting, corresponding to peaks in ellipticity for constant PA. All the discussion refers to deprojected bar parameters (see Jungwiert et al. 1997).
Regarding the primary, large scale bars, we find that the distribution of sizes is similar for both active and control galaxies. The same result is found when considering relative bar sizes (with respect to the disk size), as shown in Fig. 5. Both histograms are rather flat, and a KS test indicates that they are similar with a probability higher than 99.5%. The strength of the bars has been parametrized with (b/a), with smaller values of (b/a) for stronger bars as discussed by Martinet & Friedli (1997), and Chapelon et al. (1999). It has to be noted that their the criteria to estimate bar lengths and ellipticities are not exactly the same as ours. Nevertheless, we will be comparing active and control galaxies for which the estimations come from the same procedure (ellipse fitting, see above). In Fig. 5 we have plotted the bar strength as a function of the normalized length ($`a/R_{25}`$). Our results agree with those presented by the quoted authors. The comparison for active and control galaxies, with mean strengths of $`(b/a)`$ = 0.68 $`\pm `$ 0.19 and 0.51 $`\pm `$ 0.24, and mean lengths of $`(a/R_{25})`$ = 0.29 $`\pm `$ 0.18 and 0.32 $`\pm `$ 0.13 respectively, don’t show any significant difference. Martinet & Friedli (1997) defined strong bars for $`(b/a)`$ $``$ 0.6 and long bars for $`(a/R_{25})0.18`$. According to this definition, Fig. 5 also shows that the whole range from weak to strong bars is equally represented for both active and control galaxies. The majority of barred galaxies in both samples harbor long bars (this seems to be specially the case for control galaxies).
The fraction of detections of secondary bars is similar for both active and control galaxies. As suggested by Friedli & Martinet (1993), the formation of the secondary bar is mainly driven by the main bar, so in this case we calculate relative sizes of secondary bars with respect to primary ones. The distribution of the relative sizes is shown in Fig. 7. For active galaxies the distribution is more extended, but the differences are not significant. We stress the fact that the presence of secondary bars is not exceptional in non active galaxies.
Finally, the consideration of the primary and secondary bars together doesn’t show any correlation between the properties of both structures, or with the effective radii of the bulges and disks, in agreement with previous findings (Seygar & James 1998). The only hint we find is for a positive correlation between the sizes of the primary and secondary bars, but the scatter is too large and the sample small. In any case this would not be a difference between active and non active galaxies, since it would be present for the whole sample here.
Color gradients as a function of distance to the center (in kpc) are shown in Fig. 7. Disk colors are similar for both samples. Central colors of control galaxies agree with those of typical spirals determined by Griersmith et al. (1982) and Forbes et al. (1992). We have computed a normalized color gradient $`\delta `$(J-K’)(u) = (J-K’)(u) - (J-K’)$`(0.5)`$, where $`u`$ is the distance to the center in units of R<sub>25</sub>, $`u`$=r/R<sub>25</sub> so that (J-K’)$`(0.5)`$ is the color at a fixed relative distance of 0.5$`\times `$R<sub>25</sub>, which corresponds to the region were color gradients are reliable (see Paper I). This color gradient is plotted in Fig. 8 for active and control galaxies. Note than only three galaxies harbor bars longer than 0.5$`\times `$R<sub>25</sub> (namely NGC 4785, NGC 5728 and NGC 151), which have been excluded from the plot, so Fig. 8 shows colors inside the bar region. It can be seen that $`\delta `$(J-K’)(0.5) = 0 within the photometrical errors. $`\delta `$(J-K’)(u) values depart from $`\delta `$(J-K’)(u)=0 further out, mostly due to the background subtraction difficulties reported in Paper I. The differences in the behaviours of central regions of active and control galaxies are well visualized in this plot. In the innermost 0.1$`\times `$R<sub>25</sub>, active galaxies are generally redder than control galaxies (the reddest active galaxy is $``$1 magnitude redder than the reddest control galaxy). We also note that three control galaxies (namely NGC 2712, NGC 3835 and NGC 6155) show bluer colors inside 1 kpc, whereas none of the active galaxies becomes bluer in the center. Shaw et al. (1995) found that 19 out of a non-complete sample of 32 large barred galaxies show blue nuclei. Among them, only four were Seyferts. The result of non active galaxy with blue nucleus among isolated spirals should be confirmed with larger samples.
The average J-K’ colors of the central region (the innermost 200 pc) and of the primary bar are given in Table LABEL:decomp for all the galaxies of our sample that have a bar. The colors of the primary bars appear to be similar in active and non active galaxies, and also to be similar to the J-K’ value of the central region of the control sample. On the other hand, the central regions of the active sample have larger J-K’ values than the control one, in agreement with Kotilainen & Ward (1994). This difference cannot be explained as due to differential contributions of the nebular continuum in the J and K’ bands produced by the NLR of active galaxies. <sup>1</sup><sup>1</sup>1We have estimated that the central J-K’ colors change only by less than 10% if the color of each active galaxy is corrected for the nebular continuum contribution, estimated from the H$`\alpha `$ or $`H\beta `$ fluxes of the nucleus reported in the literature. It could be understood as coming either from an important contribution of active star forming regions usually found associated with the presence of the active nucleus in active galaxies (Rodríguez-Espinosa et al. 1987; Wilson 1988; Hunt & Giovanardi 1992) through the contribution of giants/supergiants (Shaw et al. 1995) or from dust re-emission in a dusty starburst or from the AGN illumination (Seygar & James 1999).
## 5 Discussion and Conclusions
Our analysis is intended to detect the differential properties between isolated active and non active spiral galaxies. We have considered here the large scale, global properties as well as the detailed morphology of the central regions, down to 100-300 pc. Our results show that there are no sizeable differences between active and non active isolated spiral galaxies in the volume limited sample considered here. By this, we mean that the global properties are similar, and none of the detected structures is exceptionally present or absent in one of the groups. Our results concerning the global properties, refer to a rather small sample and are in general agreement with previous ones, reinforcing our conclusion that hosts of isolated Seyfert galaxies have bulge and disk properties comparable to those of isolated non active spirals. In particular:
\- both samples define the same (Kormendy) relation between $`\mu _{eff}`$ and $`r`$ for bulges ;
\- disk components also share the same properties. This contradicts the result by Hunt et al. (1999a, 1999b) that Seyfert disks are about 1 magnitude brighter than the disks of non active spirals. This discrepancy may be explained in terms of a possible contamination by interacting objects in their sample;
\- bulge and disk scale lengths are correlated, with $`r_{bulge}0.2r_{disk}`$;
\- central colors of active galaxies are redder than the centers of non active spirals, most probably due to the AGN light re-emitted by the hot dust and/or to the presence of active star formation in circumnuclear regions.
It is generally admitted that the mechanism responsible for the transport towards the center and the possible onset of nuclear activity could be related with the presence of bars or other non axysimmetric structures. Regarding the primary bars, we know that the fraction of barred galaxies is not different among active and non active spiral galaxies (Mc Leod & Rieke 1995, Moles et al. 1995, Ho et al. 1997, Mulchaey & Regan 1997, Hunt & Malkan 1999). We notice that only one of the active galaxies in our sample, namely ESO 139-12, does not harbor a primary bar. This exception could however be of some interest since it raises the question of what kind of mechanism could produce non circular gas motions in an isolated object with no large-scale bar. Our results also show that primary bars have the same mean strength and length in both families.
The difference could reside in the properties of secondary central elongations (bars, inclined disks or rings, see Paper I). The complete analysis has to include photometric and kinematic information, in order to see what specific processes are taking place and whether they differ between active and control galaxies. But even with only the morphological information we report here, we can already derive some conclusions, keeping in mind the limitations imposed, in particular by the resolution reached by the IR images we have studied. Our result is that, down to scales of 100-300 pc, the detection rate of secondary bars is not different for active and control galaxies. Admittedly our sample is too small to allow strong statistical conclusions. But the fact is that the presence of secondary bars is not exceptional among non active galaxies. On the other hand, there is a number of active spirals in our sample with no detected secondary bar.
There remains the question of whether differences would be found at smaller scales, in particular those related to the presence of different nuclear structures. Even if this point is out of the scope of this paper, we note that HST images are only available for one of the galaxies in the control sample, so no comparison between active and non active galaxies can be attempted at this stage. In this respect, we stress that the feedback from numerical simulations is crucial to understand the mechanisms that are at the origin of the nuclear activity, and the scale at which they should operate. Thus, nuclear bars have been searched for in active galaxies because they were needed to provide the second step to transport gas from the circumnuclear region to the AGN (Shlosman et al. 1989; Friedli & Martinet 1993; Combes 1994; Heller & Shlosman 1994). As stated by Maiolino et al. (2000), the interpretation of the data depends on both the spatial resolution and the colours used to detect nuclear bars. In addition to this, the interpretation of the data within numerical model predictions is not straightforward. Regan & Mulchaey (1999) and Martini & Pogge (1999) indeed searched for straight dust lanes, considered as tracers of nuclear bars; however, the gas morphology strongly depends on the bar pattern speed, with straight shocks occurring only when the bar is rapidly rotating (Maciejewski & Sparke 1999). Therefore, slow rotating nuclear bars could also lead to the “spiraling” structure described by Regan & Mulchaey (1999) and Martini & Pogge (1999), therefore increasing the number of Seyfert galaxies hosting nuclear bars.
For Seyfert galaxies with no nuclear bar, different mechanisms have to be invoked, such as the presence of nuclear spirals (Martini & Pogge 1999; Englmaier & Shlosman 2000). Recent numerical simulations are beginning to predict an efficient enough transport of matter to the center in spiral galaxies with only a primary bar (Maciejewski, private communication), as should be the case for barred Seyferts with no nuclear bar.
The detailed analysis of such nuclear features requires in any case, in addition to the morphological information, a full kinematic characterization. In particular, in the case of NGC 6951 (Pérez et al. 2000) we have shown that no nuclear bar is detected in the HST images, but that there seems to be a nuclear spiral; this spiral structure is most probably residing in a nuclear disk, kinematically decoupled from the large-scale disk of the galaxy; the strong molecular gas accumulation could have destroyed a pre-existing nuclear bar and eventually may result in the dilution of the primary bar.<sup>2</sup><sup>2</sup>2We note that the question of gas concentration cannot be addressed through molecular gas mapping since at this stage, no CO maps are available for the remaining galaxies in our samples. The kinematic characterization of the central regions is expected to shed light on the question of how the central kinematics are related to the fuelling mechanisms in active galaxies. In the same fashion as for NGC 6951, we are now analyzing in detail the morphological and kinematical properties of the galaxies in our sample.
###### Acknowledgements.
I. Márquez acknowledges financial support from the Spanish Ministerio de Educación y Ciencia (EX94-8826734). This work is financed by DGICyT grants PB93-0139, PB96-0921, PB98-0521 and PR95-329. Financial support to develop the present investigation has been obtained through the Junta de Andalucía, the French-Spanish grants HF1996-0104 and HF1998-0052, from the Picasso program of the French Ministry of Foreign Affairs, and from the Chilean-Spanish bilateral agreement CSIC-CONICYT 99CL0018. We also acknowledge financial support from INSU-CNRS for several observing trips.
|
warning/0006/cond-mat0006262.html
|
ar5iv
|
text
|
# Phase diagram of an Ising model with long-range frustrating interactions: a theoretical analysis.
## I Introduction
There are many physical examples in which a short-ranged tendency to order is opposed by a long-range frustrating interactions: In diblock copolymers formed by two mutually incompatible polymer chains attached to each other, the repulsive short-range forces between the two types of components tend to induce phase separation of the melt, but total segregation is forbidden by the covalent bonds that link the subchains together. A microphase separation transition occurs instead at low enough temperature, and the system then forms phases with a periodically modulation of structures rich in one component or the other, such as lamellar, hexagonal, or cubic phases. In a similar way, self-assembly in water-oil-surfactant mixtures results from the competition between the short-range tendency of water and oil to phase separate and the stoichiometric constraints generated by the presence of surfactant molecules, constraints that act as the electroneutrality condition in a system of charged particles . The same kind of physics also arises in quite different fields. For instance, stripe formation in doped antiferromagnets like cuprates has been ascribed to a frustrated electronic phase separation, by which a strong local tendency of the holes to phase separate into a hole-rich “metallic” phase and a hole-poor antiferromagnetic phase is prohibited by the long-range Coulombic repulsion between the holes . A last example is provided by the structural or topological frustration in glass-forming liquids: the dramatic slowing down of the relaxation that leads to the glass formation has been interpreted as resulting from the presence of frustration-limited domains whose formation comes from the inability of the locally preferred arrangement of the molecules in the liquid to tile space periodically; topological frustration may also lead to low-temperature defect-ordered phases such as the Frank-Kasper phases in bimetallic systems.
A coarse-grained description of the above mentioned situations<sup>*</sup><sup>*</sup>*additional examples include cross-linked polymer mixture, interpenetrating networks, ultra-thin films involves lattice or continuum models with competing short-range and Coulombic interactions. The purpose of the present work is to study the phase diagram of such a model, namely the Coulomb frustrated Ising ferromagnet in which Ising spins placed on a 3-dimensional cubic lattice interact via both nearest-neighbor ferromagnetic couplings and long-range Coulomb-like antiferromagnetic terms. the model is introduced in more detail in section II and its ground states as a function of the frustration parameter, i.e. of the ratio of the antiferromagnetic coupling strength over the ferromagnetic one, is studied in section III. In section IV, we investigate the finite-temperature phase diagram in the mean-field approximation. Finally, the effect of the long-range nature of the frustrating forces, when comparing the phase behavior of the prototypical model with competing, but short-ranged interactions, the axial next-nearest neighbor Ising (ANNNI) model, as well as the limitations of the mean-field approach are discussed in section V.
## II The Coulomb frustrated Ising ferromagnet
The model is described by the Hamiltonian
$$H=J\underset{<ij>}{}S_iS_j+\frac{Q}{2}\underset{ij}{}v(𝐫_{ij})S_iS_j,$$
(1)
where, $`J,Q>0`$ are the ferromagnetic and antiferromagnetic coupling strengths, $`S_i=\pm 1`$ are Ising spin variables placed on the sites of a three-dimensional cubic lattice, $`<ij>`$ denotes a sum restricted to nearest neighbors, $`𝐫_{ij}`$ is the vector joining sites i and j, and $`v(𝐫)`$ represents a Coulomb-like interaction term with $`v(𝐫)_{}\frac{1}{|𝐫|}`$when $`|𝐫|\mathrm{}`$.(In all the paper, the lattice spacing is taken as the unit length).
In addition to considering the true Coulombic term, $`v(𝐫)=\frac{1}{|𝐫|}`$, we have also studied, for mathematical convenience in the analytical treatment, an expression of $`v(𝐫)`$ in terms of the lattice Green’s function that satisfies the Poisson equation on the three-dimensional cubic lattice. In the latter case, $`v(𝐫)`$ is then simply given, up to a multiplying factor of $`4\pi `$, as the inverse Fourier transform of the inverse lattice Laplacian,
$$v(𝐫)=\frac{4\pi }{N}\underset{𝐤}{}\frac{\mathrm{exp}(i\mathrm{𝐤𝐫})}{2_{\alpha =x,y,z}(1\mathrm{cos}(k_\alpha ))},$$
(2)
where $`N`$ is the number of lattice sites and the sum over $`𝐤=(k_x,k_y,k_z)`$ is restricted to the first Brillouin zone. For large $`|𝐫|`$ the lattice Green’s function behaves as $`1/(4\pi |𝐫|)`$ so that the expression in Eq. (2) has the proper asymptotic behavior. In practice, even at the next-neighbor distance, the difference between the true Coulombic form and Eq. (2) is very small . One has however to be careful about two points. The first one is that the $`v(𝐫)`$ defined in Eq. (2) has a nonzero, finite value at $`𝐫=0`$, $`v(\mathrm{𝟎})=0.25273100986\mathrm{}`$, a value that must of course be excluded when considering the Hamiltonian in Eq. (1).The correction term involving $`v(\mathrm{𝟎})`$ in Eq. (4) has been omitted in previous papers , but it leads only to very small corrections when the parameter $`Q/J`$ is small. For instance, Eq. (1) when expressed in Fourier space, can be written as
$$H=\frac{J}{2}\underset{𝐤}{}\widehat{V}(𝐤)|\widehat{S}(𝐤)|^2$$
(3)
where
$$\widehat{V}(𝐤)=2\underset{\alpha =x,y,z}{}\mathrm{cos}(k_\alpha )+\frac{4\pi Q}{J}\left[\frac{1}{2_{\alpha =x,y,z}(1\mathrm{cos}(k_\alpha ))}v(\mathrm{𝟎})\right]$$
(4)
and $`\widehat{S}(𝐤)`$ is the lattice Fourier transform of the Ising spin variable $`S_i`$.
To assess the quantitative difference between the Coulombic form and that involving the lattice Green’s function, we have calculated the energy of the system in several periodic configurations of the spins. For periodic lamellar patterns of large width $`m`$, the difference is negligible, but it increases when $`m`$ decreases. For a Néel antiferromagnetic state, the Coulombic energy is related to the Madelung constant and is equal to $`0.873782\mathrm{}`$, whereas the lattice Green’s function expression gives $`1.064356992`$. The second point worth mentioning is that the $`v(𝐫)`$ expressed in terms of the lattice Green’s function has the same discrete symmetry as the nearest-neighbor ferromagnetic interaction whereas the true Coulombic form has continuous rotational symmetry instead. The consequences are negligible for the ground-state and mean-field analyses, but may become important for the finite-temperature behavior of the system in other approximations.
## III Phase diagram at zero temperature
In the absence of the antiferromagnetic interaction ($`Q=0`$) the model reduces to the standard Ising ferromagnet, and the ground state of the system is obtained when all spins are aligned in a ferromagnetic state. Oppositely, when the ferromagnetic interaction is set to zero ($`J=0`$), the model is equivalent to a Coulomb lattice gas and the ground-state of the system is a Néel antiferromagnetic state. When $`Q0`$, the Coulombic interaction prevents the existence of a ferromagnetic phase, and in the thermodynamic limit, the total magnetization (charge) is constrained to be zero. Instead, phases with modulated order, i.e., with periodic patterns of “up” and “down” spins subject to the constraint of zero magnetization, are formed. We have studied these phases both analytically with the long-range interactions modeled by the lattice Green’s function and numerically with the true Coulombic form.
For small values of the frustration parameter $`Q/J`$, the ground state consists of lamellar phases in which parallel planes of ferromagnetically aligned spins form a periodic structure along the orthogonal direction. The system in such state can mapped onto a finite one-dimensional system of length $`2m`$ where $`m`$ denotes the width of the lamellae.
The short-range ferromagnetic contribution to the energy per spin of a lamellar phase can be readily calculated, and one gets
$$E_{SR}=J\left(3\frac{2}{m}\right).$$
(5)
The Coulombic energy due to the long-range competing forces can be calculated in reciprocal space by using the expression in terms of the inverse lattice Laplacian (see section II). For a lamellar phase of period $`2m`$ the wave-vectors to be considered have only one nonzero component that takes the values $`k=\pi (2n+1)/m`$ with $`0n<m1`$. Correspondingly, the lattice Fourier transform of the Ising spin variable, $`\widehat{S}(𝐤)=\frac{1}{\sqrt{N}}_{i=1}^NS_i\mathrm{exp}(i\mathrm{𝐤𝐫}_i)`$, where N is the total number of lattice site, has its modulus given by
$$|\widehat{S}(𝐤)|=|\widehat{S}(k)|=\frac{\sqrt{N}}{m\mathrm{sin}(k/2)}.$$
(6)
Using then the identities
$`{\displaystyle \underset{n=0}{\overset{m1}{}}}{\displaystyle \frac{1}{\mathrm{sin}(\pi (2n+1/(2m))^2}}`$ $`=`$ $`m^2,`$ (7)
$`{\displaystyle \underset{n=0}{\overset{m1}{}}}{\displaystyle \frac{1}{\mathrm{sin}((\pi (2n+1/(2m))^4}}`$ $`=`$ $`{\displaystyle \frac{m^4}{3}}+{\displaystyle \frac{2m^2}{3}},`$ (8)
one finds that the sum rule $`_k|\widehat{S}(k)|^2=N`$ is properly satisfied and that the Coulombic energy per site, $`E_c`$, is equal to
$`E_c`$ $`=`$ $`Q\left[{\displaystyle \underset{n=0}{\overset{m1}{}}}{\displaystyle \frac{\pi }{2\mathrm{sin}(\pi (2n+1)/(2m))^4}}2\pi v(0)\right]`$ (9)
$`=`$ $`Q\left[{\displaystyle \frac{\pi m^2}{6}}+\pi \left({\displaystyle \frac{1}{3}}2v(0)\right)\right].`$ (10)
In appendix A, we show that the same expression for the Coulombic energy is obtained when performing the calculation in real space with an effective one-dimensional potential, this latter includes a convergence factor that helps handling conditionally convergent sums appearing as a result of the long-range nature of the forces and that is taken to zero at the end of the calculation. It is worth mentioning that the above calculation in reciprocal space implicitly assumes that the contribution from the $`k=0`$ term is zero; physically, this means that the periodic system is embedded by a medium with infinite dielectric constant. (For a more detailed discussion, see Ref.) Since for simulations on ionic systems, a similar choice is generally adopted, both numerical and analytical calculations for our model have been performed using such metallic boundary conditions.
Figure. 1 shows the total energy per spin, $`E/J`$, which is the sum of the two contributions in Eq. (5) and (10) for the inverse lattice Laplacian expression, as a function of $`Q/J`$ for small values of frustration parameter $`Q/J`$. (The plot for the true Coulombic potential is very similar.) The slope of each straight line corresponds to the Coulombic energy and the intercept with the y-axis corresponds to the short-range ferromagnetic energy. For a given frustration parameter $`Q/J`$, the ground state is given by the straight-line that has the smallest energy. When $`Q/J0`$, the ground state consists of lamellar phases whose period becomes larger and larger and whose range of stability decreases. The values of $`(Q/J)`$ corresponding to transitions between two successive ground-state structures are obtained by solving the equation $`E(m)=E(m+1)`$. Therefore, when $`Q/J`$ goes to zero, $`m`$ is asymptotically given by
$$m(Q/J)^{1/3},$$
(11)
a behavior that is analogous to that observed for lamellar phases of diblock copolymer systems at low temperature in the so-called strong segregation limit. Note that for a large width $`m`$, the difference between Eq. (10) and the value obtained fro the true Coulombic term is only weakly varying with $`m`$ and can be taken as a constant, so that Eq. (11) provides also a very accurate estimate both for the sequence of lamellar phases and for the transition values of $`(Q/J)`$ for the true Coulombic potential.
When $`Q/J`$ increases, the period of the lamellar phases decreases until one reaches $`m=1`$ for $`Q/J=0.637`$. For a region of the frustration parameter between $`0.637`$ and $`1.800`$, this lamellar phase is then the most stable phase. In a narrow interval between $`Q/J=1.800`$ and $`Q/J=2.122`$ there is a cascade of phases $`(1\times m_2\times \mathrm{})`$ with $`m_2`$ decreasing until $`m_2=2`$ as the frustration parameter increases (see Fig. 2a). For $`2.122<Q/J<3.820`$, the stable phase is $`(1\times 2\times \mathrm{})`$. Note that tubular phases of type $`(m_1\times m_2\times \mathrm{})`$ with both $`m_1`$ and $`m_2>1`$ are never stable. For $`3.820<Q/J<6.237`$, the ground state is a $`1\times 1\times \mathrm{}`$ tube. In another interval $`6.237<Q/J<6.611`$, the system looses translational invariance in the third direction, and one observes a cascade of “parallepipedic” phases $`(1\times 1\times m_3)`$, with $`m_3`$ decreasing as the frustration increases until $`m_3=2`$. Between $`6.611`$ and $`9.549`$, the stable phase is $`(1\times 1\times 2)`$. For $`Q/J>9.549`$, the ground state is the standard Néel state $`(1\times 1\times 1)`$ (see Fig. 3a). It is noticeable that periodic structures of the type $`(m_1\times m_2\times m_3)`$ with $`m_1,m_2,m_3>1`$ and at least of the $`m_\alpha `$’s finite always have higher energies than those of the sequences $`(m_1\times \mathrm{}\times \mathrm{})`$, $`(\mathrm{}\times m_2\times \mathrm{})`$, $`(\mathrm{}\times \mathrm{}\times m_3)`$, whatever the value of the frustration parameter (see Appendix B).
The same analysis can be repeated with $`v(𝐫)`$ given by the true Coulomb interaction. The exact same sequence of ground states as before is obtained, but the values of the frustration parameter at the transition points are somewhat shifted: For instance, the lamellar phase $`m=1`$ is the most stable phase when $`0.627<Q/J<5.21`$, the cascade of “tubular” phases $`(1\times m_2\times \mathrm{})`$ occurs around $`Q/J=5.22`$, and $`(1\times 2\times \mathrm{})`$ is stable for $`5.23<Q/J<6.17`$ (see Fig. 2b). The cascade of “orthorhombic” phases $`(1\times 1\times m_3)`$, with $`m_3>1`$, appears around $`Q/J=14.63`$ (See Fig. 3b). The standard Néel state is stable for $`Q/J>15.33`$. It is worth mentioning that the counterparts of the lamellar, tubular, and orthorhombic phases in diblock copolymer systems (systems that are described at a coarse-grained level by the scalar field theory associated with the Hamiltonian in Eq. (1)) are the lamellar, columnar, and cubic phases, respectively.
In the region of stability of the lamellar phases, we have also investigated if more complex structures involving several one-dimensional modulations of ferromagnetically ordered layers could be present and if multiphase transition points at which more than simply two phases coexist could occur. For both questions, the answer is negative. As an example, we give in Appendix C the analytical expressions for the energy of the mixed lamellar phases that are formed by mixing lamellae of width $`m=1`$ and lamellae of width $`m=2`$ with some given periodic modulation. The simplest of such phases, that following the notation used by Fisher and Selke in their study of the axial next-nearest-neighbor Ising (ANNNI) model we denote $`<1^n2^p>`$, consists of a periodic repetition of a fundamental pattern formed by a succession of $`n`$ lamellae of width $`1`$ followed by $`p`$ lamellae of width $`2`$. (When $`n=0`$, one recovers the simple lamellar phase of width $`2`$, denoted $`<2>`$, and when $`p=0`$, one recovers the simple lamellar phase of width $`1`$, denoted $`<1>`$). It is easy to check that such mixed lamellar phases are never stable at zero temperature. In particular, they have a (strictly) higher energy than the pure lamellar phases at the zero-temperature transition point between the $`<1>`$ and the $`<2>`$ phases, at $`Q/J=2/\pi `$, so that this latter is a simple two-phase coexistence point. This conclusion remains unchanged when one considers even more complex mixed lamellar phases, such as $`<1^{n_1}2^{p_1}1^{n_2}2^{p_2}\mathrm{}1^{n_s}2^{p_s}>`$, with $`n_\alpha ,p_\alpha `$ integers for $`\alpha =1,2,\mathrm{}s`$, whose fundamental period is formed by $`n_1`$ $`1`$ layer lamellae followed by $`p_1`$ $`2`$ layer lamellae, then by $`n_2`$ $`1`$ layer lamellae, and by $`p_2`$ $`2`$ layer lamellae, etc…, two successive lamellae being composed of spins of opposite signs. We have also studied one-dimensional quasi-crystalline arrangements of lamellae: by using binary substitution rules, we have built quasi-periodic structures by iteration, but we have always found (numerically) that their energy is higher than that of the pure lamellar phases. Note finally that from the calculation of the energy of the mixed lamellar phases at zero temperature, one can also study the change of energy induced by adding defects in the pure lamellar phase $`<2>`$. As shown in Appendix C 2, adding defects always cost energy, the dominant effect being the increase in the short-range energy. The same analysis, leading to similar conclusion, can be repeated for the whole range $`0<Q/J<0.637`$ over which lamellar phases are favored.
## IV Mean-field theory
To describe phases with a spatial modulation of the magnetization, we consider a local mean-field approximation. If $`m_i=<S_i>`$ denotes the local magnetization at site $`i`$ and $`\widehat{m}(k)`$ its lattice Fourier transform, the mean-field free energy $`F_{mf}`$ is given by:
$$\frac{\beta F_{mf}}{N}=\frac{\beta J}{2N}\underset{𝐤0}{}\widehat{V}(𝐤)|\widehat{m}(𝐤)|^2\frac{1}{N}\underset{i}{}\mathrm{ln}(2\mathrm{cosh}(\beta H_i)),$$
(12)
where $`\beta =1/k_BT`$ and the effective field on site $`i`$, $`H_i`$, is equal to $`J_{ji}V_{ij}m_j`$, or in Fourier transformed space,
$$\widehat{H}(𝐤)=J\widehat{V}(𝐤)\widehat{m}(𝐤)$$
(13)
Minimization of the free energy with respect to the local magnetizations leads to the self-consistent set of equations:
$$m_i=\mathrm{tanh}(\beta H_i),$$
(14)
for each lattice site. One must then solve the coupled equations, Eqs.(13) and (14) simultaneously, and subsequently insert the solution in the expression of the free energy, Eq. (12). One finally searches for the configuration of the $`m_i`$’s that leads to the deepest minimum of the free energy for a given temperature and a given value of the frustration parameter.
To simplify the notation in the rest of the paper, units of temperature, energy, etc…, will be chosen such that $`k_B=J=1`$.
### A Order-disorder transition line
Close to the transition between the disordered and the ordered phases, the magnetizations $`m_i`$ are small and Eq. (14) can be linearized,
$$m_i\frac{H_i}{T}.$$
(15)
For a given value of the frustration parameter, the critical temperature $`T_c(Q)`$ is then given by
$$T_c(Q)=\mathrm{min}_𝐤\widehat{V}(𝐤)=V_c(𝐐),$$
(16)
where the minimum of $`\widehat{V}(𝐤)`$, $`V_c(𝐐)`$, is attained for a set of nonzero wave-vectors $`\{𝐤_c(Q)\}`$ that characterize the ordering at $`T_c(Q)`$. For the inverse lattice Laplacian expression of the long-range frustrating interaction, the $`𝐤_c(Q)^{}`$s vary continuously with $`Q`$ as follows
$`𝐤_c=(\pm \mathrm{arccos}(1\sqrt{\pi Q}),0,0),`$ for $`0Q<4/\pi ,`$ (17)
$`𝐤_c=(\pi ,\pm \mathrm{arccos}(3\sqrt{\pi Q}),0),`$ for $`4/\pi Q<16/\pi ,`$ (18)
$`𝐤_c=(\pi ,\pi ,\pm \mathrm{arccos}(5\sqrt{\pi Q})),`$ for $`16/\pi Q<36/\pi ,`$ (19)
$`𝐤_c=(\pi ,\pi ,\pi ),`$ for $`36/\pi Q.`$ (20)
One should of course add all vectors obtained by permuting the $`x,y,z`$ coordinates in Eqs. (17)-(20). The above ordering wave-vectors correspond, respectively, to lamellar (Eq. (17)), tubular (Eq. (18), orthorhombic (Eq. (19)), and cubic or Néel (Eq. (20)) phases. The corresponding critical temperature $`T_c(Q)`$ is then given by (recall that both $`T`$ and $`Q`$ are expressed in units of $`J`$)
$`T_c(Q)=64\sqrt{\pi Q}+4\pi Qv(0)`$ for $`0Q36/\pi `$ (21)
$`T_c(Q)=6+4\pi Q\left(v(0){\displaystyle \frac{1}{12}}\right)`$ for $`36/\pi Q`$ (22)
The mean-field approximation gives a line of second-order phase transition from the disordered to the modulated phases, with $`T_c(Q)`$ first decreasing with $`Q`$, reaching a minimum for $`Q_{min}=1/(4\pi v(0)^2)=1.245871`$ at $`T_c(Q_{min})=61/v(0)`$ and then increasing again for finally reaching a regime of linear increase with $`Q`$ when $`Q36/\pi `$. It is worth noting than the term $`4\pi Qv(0)`$ is important for large frustration: indeed, since $`v(0)>1/6`$, it allows to obtain a positive critical temperature for all $`Q^{}`$s. (In general, $`v(0)`$ should be larger than the inverse of the number of nearest neighbors on the lattice, e.g.. $`6`$ on a simple cubic lattice). For vanishing small frustrations, the critical temperature goes continuously to $`T_c^0`$, the critical temperature of the pure Ising ferromagnet.
### B Structure combination branching processes
At zero temperature, we showed that the system exists in pure modulated phases, whose modulation is commensurate with the underlying lattice. At the transition between the modulated and disordered phases, we have just seen that the ordering wave-vector varies continuously with the frustration parameter, hence indicating that a succession of incommensurate modulated phases is observed along the critical line. As the temperature increases from $`T=0`$ to $`T_c(Q)`$, one expects a cascade of ordered phases with commensurate spatial modulations of increasing complexity until a point is reached at which incommensurate phases appears; this is what is observed for instance in the much studied axial next-nearest neighbor Ising (ANNNI) model, in which the cascade of phases is produced by “structure combination branching processes”. To illustrate how such branching processes proceed, we consider first the low-temperature region of the phase diagram in which the simplest modulated phases, the $`<1>`$ and $`<2>`$ lamellar phases, are stable (see Fig. 4). Eqs. (13) and (14) must now be solved beyond the linear approximation. The $`<1>`$ phase corresponds to an alternate configuration of layers of up and down spins, with $`m_i=m_1(1)^i`$, which implies that the wave-vector characterizing the modulation has only one nonzero component equal to $`k=\pi `$. Using Eq. (13) and Eq. (14), one gets
$$m_1=\mathrm{tanh}\left[\frac{m_1}{T}(2\pi Q+4\pi Qv(0))\right],$$
(23)
which has a nonzero solution when
$$T<2\pi Q+4\pi Qv(0).$$
(24)
From Eq. (12), the corresponding free energies of the two phases is obtained as
$$\frac{F_{<1>}}{N}=\left(1\frac{\pi Q}{2}+2\pi Qv(0)\right)m_1^2T\mathrm{ln}\left[2\mathrm{cosh}\left(\frac{m_1}{T}(2\pi Q+4\pi Qv(0))\right)\right].$$
(25)
The $`<2>`$ phase corresponds to an alternate sequence of pairs of ferromagnetically ordered layers and is characterized by a wave-vector whose only nonzero component is $`k=\frac{\pi }{2}`$. The corresponding order parameter $`m_2`$ satisfies the following equation:
$$\frac{\sqrt{2}}{2}m_2=\mathrm{tanh}\left[\frac{\sqrt{2}m_2}{T}(2\pi Q+2\pi Qv(0))\right],$$
(26)
which has a nonzero solution when
$$T<2\pi Q+2\pi Qv(0).$$
(27)
The associated free energy is
$$\frac{F_{<2>}}{N}=\left(1\frac{\pi Q}{2}+\pi Qv(0)\right)m_2^2T\mathrm{ln}\left[2\mathrm{cosh}\left(\frac{m_2\sqrt{2}}{T}(2\pi Q+2\pi Qv(0))\right)\right].$$
(28)
The line of first-order transition at which the two phases coexist is defined by $`F_{<1>}=F_{<2>}`$. In the $`TQ`$ phase diagram, this line is almost vertical with $`Q2/\pi `$. As we have already stressed, no mixed phases coexist with the $`<1>`$ and $`<2>`$ phases at $`T=0`$ and $`Q=2/\pi `$. However, the mixed $`<12>`$ phase may become more stable at a nonzero temperature. This phase is characterized by the modulation $`m_i=\frac{2\sqrt{3}}{3}m_{12}\mathrm{cos}(2\pi i/3+\pi /2)`$, with the order parameter $`m_{12}`$ determined through the self-consistent equation
$$m_{12}=\mathrm{tanh}\left[\frac{m_{12}}{T}\left(3\frac{4\pi }{3}Q+4\pi Qv(0)\right)\right],$$
(29)
which has a nonzero solution for
$$T<3\frac{4\pi }{3}Q+4\pi Qv(0).$$
(30)
The corresponding free energy is given by
$$\frac{F_{<12>}}{N}=\frac{1}{3}\left(3\frac{4\pi Q}{3}\pi Qv(0)\right)m_{12}^2\frac{2T}{3}\mathrm{ln}\left[2\mathrm{cosh}\left(\frac{m_{12}}{T}(3\frac{4\pi }{3}Q+4\pi Qv(0))\right)\right].$$
(31)
Comparing now the free energy of the three phases, $`<1>`$,$`<2>`$, $`<12>`$, one can show that the $`<12>`$ phase become more stable than the other two in a wedge above a branching point at $`T_b=1.03`$ and $`Q=0.63`$, at which the three phases coexist. This represents the first step of a structure combination branching process by which two adjacent phases, here $`<1>`$ and $`<2>`$, get separated above a given branching point at finite temperature by a phase corresponding to the simplest combination structure, here the $`<12>`$ phase. A careful examination of the thermodynamic quantities shows that the entropy of the $`<12>`$ phase increases more rapidly with temperature than that of the $`<1>`$ and $`<2>`$ phases; although it has an unfavorable energy contribution, the $`<12>`$ phase becomes thermodynamically stable when the temperature becomes high enough ($`T_b=1.03`$). Moreover, for all temperatures between $`T=0`$ and $`T_b`$, both the entropy and the energy of the $`<1>`$ and $`<2>`$ phases are identical along the coexistence line between the two phases.
To study in more detail the branching processes, one can no longer rely on explicit analytical calculations as done above, because they become rapidly intractable. To obtain the phase diagram in a more systematic way , the mean-field equations, Eqs. (12)- (14), can be solved iteratively for finite lattices with periodic boundary conditions. The thermodynamically stable solution corresponds to that with the smallest free energy. For each temperature it is assumed that the spin structure repeats itself after $`N`$ layers (only commensurate lamellar structures are considered). The iterative procedure converges whenever the initial configuration is not too far from the equilibrium one. The iterative sequence is as follows. The effective fields $`H_i`$ are calculated from the set of initial magnetizations $`\{m_i^0\}`$ via Eq. (14). One computes the Fourier transform of the fields $`H_i`$, and using Eq. (13) yields then a new set of magnetizations $`\{m_i^1\}`$ which is used as the input for the next iteration. The calculation should be performed for various values of $`N`$ for examining many different commensurate structures. We have carried out the calculation for N up to 16. The phases with large-N periodicities, like in the ANNNI phase diagram, are only stable in a small neighborhood of the critical line. Therefore, the most of the phase diagram, except in the region close to the critical line, can be drawn by considering simple commensurate wave-vectors.
The results are shown in Figs. 4 and 5. Figure 4 illustrates the structure combination process in the region between the $`<1>`$ and $`<2>`$ phases. The $`<12>`$ and $`<1>`$ phases are separated by the $`<1^22>`$ phase which is itself separated at higher temperature from the $`<1>`$ phase by the $`<1^32>`$ phase, and so on, until one presumably reaches an accumulation point of the branching process along the transition line above which the $`<1>`$ phase melts. Beyond this accumulation point corresponding to a sequence of $`<1^n2>`$ phases when $`n\mathrm{}`$, devil’s staircases and incommensurate phases are expected. Figure 5 provides a broader picture of the phase diagram in the region where the modulated phases are lamellar: the phase diagram appears as a succession of flowers of complex modulated phases, separated by regions in which the pure lamellar phases $`<1>`$, $`<2>`$, etc…, are stable; the flowers get closer and closer when the frustration decreases. Note that the first branching point at which the simplest mixed phase appears is always at a non-zero temperature.
### C Soliton approach
At high enough temperature, near the critical line, the modulated phases are incommensurate since we saw in section IV A that the ordering wave-vector varies continuously with frustration. More insight in the phase behavior can be then provided by employing the soliton approach developed by Bak and coworkers. More precisely, this approach allows one to study analytically the melting of a commensurate phase to incommensurate phases by focusing on the behavior of the domain walls that separate commensurate regions, domain walls that can be considered as “solitons”. In the following, we use the soliton method to investigate the stability of commensurate phases at high temperature.
In the vicinity of the (upper) melting line of the $`<2>`$ phase, one can expand the mean-field equations in the appropriate order variables, which leads to the following free-energy functional
$`{\displaystyle \frac{F}{N}}={\displaystyle \frac{1}{2}}{\displaystyle \underset{𝐤}{}}[\widehat{V}(𝐤)+{\displaystyle \frac{1}{\beta }}]|m_𝐤|^2+{\displaystyle \frac{T}{12}}{\displaystyle \underset{𝝉}{}}{\displaystyle \underset{𝐤_\mathrm{𝟏}\mathrm{}𝐤_\mathrm{𝟒}}{}}\delta (𝐤_\mathrm{𝟏}+𝐤_\mathrm{𝟐}+𝐤_\mathrm{𝟑}+𝐤_\mathrm{𝟒}𝝉)\widehat{m}_{𝐤_\mathrm{𝟏}}\widehat{m}_{𝐤_\mathrm{𝟐}}\widehat{m}_{𝐤_\mathrm{𝟑}}\widehat{m}_{𝐤_\mathrm{𝟒}}`$ (32)
$`+{\displaystyle \frac{T}{30}}{\displaystyle \underset{𝝉}{}}{\displaystyle \underset{𝐤_\mathrm{𝟏}\mathrm{}𝐤_\mathrm{𝟔}}{}}\delta (𝐤_\mathrm{𝟏}+𝐤_\mathrm{𝟐}+𝐤_\mathrm{𝟑}+𝐤_\mathrm{𝟒}+𝐤_\mathrm{𝟓}+𝐤_\mathrm{𝟔}𝝉)\widehat{m}_{𝐤_\mathrm{𝟏}}\widehat{m}_{𝐤_\mathrm{𝟐}}\widehat{m}_{𝐤_\mathrm{𝟑}}\widehat{m}_{𝐤_\mathrm{𝟒}}\widehat{m}_{𝐤_\mathrm{𝟓}}\widehat{m}_{𝐤_\mathrm{𝟔}}+\mathrm{}`$ (33)
where a constant term has been discarded and $`𝝉`$ is a reciprocal-lattice vector. The above expression, Eq. (32), contains both regular and “umklapp” terms; these latter, represented by the second and third contributions in the right-hand side of Eq. (32), correspond to terms in which the sum of the wave-vectors is equal to a reciprocal-lattice vector, i.e., they keep track of the underlying lattice structure and are responsible for the stability of the commensurate phases.
For studying the stability of the $`<2>`$ phase near the critical line, we consider wave vectors that are close to the ordering wave vector $`k_{\pi /2}=(\pi /2,0,0)`$ with small fluctuations $`q_x`$ in the direction of the modulation, here along the $`x`$axis, and $`𝐪_{}`$ in the perpendicular layer. For the present case, it is sufficient to truncate Eq. (32) after the fourth order, and after expanding the interaction term to second order in the fluctuations the free-energy functional can be rewritten as
$`{\displaystyle \frac{F}{N}}={\displaystyle \underset{𝐪=q_x,𝐪_{},}{}}\left(r+ak_x+cq_x^2+c^{}𝐪_{}^2\right)|m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪}|^2+T{\displaystyle \underset{𝐪_\mathrm{𝟏}𝐪_\mathrm{𝟐}𝐪_\mathrm{𝟑}𝐪_\mathrm{𝟒}}{}}\delta (𝐤_\mathrm{𝟏}+𝐤_\mathrm{𝟐}+𝐤_\mathrm{𝟑}+𝐤_\mathrm{𝟒})`$ (34)
$`[{\displaystyle \frac{1}{2}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟏}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟐}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟑}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟒}}+{\displaystyle \frac{1}{12}}(m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟏}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟐}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟑}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟒}}`$ (35)
$`+m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟏}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟐}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟑}}m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪_\mathrm{𝟒}})],`$ (36)
where $`r=T4+2\pi Q4\pi Qv(0)0`$, $`a=22\pi Q`$, $`c=2\pi Q`$, and $`c^{}=\pi Q+1`$. Introducing now two continuous order parameters,
$`m_+(𝐫)`$ $`=`$ $`\sqrt{2}{\displaystyle \frac{d^3q}{(2\pi )^3}\mathrm{exp}(i\mathrm{𝐪𝐫})m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪}}`$ (37)
$`m_{}(𝐫)`$ $`=`$ $`\sqrt{2}{\displaystyle \frac{d^3q}{(2\pi )^3}\mathrm{exp}(i\mathrm{𝐪𝐫})m_{𝐤_{\pi /\mathrm{𝟐}}+𝐪}},`$ (38)
where $`m_{}(𝐫)`$ is the complex conjugate of $`m_+(𝐫)`$, one can express the free-energy functional in the following Ginzburg-Landau form:
$`{\displaystyle \frac{F}{N}}`$ $`=`$ $`{\displaystyle }d^3𝐫\left({\displaystyle \frac{1}{2}}c\right|({\displaystyle \frac{}{x}}i{\displaystyle \frac{a}{2c}})m_+|^2+{\displaystyle \frac{1}{2}}c^{}|_{}m_+|^2+{\displaystyle \frac{1}{2}}(r{\displaystyle \frac{a^2}{4c}})|m_+|^2`$ (39)
$`+`$ $`{\displaystyle \frac{T}{8}}|m_+|^4+{\displaystyle \frac{T}{8}}(m_+^4+m_{}^4)),`$ (40)
where $`_{}=(0,\frac{}{y},\frac{}{z})`$ and the last term is generated by the umklapp terms. Following Bak and Boehm, we choose the following ansatz for the order parameters,
$$m_\pm =A\mathrm{exp}(\pm i\varphi (x)),$$
(41)
where the amplitude $`A`$ is a constant. Note that $`\varphi (x)`$ is constant in the commensurate $`<2>`$ phase and that $`A`$ can be obtained by minimizing the free energy in that phase, which gives $`A^2=3|r|/T`$. Inserting the above expression in Eq. (39) leads, up to a constant, to the following free-energy functional per unit area perpendicular to the $`x`$direction:
$$F[\varphi ]=\frac{cA^2}{2}𝑑x\left[\varphi ^{}_{}{}^{}2(x)\frac{a}{c}\varphi ^{^{}}(x)\frac{TA^2}{12c}(1\mathrm{cos}4\varphi (x))\right],$$
(42)
where the first term is minimized for $`\varphi (x)=Ax/(2c)`$, which corresponds to an incommensurate spatial modulation, whereas the second term is minimized for $`\varphi (x)=0`$, i.e., in the commensurate $`<2>`$ phase. The overall minimization of $`F[\varphi ]`$ is attained when the phase function $`\varphi (x)`$ obeys the sine-Gordon equation:
$$\varphi ^{\prime \prime }(x)+4v\mathrm{sin}(4\varphi (x))=0,$$
(43)
where $`v=TA^2/(24c)`$. The solution consists of regions of constant phase separated by solitons in which $`\varphi `$ increases by $`\frac{\pi }{2}`$ over a short distance. More generally, one can lock for a solution over a (large) distance $`L`$ that consists of $`n`$ equally spaced solitons where $`\frac{n\pi }{2}=L\varphi ^{}`$. The corresponding free-energy is given by
$$\frac{F}{cA^2L}=\left[\frac{4}{\pi }v^{\frac{1}{2}}|\frac{a}{2c}|\right]\varphi ^{}+\frac{16}{\pi }v^{\frac{1}{2}}\varphi ^{}\mathrm{exp}[\frac{2\pi }{\varphi ^{}}v^{\frac{1}{2}}].$$
(44)
where $`\varphi ^{}=n\pi /(2L)`$. The first term, proportional to the soliton density, is the formation energy; the second term corresponds to a weak repulsion between solitons. The commensurate phase is stable as long as the first term remains positive; otherwise, the $`<2>`$ phase becomes unstable with the respect to soliton formation. The critical temperature corresponding to this melting transition is given by
$$T_{<2>IC}=42\pi Q+4\pi Qv(0)\frac{\pi }{4}\frac{(1\pi Q)^2}{Q}.$$
(45)
and is shown as a dashed line in Figs. 4 and 5.
A similar analysis can be performed to study, for instance, the stability of the $`<3>`$ phase near the critical line. The main changes are that the relevant ordering wave-vector is now $`(\pi /3,0,0)`$ and that the sixth-order umklapp terms should be kept in Eq. (32). A transition between the $`<3>`$ phase and incommensurate phases is found for
$$T_{<3>IC}=54\pi Q+4\pi Qv(0)\frac{3\pi }{4}|14\pi Q|\sqrt{\frac{5(54\pi Q)}{1+20\pi Q}}.$$
(46)
The result is shown as dotted curve in Fig. 5. The deviation from the numerical solution of the mean-field equations that is seen when moving further away from the critical line comes from the truncation of the free-energy functional that is used in the soliton approach.
## V Discussion
The fact that spin models with competing interactions can give rise to complex spatially modulated phases has been known for several decades. The ANNNI model is one of the simplest and best-studied such system, in which Ising spin variables situated on a lattice are coupled via nearest-neighbor ferromagnetic interactions in ($`D1`$)-dimensional layers orthogonal to, say, the $`x`$ axis and via next-nearest neighbor antiferromagnetic interactions along the $`x`$ axis. The major additional ingredient that is present in the Coulomb frustrated Ising ferromagnet and not is the previously studied models is the long-range nature of the competing antiferromagnetic interaction. It is then worth reviewing some of the differences between the phase behavior of the three-dimensional Coulomb frustrated ferromagnet and that of the three-dimensional ANNNI and related models. First, the long $`1/r`$ range of the frustrating interaction forbids ferromagnetic ordering for any nonzero value of the frustration parameter $`Q`$, and, as a result, there is no Lifshitz point in the model. Secondly, there is no highly degenerate multiphase point at zero temperature; this is in contrast with the ANNNI model that possesses an infinitely degenerate multiphase point at zero temperature, a point from which springs an infinite number of distinct commensurate modulated phases. Thirdly, the phase diagram of the Coulomb frustrated Ising ferromagnet, as illustrated in Fig. 5, displays an infinite number of distinct flowers of complex spatially modulated phases that emerge from zero-temperature two-phase coexistence points, the extent of the flowers in $`TQ`$ phase diagram decreasing as the frustration parameter decreases.
The long range of the antiferromagnetic interaction in the Coulomb frustrated ferromagnet also brings about additional limitations on the mean-field approach. On general grounds, one can expect the mean-field description to reproduce the topology of the phase diagram correctly, but to become increasingly inaccurate as one approaches the transition line, both from below and from above, because it overlooks the role of fluctuations. In the present case, the fluctuations have a major effect on the transition line: as argued by Brazovskii on the basis of a self-consistent Hartree treatment of a field-theoretical model analogous to the Coulomb frustrated Ising ferromagnet, and confirmed by Monte Carlo simulations, the fluctuations drive the order-disorder transition from second to first-order. The mean-field approach is thus questionable in the vicinity of the transition line (which is why it may not be worth pursuing the search for devil’s staircases as was done for the ANNNI model). However, the main points reviewed above are not affected.
## A Calculation of the Coulombic energy of lamellar phases in real space
The calculation of the ground-state energy due to the long-range Coulombic interaction can be performed as follows. For lamellar phases, the sum over the reciprocal vectors is performed along one direction. One then obtains the one-dimensional potential corresponding to the inverse lattice Laplacian,
$$W(i)=2\pi Q\frac{\mathrm{exp}(\alpha |i|)1}{\alpha },$$
(A1)
where $`\alpha `$ is a convergence factor that will taken to zero at the end of the calculation. The introduction of $`\alpha >0`$ in the calculation allows one to handle conditionally convergent sums. The average Coulombic energy per site of a cell $`2m`$ is given by
$$E_C=\frac{1}{4m}\underset{N^{}=\mathrm{}}{\overset{+\mathrm{}}{}}\underset{i,j}{\overset{2m}{}}W(|ij+2mN^{}|)S_iS_j,$$
(A2)
where $`N^{}`$ is the index for labelling the right and left cells and $`i,j`$ are denote sites within the cell. The energy per site can be divided into two parts: the first one comes from the interaction between the spins within the cell and the second one comes from the interaction between a spin and its images in the other cells; this reads
$`E_C`$ $`=`$ $`E_1+E_2`$ (A3)
$`=`$ $`{\displaystyle \frac{1}{4m}}{\displaystyle \underset{i,j}{\overset{2m}{}}}W(|ij|)S_iS_j+{\displaystyle \frac{1}{4m}}{\displaystyle \underset{N^{}0}{}}{\displaystyle \underset{i,j}{\overset{2m}{}}}W(|ij+2mN^{})S_iS_j.`$ (A4)
By considering all contributions between pairs of sites within the cell, one gets for $`E_1`$
$$E_1=\frac{1}{2m}\left(m(W(0)W(m))+\underset{n=1}{\overset{m1}{}}(2(mn)n)W(n)\underset{n=m+1}{\overset{2m1}{}}((2mn)n)W(n)\right).$$
(A5)
After some calculation, and taking the limit $`\alpha 0`$ at the end, one gets
$$E_1=Q\left(\frac{2\pi m^2}{3}+\frac{\pi }{3}\right).$$
(A6)
The sum over the right and the left cells can also be performed, and $`E_2`$ is then given by
$$E_2=\frac{\pi Q}{2m}\underset{i,j}{}\frac{e^{\alpha (2m+ij)}+e^{\alpha (2m+ji)}}{\alpha (1e^{\alpha 2m})}S_iS_j.$$
(A7)
By using the electro-neutrality condition ($`_iS_i=0`$), and in the limit $`\alpha 0`$, one obtains $`E_2`$, which has a finite value:
$`E_2`$ $`=`$ $`{\displaystyle \frac{\pi Q}{4m^2}}{\displaystyle \underset{i,j}{}}(ij)^2S_iS_j`$ (A8)
$`=`$ $`{\displaystyle \frac{\pi Q}{2m^2}}|{\displaystyle \underset{i=1}{\overset{2m}{}}}iS_i|^2`$ (A9)
$`=`$ $`{\displaystyle \frac{\pi Qm^2}{2}}.`$ (A10)
This gives for the Coulombic energy $`E_c`$
$$E_c=Q\left(\frac{\pi m^2}{6}+\frac{\pi }{3}\right).$$
(A11)
By subtracting the self-energy of the inverse lattice Laplacian potential to Eq. (A11), one exactly recovers Eq. (10).
## B Ground-state energy of tubular and orthorhombic phases
Let us calculate the energy per site for configurations whose phases are periodic with a orthorhombic cell ($`m_1\times m_2\times m_3`$). The short-range energy per spin is obtained as
$$E_{SR}=J\left(3\frac{2}{m_1}\frac{2}{m_2}\frac{2}{m_3}\right)$$
(B1)
In the reciprocal space, the allowed wave-vectors have components $`((2n_1+1)\pi /m_1,(2n_2+1)\pi /m_2,(2n_3+1)\pi /m_3)`$, with $`0n_1<m_11`$, $`0n_2<m_21`$, $`0n_3<m_31`$, so that the Coulombic energy per spin $`E_c`$ is given by
$$E_c=Q\left[\underset{n_1=0}{\overset{m_11}{}}\underset{n_2=0}{\overset{m_21}{}}\underset{n_3=0}{\overset{m_31}{}}\left(\frac{\pi }{(_{\alpha =1}^3\mathrm{sin}\left(\frac{(2n_\alpha +1)\pi }{2m_\alpha }\right)^2)_{\alpha =1}^3\mathrm{sin}\left(\frac{(2n_\alpha +1)\pi }{2m_\alpha }\right)^2}\right)2\pi v(0)\right]$$
(B2)
When both $`m_2`$ and $`m_3`$ go to infinity, one obtains lamellar phases of period $`2m`$, and Eq. (B2) reduces to Eq. (10).
When the periodic structure looses translational invariance in a second direction, one obtains a lattice of tubes whose perpendicular section is a rectangular cell $`m_1\times m_2`$. Although we have not obtained a fully analytical expression for such phases, some results can be derived. The Coulombic energy for tubes of section $`m_1\times m_2`$ is bounded as follows:
$$E_c(inf(m_1,m_2))E_c(m_1,m_2)E_c(sup(m_1,m_2)).$$
(B3)
If $`p\times m_1=m_2=m`$ where $`p`$ is a positive integer, the energy $`E_c(m)`$ behaves as
$$E_c(m^2)C(p)m^2+O(m),$$
(B4)
where the numerical coefficients $`C(p)`$ are summarized in table I.
We have also calculated numerically, via Eqs. (B1) and (B2) the total energy of modulated orthorhombic phases ($`m_1\times m_2\times m_3`$) when $`\mathrm{}>m_1,m_2,m_3>1`$. It always higher than that of the phases ($`1\times m_2\times m_3`$) or ($`1\times 1\times m_3`$).
## C Mixed Lamellar phases
The inverse Laplacian approximation allows to calculate exactly the energy of a large number of periodic structures at zero temperature. As an illustration we present here the results for mixed lamellar phases that are potential ground-state candidates in the region of $`Q/J`$ where the most stable among the simple lamellar phases involve lamellae of width $`m=1`$ (phase $`<1>`$) and $`m=2`$ (phase $`<2>`$).
### 1 The $`<1^n2^p>`$ phases.
The $`<1^n2^p>`$ phases are the simplest mixed phases that one can construct with the two elementary bricks formed by lamellae of width $`m=1`$ and $`m=2`$: they are formed by a periodic sequence of ferromagnetically aligned layers whose fundamental period consists of $`n`$ one-layer lamellae followed by $`p`$ two-layer lamellae, two successive lamellae being formed by spins of opposite signs. Because of the electoneutrality (zero magnetization) condition one has to distinguish three different families: ($`n=2q`$, $`p=2r`$), ($`n=2q+1`$, $`p=2r`$) and ($`n=2q`$, $`p=2r+1`$), where $`q`$ and $`r`$ are integers. The $`<1^{2q+1}2^{2r+1}>`$ phase is not allowed at zero temperature because they do not satisfy the electoneutrality condition. Because of the two-dimensional in-layer ferromagnetic ordering, the wave-vectors characterizing lamellar phases have only one nonzero component.
#### a The $`<1^{2q}2^{2r}>`$ phases.
The size of the one-dimensional unit cell is $`L=2q+4r`$ and the allowed values of the nonzero components of the wave vector are $`k=\frac{2\pi l}{2q+4r}`$ where $`l`$ is an integer such that $`l=0,\mathrm{},2q+4r1`$. Summing over all sites of the lattice, one finds
$$\widehat{S}(k)=\frac{\sqrt{N}}{2q+4r}\frac{\mathrm{sin}(2rk)}{\mathrm{cos}(k)cos(\frac{k}{2})}=\frac{\sqrt{N}}{2q+4r}\frac{\mathrm{sin}(qk)}{\mathrm{cos}(k)cos(\frac{k}{2})},$$
(C1)
if $`k\pm \frac{\pi }{2}`$ and $`k\pi `$. Otherwise, one obtains $`\widehat{S}(\pm \frac{\pi }{2})=\frac{\sqrt{N}}{2q+4r}\sqrt{8}r`$ and $`\widehat{S}(\pi )=\frac{\sqrt{N}}{2q+4r}2q`$.
Using the identities
$$\underset{l=1}{\overset{q+2r1}{}}\frac{\mathrm{sin}(\frac{q\pi l}{q+2r})^2}{\mathrm{cos}(\frac{\pi l}{q+2r})^2\mathrm{cos}(\frac{\pi l}{2(q+2r)})^2}=8r(r+q)$$
(C2)
and
$$\underset{l=1}{\overset{q+2r1}{}}\frac{\mathrm{sin}(\frac{q\pi l}{q+2r})^2}{\mathrm{sin}(\frac{2\pi l}{q+2r})^2}=r(r+q),$$
(C3)
one can express the Coulombic energy per spin as
$$E_c=Q\left[\pi \frac{2q^2+16r^2+16qr}{(2q+4r)^2}2\pi v(0)\right].$$
(C4)
The short-range contribution can be easily calculated and the total energy per spin is equal to
$$E_{<1^{2q}2^{2r}>}=J\left(2+\frac{q}{q+2r}\right)+Q\left(\pi \frac{2q^2+16r^2+16qr}{(2q+4r)^2}2\pi v(0)\right).$$
(C5)
It is now easy to show that the above energy, whatever the strictly positive values of $`q`$ and $`r`$ and whatever the value of $`Q/J`$, cannot be less than either the energy of the $`<1>`$ phase or that of the $`<2>`$ phase. Indeed, for the conditions $`E_{<1^{2q}2^{2r}>}E_{<1>}`$ and $`E_{<1^{2q}2^{2r}>}E_{<2>}`$ to be simultaneously satisfied, one must have
$$\frac{2q+4r}{\pi q}Q/J\frac{q+2r}{\pi (q+r)},$$
(C6)
which is impossible.
#### b The $`<1^n2^{2m}>`$ and $`<1^{2n}2^m>`$ phases.
To satisfy the requirement of the global electroneutrality, the unit cell of such phases is built as follows. For the $`<1^n2^{2m}>`$ phases, the unit cell is $`\underset{n}{\underset{}{\mathrm{}}}\underset{m}{\underset{}{\mathrm{}}}\underset{n}{\underset{}{\mathrm{}}}\underset{m}{\underset{}{\mathrm{}}}`$ and the nonzero components of the wave-vector are given by $`k=\frac{2\pi l}{(8m+2n)}`$ where $`l=1\mathrm{}8m+2n1`$. After some algebra, the Fourier transform of the spin variable $`\widehat{S}(k)`$ is obtained as
$$\widehat{S}(k)=\frac{\sqrt{N}}{2n+8m}\frac{2\mathrm{cos}(\frac{nk}{2})}{\mathrm{cos}(k)cos(\frac{k}{2})}$$
(C7)
if $`k\pi `$ and $`\widehat{S}(\pi )=\frac{\sqrt{N}}{2n+8m}2n`$. The total energy per spin is then
$$E_{<1^n2^{2m}>}=J\left(2+\frac{2n}{2n+8m}\right)+Q\left(\pi \frac{2n^2+64m^2+32mn}{(2n+8m)^2}2\pi v(0)\right).$$
(C8)
The $`<1^{2n}2^m>`$ phase is characterized by the sequence $`\underset{n}{\underset{}{\mathrm{}}}`$$`\underset{m}{\underset{}{\mathrm{}}}\underset{n}{\underset{}{\mathrm{}}}\underset{m}{\underset{}{\mathrm{}}}`$ and the nonzero components of the wave-vector are given by $`k=\frac{2\pi l}{(4m+4n)}`$ with $`l=1\mathrm{}4m+4n1`$. This leads to
$$\widehat{S}(k)=\frac{\sqrt{N}}{4n+4m}\frac{2\mathrm{sin}(nk)}{\mathrm{cos}(k)cos(\frac{k}{2})}$$
(C9)
if $`k\pm \frac{\pi }{2}`$, and $`\widehat{S}(\pm \frac{\pi }{2})=\frac{\sqrt{N}}{4n+4m}2\sqrt{2}n`$. The total energy per spin is finally given by
$$E_{<1^n2^{2m}>}=J\left(2+\frac{n}{n+m}\right)+Q\left(\pi \frac{32n^2+64m^2+32mn}{(4n+4m)^2}2\pi v(0)\right).$$
(C10)
As before, there is no range of the frustration parameter $`Q/J`$ for which these phases become ground states of the system.
### 2 Energy of defects in the $`<2>`$ phase.
The issue of the stability of the mixed versus simple lamellar phases can be addressed in a different way. Starting from the $`<2>`$ phase, one can calculate the energy change brought by inserting one or several defects of type $`<1>`$ within the periodic structure. The creation of one such defect results from the flip of a pair of up-down spins. This corresponds to the simplest excitation that one expects in the $`<2>`$ phase: $`\mathrm{}\mathrm{}\mathrm{}\mathrm{}`$ The resulting defective structure can also be viewed as a $`<1^42^{2m}>`$ phase where $`m\mathrm{}`$. The Coulombic energy of a $`<1^42^{2m}>`$ structure for a cell of size $`4+4m`$ is derived from Eq. (C4) and reads
$$e_{c<1^42^{2m}>}=(4+4m)E_{c<1^42^{2m}>}=Q\left[\pi \frac{2+4m^2+8m}{1+m}2\pi v(0)(4+4m)\right].$$
(C11)
If one subtracts the Coulombic energy of the $`<2>`$ phase for the same unit cell ($`l=4+4m`$) from the above equation, one obtains the Coulombic energy for one excitation in a $`<2>`$ phase. Strikingly, this energy goes to zero when $`m`$ goes to infinity, which means that the presence of one defect in the $`<2>`$ phase does not change at all the Coulombic energy of this phase. (This is not a result of the macroscopic limit.) Conversely, this defect has a short-range energy cost that is easily obtained as $`E_{SR}=4`$. In all the region of frustration parameters $`Q/J`$ where the $`<2>`$ phase is more stable than any other simple lamellar phase, the presence of one defect is then energetically unfavorable for the system.
We have also calculated the energy of a $`<2>`$ phase where two defects have been introduced. We have first calculated the energy of the $`<2^{2m}1^{2p}2^{2n}1^{2q}>`$ and $`<2^{2m+2n}1^{2p+2q}>`$ phases which are identical. Setting $`p=q=2`$ whereas $`n,m\mathrm{}`$, one obtains the energy of the $`<2>`$phase in the presence of two defects. We have also found that the introduction of the two defects is energetically unfavorable for the system for any value of the frustration. Although we have not obtained a general proof, the phase-locking into simple lamellar phases at zero temperature seems to be well established for the model.
|
warning/0006/astro-ph0006455.html
|
ar5iv
|
text
|
# Empirical Determinations of Key Physical Parameters Related to Classical Double Radio Sources
## 1 Introduction
Classical double radio sources, also known as FRII radio sources (Fanaroff & Riley, 1974), have been extensively studied. Numerous surveys of radio sources as well as detailed studies of individual sources have provided a wealth of radio data on many sources. Theoretical models for the sources developed and discussed by various authors (e.g. Blandford & Rees, 1974; Scheuer, 1974, 1982; Begelman, Blandford, & Rees, 1984; Begelman & Cioffi, 1989; Eilek & Shore 1989; Gopal-Krishna & Wiita 1991; Daly 1990, 2000) lend much understanding to the physics of the sources. It is believed that FRII sources are powered by highly collimated outflows from an active galactic nucleus (AGN), and that the radio emission from these sources is the result of an interaction between the beam, or jet, and the ambient medium. As a result, careful study of FRII sources can yield very useful information about the FRII sources, their environments, and their central engines (e.g. Daly 2000).
Daly (1994, 1995) presents a model for powerful extended FRII sources and showed how key parameters of an FRII source and its environment, such as the beam power and the ambient gas density, can be estimated using multi-frequency radio data. Results for a sample of powerful 3CR radio sources are also presented by the same author (Daly 1994, 1995). A larger sample was compiled and studied in detail; results from these studies have been presented in a series of papers (Wellman & Daly 1996a,b; Wellman 1997; Wellman, Daly, & Wan 1997a, b \[hereafter WDW97a,b\]; Wan & Daly 1998a,b; Guerra & Daly 1996, 1998; Guerra 1997; Daly, Guerra, & Wan 1998; Guerra, Wan, & Daly 1998; Wan 1998; Daly 2000). For example, detailed results on the density and temperature of the ambient medium of the FRII sources are presented by WDW97a,b, and the redshift evolution of the characteristic size of an FRII source and its application for cosmology are presented by Guerra & Daly (1996, 1998), and Guerra, Daly, & Wan (2000).
This paper presents results on several key parameters of the FRII sources and their environments that were not presented and discussed in previous papers. These parameters include the luminosity in directed kinetic energy of the jet, also known as the beam power, $`L_j`$, the total time the AGN will produce the collimated outflow, $`t_{}`$, the thermal pressure of the ambient medium, $`P_{th}`$, the effect of the hot gas on the microwave background radiation, and the total gravitational mass of the host cluster (we find that the radio sources lie at the centers of gas rich clusters of galaxies). The way that these parameters can be estimated from radio data, and the empirical results for the sample, are presented here.
The paper is structured as follows. A brief description of the sample and data for the sample are presented in §2. Discussions on the jet luminosity and timescale of jet activity are given in §3, the thermal pressure of the ambient medium and estimates of the effect of the hot gas on the microwave background radiation are presented in §4, and estimating the total gravitational mass of the the host cluster of the radio source is discussed in §5. Each of these sections are divided into two subsections, with the first subsection presenting theory and the second one presenting empirical results for the sample. Further discussions of the results and a summary of the paper are presented in §6.
Values of Hubble’s constant $`H_o=100h\text{km s}^1\text{Mpc}^1`$ and the de/acceleration parameter $`q_o`$=0 (an open empty universe with zero cosmological constant) are used to estimate all the parameters in this study. Different choices of cosmological parameters, within reasonable limits, give results that are very similar to, and consistent with, those presented here.
## 2 Sample Description and Data
The sample used in this study has been described in detail in several previous papers (e.g. WDW97a,b; Guerra & Daly 1998). The readers are referred to these papers for a full description of the sample; a brief description of it follows. All the FRII sources in this sample are very powerful 3CR sources, having 178 MHz radio power greater than $`3\times 10^{26}\text{ }h^2\text{ W Hz}^1\text{sr}^1`$; such powerful FRII sources are referred to as FRIIb sources (Daly 2000), or Type 1 FRII sources (Leahy & Williams 1984). They are drawn from the samples of Leahy, Muxlow, & Stephens (1989; hereafter LMS89), which contains sources with large angular sizes, and Liu, Pooley, & Riley (1992; hereafter LPR92), which contains sources with smaller angular extent. The final sample contains 27 radio lobes from 14 FRIIb radio galaxies, and 14 lobes from 8 FRIIb radio loud quasars. These sources have redshifts ranging from zero to 2, and projected core-hot spot separations between $`25h^1`$ kpc and $`200h^1`$ kpc. Note that the questions of projection effects and radio power selection effects have been addressed in great detail by Wan & Daly (1998a,b).
WDW97a,b used the radio maps from LMS89 and LPR92 to estimate the width of the radio bridge behind the radio hot spot, $`a_L`$, the non-thermal pressure of the radio lobe that drives the forward shock front, $`P_L`$, and the rate at which the bridge is lengthening, referred to as the lobe propagation velocity $`v_L`$. The radio information may also be used to estimate the beam power and total time the source will be active, as described in §3.
In order to estimate the Mach number of lobe advance, high quality radio maps which image large portions of the radio bridge are required; the radio bridge is defined as the radio emitting region that lies between the radio hot spot and the origin of the host galaxy or quasar. There are 16 bridges in the sample for which we have a detection of the Mach number of lobe advance, which includes 13 radio galaxy bridges and 3 radio loud quasar bridges. The Mach number may be combined with the lobe propagation velocity $`v_L`$ to estimate the temperature of the ambient gas (WDW97a). Thus, the ambient gas temperature and pressure may be estimated for these sources. Lower bounds on the Mach number, and hence upper bounds on the ambient gas temperature and pressure, are available for the 14 other bridges, including 8 galaxy bridges and 6 quasar bridges, as described in WDW97a,b. The maps of the sources do not have sufficient dynamic range to allow detailed analyses of their bridge structure, and thus do not have estimates on the Mach number of lobe advance, nor on the ambient gas pressure and cluster mass. As a result, the sample of sources with estimates of the ambient gas pressure and cluster mass is rather small.
Results for radio-loud quasars and radio galaxies are analyzed separately when there are enough data points, as is the case in the study of the beam power. As noted above, the number of sources with thermal pressure and cluster mass estimates is rather small (16 lobes with detections and 14 with bounds). Thus it is not practical to separate quasars and galaxies in the study of these parameters.
Cygnus A (3C 405) is the only source with a redshift $`0`$ in our sample, and in some cases, results seem to depend on whether or not this source is included. Thus, subsamples both with and without Cygnus A are analyzed, and the results included here. Note that Cygnus A is not used to normalize any of the quantities studied here, with the exception of the total source lifetime. Note, however, that an identical normalization of the total source lifetime is obtained when Cygnus A is excluded from the analysis (see the companion paper by Guerra, Daly, & Wan 2000). Thus, Cygnus A is not needed to normalize any of the quantities studied here.
## 3 Luminosity in Directed Kinetic Energy of the Jet
### 3.1 Theory
The jet or beam of an FRII radio source refers to the collimated outflow which carries energy from the AGN. When the jet impinges upon the external medium, a strong shock is formed and the kinetic energy of the jet is deposited in the vicinity of the shock front, which is marked by the radio hot spot. In principle, one can calculate the beam power, $`dE/dt`$, carried by the jet by studying the propagation of the hot spots. However, in practice it is difficult to use hot spot properties to estimate the beam power, since the hot spot is generally not resolved, and the position and properties of radio hot spots vary on short time scales (Laing 1989; Carilli, Perley, and Dreher 1988; Black et al. 1992.).
The model presented and discussed by (Daly 1994, 1995) bypasses these difficulties by using properties of the more stable radio lobes to estimate the luminosity in directed kinetic energy, or the beam power, $`L_j`$. This model is a natural progression following the work of Blandford & Rees (1974), Scheuer (1974), Eilek & Shore (1989), Gopal-Krishna & Wiita (1991), Rawlings & Saunders (1991), and many others. The method and results presented here are complementary to those of Eilek & Shore (1989) in the sense that Eilek & Shore (1989) considered the case of a non-supersonic expansion of the radio emitting region, while we consider the supersonic expansion of this region, and the shape of the radio emitting region considered here is different from that considered by Eilek & Shore. The results presented here are also complementary to those of Gopal-Krishna & Wiita (1991), who consider likely explanations for the decrease in FRII radio source size with redshift, and who present a consistent picture in which the increase in radio luminosity and decrease in radio source size with redshift can be understood. The work presented here differs from that of Gopal-Krishna & Wiita (1991) in that neither an efficiency nor an evolution of the ambient gas density with redshift are adopted. Instead, we use the radio properties of a source to solve directly for the beam power of the source, as well as several other parameters, including the ambient gas density. The work presented here on the beam power is similar in spirit to that of Rawlings & Saunders (1991). The key difference is that here the radio bridge is assumed to be cylindrically symmetric, and thus has a volume that depends on the length of the bridge times the cross-sectional area of the bridge, while Rawlings & Saunders (1991) seem to take a bridge volume that is proportional to the cube of the bridge length.
Observations and numerical simulations indicate that radio lobes of powerful extended radio sources propagate supersonically relative to the ambient medium (Alexander & Leahy 1987; Prestage & Peacock 1988; Cox, Gull, & Scheuer 1991; Daly 1994; WDW97a,b). Leahy (1990) shows that the properties of the radio lobe can be used to estimate the rate of energy input. In §3.4.4.3, Leahy (1990) shows that the jet power may be written $`(3P+P)V/\tau `$, since the energy density of a relativistic plasma is 3 times the pressure P; note that the time rate of change of the volume may be written: $`V/\tau =\pi a_L^2v_L`$ since the forward region of the source grows with a velocity $`v_L`$ and has cross-sectional area $`\pi a_L^2`$. His equation may be rewritten as:
$$L_j=4\pi a_L^2v_LP_L.$$
(1)
Here $`a_L`$ is the cross-sectional radius of the radio lobe, $`v_L`$ is the lobe propagation velocity, and the FRII source is assumed to be cylindrically symmetric about the radio axis. Using typical units, $`L_j`$ can be expressed as
$$L_j=3.6\times 10^{44}\text{erg s}^1\left(\frac{a_L}{\text{kpc}}\right)^2\left(\frac{v_L}{c}\right)\left(\frac{P_L}{10^{10}\text{dyne cm}^2}\right),$$
(2)
All three parameters used to estimate $`L_j`$ in eq. (2) can be estimated from radio data (see, for example, Daly 1995, or WDW97a,b). The lobe radius $`a_L`$ can be measured from the radio map (see WDW97a,b for detail). The lobe propagation velocity $`v_L`$ can be estimated using multi-frequency observations of radio lobes and bridges, by modeling the spectral aging of relativistic electrons due to synchrotron and inverse Compton losses (cf. WDW97a,b; Wan & Daly 1998a,b; Myers & Spangler 1985; Alexander & Leahy 1987; LMS89; LPR92; Carilli et al. 1991; Perley & Taylor 1991). The lobe pressure $`P_L`$ is given by
$$P_L=\left(\frac{4}{3}b^{1.5}+b^2\right)\frac{B_{min}^2}{24\pi },$$
(3)
where $`B_{min}`$ is the magnetic field strength in the radio lobe under the minimum-energy condition (cf. Miley 1980; Pacholczyk 1970), $`b`$ is the ratio of the true magnetic field strength $`B`$ to the minimum-energy magnetic field strength: $`B=bB_{min}`$, and the equation is in cgs units. WDW97b estimate $`b`$ to be about 0.25, with a source-to-source dispersion less than about $`15\%`$. This value of $`b`$ is consistent with values obtained by other authors (cf. Carilli et al. 1991; Perley & Taylor 1991; Feigelson et al. 1995), and is used throughout this study.
The radio spectral index of an FRII source is used in the spectral aging analysis to estimate the lobe propagation velocity $`v_L`$. It is shown in WDW97b that there is a clear correlation between the radio spectral indices of the sources in this sample and redshift. The spectral index $`\alpha `$ can be expressed as $`\alpha (1+z)^s`$. The best fitted value of $`s`$ is $`0.8\pm 0.2`$ with a reduced $`\chi ^2`$ of 1.7. Such a correlation could be caused by intrinsic curvature in the initial electron energy spectrum or effects of inverse Compton cooling on the hot spot spectral index (see WDW97b). Thus, the observed spectral index may not be the appropriate index to used in spectral aging analysis. This lead WDW97b to also consider correcting the observed spectral indices to zero redshift using the empirical correlation between $`\alpha `$ and $`(1+z)`$ for the sample. Results obtained with and without this correction are included here.
The beam power $`L_j`$ can be used to estimate the total time that the AGN will be active, and will produce highly collimated outflows. Following Daly (1994, 1995) and Guerra & Daly (1996, 1998), the total lifetime of an outflow, defined as $`t_{}`$, is related to the energy extraction rate $`L_j`$ by a power law:
$$t_{}=CL_j^{\beta _{}/3},$$
(4)
where $`\beta _{}`$ is a parameter that has implications for models of energy extraction from the central engine. As described in detail in a companion paper (see Guerra, Daly, & Wan 2000), eq. (4) is equivalent to assuming that the source will have an average size close to the average size of the full population located at a similar redshift, and that the average rate of growth of the source will be close to the current rate of growth of the source. Thus, we could have written $`t_{}<D>/v_L`$ where $`<D>`$ is the average size of the full population of FRIIb sources at similar redshift, and $`v_L`$ is the rate of growth of the source under consideration. Since this is precisely the manner in which the value of $`\beta _{}`$ is obtained, it is an equivalent approach. Note that the value of $`\beta _{}`$ is, fortuitously, independent of choice of cosmological parameters, and it is independent of the way the relation is normalized (see the companion paper of Guerra, Daly, & Wan 2000).
The current best estimate of $`\beta _{}`$ is $`1.75\pm 0.25`$ (see the compansion paper of Guerra, Daly, & Wan 2000). (Note that the $`\beta _{}`$ used here is not related to $`\beta _0`$ used in §4 and §5 for the King density profile.) The characteristic size of an FRII source is related to its lifetime as $`D_{}=v_Lt_{}=Cv_LL_j^{\beta _{}/3}`$, where $`C`$ is the normalization in eq. (4). This normalization factor is chosen so that at $`z0`$, the characteristic size of Cygnus A matches the observed average lobe-lobe size for sources at this redshift. That is, $`(v_L)CL_j^{\beta _{}/3}`$ for one side of Cygnus A is added to $`(v_L)CL_j^{\beta _{}/3}`$ on the other side of Cygnus A, and this is set equal to $`2(<D>_{z=0})`$ = $`2\times (68\pm 14)`$ kpc. This equation is then solved for the normalization factor $`C`$; the uncertainty on $`C`$ is indicated in Table 1. This normalization is then used in equation (4) to obtain an estimate of the total time that the AGN will be producing large-scale jets. Note that a nearly identical normalization is obtained when Cygnus A is excluded from the data analysis, as shown in a compansion paper (Guerra, Daly, & Wan 2000). Note also that the constraints on $`\beta `$ and on cosmological parameters are independent of this normalization, as described in Guerra, Daly, & Wan (2000). The value for $`t_{}`$ estimated using independent information from each side of the source should of course be equal, and are generally very close in value. The value adopted for the beam power, $`L_j`$, of Cygnus A used above is listed in Table 1, $`v_L`$ for Cygnus A is obtained from Table 1 from WDW97b, and a value of $`<D>_{z=0}`$ = $`(68\pm 14)`$ kpc is adopted from Table 3 of Guerra, Daly, & Wan (2000).
The results on $`t_{}`$ do not depend on the normalization adopted here; as shown in a companion paper, nearly identical results obtain when Cygnus A is excluded from the analysis. And, Cygnus A is not used to normalize any other quantity. Thus, none of the results presented in this paper rely upon a normalization based on Cygnus A.
### 3.2 Empirical Results
The beam power, $`L_j`$, has been estimated using eq. (2) for the lobe propagation velocity estimated using the observed radio spectral index, and the redshift-corrected radio spectral index; both are listed in Table 1. Typical values of $`L_j`$ are $`10^{45}h^2\text{ erg s}^1`$.
No strong correlation is found between the beam power $`L_j`$ and the linear size of the source, represented by the core-hot spot separation $`r`$. Figure 1 is a log-log plot of $`L_j`$ vs. $`r`$, where the observed radio spectral index $`\alpha `$ is used. The figure of $`L_j`$ vs. $`r`$ after applying the redshift-correction on $`\alpha `$ is almost identical to Figure 1, and is thus not shown here.
For radio-loud quasars, the best-fitted line in Fig. 1 has a slope of about zero. A Spearman rank correlation analysis shows that the correlation coefficient between $`L_j`$ and $`r`$ is $`0.03(7.72\%)`$ when no redshift-correction on the spectral index is applied, and $`0.01(1.78\%)`$ when the spectral index correction is applied. Here the number in parentheses is the significance of the correlation. These results suggest that the correlation between $`L_j`$ and $`r`$ for radio-loud quasars is insignificant, either with or without the $`\alpha z`$ correction. That is, $`L_j`$ is independent of $`r`$ for the radio-loud quasars.
For radio galaxies, $`L_j`$ appears to increase slightly with $`r`$ in Fig. 1, and results from the Spearman rank correlation analysis hint a marginally significant correlation between $`L_j`$ and $`r`$. The correlation coefficient is about 0.3, with a significance level of about $`90\%`$, which holds with or without the $`\alpha z`$ correction. However, note that the best-fitted line for radio galaxies in Fig. 1 has a slope only $`2\sigma `$ away from zero, and that a correlation coefficient of about 0.3 is rather weak. Thus it appears that $`L_j`$ is at most weakly dependent on $`r`$ for radio galaxies. These results for radio-loud quasars and radio galaxies are consistent with the assumption that $`L_j`$ is roughly constant over a source’s lifetime.
The relationship between $`L_j`$ and redshift and radio power ($`P_r`$) has been studied in detail in Wan & Daly (1998a). Increases of $`L_j`$ with $`z`$ and with $`P_r`$ are observed (see Figures 2 and 3). However, the radio power of the sources in the sample also increases with redshift since the sources used in this study all come from the flux-limited 3CR survey. As a result, which of the $`L_jz`$ and $`L_jP_r`$ correlations is more significant, and the role of radio power selection effects, needs to be considered carefully. Wan & Daly (1998a) used two parameter fitting and partial rank analysis to investigate this. The two parameter fit $`L_jP_{178}^{n_p}(1+z)^{n_z}`$ yields the following $`n_z`$ values: $`n_z=1.45\pm 0.32(0.60)`$ for all galaxies, $`n_z=3.83\pm 0.93(1.73)`$ for all galaxies except for Cygnus A, and $`n_z=9.56\pm 1.37(3.91)`$ for all radio-loud quasars.
The number in parenthesis is the uncertainty on $`n_z`$ times $`\sqrt{\chi _r^2}`$, which includes the effect of a reduced $`\chi ^2`$ $`(\chi _r^2)`$ that is greater than one for the fit. This notation will be used throughout this paper.
The values of $`n_z`$ are more that $`2\sigma `$ away from zero, for both radio galaxies and radio-loud quasars. This means that the $`L_jz`$ correlation is unlikely to be caused purely by radio power selection effects. That is, a real increase of $`L_j`$ with redshift exists, though the magnitude of this redshift evolution is not well determined.
The $`L_jP_r`$ correlation, on the other hand, is rather poorly constrained. Results from the one parameter fit suggest $`L_j`$ goes roughly proportional to $`P_{178}`$ (Figure 3). The two parameter fit $`L_jP_{178}^{n_p}(1+z)^{n_z}`$, which takes into account the redshift dependence of radio power, gives values of $`n_p`$ with large uncertainties: $`n_p=0.76\pm 0.11(0.20)`$ for all galaxies, $`n_p=0.25\pm 0.22(0.41)`$ for all galaxies except for Cygnus A, and $`n_p=0.54\pm 0.29(0.82)`$ for all radio-loud quasars. Thus, the exact relationship between $`L_j`$ and radio power is not clear at the present time. Note, however, large amounts of scatter are seen in the $`L_j`$ vs. $`P_{178}`$ plot (see Figure 3), which is also reflected in the large reduced $`\chi ^2`$ of the one parameter fit. For the same beam power, the radio power can vary by about a factor of ten. This suggests that radio power is not an accurate measure of the beam power, and that the apparent correlation between $`L_j`$ and $`P_r`$ may result from the fact that both increase with redshift.
The properties of FRIIb radio sources suggest that the total lifetime of a source decreases systematically with redshift. This follows because the average size of FRIIb sources decreases with redshift, while the rate of growth of the sources increases with redshift. Another way to state this is: as $`L_j`$ increases with redshift, the total lifetime of an outflow, $`t_{}`$, estimated using eq. (4), decreases with redshift. The values of the full lifetime of the source $`t_{}`$ are listed in Table 1. Typical values of $`t_{}`$ range from about $`10^7`$ to $`10^8`$ years. Figure 4 plots $`t_{}`$ as a function of redshift. It is clear that $`t_{}`$ decreases with $`z`$, which can explain the fact that the average size of powerful extended 3CR sources decreases with redshift for $`z>0.3`$ (see Guerra & Daly 1998), while the lobe propagation velocity $`v_L`$ increases with redshift. Clearly if $`v_L`$ increases with redshift, and the mean source size $`D_{}`$ decreases with redshift, then $`t_{}`$ must decrease with redshift since $`D_{}v_Lt_{}`$. The sources at high redshift produces more powerful jets (with larger $`L_j`$) for a shorter period of time (with smaller $`t_{}`$), which results in smaller average sizes than low-redshift sources.
## 4 Thermal Pressure of the Ambient Gas
### 4.1 Theory
The studies of WDW97a and b show that both the ambient gas density and temperature of an FRIIb source can be estimated from radio data. The thermal pressure of the ambient gas $`P_{th}`$ is obviously proportional to the product of the density $`n_a`$ and the temperature $`T`$. Interestingly, as shown below, although the lobe propagation velocity enters into $`n_a`$ and $`T`$, it cancels out in the product, so the thermal pressure of the ambient gas can be estimated using single frequency radio data if the radio emission from the bridge is mapped over a large enough region of the bridge. This offers a completely new method to estimate the thermal pressure of the ambient gas surrounding high-redshift radio sources using only single frequency radio data.
It can be complicated and time consuming to estimate the ambient gas density in the vicinity of classical double radio sources using X-ray observations, especially for sources at high redshift. An attractive alternative method of estimating $`n_a`$ in the vicinity of a radio source is to use the radio data. Given that the radio lobe represents a strong shock front, the lobe pressure $`P_L`$, the ambient gas density $`\rho _a`$, and the lobe propagation velocity are related: the strong shock jump conditions imply that $`P_L0.75\rho _av_L^2`$ holds, where $`P_L`$ is the non-thermal pressure inside the radio lobe, and $`\rho _a`$ is the mass density of the ambient gas. The electron number density of the ambient gas $`n_a`$ can be estimated using
$$n_a\frac{P_L}{1.4\mu m_pv_L^2},$$
(5)
where $`\mu `$ is the mean molecular weight of the gas in AMU, with a value of 0.63 when solar abundances are assumed, and $`m_p`$ is the proton mass.
The ambient gas temperature can be estimated by careful studies of the radio bridge, as discussed in detail in WDW97a. A brief description of the theory is as follows. When the non-thermal pressure of the radio bridge is much greater than the ambient gas pressure, the bridge undergoes supersonic expansion. This lateral expansion can be treated as a blast wave and the expansion velocity is determined by ram pressure confinement. This causes the width of the bridge, $`a`$, to vary with time approximately as $`t^{1/2}`$ (e.g. Begelman & Cioffi 1989; Daly 1990; WDW97a). Since the lobe propagation velocity $`v_L`$ is roughly constant during a given source’s lifetime (e.g. LPR92; WDW97a,b), the lobe front will advance a distance of $`v_Lt`$ during $`t`$ as the bridge expands. This means that at a distance $`x`$ from the hot spot, the bridge expansion time is $`t=x/v_L`$. Thus the shape of the radio bridge roughly follows a square root law, with the bridge width $`a(x)x^{1/2}`$. This behavior was in fact found by WDW97a. As the bridge expands, the pressure inside decreases until it becomes comparable to the ambient gas pressure. When this occurs, the lateral expansion velocity becomes sonic, causing a break in the functional form of $`a(x)`$. The width of the bridge starts to deviate from the square root law and becomes roughly constant as equilibrium is being reached. At the point where such a break occurs, the lateral expansion velocity is approximately sonic. Thus the sound speed of the ambient gas ($`c_s`$) can be estimated using
$$c_sv_s|_b=\frac{da}{dt}|_b=\frac{da}{dx}|_b\frac{dx}{dt}=\frac{da}{dx}|_bv_L,$$
(6)
where $`v_s|_b`$ denotes the lateral expansion velocity at the break.
The Mach number of lobe propagation is defined as $`Mv_L/c_s`$. Given the expression for $`c_s`$ in eq. (6), the Mach number of lobe propagation is:
$$M\frac{v_L}{c_s}\left(\frac{da}{dx}|_b\right)^1=\frac{2x_b}{a_b},$$
(7)
where $`a_b`$ is the width of the bridge at the break, and $`x_b`$ is the position of the break relative to the hot spot (see WDW97a). It can be seen from eq. (7) that $`M`$ depends only on the geometrical shape of the radio bridge.
The ambient gas temperature T can be expressed as:
$$T=\frac{\mu m_p}{\gamma k}c_s^2\left(\frac{da}{dx}|_b\right)^2v_L^2v_L^2M^2,$$
(8)
where $`\gamma `$ is the ratio of specific heats of the ambient gas.
Multi-frequency radio maps are required in order to use eqs. (5) and (8) to estimate the ambient gas density and temperature of an FRII source, since both $`n_a`$ and $`T`$ depend on $`v_L`$. The product $`n_aT`$ is independent of $`v_L`$, so $`v_L`$ is not needed to estimate the thermal pressure of the ambient gas. The ambient gas pressure in the vicinity of the radio lobe is simply
$$P_{th}=(n_a+n_i)kT\left(\frac{P_L}{v_L^2}\right)v_L^2M^2\frac{P_L}{M^2},$$
(9)
with $`P_L`$ estimated using eq. (3) and $`M`$ estimated using eq. (7). Here $`n_i`$ is the number densities of ions, and for a gas with solar abundances, the electron density is $`n_a=1.21n_i`$. Substituting in the numerical constants in eqs. (5) and (8), the thermal pressure is given by
$$P_{th}=0.8\frac{P_L}{M^2},$$
(10)
for a gas with a specific heat ratio $`\gamma =5/3`$.
The thermal pressure estimated using eq. (10) depends only on the lobe pressure $`P_L`$ and the geometrically determined Mach number $`M`$, both of which can be obtained from single frequency observations of the radio lobe and bridge. No spectral aging analysis is needed in order to estimate $`P_{th}`$, since the dependences of $`n_a`$ and $`T`$ on $`v_L`$ cancel. As a result, $`P_{th}`$ is also independent of whether or not the redshift-correction on the radio spectral index is applied.
At low-redshift, X-ray data can often be used to estimate $`n_a`$ and $`T`$, and hence $`P_{th}`$. However, high-resolution X-ray data are often difficult to obtain for sources at high redshift. This new method of estimating $`P_{th}`$, using single frequency radio data, offers an attractive alternative. It provides a powerful tool to study the environments of powerful classical double radio sources, especially those at high-redshift. Since it has been shown that these sources are in cluster-like gaseous environments, the sources may be used to study evolution of gas in clusters of galaxies (as discussed by Daly 1995; WDW97a,b; Daly 2000).
### 4.2 Empirical Results
The value of $`P_{th}`$, the thermal pressure of the ambient gas in the vicinity of the radio lobe, is listed in Table 1. Most sources in the sample have $`P_{th}`$ on the order of ($`10^{11}`$ to $`10^{10})h^{4/7}\text{dyne cm}^2`$, which is typical of gas in low-redshift clusters of galaxies. This is consistent with results obtained by WDW97(a,b), who find cluster-like density and temperature for the ambient gas of the FRII sources in the sample. Note that the thermal pressures obtained here do not depend on a spectral aging analysis, whereas density and temperature estimates do.
The X-cluster around Cygnus has been observed by Arnaud et. al. (1984) and Carilli, Perley, & Harris (1994). The thermal pressure of the ambient gas near its radio lobe is estimated to be about $`10^{10}\text{dyne cm}^2`$ (see Carilli, Perley, & Harris 1994). This estimate from X-ray data is consistent with the thermal pressure estimated here for Cygnus A (3C 405) using the new method just described (see Table 1).
The composite pressure profile, $`P_{th}`$ as a function of the core-hot spot separation $`r`$, is shown in Figure 5. It can be seen that $`P_{th}`$ decreases with $`r`$. A negative pressure gradient is expected for an isothermal gas distribution that follows the King density profile, which is suggested to be the ambient gas distribution for sources in our sample (WDW97a,b). For this gas distribution, the thermal pressure decreases with $`r`$ as
$$P_{th}(r)=P_{th,c}\left[1+\left(\frac{r}{r_c}\right)^2\right]^{\frac{3}{2}\beta _0},$$
(11)
where $`P_{th,c}`$ is the thermal pressure at the center of the cluster, and $`r_c`$ is cluster core radius.
The pressure gradient seen Figure 5 appears to be consistent with that expected for an cluster gas distribution that is isothermal and follows a King density profile, with a cluster core radius $`r_c`$ of $`(50\text{ to }150)h^1`$ kpc. The sources in our sample have values of $`r`$ ranging from (25 to 250) $`h^1`$ kpc. Within this radius range, the pressure profile given by eq. (11) has an average slope of ($`1.4\text{ to }0.7`$), for $`r_c`$ values of (50 to $`150h^1`$ kpc), assuming $`\beta _0=2/3`$. These expected slopes are consistent with the best-fit slopes of $`1.0\pm 0.3`$ in Figure 5.
There is some hint of a redshift evolution of the cluster core radius from the data. Figure 6 plots $`P_{th}`$ vs. $`r`$ in two redshift bins, $`z<0.9`$ and $`z>0.9`$. The division at $`z=0.9`$ is chosen so that the two redshift bins cover about the same range in redshift and contain about the same number of data points. It appears that $`P_{th}`$ decreases less rapidly with $`r`$ at high redshift than at low redshift, suggesting a larger $`r_c`$ at high redshift. The best-fit slope of $`1.4`$ in the low-redshift bin is consistent with $`r_c50h^1`$ kpc for the pressure profile given by eq. (11), whereas the best-fit slope of about $`0.4`$ in the high-redshift bin is consistent with $`r_c250h^1`$ kpc. This result is still preliminary since only sources with detections of $`P_{th}`$ are included in the fits. Better estimates of $`P_{th}`$ for sources currently with only upper bounds on $`P_{th}`$ will help to better determine whether $`r_c`$ is evolving with redshift.
The results obtained above are consistent with results obtained by WDW97b. They study the ambient gas density profile and find that $`r_c50h^1`$ kpc when high- and low-redshift source are considered together. They also find that that data are consistent with a constant core gas mass model where $`r_c`$ increases with redshift roughly as $`r_c(1+z)^{1.6}`$. Note that results from the study of $`P_{th}`$ do not depend on spectral aging analysis, whereas those from the study of $`n_a`$ do. It is thus encouraging that consistent results are obtained from the two studies.
For the following analysis, we use both a non-evolving core radius of $`50h^1`$ kpc, and an evolving core radius of $`50(1+z)^{1.6}h^1`$ kpc, with a focus on the latter.
The thermal pressure at the center of the cluster ($`P_{th,c}`$) can be estimated by scaling the pressure in the vicinity of the radio lobe ($`P_{th}`$) to the cluster center. This follows because the studies of Daly (1995) and WDW97a,b indicate that these radio sources are located at the centers of clusters of galaxies, so the core-hot spot separation can be used as an estimate of distance from the cluster center. The current data on $`T`$, $`n_a`$, and $`P_{th}`$ are all consistent with the gas distribution being isothermal and following a King profile. Thus we use the pressure profile for such a gas distribution, as given by eq. (11), to estimate the central thermal pressure. The central pressure is simply
$$P_{th,c}=P_{th}(r)\left[1+\left(\frac{r}{r_c}\right)^2\right]^{\frac{3}{2}\beta _0},$$
(12)
The values of $`P_{th,c}`$ for the sources in our sample are listed in Table 1, where a value of $`\beta _0=2/3`$ has been assumed. Most sources have values of $`P_{th,c}`$ around $`10^{10}h^{4/7}\text{dyne cm}^2`$, which is rather typical of gas pressure in the core regions of low-redshift galaxies clusters.
Results on the redshift evolution of the central gas pressure $`P_{th,c}`$ are somewhat uncertain because $`P_{th,c}`$ seems to correlate with both redshift and radio power when either correlation is considered separately (see Figures 7 and 8), and it is not clear at present which correlation is more significant. For the non-evolving core radius model, a two-parameter fit of $`P_{th,c}(1+z)^{n_z}P_{178}^{n_p}`$ yields $`n_z=0.29\pm 0.49(0.67)`$ and $`n_p=0.66\pm 0.19(0.26)`$ when all sources are included, and $`n_z=1.51\pm 2.11(2.99)`$ and $`n_p=0.34\pm 0.58(0.82)`$ when Cygnus A is not included. For the evolving core radius model, the two-parameter fit gives $`n_z=1.22\pm 0.49(0.65)`$ and $`n_p=0.88\pm 0.19(0.25)`$ when all sources are included, and $`n_z=1.56\pm 2.11(2.82)`$ and $`n_p=0.15\pm 0.58(0.78)`$ when Cygnus A is not included. Note that Cygnus A appears to lie in a cooling-flow region while most of the other sources are not in cooling-flow regions (see WDW97a; §5). Thus fits with and without Cygnus A are both performed. In any case, the large uncertainties on $`n_z`$ and $`n_p`$ make it hard to determine the magnitude of the redshift evolution of $`P_{th,c}`$.
The thermal pressure can be used to predict the amount of cosmic microwave background (CMB) diminution, also known as the Sunyaev-Zel’dovich (S-Z) effect, that is expected from the cluster. The reduction in the cosmic radiation ($`\mathrm{\Delta }I_\nu `$) in the direction of a cluster is given by (e.g. Sunyaev & Zel’dovich 1980; Sarazin 1988; Rephaeli 1995)
$$\mathrm{\Delta }I_\nu /I_\nu =G(\omega )y.$$
(13)
Here $`I_\nu `$ is the specific intensity of the CMB at the observing frequency $`\nu `$, $`y`$ is the Comptonization parameter, and the function $`G(\omega )`$ is
$$G(\omega )=\frac{\omega e^\omega }{e^\omega 1}\left[\omega \left(\frac{e^\omega +1}{e^\omega 1}\right)4\right],$$
(14)
where $`\omega h\nu /kT_r`$, and $`T_r=2.73\text{K}`$ is the CMB temperature (Mather et al. 1990).
The Compton $`y`$ parameter is given by
$$y=\frac{kT}{m_ec^2}\sigma _Tn_e𝑑l=\frac{\sigma _T}{m_ec^2}\left(\frac{n_e}{n_e+n_i}\right)P_{th}𝑑l.$$
(15)
where $`\sigma _T=(8\pi /3)[e^2/(m_ec^2)]^2`$ is the Thompson electron scattering cross section, $`T`$ is the cluster gas temperature, $`n_e`$ is the electron density, and $`n_i`$ is the ion density. Using the thermal pressure in a cluster with an isothermal-King gas distribution (eq. ), this gives
$$y6.3\times 10^5\left(\frac{P_{th,c}\text{ }}{10^{10}\text{ dyne cm}\text{-2}\text{ }}\right)\left(\frac{r_c\text{ }}{\text{0.25\hspace{0.17em}Mpc}}\right)\frac{\mathrm{\Gamma }\left(3\beta _0/21/2\right)\text{ }}{\mathrm{\Gamma }(3\beta _0/2)\text{ }}(1+x^2)^{3\beta _0/2+1/2},$$
(16a)
where $`xr/r_c`$, and the gas is taken to have solar abundance so that $`n_e=1.21n_i`$. For a typical value of $`\beta _0=2/3`$,
$$y1.12\times 10^4\left(\frac{P_{th,c}}{10^{10}\text{dyne cm}^2}\right)\left(\frac{r_c}{0.25\text{Mpc}}\right)(1+x^2)^{1/2}.$$
(16b)
At low frequency, clusters with strong radio sources are generally avoided for measurements of the Sunyaev-Zel’dovich effect because emission from the radio source masks the microwave diminution. Such contaminations from radio sources are reduced at high frequency since emission from steep-spectrum radio sources, such as the FRII sources studied here, decreases rapidly with increasing frequency. The Sunyaev-Zel’dovich Infrared Experiment (SuZIE) can measure the S-Z effect around 140 GHz, where the amount of CMB intensity diminution is near its peak. At this frequency, a cluster with an isothermal-King density profile will cause a CMB intensity diminution of about
$$\begin{array}{cc}\mathrm{\Delta }I_\nu 7.0\times 10^{19}\text{erg s}\text{-1}\text{Hz}\text{-1}\text{cm}\text{-2}\text{sr}\text{-1}\text{ }\left(\frac{P_{th,c}\text{ }}{10^{10}\text{ dyne cm}\text{-2}\text{ }}\right)\left(\frac{r_c\text{ }}{\text{0.25\hspace{0.17em}Mpc}}\right)\hfill & \\ \times \frac{\mathrm{\Gamma }(3\beta _0/21/2)\text{ }}{\mathrm{\Gamma }(3\beta _0/2)\text{ }}(1+x^2)^{3\beta _0/2+1/2},\hfill & \end{array}$$
(17)
and for a typical value of $`\beta _0=2/3`$,
$$\mathrm{\Delta }I_\nu 1.24\times 10^{18}\text{erg s}\text{-1}\text{Hz}\text{-1}\text{cm}\text{-2}\text{sr}\text{-1}\text{ }\left(\frac{P_{th,c}}{10^{10}\text{dyne cm}^2}\right)\left(\frac{r_c}{0.25\text{Mpc}}\right)(1+x^2)^{1/2}.$$
(18)
For a detector with a FWHM beam size $`\theta _b`$, the total amount of CMB diminution within the beam, defined as $`\mathrm{\Delta }F_\nu `$, is roughly
$$\mathrm{\Delta }F_\nu f24\text{mJy}\left(\frac{\theta _b}{\text{1.7 arcmin}}\right)^2\left(\frac{P_{th,c}}{10^{10}\text{dyne cm}^2}\right)\left(\frac{r_c}{0.25\text{Mpc}}\right),$$
(19)
where $`f`$ is the beam dilution factor, defined as the ratio of the average $`\mathrm{\Delta }I_\nu `$ within the beam to the peak value at the cluster center; and the SuZIE FWHM beam size at 140 GHz of 1.7 arcmin (e.g. Holzapfel et al. 1997) is used to calculate the numerical value. For a Gaussian beam, when the HWHM beam size corresponds to 1$`r_c`$, $`f0.7`$, and when the HWHM beam size corresponds to 2$`r_c`$, $`f0.5`$ (see Rephaeli 1987). At $`z<2`$, the SuZIE beam radius is less than $`320h^1\text{kpc}`$ for $`q_0=0`$ with no cosmological constant. Thus a cluster with $`P_{th,c}10^{10}\text{dyne cm}^2`$ and $`r_c\mathrm{\hspace{0.17em}0.25}\text{Mpc}`$ will have a $`\mathrm{\Delta }F_\nu `$ of about (15 to 20) mJy within the SuZIE beam. The clusters surrounding the FRII sources in our sample have $`P_{th,c}`$ on the order of $`10^{10}h^{4/7}\text{dyne cm}^2`$ and the data also suggest that $`r_c`$ increases with redshift, reaching about $`300h^1\text{kpc}`$ at $`z2`$. The expected CMB diminution within the SuZIE beam at 140 GHz for these clusters ranges from several to tens of mJy. These clusters make good candidates for SuZIE observations if the fluxes from the radio sources at 140 GHz or their uncertainties are small compared with the expected S-Z effect signals.
We are currently in the process of searching for high frequency data on the radio sources in our sample in oder to identify possible SuZIE observation candidates. One likely candidate that comes from a preliminary search in the published literature is 3C239. The expected CMB diminution for the cluster surrounding it is about $`30h^{3/7}`$ mJy within the SuZIE beam at 140 GHz. In contrast, extrapolation of the observed spectrum of 3C239 as given by observations at 178 MHz, 10.7 GHz, and 14.9 GHz (Kellermann & Pauliny-Toth 1973; Genzel et al. 1976; Laing, Riley, and Longair 1983), gives a 140 GHz flux of only 6 mJy. This is likely to be an overestimate since synchrotron aging and inverse Compton cooling can both cause the high frequency spectral index to become steeper than that at low frequency. The effects of inverse Compton cooling can be especially important since this source is at high redshift ($`z=1.79`$), where the energy density of the microwave background is much higher than at low redshift. Thus it appears that emission from this radio source is weak compared with the expected CMB diminution. Observations of the radio source at more frequencies above 14.9 GHz can help to better constrain its 140 GHz flux.
## 5 Gravitational Mass of the Host Cluster
### 5.1 Theory
The powerful classical double radio sources studied here are in cluster-like gaseous environments (see WDW97a, b). The studies of these sources provide information on density and temperature of the ambient gas. Thus it is possible to estimate one of the most important parameters of the cluster, the total gravitational mass, including dark matter.
The total mass can be estimated using the density and temperature profile of the intracluster medium (ICM) if the gas is in hydrostatic equilibrium. The sound crossing time in the ICM is usually short compared to the age of a high-redshift cluster, and the morphology of the X-ray emission from many low-redshift clusters is often smooth. Thus it is generally believed that hydrostatic equilibrium is a good approximation of the state of the ICM for many clusters that are not cooling flow clusters. Assuming spherical symmetry, hydrostatic equilibrium requires
$$\frac{1}{\rho _g}\frac{dP}{dr}=\frac{GM_t(r)}{r^2},$$
(20)
where $`\rho _g`$ is the gas density and $`M_t(r)`$ is the total gravitating mass within $`r`$. This means that
$$M_t(r)=\frac{kT(r)}{G\mu m_p}r\left[\frac{d\mathrm{log}\rho _g(r)}{d\mathrm{log}r}+\frac{d\mathrm{log}T(r)}{d\mathrm{log}r}\right],$$
(21)
where $`k`$ is the Boltzman constant, $`\mu `$ is the mean molecular weight of the gas in amu, $`m_p`$ is the proton mass, and $`T(r)`$ and $`\rho _g(r)`$ are the temperature and density profiles of the cluster gas.
The most commonly used hydrostatic model of the ICM is the isothermal $`\beta `$-model (Cavaliere & Fusco-Femiano, 1976,1978; Sarazin & Bahcall 1977), where the clusters gas is isothermal and the density profile follows a modified King model, i.e., $`n_a=n_c[1+(r/r_c)^2]^{3/2\beta _0}`$. Here $`n_c`$ is the core density and $`r_c`$ is core radius of the gas distribution. Using this model for the ICM, eq. (21) becomes
$$M_t(r)=\frac{3\beta _0k}{G\mu m_p}Tr\frac{(r/r_c)^2}{1+(r/r_c)^2}.$$
(22)
Results from numerical simulations of cluster formation suggest that departures of the ICM from hydrostatic equilibrium are usually small, and the mass estimated using the standard $`\beta `$-model is rather accurate (e.g. Navarro, Frenk & White 1995; Schindler 1996; Evrard, Metzler, & Navarro 1996). Note however, significant temperature decline is observed to occur at outer regions of clusters (e.g. Markevitch et al. 1998). Thus, we only use eq. (22) to estimate the cluster mass within $`r`$ rather than extrapolating to large radii.
The mass estimated using eq. (22) is not accurate if the cluster is cooling at the point where the cluster temperature is measured. Inside a cooling flow region, the temperature measured does not indicate the gravitational potential, since hydrostatic equilibrium conditions do not apply. Considering cooling only by thermal bremsstrahlung, which dominates other mechanisms at typical cluster temperatures, the cooling time of the cluster gas can be estimated using
$$t_{cool}2.1\times 10^2\left(\frac{T}{10^7\text{K}}\right)^{1/2}\left(\frac{n_a}{\text{cm}^3}\right)^1\text{Gyr},$$
(23)
where $`T`$ and $`n_a`$ are the ambient gas temperature and electron density, respectively (see WDW97a). Note this equation does not depend on the Hubble’s constant, provided that the estimates of $`n_a`$ and $`T`$ do not depend on Hubble’s constant.
Given the ambient gas temperature and density estimates, the cooling time for the sources in our sample can be estimated (see WDW97a). A cooling time less than the age of the universe at that redshift means that the FRII source is in a cooling flow region. To estimate the age of the universe at a given redshift, an open empty universe is assumed, with a current age of 14 Gyr, which corresponds to $`h0.7`$ for $`q_0=0`$ with zero cosmological constant.
Most sources in our sample do not appear to be cooling at the position where the temperature is measured. Thus their mass estimates are likely to be reliable. A few sources, including the low-redshift source Cygnus A, appear to lie in cooling flow regions.
Figure 9 shows the total cluster mass, including dark matter, within radius $`r`$, $`M_t(r)`$, as a function of the core-hot spot separation $`r`$. Since the sources appear to lie at the centers of clusters of galaxies (Daly 1995; WDW97a,b), the core-hot spot separation is used to estimate the distance from the cluster center. The radio spectral index is redshift-corrected, and the cluster core radius is taken to be $`r_c=50(1+z)^{1.6}h^1`$ kpc (see §4.2). Clusters that are cooling at the position where the temperature is measured are marked with pentagons. The plots of $`M_t(r)`$ vs. $`r`$ for other models, either with or without an $`\alpha z`$ correction and/or different core radius evolution, are presented in Figures 10 through 12.
The cluster mass increases with $`r`$ for all models. The increase is more rapid for models with an evolving core radius than models with a fixed core radius. This merely reflects the fact that with an increasing $`r_c`$, more sources lie within the core region, where the increase of mass with radius is rapid, and varies roughly as $`Mr^3`$.
To determine the redshift evolution of the cluster mass, a three parameter fit of $`M_t(r)r^{n_1}(1+z)^{n_2}P_{178}^{n_3}`$ is performed since the cluster mass within $`r`$, $`M_t(r)`$, is a function of $`r`$, and may also be affected by radio power selection effects. Results are listed in Table 2 for all models considered. A two parameter fit of $`M_t(r)r^{n_1}(1+z)^{n_2}`$ is also performed, and results are listed in Table 3. For completeness, fits including all sources, all sources except Cygnus A, and all sources that are not in cooling flow regions, are all listed in the table. Note that Cygnus A appears to be in a cooling flow region for all the models considered. Results obtained excluding cooling flow clusters are probably the correct ones to consider.
The sample of clusters with mass estimates is rather small, especially when only non-cooling flow clusters are considered. Thus it is not surprising to see that the best-fit values of $`n_2`$ have rather large uncertainties, which makes it hard to draw definitive conclusions about the redshift evolution of the cluster mass. Note though, the current data do not indicate any negative evolution of the cluster mass out to a redshift of about two. This is consistent with results obtained by this group (Daly 1994; Guerra & Daly 1996, 1998; Guerra, Daly, & Wan 2000; and Daly, Guerra, & Wan 1998), who study the characteristic size of powerful classical double radio sources and find the data suggest a low value of $`\mathrm{\Omega }_m`$; a universe with $`\mathrm{\Omega }_m=1`$ is ruled out at 99 % confidence (see Guerra, Daly, & Wan 2000).
There are some indications from the data that the redshift evolution of the cluster core radius, and the redshift-correction on the radio spectral index are favored. The gas mass within $`r`$, defined as $`M_g(r)`$, can be estimated using
$$M_g(r)=4\pi \rho _cr_c^3[x\text{tan}^1(x)],$$
(24)
where $`\rho _c`$ is the central gas density, $`r_c`$ is the core radius, $`xr/r_c`$, and a typical value of $`\beta _0=2/3`$ is used for the King density profile. Knowing the gas mass and the total mass within $`r`$, the gas mass fraction within $`r`$ can be estimated. The gas mass fraction within radius $`r`$ as a function of redshift is shown in Figure 13, where the radio spectral index is redshift-corrected and the core radius is taken to be $`r_c=50(1+z)^{1.6}h^1`$ kpc.
The values of gas fraction shown in the figure are consistent with observed values for inner regions of many clusters (e.g. Donahue 1996), whereas those for models without a redshift-correction on the radio spectral index, and without a core radius evolution seem to be low compared with observed values.
Several factors may contribute to the slow decrease of gas fraction with redshift that is seen in Figure 13. First, the increase of cluster core radius with redshift means that the gas fraction estimated for the high-redshift clusters is over a larger fraction of the cluster core than for the low-redshift clusters. The gas fraction estimated at high redshift mainly represents the gas fraction inside the cluster core, and it is known that gas fraction increases with increasing radius in clusters (e.g. David, Jones, & Forman 1995). Second, the mass contribution from galaxies is higher at the cluster center than at large radii (e.g. Loewenstein & Mushotzky 1996). Thus by sampling a larger fraction of the cluster core at higher redshift, a larger fraction of the total baryon mass is not taken into account at higher redshift. Further, the decrease in gas fraction with redshift could be due to the stripping of gas from galaxies in the cluster core over time. Finally, the cluster gas becomes more concentrated toward the cluster center as it cools, which also causes the gas fraction in the cluster center to increase with time.
These results on cluster mass and gas fraction obtained above are still preliminary due to the small size of the sample. More sources with estimates on the ambient gas temperature, and hence cluster mass, will help to test these results.
## 6 Summary and Discussion
Several key parameters of an FRII source and its gaseous environment are studied in this paper.
Direct estimates of the beam power, which measures the energy extraction rate from the AGN by the jet, are obtained. Typical beam powers of about $`10^{45}h^2\text{erg s}^1`$ are found for the sources in the sample. No strong correlation is seen between the beam power and the linear size of the source, which is consistent with the beam power being roughly constant throughout the lifetime of a source.
There is a trend for the beam power to increase with redshift, which is significant even after excluding radio power selection effects. The magnitude of this redshift evolution, however, is not well determined by the current data. It is well known that the quasar luminosity function undergoes strong evolution between $`z0`$ and $`z2`$, with the high-redshift quasars being more luminous than their low-redshift counterparts (e.g. Schmidt & Green 1983; Yee & Ellingson 1993; La Franca & Cristiani 1997). This suggests that high-redshift AGNs are more powerful than low redshift ones. Thus it is perhaps not surprising to find that the beam power, or energy per unit time channeled into the jet by the AGN, is also higher at high redshift.
The relationship between the beam power and the radio power is not well constrained after their correlations with redshift are taken into account. The two parameter fit of $`L_j(1+z)^{n_z}P_{178}^{n_p}`$ yields values of $`n_p`$ with large uncertainties. Thus it is not clear at present how the beam power and the radio power are related. However, the large amount of scatter seen in the beam power-radio power relation suggests that radio power is not an accurate measure of beam power.
The beam power $`L_j`$ can be used to estimate the total lifetime of an outflow produced by an AGN. Following Daly (1994) and Guerra & Daly (1996, 1998), the total time for which the outflow will occur, $`t_{}`$, is related to $`L_j`$: $`t_{}L_j^{\beta _{}/3}`$, where $`\beta _{}`$ is estimated to be about $`1.75\pm 0.25`$ (Guerra, Daly, & Wan 2000). Typical lifetimes of about $`10^7`$ to $`10^8`$ years are obtained for the sources studied here. This would is almost precisely the same lifetime as would be obtained by dividing the average size of all FRIIb sources at similar redshift by the rate of growth of the source under consideration. The lifetime of the outflow decreases with redshift, which explains the decrease of the average size of powerful extended 3CR sources with redshift. The sources at high redshift produces more powerful jets for a shorter period of time, which results in smaller average sizes than low-redshift sources.
A new method of estimating the thermal pressure of the ambient gas in the vicinity of powerful classical double radio source using only single frequency radio data is presented. The pressure is given by the product of the ambient gas density, estimated using ram pressure confinement of the radio lobe, and the ambient gas temperature, estimated using the Mach number for the source and the lobe advance speed. It turns out that the lobe propagation velocity cancels out of the product $`n_aT`$, and the ambient gas pressure can be estimated by studying the shape of the radio bridge, and the non-thermal pressure of the radio lobe (see §4). Thus, the thermal pressure, the product of the ambient gas density and temperature, depends only on the non-thermal pressure in the radio lobe and the geometrically determined Mach number of lobe advance, both of which can be estimated using single frequency radio data. This new method to estimate the thermal pressure does not require a spectral aging analysis, and provides a powerful tool to probe the environments of FRII sources. The thermal pressure estimated for Cygnus A using this new method agrees with that obtained using X-ray data (see §4.2.).
Thermal pressures on the order of $`10^{10}h^{4/7}\text{dyne cm}^2`$, typical of gas in low-redshift clusters of galaxies, are found for the gaseous environments of the FRII sources studied here. There are hints from the current data that the gradient of the composite pressure profile is less steep at high redshift than at low redshift, which can be explained by an increase of the cluster core radius with redshift. The current data are consistent with a core radius evolution from $`50h^1\text{kpc}`$ at $`z0`$ to $`250h^1\text{kpc}`$ at $`z2`$, which agrees with results obtained from studies of the ambient gas density (WDW97b). WDW97b find that the data can be described by a model where the core gas density decreases and the core radius increases so that the core gas mass remains roughly constant.
The thermal pressures obtained here can be used to estimate the amount of CMB diminution expected from the clusters surrounding the FRII sources in the sample. Contaminations from the radio sources can be reduced by observing at high frequency, such as 140 GHz, since emission from the radio sources decreases rapidly with increasing frequency. The cluster surrounding a source in our sample can be detected by SuZIE observations at 140 GHz if the flux from the radio source at this frequency or its uncertainty is small compared with the expected S-Z effect signal. A search for high-frequency data of the sources in our sample is ongoing in oder to identify possible SuZIE observation candidates. Preliminary results suggest that 3C239 is a good candidate.
The gravitational or total mass of the surrounding cluster is estimated for the sources in the sample, assuming hydrostatic equilibrium conditions for the gas. Masses of up to $`10^{14}M_{}`$ are found for the central regions ($`r250h^1\text{kpc}`$) of the clusters, consistent with typical values for low-redshift clusters. The redshift evolution of the cluster mass is not well determined. Current data do not indicate negative evolution of the cluster mass. This is consistent with results obtained by members of this group (Guerra & Daly 1996, 1998; Guerra, Daly, & Wan 2000; Daly, Guerra, & Wan 1998), who study the characteristic size of FRII sources and find the data strongly favor a low value for the density parameter $`\mathrm{\Omega }_m`$; a universe with $`\mathrm{\Omega }_m=1`$ is ruled out at 99 % confidence. Note, however, that with the study presented here, we only study the central regions of clusters.
The values of gas mass fraction obtained for the clusters surrounding the FRII sources studied here are consistent with observed values for central regions of clusters, after the correlation between the radio spectral index and redshift is taken into account, and a redshift evolution of the cluster core radius is considered. This suggests that the cluster core radius evolution and the effects of the radio spectral index-redshift correlation are important.
The gas mass fraction seems to decrease slowly with redshift. The increase of cluster core radius with redshift can be one cause. Another possible explanation is that a large fraction of the gas is still in the galaxies at high redshift, and is later stripped from the galaxies into the cluster at low redshift. Cooling of the cluster gas also tends to cause the gas mass fraction in the cluster center to increase with time.
The results on mass and gas fraction should be taken as preliminary because of the small size of the sample.
It is a pleasure to thank Greg Wellman for important discussions. We would like to thank Paddy Leahy for numerous helpful discussions. The referee deserves special thanks for very helpful comments and suggestions, which have significantly improved the paper. This work was supported in part by the US National Science Foundation and the College of Liberal Arts and Sciences at Rowan University.
|
warning/0006/hep-th0006095.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
The results that this paper presents are about gravitational couplings for generalized orientifold planes (GOp-planes) and y-deformed orientifold planes (yOp-planes). The usual orientifold planes do not have gauge fields on their worldvolumes and . The generalized orientifold planes that this paper consider have SO(2k) Yang-Mills gauge fields-bundles over their corresponding worldvolumes. The aim of the present paper is to display the Wess-Zumino part of the effective action for such generalized orientifold planes.
The usual orientifold planes do not have any kind of topological y-deformation.This paper presents the Wess-Zumino action for y-deformed orientifold planes.
For the usual orientifold planes the Wess-Zumino action has the following form,which can be derived both from anomaly cancellation arguments and from direct computation on string scattering amplitudes:
$`𝑺_{𝑾𝒁\mathbf{(}𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
Where the Mukai vector of RR charges for the usual orientifold p-plane is given by:
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}\mathbf{=}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
In this formula C is the vector of the RR potential forms. L is the Hirzebruch genus that generates the Hirzebruch polynomials which are given in terms of Pontryaguin classes for real bundles. The Pontryaguin classes are given in terms of the 2-form curvature of the corresponding real bundle. The formula for Q involves two real bundles over the worldvolume of the usual orientifold plane. These two bundles are the tangent bundle for the worldvolume and the normal bundle by respect to space-time for such worldvolume. Q is given then in terms of the curvatures for the tangent and normal bundles and does not have contributions from the others real bundles such as SO(2k) Yang-Mills gauge bundles and does not have any kind of topological deformation.
In this paper is presented the Mukay vector of RR charges for a generalized orientifold planes which have two SO(2k) Yang-Mills gauge bundles on their worldvolumes. Such vector of RR charges is given by the following formula:
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}`$
For the generalized orientifold planes the Wess-Zumino action has the following form:
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}`$
The formula for the vector of RR charges corresponding to a generalized orientifold plane involves now four real bundles over the worldvolume: the tangent bundle, the normal bundle and two new SO(2k) YM gauge bundles. When one of these new SO(2k) bundles is the tangent bundle and the other is the normal bundle, one obtain the usual formula for Q corresponding to the usual orientifold planes using the following identity:
$`𝑨\mathbf{(}\frac{𝐑}{\mathrm{𝟐}}\mathbf{)}𝑴𝒂𝒚𝒆𝒓\mathbf{(}\frac{𝐑}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}𝑳\mathbf{(}\frac{𝐑}{\mathrm{𝟒}}\mathbf{)}`$
Then, one has:
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}}}`$
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{=}𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}`$
In these formulas, A denotes the roof-Dirac genus and Mayer denotes the Mayer class for one SO(2k) YM gauge bundle.
In this paper,also, is presented the Mukay vector of RR charges for the y-deformed orientifold planes which have topological y-deformations on their worldvolumes. Such vector of RR charges is given by the following formula:
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
For the y-deformed orientifold planes the Wess-Zumino action has the following form:
$`𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
The formula for the vector of RR charges corresponding to a y-deformed orientifold plane involves now two real bundles over the worldvolume: the tangent bundle and the normal bundle, but in this case one has a topological y-deformation over the worldvolume. When the parameter y of the topological y-deformation is 1, then one obtain the usual formula for Q corresponding to the usual orientifold planes using the following identity:
$`𝑪𝑯𝑰_\mathrm{𝟏}\mathbf{(}𝑹\mathbf{)}\mathbf{=}𝑳\mathbf{(}𝑹\mathbf{)}`$
Then, one has:
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
$`𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{,}\mathrm{𝟏}\mathbf{,}\mathbf{)}\mathbf{=}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{=}𝑸\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{,}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}`$
In these formulas, CHI sub y denotes the chi-y- genus which when y=1 is the Hirzebruch-genus and when y=0, is the Todd genus.
In the following section the Mukay vector of RR charges for a such generalized orientifold p-plane (GOp-plane) and such y-deformed orientifold p-plane (yOp-plane), will be given in terms of the powers of the curvatures for the four real bunldes involved over the worldvolume for the GOp-planes and in terms of the powers of the curvatures for the two y-deformed bundles in the case of yOp-planes.
In the third section are presented the elementary processes corresponding to the power expansion for the two Q’s. In the final four section some conclutions are presented about other GOp-planes and yOp-planes and non-BPS GOp-planes and yOp-planes.
## 2 The Power Expantion for Q’s
Let E be a SO(2k)-bundle over the worldvolume of a generalized orientifold plane and consider a formal factorisation for the total Pontryaguin classs of the real bundle E, which has the following form:
$`𝒑\mathbf{(}𝑬\mathbf{)}\mathbf{=}\mathbf{}_{𝒊\mathbf{=}\mathrm{𝟏}}^𝒌\mathbf{(}\mathrm{𝟏}\mathbf{+}𝒚_𝒊^\mathrm{𝟐}\mathbf{)}`$
The total Pontryaguin classs of the real bundle E,has the following formal sumarisation in terms of the corresponding Pontryaguin classes:
$`𝒑\mathbf{(}𝑬\mathbf{)}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟎}}^{\mathbf{}}𝒑_𝒋\mathbf{(}𝑬\mathbf{)}`$
The total Mayer class for the real bundle E has the following formal factorisation:
$`𝑴𝒂𝒚𝒆𝒓\mathbf{(}𝑬\mathbf{)}\mathbf{=}\mathbf{}_{𝒊\mathbf{=}\mathrm{𝟏}}^𝒌𝒄𝒐𝒔𝒉\mathbf{(}\frac{𝐲_𝐢}{\mathrm{𝟐}}\mathbf{)}`$
The total Mayer class for the real bundle E has the following formal sumarisation in terms of the Mayer polynomials which are formed from the corresponding Pontryaguin classes :
$`𝑴𝒂𝒚𝒆𝒓\mathbf{(}𝑬\mathbf{)}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟎}}^{\mathbf{}}𝑴𝒂𝒚𝒆𝒓_𝒋\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑬\mathbf{)}\mathbf{,}\mathbf{}\mathbf{,}𝒑_𝒋\mathbf{(}𝑬\mathbf{)}\mathbf{)}`$
The Mayer polynomials are given by:
$`𝑴𝒂𝒚𝒆𝒓_\mathrm{𝟎}\mathbf{(}𝒑_\mathrm{𝟎}\mathbf{(}𝑬\mathbf{)}\mathbf{)}\mathbf{=}𝑴𝒂𝒚𝒆𝒓_\mathrm{𝟎}\mathbf{(}\mathrm{𝟏}\mathbf{)}\mathbf{=}\mathrm{𝟏}`$
$`𝑴𝒂𝒚𝒆𝒓_\mathrm{𝟏}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑬\mathbf{)}\mathbf{)}\mathbf{=}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐄\mathbf{)}}{\mathrm{𝟖}}`$
$`𝑴𝒂𝒚𝒆𝒓_\mathrm{𝟐}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑬\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟐}\mathbf{(}𝑬\mathbf{)}\mathbf{)}\mathbf{=}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐄\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟒}𝐩_\mathrm{𝟐}\mathbf{(}𝐄\mathbf{)}}{\mathrm{𝟑𝟖𝟒}}`$
$`𝑴𝒂𝒚𝒆𝒓_\mathrm{𝟑}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑬\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟐}\mathbf{(}𝑬\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟑}\mathbf{(}𝑬\mathbf{)}\mathbf{)}\mathbf{=}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐄\mathbf{)}^\mathrm{𝟑}\mathbf{+}\mathrm{𝟏𝟐}𝐩_\mathrm{𝟏}\mathbf{(}𝐄\mathbf{)}𝐩_\mathrm{𝟐}\mathbf{(}𝐄\mathbf{)}\mathbf{+}\mathrm{𝟒𝟖}𝐩_\mathrm{𝟑}\mathbf{(}𝐄\mathbf{)}}{\mathrm{𝟒𝟔𝟎𝟖𝟎}}`$
The pontryaguin classes of the real bundle E have the following realizations in terms of the powers of the 2-form curvature for such bundle. For this curvature the y’s are the eigenvalues:
$`𝒑_\mathrm{𝟏}\mathbf{(}𝑬\mathbf{)}\mathbf{=}𝒑_\mathrm{𝟏}\mathbf{(}𝑹_𝑬\mathbf{)}\mathbf{=}\mathbf{}\frac{\mathrm{𝟏}}{\mathrm{𝟖}𝐩𝐢^\mathrm{𝟐}}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}`$
$`𝒑_\mathrm{𝟐}\mathbf{(}𝑬\mathbf{)}\mathbf{=}𝒑_\mathrm{𝟐}\mathbf{(}𝑹_𝑬\mathbf{)}\mathbf{=}\frac{\mathrm{𝟏}}{\mathrm{𝟏𝟔}𝐩𝐢^\mathrm{𝟒}}\mathbf{[}\frac{\mathrm{𝟏}}{\mathrm{𝟖}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathrm{𝟏}}{\mathrm{𝟒}}𝒕𝒓𝑹_𝑬^\mathrm{𝟒}\mathbf{]}`$
$`𝒑_\mathrm{𝟑}\mathbf{(}𝑬\mathbf{)}\mathbf{=}𝒑_\mathrm{𝟑}\mathbf{(}𝑹_𝑬\mathbf{)}\mathbf{=}\frac{\mathrm{𝟏}}{\mathrm{𝟔𝟒}𝐩𝐢^\mathrm{𝟔}}\mathbf{[}\mathbf{}\frac{\mathrm{𝟏}}{\mathrm{𝟒𝟖}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟑}\mathbf{}\frac{\mathrm{𝟏}}{\mathrm{𝟔}}𝒕𝒓𝑹_𝑬^\mathrm{𝟔}\mathbf{+}\frac{\mathrm{𝟏}}{\mathrm{𝟖}}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}𝒕𝒓𝑹_𝑬^\mathrm{𝟒}\mathbf{]}`$
Using all these expretions one can to obtain the following expantion:
$`𝑴𝒂𝒚𝒆𝒓\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐄\mathbf{)}}{\mathrm{𝟑𝟐}}\mathbf{+}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐄\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟒}𝐩_\mathrm{𝟐}\mathbf{(}𝐑_𝐄\mathbf{)}}{\mathrm{𝟔𝟏𝟒𝟒}}\mathbf{+}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐄\mathbf{)}^\mathrm{𝟑}\mathbf{+}\mathrm{𝟏𝟐}𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐄\mathbf{)}𝐩_\mathrm{𝟐}\mathbf{(}𝐑_𝐄\mathbf{)}\mathbf{+}\mathrm{𝟒𝟖}𝐩_\mathrm{𝟑}\mathbf{(}𝐑_𝐄\mathbf{)}}{\mathrm{𝟐𝟗𝟒𝟗𝟏𝟐𝟎}}\mathbf{+}\mathbf{}`$
Now one has the following expantions:
$`𝑨\mathbf{(}\frac{𝐑}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}\mathrm{𝟏}\mathbf{}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐑\mathbf{)}}{\mathrm{𝟗𝟔}}\mathbf{+}\frac{\mathrm{𝟕}𝐩_\mathrm{𝟏}\mathbf{(}𝐑\mathbf{)}^\mathrm{𝟐}\mathbf{}\mathrm{𝟒}𝐩_\mathrm{𝟐}\mathbf{(}𝐑\mathbf{)}}{\mathrm{𝟗𝟐𝟏𝟔𝟎}}\mathbf{+}\mathbf{}`$
$`𝑳\mathbf{(}\frac{𝐑}{\mathrm{𝟒}}\mathbf{)}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐑\mathbf{)}}{\mathrm{𝟒𝟖}}\mathbf{+}\frac{\mathbf{}𝐩_\mathrm{𝟏}\mathbf{(}𝐑\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟕}𝐩_\mathrm{𝟐}\mathbf{(}𝐑\mathbf{)}}{\mathrm{𝟏𝟏𝟓𝟐𝟎}}\mathbf{+}\mathbf{}`$
Using these three expantions it is easy to obtain the following identities:
$`𝑨\mathbf{(}\frac{𝐑}{\mathrm{𝟐}}\mathbf{)}𝑴𝒂𝒚𝒆𝒓\mathbf{(}\frac{𝐑}{\mathrm{𝟐}}\mathbf{)}\mathbf{=}𝑳\mathbf{(}\frac{𝐑}{\mathrm{𝟒}}\mathbf{)}`$
$`𝑨\mathbf{(}𝑹\mathbf{)}𝑴𝒂𝒚𝒆𝒓\mathbf{(}𝑹\mathbf{)}\mathbf{=}𝑳\mathbf{(}\frac{𝐑}{\mathrm{𝟐}}\mathbf{)}`$
$`𝑨\mathbf{(}\mathrm{𝟐}𝑹\mathbf{)}𝑴𝒂𝒚𝒆𝒓\mathbf{(}\mathrm{𝟐}𝑹\mathbf{)}\mathbf{=}𝑳\mathbf{(}𝑹\mathbf{)}`$
$`𝑨\mathbf{(}\mathrm{𝟐}^𝒒𝑹\mathbf{)}𝑴𝒂𝒚𝒆𝒓\mathbf{(}\mathrm{𝟐}^𝒒𝑹\mathbf{)}\mathbf{=}𝑳\mathbf{(}\mathrm{𝟐}^{𝒒\mathbf{}\mathrm{𝟏}}𝑹\mathbf{)}`$
$`\mathbf{[}𝑨\mathbf{(}𝑹\mathbf{)}\mathrm{𝟐}^𝒌𝑴𝒂𝒚𝒆𝒓\mathbf{(}𝑹\mathbf{)}\mathbf{]}_{𝒕𝒐𝒑𝒇𝒐𝒓𝒎}\mathbf{=}𝑳\mathbf{(}𝑹\mathbf{)}_{𝒕𝒐𝒑𝒇𝒐𝒓𝒎}`$
With the help from these identities one has that:
$`\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{=}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
Using all these equations it is easy to obtain the following power expantion for Q:
$`\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟏𝟓𝟑𝟔}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟓𝟏𝟐}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟐}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟒𝟕𝟏𝟖𝟓𝟗𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟐𝟗𝟒𝟗𝟏𝟐𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟓𝟐𝟒𝟐𝟖𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟗𝟔𝟔𝟎𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟒}\mathbf{)}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟕𝟖𝟔𝟒𝟑𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟐}\mathbf{)}`$
When the bundle E is the tangent bundle and the bundle F is the normal bundle one obtain the usual power expantion for Q corresponding to the usual orientifold plane:
$`\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟏𝟓𝟑𝟔}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟓𝟏𝟐}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟒𝟕𝟏𝟖𝟓𝟗𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟐𝟗𝟒𝟗𝟏𝟐𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟓𝟐𝟒𝟐𝟖𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟗𝟔𝟔𝟎𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟕𝟖𝟔𝟒𝟑𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}`$
$`\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{=}\mathrm{𝟏}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟕𝟔𝟖}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟏𝟕𝟗𝟔𝟒𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathrm{𝟕}\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟒𝟕𝟒𝟓𝟔𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}`$
$`\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{=}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
For the other hand in the case of the y-Op-plane, the total Chern Class for a complex n-dimensional bundle V over the worldvolume has the following sumarization:
$`𝒄\mathbf{(}𝑽\mathbf{)}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟎}}^{\mathbf{}}𝒄_𝒋\mathbf{(}𝑻\mathbf{)}`$
also, the total Chern Class for the such bundle has the following factorization:
$`𝒄\mathbf{(}𝑽\mathbf{)}\mathbf{=}\mathbf{}_{𝒊\mathbf{=}\mathrm{𝟏}}^𝒏\mathbf{(}\mathrm{𝟏}\mathbf{+}𝒙_𝒊\mathbf{)}`$
The CHI-y- genus for the complex bundle V has the following formal factorisation:
$`𝑪𝑯𝑰_𝒚\mathbf{(}𝑽\mathbf{)}\mathbf{=}\mathbf{}_{𝒊\mathbf{=}\mathrm{𝟏}}^𝒏\frac{\mathbf{(}\mathrm{𝟏}\mathbf{+}\mathrm{𝐲𝐞𝐱𝐩}\mathbf{(}\mathbf{}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}𝐱_𝐢\mathbf{)}\mathbf{)}𝐱_𝐢}{\mathrm{𝟏}\mathbf{}\mathrm{𝐞𝐱𝐩}\mathbf{(}\mathbf{}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}𝐱_𝐢\mathbf{)}}`$
The CHI-y- genus for the complex bundle V has the following formal sumarisation in terms of the y-deformed Todd polynomials which are formed from the corresponding Chern classes and from the polynomials on y :
$`𝑪𝑯𝑰_𝒚\mathbf{(}𝑽\mathbf{)}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟎}}^{\mathbf{}}𝑻_𝒋\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}\mathbf{}\mathbf{,}𝒄_𝒋\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}`$
The y-Todd polynomials are given by:
$`𝑻_\mathrm{𝟎}\mathbf{(}𝒄_\mathrm{𝟎}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}𝑻_\mathrm{𝟎}\mathbf{(}\mathrm{𝟏}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\mathrm{𝟏}`$
$`𝑻_\mathrm{𝟏}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{(}\mathrm{𝟏}\mathbf{}𝐲\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟐}}`$
$`𝑻_\mathrm{𝟐}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟏𝟐}}`$
$`𝑻_\mathrm{𝟑}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟐}𝐲\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟐𝟒}}`$
$`𝑻_\mathrm{𝟒}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟒}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{(}\mathbf{}𝐲^\mathrm{𝟒}\mathbf{+}\mathrm{𝟒𝟕𝟒}𝐲^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟐𝟒}𝐲\mathbf{}\mathrm{𝟏}\mathbf{}\mathrm{𝟏𝟐𝟒}𝐲^\mathrm{𝟑}\mathbf{)}𝐜_\mathrm{𝟒}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{}\mathrm{𝟓𝟖}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟒}\mathbf{(}\mathrm{𝟑}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟒}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟒}\mathbf{)}}{\mathrm{𝟕𝟐𝟎}}`$
Now the relations between the Pontryaguin classes and the Chern Classes for the bundle V are given by the following formulas:
$`𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{=}\mathbf{}\mathrm{𝟐}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{+}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}^\mathrm{𝟐}`$
$`𝒑_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{=}\mathrm{𝟐}𝒄_\mathrm{𝟒}\mathbf{(}𝑽\mathbf{)}\mathbf{}\mathrm{𝟐}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{+}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}^\mathrm{𝟐}`$
Using these relations the y-deformed Todd polynomials can be written as follows:
$`𝑻_\mathrm{𝟎}\mathbf{(}𝒄_\mathrm{𝟎}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}𝑻_\mathrm{𝟎}\mathbf{(}\mathrm{𝟏}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\mathrm{𝟏}`$
$`𝑻_\mathrm{𝟏}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{(}\mathrm{𝟏}\mathbf{}𝐲\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟐}}`$
$`𝑻_\mathrm{𝟐}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐩_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟑}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟏𝟐}}`$
$`𝑻_\mathrm{𝟑}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟐}𝐲\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟐𝟒}}`$
$`𝑻_\mathrm{𝟒}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟒}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒚\mathbf{)}\mathbf{=}\frac{\mathbf{}\mathrm{𝟏𝟓}\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{+}\mathrm{𝟏𝟒}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟒}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟓}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟒}\mathbf{(}\mathrm{𝟕}𝐩_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{}𝐩_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟐}\mathbf{)}}{\mathrm{𝟕𝟐𝟎}}`$
When y=1 the y-deformed Todd polynomials are the same Hirzebruch polynomials:
$`𝑻_\mathrm{𝟎}\mathbf{(}𝒄_\mathrm{𝟎}\mathbf{(}𝑽\mathbf{)}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}𝑻_\mathrm{𝟎}\mathbf{(}\mathrm{𝟏}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}\mathrm{𝟏}\mathbf{=}𝑳_\mathrm{𝟎}`$
$`𝑻_\mathrm{𝟏}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}\frac{\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟐}}\mathbf{=}\mathrm{𝟎}`$
$`𝑻_\mathrm{𝟐}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}\frac{\mathbf{(}\mathrm{𝟏}\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐩_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟑}\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟏𝟐}}\mathbf{=}\frac{𝐩_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟑}}\mathbf{=}𝑳_\mathrm{𝟏}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{)}`$
$`𝑻_\mathrm{𝟑}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}\frac{\mathbf{}\mathbf{(}\mathrm{𝟏}\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}𝐜_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟐}\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐕\mathbf{)}}{\mathrm{𝟐𝟒}}\mathbf{=}\mathrm{𝟎}`$
$`𝑻_\mathrm{𝟒}\mathbf{(}𝒄_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟑}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒄_\mathrm{𝟒}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{,}\mathrm{𝟏}\mathbf{)}\mathbf{=}\frac{\mathbf{}\mathrm{𝟏𝟓}\mathbf{(}\mathrm{𝟏}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟏𝟒}\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟒}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟓}\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}\mathrm{𝟏}\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐕\mathbf{)}\mathbf{+}\mathbf{(}\mathrm{𝟏}\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟒}\mathbf{(}\mathrm{𝟕}𝐩_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{}𝐩_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟐}\mathbf{)}}{\mathrm{𝟕𝟐𝟎}}\mathbf{=}\frac{\mathrm{𝟕}𝐩_\mathrm{𝟐}\mathbf{(}𝐕\mathbf{)}\mathbf{}𝐩_\mathrm{𝟏}\mathbf{(}𝐕\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟒𝟓}}\mathbf{=}𝑳_\mathrm{𝟐}\mathbf{(}𝒑_\mathrm{𝟏}\mathbf{(}𝑽\mathbf{)}\mathbf{,}𝒑_\mathrm{𝟐}\mathbf{(}𝑽\mathbf{)}\mathbf{)}`$
Using all these expretions one can to obtain the following expantion:
$`𝑪𝑯𝑰_𝒚\mathbf{(}\frac{𝐑_𝐕}{\mathrm{𝟒}}\mathbf{)}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟏}\mathbf{}𝐲\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}}{\mathrm{𝟖}}\mathbf{+}\frac{\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}\mathbf{+}\mathrm{𝟑}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟐}\mathbf{(}𝐑_𝐕\mathbf{)}}{\mathrm{𝟏𝟗𝟐}}\mathbf{+}\frac{\mathbf{}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}𝐜_\mathrm{𝟐}\mathbf{(}𝐑_𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟐}𝐲\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐑_𝐕\mathbf{)}}{\mathrm{𝟏𝟓𝟑𝟔}}\mathbf{+}\frac{\mathbf{}\mathrm{𝟏𝟓}\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{+}\mathrm{𝟏𝟒}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟒}\mathbf{(}𝐑_𝐕\mathbf{)}\mathbf{+}\mathrm{𝟏𝟓}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}𝐜_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐑_𝐕\mathbf{)}\mathbf{+}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟒}\mathbf{(}\mathrm{𝟕}𝐩_\mathrm{𝟐}\mathbf{(}𝐑_𝐕\mathbf{)}\mathbf{}𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}^\mathrm{𝟐}\mathbf{)}}{\mathrm{𝟏𝟖𝟒𝟑𝟐𝟎}}\mathbf{+}\mathbf{}`$
When the first chern class of V is trivial, one obtain, using again the relations between pontryaguin classes and Chern classes, the following result:
$`𝑪𝑯𝑰_𝒚\mathbf{(}\frac{𝐑_𝐕}{\mathrm{𝟒}}\mathbf{)}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟐}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{}\mathrm{𝟑}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{)}𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}}{\mathrm{𝟑𝟖𝟒}}\mathbf{+}\frac{\mathrm{𝟏𝟐}𝐲\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}𝐜_\mathrm{𝟑}\mathbf{(}𝐑_𝐕\mathbf{)}}{\mathrm{𝟏𝟓𝟑𝟔}}\mathbf{+}\frac{\mathbf{(}\mathbf{}\mathrm{𝟔𝟎}\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{+}\mathrm{𝟏𝟒}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟓𝟔}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟒}\mathbf{)}𝐩_\mathrm{𝟐}\mathbf{(}𝐑_𝐕\mathbf{)}\mathbf{}\mathbf{(}\mathbf{}\mathrm{𝟏𝟓}\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{+}\mathrm{𝟏𝟒}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝐲\mathbf{}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟖}\mathbf{(}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟒}\mathbf{)}𝐩_\mathrm{𝟏}\mathbf{(}𝐑_𝐕\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟏𝟒𝟕𝟒𝟓𝟔𝟎}}\mathbf{+}\mathbf{}`$
Finally using this last expansion and the relations between the Pontryaguin classes and the 2-form curvature, one can to obtain the following development for the Q of the yOp-planes:
$`\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟔𝟏𝟒𝟒}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟑}}{\mathrm{𝟐𝟓𝟔}}𝒚\mathbf{(}𝒚\mathbf{}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝒄_\mathrm{𝟑}\mathbf{(}𝑹_𝑻\mathbf{)}\mathbf{}𝒄_\mathrm{𝟑}\mathbf{(}𝑹_𝑵\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟕𝟓𝟒𝟗𝟕𝟒𝟕𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒚^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}𝒚\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟖𝟖𝟕𝟒𝟑𝟔𝟖𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}\mathbf{}\mathrm{𝟒}𝒚^\mathrm{𝟒}\mathbf{}\mathrm{𝟒𝟗𝟔}𝒚^\mathrm{𝟑}\mathbf{+}\mathrm{𝟏𝟖𝟗𝟔}𝒚^\mathrm{𝟐}\mathbf{}\mathrm{𝟒𝟗𝟔}𝒚\mathbf{}\mathrm{𝟒}\mathbf{)}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}`$
When y=1, then one obtain the development for the Q of the usual Op-plane:
$`\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{=}\mathrm{𝟏}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟏}^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟔𝟏𝟒𝟒}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟑}}{\mathrm{𝟐𝟓𝟔}}\mathrm{𝟏}\mathbf{(}\mathrm{𝟏}\mathbf{}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝒄_\mathrm{𝟑}\mathbf{(}𝑹_𝑻\mathbf{)}\mathbf{}𝒄_\mathrm{𝟑}\mathbf{(}𝑹_𝑵\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟕𝟓𝟒𝟗𝟕𝟒𝟕𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}\mathrm{𝟏}^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟖𝟖𝟕𝟒𝟑𝟔𝟖𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}\mathbf{}\mathrm{𝟒}\mathbf{}\mathrm{𝟒𝟗𝟔}\mathbf{+}\mathrm{𝟏𝟖𝟗𝟔}\mathbf{}\mathrm{𝟒𝟗𝟔}\mathbf{}\mathrm{𝟒}\mathbf{)}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}`$
$`\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{=}\mathrm{𝟏}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟕𝟔𝟖}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟏𝟕𝟗𝟔𝟒𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathrm{𝟕}\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟒𝟕𝟒𝟓𝟔𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}`$
$`\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_\mathrm{𝟏}\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{=}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}`$
## 3 The Elementary Processes
The WZ action for the usual orientifold p-plane can be writen as a sum of the WZ actions for three elementary processes:
$`𝑺_{𝑾𝒁\mathbf{(}𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟏}}^\mathrm{𝟑}𝑺_{𝑾𝒁\mathbf{(}𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}𝒋}`$
The WZ actions for the three elementary processes are given by the following expretions:
$`𝑺_{𝑾𝒁\mathbf{(}𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟏}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{+}\mathrm{𝟏}}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟐}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟑}}\mathbf{[}\mathbf{}\mathbf{(}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟕𝟔𝟖}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{)}\mathbf{]}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟑}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟕}}\mathbf{(}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟏𝟕𝟗𝟔𝟒𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathrm{𝟕}\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟒𝟕𝟒𝟓𝟔𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}\mathbf{)}`$
The first WZ action describes an elementary process for which the usual orientifold p-plane emites one (p+1)-form RR potential. The second WZ action describes an elementary process for which the usual Op-plane absorbs two gravitons and emits one (p-3)-form RR potential. The third WZ action describes an elementary process for which the Op-plane absorbs four gravitons and emits one (p-7)-form RR potential.
From the result of the section two, the WZ action for a generalized orientifold p-plane can be writen as a sum of the WZ actions for some elementary processes:
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟏}}^\mathrm{𝟔}𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}𝒋}`$
The WZ actions for the six elementary processes are given by the following expretions:
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟏}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{+}\mathrm{𝟏}}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟐}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟑}}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟏𝟓𝟑𝟔}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟑}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟑}}\mathbf{(}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟓𝟏𝟐}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟐}\mathbf{)}\mathbf{)}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟒}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟕}}\mathbf{(}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟒𝟕𝟏𝟖𝟓𝟗𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{+}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟐𝟗𝟒𝟗𝟏𝟐𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}\mathbf{)}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟓}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟕}}\mathbf{(}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟓𝟐𝟒𝟐𝟖𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟗𝟔𝟔𝟎𝟖}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟒}\mathbf{)}\mathbf{)}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝑮𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟔}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟕}}\mathbf{(}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟕𝟖𝟔𝟒𝟑𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{(}𝒕𝒓𝑹_𝑬^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑭^\mathrm{𝟐}\mathbf{)}\mathbf{)}`$
The first WZ action describes an elementary process for which the generalized orientifold p-plane emites one (p+1)-form RR potential. The second WZ action describes an elementary process for which the generalized Op-plane absorbs two gravitons and emits one (p-3)-form RR potential. The third WZ actuib describes an elementary process for which the generalized Op-plane absorbs two gaugeons and emits one (p-3)-form RR potential. The fourth WZ action describes an elementary process for which the GOp-plane absorbs four gravitons and emits one (p-7)-form RR potential. The fifth WZ action describes an elementary process for which the GOp-plane absorbs four gaugeons and emits one (p-7)-form RR potential. The sixth WZ action describes an elementary process for which the GOp-planes absorbs two gravitons and two gaugeons and emits one (p-7)-form RR potential.
When the gaugeons corresponding to the bundles E and F are the same gravitons corresponding to the bundles T and N respectively, then the six elementary process for the GOp-plane are reduced to the usuals three elementary process for the usual Op-plane: Op-plane emites one (p+1)-form RR potential,Op-plane absorbs two gravitons and emits one (p-3)-form RR potential; and, Op-plane absorbs four gravitons and emits one (p-7)-form RR potential.
Of other hand, from the result of the section two, the WZ action for a y-deformed orientifold p-plane can be writen as a sum of the WZ actions for some elementary processes:
$`𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}}\mathbf{=}\mathbf{}_{𝒋\mathbf{=}\mathrm{𝟏}}^\mathrm{𝟒}𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}𝒋}`$
The WZ actions for the four elementary processes are given by the following expretions:
$`𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟏}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{+}\mathrm{𝟏}}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟐}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟑}}\mathbf{[}\mathbf{}\mathbf{(}\frac{\mathbf{(}𝐲^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}𝐲\mathbf{+}\mathrm{𝟏}\mathbf{)}\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟐}}{\mathrm{𝟔𝟏𝟒𝟒}𝐩𝐢^\mathrm{𝟐}}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}\mathbf{)}\mathbf{]}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟑}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟓}}\mathbf{(}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟑}}{\mathrm{𝟐𝟓𝟔}}𝒚\mathbf{(}𝒚\mathbf{}\mathrm{𝟏}\mathbf{)}\mathbf{(}𝒄_\mathrm{𝟑}\mathbf{(}𝑹_𝑻\mathbf{)}\mathbf{}𝒄_\mathrm{𝟑}\mathbf{(}𝑹_𝑵\mathbf{)}\mathbf{)}\mathbf{)}\mathbf{}`$
$`𝑺_{𝑾𝒁\mathbf{(}𝒚𝑶𝒑\mathbf{}𝒑𝒍𝒂𝒏𝒆\mathbf{)}\mathbf{,}\mathrm{𝟒}}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪_{𝒑\mathbf{}\mathrm{𝟕}}\mathbf{(}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟕𝟓𝟒𝟗𝟕𝟒𝟕𝟐}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}𝒚^\mathrm{𝟐}\mathbf{}\mathrm{𝟏𝟎}𝒚\mathbf{+}\mathrm{𝟏}\mathbf{)}^\mathrm{𝟐}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟐}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟐}\mathbf{)}^\mathrm{𝟐}\mathbf{}\frac{\mathbf{(}\mathrm{𝟒}𝐩𝐢^\mathrm{𝟐}\mathrm{𝐚𝐥𝐟𝐚}\mathbf{)}^\mathrm{𝟒}}{\mathrm{𝟏𝟖𝟖𝟕𝟒𝟑𝟔𝟖𝟎}𝐩𝐢^\mathrm{𝟒}}\mathbf{(}\mathbf{}\mathrm{𝟒}𝒚^\mathrm{𝟒}\mathbf{}\mathrm{𝟒𝟗𝟔}𝒚^\mathrm{𝟑}\mathbf{+}\mathrm{𝟏𝟖𝟗𝟔}𝒚^\mathrm{𝟐}\mathbf{}\mathrm{𝟒𝟗𝟔}𝒚\mathbf{}\mathrm{𝟒}\mathbf{)}\mathbf{(}𝒕𝒓𝑹_𝑻^\mathrm{𝟒}\mathbf{}𝒕𝒓𝑹_𝑵^\mathrm{𝟒}\mathbf{)}\mathbf{)}`$
The first WZ action describes an elementary process on which the yOp-plane emites one (p+1)-form RR potential. The second WZ action describes an elementary process for which the y-deformed Op-plane absorbs two gravitons and emits one (p-3)-form RR potential. The third WZ action describes an elementary process for which the y-deformed Op-plane absorbs three gravitons and emits one (p-5)-form RR potential. The fourth WZ action describes an elementary process for which the yOp-plane absorbs four gravitons and emits one (p-7)-form RR potential. When y=1,then the four elementary process for the yOp-plane are reduced to the usuals three elementary process for the usual Op-plane: Op-plane emites one (p+1)-form RR potential,Op-plane absorbs two gravitons and emits one (p-3)-form RR potential; and, Op-plane absorbs four gravitons and emits one (p-7)-form RR potential.
## 4 Conclutions
The WZ action for the GOp-planes can be modified or extended by various ways. When the bundles haven non-trivial second Stiefel-Whitney classes one can to write the following WZ action which incorporates an effect of the magnetic monopoles:
$`𝑺_{𝑾𝒁}\mathbf{=}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}𝐞^{\frac{𝐝_\mathrm{𝟏}}{\mathrm{𝟐}}}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}𝐞^{\frac{𝐝_\mathrm{𝟐}}{\mathrm{𝟐}}}}}`$
where:
$`𝒅_\mathrm{𝟏}\mathbf{=}𝒓𝒆𝒅𝒖𝒄𝒕𝒊𝒐𝒏\mathbf{.}𝒎𝒐𝒅\mathbf{.2}\mathbf{(}𝒘_\mathrm{𝟐}\mathbf{(}𝑻\mathbf{)}\mathbf{+}𝒘_\mathrm{𝟐}\mathbf{(}𝑬\mathbf{)}\mathbf{)}`$
$`𝒅_\mathrm{𝟐}\mathbf{=}𝒓𝒆𝒅𝒖𝒄𝒕𝒊𝒐𝒏\mathbf{.}𝒎𝒐𝒅\mathbf{.2}\mathbf{(}𝒘_\mathrm{𝟐}\mathbf{(}𝑵\mathbf{)}\mathbf{+}𝒘_\mathrm{𝟐}\mathbf{(}𝑭\mathbf{)}\mathbf{)}`$
This action describes processes on which the GOp-plane emites RR-forms and absorbs gravitons, gaugeons and magnetic monopoles.
From the other side one can to write the following actions for GOp-planes non standard:
$`𝑺_{𝑾𝒁}\mathbf{=}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\mathbf{(}\mathrm{𝟐}\sqrt{\frac{𝐀\mathbf{(}𝐑_𝐓\mathbf{)}}{𝐀\mathbf{(}𝐑_𝐍\mathbf{)}}}\mathbf{}\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{)}`$
$`𝑺_{𝑾𝒁}\mathbf{=}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\mathbf{(}\sqrt{\frac{𝐀\mathbf{(}𝐑_𝐓\mathbf{)}}{𝐀\mathbf{(}𝐑_𝐍\mathbf{)}}}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{)}`$
In the same way, one can to write the following actions for yOp-planes non standard:
$`𝑺_{𝑾𝒁}\mathbf{=}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\mathbf{(}\mathrm{𝟐}\sqrt{\frac{𝐀\mathbf{(}𝐑_𝐓\mathbf{)}}{𝐀\mathbf{(}𝐑_𝐍\mathbf{)}}}\mathbf{}\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{)}`$
$`𝑺_{𝑾𝒁}\mathbf{=}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\mathbf{(}\sqrt{\frac{𝐀\mathbf{(}𝐑_𝐓\mathbf{)}}{𝐀\mathbf{(}𝐑_𝐍\mathbf{)}}}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{)}`$
These actions correspond respectively to the Sp-type yOp-planes and the yOp-planes that give rise to gauge symmetries of type SO(2n+1). Such non-standard yOp-planes are building from combinations of the D-p-branes and standard yOp-planes.
By combination of Dp-branes, Op-planes, GOp-planes and yOp-planes one can to have gauge teories with symmetries Sp and SO-odd whose WZ actions are give respectively by:
$`𝑺_{𝑾𝒁}\mathbf{=}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\mathbf{(}\mathrm{𝟐}\sqrt{\frac{𝐀\mathbf{(}𝐑_𝐓\mathbf{)}}{𝐀\mathbf{(}𝐑_𝐍\mathbf{)}}}\mathbf{}\frac{\mathrm{𝟏}}{\mathrm{𝟑}}\mathbf{(}\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{+}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{+}\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{)}\mathbf{)}`$
$`𝑺_{𝑾𝒁}\mathbf{=}\frac{𝐓_𝐩}{\mathrm{𝐤𝐚𝐩𝐩𝐚}}\mathbf{}_{𝒑\mathbf{+}\mathrm{𝟏}}𝑪\mathbf{(}\sqrt{\frac{𝐀\mathbf{(}𝐑_𝐓\mathbf{)}}{𝐀\mathbf{(}𝐑_𝐍\mathbf{)}}}\mathbf{}\mathrm{𝟐}^{𝒑\mathbf{}\mathrm{𝟒}}\frac{\mathrm{𝟏}}{\mathrm{𝟑}}\mathbf{(}\sqrt{\frac{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{\mathrm{𝐂𝐇𝐈}_𝐲\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{+}\sqrt{\frac{𝐋\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟒}}\mathbf{)}}{𝐋\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟒}}\mathbf{)}}}\mathbf{+}\sqrt{\frac{𝐀\mathbf{(}\frac{𝐑_𝐓}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐄}{\mathrm{𝟐}}\mathbf{)}}{𝐀\mathbf{(}\frac{𝐑_𝐍}{\mathrm{𝟐}}\mathbf{)}\mathrm{𝐌𝐚𝐲𝐞𝐫}\mathbf{(}\frac{𝐑_𝐅}{\mathrm{𝟐}}\mathbf{)}}}\mathbf{)}\mathbf{)}`$
Finally one can to think about non-BPS GOp-planes and non-BPS yOp-planes with the tachyon effect.
In conclution gauge theories with symmetries SO-even,Sp and SO-odd can be obtained from the combination of the Dp-branes,Op-planes, GOp-planes and yOp-planes of the string theory.
## 5 References
About WZ action for usual orientifold planes:
K. Dasgupta, D. Jatkar and S. Mukhi, Gravitational couplings and Z2 orientifolds, Nucl. Phys. B523 (1998) 465, hep-th/9707224.
K. Dasgupta and S. Mukhi, Anomaly inflow on orientifold planes, J. High Energy Phys. 3 (1998) 4, hep-th/9709219.
J. Morales, C. Scrucca and M. Serone, Anomalous couplings for D-branes and O-planes, hep-th/9812071.
B.Stefanski,Jr., Gravitational Couplings of D-branes and O-planes, hep-th/9812088
Ben Craps and Frederik Roose, (Non-)Anomalous D-brane and O-plane couplings:the normal bundle, hep-th/9812149.
About WZ action for non-standard orientifold planes:
Sunil Mukhi and Nemani V. Suryanarayana, Gravitational Couplings, Orientifolds and M-Planes, hep-th/9907215
About Mayer class , Mayer integrality theorem and CHI-y-genus:
F. Hirzebruch, Topological Methods in Algebraic Geometry, 1978
Christian Bar, Elliptic Symbols, december 1995, Math. Nachr. 201, 7-35 (1999)
Taras E. Panov, Calculation of Hirzebruch Genera for manifolds acted on by the Group Z/p via invariantes of the action, math.AT/9909081
|
warning/0006/cond-mat0006443.html
|
ar5iv
|
text
|
# Projecting the Kondo Effect: Theory of the Quantum Mirage
\[
## Abstract
A microscopic theory is developed for the projection (quantum mirage) of the Kondo resonance from one focus of an elliptic quantum corral to the other focus. The quantum mirage is shown to be independent of the size and the shape of the ellipse, and experiences $`\lambda _F/4`$ oscillations ($`\lambda _F`$ is the surface-band Fermi wavelength) with an increasing semimajor axis length. We predict an oscillatory behavior of the mirage as a function of a weak magnetic field applied perpendicular to the sample.
\]
In a recent experiment, Manoharan et al. used an elliptic quantum corral to project the image of a Kondo resonance over a distance of tens of angstroms, from one focus of the ellipse to the other focus. By placing a magnetic Co atom at one focus of the ellipse and measuring the tunneling current to a close-by scanning tunneling microscope (STM) tip, a distinctive Kondo resonance was seen in the $`I`$-$`V`$ curve when the tip was brought directly above the Co adatom. Remarkably, a similar Kondo signature was observed when the tip was placed above the empty focus, indicating coherent refocusing of the spectral image by the surrounding corral. This should be contrasted with STM measurements of isolated magnetic adatoms on open surfaces , where a limited spatial extent of $`10`$Å was observed for the Kondo effect.
Semiclassically, one can attribute this refocusing phenomena to the property that all classical paths leaving one focus of the ellipse bounce specularly off the perimeter and converge onto the second focus with the same acquired phase (see Fig. 1). However, this simple picture does not explain the quantitative features of the experiment. For example, the complex interference patterns in the $`dI/dV`$ difference map throughout the ellipse, or the $`\lambda _F/4`$ oscillations of the mirage with an increasing semimajor axis length $`a`$ ($`\lambda _F`$ is the Fermi wavelength). Explanation of these features requires a quantitative theory, which is the objective of the present Letter.
Starting with a microscopic picture of Kondo scattering off the Co adatoms we obtain good qualitative and quantitative agreement with the experiment. We establish a remarkable feature of the quantum mirage which, aside from the $`\lambda _F/4`$ oscillations mentioned above, is independent of the size and the shape of the ellipse, provided the ellipses is not too small. In particular, there is no dependence on the ellipse eccentricity $``$, see Fig. 1. In the presence of a weak perpendicular magnetic field, we predict an oscillatory behavior of the quantum mirage as a function of the magnetic flux encircled by the ellipse.
The Cu(111) surface has a band of surface states, which acts as a two-dimensional electron gas. The surface band starts $`450\mathrm{m}\mathrm{e}\mathrm{V}`$ below the Fermi energy, and has a Fermi wave number of $`k_F^14.75`$Å. When a Co adatom is placed on the surface, it scatters both the surface electrons and the underlying bulk electrons. As recently shown by Újsághy et al. for Co on Au(111), one may model the Co adatom by an effective nondegenerate Anderson impurity , characterized by an effective energy level $`ϵ_d`$, an on-site repulsion $`U`$, and two hybridization matrix elements $`t_s`$ and $`t_b`$ to the underlying surface and bulk conduction electrons. In this manner, each of the Co atoms forming the ellipse in the experiment of Manoharan et al. acts as an Anderson impurity, as does the adatom placed inside the ellipse.
Denoting the creation of a surface-state and a bulk conduction electron by $`c_{\stackrel{}{k}\sigma }^{}`$ and $`a_{\stackrel{}{q}\sigma }^{}`$, respectively (here $`\stackrel{}{k}`$ labels a two-dimensional surface vector while $`\stackrel{}{q}`$ is a three-dimensional vector), we model the system by the Hamiltonian $`=_{\mathrm{surf}}+_{\mathrm{bulk}}+_{i=0}^N_{\mathrm{imp}}(\stackrel{}{R}_i)`$, where
$`_{\mathrm{surf}}={\displaystyle \underset{\stackrel{}{k}\sigma }{}}ϵ_\stackrel{}{k}c_{\stackrel{}{k}\sigma }^{}c_{\stackrel{}{k}\sigma }\text{and}_{\mathrm{bulk}}={\displaystyle \underset{\stackrel{}{q}\sigma }{}}E_\stackrel{}{q}a_{\stackrel{}{q}\sigma }^{}a_{\stackrel{}{q}\sigma }`$ (1)
describe the free surface and bulk conduction bands, respectively, and
$`_{imp}(\stackrel{}{R}_i)=`$ $`ϵ_d{\displaystyle \underset{\sigma }{}}d_{i\sigma }^{}d_{i\sigma }+Ud_i^{}d_id_i^{}d_i`$ (3)
$`+{\displaystyle \underset{\sigma }{}}\{t_sd_{i\sigma }^{}\psi _\sigma (\stackrel{}{R}_i)+t_bd_{i\sigma }^{}\chi _\sigma (\stackrel{}{R}_i)+\mathrm{h}.\mathrm{c}.\}`$
describes a Co adatom at point $`\stackrel{}{R}_i`$ on the surface. Here $`d_{i\sigma }^{}`$ creates a localized Co electron at site $`\stackrel{}{R}_i`$ ($`i=0`$ for the inner adatom and $`i=1,\mathrm{},N`$ for the perimeter adatoms), while $`\psi _\sigma (\stackrel{}{R}_i)`$ and $`\chi _\sigma (\stackrel{}{R}_i)`$ annihilate, respectively, a surface and a bulk conduction electron at site $`\stackrel{}{R}_i`$. For simplicity, we have taken the different adatoms to be identical, and neglected any momentum dependence of $`t_s`$ and $`t_b`$. In what follows, we shall mainly be interested in the case where the inner adatom is located at the left focus, i.e., $`\stackrel{}{R}_0=\stackrel{}{R}`$ in the notations of Fig. 1.
Consider now an STM tip placed directly above the surface point $`\stackrel{}{r}`$. If the tip couples predominantly to the underlying surface-state electrons at $`\stackrel{}{r}`$, then the differential conductance for the current through the STM tip measures, up to thermal broadening, the local surface-electron density of states at point $`\stackrel{}{r}`$, $`\rho (\stackrel{}{r},ϵ)`$. For an isolated Co impurity, $`\rho (\stackrel{}{r},ϵ)`$ depends on the Kondo scattering from the impurity as described, e.g., in Ref. . For the multiple-impurity configuration considered here there are two main modifications: (i) There are multiple scattering off the different Co adatoms; (ii) Intersite correlations alter the Kondo scattering off each Co adatom. Due to the relatively large distance between Co atoms ($`10`$Å for neighboring atoms on the ellipse perimeter), we expect the latter effect to be small, and therefore neglect it hereafter.
Since we are mostly interested in the effect of the Co atom placed inside the ellipse, we distinguish it from the other Co atoms on the perimeter of the ellipse. Neglecting intersite correlations, $`\rho (\stackrel{}{r},ϵ)`$ takes the form $`\rho (\stackrel{}{r},ϵ)=\overline{\rho }(\stackrel{}{r},ϵ)+\delta \rho (\stackrel{}{r},ϵ)`$, where
$`\overline{\rho }(\stackrel{}{r},ϵ)={\displaystyle \frac{1}{\pi }}\text{Im}\{G(\stackrel{}{r},\stackrel{}{r};ϵ)\}`$ (4)
is the density of states of an empty ellipse (i.e., in the absence of the inner adatom), and
$$\delta \rho (\stackrel{}{r},ϵ)=\frac{1}{\pi }\mathrm{Im}\{t_s^2G(\stackrel{}{r},\stackrel{}{R}_0;ϵ)G_d(ϵ)G(\stackrel{}{R}_0,\stackrel{}{r};ϵ)\}$$
(5)
is the additional contribution due to the extra Co atom at $`\stackrel{}{R}_0`$. Here $`G(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{};ϵ)`$ is the retarded Green function of the surface electrons for an empty ellipse, and $`G_d(ϵ)`$ is the fully dressed retarded Green function of the $`d`$ electrons of the inner adatom. Note that in writing Eq. (5) we have assumed that the three-dimensional propagation of bulk electrons between different Co sites on the surface is small compared to the two-dimensional propagation of the surface electrons. This assumption is quite reasonable considering that the three-dimensional propagation near the surface decays as $`1/r^2`$, compared to $`1/\sqrt{r}`$ for the two-dimensional surface propagation .
Experimentally, $`\delta \rho (\stackrel{}{r},ϵ)`$ is extracted by first measuring the local density of states of the empty ellipse, and then subtracting it from the measured density of states with the extra Co atom. To compute $`\delta \rho (\stackrel{}{r},ϵ)`$, one needs to evaluate $`G(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{};ϵ)`$, which is our next goal. To this end, we introduce the $`N\times N`$ matrix $`g_{ij}(1\delta _{i,j})G_0(\stackrel{}{R}_i,\stackrel{}{R}_j)`$, along with the two vector quantities, $`v_i=G_0(\stackrel{}{r},\stackrel{}{R}_i)`$ and $`u_i=G_0(\stackrel{}{R}_i,\stackrel{}{r}{}_{}{}^{})`$. Here $`G_0(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{})`$ is the free surface Green function without the corral, and $`i`$ and $`j`$ run over $`1,\mathrm{},N`$. Using these quantities, $`G(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{};ϵ)`$ is compactly expressed as
$$G(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{};ϵ)=G_0(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{})+\underset{i,j=1}{\overset{N}{}}v_i\left[\frac{1}{1Tg}T\right]_{ij}u_j,$$
(6)
where $`T(ϵ)=t_s^2G_d(ϵ)`$ is the surface-to-surface component of the conduction-electron scattering $`T`$-matrix at each Co site. Again, Eq. (6) omits the intersite correlations and the bulk propagation between different Co sites. Finally, for temperatures and energies below the Kondo temperature, the Kondo part of $`G_d(ϵ)`$ may be well approximated by the Lorentzian form $`Z_K/(ϵϵ_F+iT_K)`$, where $`T_K`$ is the Kondo temperature, $`ϵ_F`$ is the Fermi energy, and
$$Z_K=\frac{T_K}{\pi \rho _st_s^2+\pi \rho _bt_b^2}$$
(7)
is the corresponding weight. Here $`\rho _s`$ and $`\rho _b`$ are the surface and bulk density of states at the Fermi level.
A key parameter that enters the quantum mirage is the ratio of scattering rates
$$t=\frac{\pi \rho _st_s^2}{\pi \rho _st_s^2+\pi \rho _bt_b^2}.$$
(8)
Physically, $`t`$ represents the probability that a surface-state electron impinging on a Co adatom will be scattered to a surface-state electron rather than a bulk electron. Hence $`t`$ is a measure of the in-elasticity of the scattering of surface waves from the Co impurities. In the theory of Heller et al. for the standing waves formed in a quantum corral, $`t`$ is found to be $`1/2`$. Hereafter we shall use the same value for $`t`$.
In Fig. 2 we depict $`\delta \rho (\stackrel{}{r},ϵ_F)`$ for various configurations of the Co atoms, as measured by Manoharan et al. . The upper panels correspond to an ellipse with eccentricity $`=0.786`$ and 34 adatoms (ellipse b of Ref. ), while the lower panels correspond to an ellipse with eccentricity $`=0.5`$ and 36 adatoms (ellipse a in Ref. ). The quantum mirage is clearly seen in each of the left two panels, where an additional adatom has been placed at the left focus of the ellipse. For both ellipses, there is a strong signal in the tunneling density of states right above the right focus, in accordance with the experimental data. By contrast, the quantum mirage disappears when the additional adatom is places off the focus, as shown in the right two panels. These results are in good agreement with the experimental measurements, reproducing even fine details of the experimental patterns.
The results of Fig. 2 were obtained from Eqs. (5) and (6), by setting $`t=1/2`$ and approximating $`G_0`$ with the free two-dimensional Green function:
$`G_0(\stackrel{}{r},\stackrel{}{r}{}_{}{}^{})=i\pi \rho _sH_0^{(1)}(k|\stackrel{}{r}\stackrel{}{r}{}_{}{}^{}|).`$ (9)
Here $`k`$ is the wave number, and $`H_0^{(1)}(x)`$ is the Hankel function of zeroth order, which for $`x1`$ takes the asymptotic form
$`H_0^{(1)}(x)\sqrt{{\displaystyle \frac{2}{\pi x}}}\mathrm{exp}\left(ixi{\displaystyle \frac{\pi }{4}}\right).`$ (10)
The good agreement between our calculations with $`t=1/2`$ and the experimental data indicates that the number of scattering events that a particle undergoes before leaving the surface bounded by the ellipse is small. This suggests the possibility of calculating the quantum mirage at the right focus at the Fermi level, $`\delta \rho (\stackrel{}{R},ϵ_F)`$, using perturbation theory in $`t`$. Thus, to linear order in $`t`$, the Green function between the left and the right foci is given by
$`G(\stackrel{}{R},\stackrel{}{R};ϵ_F)G_0(\stackrel{}{R},\stackrel{}{R})+G_1(\stackrel{}{R},\stackrel{}{R})+\mathrm{},`$ (11)
where
$`G_0(\stackrel{}{R},\stackrel{}{R})i\rho _s\sqrt{{\displaystyle \frac{\pi }{k_Fa}}}e^{i2k_Fai\frac{\pi }{4}}`$ (12)
is the contribution of the direct path connecting the two foci (illustrated by the dashed line in Fig. 1), and
$`G_1(\stackrel{}{R},\stackrel{}{R})={\displaystyle \frac{t}{i\pi \rho _s}}{\displaystyle \underset{j=1}{\overset{N}{}}}G_0(\stackrel{}{R},\stackrel{}{R}_j)G_0(\stackrel{}{R}_j,\stackrel{}{R})`$ (13)
comes form all trajectories which scatter from a single Co adatom on the ellipse perimeter (solid lines in Fig. 1).
Conventionally, the second term $`G_1`$ is smaller than $`G_0`$ for several reasons. First, each scattering event is inelastic, and therefore introduces a reduction factor of $`t`$. Second, the scattered orbits are longer. Third, the various orbits have generally different lengths, and therefore their corresponding phases add up incoherently. The situation is quite different in the present case. Due to the defining property of the ellipse, the length of all scattered orbits between the two foci is precisely the same, and hence their contributions add up coherently. Consequently, the second term in Eq. (11) takes the form
$`G_1(\stackrel{}{R},\stackrel{}{R})i\rho _s{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{2t}{k_F\sqrt{\xi _{1,j}\xi _{2,j}}}}e^{i2k_Fai\frac{\pi }{2}},`$ (14)
where $`2a`$ is the length of each orbit, and $`\xi _{1,j}`$ and $`\xi _{2,j}`$ are the distances between the impurity at $`\stackrel{}{R}_j`$ and the right and left foci, respectively (see Fig. 1). Finally, we approximate the sum in Eq. (14) by an integral:
$`{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{1}{\sqrt{\xi _{1,j}\xi _{2,j}}}}{\displaystyle \frac{1}{d}}{\displaystyle 𝑑s\frac{1}{\sqrt{\xi _1(s)\xi _2(s)}}},`$ (15)
where $`s`$ denotes the coordinate along the ellipse contour, and $`d`$ is the mean distance between adjacent adatoms. The result of the integral is independent of the eccentricity of the ellipse, and is simply $`2\pi `$. Thus, the contribution of orbits scattered from a single perimeter adatom is
$`G_1(\stackrel{}{R},\stackrel{}{R})\rho _s{\displaystyle \frac{4\pi t}{k_Fd}}e^{i2k_Fa}.`$ (16)
Comparing $`G_0(\stackrel{}{R},\stackrel{}{R})`$ and $`G_1(\stackrel{}{R},\stackrel{}{R})`$ at the Fermi energy, one sees that the leading contribution to the quantum mirage comes from $`G_1`$ , provided
$$\frac{d}{a}\frac{16\pi t^2}{k_Fd}.$$
(17)
Substituting the experimental parameters, $`d10`$Å, $`a70`$Å, and $`k_F1/4.75`$Å<sup>-1</sup>, and setting $`t=1/2`$, it is straightforward to verify that Eq. (17) holds for all ellipses with eccentricity $`0.05<<1`$. Furthermore, Eq. (17) is always satisfied for sufficiently large ellipses, provided the mean distance between adjacent adatoms is kept fixed ($`ad`$).
Neglecting the contribution of the direct path, $`G_0(\stackrel{}{R},\stackrel{}{R})`$, and using Eqs. (5) and (16) with $`\stackrel{}{r}=\stackrel{}{R}`$, the resulting local density of states at the Fermi energy takes the form
$$\delta \rho (\stackrel{}{R},ϵ_F)\rho _s\frac{16t^3}{(k_Fd)^2}\mathrm{cos}(4k_Fa).$$
(18)
The main feature of the above result is the robustness of the quantum mirage: As long as condition (17) is satisfied, the amplitude of the mirage is independent of the size of the ellipse, $`a`$, and its eccentricity, $``$. Rather, the amplitude is determined by $`t`$, which characterizes the in-elasticity of the scattering of surface waves from adatoms, and the dimensionless mean distance between adjacent adatoms along the ellipse, $`k_Fd`$. The oscillations of the mirage as function of $`a`$ are indeed periodic with a period of $`\lambda _F/4`$, as seen experimentally .
Next we consider the effect of a weak uniform magnetic field, $`B`$, applied perpendicular to the surface. As shown below, the quantum mirage experiences a distinctive oscillatory behavior as a function of the magnetic field, which depends on the size and the shape of the ellipse. Here we assume that the ellipse is sufficiently large and that the magnetic field is sufficiently weak so that (i) Zeeman splitting of the Kondo resonance can be neglected, and (ii) the cyclotron radius of the conduction electrons is much larger than the ellipse size.
Under these circumstances, the main effect of the magnetic field is to introduce an additional Aharonon-Bohm phase to the contribution of each path. This phase is the magnetic flux encircled by the orbit, measured in units of the quantum flux, $`\varphi _0=hc/e`$. Here $`h`$ is Planck’s constant, $`c`$ is the velocity of light, and $`e`$ is the electron charge. To compute $`G_1(\stackrel{}{R},\stackrel{}{R})`$ we fix the gauge by calculating the flux encircled by the path which goes from $`\stackrel{}{R}`$ to $`\stackrel{}{R}_j`$, to $`\stackrel{}{R}`$, and then back to $`\stackrel{}{R}`$ along the semimajor axis of the ellipse. Accordingly, the sum of Eq. (14) is modified to $`_j(\xi _{1,j}\xi _{2,j})^{1/2}e^{i2\pi \phi _j/\varphi _0}`$, where $`\phi _j`$ is the flux of each trajectory. Using the continuum approximation of Eq. (15) this sum gives $`2\pi J_0(2𝒜B/\varphi _0)`$, where $`J_0(x)`$ is the Bessel function of zeroth order, and $`𝒜=\pi a^2\sqrt{1^2}`$ is the area of the ellipse. The quantum mirage is, thus, modified according to
$`\delta \rho (B)\delta \rho (0)J_0^2\left(2{\displaystyle \frac{𝒜B}{\varphi _0}}\right),`$ (19)
where $`\delta \rho (0)`$ is the zero-field result of Eq. (18). Note that, for a given $`a`$, the sensitivity to a magnetic field is largest for $`=1/\sqrt{2}`$.
Finally, we discuss the role of imperfections in the ellipse. Clearly, a combined effect of many coherent trajectories is very sensitive to imperfections and dephasing. At 4K, dephasing effects due to electron-electron and electron-phonon interactions are negligible over distances of the order of hundreds of angstroms. In what follows we show that effect of imperfections is small too.
The main source of imperfections in the ellipse comes from the position of the adatoms forming the ellipse. These are constrained to sit on a triangular lattice imposed by the underlying Cu(111) surface. Consequently, the lengths of the orbits contributing to $`\delta \rho `$ are not those of an ideal ellipse. To estimate the effect of the deviations, we consider a random distribution of the trajectory lengths, and average over the distribution. We first notice that each contribution to $`\delta \rho `$ is composed of 4 segments: $`\xi _1`$, $`\xi _2`$, $`\xi _3`$, and $`\xi _4`$, as illustrated in Fig. 1. Each of these segments is regarded as an independent random variable uniformly distributed in the range $`(\overline{\xi }b/2,\overline{\xi }b/2)`$, where $`\overline{\xi }`$ is the exact distance from the focus to the ellipse boundary, and $`b`$ is the triangular lattice spacing. The total length $`\eta =_{n=1,4}\xi _n`$ is, therefore, approximately a Gaussian random variable with mean $`4a`$ and variance $`b^2/3`$. Averaging the cosine term in Eq. (18), $`\mathrm{cos}(k_F\eta )`$, the disorder-averaged value of the quantum mirage is reduced by a factor of approximately $`Q=\mathrm{exp}(k_F^2b^2/6)`$. Substituting the experimental values $`k_F=1/4.75`$Å<sup>-1</sup> and $`b=2.55`$Å, one finds $`Q0.95`$, meaning that the effect of imperfections is negligible.
In conclusion, in this Letter we have studied the phenomenon of quantum mirage, and clarified its relation to the classical orbits of a particle in an ellipse. Our theory also predicts a distinctive behavior of the quantum mirage in the presence of a perpendicular magnetic field, which could be tested experimentally. Finally our approach clearly shows that the phenomenon of quantum mirage is not unique for magnetic adatoms. It will also appear for nonmagnetic atoms (on the ellipse perimeter or at its focus) with strong scattering, e.g. adatoms with resonant tunneling states at the Fermi level.
A.S. is grateful to Hari Manoharan for useful discussions. This research was supported by the Israel science foundation founded by The Israel Academy of Science and Humanities, and by Grant No. 9800065 from the USA-Israel Binational Science Foundation (BSF).
|
warning/0006/hep-lat0006024.html
|
ar5iv
|
text
|
# An illustration of chiral fermions on a 1+1 dimensional lattice
## Abstract
The vectorlike doubling of low-energy excitations is in fact a natural consequence of the pair-production around the zero-energy ($`E=0`$) due to the quantum field fluctuations of the lattice regularized vacuum. On the 1+1 dimensional lattice, we study an anomaly-free chiral model (11112) of four left-movers and one right-mover with strong interactions. Exact computations of relevant $`S`$-matrices illustrate that for high-momentum states, a negative energy-gap ($`E<0`$) develops; the bound state and its constituents, which have the same quantum numbers but opposite chiralities, fill the same energy-state so that chiral symmetries are preserved; for low-momentum states, the negative energy-gap vanishes and the bound state dissolves into its constituents near zero energy. As a consequence of the gauge-anomaly cancellation and the index theorem for flavor-singlet anomalies, the net number of zero-modes pushed down into and pumped out from the zero-energy level by the gauge field is zero.
xue@icra.it
11.15Ha, 11.30.Rd, 11.30.Qc
The parity-violating feature in the low-energy is strongly phenomenologically supported. Based on this feature, the successful standard model for particle physics is constructed in the form of a renormalizable quantum field theory with chiral gauge symmetries on the space-time. While, the very-small-scale structure of the space-time, the arena of physical reality, can exhibit rather complex structure of a space-time “string” or “foam”, instead of a simple space-time point. As the consequence of these fundamental constituents of the space-time, the physical space-time gets endowed with a fundamental length and fundamental theory must be finite. However, with very generic axioms, the “no-go” theorem demonstrates that the quantum field theories with parity-violating gauge symmetries, as the standard model, cannot be consistently regularized on the lattice, i.e., the discrete space-time. This paradox may imply a new physics beyond the standard model. Searching for a chiral gauge symmetric approach to properly regularize the standard model on the lattice has been greatly challenging to particle physicists for the last two decades. One of the classes of approaches is the modeling by appropriately introducing local interactions-. However, the phenomenon of spontaneous symmetry breakings in the intermediate coupling and the argument of anomaly-cancellation within vectorlike spectra in the strong coupling prevent such modelings from a scaling region for low-energy chiral gauged fermions. It was then a general belief that it seems impossible to formulate a theory of exact chiral gauge symmetries on the lattice. Nevertheless, in refs., a model with peculiar interactions and a plausible scaling region were advocated and the dynamics of realizing chiral gauged fermions was studied. In this paper, based on a simple chiral model (11112) on the 1+1 dimensional lattice, we attempt to give an exact illustration of the dynamics realizing chiral gauge theories in the low-energy scaling region.
The chiral model (11112) is made of the $`U(1)`$ gauge field, four left-movers $`\psi _L^i`$ with charge $`Q_L^i=1`$ ($`i=1,2,3,4)`$ and one right-mover $`\psi _R`$ with charge $`Q_R=2`$. The t’Hooft condition for gauge anomaly cancellation $`_i(Q_L^i)^2=Q_R^2`$ is satisfied. With the fixed spatial and temporal lattice spacings $`a`$ and $`a_t`$ $`(aa_t)`$, the free Hamiltonian is given by (we henceforth omit the index $`i=1,2,3,4`$),
$$H_{}=\frac{1}{2a}\underset{x}{}\left(\overline{\psi }_L(x)D^L\gamma \psi _L(x)+\overline{\psi }_R(x)D^R\gamma \psi _R(x)\right),$$
(1)
where all fermionic fields are two-component and dimensionless Weyl fields, $`x`$ is the integer label of space sites, $`\gamma `$-matrix ($`\gamma ^2=1`$) and $`D^{L,R}`$ are ($`\delta _{x,x\pm 1}\psi _L(x)=\psi _L(x\pm 1)`$),
$$D^{L,R}=([U_1(x)]^{Q_{L,R}}\delta _{x,x+1}[U_1^{}(x)]^{Q_{L,R}}\delta _{x,x1}),$$
(2)
where $`U_1(x)`$ is the gauge field at a spatial link and the temporal gauge fixing $`U_{}(x)=1`$. This is a chiral gauge model that cannot be naively quantized on the lattice due to the doubling problem.
The vectorlike doubling phenomenon is in fact the pair-production around zero energy ($`E=0`$) due to quantum field fluctuations of the lattice regularized vacuum (lattice vacuum) of eq.(1). Since the 1+1 dimensional space-time is discretized with the lattice spacings $`a`$ and $`a_t`$, the volume of the lattice vacuum in the energy-momentum space ($`E\stackrel{~}{p}`$ plan), $`0a_tE\pi `$ and $`|a\stackrel{~}{p}|\pi `$ ($`\mathrm{}=1`$), must be finite. The total number of negative energy states must be finite for the reason that each state occupies a quantum volume of $`h`$. Each negative energy state is filled by both particle and antiparticle states so that the lattice vacuum exactly preserve chiral symmetries of (1). As an example, the low-energy excitation of the right-mover ($`\psi _R`$) and its antiparticle can be created from just below zero energy to just above by quantum field fluctuations of the lattice vacuum, since these modes are massless. The energy and momentum fluctuations $`\mathrm{\Delta }E`$ and $`\mathrm{\Delta }\stackrel{~}{p}`$ of the lattice vacuum in such pair production process obey the Heisenberg uncertainty principle
$$a_t|\mathrm{\Delta }E|\pi ,a|\mathrm{\Delta }\stackrel{~}{p}|\pi .$$
(3)
These energy and momentum fluctuations $`|\mathrm{\Delta }E|,|\mathrm{\Delta }\stackrel{~}{p}|`$ of the lattice vacuum are apparently equal to the energy and momentum differences between particle and antiparticle created. Assuming the right-mover ($`\psi _R`$) with the energy $`E=+\stackrel{~}{p}0`$ at the momentum state $`a\stackrel{~}{p}0`$, its antiparticle (the left-mover) must be at the energy-momentum state $`E=\stackrel{~}{p}\pi /a`$ and $`a\stackrel{~}{p}+\pi `$ according to eq.(3). However, the energy state $`E=\pi /a`$ is actually an empty state (hole) of the lattice vacuum. All fully filled negative energy states with $`E(0,\pi /a)`$ must fluctuate down to fill the empty states of lower negative energy levels, as a result the empty state moves to $`E0`$, that is just a low-energy excitation (doubler) of the antiparticle (left-mover) at $`a\stackrel{~}{p}\pi `$ and $`E0`$. Chiral symmetries are preserved for all negative energy states and the zero-energy state filled by pairs of left- and right- movers. All these discussions are applicable to four dimensions. The total number of negative energy states of the vacuum is infinite in the continuum limit $`a_t0,a0`$ and the low-energy excitation (doubler) of the left-mover appear at $`p\mathrm{}`$.
We introduce a neutral and massless spectator $`\chi =\chi _L+\chi _R`$, $`\chi _R`$ couples to four left-movers $`\psi _L^i`$ and $`\chi _L`$ couples to the right-mover $`\psi _R`$ as follow,
$`H_i^L`$ $`=`$ $`g{\displaystyle \underset{x}{}}\overline{\psi }_L(x)\left[\mathrm{\Delta }\chi _R(x)\right]\left[\mathrm{\Delta }\overline{\chi }_R(x)\right]\psi _L(x),`$ (4)
$`H_i^R`$ $`=`$ $`g{\displaystyle \underset{x}{}}\overline{\psi }_R(x)\left[\mathrm{\Delta }\chi _L(x)\right]\left[\mathrm{\Delta }\overline{\chi }_L(x)\right]\psi _R(x),`$ (5)
where the multifermion coupling $`g`$ has dimension $`[a^1]`$ and the operator $`\mathrm{\Delta }`$ is given as,
$`\mathrm{\Delta }\chi _{L,R}(x)`$ $``$ $`\left[\chi _{L,R}(x+1)+\chi _{L,R}(x1)2\chi _{L,R}(x)\right],`$ (6)
$`w(p)`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{x}{}}e^{ipx}\mathrm{\Delta }(x)=\left(\mathrm{cos}(p)1\right),`$ (7)
where the dimensionless momentum $`p=\stackrel{~}{p}a`$. Eq.(7) indicates that large momentum states of $`\chi _R(\chi _L)`$ strongly couple to $`\psi _L(\psi _R)`$, while small momentum states of $`\chi _R(\chi _L)`$ weakly couple to $`\chi _R(\chi _L)`$. For the convenience of computations, we rescale the fermion fields $`\psi (a_tg)^{\frac{1}{4}}\psi `$ and rewrite,
$`a_tH={\displaystyle \frac{1}{2a}}({\displaystyle \frac{a_t}{g}})^{\frac{1}{2}}{\displaystyle \underset{x}{}}\left(\overline{\psi }_L(x)D^L\gamma \psi _L(x)+\overline{\psi }_R(x)D^R\gamma \psi _R(x)+\mathrm{}\right)+a_tH_i^L+a_tH_i^R,`$ (8)
where $`g`$ in $`H_i^{L,R}`$ is rescaled away and “$`\mathrm{}`$” stands for the kinetic terms for $`\chi `$. We consider the limit $`a_t/a0`$, $`ga\mathrm{}`$ and $`ga_t`$ is fixed.
This Hamiltonian system (8) possesses the continuous Abelian chiral gauge symmetry, global chiral symmetries $`U_{L,R}(1)`$ and the shift-symmetries of $`\chi _R`$ and $`\chi _L`$. Due to the Mermin and Wagner theorem, these continuous symmetries cannot be spontaneously broken for any values of the coupling $`ga`$. In addition, the shift-symmetries protect the right-mover $`\psi _R`$ sector and left-movers $`\psi _L`$ sector from coupling each other and guarantee the spectators $`\chi _R,\chi _L`$ decoupled as free particles. These features greatly simply our illustrations.
In this paper, we take the left-moving sector $`\psi _L(x)`$ as an example and analyze only the charged spectrum of Hamiltonian. For the strong coupling $`ga1`$, the three-fermion states $`\mathrm{\Psi }_R`$ with the same quantum numbers of $`\psi _L(x)`$ is formed,
$$\mathrm{\Psi }_R=\frac{1}{2}(\overline{\chi }_R\psi _L)\chi _R,$$
(9)
which is a two-component Wely fermion state. To show this, we compute the S-matrix for this three-fermion state $`\mathrm{\Psi }_R(x,t)`$,
$$S_{33}(x)\underset{t_{f,i}\pm \mathrm{}}{lim}\mathrm{\Psi }_R(0,t_f)|\mathrm{\Psi }_R(x,t_i),$$
(10)
where
$$|\mathrm{\Psi }_R(x,t_i)=e^{iH_i^Lt_i}|\mathrm{\Psi }_R(x,\mathrm{});\mathrm{\Psi }_R(0,t_f)|=\mathrm{\Psi }_R(0,+\mathrm{})|e^{iH_i^Lt_f},$$
(11)
and $`|\mathrm{\Psi }_R(x,\pm \mathrm{})`$ are the asymptotical states. We have then
$$S_{33}(x)=\underset{t_{f,i}\pm \mathrm{}}{lim}\mathrm{\Psi }_R(0,+\mathrm{})|e^{iH_i^Lt_f}e^{iH_i^Lt_i}|\mathrm{\Psi }_R(x,\mathrm{}).$$
(12)
Using ($`\psi `$ indicates $`\psi _L`$ or $`\chi _R`$)
$$𝑑\psi (x,t)𝑑\overline{\psi }(x,t)\overline{\psi }(x,t)\psi (x,t)|\psi (x,t)\psi (x,t)|=1,$$
(13)
we have
$$S_{33}(x)=\underset{t_{f,i}\pm \mathrm{}}{lim}\mathrm{\Psi }_R(0,+\mathrm{})|e^{iH_i^Lt_f}|\psi (0,t_f)\psi (0,t_f)|\psi (x,t_i)\psi (x,t_i)|e^{iH_i^Lt_i}|\mathrm{\Psi }_R(x,\mathrm{}).$$
(14)
We define the form factor of the three-fermion state $`\mathrm{\Psi }_R(x)`$:
$`Z_{t_i}(x)`$ $``$ $`\psi (x,t_i)|e^{iH_i^Lt_i}|\mathrm{\Psi }_R(x,\mathrm{})=\mathrm{\Pi }_{\mathrm{}}^{t_i}{\displaystyle 𝑑\psi (x,t)𝑑\overline{\psi }(x,t)e^{a_tH_i^L}\overline{\psi }_L(x,t_i)\mathrm{\Psi }_R(x,\mathrm{})}`$ (15)
$`Z_{t_f}(0)`$ $``$ $`\mathrm{\Psi }_R(0,+\mathrm{})|e^{iH_i^Lt_f}|\psi (0,t_f)=\mathrm{\Pi }_{t_f}^+\mathrm{}{\displaystyle 𝑑\psi (0,t)𝑑\overline{\psi }(0,t)e^{a_tH_i^L}\overline{\mathrm{\Psi }}_R(0,+\mathrm{})\psi _L(0,t_f)},`$ (16)
where we make the Wick rotation to the Euclidean space. The transfer matrix $`S_{11}(x)`$ of two intermediate states $`|\psi (0,t_f),|\psi (x,t_i)`$ is given by
$$S_{11}(x)\psi (0,t_f)|\psi (x,t_i)=\mathrm{\Pi }_{t_f,x}^{t_i,y}𝑑\psi (x,t)𝑑\overline{\psi }(x,t)e^{a_tH}\overline{\psi }_L(0,t_i)\psi _L(x,t_f),$$
(17)
where $`t_{i,f}`$ are finite and the $`H`$ is the total Hamiltonian. As a result, the S-matrix can be written as,
$$S_{33}(x)=\underset{t_{f,i}\pm \mathrm{}}{lim}Z_{t_i}(x)S_{11}(x)Z_{t_f}(0).$$
(18)
By eq.(16) we compute the form factors $`Z_{t_i}(0)`$, $`Z_{t_f}(x)`$:
$$Z_{t_i}(0)=a_tg\mathrm{\Delta }^2(0)Z_{t_f}(x)=a_tg\mathrm{\Delta }^2(x).$$
(19)
To compute $`S_{11}(x)`$, we define another transfer matrix $`S_{31}(x)`$ of two intermediate states $`|\mathrm{\Psi }_R(x,t_i),|\psi _L(0,t_f)`$,
$$S_{31}(x)\psi _L(0,t_f)|\mathrm{\Psi }_R(x,t_i).$$
(20)
The exact recursion relations for $`S_{11}(x)`$ and $`S_{31}(x)`$ can be obtained (cf. eqs.(C15), (C17) and (C18) in ref.)
$`S_{11}(x)`$ $`=`$ $`{\displaystyle \frac{1}{a_tg\mathrm{\Delta }^2(x)}}\left({\displaystyle \frac{a_t}{2a}}\right)^3{\displaystyle \stackrel{}{}}S_{31}(x)\gamma ,`$ (21)
$`S_{31}(x)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{\delta (x)}{2a_tg\mathrm{\Delta }^2(x)}}+{\displaystyle \frac{1}{a_tg\mathrm{\Delta }^2(x)}}\left({\displaystyle \frac{a_t}{2a}}\right){\displaystyle \stackrel{}{}}S_{11}(x)\gamma \right),`$ (22)
where $`^{}f(x)=f(x+1)f(x1)`$. These recursion equations can be solved by the Fourier transformations for $`p0`$,
$`S_{11}(p)`$ $`=`$ $`{\displaystyle \underset{x}{}}e^{ipx}S_{11}(x)={\displaystyle \frac{\frac{ia_t}{2a}\mathrm{sin}(p)\gamma }{(\frac{a_t}{a})^2\mathrm{sin}^2(p)+M^2(p)}},`$ (24)
$`S_{31}(p)`$ $`=`$ $`{\displaystyle \underset{x}{}}e^{ipx}T_{31}(x)={\displaystyle \frac{\frac{1}{2}M(p)}{(\frac{a_t}{a})^2\mathrm{sin}^2(p)+M^2(p)}},`$ (25)
$`M(p)`$ $`=`$ $`8agw^2(p).`$ (26)
As a result, for $`t_{f,i}\pm \mathrm{}`$ we obtain the exact S-matrix (14)
$$S_{33}(p)=Z(p)\frac{\frac{ia_t}{2a}\mathrm{sin}(p)\gamma }{(\frac{a_t}{a})^2\mathrm{sin}^2(p)+M^2(p)}Z(p),$$
(27)
where the form factor (19) $`Z(p)=4a_tgw^2(p)`$ in the momentum space. The three-fermion state $`\mathrm{\Psi }_R`$ is represented by the pole and its residual(form factor) in the S-matrix (27). We make a wave-function renormalization of three-fermion states $`\mathrm{\Psi }_R`$ with respect to the doubler $`p=\pi `$,
$$\mathrm{\Psi }_R|_{ren}=Z^1(\pi )\mathrm{\Psi }_R;Z(\pi )=16a_tg.$$
(28)
$`\mathrm{\Psi }_R|_{ren}`$ mixes with $`\psi _L`$ to form a four-component massive Dirac fermion $`\mathrm{\Psi }_c`$,
$$\mathrm{\Psi }_c=(\mathrm{\Psi }_R|_{ren},\psi _L),$$
(29)
represented by the pole of the S-matrix ,
$`S_c(x)`$ $`=`$ $`\underset{t_{f,i}\pm \mathrm{}}{lim}\mathrm{\Psi }_c(0,t_f)|\overline{\mathrm{\Psi }}_c(x,t_i)`$ (30)
$`S_c(p)`$ $`=`$ $`{\displaystyle \frac{\frac{ia_t}{a}\mathrm{sin}(p)\gamma +M(p)}{(\frac{a_t}{a})^2\mathrm{sin}^2(p)+M^2(p)}}={\displaystyle \frac{1}{\frac{ia_t}{a}\mathrm{sin}(p)\gamma +M(p)}}`$ (31)
at $`p=\pi `$ and mass $`8ag/a_t`$.
The Hamiltonian $`H_{\mathrm{bound}}`$ in the basis of the eigen-states of $`\psi _L`$ and three-fermion state $`\mathrm{\Psi }_R`$, i.e., Dirac fermion $`\mathrm{\Psi }_c`$, can be obtained from eq.(31):
$$a_tH_{\mathrm{bound}}=\frac{ia_t}{a}\mathrm{sin}(p)\gamma +M(p),H_{\mathrm{bound}}=\frac{i}{a}\mathrm{sin}(p)\gamma \frac{M(p)}{a_t}.$$
(32)
This clearly indicates that the binding energy of the three-fermion state $`\mathrm{\Psi }_R`$ is $`M(p)/a_t`$. Due to the locality of the theory, the vector-like spectrum (32) obtained by the strong coupling for large momentum states, can be analytically continued to small momentum states. However, eq.(31) does not represent a massless pole at $`p=0`$, since the S-matrix $`S_{33}(p)`$ (27) is no longer singular at $`p=0`$ for the form factor $`Z(p)`$ positively vanishing $`O(p^4)`$ and we are not allowed to make wave-function renormalization (28).
$`\mathrm{\Psi }_R(p)`$ consists of three constituents $`\psi _L(p^{})`$, $`\chi _R(q)`$ and $`\overline{\chi }_R(q^{})`$ with the total momentum $`p=p^{}+qq^{}`$. $`\overline{\mathrm{\Psi }}_R(p)`$ consists of three constituents $`\overline{\psi }_L(p^{})`$, $`\chi _R(q)`$ and $`\overline{\chi }_R(q^{})`$ with the total momentum $`p=p^{}+qq^{}`$. The relative momentum $`|\mathrm{\Delta }p|`$ of three constituents within the bound state $`\mathrm{\Psi }_R`$ is of $`O(|p|)`$. According to the Heisenberg uncertainty principle $`|\mathrm{\Delta }p||\mathrm{\Delta }x|2\pi a`$, where $`|\mathrm{\Delta }x|`$ is the size of the bound state $`\mathrm{\Psi }_R`$. If $`|\mathrm{\Delta }p|\pi `$, the size $`|\mathrm{\Delta }x|`$ of the bound state is a few of the lattice spacing $`a`$. As “$`|p|`$” goes to zero, the size $`|\mathrm{\Delta }x|`$ of $`\mathrm{\Psi }_R`$ increase, the negative energy-gap $`M(p)/a_t`$ and form factor $`Z(p)`$ in (27) go to zero, the bound state $`\mathrm{\Psi }_R`$ dissolves into its three constituents.
To illustrate this, we first define the notion of the three-fermion cut $`𝒞[\mathrm{\Psi }_R]`$, which is a virtual state rather than a particle state. For a given total momentum “$`\stackrel{~}{p}`$” in the low-energy region, this virtual state $`𝒞[\mathrm{\Psi }_R](\stackrel{~}{p})`$ has the same total momentum “$`\stackrel{~}{p}`$” and contains the same constituents as $`\mathrm{\Psi }_R(\stackrel{~}{p})`$: $`\psi _L(\stackrel{~}{p}^{})`$, $`\chi _R(\stackrel{~}{q})`$ and $`\overline{\chi }_R(\stackrel{~}{q}^{})`$, where $`\stackrel{~}{p}=\stackrel{~}{q}\stackrel{~}{q}^{}\stackrel{~}{p}^{}`$. $`\psi _L(\stackrel{~}{p}^{})`$, $`\chi _R(\stackrel{~}{q})`$ and $`\overline{\chi }_R(\stackrel{~}{q}^{})`$ are low-energy excitations.
This virtual state $`𝒞[\mathrm{\Psi }_R]`$ has the same quantum numbers as the bound state $`\mathrm{\Psi }_R`$. The total energy of such a virtual state $`𝒞[\mathrm{\Psi }_R]`$ is given by
$`E_t`$ $`=`$ $`E_1(\stackrel{~}{q})+E_2(\stackrel{~}{q}^{})+E_3(\stackrel{~}{p}^{}),`$ (33)
$`E_1(\stackrel{~}{q})`$ $`=`$ $`\stackrel{~}{q}>0,\stackrel{~}{q}>0\mathrm{f}\mathrm{o}\mathrm{r}\chi _R,`$ (34)
$`E_2(\stackrel{~}{q}^{})`$ $`=`$ $`\stackrel{~}{q}^{}>0,\stackrel{~}{q}^{}<0\mathrm{f}\mathrm{o}\mathrm{r}\overline{\chi }_R,`$ (35)
$`E_3(\stackrel{~}{p}^{})`$ $`=`$ $`\stackrel{~}{p}^{}>0,\stackrel{~}{p}^{}<0\mathrm{f}\mathrm{o}\mathrm{r}\psi _L.`$ (36)
There is no any definite one to one relationship between the total energy $`E_t`$ and the total momentum “$`\stackrel{~}{p}`$”. The total energy spectrum $`E_t`$ of such a virtual state $`𝒞[\mathrm{\Psi }_R]`$ is continuum with respect to the given total momentum “$`\stackrel{~}{p}`$”. The lowest energy $`E_t^{\mathrm{min}}(\stackrel{~}{p})`$ (the energy-threshold) of the virtual state $`𝒞[\mathrm{\Psi }_R]`$ is
$$E_t^{\mathrm{min}}(\stackrel{~}{p})=|\stackrel{~}{p}|E_t=|\stackrel{~}{p}^{}|+|\stackrel{~}{q}|+|\stackrel{~}{q}^{}|.$$
(37)
Given the same total momentum “$`p`$” in the low-energy, the bound state $`\mathrm{\Psi }_R(\stackrel{~}{p})`$ is stable, only if only there is an energy gap $`\mathrm{\Delta }(\stackrel{~}{p})`$ between the energy-threshold (37) of $`𝒞[\mathrm{\Psi }_R](\stackrel{~}{p})`$ and the negative binding-energy (32) of $`\mathrm{\Psi }_R(\stackrel{~}{p})`$, i.e.,
$$\mathrm{\Delta }(\stackrel{~}{p})=E_t^{\mathrm{min}}(\stackrel{~}{p})(\frac{M(\stackrel{~}{p})}{a_t})>0.$$
(38)
As the energy gap $`\mathrm{\Delta }(\stackrel{~}{p})`$ vanishes for $`a\stackrel{~}{p}0`$, the three-fermion state $`\mathrm{\Psi }_R(\stackrel{~}{p})`$ must dissolve into its virtual state $`𝒞[\mathrm{\Psi }_R](\stackrel{~}{p})`$. $`\mathrm{\Delta }(\stackrel{~}{p})=0`$ determining the critical momentum threshold $`|\stackrel{~}{p}_c|`$ for such a dissolving phenomenon in the low-energy limit,
$$|\stackrel{~}{p}_c|=\frac{1}{a}\left(\frac{a_t}{2a^2g}\right)^{\frac{1}{3}}\frac{\pi }{2a},ag1,\frac{a_t}{a}1,$$
(39)
and the energy-threshold $`ϵ=\sqrt{2}|\stackrel{~}{p}_c|`$. As an example, for $`ag=100`$ and $`a_t/a=10^1`$, $`ϵ10^1a`$. This dissolving phenomenon is chiral symmetric, since $`\mathrm{\Psi }_R(\stackrel{~}{p})`$ and $`𝒞[\mathrm{\Psi }_R](\stackrel{~}{p})`$ have the same quantum numbers.
Let us discuss how the lattice vacuum is filled by the large momentum states of $`\mathrm{\Psi }_R`$. For simplicity, we introduce the notation $`\psi (p)`$ indicating fermion $`\psi `$ at the momentum state “$`p`$” ($`\pi >p>0`$). As $`\psi _L(p)`$ mixes up with $`\mathrm{\Psi }_R(p)`$ to form the Dirac fermion $`\mathrm{\Psi }_c(p)`$ (29), $`\psi _L(p)`$ and $`\mathrm{\Psi }_R(p)`$ fill into the same positive energy state. $`\overline{\psi }_L(p)`$ and $`\overline{\mathrm{\Psi }}_R(p)`$ fill into the same negative energy state. Thus, each energy state is filled by both left- and right- moving states with the same quantum numbers so that chiral symmetries is preserved by the vectorlike spectrum $`(\psi _L(p),\mathrm{\Psi }_R(p))`$. In addition, a negative energy-gap $`M(p)/a_t`$ is formed. The gauge potential $`A_{}=0`$ and $`A_1`$ give the electric field $`=_{}A_1`$ in the direction of $`p>0`$. The electric force $`\pm Q`$, sign “$`+()`$” for the right(left)-mover, is the rate of changing momentum states $`\mathrm{`}\mathrm{`}p\mathrm{"}`$ by the unit of $`2\pi `$ per unit volume of space-time. The electric force drives fermions spectra of high-momentum states $`\mathrm{`}\mathrm{`}p\mathrm{"}`$ flowing along their dispersion relations. In $`pa|\stackrel{~}{p}_c|`$, the electric field $``$ pushes the four left-movers $`\psi _L^i(p)`$ down to the energy-threshold $`ϵ`$ and pumps the four three-fermion states $`\overline{\mathrm{\Psi }}_R^i(p)`$ up to the energy-threshold $`ϵ`$. In $`pa|\stackrel{~}{p}_c|`$, the electric field $``$ pumps the four three-fermion states $`\mathrm{\Psi }_R^i(p)`$ away from the energy-threshold $`ϵ`$ and pushes the four left-movers $`\overline{\psi }_L^i(p)`$ down below the energy-threshold $`ϵ`$. The rate of pumping out four three-fermion states $`\overline{\mathrm{\Psi }}_R^i(p)`$ in $`p<a|\stackrel{~}{p}_c|`$ and pushing down four left-movers $`\overline{\psi }_L^i(p)`$ for $`p>a|\stackrel{~}{p}_c|`$ are the exactly same, consistently with the finite number of states of the lattice vacuum.
The discussions and computations for the right-mover $`\psi _R`$, the corresponding three-fermion state $`\mathrm{\Psi }_L=\frac{1}{2}(\overline{\chi }_L\psi _R)\chi _L`$ and cut $`𝒞[\mathrm{\Psi }_L]`$ are the exactly same as that for the left-moving sector $`\psi _L`$.
Now we focus the spectral flows of low-energy excitations within $`ϵ<\stackrel{~}{p}<ϵ`$. The neutral spectator fermion $`(\chi _L,\chi _R)`$ is a free and massless Dirac particle ($`E=\pm \stackrel{~}{p}`$). Since the vacuum energy states including the zero-energy state ($`E[0,ϵ)`$) are only filled by the left(right)-movers $`\psi _L^i(\stackrel{~}{p})(\psi _R(\stackrel{~}{p}))`$ without their partners of the same charge and opposite chirality, the $`U(1)`$ chiral gauge symmetry is broken and gauge anomalies must appear. In the gauge fixing $`A_{}=0`$, the gauge anomalies are given by $`(Q_L^i)^2/(4\pi )`$ for the gauge current $`J_L^i=Q_L^i\overline{\psi }_L^i\gamma \psi _L^i`$ of four left-movers $`\psi _L^i`$ and $`+(Q_R)^2/(4\pi )`$ for the gauge current $`J_R=Q_R\overline{\psi }_R\gamma \psi _R`$ of right-mover $`\psi _R`$, which are proportional to the rate of pushing the charges $`Q_L^i`$ of four $`\psi _L^i`$ down into and pumping the charge $`Q_R`$ of one $`\psi _R`$ out from the zero-energy level $`E=0`$. Due to the fact that the t’Hooft condition is obeyed and gauge anomalies are exactly canceled, the corresponding net charges pushed down into and pumped out from the lattice vacuum by the electric field $``$ are zero. However, the corresponding net number of zero modes pushed down into and pumped out from the lattice vacuum is not zero, i.e., 4-1=3. It seems to be inconsistent for the finiteness of the lattice vacuum for no extra rooms accommodating 3 zero modes.
On the other hand, the vacuum energy states ($`E[0,ϵ)`$) that are filled by the left(right)-movers $`\psi _L^i(\stackrel{~}{p})(\psi _R(\stackrel{~}{p}))`$ break the global chiral symmetries $`U_{L,R}(1)`$. The divergence of Noether currents of these global symmetries receives anomalies. Such flavor-singlet anomalies are given by $`\pm /(4\pi )`$, “+” for $`j_L=_i\overline{\psi }_L^i\gamma \psi _L^i`$ and “-” for $`j_R=\overline{\psi }_R\gamma \psi _R`$. $`\pm /(4\pi )`$ are proportional to the rate of pushing the number of state of left-movers down into and pumping the number of state of right-mover out from zero energy $`E=0`$. By the index theorem, the axial anomaly of the current $`j^5=j_Rj_L`$ is given by $`\mathrm{\Delta }n=n_{}n_+`$, where $`n_{}(n_+)`$ is the number of right(left)-movers. $`\mathrm{\Delta }n`$ are the fermion numbers carried by topological gauge fields. $`\mathrm{\Delta }n=14=3`$ indicating three zero modes flowing out from the lattice vacuum and three states are emptied for accommodating 3 zero modes pushed in, consistently with the finiteness of the lattice vacuum.
In fact, for the reason that the dissolving energy-scale $`ϵ\pi /a\pi /a_t`$ and the asymmetry $`aa_t`$ in the space-time turns out to be irrelevant at the low-energy scale $`ϵ`$, we can define a low-energy effective Lagrangian for the 11112 model with a continuous regularization at the energy scale $`ϵ`$. The asymmetry of filling the vacuum energy states ($`E[0,ϵ)`$), as discussed in above, must appear as explicit symmetry breaking terms in the low-energy effective Lagrangian at the scale $`ϵ`$. As results we have, (i) the gauge anomalies $`Q_L^iϵ^{\mu \nu }_\mu A_\nu /(4\pi )`$ and $`Q_Rϵ^{\mu \nu }_\mu A_\nu /(4\pi )`$ of the $`U(1)`$ gauge symmetry; (ii) the flavor-singlet anomalies $`\pm ϵ^{\mu \nu }F_{\mu \nu }/(4\pi )`$ of the $`U_{L,R}(1)`$ global chiral symmetries and the axial anomaly is given by $`\mathrm{\Delta }n=n_{}n_+`$; (iii) local counterterms at the scale $`ϵ`$.
|
warning/0006/cond-mat0006218.html
|
ar5iv
|
text
|
# Spin-triplet ‘f-wave’ pairing proposed for an organic superconductor (TMTSF)2PF6
## Abstract
By examining how the spin- and/or charge-fluctuation exchange can contribute to pairing instabilities, we propose that a spin-triplet f-wave-like pairing with a $`d`$-vector perpendicular to the $`b`$-axis may be realized in (TMTSF)<sub>2</sub>PF<sub>6</sub> due to (i) a quasi-one-dimensional Fermi surface, (ii) a coexistence of $`2k_F`$ charge fluctuations and spin fluctuations, and (iii) an anisotropy in spin fluctuations. Fluctuation-exchange study for the Hubbard model confirms the point (i), while a phenomelogical analysis is given for (ii) and (iii). The proposed pairing is consistent with various experiments.
Spin-triplet pairing is conceptually fascinating, but there seem to be few examples. Recently an organic superconductor, (TMTSF)<sub>2</sub>PF<sub>6</sub>, has attracted much attention since a triplet pairing is suggested from an observation of large $`H_{c2}`$ as well as from a Knight shift experiment, while the absence of Hebel-Slichter peak and the power-law decay of $`T_1^1`$ below $`T_c`$ suggest an anisotropic pairing with nodes in the gap.
If triplet pairing is indeed realized in (TMTSF)<sub>2</sub>PF<sub>6</sub>, its mechanism is a challenging theoretical puzzle: in the pressure-temperature phase diagram for this material the superconductivity lies right next to the $`2k_F`$ spin density wave (SDW), so that if one seeks an electronic origin, a spin-singlet d-wave-like pairing mediated by spin fluctuations is most naturally expected as proposed by several authors. If on the other hand one assumes a phonon-mediated attractive interaction, triplet pairing, with nodes in the gap in general, would seem less favorable compared to singlet s-wave pairing without nodes. Recently, Kohmoto and Sato have proposed that this difficulty in the phonon mechanism may be circumvented for TMTSF compounds, where they show that the quasi-one dimensionality of the Fermi surface, along with the presence of spin fluctuations, makes a triplet p-wave pairing without nodes on the Fermi surface dominate over s-wave pairing. However, it is not clear if such a nodeless gap can be reconciled with the absence of Hebel-Slichter peak and the power-law $`T_1^1`$ in (TMTSF)<sub>2</sub>PF<sub>6</sub>.
If we turn to another prominent candidate for triplet superconductivity accompanied by SDW fluctuations, Sr<sub>2</sub>RuO<sub>4</sub>, Takimoto recently proposed that charge fluctuations (or more precisely orbital fluctuations) should arise from repulsions between degenerate 4d orbitals, and that the coexistence of spin and charge fluctuations may lead to a triplet pairing. This makes us recall an experimental fact that a $`2k_F`$ charge density wave (CDW) actually coexists with the SDW in (TMTSF)<sub>2</sub>PF<sub>6</sub> as suggested from X-ray diffuse scattering. In another theory for Sr<sub>2</sub>RuO<sub>4</sub>, Sato and Kohmoto, and independently Kuwabara and Ogata, have proposed that anisotropy of the spin fluctuations, known to be present experimentally, may give rise to a triplet p-wave pairing. The anisotropy of spin fluctuations is also present in (TMTSF)<sub>2</sub>PF<sub>6</sub>. Moreover, Sr<sub>2</sub>RuO<sub>4</sub> has two quasi-1D Fermi surfaces (although they are weakly hybridized to result in two 2D Fermi surfaces), so the ruthenate seems to share several features with the TMTSF compound, although the strong charge fluctuation employed in Takimoto’s mechanism is yet to be detected experimentally in Sr<sub>2</sub>RuO<sub>4</sub>.
However, the triplet pairing mechanism of (TMTSF)<sub>2</sub>PF<sub>6</sub> cannot be the same with that of Sr<sub>2</sub>RuO<sub>4</sub> since $`\stackrel{}{d}`$ (the d-vector characterizing the triplet pairing) $`\stackrel{}{z}`$ (easy axis of the spins) is experimentally suggested in the former, while $`\stackrel{}{d}\stackrel{}{z}`$ in the latter. In the present paper, we propose that a triplet f-wave-like pairing with $`\stackrel{}{d}\stackrel{}{z}`$ can take place in (TMTSF)<sub>2</sub>PF<sub>6</sub> due to a combination of (i) the quasi-one-dimensionality of the Fermi surface, (ii) coexistence of $`2k_F`$ spin and charge fluctuations, and (iii) the anisotropy in the spin fluctuations. In the first part of the paper, we focus on how the quasi-one-dimensionality works favorably for the triplet pairing, and perform a fluctuation-exchange (FLEX) calculation for the on-site repulsion Hubbard model on a lattice for (TMTSF)<sub>2</sub>PF<sub>6</sub>. Then, in the second part, we discuss phenomelogically how the triplet pairing can become competitive against the singlet when charge fluctuations coexist with spin fluctuations. We finally point out that the anisotropy in the spin fluctuations should further favor triplet pairing with $`\stackrel{}{d}\stackrel{}{z}`$.
We first consider the on-site Hubbard model, $`=_{i,j\sigma }t_{ij}c_{i\sigma }^{}c_{j\sigma }+U_in_in_i`$, in the hole picture on a quasi-1D lattice ($`|t_S|>|t_I|`$) depicted in Fig.1. There are $`n=0.5`$ holes per site. Since sites A and B are inequivalent for $`t_{S1}t_{S2}`$ and $`t_{I1}t_{I2}`$ (dimerization of the molecules), we adopt the two-band version of the FLEX (although we shall see that the dimerization is not essential in our argument).
For later discussions, we first recapitulate the one-band version of FLEX in a general fashion, where we proceed: (i) Dyson’s equation is solved to obtain the renormalized
Green’s function $`G(k)`$, where $`k`$ is a shorthand for the wave vector $`𝐤`$ and the Matsubara frequency, $`iϵ_n`$, (ii) the fluctuation-exchange interaction $`V^{(1)}(q)`$, given as,
$$V^{(1)}(q)=\frac{1}{2}V_{\mathrm{sp}}^{zz}(q)+V_{\mathrm{sp}}^+(q)+\frac{1}{2}V_{\mathrm{ch}}(q),$$
(1)
consists of the contribution from longitudinal $`(zz)`$ and transverse $`(+)`$ spin fluctuations (sp) and that from charge fluctuations (ch). For the on-site Hubbard model in particular, $`V_{\mathrm{sp}}^{zz}=V_{\mathrm{sp}}^+(V_{\mathrm{sp}})=U^2\chi ^{\mathrm{sp}}`$ and $`V_{\mathrm{ch}}=U^2\chi ^{\mathrm{ch}}`$, where the spin and the charge susceptibilities are given as $`\chi ^{\mathrm{sp}}(q)=\chi ^{\mathrm{irr}}(q)/[1U\chi ^{\mathrm{irr}}(q)]`$ and $`\chi ^{\mathrm{ch}}(q)=\chi ^{\mathrm{irr}}(q)/[1+U\chi ^{\mathrm{irr}}(q)]`$, respectively, using the irreducible susceptibility $`\chi ^{\mathrm{irr}}(q)=\frac{1}{N}_kG(k+q)G(k)`$ ($`N`$:number of $`k`$-point meshes). (iii) $`V^{(1)}`$ then brings about the self-energy, $`\mathrm{\Sigma }(k)=\frac{1}{N}_qG(kq)V^{(1)}(q)`$, which is fed back to Dyson’s equation, and the self-consistent loop is repeated until convergence is attained.
$`T_c`$ is the temperature where the eigenvalue $`\lambda `$ of the following Éliashberg equation for the superconducting order parameter $`\varphi (k)`$ reaches unity.
$`\lambda _\mu \varphi _\mu (k)`$ $`=`$ $`{\displaystyle \frac{T}{N}}{\displaystyle \underset{k^{}}{}}\varphi _\mu (k^{})|G(k^{})|^2V_\mu ^{(2)}(kk^{})`$ (2)
Here, the pairing interaction $`V_\mu ^{(2)}(q)`$ is given as
$`V_s^{(2)}(q)={\displaystyle \frac{1}{2}}V_{\mathrm{sp}}^{zz}(q)+V_{\mathrm{sp}}^+(q){\displaystyle \frac{1}{2}}V_{\mathrm{ch}}(q)`$ (3)
for singlet pairing,
$`V_t^{(2)}(q)={\displaystyle \frac{1}{2}}V_{\mathrm{sp}}^{zz}(q){\displaystyle \frac{1}{2}}V_{\mathrm{ch}}(q)`$ (4)
for triplet pairing with total $`S_z=\pm 1`$ ($`\stackrel{}{d}\stackrel{}{z}`$), and
$`V_t^{(2)}(q)={\displaystyle \frac{1}{2}}V_{\mathrm{sp}}^{zz}(q)V_{\mathrm{sp}}^+(q){\displaystyle \frac{1}{2}}V_{\mathrm{ch}}(q)`$ (5)
for triplet pairing with $`S_z=0`$ ($`\stackrel{}{d}\stackrel{}{z}`$). In the on-site Hubbard model, $`V_{\mathrm{sp}}V_{\mathrm{ch}}`$ is satisfied, so that $`|V_t^{(2)}|(1/3)|V_s^{(2)}|`$ holds with $`V_t^{(2)}=V_t^{(2)}V_t^{(2)}`$.
In the two-band version of FLEX, $`G`$, $`\chi `$, $`\mathrm{\Sigma }`$, and $`\varphi `$ all become $`2\times 2`$ matrices, whose elements are denoted as $`G_{\alpha \beta }`$ etc with $`\alpha ,\beta =`$A or B in the site representation, which may be converted to the band representation with a unitary transform. Since the Fermi surface lies in the lower band for quarter filling, we concentrate on Green’s function and the order parameter in that band, denoted as $`G`$ and $`\varphi _s,\varphi _t`$, respectively. As for the spin susceptibility, we diagonalize the $`2\times 2`$ matrix $`\chi ^{\mathrm{sp}}`$ and concentrate on the larger eigenvalue, denoted as $`\chi `$. To ensure convergence at low temperatures in the two-band system we had to take $`64\times 64`$ $`k`$-points and $`ϵ_n`$ from $`(2N_c1)\pi T`$ to $`(2N_c1)\pi T`$ with $`N_c`$ up to 8192.
In Fig.2, we show contour plots of $`|G(𝐤,i\pi k_BT)|^2`$(a), $`\chi (𝐤,0)`$ (b), $`\varphi _s(𝐤,i\pi k_BT)`$ (c), and $`\varphi _t(𝐤,i\pi k_BT)`$ (d) for $`T=0.015`$. The Fermi surface as identified from the ridge in $`|G(𝐤)|^2`$ is a pair of warped quasi 1D pieces. The spin susceptibility $`\chi (𝐪,0)`$ has a peak at $`𝐐(\pi ,\pi /2)`$ (or $`(\pi /2,\pi /2)`$ in the unfolded Brillouin zone in the absence of dimerization), as expected from the nesting vector and in agreement with experimental results. The singlet pairing order parameter is seen to change sign in such a way that (i) $`\varphi _s(𝐤)=\varphi _s(𝐤)`$, and (ii) $`\varphi _s(𝐤)`$ has opposite signs across the nesting vector Q so that the repulsive $`V_s^{(2)}(𝐐)`$ (eq.(3)) acts as an attractive interaction in the gap equation. We call the singlet pairing a ‘d-wave’ in that the sign of $`\varphi _s(𝐤)`$ changes like $`++`$ if we rotationally scan the Fermi surface, which is consistent with previous studies.
For the triplet pairing, by contrast, $`V_t^{(2)}(𝐐)`$ is attractive (eq.(4) or (5) with $`V_{\mathrm{sp}}^+=V_{\mathrm{sp}}^{zz}`$), so that the order parameter should have the same sign across $`𝐐`$. This requirement, along with the condititon for a triplet order parameter $`\varphi _t(𝐤)=\varphi _t(𝐤)`$, can be satisfied by adding extra nodal lines along $`k_a0`$ and $`k_a\pi `$ (mod $`2\pi `$). We call this pairing an ‘f-wave’ in that $`\varphi _s`$ behaves this time like $`+++`$ along the Fermi surface.
A virtue of the quasi-one dimensionality is that the magnitudes, $`|\varphi _s(𝐤)|`$ and $`|\varphi _t(𝐤)|`$, are almost identical around the Fermi surface as seen from Figs.2(a) and (c),(d). In fact, the ‘f-wave’ here is to ‘d-wave’ what p-wave is to s-wave in Kohmoto-Sato’s picture in that a singlet is converted into a triplet by introducing extra nodes that do not affect $`|\varphi (𝐤)|`$ on the Fermi surface. In such a situation, the difference between $`\lambda _s`$ and $`\lambda _t`$ comes almost entirely from the difference between $`|V_s^{(2)}(𝐐)|`$ and $`|V_t^{(2)}(𝐐)|`$ in the Éliashberg equation (2).
Our result shows that the ratio $`\lambda _t/\lambda _s`$ indeed tends to $`1/3`$ at low temperatures (Fig.3; solid line), which reflects the ratio $`|V_t^{(2)}(𝐐)/V_s^{(2)}(𝐐)|1/3`$ in the on-site Hubbard model. The ratio approaches to $`1/3`$ as the temperature is lowered, because the ridge in $`|G|^2`$ sharpens so that the triplet pairing, with the order parameter vanishing around $`k_a=0,\pi `$, becomes more favorable. We can also confirm that quasi-1D is exploited in realizing $`\lambda _t/\lambda _s1/3`$ by pushing the system toward 2D with larger value of $`t_{I2}`$ and $`|t_{I3}|`$, for which the ratio $`\lambda _s/\lambda _t`$ deviates from 1/3 even at low temperatures (Fig.3; dashed line).
We have so far seen that the difference in $`\lambda `$ between singlet ‘d’ and triplet ‘f’ can directly reflect the difference between $`|V_s^{(2)}(𝐐)|`$ and $`|V_t^{(2)}(𝐐)|`$ in a quasi-1D system. The ‘f-wave’ proposed here is an appealing candidate, because it can account for both experimentally suggested triplet pairing and the nodes in the superconducting gap. However, even for a quasi-1D system, ‘f’ is only 1/3 competitive against ‘d’ as far as the on-site repulsion Hubbard model is concerned — to make ‘f’ dominate over ‘d’, we need to have $`|V_t^{(2)}(𝐐)|>|V_s^{(2)}(𝐐)|`$. So at this stage we depart from the Hubbard model to argue phenomelogically how some factors in the actual (TMTSF)<sub>2</sub>PF<sub>6</sub> that are not taken into account in the simple Hubbard model can indeed make ‘f’ competitive against ‘d’.
An important experimental fact for (TMTSF)<sub>2</sub>PF<sub>6</sub> that cannot be explained by the on-site Hubbard model is that a $`2k_F`$ CDW actually coexists with the $`2k_F`$ SDW. The coexistence of spin and charge fluctuations can favor triplet pairing as pointed out for Sr<sub>2</sub>RuO<sub>4</sub> by Takimoto mentioned above. This can be seen in eqs.(3),(4), where an increase in $`V_{\mathrm{ch}}`$ enhances $`|V_t^{(2)}|`$ and suppresses $`|V_s^{(2)}|`$ (as far as $`V_{\mathrm{sp}}>3V_{\mathrm{ch}}`$ for isotropic spin fluctuations assumed for the time being). Now, if we take the coexistence of $`2k_F`$ SDW and $`2k_F`$ CDW in (TMTSF)<sub>2</sub>PF<sub>6</sub> to be $`V_{\mathrm{sp}}(𝐐)V_{\mathrm{ch}}(𝐐)`$, eqs.(3) and (4) dictate that $`|V_t^{(2)}(𝐐)||V_s^{(2)}(𝐐)|`$, so ‘f’ does indeed compete with ‘d’, but the competition is still subtle.
Is there any mechanism that further favors the triplet pairing? Magnetic anisotropy is, in our view, one. It has actually been revealed experimentally for (TMTSF)<sub>2</sub>PF<sub>6</sub> that the SDW has an easy axis in the $`b`$-direction, which implies that $`V_{\mathrm{sp}}^{zz}(𝐐)>V_{\mathrm{sp}}^+(𝐐)`$ is satisfied for $`z`$ taken to be $`b`$. In such a situation, the ‘f’ pairing is more favorable in the $`S_z=\pm 1`$ channel since $`|V_t^{(2)}|>|V_t^{(2)}|`$. The condition for ‘f’ dominating over ‘d’ now reads $`|V_t^{(2)}(𝐐)|>V_s^{(2)}(𝐐)`$, or
$$V_{\mathrm{ch}}(𝐐)>V_{\mathrm{sp}}^+(𝐐)$$
(6)
from eqs.(3),(4). This last condition should be satisfied in (TMTSF)<sub>2</sub>PF<sub>6</sub> because the spins do not order in the transverse direction even in the SDW phase, while the charges do. We stress that $`\stackrel{}{d}\stackrel{}{z}`$ with $`\stackrel{}{z}\stackrel{}{b}`$ is consistent with the experimental result: it is when the magnetic field is applied parallel to the $`b`$-axis that (i)$`H_{c2}`$ becomes largest at low temperatures, and (ii) the Knight shift is unchanged across $`T_c`$.
The mechanism in which the anisotropy of the spin fluctuations favors triplet pairing is reminiscent of the one proposed in refs. for Sr<sub>2</sub>RuO<sub>4</sub>, but a crucial difference is that refs., which do not consider charge fluctuations, conclude a p-wave pairing with $`S_z=0`$ in agreement with the experimental results suggesting $`\stackrel{}{d}\stackrel{}{z}`$ in Sr<sub>2</sub>RuO<sub>4</sub>. Let us see how this would occur in the present context. If $`V_{\mathrm{sp}}^{zz}>(2V_{\mathrm{sp}}^++V_{\mathrm{ch}})`$, we can see from eq.(5) that $`V_t^{(2)}(𝐐)`$ becomes repulsive, which will mediate a triplet pairing having an order parameter with opposite signs across $`𝐐`$. This requirement, along with the triplet condition $`\varphi _t(𝐤)=\varphi _t(𝐤)`$, can be satisfied by putting nodes only at $`k_a0`$ and $`k_a\pi `$, thereby making the order parameter nodeless on the Fermi surface as in refs.. Specifically, if $`V_{\mathrm{ch}}/V_{\mathrm{sp}}^{zz}`$ and $`V_{\mathrm{sp}}^+/V_{\mathrm{sp}}^{zz}`$ are both sufficiently small, the pairing interactions $`V_s^{(2)}`$ (favoring ‘d’), $`V_t^{(2)}`$ (‘f’), and $`V_t^{(2)}`$ (p) will all have similar magnitudes, so the p-wave pairing, with no nodes on the Fermi surface, should dominate over the others. Thus, the ‘f’ pairing is not realized unless $`V_{\mathrm{ch}}/V_{\mathrm{sp}}^{zz}`$ is significant even if eq.(6) is satisfied (see Fig.4).
In this context, a possibly related problem is the pairing symmetry in (TMTSF)<sub>2</sub>ClO<sub>4</sub>, another candidate for a triplet superconductor. For this compound an NMR experiment suggests a presence of nodes on the Fermi surface, while a recent thermal conductivity
measurement suggests a nodeless gap. If we adopt the latter result, the nodeless p-wave pairing should become a strong candidate. Then a comparison of the magnitude of the charge fluctuation as well as the direction of $`\stackrel{}{d}`$ between (TMTSF)<sub>2</sub>ClO<sub>4</sub> and (TMTSF)<sub>2</sub>PF<sub>6</sub> will be a crucial test.
Having discussed the lower bound of $`V_{\mathrm{ch}}/V_{\mathrm{sp}}^{zz}`$ for ‘f’ pairing, how about an upper bound? When $`V_{\mathrm{ch}}/V_{\mathrm{sp}}^{zz}1`$, singlet s-wave pairing with $`\varphi _s(𝐤)\mathrm{constant}`$ should enter as the dominant pairing. This is because $`V_s^{(2)}`$ becomes attractive for $`V_{\mathrm{ch}}>(V_{\mathrm{sp}}^{zz}+2V_{\mathrm{sp}}^+)`$, so that $`\varphi _s(𝐤)`$ no longer has to change sign. As $`V_s^{(2)}/V_t^{(2)}`$ tends to unity with the increase of charge fluctuations, the ‘f’, with its nodes on the Fermi surface, should thus give way to the nodeless s. All the above reasoning is schematically summarized as a generic phase diagram in Fig.4.
An additional bonus from the coexistence of strong spin/charge fluctuations and the anisotropic spin fluctuations is that these effects may serve to enhance the transition temperature for triplet pairing. Namely, a flaw in triplet superconductivity mediated by isotropic spin fluctuations is that the absolute value of the triplet pairing interaction $`|V_t^{(2)}|`$ is only one third of the effective interaction $`V^{(1)}`$ that determines the normal self-energy as seen from eqs.(1) and (4) (or (5)) with $`V_{\mathrm{sp}}^{zz}=V_{\mathrm{sp}}^+V_{\mathrm{ch}}`$. This is in contrast with the case of singlet pairing, where $`V_s^{(2)}`$ is nearly identical to $`V^{(1)}`$. Since a large self-energy correction results in a short quasi-particle lifetime, $`V^{(1)}`$ suppresses $`T_c`$ while $`V^{(2)}`$ enhances it, and $`V^{(1)}3|V_t^{(2)}|`$ in the Hubbard model generally results in a $`T_c`$, if any, too low to be detected in FLEX calculations. This difficulty is resolved for large $`V_{\mathrm{ch}}`$ and/or small $`V_{\mathrm{sp}}^+`$, for which $`|V_t^{(2)}|`$ approaches $`V^{(1)}`$.
The microscopic origin of the charge fluctuation remains to be identified. Since there is no orbital degeneracy in (TMTSF)<sub>2</sub>X, the origin cannot be the one proposed by Takimoto for Sr<sub>2</sub>RuO<sub>4</sub>. In fact, the mechanism of $`2k_F`$ SDW-CDW coexistence in (TMTSF)<sub>2</sub>PF<sub>6</sub> has been investigated by several authors. Some assume electron-lattice coupling, while others envisage a purely electronic origin in terms of off-site repulsions up to second nearest neighbors. It would be an interesting future problem to investigate microscopically the singlet-triplet competition and to evaluate $`T_c`$ by taking these effects into account.
We thank Tetsuya Takimoto for illuminating discussions and sending ref. prior to publication. We are also indebted to Mahito Kohmoto, Masao Ogata, and Masatoshi Sato for valuable discussions. One of us (H.A.) also thanks John Singleton for discussions on the SDW/CDW problem. Numerical calculations were performed at the Computer Center, University of Tokyo, and at the Supercomputer Center, ISSP, University of Tokyo.
|
warning/0006/hep-ph0006051.html
|
ar5iv
|
text
|
# 1 Introduction
## 1 Introduction
Besides its conceptual elegance, the prime motivation to consider low energy supersymmetry (SUSY) is the stabilization of the electroweak scale under radiative corrections. However, the Minimal Supersymmetric Standard Model (MSSM) provides no explanation to the $`\mu `$-problem , i.e., to why the scales of SUSY breaking and of the supersymmetric bilinear Higgs $`\mu `$-term,
$$\mu h_dh_u,$$
(1)
are of comparable magnitude. The MSSM should therefore be extended by at least some sector solving the $`\mu `$-problem. In this sense, the term MSSM is a misnomer.
Moreover, in sharp contrast to the non-supersymmetric Standard Model (SM) where the stability of the proton can be understood entirely by gauge invariance, the MSSM predicts rapid proton decay via dimension 4 operators, unless an additional discrete symmetry, such as $`R`$-parity, is imposed ad hoc. While this is a logical possibility, it severely compromises the initial elegance and one prefers to save the successful features of the SM when exploring its alternatives. It has also been argued that those discrete symmetries<sup>1</sup><sup>1</sup>1Continuous global symmetries are not expected to arise from string theory . which survive as remnants of some underlying string model, might either be unsuited to forbid (at least) the dangerous dimension 4 operators, or arise only along some fine-tuned flat scalar field direction. Indeed, baryon number conservation by the principles of gauge invariance (and renormalizability) has been on the wish list of SUSY practitioners for almost two decades .
It is well known that either of the problems addressed in the previous two paragraphs can be solved by the introduction of an extra $`U(1)^{}`$ symmetry. Examples for either case are reviewed in the next section using $`U(1)`$ groups motivated by $`E_6`$ Grand Unified Theories (GUTs). Additional $`U(1)`$s have long been considered as very well motivated extensions of the MSSM ; they are predicted in most GUTs and appear copiously in superstring theories. In this article, I address the question whether it is possible to find models in which an extra $`U(1)^{}`$ solves both problems simultaneously.
If such an extension is free of gauge and gravitational anomalies, one can allow a grand desert scenario between the electroweak scale and a more fundamental high mass scale. This is desirable since our confidence in the MSSM is boosted by the observation of approximate gauge coupling unification (as predicted by the simplest and most economic GUT and string models) at a scale somewhat below the Planck scale. An extra $`U(1)^{}`$ does not affect the renormalization group equations at the one-loop level, and it is conceivable that the quality of unification even improves at two-loop precision. In order to rigorously protect the electroweak scale from higher mass scales, I also demand that the field content of the models be completely chiral before gauge and supersymmetry breaking. I.e., all superfields should transform non-trivially with respect to the low-energy $`SU(3)\times SU(2)\times U(1)_Y\times U(1)^{}`$ gauge symmetry, and all bilinear (mass) terms must be forbidden by gauge invariance. Note, that these requirements present a strong set of constraints, and it is not clear a priori, whether any solutions exist. The key ingredient necessary to arrive at a solution is that at least two MSSM singlet fields which are charged under the $`U(1)^{}`$ have to acquire vacuum expectation values (VEVs).
Electroweak precision data imply further hints and constraints. The predictions of the SM are generally in very good agreement with experiments provided the Higgs boson mass, $`M_H`$, is smaller than about 250 GeV . This is also predicted by the MSSM and many extensions ($`M_H\stackrel{<}{_{}}150`$ GeV) . Moreover, the superpartners and extra Higgs fields decouple from the precision observables over large parts of the MSSM parameter space, so that the MSSM is likewise in agreement with observations . Addition of an extra $`Z^{}`$ boson can even improve the global fit to all data , driven mostly by the parity violation experiments in Cs , but the results in Tl , and from the $`Z`$ lineshape measurements at LEP 1 also play a role. Note, that these fits suggest a $`Z^{}`$ boson mass $`M_Z^{}\stackrel{<}{_{}}𝒪(\text{1 TeV})`$ and continue to yield results on $`M_H`$ consistent with SUSY .
Fits to the oblique parameters, $`S`$, $`T`$, and $`U`$ (or $`ϵ_1`$, $`ϵ_2`$, and $`ϵ_3`$ ), describing new physics contributions to vector boson self-energies, yield results consistent with zero (for Higgs masses in the MSSM range). This implies restrictions on extra matter fields which can contribute significantly to $`S`$ and $`T`$, but if they are non-chiral (with respect to the SM gauge group) and approximately mass degenerate their effects are rendered small. If they arise in complete representations of $`SU(5)`$, they do not affect gauge coupling unification at the one-loop level, nor its scale. However, in the bottom-up approach of this paper I do not assume an $`SU(5)`$-GUT symmetry. I will show that other matter configurations can preserve unification, as well.
To reiterate, I will assume that there is an additional $`U(1)^{}`$ symmetry with a mechanism built in to solve the SUSY $`\mu `$-problem. I will classify the solutions to the conditions of anomaly cancellation, gauge coupling unification, and chirality. To avoid potential problems with fractionally charged states, I will also assume $`SU(5)`$-type charge quantization. About half of the resulting models predict a sufficiently long proton lifetime.
In the next section, I review the situation for the most popular $`U(1)^{}`$ extensions arising from $`SO(10)`$ and $`E_6`$ GUTs . In Section 3, I introduce the anomaly conditions and show that it is not possible to find chiral models if only one MSSM singlet field develops a VEV. Section 4 shows that there are 33 solutions (family universal with respect to the ordinary fermions) if there are two singlets acquiring VEVs in the course of $`U(1)^{}`$ breaking. Five of these models involve a number of $`\mathrm{𝟓}+\overline{\mathrm{𝟓}}`$ representations and singlets of $`SU(5)`$, but no extra matter beyond that. Allowing incomplete representations of $`SU(5)`$ and the weaker requirements of gauge unification and absence of fractionally charged states yield the remaining 28 solutions. I will also briefly address the prospects to derive these models from the $`E_8\times E_8`$ heterotic string theory. Some of the phenomenological implications of the type of models obtained in Section 4 are discussed in Section 5. My conclusions are presented in Section 6.
## 2 $`E_6`$ inspired $`U(1)^{}`$ models
Consider, as an illustrative example, the case of an $`E_6`$ gauge group which is free of anomalies in all its representations. A fundamental 27 representation contains a pair of $`SU(2)`$ doublets, $`h_d`$ and $`h_u`$, carrying the appropriate $`U(1)_Y`$ hypercharges to serve as the MSSM Higgs doublets, i.e., allowing the Yukawa couplings,
$$qh_d\overline{d},qh_u\overline{u},lh_de^+,lh_u\overline{\nu }.$$
(2)
Here $`l`$ and $`q`$ are the lepton and quark doublet superfields<sup>2</sup><sup>2</sup>2I denote the MSSM multiplets by lower case letters, while additional fields will be capitalized., $`e^+`$, $`\overline{d}`$, and $`\overline{u}`$ denote the $`SU(2)`$ singlets, and $`\overline{\nu }`$ refers to the right-handed neutrino superfield. If one identifies the $`U(1)^{}`$ with the $`U(1)_\psi `$ defined by $`E_6SO(10)\times U(1)_\psi `$, then $`h_d`$ and $`h_u`$ have equal $`U(1)^{}`$ charges, excluding an elementary $`\mu `$-term. A 27 also includes an MSSM singlet superfield, $`S`$, which has the right $`U(1)^{}`$ charge to allow a trilinear Yukawa term,
$$Sh_dh_u,$$
(3)
in the superpotential. If the scalar component of $`S`$ develops a VEV triggered by SUSY breaking, it breaks the $`U(1)^{}`$ gauge symmetry and induces an effective $`\mu `$-term, linking these scales. Since it is assumed that electroweak symmetry breaking is triggered by SUSY breaking, as well, all scales are related and the $`\mu `$-problem is solved . Moreover, one predicts $`M_Z^{}\stackrel{<}{_{}}𝒪(\text{1 TeV})`$ if one wants to avoid excessive fine-tuning. The $`U(1)_\psi `$ symmetry also forbids all baryon number, B, and lepton number, L, violating terms in the renormalizable superpotential of the MSSM,
$$lh_u,lle^+,lq\overline{d},\overline{d}\overline{d}\overline{u}.$$
(4)
However, each 27 also contains an extra pair of $`SU(2)`$ singlet quarks, $`D`$ and $`\overline{D}`$. Upon $`U(1)^{}`$ breaking they are expected to receive TeV scale masses through terms of the form<sup>3</sup><sup>3</sup>3Terms of this type can also trigger radiative $`U(1)^{}`$ symmetry breaking in analogy to electroweak symmetry breaking in the MSSM, provided the coupling strength is comparable to the top quark Yukawa coupling .,
$$SD\overline{D},$$
(5)
but they also allow new B and L violating (though BL conserving) terms,
$$qqD,\overline{d}\overline{D}\overline{u}.$$
(6)
Together with the B and L conserving operators,
$$lq\overline{D},e^+D\overline{u},\overline{\nu }D\overline{d},$$
(7)
dimension 4 proton decay is reintroduced through the exchange of $`D`$ and $`\overline{D}`$ quarks.
Clearly, one can avoid this problem by restricting oneself to $`SO(10)`$ and identifiying the $`U(1)^{}`$ with the $`U(1)_\chi `$ appearing in $`SO(10)SU(5)\times U(1)_\chi `$, which likewise forbids the terms in Eq. (4). The ordinary fermions and the $`\overline{\nu }`$ complete anomaly free 16 representations (avoiding the extra singlet quarks), and one has to add a pair of Higgs doublets which by anomaly cancellation are required to have equal and opposite $`U(1)^{}`$ charges. As a result, a primordial $`\mu `$-term is allowed, and any term of the form (3) can only include a $`U(1)^{}`$ neutral singlet field, $`S`$. Therefore, both aspects of the $`U(1)^{}`$ solution to the $`\mu `$-problem (exclusion of an elementary $`\mu `$-term and generation of an effective $`\mu `$-term at the TeV scale) are lost, while there is no problem with too rapid proton decay. This is the reversed situation compared to the $`U(1)_\psi `$ case.
One possibility is to ignore the $`\mu `$-problem. Models containing an anomaly free $`U(1)^{}`$ symmetry designed to stabilize the proton have been discussed by various authors . Conversely, non-anomalous $`U(1)^{}`$s addressing the $`\mu `$-problem without reference to proton decay are discussed in Ref. . An interesting solution to both problems is to flip the signs of the hypercharge assignments of the $`D`$ and $`\overline{D}`$ quarks relative to the $`E_6`$ case . This does not alter the SM anomaly conditions, and forbids the dangerous terms in Eq. (6). However, it implies the existence of stable fractionally charged baryons with TeV scale masses. Such states have not been observed, yet, and lead to serious cosmological problems<sup>4</sup><sup>4</sup>4I am indebted to Michael Plümacher for discussions on this point.. If the reheating temperature after inflation is of the order of the mass of these exotic particles or above, their hadronic interactions give rise to rather large number densities of these stable hadrons , orders of magnitude above experimental bounds . Lowering the reheating temperature does not solve the problem either, since successful baryogenesis probably requires a reheating temperature after inflation at least of the order of the critical temperature of the electroweak phase transition of about 100 GeV . Then a significant number density of TeV scale particles can still be produced by the high-energy tail of thermal particle distributions, or during reheating itself . The present work is therefore devoted to models without fractionally charged states, i.e., $`SU(5)`$ charge quantization is imposed,
$$Q+\frac{𝒞_3}{3}=Q_Y+\frac{𝒞_2}{2}+\frac{𝒞_3}{3}\mathrm{𝖹𝖹},$$
(8)
where $`Q`$ is electric charge, $`Q_Y`$ is hypercharge, and where $`𝒞_2`$ is an even (odd) integer for vector (spinor) representations of $`SU(2)`$. Similarly, $`𝒞_3`$ is the triality class of $`SU(3)`$ (e.g. $`\pm 1`$ for quarks and antiquarks).
All of the $`E_6`$ inspired models discussed above have another unwanted feature in the context of gauge coupling unification, which requires the addition of an extra pair of $`SU(2)`$ doublets. By virtue of anomaly cancellation these must be non-chiral and one would therefore need a mechanism preventing them from receiving high scale tree-level masses. This is not a problem in the context of string models which predict the absence of fundamental mass parameters in the low energy effective Lagrangian by conformal invariance. However, as these models stand, the extra doublets would then be massless. One could couple them to an overall singlet receiving a VEV, but in general this is not protected to be of electroweak size. Alternatively, one can assume that the extra doublets develop VEVs, as well, but it seems difficult to arrange that all chargino and neutralino masses are compatible with the experimental lower mass limits as some masses would be predicted to be strictly of order the $`Z`$ boson mass, $`\stackrel{<}{_{}}M_Z`$. This paper is therefore devoted to stabilize electroweak scale physics by the principles of gauge and supersymmetry, while assuring that all fields receive large enough masses after symmetry breaking.
An alternative way to remove the dangerous terms in Eq. (6) would be to choose a particular linear combination of $`U(1)_\chi `$, $`U(1)_\psi `$, and $`U(1)_Y`$. Table 1 shows that the condition,
$$\stackrel{~}{a}+2\stackrel{~}{b}+\frac{5}{3}=0,$$
(9)
allows a Dirac mass term, $`D\overline{d}`$, so that $`D`$ and $`\overline{d}`$ can be removed from the spectrum. Baryon number is now conserved because the trilinear terms in Eq. (4) are still forbidden, and effects from the ones in Eq. (6) decouple. The role of $`\overline{d}`$ is now played by $`\overline{D}`$, but since $`lq\overline{D}`$ is the only type of operator involving it, $`l`$ must act as a Higgs superfield, and trade its role with $`h_d`$. However, Eq. (9) also implies that $`l`$ and $`h_u`$ have equal and opposite $`U(1)^{}`$ charges and furthermore allows $`\overline{\nu }`$ to decouple from the spectrum. Hence, we recover the $`SO(10)`$ case (with $`\overline{\nu }`$ replaced by $`S`$) and the $`\mu `$-problem which comes with it. Eq. (9) defines therefore an alternative set of $`SO(10)`$ models within $`E_6`$, in analogy with the alternative left-right model .
## 3 Anomaly cancellation in the presence of an extra $`U(1)^{}`$
Consider the $`U(1)^{}`$ charge assignment in Table 2. Indicated are three generations of ordinary matter fields, one pair of Higgs doublets, $`h_d`$ and $`h_u`$, appropriate to break electroweak symmetry, one singlet Higgs, $`S`$, for $`U(1)^{}`$ symmetry breaking, and further singlet fields<sup>5</sup><sup>5</sup>5I will refer to a singlet as a right-handed neutrino, $`\overline{\nu }`$, if it admits the Yukawa coupling in Eq. (2)., $`T_i`$, which are (initially) assumed not to receive VEVs. Shown are also examples of extra matter fields, where $`E^{}`$ are charged lepton singlets, $`L`$ are lepton doublets, $`U`$ and $`D`$ are up-type and down-type quark singlets, $`Q`$ are quark doublets, and where $`E^+`$, $`\overline{L}`$, $`\overline{U}`$, $`\overline{D}`$, and $`\overline{Q}`$ are their mirror partners. $`W`$ and $`G`$ are neutral fields under the $`U(1)_Y`$ transforming in the adjoint representations of $`SU(2)`$ and $`SU(3)`$, respectively. These are the fields which will play a role in the models surviving the analysis in Section 4, but initially any field respecting Eq. (8) is allowed. I assume family universality for the ordinary fermions and also the exotic fields, i.e., the new fields have identical $`U(1)^{}`$ charges in the cases of multiplicities greater than one, but this assumption is not crucial and can be relaxed easily.
As usual, all ordinary charged leptons and quarks are expected to become massive through the Yukawa couplings<sup>6</sup><sup>6</sup>6While $`m_u=0`$ is strongly disfavored, it is not firmly ruled out. I will comment on this case later in this Section. in Eq. (2). On the other hand, the extra fields, such as $`D`$ and $`E`$, are assumed to get vector-like (with respect to the SM group) masses induced by trilinear couplings with $`S`$ and their respective $`SU(5)`$ mirror partners, as for example in Eq. (5). This way their contributions to oblique parameters decouple. If an irreducible representation is real, $`𝐫=\overline{𝐫}`$, and has $`Q^{}=s/2`$, it can acquire a supersymmetric Majorana mass through a trilinear coupling to $`S`$. Examples are the fields $`W`$ and $`G`$, and in this case it suffices to add a single copy. Finally, I demand,
$$s=(c_1+c_2)0,$$
(10)
to guarantee a solution to both aspects of the $`\mu `$-problem.
Since the exotic matter is vector-like with respect to the SM gauge group it does not spoil the anomaly cancellation present there. Triangle diagrams with two SM (i.e., either $`SU(3)`$, or $`SU(2)`$, or $`U(1)_Y`$) and one $`U(1)^{}`$ gauge fields as external legs yield three conditions which have to be satisfied. At first sight one would expect that there are many solutions given the many free parameters in Table 2, but as it will turn out there are none. For example, if the exotic matter consists of $`k_5`$ copies of $`\mathrm{𝟓}+\overline{\mathrm{𝟓}}`$ representations of $`SU(5)`$, the mixed $`SU(3)`$/$`U(1)^{}`$ anomaly condition reads,
$$(k_53)(c_1+c_2)=0,$$
(11)
and together with Eq. (10) fixes $`k_5=3`$. Eq. (11) is independent of the $`d_i`$, and so are the mixed $`SU(2)`$/$`U(1)^{}`$ and $`U(1)_Y^2`$/$`U(1)^{}`$ conditions, which are, respectively,
$`(k_5+1)(c_1+c_2)+3(a+3b)=0,`$ (12)
$`({\displaystyle \frac{10}{3}}k_514)(c_1+c_2)6(a+3b)=0.`$ (13)
This is solved only by the $`SO(10)`$-type relation, $`a=3b`$, and for equal and opposite Higgs charges, $`c_1=c_2`$.
This conclusion cannot be avoided even if we generalize to arbitrary extra matter fields of the form $`𝐍+\overline{𝐍}`$, where N is any (in general reducible) representation of $`SU(5)`$. Eqs. (11)–(13) still apply, but $`k_5`$ is now defined as the index of $`𝐍`$,
$$k_5=2\underset{r}{}\frac{n_r\mathrm{dim}𝐫}{\mathrm{dim}SU(5)}C_r$$
(14)
where the sum is over the irreducible representations, $`𝐫𝐍`$. $`n_r`$ are the multiplicities ($`n_r=1/2`$ for Majorana types) and $`C_r`$ is the second-order Casimir invariant of representation r. One can also allow extra matter fields in incomplete representations of $`SU(5)`$. However, preservation of gauge coupling unification enforces the differences of the one-loop $`\beta `$-function coefficients to be the same as in the MSSM<sup>7</sup><sup>7</sup>7I will refer to such configurations as quasi-complete $`SU(5)`$ representations.. This leaves only one adjustable parameter, the index $`k_5`$, and the anomaly conditions remain unchanged.
The conclusion, $`c_1+c_2=0`$, is also independent of the assumption of family universality. If, for example, the $`U(1)^{}`$ charges of the third generation were different from the two lighter ones and associated with parameters $`a^{}`$ and $`b^{}`$ (cf. Table 2), then this would merely result in the replacement,
$$3(a+3b)2a+a^{}+3(2b+b^{}),$$
(15)
in Eqs. (12) and (13).
A more interesting case can be constructed from the one in the previous paragraph by replacing $`(b+c_2)`$ by $`(b^{}+c_2)`$ for the quark singlets $`\overline{u}`$ and $`\overline{c}`$, and conversely for $`\overline{t}`$. This results in a zero eigenvalue for the up-type quark mass matrix, and would either require $`m_u=0`$, or that a non-vanishing $`m_u`$ can be generated radiatively, non-perturbatively, or by some other mechanism. Family non-universal $`U(1)^{}`$ charges can also induce significant flavor changing neutral current effects . If one allows for this possibility, Eqs. (11)–(13) are replaced by,
$$\left(\begin{array}{ccc}0& +1& k_53\\ +1& 3& k_5+1\\ 2& \frac{34}{3}& \frac{10}{3}k_514\end{array}\right)\left(\begin{array}{c}2a+a^{}+9b\\ bb^{}\\ c_1+c_2\end{array}\right)=0.$$
(16)
However, the determinant of this matrix is non-vanishing ($`=4`$) independently of $`k_5`$, and there is only the trivial solution implying $`b=b^{}`$, i.e., nothing new.
The setup discussed in this section does not seem to provide solutions to the $`\mu `$-problem. It should be stressed, however, that this conclusion depends crucially on the additional assumptions of gauge coupling unification and chirality. For example, the model in Ref. is chiral, but fails to produce unification; alternatively, if unification is enforced by adding an extra pair of doublets it turns non-chiral. Another example can be constructed with the assignments in Table 2. Suppose the exotic matter consists of the fields $`(Q+\overline{Q})+(U+\overline{U})`$. Eq. (11) is then satisfied ($`k_5=3`$), and Eq. (13) becomes equivalent to Eq. (12), which are solved by, $`4(c_1+c_2)=3(a+3b)`$. The quadratic anomaly condition from triangle graphs with one hypercharge and two $`U(1)^{}`$ gauge bosons in the external legs depends on $`d_3`$ and $`d_4`$ and can easily be solved. Extra singlets, $`T_i`$, can be added to satisfy the cubic (pure $`U(1)^{}`$) and trace (mixed gravitational/$`U(1)^{}`$) anomaly conditions. This model is chiral but the gauge couplings do not unify. Unification can be arranged by adding an $`E^\pm `$ pair to complete a $`\mathrm{𝟏𝟎}+\overline{\mathrm{𝟏𝟎}}`$ representation of $`SU(5)`$, but it would have to carry equal and opposite $`U(1)^{}`$ charges, spoiling the chirality of the model. Thus, this model mirrors the situation in Ref. except that it has no fractionally charged baryons.
## 4 Two Higgs singlet solutions
Suppose now that besides $`S`$ another MSSM singlet field, $`T=T_0`$, with $`Q^{}=t=t_0`$, develops a VEV. A subset of the exotic matter fields, $`𝐌+\overline{𝐌}𝐍+\overline{𝐍}`$, could then acquire TeV scale masses through trilinear couplings to $`T`$, provided the $`U(1)^{}`$ charge assignments of the fields in $`𝐌`$ and $`\overline{𝐌}`$ are changed<sup>8</sup><sup>8</sup>8In general, this explicitly violates the assumption of universality w.r.t. the exotic fields which was made in Section 3. from $`d_i`$ and $`(s+d_i)`$ to $`e_i`$ and $`(t+e_i)`$, respectively. The anomaly conditions in Eqs. (11)–(13) now change to,
$$\left(\begin{array}{ccc}0& k_5k_33& k_3\\ 1& k_5k_2+1& k_2\\ +2& \frac{10}{3}(k_5k_1)14& \frac{10}{3}k_1\end{array}\right)\left(\begin{array}{c}3(a+3b)\\ s\\ t\end{array}\right)=0,$$
(17)
where $`k_3`$ is the $`SU(3)`$ index of $`𝐌`$, $`k_3\mathrm{dim}SU(3)=2_rm_r\mathrm{dim}𝐫C_r`$. Similarly, $`k_2`$ is the $`SU(2)`$ index of $`𝐌`$, and $`k_1=6/5\mathrm{Tr}_𝐌Q_Y^2`$. The $`SU(5)`$ charge quantization condition (8) implies,
$$\frac{5}{6}k_1\frac{1}{2}k_2\frac{1}{3}k_3k_0\mathrm{𝖹𝖹}.$$
(18)
Notice, that $`k_3`$, $`k_2`$, and $`k_0`$ are integers, and that by definition, $`0k_jk_5`$ for $`1j3`$, which implies the bound $`|k_0|5k_5/6`$. If the determinant of the matrix in Eq. (17) is non-vanishing, there is only the unwanted solution, implying $`c_1+c_2=0`$. The matrix becomes singular for,
$$k_3=(k_53)(k_0+k_2k_3).$$
(19)
Using the definitions,
$$\sigma \frac{(a+3b)}{s},\tau \frac{t}{s},$$
(20)
the solutions can be classified as follows:
* The case $`k_3=k_0+k_2=0`$ implies $`5k_1=3k_2`$, and therefore $`k_0=k_1=k_2=\sigma =\tau =0`$ independently of $`k_5`$, which will be rejected because it corresponds essentially to the non-chiral situation in Section 3. This is the only possibility for $`k_5=1`$.
* If $`k_55`$, Eq. (19) implies that $`k_3`$ is an integer multiple of $`k_53`$, and $`k_jk_5`$ requires $`k_3=k_53`$ and $`k_0+k_2=k_52`$. One has $`3\sigma =(k_5k_2+1)`$ and $`\tau =0`$, i.e., these models are also non-chiral since the fields within the $`𝐌+\overline{𝐌}`$ representation are not tied to $`U(1)^{}`$ symmetry breaking. Such solutions exist for $`k_5=3`$ (e.g., the three generation $`E_6`$ model with an extra pair of doublets) and $`k_5=4`$, as well, but they will be discarded.
* $`k_5=2`$ implies $`k_0+k_2=0`$, $`k_3>k_2`$, $`3\sigma =3+k_2/k_3`$ and $`\tau =1+1/k_3`$. There are three solutions, but the case $`(k_3,k_2,k_1,k_0)=(2,1,1/5,1)`$ cannot be made consistent with the $`SU(5)`$ quantization condition.
* $`k_5=3`$ implies $`k_3=0`$, $`3\sigma =3+k_0/(k_0+k_2)`$ and $`\tau =11/(k_0+k_2)`$. There are eight solutions of this type.
* $`k_5=4`$ implies $`k_3=2`$, $`k_0+k_2=4`$, $`\sigma =5/3k_2/6`$ and $`\tau =1/2`$. There are two solutions, but the case $`(k_3,k_2,k_1,k_0)=(2,3,19/5,1)`$ cannot be made consistent with the $`SU(5)`$ quantization condition.
Thus, there are an infinite number of non-chiral solutions, but only 11 chiral ones, which are collected in Table 3. All of them predict $`\tau =t/s0`$, implying that there are no $`D`$-flat scalar field directions. Furthermore, all of them predict $`k_54`$, which guarantees that the gauge couplings remain perturbative up to the unification scale, while $`k_5=5`$ would yield non-perturbative unification .
In general, each solution in Table 3 can be realized with various exotic matter contents. Observe that the configurations of $`k_i`$ in the Table restrict the possible matter content to the fields displayed in Table 2. These fields can be combined into quasi-complete $`SU(5)`$ representations shown in Table 4.
In the remainder of this Section and in the next Section I will focus on exotic matter in complete $`\mathrm{𝟓}+\overline{\mathrm{𝟓}}`$ representations of $`SU(5)`$, unless noted otherwise. In this case one can write the sum over the other singlet charges, which is fixed by the trace condition, $`\mathrm{Tr}Q^{}=0`$, as
$$\omega _1\frac{1}{s}\underset{i=1}{}t_i=5k_5+(\tau 1)(3k_3+2k_21)12.$$
(21)
Interestingly, when 3 right-handed neutrinos are included, $`\omega _1`$ is completely determined within each model. For $`\mathrm{𝟓}+\overline{\mathrm{𝟓}}`$ representations one has $`k_0=0`$, and then Eqs. (10), (17), and (18) imply the relation,
$$s=(c_1+c_2)=(a+3b),$$
(22)
which is familiar from the $`E_6`$ assignment in Table 1. Thus, most of the phenomenological studies of $`E_6`$ inspired $`Z^{}`$ bosons apply here, and also to the other solutions with $`k_0=0`$. To simplify the algebra, I impose the orthogonality condition, $`\mathrm{Tr}Q_YQ^{}=0`$, which can be relaxed at a later stage. Combined with the anomaly condition quadratic in the $`U(1)^{}`$ charges<sup>9</sup><sup>9</sup>9$`\mathrm{Tr}Q_YQ_{}^{}{}_{}{}^{2}`$ also vanishes if $`s=0`$., $`\mathrm{Tr}Q_YQ_{}^{}{}_{}{}^{2}=0`$, I find,
$`2(k_3e_1+k_2e_2)`$ $`=`$ $`9{\displaystyle \frac{(ab)+(c_1c_2)}{\tau 1}}\tau (k_3+k_2)s,`$ (23)
$`2[(k_3k_5)d_1+(k_2k_5)d_2]`$ $`=`$ $`{\displaystyle \frac{(6\tau +3)(ab)+(7\tau +2)(c_1c_2)}{\tau 1}}+[2k_5(k_3+k_2)]s.`$ (24)
Similarly, the cubic condition, $`\mathrm{Tr}Q_{}^{}{}_{}{}^{3}=0`$, reads,
$$\omega _3\frac{1}{s^3}\underset{i=1}{}t_i^3=5k_5+(\tau ^31)(3k_3+2k_21)\frac{3}{4}\left[3+9\left(\frac{ab}{s}\right)^2+4\left(\frac{c_1c_2}{s}\right)^2\right]+$$
(25)
$$3\left[3\tau k_3\frac{e_1}{s}\left(\frac{e_1}{s}+\tau \right)+2\tau k_2\frac{e_2}{s}\left(\frac{e_2}{s}+\tau \right)+3(k_5k_3)\frac{d_1}{s}\left(\frac{d_1}{s}+1\right)+2(k_5k_2)\frac{d_2}{s}\left(\frac{d_2}{s}+1\right)\right],$$
and one can show that,
$$\omega _3\frac{1}{4}\left[5(k_51)+(\tau ^31)(3k_3+2k_2)\right]\tau ^3.$$
(26)
The inequality (26) is saturated when,
$$a=b=\frac{s}{4},c_1=c_2=d_1=d_2=\frac{s}{2},e_1=e_2=\frac{\tau }{2}s,$$
(27)
which yields a class of $`Z^{}`$ bosons which couple like the $`Z_\psi `$ to the ordinary fermions and the Higgs doublets. This is encouraging as the recent analysis of precision data revealed an excellent fit to the $`Z_\psi `$ assuming some amount of mixing with hypercharge. Recall that $`U(1)^{}`$ subgroups from $`E_6`$ are frequently discussed because they represent manifestly anomaly free examples (and because they arise in popular GUT and string models), but the $`E_6`$ gauge symmetry and Yukawa coupling relations are usually explicitly broken to avoid serious conflicts with observation (such as the proton lifetime). Here we recover $`U(1)^{}U(1)_{\stackrel{~}{\psi }}`$ symmetries which are very similar to the $`U(1)_\psi `$ from a very different bottom-up approach.
Relaxing the orthogonality condition, $`\mathrm{Tr}Q_YQ^{}=0`$, and using the trivial solution to Eq. (17), $`a+3b=c_1+c_2=s=t=0`$, one can construct another family of special solutions, $`U(1)_{\stackrel{~}{\chi }}`$, where the MSSM fields have charges as in the $`U(1)_\chi `$ case. While they would yield non-chiral pairs of Higgs doublets, these solutions could be interesting if they enter linear combinations with the $`U(1)_{\stackrel{~}{\psi }}`$ or if the corresponding $`Z_{\stackrel{~}{\chi }}`$ coexists with the $`Z_{\stackrel{~}{\psi }}`$. The quadratic anomaly condition now reduces to
$$(k_3k_5)d_1+(k_2k_5)d_2\tau (k_3e_1+k_2e_2)+2(a+c_1)=0,$$
(28)
where solutions with $`c_1=a`$ correspond to hypercharge and $`c_1=2a/3`$ to the $`U(1)_{\stackrel{~}{\chi }}`$. Note, that there are other solutions to the anomaly constraints with $`s=0`$, but only those satisfying Eq. (28) can mix with the chiral $`U(1)^{}`$s.
The encounter of the $`U(1)_{\stackrel{~}{\psi }}`$ and $`U(1)_{\stackrel{~}{\chi }}`$ groups is also encouraging from a top-down perspective. There are classes of string models where gauge coupling unification is predicted without the appearance of a full GUT group in the zero-slope limit. Examples are heterotic string constructions where the gauge group factors considered are realized at the same Kac-Moody level, $`k`$, or open string models with the group factors arising from the same $`D`$-brane. The $`E_8\times E_8`$ heterotic string is particularly suited here, because it offers the prerequisite $`E_6`$ subgroups. The reference $`E_8`$ symmetry implies the relevant matching (unification) conditions, as well as quantization conditions for its $`U(1)`$ subgroups. The latter are given by
$$Q_\chi \frac{6}{5}Q_Y\mathrm{𝖹𝖹},Q_\psi \frac{5}{4}Q_\chi \mathrm{𝖹𝖹},Q_\varphi \frac{4}{3}Q_\psi \mathrm{𝖹𝖹},$$
(29)
where the last relation refers to the $`U(1)_\varphi `$ defined by $`E_7E_6\times U(1)_\varphi `$. The normalizations are chosen such that $`Q_Y=1`$ for $`e^+`$ ($`a=1/2`$), $`Q_\chi =1`$ for $`\overline{\nu }`$ ($`a=3/5`$), $`Q_\psi =1`$ for $`S`$ ($`s=1`$), and $`Q_\varphi =1`$ for one of the $`E_6`$ singlets within the 56 of $`E_7`$. Eq. (28) implies for the $`U(1)_{\stackrel{~}{\chi }}`$ of solution II, $`2d_2+3e_1+2/5=0`$, or when combined with the first Eq. (29), $`e_1+2/52\mathrm{𝖹𝖹}`$. With the exception of two fields, the second Eq. (29) is also satisfied. But the $`D`$ quarks have the sign of their $`U(1)_{\stackrel{~}{\psi }}`$ charges reversed and the $`T`$ field ($`Q_{\stackrel{~}{\chi }}=0`$) has half-integer instead of integer $`U(1)_{\stackrel{~}{\psi }}`$ charge. It is not clear at present<sup>10</sup><sup>10</sup>10In Kac-Moody level 1 models $`SU(5)`$ charge quantization can never be obtained unless a complete $`SU(5)`$ GUT group survives . On the other hand in the level 2 models or Ref. the third Eq. (29) is respected by all fields. whether $`E_8`$ charge quantization, i.e., Eqs. (29), are predicted in string models with gauge coupling unification and $`SU(5)`$ charge quantization. If so one would have to impose the condition $`\tau \mathrm{𝖹𝖹}`$, and only solution I would survive.
The $`E_8`$ embedding also allows one to fix the overall normalization (coupling strength) of the $`U(1)^{}`$ . In units in which an $`E_8`$ root vector $`P`$ has length $`P^2=1`$, one has $`P^2=3/5`$ for a 10 of $`SU(5)`$, $`P^2=5/8`$ for a 16 of $`SO(10)`$, $`P^2=2/3`$ for a 27 of $`E_6`$, and $`P^2=3/4`$ for a 56 of $`E_7`$. The 10 of $`SU(5)`$ is relevant here, because it contains the $`SU(3)\times SU(2)`$ singlet with unit charge (the $`e^+`$) and can therefore serve to properly normalize the hypercharge gauge coupling, $`g_Y`$, in terms of one of the non-Abelian gauge couplings, $`g`$. Assuming the normalization in Eq. (29) one finds the well-known result, $`g_Y=\sqrt{3/5}g`$. Notice, that this argument needs no reference to a complete GUT group like $`SU(5)`$, but it applies only to string constructions in which the embedding into $`E_8`$ is still traceable. Referring again to the normalization fixed by Eq. (29) one finds in a similar way, $`g_\chi =\sqrt{5/8}g`$ and $`g_\psi =\sqrt{2/3}g`$.
The reference $`E_8`$ root vector also allows to set lower limits on the possible Kac-Moody levels, $`k`$. For example, the singlet $`T`$ of solution II has $`Q_{\stackrel{~}{\psi }}=3s/2`$, which implies $`P_T^2=3/2>1`$ and means that $`T`$ cannot be obtained in the massless sector if $`k=1`$, while $`k2`$ would be allowed by this consideration. Interestingly, the theorem about fractionally charged states in level $`k=1`$ models implies the same conclusion.
## 5 Phenomenological aspects
The one-loop $`\beta `$-function coefficient for the $`U(1)^{}`$ is given by,
$$\frac{\beta _1^{}}{s^2}=\omega _2+5k_5+(\tau ^21)(3k_3+2k_2+1)+6+9\left(\frac{ab}{s}\right)^2+7\left(\frac{c_1c_2}{s}\right)^2+$$
(30)
$$2\left[3k_3\frac{e_1}{s}\left(\frac{e_1}{s}+\tau \right)+2k_2\frac{e_2}{s}\left(\frac{e_2}{s}+\tau \right)+3(k_5k_3)\frac{d_1}{s}\left(\frac{d_1}{s}+1\right)+2(k_5k_2)\frac{d_2}{s}\left(\frac{d_2}{s}+1\right)\right],$$
where $`\omega _2=s^2_{t=i}t_i^2`$. $`\beta _1^{}`$ is minimized when Eqs. (27) are satisfied, i.e., for the $`U(1)_{\stackrel{~}{\psi }}`$, in which case one finds,
$$\beta _1^{}=\frac{1}{3}\left[5(k_5+2)+(\tau ^21)(3k_3+2k_2)+2(\tau ^2+\omega _2)\right].$$
(31)
In Eq. (31) I have chosen the $`E_6`$ inspired normalization, $`s^2=2/3`$. For comparison, the $`SU(3)\times SU(2)\times U(1)_Y`$ one-loop $`\beta `$-function coefficients are $`\beta _3=k_53`$, $`\beta _2=k_5+1`$, and $`\beta _1=k_5+33/5`$ (using $`SU(5)`$ normalization, $`a^2=3/20`$), respectively.
For concreteness, I now focus on the $`Z_{\stackrel{~}{\psi }}`$ of solution II which is reminiscent of an $`E_6`$ model with three families of 27, but with one copy of $`D+\overline{D}`$ quarks removed and the $`U(1)^{}`$ charges of the other two copies altered. The trace and cubic anomaly conditions, $`\omega _1=1/2`$ and $`\omega _3=\omega _3^{\mathrm{max}}=23/16`$, can be satisfied by adding, for example, $`SU(5)`$ singlets with $`U(1)^{}`$ charges, $`t_1=t_2=s`$, $`t_3=3s/4`$, $`t_{2i+2}=s/2`$, and $`t_{2i+3}=s/4`$, where $`1i9`$, or in short-hand notation, $`t_i[1^2,3/4,(1/2)^9,(1/4)^9]`$. There are no singlets with charges $`1`$ or $`3/2`$ in this choice, assuring that the crucial fields, $`S`$ and $`T`$, cannot acquire high scale masses. The quantum numbers for this case, $`Z_{\stackrel{~}{\psi }}^{(\mathrm{II})}`$, are summarized in Table 5.
There are a total of 12 fields with quantum numbers like the right-handed neutrinos. Since these fields are protected from acquiring high scale Majorana masses, one predicts a Dirac mass matrix with three electroweak scale eigenvalues. I.e., there are unacceptably large neutrino masses, unless the Yukawa couplings are chosen zero or tiny. A clean way to cure this problem is to add the set of singlets<sup>11</sup><sup>11</sup>11By virtue of the identity, $`n^34(n1)^3+6(n2)^34(n3)^3+(n4)^3=0`$, such a configuration does not alter $`\omega _1`$ or $`\omega _3`$., $`t_i3[1,(3/4)^4,(1/2)^6,(1/4)^4]`$. Now the fields with $`t_i=\pm s/4`$ (and some others) can form bilinear Dirac mass terms, resulting in a neutrino mass matrix with three zero eigenvalues, which correspond predominantly to the left-handed neutrinos provided the bilinear masses are much larger than the Yukawa mass terms. Clearly, the bilinear masses can be taken to infinity and the corresponding fields can be smoothly removed from the spectrum, yielding the residual chiral set of singlets (now including $`S`$, $`T`$, and the three $`\overline{\nu }`$), $`[3/2,1^6,(3/4)^{11},(1/2)^9]`$.
There is another interesting solution which can be obtained by adding the fields, $`t_i3[(5/4),1^4,(3/4)^6,(1/2)^4,(1/4)]`$ to the ones in Table 5. This results in the chiral singlet set, $`[3/2,(5/4)^3,1^{15},(3/4)^{17},(1/2)^3,(1/4)^9]`$, where the fields with $`Q^{}=5s/4`$ allow trilinear couplings with $`S`$ and the $`\overline{\nu }`$. This once again results in three zero eigenvalues, and if the masses generated by $`S`$ are an order of magnitude larger than the ones from the standard Yukawa terms, these also correspond predominantly to the left-handed neutrinos, but the mixing with the massive states could induce a small amount of missing invisible $`Z`$ width, $`\mathrm{\Gamma }_{\mathrm{inv}}`$. Indeed, the LEP Collaborations currently report a shortage of about 0.5% in $`\mathrm{\Gamma }_{\mathrm{inv}}`$. Furthermore, the Fermi constant, $`G_F`$, extracted from $`\mu `$ decays could increase, significantly increasing the extracted Higgs boson mass from electroweak fits (the central value of which is currently below the lower search limit ) but potentially deteriorating the quality of the global fit to all data. There can be other effects associated with this scenario, such as a violation of weak charged current universality. There are strong limits on the non-universality between the first two families, however, driven by the well measured branching ratio, $`\frac{(\pi \mu \nu _\mu )}{(\pi e\nu _e)}`$. Thus the effect in $`\mathrm{\Gamma }_{\mathrm{inv}}`$ should be expected to be dominated by the $`\nu _\tau `$ sector, in agreement with the experience that the third family Yukawa couplings are typically the largest.
Yet another solution to the problem of neutrino mass would be to implement the see-saw mechanism which can be achieved by choosing a linear combination of the $`U(1)_{\stackrel{~}{\psi }}`$ and the $`U(1)_{\stackrel{~}{\chi }}`$, such that the $`\overline{\nu }`$ are $`U(1)^{}`$ neutral and allowed to receive high scale Majorana masses. As is well known, this also produces small but non-vanishing Majorana masses for the left-handed neutrinos, which is not the case for the scenarios in the previous two paragraphs. However, one should generally expect non-vanishing masses for the left-handed neutrinos originating from non-renormalizable terms suppressed by powers of the high mass scale . Dimension 5 operators would yield the same kind of suppression as in the see-saw mechanism, but higher order suppressions (by intermediate scales) are conceivable<sup>12</sup><sup>12</sup>12Employing the $`Z^{}`$ scale, $`M_Z^{}800`$ GeV , and the unification scale, $`M2\times 10^{16}`$ GeV, suggests neutrino masses of order $`M_Z^{}^2/M0.03`$ eV, in excellent agreement with the mass scale inferred from atmospheric neutrino oscillations . One might wonder whether this agreement not only signals the presence of a new physics scale close to the unification scale, but possibly even suggests the absence of further scales below that..
As for baryon number violating interactions, the $`U(1)_{\stackrel{~}{\psi }}`$ symmetry of solution II forbids the terms in Eq. (6), thus assuring proton stability at the renormalizable level. Dimension 5 operators can also lead to proton decay at rates conflicting with the experimental limits. However, the $`D`$-terms,
$$qq\overline{d}^{},qq\overline{D}^{},DD\overline{u}^{},\overline{u}\overline{d}D^{},\overline{u}\overline{D}D^{},$$
(32)
are also absent by virtue of the $`U(1)_{\stackrel{~}{\psi }}`$, and likewise for the $`F`$-terms,
$$lqqq,H_dqqq,Lqqq,H_uqDD,\overline{L}qDD,e^+DDD.e^+\overline{u}\overline{u}\overline{d}.$$
(33)
On the other hand, the $`F`$-terms,
$$\overline{\nu }qqD,\overline{\nu }\overline{u}\overline{d}\overline{D},T_i^{3/4}\overline{u}\overline{d}\overline{d},T_i^{+5/4}\overline{u}\overline{D}\overline{D},$$
(34)
are guaranteed to be absent only when there are no MSSM singlets with $`Q^{}=5s/4`$, $`s/4`$, or $`3s/4`$. While this can be arranged, there is another dimension 5 $`F`$-term,
$$e^+\overline{u}\overline{u}\overline{D},$$
(35)
which is similar to the operator $`e^+\overline{u}\overline{u}\overline{d}`$, present in the $`R`$-parity conserving version of the MSSM. There one has to assume that its corresponding coupling strength is small, and taking into account suppressions from first generation Yukawa couplings, Cabibbo mixing, and phase space, the resulting proton lifetime can be acceptable. In comparison, the term (35) is less concerning, because at the renormalizable level the $`D`$ quarks couple only (ignoring the possibility of a $`T_i^{+1/2}\overline{d}D`$-term) via gauge interactions and the $`TD\overline{D}`$-term, both conserving “$`D`$-number”. Therefore, observable proton decay is not expected, unless a large $`\overline{d}\overline{D}`$ mixing effect can be generated. Alternatively, the critical dimension 5 $`F`$-term in Eq. (35) can be removed if there is an admixture of the $`U(1)_{\stackrel{~}{\chi }}`$. Thus, choosing $`k_3=k_5`$ (or more generally coupling all of the exotic quarks to singlets other than $`S`$) is an efficient strategy to achieve a realistic proton lifetime.
## 6 Conclusions
I have presented a class of models with an additional non-anomalous $`U(1)^{}`$ in which gauge couplings unify, the $`\mu `$-problem can be solved, and electric charge is quantized. These models are chiral, protecting a large desert between the electroweak (SUSY breaking) and unification (string) scales. Moreover, in some of these models the $`U(1)^{}`$ is sufficient to avoid conflicts with experimental bounds on the proton lifetime. Within a deliberately traditional and conservative framework the analysis presented is rather general, but there are various directions which can be chosen to go beyond it.
For example, I assumed that two MSSM singlet fields participate in $`U(1)^{}`$ symmetry breaking. Only a finite number of solutions exist since the anomaly matrix in Eq. (17) has to be singular. The analysis generalizes straightforwardly if one allows three (or more) singlets to acquire VEVs. In this case an infinite number of solutions exist, a large (but finite) number of which being consistent with perturbative gauge coupling unification. A preliminary study of this class reveals that unlike the solutions in Table 3 there are also models with $`D`$-flat scalar field directions, a property which can be phenomenologically advantageous. There are also solutions simultaneously in accordance with $`E_8`$ charge quantization (see Eq. (29)) and the proton lifetime.
I have further assumed that all exotic fields receive their masses through couplings to MSSM singlets acquiring VEVs upon $`U(1)^{}`$ breaking. But exotic leptons participating in $`SU(2)\times U(1)_Y`$ symmetry breaking will in general receive mass contributions of $`𝒪(M_Z)`$ which could conceivably by themselves be large enough to be in agreement with current limits. This seems difficult but is a possible loophole which I have ignored in this work.
An implicit assumption has been made concerning the precise form of gauge coupling unification. It is the often made assertion that gauge coupling unification works satisfactorily in the MSSM and should not be altered in leading (one-loop) order. Extra matter multiplets are then usually restricted to appear exclusively in complete $`SU(5)`$ representations. In the present work, I have already relaxed this restriction and included quasi-complete $`SU(5)`$ representations. However, gauge coupling unification works only approximately, with the prediction for the strong coupling constant, $`\alpha _s(M_Z)`$, significantly higher than most observations and the unification scale at least one order of magnitude lower than the string scale. It is fair to ask whether some arbitrary (incomplete from the perspective of $`SU(5)`$) representation could improve the quality of unification. Indeed, by including a set of fields which contributes stronger to the $`SU(3)`$ $`\beta `$-function than to the one of $`SU(2)`$ improves the prediction for $`\alpha _s(M_Z)`$. Furthermore, by having the smallest contribution to the $`U(1)_Y`$ $`\beta `$-function, one can increase the unification scale. The minimal choice is the set $`G+W+E^\pm `$, contributing $`\mathrm{\Delta }\beta _3=3`$, $`\mathrm{\Delta }\beta _2=2`$, and $`\mathrm{\Delta }\beta _1=6/5`$, respectively. Alternatively, one could add another $`\mathrm{𝟓}+\overline{\mathrm{𝟓}}`$ of $`SU(5)`$ to this set, or choose a family plus a mirror-family of quarks, $`Q+\overline{Q}`$, $`U+\overline{U}`$, and $`D+\overline{D}`$. Both cases correspond to $`\mathrm{\Delta }\beta _3=4`$, $`\mathrm{\Delta }\beta _2=3`$, and $`\mathrm{\Delta }\beta _1=11/5`$. The case $`\mathrm{\Delta }\beta _3=5`$ (and the same $`\beta `$-function differences) would yield unification in the non-perturbative domain, but is also conceivable. In any case, these choices yield $`SU(2)\times U(1)_Y`$ unification slightly above the reduced Planck scale of $`2.4\times 10^{18}`$ GeV. The prediction for $`\alpha _s(M_Z)`$ is now below typical observed values but the discrepancy is in general neither much better nor much worse than in the MSSM. If it is found that higher order or threshold effects tend to raise the prediction for $`\alpha _s`$, these scenarios may be of potential interest, but I have not considered them in the analysis of this paper. In this context it is amusing to note that one would obtain a very good prediction for $`\alpha _s`$, as well as unification very close to the string scale, if these extra fields have intermediate scale masses near the geometric mean of the string and electroweak scales.
Even given these possible directions it is possible to draw some general conclusions from the kind of analysis introduced in this work. Having left the framework of $`E_6`$ inspired $`U(1)^{}`$ models, anomaly cancellation tends to provide $`U(1)`$ groups which have charges to the ordinary fermions just as the $`U(1)`$ subgroups of $`E_6`$. The charges of some of the extra fields typically differ, sometimes forbidding fast proton decay mediating couplings. A more thorough phenomenological investigation will be presented elsewhere.
## Acknowledgement:
It is a pleasure to thank Paul Langacker and Michael Plümacher for many fruitful discussions. This work was supported in part by U.S. Department of Energy Grant EY–76–02–3071.
|
warning/0006/hep-th0006053.html
|
ar5iv
|
text
|
# Macroscopical matter derived from Quantum Field Theory
## I Introduction
The aim of this article is to find the macroscopical expression of the energy-momentum tensor of a free scalar field derived from Quantum Field Theory (QFT) up to the first quantum order in the WKB expansion, and to study its macroscopical behavior. In previous works we established the relation among QFT variables -amplitudes and phases- with macroscopic fluid variables -proper energy and four-velocity- but considering only the lowest order in $`\mathrm{}`$ in the scheme of general curved space-times, in Riemannian and non-Riemannian geometries. In both cases, free scalar matter corresponds to perfect fluids without pressure following geodesics. In fact, this statement also holds for flat Minkowski space-time where geodesics must be understood as straight non-accelerated trajectories (inertial principle). In the present article we prove that up to the following order, free scalar fields correspond to perfect fluids with pressure. We also find the corresponding law of motion which shows dispersions from inertial motion. Up to our knowledge, these are new results which give a deeper insight in the relation between second quantization and standard description of macroscopical matter. We believe that these results contribute to place QFT as the ultimate description of matter structure.
The paper is organized as follows: in section II we expand the energy-momentum tensor of the free scalar field. In section III, we find the relation between microscopical and macroscopical variables and prove that free scalar matter must be represented by a perfect fluid with pressure when quantum corrections are taken into account. In IV we find the equation which relates the WKB amplitudes that are involved in the corresponding definitions of proper energy densities and pressures; this section is also devoted to find the first quantum correction to the classical equations of motion. Finally, in V, we make a summary of our results.
## II The WKB expansion of the energy-momentum tensor
In order to study the relationship between quantum and classical dynamics, we start considering the complex wave function $`\mathrm{\Psi }`$ in terms of the path integral formalism. In the first quantization scheme, the wave function of the particle may be expressed as:
$$\mathrm{\Psi }(x_o,x_f)=Ne^{iS_{cl}[x]/\mathrm{}}𝒟x$$
(1)
where $`S_{cl}[x]`$ is the classical action which is a functional of the trajectory $`x(t)`$ depending on the initial and final fixed data ($`x_o,x_f`$) and $`N`$ is a normalization factor. If $`S_{cl}`$ is the action of a non-relativistic particle, then $`\mathrm{\Psi }`$ corresponds to a wave function which satisfies the Schroedinger equation . On the other hand, when we work in the scheme of second quantization, $`\mathrm{\Psi }`$ must be taken to be a field satisfying the relativistic field equation obtained for the corresponding representation of the Lorentz group. Field equations arise minimizing action $`S[\mathrm{\Psi }]`$ with respect to the field:
$$\frac{\delta S[\mathrm{\Psi }]}{\delta \mathrm{\Psi }}=0$$
(2)
The expansion of eq. (1) in powers of $`\mathrm{}`$ reads:
$$\mathrm{\Psi }=Ne^{iS_{cl}[x_{cl}]/\mathrm{}}e^{{\scriptscriptstyle {\scriptscriptstyle \frac{1}{2}}\delta S_{cl}[x]/\delta x_1\delta x_2𝒟x_1𝒟x_2}+\mathrm{}}\text{ }𝒟x=\underset{n=0}{}(i\mathrm{})^n\mathrm{\Psi }_ne^{iS_{cl}/\mathrm{}}$$
(3)
with $`x_{cl}`$ the classical trajectory. This expression is known as the WKB expansion. For any group representation, this expansion means that:
$$S[\mathrm{\Psi }]=S_{cl}[x]+𝒪(\mathrm{})$$
(4)
So, the $`\mathrm{}=0`$ limit of $`S[\mathrm{\Psi }]`$ in a first quantization scheme leads to $`S_{cl}[x],`$ in the same way as the $`\mathrm{}=0`$ limit of the effective action conduces, in a second quantization scheme, to $`S[\mathrm{\Psi }].`$
We are interested in analyzing the case of a complex scalar field $`\mathrm{\Phi }.`$ Its action reads:
$$S[\mathrm{\Psi }]=(\mathrm{\Phi },\mathrm{\Phi })\text{ }d^4x=\frac{1}{2}(^\mu \mathrm{\Phi }_\mu \mathrm{\Phi }^{}+\frac{m^2}{\mathrm{}^2}\mathrm{\Phi }\mathrm{\Phi }^{})\text{ }d^4x$$
(5)
The corresponding energy-momentum tensor is:
$$T^{\mu \nu }\frac{}{_\mu \mathrm{\Phi }}^\nu \mathrm{\Phi }+\frac{}{_\mu \mathrm{\Phi }^{}}^\nu \mathrm{\Phi }^{}\eta ^{\mu \nu }$$
(6)
> which conduces to:
>
> $$T^{\mu \nu }=^{(\mu }\mathrm{\Phi }^{\nu )}\mathrm{\Phi }^{}\frac{1}{2}\eta ^{\mu \nu }^\lambda \mathrm{\Phi }_\lambda \mathrm{\Phi }^{}\frac{1}{2}\eta ^{\mu \nu }\frac{m^2}{\mathrm{}^2}\mathrm{\Phi }\mathrm{\Phi }^{}$$
> (7)
> The WKB expansion (3) for the scalar field reads:
>
> $$\mathrm{\Phi }=\underset{n=0}{}(i\mathrm{})^n\phi _ne^{iS_{_{cl}}/\mathrm{}}$$
> (8)
> Now, using expansion (8) in expression (7), and after a straightforward computation, we find that up to order $`1/\mathrm{}`$, the energy momentum tensor can be written as:
$`T^{\mu \nu }=\frac{1}{\mathrm{}^2}^\mu S^\nu S\phi _o\phi _o^{}+\frac{1}{\mathrm{}}\{^{(\mu }S`$ Im$`(\phi _o^{\nu )}\phi _o^{})2^\mu S^\nu S`$ Im($`\phi _o\phi _1^{})+`$
$$\eta ^{\mu \nu }[(^\lambda S_\lambda S)\text{ Im}(\phi _o\phi _1^{})+(^\lambda S)\text{ Im}(\phi _o_\lambda \phi _o^{})+m^2\text{ Im}(\phi _o\phi _1^{})]\}$$
(9)
> where $`S_{cl}=S,`$ and Im means “imaginary part. It is easy to verify that this symmetric quantity is conserved, i.e. satisfies:
>
> $$_\mu T^{\mu \nu }=0$$
> (10)
## III The Scalar Field as a Perfect Fluid
> Let us analyze the $`\mathrm{}^2`$ order (the first term) of expression (9). We may write the corresponding term as
>
> $$T^{(1/\mathrm{}^2)\mu \nu }=\widehat{\rho }u^\mu u^\nu $$
> (11)
> where
>
> $$\widehat{\rho }=\frac{m^2}{\mathrm{}^2}\phi _o\phi _o^{}$$
> (12)
> is the proper energy and $`u^\mu `$ is the timelike vector orthogonal to the hypersurface of constant phase $`S`$, where $`_\mu S=P^\mu `$ and $`S`$ is understood as the Hamilton principal function. In the free case, canonical momentum $`P^\mu `$ coincides with the ordinary momentum $`p^\mu `$, i.e.:
>
> $$P^\mu =p^\mu =mu^\mu $$
> (13)
> Four-velocity $`u^\mu `$ satisfies $`u^\mu u_\mu =1.`$ Relation (11) shows that $`T^{(1/\mathrm{}^2)\mu \nu }`$ is nothing but the energy-momentum tensor of a dust.
>
> Using (13) in (9) we are able to write the energy-momentum tensor showing explicitly the four-velocity components:
>
> $`T^{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{m^2}{\mathrm{}^2}}\phi _o\phi _o^{}u^\mu u^\nu 2{\displaystyle \frac{m}{\mathrm{}}}u^{(\mu }\text{Im}(\phi _o^{\nu )}\phi _o^{})`$ (15)
> $`2{\displaystyle \frac{m^2}{\mathrm{}}}u^\mu u^\nu \text{ Im(}\phi _o\phi _1^{}){\displaystyle \frac{m}{\mathrm{}}}u^\lambda \text{ Im}(_\lambda \phi _o\phi _o^{})\eta ^{\mu \nu }`$
> From (14) we can see that at a point where the observer is at rest (proper observer), i.e.: $`\stackrel{~}{u}^\mu =(1,0,0,0)`$, the components of $`T^{\mu \nu },`$ which we denote as $`\stackrel{~}{T}^{\mu \nu },`$ are:
>
> $$\stackrel{~}{T}^{oo}=\frac{m^2}{\mathrm{}^2}\phi _o\phi _o^{}\frac{m}{\mathrm{}}[2m\text{ Im}(\phi _o\phi _1^{})+\text{Im(}\phi _o\dot{\phi }_o^{})]$$
> (16)
> $$\stackrel{~}{T}^{oi}=0$$
> (17)
> $$\stackrel{~}{T}^{ij}=\frac{m}{\mathrm{}}\text{Im(}\phi _o\dot{\phi }_o^{})\text{ }\delta ^{ij}$$
> (18)
> (with $`i`$, $`j`$, $`\mathrm{}`$ $`=`$ $`1`$, $`2`$, $`3`$). This will be useful in order to find the general expression in any Lorentzian frame.
>
> Once we have written the components of $`T^{\mu \nu },`$ we see that it is possible to define:
>
> $$\rho =\frac{m^2}{\mathrm{}^2}\phi _o\phi _o^{}\frac{m}{\mathrm{}}[2m\text{ Im}(\phi _o\phi _1^{})+\text{Im(}\phi _o\dot{\phi }_o^{})]\widehat{\rho }\stackrel{˘}{\rho }$$
> (19)
> $$p=\frac{m}{\mathrm{}}\text{Im(}\phi _o\dot{\phi }_o^{})$$
> (20)
> which are the proper energy density and the pressure, respectively. Here $``$ $`d/d\tau `$, $`\tau `$ being the proper time, $`\widehat{\rho }`$ is given by (12) while $`\stackrel{˘}{\rho }`$ is $`\stackrel{˘}{\rho }=\frac{m}{\mathrm{}}[2m`$ Im$`(\phi _o\phi _1^{})+`$Im($`\phi _o\dot{\phi }_o^{})].`$ So, the components of $`\stackrel{~}{T}^{\mu \nu }`$ read:
>
> $$\stackrel{~}{T}^{oo}=\rho $$
> (21)
> $$\stackrel{~}{T}^{oi}=0$$
> (22)
> $$\stackrel{~}{T}^{ij}=p\text{ }\delta ^{ij}$$
> (23)
> Expressions (20-22) represent the fluid that a proper observer would describe as isotropic, i.e. a perfect fluid.
>
> Now we want to generalize expressions (20-22) from this system to any Lorentz frame. To do this we analyze how $`T^{\mu \nu }`$ may be ‘measured ’ in a reference frame at rest in the laboratory supposing that the fluid appears to be moving at each space-time point with velocity $`v^i`$. This velocity is related to the spatial components of the four velocity $`u^i`$ of the fluid via:
>
> $`v^i=\gamma ^1u^i\text{ }`$
> where $`\gamma =(1\overline{v}^2/c^2)^{1/2}`$ is the boost that transforms Lorentz reference frames among them. Applying this boost to components (20-22) of $`T^{\mu \nu },`$ we obtain:
>
> $$T^{oo}=(\rho +p\overline{v}^2)\gamma ^2$$
> (24)
> $$T^{oi}=(\rho +p)v^i\gamma ^2$$
> (25)
> $$T^{ij}=p\delta ^{ij}+(\rho +p)v^iv^j\gamma ^2$$
> (26)
> with $`\rho `$ and $`p`$ given by (18) and (19), respectively.
>
> Combining (23), (24) and (25), we see that in any Lorentz frame, expression (14) reads:
>
> $$T^{\mu \nu }=\rho u^\mu u^\nu +p(\eta ^{\mu \nu }+u^\mu u^\nu )$$
> (27)
> i.e.: a perfect fluid. In conclusion: expression (9) for $`T^{\mu \nu }`$ which is given by QFT, corresponds to a perfect fluid that can be written only in terms of macroscopical variables.
>
> In order to connect proper energy density and pressure we would only need an equation of state. The structure of this equation depends on the properties of the system. In the particular case in which temperature is an independent variable, pressure is a function of $`\rho _o`$ and $`\mathrm{\Pi }`$:
>
> $$p=p(\rho _o,\mathrm{\Pi })$$
> (28)
> where
>
> $$\rho =\rho _o(1+\mathrm{\Pi })$$
> (29)
> $`\rho _o`$ being the proper mass density and the product $`\rho _{o^.}\mathrm{\Pi }`$, the internal energy density.
## IV Relation between the WKB amplitudes and equations of motion
In QFT, field equations are obtained minimizing the effective action with respect to the field (equation (2)). For the scalar representation $`\mathrm{\Phi }`$, this conduces to the Klein-Gordon equation:
$$\mathrm{}\mathrm{\Phi }+\frac{m^2}{\mathrm{}^2}\mathrm{\Phi }=0$$
(30)
which in terms of the WKB expansion of the field, reads:
$$\underset{n}{}[\mathrm{}^2\mathrm{}\varphi _n+i\mathrm{}(2^\mu \varphi _n_\mu S+\varphi _n\mathrm{}S)\varphi _n^\mu S_\mu S)(i\mathrm{})^n]=_nm^2\varphi _n(i\mathrm{})^n$$
(31)
> Assuming that (30) must be satisfied order by order in $`\mathrm{},`$ we obtain the differential equations which are needed to describe the perfect fluid via eqs. (18) and (19):
>
> $$n=0^\mu S_\mu S+m^2=0$$
> (32)
> $$n=12i^\mu \varphi _o_\mu S+i\varphi _o\mathrm{}S+\varphi _1(\mathrm{}S+m^2)=0$$
> (33)
> Equation (31) provides no new information: it is trivially satisfied because $`^\mu S=P^\mu =mu^\mu `$ and $`u^\mu u_\mu =1`$. Equation (32) may be written as:
>
> $$2u^\mu _\mu \varphi _o+(\varphi _oi\varphi _1)_\mu u^\mu im\varphi _1=0$$
> (34)
> and relates the first two amplitudes of the WKB expansion of the field.
In order to find the equations of motion corresponding to scalar particles, we compute equation (10) using (19) and obtain:
$$^\mu P+_\nu [(\rho +P)u^\mu u^\nu ]=0$$
(35)
Keeping only those equations corresponding to the spatial components of the velocity $`\overline{v},`$ we find the following vectorial equation:
$$(\rho +P)[\frac{\overline{v}}{t}+(\overline{v}\overline{})\overline{v}]=\gamma ^2[\overline{}P+\overline{v}\frac{P}{t}]$$
(36)
where we have used the fact that $`u^i=v^iu^o`$. In eq. (35), different orders in $`1/\mathrm{}^n`$ are involved due to the corresponding expressions of $`\rho `$ and $`P`$ (see (18) and (19)). So, for the first two orders, we obtain two equalities. Up to the $`1/\mathrm{}^2`$ (main) contribution we have :
$$\frac{\overline{v}}{t}+(\overline{v}\overline{})\overline{v}=0$$
(37)
Taking into account the $`1/\mathrm{}`$ correction, eq. (36) becomes:
$$(\widehat{\rho }\stackrel{˘}{\rho }+P)[\frac{\overline{v}}{t}+(\overline{v}\overline{})\overline{v}]=\gamma ^2[\overline{}P+\overline{v}\frac{P}{t}]$$
(38)
The first equation tells us that, at the highest order in the WKB expansion, the free massive scalar field is represented by particles with null acceleration, i.e., their trajectories are straight lines. This is the classical limit of the approach and represents the inertial principle. But we can see from eq. (37) that quantum corrections to the straight line appear at the next order. That is to say: free matter follows trajectories that deviate from straight lines and from non-accelerated uniform motion.
We want to recall that equation (37) may be considered the relativistic second law of Newton: $`F^\mu =mdu^\mu /d\tau `$, where quantum corrections to $`\rho `$ and the appearance of $`P`$ (expressions (18) and (19) respectively), are responsible for deviations from straight lines with no need of external forces nor of quantum potentials (as, for example, in the Bohm’s scheme; see). Equation (37) reveals the nature of the internal structure of matter and how it acts in order to produce quantum effects.
## V Summary
We have developed the WKB expansion of the scalar field up to the first quantum correction and studied the relations between QFT description of matter and macroscopic fluids. Our aim was to find the behaviour of this matter in a macroscopical picture, asumming quantum corrections. Our main results are the following:
\- at the first quantum approximation, free scalar fields correspond to perfect fluids with pressure,
\- the corresponding law of motion implies deviations from straight unaccelerated motion.
>
## VI References
M. Castagnino, G. Domenech, M. Levinas and N. Umérez, Class.Quant.Grav. 6 (1989) L 1.
G. Domenech, M. Levinas and N. Umérez, Phys. Lett. 137A (1989) 17.
R. Feynman and A. Hibbs, Quantum Mechanics and Path Integral, New York, Mc.Graw-Hill, 1965, chap. 4.
D. Bohm, B. Hiley and P. Kaloyerou, Phys. Rep.144 No. 6 (1987) 321.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.