1: \documentclass[eqsecnum,preprint]{revtex4}
2:
3: %\documentclass[preprint,aps]{revtex4}
4: %\documentclass[aps,rmp]{revtex4}
5: %\documentclass[12pt,twoside]{book}
6:
7:
8:
9: \makeatletter % Macros for arabic section numbering
10: %\renewcommand\l@section{\@dottedtocline{2}{5em}{5em}}
11: \renewcommand{\thesection}{\arabic{section}}
12: \renewcommand{\p@subsection}{}
13: \renewcommand{\thesubsection}{\arabic{section}.\arabic{subsection}}
14: \renewcommand{\p@subsubsection}{}
15: \renewcommand{\thesubsubsection}{\arabic{section}.\arabic{subsection}.\arabic{subsubsection}}
16: \makeatother
17:
18:
19:
20: \newcommand{\lbfig}[1]{\refstepcounter{fig} \label{#1} }
21: \newcounter{fig}
22:
23:
24: \usepackage{amssymb}
25: \usepackage{amsmath}
26: \usepackage{bbm}
27: \usepackage{epsf}
28: \usepackage{epsfig}
29: \usepackage{psfrag}
30: \usepackage{citesort}
31: %\usepackage{here}
32: %\usepackage{refman}
33: %\usepackage{fancyheadings}
34: %\usepackage{latexsym}
35:
36: \renewcommand{\baselinestretch}{1.15}
37: \renewcommand{\topfraction}{0.85}
38: \renewcommand{\bottomfraction}{.85}
39: \renewcommand{\textfraction}{0.15}
40: \renewcommand{\floatpagefraction}{0.8}
41:
42: \parindent0em
43: \sloppy
44:
45: \textwidth 7.1in
46: \topmargin-.7in
47: %\topmargin-0.2in
48: \oddsidemargin-.3in
49: \textheight 9.6in
50:
51:
52: \newcommand{\beq}{\begin{equation}}
53: \newcommand{\eeq}{\end{equation}}
54: \newcommand{\beqa}{\begin{eqnarray}}
55: \newcommand{\eeqa}{\end{eqnarray}}
56: \newcommand{\Tr}{\text{Tr}}
57:
58: \newcommand{\del}{\partial}
59: \newcommand{\delk}{\partial_{k_z}}
60: \newcommand{\tk}{\tilde{k}_0}
61: \newcommand{\tw}{\tilde{\omega}}
62:
63: \newcommand{\deldag}{\mathbin{\partial\mkern-10.5mu\big/}}
64: \newcommand{\mDdag}{\mathbin{{\mathbf D}\mkern-14mu\big/}}
65: \newcommand{\kdag}{\mathbin{k\mkern-9.8mu\big/}}
66: \newcommand{\ndag}{\mathbin{n\mkern-9mu\big/}}
67: \newcommand{\hdag}{\mathbin{h\mkern-10mu\big/}}
68: \newcommand{\npdag}{\mathbin{n'\mkern-13.5mu\big/}}
69:
70: \newcommand{\bm}{\mathbf{m}}
71: \newcommand{\bX}{\mathbf{X}}
72: \newcommand{\bY}{\mathbf{Y}}
73:
74: \def\DBR#1#2{\bigl\{#1 \bigr\}\bigl\{#2 \bigr\}} % The diamond bracket
75: \def\Slash#1{#1\kern-0.55em\raise.05ex\hbox{/}}
76: \def\slash#1{#1\kern-0.5em\raise.05ex\hbox{{$\scriptstyle /$}}}
77:
78:
79:
80:
81:
82:
83:
84:
85:
86:
87:
88: \begin{document}
89:
90:
91:
92:
93: \pagenumbering{Roman}
94:
95: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
96: % %
97: % %
98: % OOOOO O OOOOO O OOOOO OOOO OOO OOOO OOOOO %
99: % O O O O O O O O O O O %
100: % O O O O OOOO OOOO OOOOO O OOO OOOO %
101: % O O O O O O O O O O O %
102: % O O O OOOOO OOOOO O O O OOO OOOOO %
103: % %
104: % %
105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106:
107: \begin{titlepage}
108:
109: \vskip 0.3in
110:
111: \preprint{BNL-72343-2004-JA, HD-THEP-03-62}
112:
113: \vskip 0.6in
114:
115: \centerline{\LARGE \bf Transport equations for chiral fermions to order $\hbar$}
116: \vskip 0.2in
117: \centerline{\LARGE \bf and electroweak baryogenesis: Part~I}
118:
119: \vskip 0.4in
120:
121: \begin{center}
122:
123: {\Large \bf
124: Tomislav Prokopec$^{\diamond}$,
125: Michael G. Schmidt$^\diamond$ \\
126: \vskip 0.2in
127: and Steffen Weinstock$^\bullet$}
128:
129:
130: \vskip 0.3in
131: {\it $^\diamond$ Institut f\"ur Theoretische Physik, Universit\"at Heidelberg\\
132: Philosophenweg 16, D-69120 Heidelberg, Germany} \\
133:
134: \vskip 0.1in
135: {\it $^\bullet$ Nuclear Theory Group, Brookhaven National Laboratory, Upton, NY 11973-5000, USA}
136:
137:
138: \end{center}
139:
140: \vskip 0.4in
141:
142: \centerline{\bf Abstract}
143: \vskip 0.2in
144:
145:
146:
147: This is the first in a series of two papers. In this first part,
148: we use the Schwinger-Keldysh formalism to derive semiclassical Boltzmann
149: transport equations, accurate to order $\hbar$,
150: for massive chiral fermions, scalar particles,
151: and for the corresponding CP-conjugate states.
152: Our considerations include complex mass terms and
153: mixing fermion and scalar fields, such that
154: CP-violation is naturally included, rendering the equations
155: particularly suitable for studies of baryogenesis
156: at a first order electroweak phase transition.
157: We provide a quantitative criterion for when the reduction to
158: the diagonal kinetic equations in the mass eigenbasis is justified,
159: leading to a quasiparticle picture even in the case of mixing scalar
160: or fermionic particles.
161: Within the approximations we make, it is possible
162: to first study the Boltzmann equations without the collision term.
163: In a second paper~\cite{PSW_2} we discuss the collision terms
164: and reduce the Boltzmann equations to fluid equations.
165:
166:
167:
168: \vskip 0.4in
169:
170: \footnoterule
171: \vskip 3truemm
172: {\small\tt
173: \noindent
174: $^\diamond$T.Prokopec@, M.G.Schmidt@thphys.uni-heidelberg.de \\
175: $^\bullet$weinstock@bnl.gov
176: }
177:
178:
179: \end{titlepage}
180:
181:
182:
183:
184:
185:
186:
187:
188:
189:
190:
191:
192:
193: \pagenumbering{arabic}
194:
195: %%%%%%%%%%%%%%%%%%%%%%%%%%
196: % %
197: % %
198: % OOOOO OOO OOOO %
199: % O O O O %
200: % O O O O %
201: % O O O O %
202: % O OOO OOOO %
203: % %
204: % %
205: %%%%%%%%%%%%%%%%%%%%%%%%%%
206:
207: \tableofcontents
208:
209:
210:
211:
212:
213:
214:
215:
216: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
217: % %
218: % %
219: % OOOOO O O OOOOO OOOO OOO OOO O O OOOO OOOOO O OOO O O %
220: % O OO O O O O O O O O O O O O O O O OO O %
221: % O O O O O OOOO O O O O O O O O O O O O O O %
222: % O O OO O O O O O O O O O O O O O O O OO %
223: % OOOOO O O O O O OOO OOO OOO OOOO O O OOO O O %
224: % %
225: % %
226: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
227:
228: \cleardoublepage
229: \section{Introduction}
230:
231: \vskip 0.05in
232:
233: The question of how Boltzmann transport equations emerge from
234: quantum field theory is an old
235: problem~\cite{Schwinger:1961,KadanoffBaym:1962,Keldysh:1964}. It seems that
236: a consensus has been reached, according to which a suitable
237: (loop) truncation of the self-energies in the Kadanoff-Baym equations,
238: supplied by an on-shell reduction, leads to Boltzmann transport equations
239: for the dynamics of quasiparticles in phase space.
240: While procedures leading to classical
241: Boltzmann equations -- which formally correspond to taking the
242: limit $\hbar\rightarrow 0$ of the Kadanoff-Baym equations --
243: seem to be quite clear and well
244: established~\cite{ChouSuHaoYu:1985,CalzettaHu:1986},
245: no systematic treatment has been pursued beyond the classical approximation,
246: which would lead to a kinetic description valid to linear order in an $\hbar$
247: expansion
248: (a notable exception is a study of interference between the tree-level
249: and one-loop decay rates of relevance for grand unified
250: baryogenesis~\cite{KolbWolfram:1980}.)
251: The inclusion of $\hbar$ corrections is absolutely
252: necessary for any treatment of baryogenesis, which represents the primary
253: impetus for the work presented here,
254: since CP-violating effects vanish in the limit
255: when $\hbar\rightarrow 0$. More concretely,
256: one deals with particle transport in high temperature relativistic plasmas
257: with Minimal Standard Model particle content,
258: often extended to include exotic, as-of-yet undiscovered new particles,
259: {\it e.g.} various supersymmetric partners.
260: While $\hbar$ expansions are often
261: identified with loop expansions in field theory, here we associate $\hbar$
262: primarily to a gradient expansion
263: (which is a field-theoretical generalization of the WKB expansion)
264: with respect to space-time variations of background fields.
265: The relevant examples of background fields for baryogenesis are
266: the mass terms generated by the (Higgs) order parameter
267: at a first order electroweak phase transition; other examples include
268: classical (possibly stochastic) background gauge fields.
269: One important hurdle, which must be overcome in the development of any
270: formalism that aspires to accurately describe transport in plasmas beyond
271: the classical level, is the establishment of a quasiparticle picture.
272: This entails a demonstration of the existence and the
273: identification of the relevant quasiparticle excitations in the plasma,
274: which can be used to encode the relevant dynamical information.
275: This is the problem we solve here.
276: Apart from baryogenesis, potential applications of the formalism developed
277: in this work include propagation of neutrinos
278: in dissipative media, transport of electrons and holes in quantum wires and
279: quantum semiconductor devices, dynamics of Bose-Einstein condensates,
280: {\it etc.}
281:
282:
283: \vskip 0.05in
284:
285: The standard approach to electroweak
286: baryogenesis~\cite{KuzminRubakovShaposhnikov:1985}
287: requires a study of the generation and transport of CP-violating
288: flows~\cite{CohenKaplanNelson:1990+1991} arising from interactions of fermions
289: with expanding phase transition fronts of a first order electroweak
290: phase transition. The most prominent theoretical problem in such
291: scenarios over the past few years has been to
292: find a systematic derivation of the appropriate transport equations, including
293: the CP-violating sources. In particular, there has been controversy in the
294: literature regarding what are the dominant
295: sources appearing in the transport equations and how to compute them.
296: This work is an attempt to resolve this daunting question.
297:
298: \vskip 0.05in
299:
300: %As regards equilibrium considerations,
301: It is known that a strong first order transition is a necessary requirement
302: for electroweak baryogenesis~\cite{Shaposhnikov:1986,Shaposhnikov:1987}.
303: While absent in the
304: Minimal Standard Model~\cite{KajantieLaineRummukainenShaposhnikov,
305: RummukainenTsypinKajantieLaineShaposhnikov:1998,CsikorFodorHeitger:1998,
306: AokiCsikorFodorUkawa:1999,
307: BuchmullerPhilipsen:1994,BergerhoffWetterich:1994},
308: a strong transition can be realized in its extensions, which include
309: two Higgs doublet models~\cite{LaineRummukainen:2001},
310: the Minimal Supersymmetric Standard Model
311: (MSSM)~\cite{CarenaQuirosWagner:1996,CarenaQuirosWagner:1998,
312: ClineKainulainen:1996,Laine:1996,
313: Losada97,BodeckerJohnLaineSchmidt:1997,LaineRummukainen:1998,
314: PilaftsisWagner:1999,CsikorFodorHegedusJakovacKatzPiroth:2000,
315: Espinosa:1996,deCarlosEspinosa:1997},
316: and its nonminimal extensions (NMSSM)~\cite{Pietroni:1993,
317: DaviesFroggattMoorhouse:1996,HuberSchmidt:1999}.
318: %
319: In the NMSSM, not only the possibility of electroweak baryogenesis
320: is given in wide regions of the parameter space, but it can also have stronger
321: CP-violation, increasing the amount of the actually produced baryon
322: asymmetry~\cite{HuberJohnLaineSchmidt}.
323:
324:
325: \vskip 0.05in
326:
327: Considering the derivation of the relevant transport equations,
328: there are essentially two approaches in literature. The first approach,
329: which we shall refer to as the {\it semiclassical force
330: mechanism}~\cite{JoyceProkopecTurok:1995,JoyceProkopecTurok:1996}, is based
331: on the observation that, when treated in the WKB approximation,
332: a CP-violating quadratic part of the fermionic Lagrangean
333: exhibits a CP-violating shift at first order in gradients in the dispersion
334: relation, which manifests itself as a CP-violating semiclassical
335: force in the Boltzmann equation. This force then appears
336: in the fluid transport equations, or equivalently diffusion equations
337: for the relevant particle densities. In the original
338: work~\cite{JoyceProkopecTurok:1995,JoyceProkopecTurok:1996} baryogenesis
339: from scattering of top quarks and tau leptons in two Higgs doublet models
340: was considered. In the subsequent
341: work~\cite{ClineJoyceKainulainen:1998}
342: the semiclassical force formalism was extended to include the case of fermion
343: mixing, which is relevant for example for the {\it chargino} induced
344: baryogenesis in the Minimal Supersymmetric Standard Model
345: (MSSM)~\cite{ClineJoyceKainulainen:1998,Kainulainen:1999,ClineKainulainen:2000,
346: ClineJoyceKainulainen:2000+2001,HuberJohnSchmidt:2001} and in its nonminimal
347: extensions (NMSSM)~\cite{HuberSchmidt:2000,HuberSchmidt:2001}.
348:
349: \vskip 0.05in
350:
351: The second approach, commonly referred to as
352: {\it spontaneous baryogenesis}~\cite{CohenKaplanNelson:1991}, is based
353: on the observation that in the presence of CP-violation the (fermion)
354: hypercharge currents are not conserved. As a consequence, the energy levels for
355: CP-conjugate fermionic states are mutually shifted.
356: In the presence of scatterings and transport,
357: this shift then leads to different populations
358: for chiral fermions and the corresponding antifermions,
359: thus sourcing baryogenesis. It was then pointed out in
360: Refs.~\cite{DineThomas:1994}, \cite{JoyceProkopecTurok:1996}
361: and \cite{ComelliPietroniRiotto:1995}
362: that the hypercharge current, which sources spontaneous baryogenesis,
363: is suppressed by the square of the mass.
364: (This must be so simply because the hypercharge current is conserved in the
365: limit when the mass of particles vanishes,
366: so that the spontaneous source must vanish, too.)
367: Mostly in order to incorporate the possibility of fermionic mixing,
368: the spontaneous baryogenesis mechanism was
369: subsequently developed by several groups,
370: which adopted the idea to baryogenesis studies mediated by the stops and
371: charginos of the MSSM~\cite{HuetNelson:1995+1996,Riotto:1995,Riotto:1998,
372: CarenaQuirosRiottoViljaWagner:1997,
373: CarenaMorenoQuirosSecoWagner:2000,
374: CarenaQuirosSecoWagner:2002} and NMSSM~\cite{DaviesFroggattMoorhouse:1996}.
375:
376: \vskip 0.05in
377:
378: From the current literature, it is in fact not clear whether the semiclassical
379: force and spontaneous sources represent identical or different sources,
380: or if there perhaps has been a conceptual problem in calculating (one or both of) the
381: sources. The situation was made even more controversial by a recent
382: work~\cite{CarenaQuirosSecoWagner:2002}, in which it was claimed that
383: the spontaneous baryogenesis and semiclassical force sources
384: correspond formally to the same source in fluid transport equations,
385: arguing that the method used for computing the semiclassical force source was
386: incorrect.
387:
388: \vskip 0.05in
389:
390: This is to be compared with the methodology
391: advocated in the work based on the WKB
392: method~\cite{ClineJoyceKainulainen:2000+2001},
393: and more recently in~\cite{KainulainenProkopecSchmidtWeinstock:2001,
394: KainulainenProkopecSchmidtWeinstock:2002,
395: KainulainenProkopecSchmidtWeinstock:2002b,
396: KainulainenProkopecSchmidtWeinstock:2002c}
397: based on the Schwinger-Keldysh formalism~\cite{Schwinger:1961},
398: or more precisely on a gradient expansion of the Kadanoff-Baym
399: equations~\cite{KadanoffBaym:1962} in a weak coupling regime. In order
400: to clarify the point of disagreement, we first extend the approach developed
401: in~\cite{KainulainenProkopecSchmidtWeinstock:2001,
402: KainulainenProkopecSchmidtWeinstock:2002}.
403: We consider the dynamics of non-equilibrium Wigner functions for massive chiral
404: fermions which couple to a CP-violating Higgs condensate of advancing
405: phase fronts to first order in a gradient expansion.
406: Assuming that the relevant mass parameter depends only on one
407: spatial coordinate, which models reasonably well the limit of
408: large almost planar bubble walls of a strong first order phase transition,
409: implies a conserved spin, and the constraint equation for the Wigner function
410: is solved by a spectral on-shell {\it ansatz}, thus proving
411: the validity of the {\it quasiparticle picture} of the plasma.
412: This result holds both in the case of one fermion
413: and in the case of mixing fermions.
414:
415: \vskip 0.05in
416:
417:
418: An important difference, regarding the source calculation
419: in spontaneous baryogenesis and semiclassical force baryogenesis,
420: is the choice of the basis in flavor space, which leads to qualitative
421: differences in the resulting baryon production.
422: In studies of chargino-mediated baryogenesis flavor mixing is of
423: crucial importance. While the proponents of spontaneous baryogenesis
424: argue that the weak interaction (flavor) basis (of charginos)
425: is the right basis to work in, since it is favored by the
426: interactions~\cite{CarenaQuirosSecoWagner:2002}, the semiclassical force camp
427: contends that the mass eigenbasis has apparent advantages.
428: Of course, any physical quantity must be basis independent,
429: but only when one includes flavor mixing in transport
430: equations~\cite{KonstandinProkopecSchmidtWeinstock:2004},
431: which -- as of yet -- has not been done.
432: %
433: We strongly favor the mass eigenbasis:
434: as we prove in section~\ref{Constraint and kinetic equations},
435: when working in the mass eigenbasis and to order $\hbar$ accuracy,
436: the diagonal densities decouple from the off-diagonals.
437: The decoupling works uniquely in the mass eigenbasis.
438: The proof is valid provided the mass eigenvalues are not nearly degenerate,
439: that is when $\hbar k\cdot \partial \ll \delta(m_d^2)$ is satisfied, where
440: $\delta(m_d^2)$ signifies the split in the mass eigenvalues. While this
441: strongly indicates that the mass eigenbasis is singled out, the definite
442: resolution of the controversy awaits a basis independent treatment.
443:
444:
445:
446: \vskip 0.05in
447:
448: The paper is organized as follows.
449:
450: \vskip 0.05in
451:
452: In section~\ref{From the 2PI effective action to kinetic theory}
453: we review the Schwinger-Keldysh formalism, suitable for out-of-equilibrium
454: dynamics of field theoretical correlators, and present a derivation
455: of the relevant Kadanoff-Baym equations in Wigner space
456: for our model Lagrangean of chiral fermions coupled to scalars.
457: We include the case of flavor mixing in both the fermionic and
458: the scalar sector.
459:
460:
461: \vskip 0.05in
462:
463: In section~\ref{Reduction to the on-shell limit} we study
464: an on-shell reduction of the Kadanoff-Baym equations for a single scalar field.
465: We show that it is consistent to include the (real part of the)
466: self-energy and the collision term to both the constraint and kinetic
467: equation, provided one truncates the self-energy
468: and the collision term at the same order in the coupling constant.
469: This resolves one of the important -- and up-to-now unanswered -- questions
470: of Boltzmann-like kinetics of relativistic systems.
471: Even though we do not attempt to generalize our result
472: to the kinetics of fermions, it is quite plausible that the same conclusion can
473: be reached. In the subsequent sections we assume that this is the case.
474:
475: \vskip 0.05in
476:
477: A proof that no source appears in the flow term of the kinetic equations
478: for mixing scalars at order $\hbar$ for planar walls is presented in
479: section~\ref{Kinetics of scalars: tree-level analysis}.
480: We complete the proof originally presented
481: in~\cite{KainulainenProkopecSchmidtWeinstock:2001}. Next, we discuss scalar
482: kinetic equations for the CP-conjugate scalar densities, which
483: are then used in Paper~II~\cite{PSW_2}
484: %section~\ref{Scalar collision term}
485: to identify the CP-violating source in the scalar collision term.
486:
487: \vskip 0.05in
488:
489: Section~\ref{Kinetics of fermions: tree-level analysis}
490: is devoted to the tree-level kinetics of mixing fermions in the presence
491: of a CP-violating, complex mass matrix. An important application includes
492: chargino and neutralino mediated baryogenesis in supersymmetric extensions
493: of the Minimal Standard Model. We begin by identifying a conserved quantity,
494: which for planar bubble walls and in the wall frame corresponds to
495: the spin in the direction
496: $\vec n \propto (k_xk_z,k_yk_z,k_0^2 - k_x^2 - k_y^2)$ , which {\it does not}
497: correspond to helicity, which has been erroneously identified as the conserved
498: quantity in most of the literature on semiclassical force baryogenesis.
499: Then we construct Wigner functions which are block diagonal in the
500: conserved spin. Just like in the scalar case,
501: we prove that to order $\hbar$ the diagonal and off-diagonal densities
502: decouple in the mass eigenbasis (we fill in a gap in the original derivation
503: in Ref.~\cite{KainulainenProkopecSchmidtWeinstock:2001}),
504: provided the mass eigenvalues
505: are not nearly degenerate, $\hbar k\cdot \partial \ll \delta(|m_d|^2)$.
506: %
507: The final part of the section contains the derivation of Boltzmann-like
508: transport equations for the fermion distribution functions, with
509: particular emphasis on the kinetics of CP-violating densities.
510: The relevant CP-violating sources are identified, some of which
511: are unaccounted for in the existing literature.
512: For example, we show that a CP-even deviation from equilibrium in
513: the distribution function can contribute as a CP-violating source at order
514: $\hbar$.
515:
516: \vskip 0.05in
517:
518: In Paper~II~\cite{PSW_2} we address the collision terms of the Boltzmann
519: equations, which contain further CP-violating sources. We furthermore
520: simplify the Boltzmann equations to a set of fluid equations,
521: based on which one can study the dynamics of CP-violating currents.
522:
523:
524:
525:
526:
527:
528:
529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
530: % %
531: % %
532: % OOO OOOO O %
533: % O O O O O %
534: % O OOOO O %
535: % O O O %
536: % OOOOO O O %
537: % %
538: % %
539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
540:
541: \cleardoublepage
542: \section{From the 2PI effective action to kinetic theory}
543: \label{From the 2PI effective action to kinetic theory}
544:
545:
546: \subsection{Green Functions}
547:
548:
549: The {\it in-out-formalism} of quantum field theory is well suited for
550: the description of particle scattering experiments, in which the system is
551: prepared to be in a definite {\it in-state} at $t\rightarrow-\infty$ and
552: the question is what is the probability amplitude for the system to be found
553: in a definite {\it out-state} at $t\rightarrow+\infty$.
554: In statistical physics, however, it is often of interest to follow the temporal
555: evolution of a system. Starting with definite initial conditions at some
556: time $t=t_0$, we ask for the expectation values of physical
557: quantities at finite times.
558: A theoretical framework for such problems was first suggested by Schwinger
559: in 1961~\cite{Schwinger:1961}
560: and then developed further by Keldysh~\cite{Keldysh:1964} and others.
561: An extension of field theory capable of dealing with non-equilibrium problems
562: is obtained by defining the time arguments of all quantities on a path
563: $\cal C$ that leads from the initial time $t_0$ to $t$ and then back
564: to $t_0$, as illustrated in figure~\ref{figure-CTP}.
565: All integrals and derivatives have then to be performed along
566: that path, and the usual time ordering becomes time ordering $T_{\cal C}$
567: along ${\cal C}$. This formalism is called Schwinger-Keldysh formalism
568: or Closed Time Path (CTP) formalism~\cite{ChouSuHaoYu:1985}.
569: %
570: \begin{figure}[tbp]
571: \centerline{\hspace{.in}
572: \epsfig{file=A_pics/ctp-contour.eps, width=4.in,height=0.6in}
573: }
574: \vskip -0.1in
575: \caption{\small
576: The complex time contour for the Schwinger-Keldysh non-equilibrium formalism.
577: }
578: \lbfig{figure-CTP}
579: \end{figure}
580: %
581: The definitions of the scalar and the fermionic Green functions are
582: %
583: \begin{eqnarray}
584: i\Delta(u,v)
585: &\equiv& \langle \Omega| T_C \, \phi(u) \phi^\dagger(v) |\Omega\rangle
586: \label{Green_scalar_CTP}
587: \\
588: %
589: iS_{\alpha\beta}(u,v)
590: &\equiv& \langle \Omega| T_{\cal C} \, \psi_\alpha(u) \bar{\psi}_\beta(v)
591: |\Omega\rangle
592: \,,
593: \label{Green_fermionic_CTP}
594: \end{eqnarray}
595: %
596: where $|\Omega\rangle $ denotes the physical state of the system, and
597: $\phi$ and $\psi$ are the scalar and fermionic fields, respectively.
598: The contour $\cal C$ can now be split into a ${\cal C}_+$ branch from $t_0$
599: to $t$ and a ${\cal C}_-$ branch from $t$ to $t_0$, as shown
600: in figure~\ref{figure-CTP}. An important simplification, which will
601: help us to perform the Wigner transform, instrumental for
602: gradient expansion, is an extension of the integration path
603: to $t_0\rightarrow -\infty$ and $t\rightarrow \infty$.
604: This simplification is suitable for problems of our interest, which
605: comprise plasmas close to chemical and thermal equilibrium with efficient
606: equilibration processes, as it is indeed the case with the electroweak plasma.
607: In this case the influence of initial conditions can be neglected.
608: Denoting the branch on which the time argument lies by an index
609: $a=\pm$, we can rewrite the formalism using ordinary time arguments.
610: We then have
611: %
612: \begin{eqnarray}
613: \int_{\cal C} d^4u &\rightarrow& \sum_a a \int_{-\infty}^\infty d^4u
614: \nonumber \\
615: \delta_{\cal C}(u-v) &\rightarrow& a \delta_{ab} \delta(u-v)
616: \nonumber\\
617: i\Delta(u,v) &\rightarrow& i\Delta^{ab}(u,v)
618: \nonumber\\
619: iS(u,v) &\rightarrow& iS^{ab}(u,v)
620: \,.
621: \label{translation}
622: \end{eqnarray}
623: %
624: The additional factors of $a$ in the integral and in the $\delta$-function
625: appear since the ${\cal C}_-$ branch runs backward in time.
626:
627: \vskip 0.1in
628:
629: In this Keldysh formulation one usually defines the following
630: four scalar Green functions:
631: %
632: \begin{eqnarray}
633: i\Delta^{++}(u,v) &\equiv& i\Delta^t(u,v) \;
634: = \langle\Omega|
635: T [\phi(u) \phi^\dagger(v) ]
636: |\Omega\rangle
637: \nonumber\\
638: i\Delta^{+-}(u,v) &\equiv& i\Delta^<(u,v)
639: = \langle\Omega|
640: \phi^\dagger(v) \phi(u)
641: |\Omega\rangle
642: \nonumber\\
643: i\Delta^{-+}(u,v) &\equiv& i\Delta^>(u,v)
644: = \langle\Omega|
645: \phi(u) \phi^\dagger(v)
646: |\Omega\rangle
647: \nonumber\\
648: i\Delta^{--}(u,v) &\equiv& i\Delta^{\bar t}(u,v) \;
649: = \langle\Omega|
650: \overline T [ \phi(u) \phi^\dagger(v) ]
651: |\Omega\rangle
652: \,,
653: \label{Green_scalar_index}
654: \end{eqnarray}
655: %
656: and similarly the fermionic Green functions
657: %
658: \begin{eqnarray}
659: iS_{\alpha\beta}^{++}(u,v) &\equiv& iS_{\alpha\beta}^t(u,v)
660: = \;\;\, \langle\Omega|
661: T [ \psi_\alpha(u) \bar{\psi}_\beta(v)]
662: |\Omega\rangle
663: \nonumber\\
664: %
665: iS_{\alpha\beta}^{+-}(u,v) &\equiv& iS_{\alpha\beta}^<(u,v)
666: = - \langle\Omega|
667: \bar{\psi}_\beta(v) \psi_\alpha(u)
668: |\Omega\rangle
669: \nonumber\\
670: %
671: iS_{\alpha\beta}^{-+}(u,v) &\equiv& iS_{\alpha\beta}^>(u,v)
672: = \;\;\, \langle\Omega|
673: \psi_\alpha(u) \bar{\psi}_\beta(v)
674: |\Omega\rangle
675: \nonumber\\
676: %
677: iS_{\alpha\beta}^{--}(u,v) &\equiv& iS_{\alpha\beta}^{\bar{t}}(u,v)
678: = \;\;\, \langle\Omega|
679: \overline{T} [\psi_\alpha(u) \bar{\psi}_\beta(v)]
680: |\Omega\rangle
681: \,,
682: \label{Green_fermionic_index}
683: \end{eqnarray}
684: %
685: where ${T}$ ($\overline{T}$) denotes time (`anti-time') ordering
686: and the additional minus sign in the second fermion line is due to
687: the anticommutation property of the fermionic fields.
688:
689: \vskip 0.1in
690:
691: These definitions
692: %~(\ref{Green_scalar_index}) and~(\ref{Green_fermionic_index})
693: imply immediately the following hermiticity properties for the Wightman
694: functions
695: %
696: \begin{eqnarray}
697: \label{S_herm_config}
698: \left( i \Delta^{<,>}(u,v) \right)^\dagger &=& i \Delta^{<,>}(v,u)
699: \\
700: \left( i\gamma^0 S^{<,>}(u,v) \right)^\dagger &=& i\gamma^0 S^{<,>}(v,u)
701: \,.
702: \label{Delta_herm_config}
703: \end{eqnarray}
704: %
705: The four two-point functions~(\ref{Green_scalar_index})
706: and~(\ref{Green_fermionic_index}) are not
707: completely independent. Indeed, with the definition of time and anti-time
708: ordering one finds for the chronological (Feynman) and anti-chronological
709: Green functions
710: %
711: \begin{eqnarray}
712: G^t(u,v) &=& \theta(u_0-v_0) G^>(u,v) + \theta(v_0-u_0)G^<(u,v)
713: \nonumber\\
714: %
715: G^{\bar{t}}(u,v) &=& \theta(u_0-v_0) G^<(u,v) + \theta(v_0-u_0)G^>(u,v)
716: \,,
717: \label{Gtt_by_G<>}
718: \end{eqnarray}
719: %
720: where we have used the generic notation $G = \{\Delta,S\}$.
721: From now on we use this notation in relations which are identical
722: for bosonic and fermionic Green functions.
723:
724: \vskip 0.1in
725:
726: From the definitions of the retarded and advanced propagators
727: %
728: \begin{eqnarray}
729: G^r &\equiv& G^t - G^< = G^> - G^{\bar t}
730: \nonumber\\
731: G^a &\equiv& G^t - G^> = G^< - G^{\bar t}
732: \label{G-ra}
733: \end{eqnarray}
734: %
735: and the definitions of their hermitean and antihermitean parts
736: %
737: \begin{eqnarray}
738: G_h &\equiv& \frac 12 (G^r+G^a)
739: %
740: \nonumber\\
741: %
742: G_a &\equiv& \frac {1}{2i} (G^a-G^r)
743: = \frac {i}{2} (G^>-G^<)
744: \equiv {\cal A}
745: \,,
746: \label{G:ha}
747: \end{eqnarray}
748: %
749: one obtains the spectral relation
750: %
751: \beq
752: G_h(u,v) = -i \, {\rm sign}(u^0-v^0)\, {\cal A}(u,v)
753: \,,
754: \label{G-h}
755: \eeq
756: %
757: where ${\rm sign}(u^0-v^0)=\Theta(u^0-v^0)-\Theta(v^0-u^0)$,
758: $u^\mu = (u^0,\vec u)$.
759: ${\cal A}$ is called the spectral function.
760:
761:
762:
763:
764:
765:
766:
767:
768:
769:
770:
771:
772: \vskip 0.1in
773:
774:
775: \subsection{Lagrange density}
776: \label{Lagrange density}
777:
778: We shall now consider the dynamics of fermionic and scalar particles
779: in the presence of a scalar condensate which may give space-time dependent
780: masses to both types of particles, and we shall assume
781: that scalars and fermions couple {\it via} a Yukawa interaction term.
782: The tree-level action is then
783: %
784: \begin{eqnarray}
785: I[\phi,\psi] &=& \int_{\cal C} d^4 u \, {\cal L}
786: \label{action:tree-level}
787: \\
788: {\cal L} &=&
789: i \bar{\psi} \deldag \psi
790: - \bar{\psi}_L m \psi_R - \bar{\psi}_R m^* \psi_L
791: + \left(\del_\mu\phi \right)^\dagger\left( \del^\mu\phi\right)
792: - \phi^\dagger M^2\phi
793: + {\cal L}_{\rm int}
794: \,,
795: \label{lagrangean}
796: \end{eqnarray}
797: %
798: where the Yukawa interaction Lagrangean reads
799: %
800: \begin{eqnarray}
801: {\cal L}_{\rm int}
802: &=& - \bar{\psi}_L y\phi \psi_R
803: - \bar{\psi}_R (y\phi)^\dagger \psi_L
804: \nonumber\\
805: &=& - \bar{\psi}\left( P_R \otimes y\phi + P_L \otimes (y\phi)^\dagger
806: \right) \psi
807: \,,
808: \label{Yukawa}
809: \end{eqnarray}
810: %
811: and we can rewrite the fermionic mass term as
812: %
813: \beq
814: - \bar{\psi}_L m \psi_R - \bar{\psi}_R m^* \psi_L
815: = - \bar{\psi} (\mathbbm{1}\otimes m_h + i\gamma^5 \otimes m_a) \psi
816: \,.
817: \label{lag_ferm_mass}
818: \eeq
819: %
820: Fermionic fields with definite chirality $\psi_{R,L} = P_{R,L}\psi$
821: are defined with the help of the chirality projection operator
822: $P_{R,L}=(\mathbbm{1}\pm\gamma^5)/2$, and
823: $\gamma^5 = i \gamma^0\gamma^1\gamma^2\gamma^3$.
824: Our analysis includes the possibility of particle flavor mixing, but for
825: notational simplicity we have suppressed scalar and fermionic flavor labels
826: (we shall explicitly state so when we consider the special case of the single
827: field dynamics without mixing.)
828: Thus, $\phi$ and $\psi$ are vectors in the scalar and fermionic flavor
829: space, respectively, and $y$ is a vector in the scalar
830: and a matrix in the fermionic flavor space.
831: Then $\otimes$ denotes a direct product of the spinor and fermionic flavor
832: spaces. $M$ and $m$ are complex matrices in the scalar and fermionic flavor
833: space, respectively, whose non-diagonal elements couple fields of
834: different flavors to each other.
835: %
836: Even though the Lagrangean~(\ref{lagrangean}) cannot fully describe the
837: physics of the Standard Model, since it does not contain gauge fields,
838: it contains many essential elements of the physics of CP-violation
839: and interactions of importance for baryogenesis.
840: Since the main purpose of this work
841: is to establish controlled calculational methods, the
842: Lagrangean~(\ref{lagrangean}) should be a good toy model for
843: studying transport, in particular for baryogenesis.
844:
845: \vskip 0.1in
846:
847: An important example of the scalar field condensate is the Higgs field
848: at a first order electroweak phase transition in variants of the standard
849: model. This phase transition proceeds via nucleation and growth of
850: the broken phase bubbles of a nonzero Higgs condensate,
851: which varies at the interface of the symmetric and the broken phases,
852: the so called bubble wall.
853: The phase transition bubbles quickly grow large in comparison with the
854: thickness of the phase transition front, so for studies of local transport
855: phenomena at the scale of the bubble walls one can to a good approximation
856: assume planar symmetry. We will explicitly make use of this symmetry in
857: later sections.
858: To allow for CP-violation we choose the fermion mass to be a complex matrix
859: %
860: \begin{equation}
861: m(u) = m_h(u) + i m_a(u) = |m(u)|\mbox{e}^{i\theta(u)}
862: \,,
863: \label{fermion_varying mass}
864: \end{equation}
865: %
866: where $m_h$ and $m_a$ denote the hermitean and antihermitean part of $m$,
867: respectively.
868: The last part of this equation is to be understood as an equation
869: for the components.
870: CP-violation can be mediated either through $m_a$ or
871: complex off-diagonal entries of $m_h$.
872: %
873: The scalar mass matrix is by construction hermitean,
874: ${M^2}^\dagger = M^2$.
875: Nevertheless, in the case of flavor mixing CP-violation can be
876: mediated through the off-diagonal elements of $M^2$, provided they
877: are complex.
878:
879:
880:
881:
882:
883:
884: \subsection{Two-particle-irreducible (2PI) effective action}
885: \label{Two-particle-irreducible (2PI) effective action}
886:
887: We shall now review how to derive the equation of motion
888: for the Green functions in the CTP-formalism. Note first that
889: the tree-level equation can be obtained in a straightforward manner
890: as follows. By varying the action~(\ref{action:tree-level})
891: %-\ref{lagrangean})
892: with respect to $\phi^\dagger$, one obtains the familiar Klein-Gordon
893: equation
894: %
895: \begin{equation}
896: \big( - \del^2_u - M^2(u) \big) \phi(u) = 0
897: \,.
898: \label{eom_scalar_field}
899: \end{equation}
900: %
901: After multiplying this from the left by $\phi^\dagger(v)$ and taking
902: the expectation value one gets the tree-level equation of
903: motion for the Wightman function~(\ref{Green_scalar_index}):
904: %
905: \begin{equation}
906: \big( - \del^2_u - M^2(u) \big) i\Delta^<(u,v) = 0
907: \,.
908: \label{scalar-eom:tree-level}
909: \end{equation}
910: %
911: By an analogous procedure one finds that the same equation holds
912: for $i\Delta^>$. The equations of motion for
913: $i\Delta^{t}$ and $i\Delta^{\bar t}$
914: are obtained by imposing the appropriate time ordering, such that
915: they acquire on the {\it r.h.s.} the $\delta$-function sources
916: $i\delta^4(u-v)$ and $-i\delta^4(u-v)$, respectively.
917: %
918: Similarly one finds that at tree-level the fermionic Wightman
919: functions~(\ref{Green_fermionic_index}) obey the Dirac equation
920: %
921: \begin{equation}
922: \left( i\deldag_u - m_h(u) -i\gamma^5m_a(u) \right) iS^{<,>}(u,v) = 0
923: ,
924: \label{fermionic-eom:tree-level}
925: \end{equation}
926: %
927: and the equations of motion for $iS^{t}$ and $iS^{\bar t}$
928: again acquire $i\delta^4(u-v)$ and $-i\delta^4(u-v)$
929: on the {\it r.h.s.}, respectively.
930:
931:
932: \vskip 0.1in
933:
934:
935: In order to include interactions into the equations of motion,
936: we use the 2PI effective action approach, which was
937: originally developed for equilibrium problems in condensed matter physics
938: (where it is better known as $\Phi$-derivable actions)
939: by Luttinger and Ward~\cite{LuttingerWard:1960}
940: and in relativistic field theory by
941: Cornwall, Jackiw and Tomboulis~\cite{CornwallJackiwTomboulis:1974} and
942: then adapted to non-equilibrium situations by Chou, Su, Hao and
943: Yu~\cite{ChouSuHaoYu:1985}, and by Calzetta and Hu~\cite{CalzettaHu:1986}.
944: In this approach one uses the two-particle irreducible (2PI) out-of-equilibrium
945: formulation for the CTP effective action $\Gamma$, which
946: is a functional not only of the expectation
947: value of the field $\varphi(u)\equiv \langle\Omega|\phi(u)|\Omega\rangle$,
948: but also of the two-point functions $i\Delta(u,v)$ and $iS(u,v)$
949: defined in~(\ref{Green_scalar_CTP}-\ref{Green_fermionic_CTP}).
950: The main advantage of this formalism compared to the more standard
951: 1PI approach is that, at a given order in the loop expansion, varying
952: the 2PI action probes much more accurately the field configuration space.
953: Technically speaking, at a given order in the loop expansion
954: the 2PI effective action typically resums a much larger set of
955: diagrams than the corresponding 1PI effective action.
956: In absence of sources, the equations of motion are obtained simply by
957: extremising the effective action with respect to
958: $\varphi(u)$, $\Delta(u,v)$ and $S(u,v)$:
959: %
960: \begin{eqnarray}
961: \frac{\delta\Gamma[\varphi,\Delta,S]}{\delta\varphi} &=& 0
962: \nonumber\\
963: \frac{\delta\Gamma[\varphi,\Delta,S]}{\delta\Delta} &=& 0
964: \nonumber\\
965: \frac{\delta\Gamma[\varphi,\Delta,S]}{\delta S} &=& 0
966: \,.
967: \label{eom:2PI}
968: \end{eqnarray}
969: %
970: These equations are obtained by the variation of the 2PI effective
971: action, and hence they correspond to the Schwinger-Dyson equations,
972: with the self-energy approximated by the 1PI diagrams.
973: Since we are here not interested in modeling the dynamics of
974: the scalar field condensate, we shall not study the equation of motion
975: for $\varphi$.
976: For our purposes we consider $\varphi$ as given
977: and simply absorb the effects of $\varphi$ into space-time dependent
978: mass terms $m$ and $M^2$. The quantum dynamics of a self-interacting
979: scalar field in the presence of a condensate describing the inflaton decay
980: is considered in the framework of the 2PI effective action approach
981: for example in~\cite{BergesSerreau:2002,
982: AartsAhrensmeierBaierBergesSerreau:2002,CalzettaHu:2002},
983: while the equivalent problem for classical scalar fields is studied
984: in~\cite{KhlebnikovTkachev:1996,ProkopecRoos:1996}.
985: %
986: Various aspects of the out-of-equilibrium dynamics and thermalization
987: of quantum fields are considered
988: in~\cite{Berges:2002,BergesBorsanyiSerreau:2002,AartsBerges:2001,BergesCox:2000}.
989: %
990: Here we develop a gradient approximation for the two-point function,
991: starting with the 2PI effective action, which is appropriate for studying
992: the dynamics of fields in the presence of slowly varying backgrounds.
993:
994:
995: \vskip 0.1in
996:
997: %
998: \begin{figure}[tbp]
999: \centerline{\hspace{.in}
1000: \epsfig{file=A_pics/2pi-1loop.eps, width=3.4in,height=1.3in}
1001: }
1002: \vskip -0.1in
1003: \caption{\small
1004: The one-loop diagrams contributing to the 2PI effective
1005: action~(\ref{2PI_effective_action}).
1006: }
1007: \lbfig{figure:2pi-1loop}
1008: \end{figure}
1009: %
1010: The 2PI effective action corresponding to the classical
1011: action~(\ref{action:tree-level})
1012: can be written in the form~\cite{CornwallJackiwTomboulis:1974}
1013: %
1014: \begin{equation}
1015: \Gamma[\Delta,S] = i \Tr ({\Delta^{(0)}}^{-1}\!\Delta)
1016: - i \Tr ({S^{(0)}}^{-1}\!S)
1017: + i \Tr \ln \Delta^{-1}
1018: - i \Tr \ln S^{-1}
1019: + \Gamma_2[\Delta,S]
1020: \,,
1021: \label{2PI_effective_action}
1022: \end{equation}
1023: %
1024: where the minus signs in the fermionic terms are related to Pauli statistics.
1025: We use a condensed notation, with $\Tr$ denoting both integration
1026: over space-time and summation over spinor and flavor indices.
1027: The first two terms in~(\ref{2PI_effective_action}) are the classical
1028: (tree-level) actions,
1029: the next two terms correspond to the one-loop vacuum diagrams
1030: shown in figure~\ref{figure:2pi-1loop}, while the last term
1031: $\Gamma_2$ stands for the sum of all two-particle irreducible vacuum graphs.
1032: In Paper~II we illustrate how to compute $\Gamma_2$ in practice.
1033: The inverse free propagators ${\Delta^{(0)}}^{-1}$ and ${S^{(0)}}^{-1}$
1034: can be read off from the classical
1035: action~(\ref{action:tree-level}-\ref{lagrangean}) rewritten as
1036: %
1037: \beq
1038: I[\phi,\psi]
1039: = \int_{\cal C} d^4u \, d^4v \,
1040: \phi^\dagger(u) {\Delta^{(0)}}^{-1}(u,v) \phi(v)
1041: + \int_{\cal C} d^4u \, d^4v \,
1042: \bar\psi(u) {S^{(0)}}^{-1}(u,v) \psi(v)
1043: + \int_{\cal C} d^4 u {\cal L}_{\rm int}
1044: \,.
1045: \label{scalar_class_eff_action}
1046: \eeq
1047: %
1048: They satisfy
1049: equations~(\ref{scalar-eom:tree-level}-\ref{fermionic-eom:tree-level})
1050: and are given by
1051: %
1052: \begin{eqnarray}
1053: {\Delta^{(0)}}^{-1}(u,v) &=& \big(\!-\del^2_u-M^2(u)\big)
1054: \delta^4_{\cal C}(u-v)
1055: \nonumber\\
1056: {S^{(0)}}^{-1}(u,v) &=& \left( i\deldag_u - m_h - i\gamma^5 m_a \right)
1057: \delta_{\cal C}^4(u-v)
1058: \,.
1059: \label{inverse_free_props}
1060: \end{eqnarray}
1061: %
1062: We now take the functional derivatives of the effective
1063: action~(\ref{2PI_effective_action}) with respect to $\Delta$ and $S$
1064: to obtain
1065: %
1066: \begin{eqnarray}
1067: \frac{\delta\Gamma[\Delta,S]}{\delta \Delta(v,u)}
1068: &=& i {\Delta^{(0)}}^{-1}(u,v)
1069: - i \Delta^{-1}(u,v)
1070: + \frac{\delta\Gamma_2[\Delta,S]}{\delta \Delta(v,u)}
1071: = 0
1072: \nonumber\\
1073: \frac{\delta\Gamma[\Delta,S]}{\delta S(v,u)}
1074: &=& - i {S^{(0)}}^{-1}(u,v)
1075: + i S^{-1}(u,v)
1076: + \frac{\delta\Gamma_2[\Delta,S]}{\delta S(v,u)}
1077: = 0
1078: \,.
1079: \label{functional-derivative_of_Gamma}
1080: \end{eqnarray}
1081: %
1082: By making use of the definitions for the scalar and fermionic self-energies
1083: %
1084: \begin{eqnarray}
1085: \Pi(u,v) &\equiv& \;\;\, i \frac{\delta\Gamma_2[\Delta,S]}{\delta \Delta(v,u)}
1086: \label{Pi}
1087: \\
1088: \Sigma(u,v) &\equiv& - i \frac{\delta\Gamma_2[\Delta,S]}{\delta S(v,u)}
1089: \label{Sigma}
1090: \end{eqnarray}
1091: %
1092: and multiplying from the right by $\Delta$ and $-S$, respectively,
1093: we can recast~(\ref{functional-derivative_of_Gamma}) as the
1094: Schwinger-Dyson equations
1095: %
1096: \begin{eqnarray}
1097: \big(-\del^2_u - M^2(u) \big) i\Delta(u,v)
1098: &=& i\delta_{\cal C}^4(u-v)
1099: + \int_{\cal C} d^4w \, \Pi(u,w) i\Delta(w,v)
1100: \label{eom_contour:scalar}
1101: \\
1102: \big( i\deldag_u - m_h(u) - i\gamma^5 m_a(u) \big) iS(u,v)
1103: &=& i\delta_{\cal C}^4(u-v)
1104: + \int_{\cal C} d^4w \, \Sigma(u,w) iS(w,v)
1105: \,.
1106: \label{eom_contour:fermionic}
1107: \end{eqnarray}
1108: %
1109:
1110: \vskip 0.1in
1111:
1112: So far we have written everything in the complex contour notation.
1113: In the index notation the self-energies are
1114: %
1115: \begin{eqnarray}
1116: \Pi^{ab}(u,v)
1117: &=& \;\;\, iab \frac{\delta\Gamma_2[\Delta,S]}{\delta \Delta^{ba}(v,u)}
1118: \label{Pi:ab}
1119: \\
1120: \Sigma^{ab}(u,v)
1121: &=& - iab \frac{\delta\Gamma_2[\Delta,S]}{\delta S^{ba}(v,u)}
1122: \,,
1123: \label{Sigma:ab}
1124: \end{eqnarray}
1125: %
1126: and the equations of
1127: motion~(\ref{eom_contour:scalar}-\ref{eom_contour:fermionic}) read
1128: %
1129: \begin{eqnarray}
1130: \big(-\del^2_u - M^2(u) \big) i\Delta^{ab}(u,v)
1131: &=& ai\delta_{ab}\delta^4(u-v)
1132: + \sum_c c \int d^4w \, \Pi^{ac}(u,w) i\Delta^{cb}(w,v)
1133: \label{eom_index:scalar}
1134: \\
1135: \big(i\deldag_u - m_h(u) - i\gamma^5 m_a(u) \big) iS^{ab}(u,v)
1136: &=& ai\delta_{ab}\delta^4(u-v)
1137: + \sum_c c \int d^4w \, \Sigma^{ac}(u,w) iS^{cb}(w,v)
1138: \,.
1139: \label{eom_index:fermionic}
1140: \end{eqnarray}
1141: %
1142: These are the fundamental quantum dynamical equations corresponding
1143: to the action~(\ref{action:tree-level}-\ref{lagrangean}).
1144: They look deceptively simple, since the full complexity of the problem
1145: is hidden in the self-energies
1146: %~(\ref{Pi:ab}-\ref{Sigma:ab})
1147: , which are
1148: very complicated functionals of the Green functions and in general not known
1149: completely.
1150:
1151:
1152: \vskip 0.1in
1153:
1154: From Eqs.~(\ref{Green_scalar_index}-\ref{Green_fermionic_index})
1155: and~(\ref{eom_index:scalar}-\ref{eom_index:fermionic}) we infer the
1156: bosonic equations
1157: %
1158: \begin{eqnarray}
1159: (-\partial^2-M^2) \Delta^{r,a}
1160: - \Pi^{r,a} \odot \Delta^{r,a} &=& \delta
1161: \label{Deltaeom-ra}
1162: \\
1163: (-\partial^2-M^2) \Delta^{<,>}
1164: - \Pi^r \odot \Delta^{<,>} &=& \Pi^{<,>} \odot \Delta^a
1165: \,,
1166: \label{Deltaeom-<>}
1167: \end{eqnarray}
1168: %
1169: and the fermionic equations
1170: %
1171: \begin{eqnarray}
1172: (i\partial\!\!\!/-m_h - i\gamma^5 m_a) S^{r,a}
1173: - \Sigma^{r,a}\odot S^{r,a} &=& \delta
1174: \label{Seom-ra}
1175: \\
1176: (i\partial\!\!\!/-m_h - i\gamma^5 m_a) S^{<,>}
1177: - \Sigma^r \odot S^{<,>} &=& \Sigma^{<,>}\odot S^a
1178: \,,
1179: \label{Seom-<>}
1180: \end{eqnarray}
1181: %
1182: where $\odot$ stands for integration over the intermediate variable.
1183: %
1184: We use the notation
1185: $\Pi^t \equiv \Pi^{++}$, {\it etc.}
1186: ({\it cf.} Eqs.~(\ref{Green_scalar_index}-\ref{Green_fermionic_index}))
1187: and have defined retarded and advanced self-energies
1188: in analogy with~(\ref{G-ra}).
1189: %
1190: As we have already pointed out, not all Green functions are independent:
1191: $G^r$ and $G^a$ can be expressed in terms of $G^<$ and $G^>$
1192: ({\it cf.} Eqs.~(\ref{Gtt_by_G<>}-\ref{G-ra})), implying that
1193: the equations for the retarded and advanced propagators are redundant.
1194: It is not difficult to show that Eqs.~(\ref{Deltaeom-ra}-\ref{Seom-<>})
1195: are consistent provided the relations
1196: %
1197: \begin{eqnarray}
1198: \Pi^t(u,v) &=& \delta^4(u-v)\Pi^{\tt sg}(u)
1199: + \theta(u_0-v_0) \Pi^>(u,v)
1200: + \theta(v_0-u_0) \Pi^<(u,v)
1201: \nonumber
1202: \\
1203: \Pi^{\bar{t}}(u,v) &=& \delta^4(u-v)\Pi^{\tt sg}(u)
1204: + \theta(u_0-v_0) \Pi^<(u,v)
1205: + \theta(v_0-u_0) \Pi^>(u,v)
1206: \,,
1207: \label{Pit-Pi<>}
1208: \end{eqnarray}
1209: %
1210: as well as corresponding relations for the fermionic self-energies,
1211: are satisfied.
1212: %
1213: This should be the case for any reasonable approximation of the
1214: self-energies. The additional singular terms $\Pi^{\tt sg}(u)$ and
1215: $\Sigma^{\tt sg}(u)$ we have allowed for appear naturally in some
1216: theories and can be absorbed by the mass terms (mass renormalization).
1217: An example of a singular self-energy contribution is the one-loop
1218: tadpole of a scalar theory with quartic self-interaction. The one-loop
1219: expressions for the self-energies that we will derive in
1220: Paper~II
1221: %section~\ref{The two-loop 2PI effective action and the self-energies}
1222: indeed satisfy~(\ref{Pit-Pi<>}) and the corresponding fermionic equations.
1223:
1224:
1225: \vskip 0.1in
1226:
1227:
1228: \begin{figure}[tbp]
1229: \begin{center}
1230: \epsfig{file=A_pics/contour-ret.eps, height=0.8in,width=4.in}
1231: \end{center}
1232: \vskip -0.3in
1233: \lbfig{figure:retarded-contour}
1234: \caption{%
1235: \small
1236: Integration contour for the retarded
1237: propagators $\Delta^r$ and $S^r$ in~(\ref{Delta-ra:free}-\ref{S-ra:free}).
1238: }
1239: \end{figure}
1240: %
1241: %
1242: \begin{figure}[tbp]
1243: \begin{center}
1244: \epsfig{file=A_pics/contour-adv.eps, height=0.8in,width=4.in}
1245: \end{center}
1246: \vskip -0.3in
1247: \lbfig{figure:advanced-contour}
1248: \caption{%
1249: \small
1250: Integration contour for the advanced propagators
1251: $\Delta^a$ and $S^a$ in~(\ref{Delta-ra:free}-\ref{S-ra:free}).
1252: }
1253: \end{figure}
1254: %
1255: To clarify the physical meaning of
1256: Eqs.~(\ref{Deltaeom-ra}-\ref{Seom-<>}), we observe that
1257: the equations for the retarded and advanced
1258: propagators~(\ref{Deltaeom-ra}) and~(\ref{Seom-ra}) describe mostly
1259: the spectral properties of the system. Indeed, this can be seen for example
1260: %from the homogeneous (translationally invariant) solutions
1261: from the tree-level solutions, which in the limit of constant mass terms and
1262: no flavor mixing read
1263: %
1264: \begin{eqnarray}
1265: \Delta_{\tt 0}^{r,a}(u,v)
1266: &=& \int \frac{d^4k}{(2\pi)^4}{\rm e}^{-ik\cdot(u-v)}
1267: \frac{1}{k^2 - M^2 \pm i{\rm sign}(k_0)\epsilon}
1268: \label{Delta-ra:free}
1269: \\
1270: S^{r,a}_{\tt 0}(u,v) &=& \int \frac{d^4k}{(2\pi)^4}{\rm e}^{-ik\cdot(u-v)}
1271: \frac{k\!\!\!/ + m_h - i\gamma^5 m_a}
1272: {k^2 - m_h^2 - m_a^2 \pm i{\rm sign}(k_0)\epsilon}
1273: \,,
1274: \label{S-ra:free}
1275: \end{eqnarray}
1276: %
1277: where we use the infinitesimal pole shifts $\mp i\epsilon$
1278: ($\epsilon \rightarrow 0+$)
1279: to indicate the standard integration prescription for the retarded
1280: and advanced propagators, respectively, shown in
1281: figures~\ref{figure:retarded-contour}
1282: and~\ref{figure:advanced-contour}.
1283: The spectral functions~(\ref{G:ha}) are then
1284: %
1285: \begin{eqnarray}
1286: {\cal A}_{\phi\tt 0} &=& \int \frac{d^4k}{(2\pi)^4}{\rm e}^{-ik\cdot(u-v)}
1287: \pi {\rm sign}(k_0)\delta(k^2 - M^2)
1288: \label{Aphi:free}
1289: \\
1290: {\cal A}_{\psi\tt 0} &=& \int \frac{d^4k}{(2\pi)^4}{\rm e}^{-ik\cdot(u-v)}
1291: (k\!\!\!/ + m_h - i\gamma^5 m_a)
1292: \pi {\rm sign}(k_0)\delta(k^2 - m_h^2 - m_a^2)
1293: \,.
1294: \label{Apsi:free}
1295: \end{eqnarray}
1296: %
1297: The $\delta$-functions
1298: % in~(\ref{Aphi:free}-\ref{Apsi:free})
1299: indicate
1300: that the frequencies of the plasma excitations are constrained to lie on the
1301: mass shell, which for scalars is given by
1302: %
1303: \begin{equation}
1304: k_0 = \pm \omega_\phi
1305: \,,\qquad
1306: \omega_\phi = \sqrt{\vec k^2 + M^2}
1307: \label{mass-shell:free-scalars}
1308: \end{equation}
1309: %
1310: and for fermions by
1311: %
1312: \begin{equation}
1313: \quad
1314: k_0 = \pm \omega_0 \,,
1315: \qquad
1316: \omega_0 = \sqrt{\vec k^2 + m_h^2 + m_a^2}
1317: \,.
1318: \label{mass-shell:free-fermions}
1319: \end{equation}
1320: %
1321:
1322: \vskip 0.1in
1323:
1324: On the other hand, the equations of motion for the Wightman
1325: functions~(\ref{Deltaeom-<>}) and~(\ref{Seom-<>})
1326: describe mostly statistical (kinetic) properties of the system.
1327: This can be seen for example from considering the thermal equilibrium
1328: solutions to these equations, which are given in
1329: section~\ref{Equilibrium Green functions} below.
1330: %
1331: The kinetic equations~(\ref{Deltaeom-<>}) and~(\ref{Seom-<>})
1332: can be rewritten in the form of the Kadanoff-Baym (KB) equations
1333: %
1334: \begin{eqnarray}
1335: (-\partial^2-M^2) \Delta^{<,>}
1336: - \Pi_h\odot \Delta^{<,>}
1337: - \Pi^{<,>}\odot \Delta_h
1338: &=& {\cal C}_\phi
1339: \equiv \frac 12 (\Pi^> \odot \Delta^< - \Pi^< \odot \Delta^>)
1340: \label{Deltaeom-<>b}
1341: \\
1342: (i\partial\!\!\!/ - m_h - i\gamma^5 m_a)S^{<,>}
1343: - \Sigma_h\odot S^{<,>}
1344: - \Sigma^{<,>}\odot S_h &=& {\cal C}_\psi
1345: \equiv \frac 12 ( \Sigma^> \odot S^< - \Sigma^< \odot S^>)
1346: \,,
1347: \label{Seom-<>b}
1348: \end{eqnarray}
1349: %
1350: where ${\cal C}_\phi$ and ${\cal C}_\psi$ denote the scalar and fermionic
1351: collision terms, respectively, and the hermitean parts of the self-energies
1352: are defined analogously to those for the Wightman functions (\ref{G:ha}).
1353: %
1354: With reasonable approximations for the self-energies, the Kadanoff-Baym
1355: equations
1356: %~(\ref{Deltaeom-<>b}-\ref{Seom-<>b})
1357: are suited to study kinetics of out-of-equilibrium quantum systems.
1358: %
1359: The terms $\Pi_h\Delta^<$ and $\Sigma_hS^<$ on the {\it l.h.s.}
1360: of~(\ref{Deltaeom-<>b}-\ref{Seom-<>b}) represent the self-energy contributions,
1361: which are sometimes considered as a (nonlocal) contribution to the mass terms.
1362: We postpone the discussion of how inclusion of the self-energy may change
1363: the plasma dynamics and in particular the quasiparticle picture
1364: to a future work.
1365: %
1366: The collision terms ${\cal C}_\phi$ and ${\cal C}_\psi$,
1367: as we will see in Paper~II,
1368: contain the standard gain and loss terms responsible for equilibration,
1369: but they may also contain additional CP-violating sources.
1370: %
1371: Finally, the terms $\Pi^<\Delta_h$ and $\Sigma^<S_h$ induce broadening of
1372: the on-shell dispersion relations and may cause breakdown of the quasiparticle
1373: picture. We shall consider the role of these terms in some detail
1374: in section~\ref{Reduction to the on-shell limit}.
1375: %
1376: There we will need the width
1377: %
1378: \beq
1379: \Gamma_\phi = \frac{1}{2i} (\Pi^a-\Pi^r) = \frac i2 (\Pi^>-\Pi^<)
1380: \label{Gammapsiphi}
1381: \eeq
1382: %
1383: and the hermitean part of the scalar self-energy
1384: %
1385: \beq
1386: \Pi_h = \frac 12 (\Pi^r+\Pi^a)
1387: = \Pi^t - \frac 12 (\Pi^>+\Pi^<)
1388: \,,
1389: \label{DeltaS_h}
1390: \eeq
1391: %
1392: which satisfy the spectral relation ({\it cf.} Eq.~(\ref{G-h}))
1393: %
1394: \beq
1395: \Pi_h(u,v) =
1396: \frac 12 {\rm sign}(u^0-v^0)\,[\Pi^{>}(u,v)-\Pi^{<}(u,v)]
1397: = -i {\rm sign}(u^0-v^0)\, \Gamma_\phi(u,v)
1398: \,.
1399: \label{Sigma-h}
1400: \eeq
1401: %
1402: Analogous relations hold for the fermionic self-energies.
1403:
1404:
1405:
1406:
1407:
1408:
1409:
1410:
1411:
1412:
1413:
1414:
1415: \subsection{Wigner representation and gradient expansion}
1416: \label{Wigner representation and gradient expansion}
1417:
1418: In equilibrium the Green functions $\Delta(u,v)$ and $S(u,v)$
1419: depend only on the relative (microscopic) coordinate $r = u-v$.
1420: This dependence corresponds to the internal fluctuations
1421: that typically take place on microscopic scales.
1422: In non-equilibrium situations, however, $\Delta$ and $S$
1423: depend also on the average (macroscopic) coordinate $x = (u+v)/2$.
1424: This dependence describes the system's behavior on large, macroscopic scales.
1425: For example, if the system couples to an external field which varies in space
1426: and time, the Green functions will show a corresponding dependence
1427: on the average coordinate.
1428: Thick bubble walls at a first order electroweak phase transition
1429: represent precisely an example of such an external field.
1430:
1431: \vskip 0.1in
1432:
1433: In order to separate the microscopic scale fluctuations from the behavior
1434: on macroscopical scales one typically performs a Wigner transform, which
1435: is a Fourier transform with respect to the relative coordinate $r$.
1436: The Green function in the Wigner representation, called Wigner function, is
1437: %
1438: \begin{equation}
1439: G(k,x) = \int d^4r \, {\rm e}^{ik\cdot r} \, G( x+r/2 , x-r/2 )
1440: \,.
1441: \label{G_Wigner}
1442: \end{equation}
1443: %
1444: In the Wigner representation the hermiticity properties (\ref{S_herm_config})
1445: and~(\ref{Delta_herm_config}) become simply
1446: %
1447: \beqa
1448: \big( i\Delta^{<,>}(k,x) \big)^\dagger &=& i\Delta^{<,>}(k,x)
1449: \label{herm_scal_Wigner}
1450: \\
1451: \big( i\gamma^0S^{<,>}(k,x) \big)^\dagger &=& i\gamma^0S^{<,>}(k,x)
1452: \,.
1453: \label{herm_ferm_Wigner}
1454: \eeqa
1455: %
1456: In order to transform the equations of motion, we make use of
1457: the general relation
1458: %
1459: \begin{equation}
1460: \int d^4(u-v) \, {\rm e}^{ik\cdot(u-v)}
1461: \int d^4w A(u,w) B(w,v)
1462: = {\rm e}^{-i\diamond}\{A(k,x)\}\{B(k,x)\} \,,
1463: \label{convolution_Wigner}
1464: \end{equation}
1465: %
1466: where $x$ is the average coordinate and the diamond operator
1467: is defined by
1468: %
1469: \begin{equation}
1470: \diamond \left\{\cdot\right\} \left\{\cdot\right\}
1471: = \frac{1}{2} \big( \del^{(1)}\cdot\del^{(2)}_k
1472: - \del^{(1)}_k\cdot\del^{(2)}
1473: \big)\{\cdot\}\{\cdot\}
1474: \,.
1475: \label{diamond}
1476: \end{equation}
1477: %
1478: The superscripts $(1)$ and $(2)$ refer to the first and second argument,
1479: respectively, and $\del \equiv \del_x$.
1480: Since the nonlocal terms have the form of~(\ref{convolution_Wigner}),
1481: it is a simple exercise to show that
1482: the Kadanoff-Baym equations~(\ref{Deltaeom-<>b}-\ref{Seom-<>b}),
1483: when written in the Wigner representation, become
1484: %
1485: \begin{eqnarray}
1486: \Big( k^2 - \frac{1}{4} \del^2 + ik\cdot\del
1487: - M^2{\rm e}^{-\frac{i}{2}\stackrel{\leftarrow}{\del}\cdot\del_k}
1488: \Big)\Delta^{<,>}
1489: - {\rm e}^{-i\diamond}\{\Pi_h\}\{ \Delta^{<,>}\}
1490: - {\rm e}^{-i\diamond}\{\Pi^{<,>}\}\{\Delta_h\}
1491: = {\cal C}_\phi
1492: \label{Wigner-space:scalar_eom}
1493: \\
1494: \Big( \kdag
1495: + \frac i2 \deldag
1496: - m_h
1497: {\rm e}^{-\frac{i}{2}\stackrel{\leftarrow}{\del}\cdot\del_k}
1498: - i\gamma^5m_a
1499: {\rm e}^{-\frac{i}{2}\stackrel{\leftarrow}{\del}\cdot\del_k}
1500: \Big)S^{<,>}
1501: - {\rm e}^{-i\diamond}\{\Sigma_h\}\{ S^{<,>}\}
1502: - {\rm e}^{-i\diamond}\{\Sigma^{<,>}\}\{ S_h\}
1503: = {\cal C}_\psi
1504: \,,
1505: \label{Wigner-space:fermionic_eom}
1506: \end{eqnarray}
1507: %
1508: where the collision terms are
1509: %
1510: \begin{eqnarray}
1511: {\cal C}_\phi &=& \frac 12 {\rm e}^{-i\diamond}
1512: \big( \{\Pi^>\} \{\Delta^<\}
1513: - \{\Pi^<\} \{\Delta^>\}
1514: \big)
1515: \label{Cphi}
1516: \\
1517: {\cal C}_\psi &=& \frac 12 {\rm e}^{-i\diamond}
1518: \big( \{\Sigma^>\}\{S^<\}
1519: - \{\Sigma^<\}\{S^>\}
1520: \big)
1521: \,.
1522: \label{Cpsi}
1523: \end{eqnarray}
1524: %
1525: We shall analyze these equations in gradient expansion:
1526: we assume that the variation of the background field,
1527: and therefore also the variation of the Green functions and self-energies,
1528: with respect to the macroscopic coordinate $x$ is small when compared with
1529: the momenta of plasma excitations. Then we can perform an expansion
1530: in derivatives with respect to $x$.
1531: Formally, this criterion reads
1532: %
1533: \begin{equation}
1534: \del \ll k
1535: .
1536: \label{gradient-expansion:criterion}
1537: \end{equation}
1538: %
1539: That is, the background field is assumed to have a characteristic length scale
1540: that is large in comparison to the de~Broglie wavelength of the particles
1541: in the plasma.
1542:
1543: \vskip 0.1in
1544:
1545: This assumption of a slowly varying
1546: background field seems to be justified in our case~\cite{ThickWalls}:
1547: in the MSSM, for instance,
1548: the width of the bubble wall $L_w$ is roughly $10/T$, where $T$ is the
1549: temperature of the plasma. The typical momentum of a particle in the plasma
1550: is of the order $T$, so that the de~Broglie wavelength $l_{dB} \sim 1/T$
1551: is indeed small when compared to $L_w$.
1552: Since the expansion in powers of derivatives can be viewed as an expansion in
1553: powers of the Planck constant $\hbar$, the gradient expansion is equivalent to
1554: a semiclassical expansion. We expect that the leading order terms in the
1555: gradient expansion correspond to classical kinetic equations,
1556: while higher order derivatives represent quantum corrections.
1557:
1558:
1559:
1560:
1561:
1562:
1563:
1564:
1565: \subsection{Sum rules and Equilibrium Green functions}
1566: \label{Equilibrium Green functions}
1567:
1568:
1569:
1570: It is easily shown that, as a consequence of the equal-time
1571: commutation relation for scalars, the Wigner transform of their
1572: spectral function ${\cal A}_\phi$ (\ref{G:ha})
1573: obeys the spectral sum rule
1574: %
1575: \begin{equation}
1576: \int_{-\infty}^\infty \frac{dk_0}{\pi}k_0 {\cal A}_\phi(k,x) = 1
1577: \,,
1578: \label{sum_rule_scalars}
1579: \end{equation}
1580: %
1581: while the anticommutation rule for fermions leads to
1582: %
1583: \begin{equation}
1584: \int_{-\infty}^\infty \frac{dk_0}{\pi} \gamma^0 {\cal A}_\psi(k,x)
1585: = \mathbbm{1}
1586: \label{sum_rule_fermions}
1587: \end{equation}
1588: %
1589: for the fermionic spectral function.
1590: These sum rules are important, since they have to be imposed as
1591: consistency conditions on the solutions for the Wigner functions
1592: $G^<$ and $G^>$, or for $G^a$ and $G^r$, as indicated in~(\ref{G:ha}).
1593: %
1594: Furthermore, making use of these spectral sum rules
1595: and the Kubo-Martin-Schwinger (KMS) relations
1596: %
1597: \beqa
1598: \Delta^>_{\rm eq}(k) &=& \;\;\, {\rm e}^{\beta k_0} \Delta^<_{\rm eq}(k)
1599: \nonumber\\
1600: S^>_{\rm eq}(k) &=& -{\rm e}^{\beta k_0} S^<_{\rm eq}(k)
1601: \,,
1602: \label{KMS}
1603: \eeqa
1604: %
1605: where $\beta=1/T$ denotes the inverse temperature,
1606: one can unambiguously obtain the thermal equilibrium
1607: propagators~\cite{LeBellac:1996}.
1608: They are translationally invariant in space and time
1609: and hence, when written in the Fourier space, defined by the Fourier
1610: transform with respect to the relative coordinate $r = u-v$,
1611: they display a dependence on the momentum only:
1612: %
1613: \begin{equation}
1614: G_{\rm eq}(k) = \int d^4r \, {\rm e}^{ik\cdot r}G_{\rm eq}(r)
1615: \,.
1616: \label{Green_fermionic_Fourier}
1617: \end{equation}
1618: %
1619: The bosonic and fermionic equilibrium propagators are the solutions of
1620: the Klein-Gordon~(\ref{scalar-eom:tree-level}) and the Dirac
1621: equation~(\ref{fermionic-eom:tree-level}), respectively, which in the
1622: Fourier space read
1623: %
1624: \begin{eqnarray}
1625: \big( k^2 - M^2 \big) i\Delta_{\rm eq}^{<,>}(k) &=& 0
1626: \,,\qquad\qquad\quad\;\,
1627: \big( k^2 - M^2 \big) i\Delta_{\rm eq}^{r,a}(k) = i
1628: \nonumber
1629: \\
1630: \left( \kdag - m_h -i\gamma^5m_a \right) iS_{\rm eq}^{<,>}(k) &=& 0
1631: \,,\qquad
1632: \left( \kdag - m_h -i\gamma^5m_a \right) iS_{\rm eq}^{r,a}(k) = i
1633: \,,
1634: \label{kspace:fermionic-bosonic:tree-level}
1635: \end{eqnarray}
1636: %
1637: with the normalization~(\ref{sum_rule_scalars}-\ref{sum_rule_fermions})
1638: and the boundary conditions~(\ref{KMS}).
1639: The scalar equilibrium Green functions are
1640: %
1641: \begin{eqnarray}
1642: i\Delta_{\rm eq}^t(k) &=& \;\;\,
1643: \frac{i}{k^2-M^2+i{\rm sign}(k_0)\epsilon}
1644: + 2\pi \delta(k^2-M^2) {\rm sign}(k_0)
1645: n^\phi_{\rm eq}(k_0)
1646: \label{Green_scalar_eq_t}
1647: \\
1648: %
1649: i\Delta_{\rm eq}^{\bar{t}}(k) &=& -\frac{i}{k^2-M^2+i{\rm sign}(k_0)\epsilon}
1650: + 2\pi \delta(k^2-M^2) {\rm sign}(k_0)
1651: (1 + n^\phi_{\rm eq}(k_0))
1652: \label{Green_scalaw_eq_tbar}\\
1653: %
1654: i\Delta_{\rm eq}^<(k) &=& \;\; 2\pi \delta(k^2-M^2) {\rm sign}(k_0)
1655: n^\phi_{\rm eq}(k_0)
1656: \label{Green_scalar_eq_<}
1657: \\
1658: %
1659: i\Delta_{\rm eq}^>(k) &=& \;\; 2\pi \delta(k^2-M^2) {\rm sign}(k_0)
1660: (1+ n^\phi_{\rm eq}(k_0))
1661: \,,
1662: \label{Green_scalar_eq_>}
1663: \end{eqnarray}
1664: %
1665: where the Bose-Einstein distribution appears:
1666: %
1667: \begin{equation}
1668: n^\phi_{\rm eq}(k_0) = \frac{1}{{\rm e}^{\beta k_0} - 1}
1669: \,.
1670: \label{BoseEinstein}
1671: \end{equation}
1672: %
1673: Similarly, for the fermions we have
1674: %
1675: \begin{eqnarray}
1676: iS_{\rm eq}^t(k) &=& \;\;\,\frac{i(\kdag + m_h - i\gamma^5m_a)}
1677: {k^2 - |m|^2 + i{\rm sign}(k_0)\epsilon}
1678: - 2\pi (\kdag + m_h - i\gamma^5m_a)
1679: \delta(k^2-|m|^2)
1680: {\rm sign}(k_0) n_{\rm eq}(k_0)
1681: \label{Green_fermionic_eq_t}\\
1682: iS_{\rm eq}^{\bar{t}}(k)
1683: &=& -\frac{i(\kdag + m_h - i\gamma^5m_a)}
1684: {k^2 - |m|^2 + i{\rm sign}(k_0)\epsilon}
1685: + 2\pi (\kdag + m_h - i\gamma^5m_a)
1686: \delta(k^2 - |m|^2) {\rm sign}(k_0)
1687: (1- n_{\rm eq}(k_0))
1688: \qquad\,
1689: \label{Green_fermionic_eq_tbar}
1690: \\
1691: iS_{\rm eq}^<(k) &=& -2\pi (\kdag+ m_h - i\gamma^5m_a)
1692: \delta(k^2 - |m|^2) {\rm sign}(k_0)
1693: n_{\rm eq}(k_0)
1694: \label{Green_fermionic_eq_<}
1695: \\
1696: iS_{\rm eq}^>(k) &=& \;\;\, 2\pi (\kdag + m_h - i\gamma^5m_a)
1697: \delta(k^2 - |m|^2) {\rm sign}(k_0)
1698: (1 - n_{\rm eq}(k_0))
1699: \,,
1700: \label{Green_fermionic_eq_>}
1701: \end{eqnarray}
1702: %
1703: with the Fermi-Dirac distribution function
1704: %
1705: \begin{equation}
1706: n_{\rm eq}(k_0) = \frac{1}{{\rm e}^{\beta k_0} + 1}
1707: \label{FermiDirac}
1708: \end{equation}
1709: %
1710: and $ |m|^2 \equiv m_h^2 + m_a^2$.
1711: For the case of several mixing particle species, the equilibrium Green
1712: functions are diagonal matrices in flavor space, and the given relations hold
1713: for the diagonal elements.
1714:
1715:
1716:
1717:
1718:
1719:
1720:
1721:
1722:
1723:
1724:
1725:
1726:
1727: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1728: % %
1729: % %
1730: % OOOO OOOOO OOO O O OOOO OOOOO O OOO O O %
1731: % O O O O O O O O O O O O OO O %
1732: % OOOO OOOO O O O O O O O O O O O O %
1733: % O O O O O O O O O O O O O OO %
1734: % O O OOOOO OOO OOO OOOO O O OOO O O %
1735: % %
1736: % %
1737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1738:
1739: \cleardoublepage
1740: \section{Reduction to the on-shell limit}
1741: \label{Reduction to the on-shell limit}
1742:
1743: In this section~\footnote{Based on work with Kimmo Kainulainen.}
1744: we pay particular attention to the self-consistency
1745: of the on-shell limit as a controlled expansion in the coupling constants of
1746: the theory. To this aim we consider the role of the dissipative terms
1747: on the {\it l.h.s.}
1748: of the propagator equations~(\ref{Deltaeom-ra})
1749: and the Kadanoff-Baym
1750: equation~(\ref{Wigner-space:scalar_eom}).
1751: We shall see that including these terms into the retarded and
1752: advanced propagators results in a Breit-Wigner form for the spectral function.
1753: By making use of the spectral decomposition ansatz for the Wigner function
1754: with the Breit-Wigner form for the spectral function, we shall study
1755: under which conditions the kinetic equation reduces to the on-shell form.
1756:
1757: \vskip 0.1in
1758:
1759:
1760: For simplicity, we consider here the scalar case.
1761: It is quite plausible that similar techniques can be used to extend
1762: the analysis presented in this section to the case of fermions.
1763: Let us begin by rewriting equations~(\ref{Deltaeom-ra})
1764: and~(\ref{Wigner-space:scalar_eom}) more compactly,
1765: %
1766: \begin{eqnarray}
1767: {\rm e}^{-i\diamond}\{\Omega_\phi^2 \pm i\Gamma_\phi\}\{\Delta^{r,a}\} &=& 1
1768: \label{Deltaeom-ra2}
1769: \\
1770: {\rm e}^{-i\diamond}\{\Omega_\phi^2\}\{\Delta^{<,>}\}
1771: - {\rm e}^{-i\diamond}\{\Pi^{<,>}\}\{\Delta_h\}
1772: &=& {\cal C}_\phi
1773: \,,
1774: \label{Wigner-space:scalar:compact}
1775: %\\
1776: % {\rm e}^{-i\diamond}\{\Omega\!\!\!\!/\;\pm i \Gamma_\psi\}\{S^{r,a}\} &=& 1
1777: %\label{Seom-ra2}
1778: %\\
1779: % {\rm e}^{-i\diamond}\{\Omega\!\!\!\!/\;\}\{S^{<,>}\} \,
1780: % - {\rm e}^{-i\diamond}\{\Sigma^{<,>}\}\{S_h\} \,
1781: % &=& {\cal C}_\psi
1782: %,
1783: %\label{Wigner-space:fermionic:compact}
1784: \end{eqnarray}
1785: %
1786: where ${\cal C}_\phi$ is defined in Eq.~(\ref{Cphi}), and we defined
1787: %
1788: \begin{eqnarray}
1789: \Omega_\phi^2 &=& k^2 - M^2 - \Pi_h
1790: \,.
1791: \label{Omegas}
1792: \end{eqnarray}
1793: %
1794: Upon adding and subtracting the complex conjugate
1795: of~(\ref{Deltaeom-ra2}), we arrive at the following
1796: propagator equations for the case of a single scalar
1797: field up to second order in gradients
1798: %
1799: \begin{eqnarray}
1800: \cos\diamond \DBR{\Omega_\phi^2\pm i\Gamma_\phi}{\Delta^{r,a}} &=& 1
1801: \label{spec1}
1802: \\
1803: \sin\diamond \DBR{\Omega_\phi^2\pm i\Gamma_\phi}{\Delta^{r,a}} &=& 0
1804: ,
1805: \label{spec2}
1806: \end{eqnarray}
1807: %
1808: where we used the fact that $\Delta^{r,a}$ commute with
1809: $\Omega_\phi^2$ and $\Gamma_\phi$. This commutation property
1810: is special to the single field case however. Indeed, in the case of mixing
1811: scalars, $\Delta^{r,a}$ are matrices which in general
1812: do not commute with $\Omega_\phi^2$ and $\Gamma_\phi$.
1813:
1814: \vskip 0.1in
1815:
1816: Now observe that Eq.\ (\ref{spec2}) can be obtained from Eq.\
1817: (\ref{spec1}) by
1818: an application of the differential operator $\tan\diamond$, and hence it gives
1819: no new information. Because $\cos\diamond$ is an even function, we then see
1820: that the corrections to the propagator equation appear first time
1821: at the second order in gradients~\cite{Henning:1995}.
1822: Therefore, to the first order in gradients, we have simply
1823: %
1824: \begin{equation}
1825: \Delta^{r,a} = (\Omega_\phi^2 \pm i\Gamma_\phi)^{-1}
1826: ,
1827: \label{Delta:ra1}
1828: \end{equation}
1829: %
1830: or equivalently ({\it cf.}~(\ref{G:ha}))
1831: %
1832: \begin{eqnarray}
1833: \Delta_h &=& \frac{\Omega_\phi^2}
1834: {\Omega_\phi^4 + \Gamma_\phi^2}
1835: \label{Delta_h}
1836: \\
1837: {\cal A}_\phi &=& \frac{\Gamma_\phi}
1838: {\Omega_\phi^4 + \Gamma_\phi^2}.
1839: \label{A_phi}
1840: \end{eqnarray}
1841: %
1842: The kinetic and constraint equations can be obtained from the
1843: Kadanoff-Baym equations~(\ref{Wigner-space:scalar:compact})
1844: as the hermitean and antihermitean parts,
1845: %
1846: \begin{eqnarray}
1847: - \diamond\DBR{\Omega^2_\phi}{i\Delta^{<,>}}
1848: + \diamond\DBR{i\Pi^{<,>}}{\Delta_h}
1849: &=& \frac{1}{2} \big(\Pi^> \Delta^<- \Pi^< \Delta^>\big) .
1850: \label{ke:scalar}
1851: \\
1852: \Omega_\phi^2 \Delta^{<,>}
1853: + \frac{i}{2} \diamond\big(\DBR{\Pi^>}{\Delta^<}
1854: -\DBR{\Pi^<}{\Delta^>}\big)
1855: &=& \Pi^< \Delta_h
1856: ,
1857: \label{ce:scalar}
1858: \end{eqnarray}
1859: %
1860: where we truncated to second and first order in gradients, respectively.
1861: Because of the hermiticity property~(\ref{herm_scal_Wigner})
1862: of the scalar Wigner function, both kinetic and constraint equations must
1863: be satisfied simultaneously. The constraint equation selects the physical
1864: set of solutions among all solutions of the kinetic equation.
1865:
1866: \vskip 0.1in
1867:
1868: Note that the equations (\ref{Delta_h}-\ref{A_phi}) are only formal solutions
1869: in the sense that $\Pi_h$ and $\Gamma_\phi$ are functionals
1870: of $i\Delta^{<,>}$ and $iS^{<,>}$.
1871: Consequently, equations~(\ref{ke:scalar}-\ref{ce:scalar}) are
1872: complicated integro-differential equations.
1873: However, in many cases deviations from equilibrium are small,
1874: so that one can linearize them
1875: %~(\ref{ke:scalar}-\ref{ce:scalar})
1876: in deviation from equilibrium, such that $\Pi_h$ and $\Gamma_\phi$
1877: can be computed by using the equilibrium distribution
1878: functions~(\ref{Green_scalar_eq_t}-\ref{FermiDirac}).
1879:
1880: \vskip 0.1in
1881:
1882: Inspired by the spectral form of the equilibrium solutions,
1883: %~(\ref{Green_scalar_eq_t}-\ref{BoseEinstein}),
1884: we shall assume that close to equilibrium
1885: the following spectral decomposition holds
1886: %
1887: \begin{equation}
1888: i\Delta^< = 2{\cal A}_\phi n_\phi
1889: ,\qquad\quad
1890: i\Delta^> = 2{\cal A}_\phi (1+ n_\phi)
1891: \,,
1892: \label{spectral-decomposition:scalar}
1893: \end{equation}
1894: %
1895: where $n_\phi=n_\phi(k,x)$ is some unknown function representing
1896: a generalized distribution function on phase space.
1897: Similarly, for the self-energies, we can formally write
1898: %
1899: \begin{equation}
1900: i\Pi^< \equiv 2\Gamma_\phi n_{\Pi}
1901: ,\qquad
1902: i\Pi^> \equiv 2\Gamma_\phi (1+n_{\Pi}).
1903: \label{spectral-decomposition:Pi}
1904: \end{equation}
1905: %
1906: We have adopted a somewhat nonstandard definition for $\Gamma_\phi$;
1907: the more standard definition is obtained by the simple replacement
1908: $\Gamma_\phi\rightarrow 2 k_0\Gamma_\phi$.
1909: At the moment $n_{\Pi}$ is just an arbitrary variable replacing
1910: $\Pi^{<,>}$, but it will become closely related to the particle
1911: distribution function in the end. It is important to note that unlike
1912: $n_\phi$, it is {\em not} to be considered a free variable in the equations.
1913:
1914: Then using the definitions
1915: (\ref{spectral-decomposition:scalar}-\ref{spectral-decomposition:Pi})
1916: and the identity
1917: %
1918: \begin{equation}
1919: \diamond \DBR{f}{gh} = g \diamond\DBR{f}{h} + h \diamond\DBR{f}{g}
1920: ,
1921: \label{identity}
1922: \end{equation}
1923: %
1924: we rewrite Eqs.\ (\ref{ke:scalar}-\ref{ce:scalar}) as
1925: %
1926: \begin{eqnarray}
1927: {\cal A}_\phi\diamond\DBR{\Omega_\phi^2}{n_\phi}
1928: + \Gamma_\phi\diamond\DBR{\Delta_h}{n_\phi}
1929: &=& \Gamma_\phi{\cal A}_\phi \big(n_\phi-n_{\Pi}\big)
1930: \label{ke:scalar2}
1931: \\
1932: \Omega^2_\phi{\cal A}_\phi n_\phi
1933: + \diamond\DBR{\Gamma_\phi{\cal A}_\phi}{n_\phi}
1934: &=& \Gamma_\phi n_{\Pi} \Delta_h.
1935: \label{ce:scalar2}
1936: \end{eqnarray}
1937: %
1938: We can now see that taking $n_{\Pi}\rightarrow n_\phi$ solves
1939: the kinetic equation~(\ref{ke:scalar2}) to the zeroth order in gradients,
1940: establishing the promised connection between $n_{\Pi}$ and $n_\phi$.
1941: This also shows that, to the first order in gradients, it is consistent
1942: to replace $n_{\Pi}$ by the zeroth order solution $n_\phi$,
1943: whenever the $\diamond$ operator acts on $n_{\Pi}$. In fact, we
1944: already took this into account by replacing $n_{\Pi}$ by $n_\phi$ in the
1945: $\Gamma_\phi$-terms on the {\it l.h.s.}
1946: of Eqs.~(\ref{ke:scalar2}-\ref{ce:scalar2}).
1947:
1948: \vskip 0.1in
1949:
1950: We shall now see how to reduce these equations
1951: %~(\ref{ke:scalar2}-\ref{ce:scalar2})
1952: to an on-shell approximation, and in particular ask to what extent
1953: the on-shell limit is a good approximation to the plasma dynamics
1954: close to equilibrium.
1955: The equations
1956: %~(\ref{ce:scalar2}-\ref{ke:scalar2})
1957: can be simplified further when the width $\Gamma_\phi$ is small.
1958: While this is in general true in the weak coupling limit, due care
1959: must be exercised in making this approximation, because both $\Pi_h$ and
1960: $\Gamma_\phi$ are often controlled by the same coupling constants.
1961:
1962: \subsection{Propagator equation}
1963: \label{sec: Propagator equation}
1964:
1965: In the weak coupling limit $\Gamma_\phi \rightarrow 0$ and
1966: to first order in gradients, the spectral function~(\ref{A_phi})
1967: reduces to the singular spectral form,
1968: %
1969: \begin{equation}
1970: {\cal A}_\phi { \;\; \stackrel{\Gamma_\phi\rightarrow 0}{\longrightarrow} \;\;}
1971: {\cal A}_s \; = \; \pi\; {\rm sign}(k_0) \;
1972: \delta\big(\Omega_\phi^2\big)
1973: ,
1974: % + O(\Gamma_\phi)
1975: \label{As}
1976: \end{equation}
1977: %
1978: where the correction is of order $\Gamma_\phi$.
1979: This singular hypersurface defines the usual quasiparticle dispersion
1980: relation
1981: %
1982: \begin{equation}
1983: \Omega^2_\phi \equiv k_0^2 - \vec k^2 - M^2 - \Pi_h(k_0,\vec k,x) = 0
1984: ,
1985: \label{quasiDR2}
1986: \end{equation}
1987: %
1988: which defines the spectrum of the quasiparticle excitations of the system.
1989: This equation
1990: % (\ref{quasiDR2})
1991: has in general two distinct solutions
1992: $k_0 = \omega_k$ and $k_0 = -\bar \omega_k$, corresponding to particles and
1993: antiparticles, respectively. Thus the spectral function (\ref{As}) breaks
1994: up into two clearly separate contributions:
1995: %
1996: \begin{equation}
1997: {\cal A}_s =
1998: \frac{\pi }{2 \omega_k} Z_k\; \delta(k_0 - \omega_k)
1999: - \frac{\pi }{2 \bar \omega_k} \bar Z_k \;
2000: \delta(k_0 + \bar\omega_k),
2001: \label{As2}
2002: \end{equation}
2003: %
2004: where the wave function renormalization factors $Z_k$ and $\bar Z_k$ are
2005: %
2006: \begin{equation}
2007: 2\omega_k Z^{-1}_k =
2008: \left | {\partial \Omega_\phi^2}/{\partial k_0}
2009: \right|_{k_0 = \omega_k}
2010: ,\qquad
2011: 2\bar \omega_k \bar Z^{-1}_k =
2012: \left | {\partial \Omega_\phi^2}/{\partial k_0}
2013: \right|_{k_0=-\bar\omega_k}
2014: .
2015: \label{Zpm}
2016: \end{equation}
2017: %
2018:
2019: \vskip 0.1in
2020:
2021: The definition of $\Delta_h$ in~(\ref{Delta_h})
2022: becomes more problematic in the limit $\Gamma_\phi \rightarrow 0$ because
2023: of the appearance of the so called Landau ghost poles, when perturbative
2024: expressions are used for the self-energy $\Pi_h$~\cite{Henning:1995}.
2025: Therefore, when needed, we shall relate $\Delta_h$ to ${\cal A}_\phi$
2026: {\it via} equations~(\ref{spec1}-\ref{spec2}) before taking
2027: the zero width limit. Let us note however that the spectral representation
2028: %
2029: \begin{equation}
2030: \Delta_h(k,x) { \;\; \stackrel{\Gamma\rightarrow 0}{\longrightarrow} \;\;}
2031: {\rm Re} \int d k_0'
2032: \frac{{\cal A}_\phi(k_0',\vec k,x)}
2033: {k_0 - k_0' + i\epsilon}
2034: \label{PGR}
2035: \end{equation}
2036: %
2037: remains valid for $\Delta_h$ even in the on-shell limit \cite{Henning:1995}.
2038:
2039: \vskip 0.1in
2040:
2041: The full spectral function ${\cal A}_\phi$ must obey the
2042: sum-rule~(\ref{sum_rule_scalars}).
2043: The fact that the singular function ${\cal A}_s$~(\ref{As2}) fails this rule
2044: when interactions are included, tells us something important about the physical
2045: nature of the quasiparticle approximation. Indeed, one is here describing the
2046: spectrum of inherently collective plasma excitations by a simple single
2047: particle picture. The amount by which ${\cal A}_s$ fails the sum-rule can then
2048: be attributed to the effect of collective plasma excitations. This gives a
2049: quantitative estimate for the goodness, but also indicates the limits of
2050: applicability of the quasiparticle approximation.
2051:
2052: \vskip 0.1in
2053:
2054: Since both $\Pi_h$ and $\Gamma_\phi$ are controlled
2055: by the same coupling constants, they typically acquire contributions
2056: at the same order in the coupling constant, so that one cannot take the limit
2057: $\Gamma_\phi\rightarrow 0$ as an independent approximation.
2058: The theories containing tadpoles, such as the $\lambda \phi^4$-theory,
2059: seem at a first sight to be an exception. However, the tadpoles give rise
2060: to a singular non-dissipative contribution to the self-energy, which
2061: can be absorbed by the mass term, implying that they cannot be responsible
2062: for thermalization. This means that in order to study thermalization properly,
2063: one has to include the higher order nonlocal dissipative contributions,
2064: which typically contribute to both the self-energy and width at the same order
2065: in the coupling constant. Consequently, there is a danger that one might lose
2066: the consistency of the single particle picture. Nevertheless, as we shall now see,
2067: this is not the case. Namely, the two quantities appear in the spectral function
2068: in quite distinct ways such that the kinetic equation allows a consistent on-shell
2069: limit assuming equal accuracy in the computation of $\Pi_h$ and $\Gamma_\phi$.
2070:
2071: \subsection{Kinetic equation in the on-shell limit}
2072: \label{Kinetic equation in the on-shell limit}
2073:
2074: Our aim is now to study the conditions under which the kinetic and constraint
2075: equations~(\ref{ke:scalar2}-\ref{ce:scalar2}) reduce to the on-shell
2076: approximation for the effective plasma degrees of freedom found in the previous
2077: section. Bearing in mind the discussion concerning consistency of the
2078: quasiparticle limit, we start by using the
2079: expressions~(\ref{Delta_h}-\ref{A_phi}) with
2080: a finite width $\Gamma_\phi$. The term linear in $\Gamma_\phi$ on the
2081: {\it l.h.s.} of~(\ref{ke:scalar2}) is often simply dropped without
2082: a careful consideration~\cite{GreinerLeupold:1998}. However, making use
2083: of the identity $\Gamma_\phi\Delta_h = \Omega_\phi^2{\cal A}_\phi$,
2084: which is accurate to the leading order in gradients
2085: ({\it cf.} Eq.~(\ref{ce:scalar2})), and
2086: expressions~(\ref{Delta_h}-\ref{A_phi}), we can in fact combine it with the
2087: first term in the kinetic equation to obtain
2088: %
2089: \begin{equation}
2090: \frac{-2\Gamma_\phi^3}{(\Omega_\phi^4+\Gamma_\phi^2)^2} \;
2091: \diamond \DBR{\Omega_\phi^2}{n_\phi}
2092: - \frac{2\Gamma_\phi^2\Omega_\phi^2}{(\Omega_\phi^4+\Gamma_\phi^2)^2} \;
2093: \diamond \DBR{\Gamma_\phi}{n_\phi}
2094: = \Gamma_\phi{\cal A}_\phi \big(n_\phi-n_{\Pi}\big)
2095: \,.
2096: \label{ke:scalar3}
2097: \end{equation}
2098: %
2099: Note that in the weak coupling limit and to leading order in gradients
2100: %
2101: \begin{equation}
2102: \frac{-2\Gamma_\phi^3}{(\Omega_\phi^4+\Gamma_\phi^2)^2}
2103: \stackrel{\Gamma_\phi\rightarrow 0}{\longrightarrow}
2104: {\cal A}_s,
2105: \label{ke:scalar3b}
2106: \end{equation}
2107: %
2108: such that Eq.~(\ref{ke:scalar3}) can be rewritten as
2109: %
2110: %
2111: \begin{equation}
2112: {\cal A}_s\; \diamond \DBR{\Omega_\phi^2}{n_\phi}
2113: - 2\Gamma_\phi{\cal A}_s\Delta_h \;
2114: \diamond \DBR{\Gamma_\phi}{n_\phi}
2115: = \Gamma_\phi{\cal A}_\phi \big(n_\phi-n_{\Pi}\big)
2116: \,.
2117: \label{ke:scalar4}
2118: \end{equation}
2119: %
2120: Under the reasonable assumption that $\Gamma_\phi$ is a smooth function of
2121: the energy, we get from~(\ref{ke:scalar4}) a consistent expansion of
2122: the final result in powers of $\Gamma_\phi$ evaluated at the pole.
2123: In this expansion the contributions from the off-shellness cancel so that up to
2124: order ${\cal O}(\Gamma_\phi)$ the integrated equation involves only the leading
2125: on-shell excitations, defined by the singular spectral function~(\ref{As2}).
2126: In other words, the integrated kinetic equation exhibits the important property
2127: of {\em closure} onto the on-shell excitations in the weak coupling limit. We
2128: can then replace (\ref{ke:scalar4}) by the singular
2129: {\em on-shell kinetic equation}
2130: %
2131: \begin{equation}
2132: {\cal A}_s \diamond\DBR{\Omega_\phi^2}{n_\phi}
2133: = \Gamma_\phi{\cal A}_s \big(n_\phi-n_{\Pi}\big)
2134: .
2135: \label{ke:singular-onshell}
2136: \end{equation}
2137: %
2138: We again emphasize that it is actually consistent to include non-local
2139: loop contributions to $\Pi_h$ on the {\it l.h.s.} despite the fact that
2140: the singular single particle limit does not, strictly speaking, exist for
2141: the excitations. This is simply because the corrections due to off-shellness
2142: cancel up to ${\cal O}(\Gamma_\phi)$. This is fortunate, because applying
2143: the weak coupling limit in the strict sense to~(\ref{ke:singular-onshell})
2144: indeed requires either computing $\Pi_h$ and $\Gamma_\phi$ to the same order,
2145: or simply neglecting $\Gamma_\phi$, leading to a collisionless Boltzmann equation.
2146: A somewhat orthogonal approach has been taken by Leupold~\cite{Leupold:2000}
2147: in which, based on particle number conservation, the author advocates a
2148: modification in the relation between the Wigner function and the on-shell
2149: distribution function.
2150:
2151: \vskip 0.1in
2152:
2153: The final step is to show that the constraint equation~(\ref{ce:scalar2}) can
2154: be written as
2155: %
2156: \begin{equation}
2157: \Omega^2_\phi{\cal A}_s\big(n_\phi-n_{\Pi}\big)
2158: = -\diamond\DBR{\Gamma_\phi{\cal A}_s}{n_\phi}.
2159: \label{ce:scalar:singular}
2160: \end{equation}
2161: %
2162: In the limit $\Gamma_\phi \rightarrow 0$ the constraint equation becomes
2163: identically solved when the quasiparticle on-shell condition~(\ref{quasiDR2})
2164: is satisfied. In other
2165: words, the on-shell condition and constraint equations become degenerate.
2166: Note that we can here refer to the on-shell condition, not necessarily restricted
2167: to the singular approximation, but in the broader sense we derived the on-shell
2168: kinetic equation~(\ref{ke:singular-onshell}).
2169: Indeed, the constraint equation can always be
2170: satisfied by the on-shell solutions as given by Eq.~(\ref{ke:singular-onshell}),
2171: whereby it only gives rise to an integral constraint for the off-shell
2172: excitations.
2173: These however, do not belong to the set of dynamical degrees of freedom
2174: for the kinetic equation~(\ref{ke:singular-onshell}), which only contains
2175: the on-shell excitations as defined by the singular overall projection
2176: operator ${\cal A}_s$. In this sense the kinetic
2177: equation~(\ref{ke:singular-onshell}) and the constraint equation decouple.
2178: Of course, the off-shell excitations in the constraint
2179: equation~(\ref{ce:scalar:singular}) become relevant when the off-shell effects
2180: are included into the kinetic equation.
2181:
2182: We emphasise that the derivation of the on shell kinetic
2183: equation~(\ref{ke:singular-onshell}) applies to the special case
2184: of one scalar field in weak coupling limit and close to equilibrium.
2185: It would be of interest to extend the analysis to the case of mixing
2186: scalar and fermionic fields.
2187:
2188:
2189:
2190:
2191:
2192:
2193:
2194:
2195:
2196:
2197:
2198:
2199:
2200:
2201:
2202:
2203:
2204: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2205: % %
2206: % OOOO OOOO OOO O OOO OOOO %
2207: % O O O O O O O O O %
2208: % OOO O OOOOO O OOOOO OOOO %
2209: % O O O O O O O O O %
2210: % OOOO OOOO O O OOOOO O O O O %
2211: % %
2212: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2213:
2214: \cleardoublepage
2215: \section{Kinetics of scalars: tree-level analysis}
2216: \label{Kinetics of scalars: tree-level analysis}
2217:
2218:
2219: In this section we analyze the tree-level dynamics of the scalar sector
2220: of our theory with mixing and emphasise its ramifications.
2221: %
2222: The hermiticity property (\ref{herm_scal_Wigner}) of the scalar Wigner
2223: function implies that the hermitean and the antihermitean part of the
2224: equation of motion
2225: %
2226: \begin{equation}
2227: \left( k^2 - \frac14 \partial^2 + i k\cdot\partial
2228: - M^2(x)e^{-\frac i2\stackrel{\leftarrow}{\partial}\,\cdot\,\partial_k}
2229: \right) i\Delta^< = 0
2230: \label{scalars-eom}
2231: \end{equation}
2232: %
2233: ought to be satisfied simultaneously.
2234: A simple analysis reveals that the hermitean part corresponds to the
2235: {\it kinetic} equation, while the antihermitean part corresponds to the
2236: {\it constraint} equation. Roughly speaking, the kinetic equation describes
2237: the dynamics of quantum fields, while the constraint equation
2238: constrains the space of solutions of the kinetic
2239: equation~\cite{KainulainenProkopecSchmidtWeinstock:2001}.
2240: %
2241: In the simple case when $N=1$ it is immediately clear that the first quantum
2242: correction to the constraint equation, the real part of (\ref{scalars-eom}),
2243: is of second order and to the kinetic equation (imaginary part) of third order
2244: in gradients or second order in $\hbar$.
2245: To extract the spectral information to second order in
2246: $\hbar$ is quite delicate since the constraint equation
2247: in~(\ref{scalars-eom}) contains derivatives.
2248: %
2249: The situation is more involved in the case of more than one mixing field.
2250: In this case it is convenient to rotate into the mass eigenbasis
2251: %
2252: \begin{equation}
2253: M_d^2 = U M^2U^\dagger ,
2254: \label{rotu}
2255: \end{equation}
2256: %
2257: where $U$ is the unitary matrix that diagonalizes $M^2$.
2258: In this propagating basis equation~(\ref{scalars-eom}) becomes
2259: %
2260: \begin{equation}
2261: \left( k^2 - \frac14 {\cal D}^2
2262: + i k\cdot{\cal D}
2263: - M_d^2e^{-\frac i2\stackrel{\leftarrow}{{\cal D}}\,\cdot\,\partial_k}
2264: \right) \Delta_d^< = 0,
2265: \label{scalars-eom-d}
2266: \end{equation}
2267: %
2268: where $\Delta^<_d \equiv U \Delta^< U^\dagger$ and
2269: the `covariant' derivative is defined as
2270: %
2271: \def\calD{{\cal{D}}}
2272: %
2273: \begin{equation}
2274: \calD_\mu = \partial_\mu -i \left[{\Xi}_\mu,\;\cdot\;\right],
2275: \qquad
2276: {\Xi}_\mu = iU\partial_\mu U^\dagger
2277: \,.
2278: \label{calD}
2279: \end{equation}
2280: %
2281: Making use of $(i\Delta_d^<)^\dagger= i\Delta_d^<$,
2282: $\calD_\mu^\dagger = \calD_\mu$, which implies
2283: $({\calD_\mu}i\Delta_d^<)^\dagger = i\Delta_d^<\overleftarrow{\cal D}_\mu
2284: = {\calD_\mu}i\Delta_d^<$,
2285: and $\hat M_{c,s}^2 i \Delta_d^<
2286: = \frac 12 \{\hat M_{c,s}^2, i \Delta_d^<\}
2287: + \frac 12 [\hat M_{c,s}^2, i \Delta_d^< ] $,
2288: we can identify the antihermitean part of~(\ref{scalars-eom-d})
2289: as the constraint equation
2290: %
2291: \begin{equation}
2292: \Big(k^2 - \frac14 {\cal D}^2\Big) i\Delta_d^<
2293: - \frac 12\left\{\hat M^2_c, i\Delta_d^< \right\}
2294: + \frac i2 \left[\hat M^2_s, i\Delta_d^< \right]
2295: = 0,
2296: \label{scalars-ce-d}
2297: \end{equation}
2298: %
2299: and the hermitean part is the kinetic equation
2300: %
2301: \begin{equation}
2302: k\cdot{\cal D}i\Delta_d^<
2303: + \frac 12 \left\{ \hat M^2_s, i\Delta_d^< \right\}
2304: + \frac i2 \left[ \hat M^2_c, i\Delta_d^< \right]
2305: = 0
2306: \,.
2307: \label{scalars-ke-d}
2308: \end{equation}
2309: %
2310: We defined
2311: %
2312: \begin{eqnarray}
2313: \hat M^2_c &=&
2314: M^2_d\cos\frac 12\stackrel{\leftarrow}{{\cal D}}\cdot\,\partial_k
2315: \nonumber\\
2316: \hat M^2_s &=&
2317: M^2_d\sin\frac 12\stackrel{\leftarrow}{{\cal D}}\cdot\,\partial_k
2318: \,,
2319: \label{McMs}
2320: \end{eqnarray}
2321: %
2322: and $[\,\cdot\,,\,\cdot] $, $\{\,\cdot\,,\,\cdot\}$ denote a commutator
2323: and an anticommutator, respectively.
2324: The constraint~(\ref{scalars-ce-d}) and kinetic equation~(\ref{scalars-ke-d})
2325: are formally still exact. Since we are interested in order $\hbar$
2326: correction to the classical approximation, it is good to keep in mind that one
2327: can always restore $\hbar$ dependencies by the simple replacements
2328: $\partial \rightarrow \hbar \partial$ and $i\Delta_d^< \rightarrow
2329: \hbar^{-1} i \Delta_d^<$. We work to order $\hbar$ accuracy,
2330: so it suffices to truncate the constraint equation~(\ref{scalars-ce-d})
2331: to first order in gradients, and the kinetic equation~(\ref{scalars-ke-d})
2332: to second order in gradients:
2333: %
2334: \begin{eqnarray}
2335: k^2 i\Delta_d^<
2336: - \frac 12\left\{M^2_d, i\Delta_d^< \right\}
2337: + \frac i4 \left[{\cal D}M^2_d, \partial_k i\Delta_d^< \right]
2338: &=& 0,
2339: \label{scalars-ce-d2}
2340: \\
2341: k\cdot{\cal D}\, i\Delta_d^<
2342: + \frac 14 \left\{ {\cal D} M^2_d, \partial_k i\Delta_d^< \right\}
2343: + \frac i2 \left[ M^2_d \Big(1
2344: - \frac 18(\stackrel{\leftarrow}{{\cal D}}\cdot\,\partial_k)^2 \Big),
2345: i\Delta_d^< \right]
2346: &=& 0
2347: \,.
2348: \label{scalars-ke-d2}
2349: \end{eqnarray}
2350: %
2351: We shall now consider these equations in more detail.
2352: The off-diagonal elements of $i\Delta^<_d$
2353: in both the constraint~(\ref{scalars-ce-d2}) and the kinetic
2354: equation~(\ref{scalars-ke-d2}) are sourced by the diagonal elements
2355: through the terms involving commutators, which are suppressed by at least
2356: $\hbar$ with respect to the diagonal
2357: elements~\footnote{Since $M_d^2$ is a diagonal matrix, the commutator
2358: $\frac i2 \left[ M^2_d, i\Delta_d^< \right]$ in~(\ref{scalars-ke-d2}),
2359: which is formally of order $\hbar^{-1}$,
2360: contributes to the off-diagonal equations {\it via} terms that contain
2361: only the off-diagonal elements.}.
2362: On the other hand,
2363: the off-diagonals source the diagonal equations through terms that
2364: are of the same order as the diagonals.
2365: For the CP-violating diagonal densities that are suppressed at least by
2366: one power of $\hbar$, this immediately implies that the CP-violating
2367: off-diagonals are at least of the order $\hbar^2$, and thus cannot source
2368: the diagonal densities at order $\hbar$. By a similar argument,
2369: the second order term in the commutator in~(\ref{scalars-ke-d2}) can be
2370: dropped, since it can induce effects only at order $\hbar^2$.
2371:
2372: This analysis is however incomplete for the following reason.
2373: When the rotation matrices
2374: ${\Xi}_\mu = iU\partial_\mu U^\dagger$ in~(\ref{calD})
2375: contain CP-violation, the off-diagonal elements can in principle
2376: be CP-violating already at order $\hbar$, and hence source the diagonals
2377: at the same order. In order to get more insight into the role of
2378: the off-diagonals, we now analyze the case of two mixing scalars.
2379: %
2380: For notational simplicity we omit the index $d$ for diagonal in the following.
2381: All quantities have to be taken in the rotated basis,
2382: where the mass is diagonal.
2383: %
2384: The constraint equations~(\ref{scalars-ce-d2}), when written in components,
2385: are
2386: %
2387: \beqa
2388: \Big(k^2 - M_{\tt ii}^2 \Big)i\Delta_{\tt ii}^<
2389: \mp \frac 14 \delta(M^2)
2390: \Big( \Xi_{\tt 12}\cdot \del_{k}i\Delta_{\tt 21}^<
2391: + \Xi_{\tt 21}\cdot \del_{k}i\Delta_{\tt 12}^<
2392: \Big)
2393: &=& 0,
2394: \label{scalars-ce-11-22}
2395: \\
2396: \Big( k^2
2397: - \bar M^2
2398: + \frac i4 (\del\delta(M^2)) \cdot \del_k
2399: \Big)
2400: i\Delta_{\tt 12}^<
2401: + \frac 14 \delta(M^2) \Xi_{\tt 12} \cdot \del_k \delta(i\Delta^<)
2402: &=& 0
2403: \,,
2404: \label{scalars-ce-12}
2405: \eeqa
2406: %
2407: where
2408: $\bar M^2 \equiv {\rm Tr}(M^2)/2 = (M_{\tt 11}^2+M_{\tt 22}^2)/2$,
2409: $\delta(M^2) \equiv M_{\tt 11}^2-M_{\tt 22}^2$,
2410: $\delta(i\Delta^<) \equiv i\Delta^<_{\tt 11}-i\Delta^<_{\tt 22}$,
2411: and the equation for $i\Delta_{\tt 21}^< = (i\Delta_{\tt 12}^<)^*$ is obtained
2412: by taking the complex conjugate of~(\ref{scalars-ce-12}).
2413: To order $\hbar$ the diagonal equations~(\ref{scalars-ce-11-22}) are solved
2414: by the spectral on-shell solution
2415: %
2416: \begin{eqnarray}
2417: i\Delta_{\tt ii}^<(k,x) &=&
2418: 2\pi \delta\Big(k^2 - M_{\tt ii}^2\Big) {\rm sign}(k_0)
2419: n^\phi_i(k,x)
2420: \nonumber\\
2421: &=& \frac{\pi}{\omega_{\phi i}}
2422: \left[
2423: \delta(k_0-\omega_{\phi i})
2424: -\delta(k_0+\omega_{\phi i})
2425: \right]n^\phi_i(k,x)
2426: \,,
2427: \label{ce-diag-solution}
2428: \end{eqnarray}
2429: %
2430: where $ \omega_{\phi i}
2431: = [\vec k^{\,2}+M_{\tt ii}^2]^{1/2}$,
2432: and $n^\phi_i(k,x)$ represents the occupation number density
2433: on phase space, which in thermal equilibrium reduces to
2434: $n^\phi_i(k,x)\rightarrow n^\phi_{\rm eq} = 1/({\rm e}^{\beta k_0} -1)$.
2435: By making use of the sum rule~(\ref{sum_rule_scalars}) for the spectral function
2436: ${\cal A}_\phi = (i/2)(\Delta^>-\Delta^<)$, one can show that
2437: the other Wigner function has to be of the form
2438: %
2439: \begin{equation}
2440: i\Delta_{\tt ii}^>(k,x) =
2441: 2\pi \delta\Big(k^2 - M_{\tt ii}^2\Big) {\rm sign}(k_0)
2442: \Big(1 + n^\phi_i(k,x)\Big)
2443: \label{ce-diag-solution>}
2444: \end{equation}
2445: %
2446: with the same density $n^\phi_i$ as in~(\ref{ce-diag-solution}).
2447: These solutions are consistent provided the off-diagonals
2448: $i\Delta_{\tt 12}^<$ and $i\Delta_{\tt 21}^<$ are of order $\hbar$,
2449: which, as we shall argue,
2450: is a self-consistent assumption. The off-diagonal constraint
2451: equation~(\ref{scalars-ce-12}) is solved by
2452: %
2453: \beq
2454: i\Delta_{\tt 12}^<
2455: = n^\phi_{12} \, \delta(k^2-\bar M^2)
2456: - \frac 12 \Xi_{\tt 12} \cdot \del_k{\rm Tr}(i\Delta^<)
2457: - \frac{2}{\delta(M^2)} \, k \cdot \Xi_{\tt 12} \, \delta(i\Delta^<)
2458: \,,
2459: \label{Delta12:solution}
2460: \eeq
2461: %
2462: where $n^\phi_{12}=n^\phi_{12}(k,x)$ is a function that can be
2463: determined from the boundary conditions. Since we are interested
2464: in situations close to
2465: equilibrium in which the off-diagonals are driven away from zero primarily
2466: by the diagonals, we can set $n^\phi_{12}$ to zero for our purposes. Note that
2467: the solution~(\ref{Delta12:solution}) contains derivatives of the delta
2468: function, and hence it does not strictly speaking represent an on-shell form.
2469:
2470:
2471:
2472: \vskip 0.1in
2473:
2474: Consider now the kinetic equations~(\ref{scalars-ke-d2}), which
2475: in components read
2476: %
2477: \beqa
2478: \Big( k \cdot \del
2479: + \frac 12 (\del M_{\tt ii}^2) \cdot \del_k
2480: \Big)
2481: i\Delta_{\tt ii}^<
2482: &=& \pm i \, k \cdot \Big( \Xi_{\tt 12} i\Delta_{\tt 21}^<
2483: - \Xi_{\tt 21} i\Delta_{\tt 12}^<
2484: \Big)
2485: \nonumber\\
2486: &-& \frac i4 \delta(M^2)\,
2487: \Big( \Xi_{\tt 12} \cdot \del_k i\Delta_{\tt 21}^<
2488: - \Xi_{\tt 21} \cdot \del_k i\Delta_{\tt 12}^<
2489: \Big)
2490: \label{scalars-ke-11-22}
2491: \\
2492: \Big( k \cdot \del
2493: + \frac 12(\del\bar M^2) \cdot \del_k \!
2494: + \! \frac i2 \delta(M^2) \!
2495: - \! i k\cdot\delta(\Xi)
2496: \Big)
2497: i\Delta_{\tt 12}^< \!\!\!
2498: &=& \!\!\! - i k \cdot \Xi_{\tt 12} \delta(i\Delta^<)
2499: \!-\! \frac i4 \delta(M^2) \Xi_{\tt 12} \cdot \del_k {\rm Tr}(i\Delta^<)
2500: \,,
2501: \nonumber\\
2502: \label{scalars-ke-12}
2503: \eeqa
2504: %
2505: and the equation for $i\Delta_{\tt 21}^<$ is again
2506: obtained by taking the complex conjugate of~(\ref{scalars-ke-12}).
2507: To leading order in gradients the off-diagonal equation~(\ref{scalars-ke-12})
2508: is solved by
2509: %
2510: \begin{equation}
2511: i\Delta_{\tt 12}^<
2512: = - \frac 12\,\Xi_{\tt 12}\cdot\partial_k{\rm Tr}(i\Delta^<)
2513: - \frac{2}{\delta(M^2)}\, k\cdot \Xi_{\tt 12}\, \delta(i\Delta^<)
2514: \,,
2515: \label{scalars-ke-12-solution}
2516: \end{equation}
2517: %
2518: and similarly for $i\Delta_{\tt 21}^<$.
2519: Remarkably, this corresponds precisely
2520: to the constraint equation solution~(\ref{Delta12:solution}),
2521: provided one sets $n^\phi_{12} =0$, representing a very nontrivial
2522: consistency check of the kinetic theory for mixing particles.
2523: By making use of the technique of Green functions, in
2524: Appendix~\ref{Gradient expansion in the off-diagonal scalar kinetic equation}
2525: we show that the leading order result~(\ref{scalars-ke-12-solution})
2526: represents a valid approximation,
2527: provided the condition for the gradient approximation
2528: $k\cdot\partial \ll \delta(M^2)$ is satisfied, that is
2529: one is not near the degenerate mass limit, $\delta(M^2) = 0$.
2530:
2531: The example of mixing scalars represents indeed a nice illustration of
2532: the workings of the constraint and kinetic equations.
2533: Unlike in the one field case, in which the kinetic flow term can be
2534: obtained by acting with the bilinear $\diamond$ operator on the constraint
2535: equation, in the mixing case the situation is more complex.
2536: All of the solutions of the kinetic equation~(\ref{scalars-ke-12}) are
2537: simultaneously solutions of the constraint equation~(\ref{scalars-ce-12}).
2538: The converse is however {\it not} true: the constraint equation contains
2539: a larger set of solutions, which include the homogeneous
2540: solutions~(\ref{Delta12:solution})
2541: that lie on a different energy shell, which are in fact excluded by
2542: imposing the kinetic equation~(\ref{scalars-ke-12}).
2543: This can be shown as follows:
2544: the constraint equation~(\ref{scalars-ce-12}) can be obtained
2545: from the kinetic equation~(\ref{scalars-ke-12}) by
2546: multiplying it by $\big(k^2 - \Tr(M^2)/2\big)/\delta(M^2)$,
2547: followed by a partial integration. The constraint equation obtained this way
2548: has of course a larger set of solutions. The additional solutions are of
2549: the type: a function multiplying $\delta\big(k^2 - \Tr(M^2)/2\big)$.
2550:
2551:
2552: Finally, upon inserting~(\ref{scalars-ke-12-solution}) into the diagonal
2553: equation~(\ref{scalars-ke-11-22}), we get
2554: %
2555: \beqa
2556: \Big( k\cdot\partial
2557: + \frac 12 (\partial M_{\tt ii}^2)\cdot \partial_k
2558: \pm i \, k\cdot
2559: \big[\Xi_{\tt 12},\Xi_{\tt 21}\big]\cdot
2560: \partial_k\,
2561: \Big)
2562: i\Delta_{\tt ii}^<
2563: = 0
2564: \,.
2565: \label{scalars-ke-11-22b}
2566: \eeqa
2567: %
2568: It is remarkable that not only the off-diagonals have disappeared
2569: from~(\ref{scalars-ke-11-22b}), but also the diagonals decouple.
2570: The commutator
2571: $k\cdot \Big[\Xi_{\tt 12},\Xi_{\tt 21}\Big]\cdot \partial_k$
2572: does {\it not} vanish in general. However, for planar walls and
2573: in the wall frame,
2574: %
2575: \begin{equation}
2576: k\cdot \Big[\Xi_{\tt 12},\Xi_{\tt 21}\Big]
2577: \cdot \partial_k \;\rightarrow \;
2578: - k_z\Big( \Xi_{z\tt 12}\,\Xi_{z\tt 21}
2579: - \Xi_{z\tt 21}\,\Xi_{z\tt 12}
2580: \Big)
2581: \partial_{k_z}
2582: = 0
2583: \label{vasnishing-commutator}
2584: \end{equation}
2585: %
2586: vanishes identically, so that Eq.~(\ref{scalars-ke-11-22b}) reduces to
2587: %
2588: \begin{equation}
2589: \Big( k \cdot \del
2590: - \frac 12 (\del_z M_{\tt ii}^2)\del_{k_z}
2591: \Big)
2592: i\Delta_{\tt ii}^<
2593: = 0
2594: \,.
2595: \label{scalars-ke-11-22c}
2596: \end{equation}
2597: %
2598: This proves that, although a naive analysis of
2599: Eqs.~(\ref{scalars-ce-d2}-\ref{scalars-ke-d2}) indicated that there might be
2600: a CP-violating source in the kinetic equation at order $\hbar$,
2601: a more detailed look at the structure of these equations shows that
2602: for planar walls considered in the wall frame {\it no} such source appears.
2603: Equation~(\ref{scalars-ke-11-22c}) thus represents a self-consistent kinetic
2604: description of scalar fields accurate to order $\hbar$, such that the only
2605: source in the kinetic equation is the classical force,
2606: $\vec F_i = - \nabla \omega_{\phi i}
2607: = - \nabla M_{\tt ii}^2/ 2\omega_{\phi i}$.
2608: In addition, we have shown that to order $\hbar$
2609: the on-shell approximation for the diagonal occupation
2610: numbers~(\ref{ce-diag-solution}) with the classical dispersion relation
2611: $k_0 = \pm \omega_{\phi i}$ still holds.
2612: Even though the on-shell approximation fails for the off-diagonals, this
2613: is only relevant for the dynamics at higher orders in gradient expansion.
2614: Further, inclusion of collisions in the off-diagonal equations cannot
2615: change our conclusions concerning the source cancellation
2616: expressed in Eq.~(\ref{scalars-ke-11-22b}), even though it may introduce new
2617: collisional contributions. (For a detailed study of collisional sources
2618: we refer to Paper~II.)
2619: %section~\ref{Collision term}.)
2620: Needless to say, this analysis easily generalizes to
2621: the case of N mixing scalars. This completes the proof that was originally
2622: presented in Ref.~\cite{{KainulainenProkopecSchmidtWeinstock:2001}},
2623: which states that in the flow term of scalars there is no source
2624: for baryogenesis at order $\hbar$ in gradient expansion.
2625:
2626:
2627:
2628:
2629:
2630:
2631: \subsection{Boltzmann transport equation for CP-violating scalar densities}
2632: \label{Boltzmann transport equation for CP-violating scalar densities}
2633:
2634:
2635: For completeness, and to make a connection between the tree-level analysis
2636: presented here and the analysis of the collision term in Paper~II,
2637: %section~\ref{Scalar collision term},
2638: we now show how -- starting with the
2639: kinetic equation for the Wightman function~(\ref{scalars-ke-11-22c})
2640: -- one obtains a Boltzmann equation for the CP-violating scalar particle
2641: densities.
2642: All flavor matrices are to be taken in the basis
2643: where the mass is diagonal.
2644:
2645: In the beginning we review some of the basics of the C (charge) and CP
2646: (charge and parity) transformations of quantum scalar fields. Under C and P,
2647: a scalar field $\phi$ transforms as
2648: %
2649: \begin{eqnarray}
2650: \phi^c(u) \equiv
2651: {\cal C}\, \phi(u) \, {\cal C}^{-1}
2652: = \xi_\phi^* \phi^*(u)
2653: &,& \, {\cal C}\, \phi^\dagger(u) \, {\cal C}^{-1} \;
2654: = \xi_\phi \phi^T(u) \,
2655: \nonumber
2656: \\
2657: \phi^p(u) \equiv
2658: {\cal P}\, \phi(u) \, {\cal P}^{-1}
2659: = \eta_\phi^* \phi(\bar u)
2660: &,& {\cal P}\, \phi^\dagger(u) \, {\cal P}^{-1}
2661: = \eta_\phi \phi^\dagger(\bar u)
2662: \,,
2663: \label{phi:C+CP-transform}
2664: \end{eqnarray}
2665: %
2666: where $|\xi_\phi|=1$ and $|\eta_\phi|=1$ are global phases,
2667: and $\bar u^\mu = (u_0,-\vec{u})$ denotes the inversion of the spatial part
2668: of $u$. Note that our definition
2669: of charge conjugation includes an additional transposition with respect
2670: to the usual definition~\cite{ItzyksonZuber:1980}, which is required in
2671: the case of mixing scalar fields, when $\phi$ is a vector
2672: in flavor space. From~(\ref{phi:C+CP-transform}) it
2673: follows that the scalar Wightman functions transform as
2674: %
2675: \begin{eqnarray}
2676: i\Delta^<(u,v)
2677: &\stackrel{{\cal C}}{\longrightarrow}&
2678: i\Delta^>(v,u)^T
2679: % = i{\Delta^c}^<(u,v)
2680: \\
2681: i\Delta^<(u,v)
2682: &\stackrel{{\cal CP} }{\longrightarrow}&
2683: i\Delta^>(\bar v,\bar u)^T
2684: % = i{\Delta^{cp}}^<(u,v)
2685: \,.
2686: \label{phi_Wigner:C+CP}
2687: \end{eqnarray}
2688: %
2689: When written in the Wigner representation, the equivalent transformations are
2690: %
2691: \beqa
2692: i\Delta^<(k,x)
2693: &\stackrel{{\cal C}}{\longrightarrow}& i\Delta^>(-k,x)^T
2694: \equiv i{\Delta^c}^<(k,x)
2695: \label{scalar_wigner_C}
2696: \\
2697: %
2698: i\Delta^<(k, x)
2699: &\stackrel{{\cal CP}}{\longrightarrow}& i\Delta^>(-\bar{k}, \bar{x})^T
2700: \equiv i{\Delta^{cp}}^<(\bar{k}, \bar{x})
2701: \,.
2702: \label{scalar_wigner_CP}
2703: \eeqa
2704: %
2705: The $CP$-conjugate of the Wigner function $i\Delta^<(k,x)$ is related
2706: to physical objects at position $-\vec{x}$ with momentum $-\vec{k}$.
2707: Therefore we introduced additional inversions of the spatial parts of
2708: the position and momentum arguments in the definition of
2709: $i{\Delta^{cp}}^<(k,x)$,
2710: so that this object indeed describes particles at position $\vec{x}$ and with
2711: momentum $\vec{k}$.
2712: Note that this definition differs from what is usually found in textbooks.
2713: %
2714: In order to study the effects of the C and CP transformations on the distribution
2715: functions, we proceed analogously to~(\ref{ce-diag-solution})
2716: and define the spectral solutions
2717: %
2718: \begin{eqnarray}
2719: i{\Delta_{\tt ii}^{c/cp}}^<(k,x)
2720: &=& \frac{\pi}{\omega_{\phi i}}
2721: \left[
2722: \delta(k_0-\omega_{\phi i})
2723: -\delta(k_0+\omega_{\phi i})
2724: \right] n^{\phi c/cp}_i(k,x) \,
2725: \label{scalar_C_CP_onshell<}
2726: \\
2727: i{\Delta_{\tt ii}^{c/cp}}^>(k,x)
2728: &=& \frac{\pi}{\omega_{\phi i}}
2729: \left[
2730: \delta(k_0-\omega_{\phi i})
2731: -\delta(k_0+\omega_{\phi i})
2732: \right] \Big(1+n^{\phi c/cp}_i(k,x)\Big)
2733: \,.
2734: \label{scalar_C_CP_onshell>}
2735: \end{eqnarray}
2736: %
2737: Now upon inserting this and~(\ref{ce-diag-solution}-\ref{ce-diag-solution>})
2738: into (\ref{scalar_wigner_C}) and (\ref{scalar_wigner_CP}), we find
2739: %
2740: \beq
2741: n^{\phi c}_i \big(\omega_{\phi i}, \vec{k}, x \big)
2742: = n^{\phi cp}_i \big(\omega_{\phi i}, \vec{k}, x \big)
2743: = -\big[1 + n^\phi_i\big(-\omega_{\phi i}, -\vec{k} , x \big)
2744: \big]
2745: \,.
2746: \label{scalar_neg-energiy_cp}
2747: \eeq
2748: %
2749: Because of the additional inversions in the definition of ${i\Delta^{cp}}^<$,
2750: the two densities are identical.
2751:
2752: \vskip 0.1in
2753:
2754: We shall now consider the kinetic equation for scalars.
2755: First we include the collision term into Eq.~(\ref{scalars-ke-11-22c}),
2756: %
2757: \begin{equation}
2758: \Big( k\cdot\partial
2759: - \frac 12 (\partial_z M_{\tt ii}^2)\partial_{k_z}
2760: \Big)
2761: i\Delta_{\tt ii}^<(k,x)
2762: = \frac 12 \big[ {\cal C}_{\phi\tt ii}(k,x)
2763: + {\cal C}^\dagger_{\phi\tt ii}(k,x)
2764: \big]
2765: \,,
2766: \label{scalars-ke-11-22c+Col}
2767: \end{equation}
2768: %
2769: where ${\cal C}_{\phi\tt ii}$ denotes the diagonal elements of
2770: the collision term defined in Eq.~(\ref{Cphi}) after rotation into
2771: the basis where the mass is diagonal. Next, we insert the spectral
2772: on-shell solution~(\ref{ce-diag-solution}) and
2773: integrate over positive frequencies, to obtain
2774: %
2775: \begin{equation}
2776: \left( \del_t
2777: +\frac{\vec{k}\cdot{\nabla}}{\omega_{\phi i}}
2778: -\frac{(\partial_z{M_{\tt ii}^2}(z))}{2\omega_{\phi i}}\partial_{k_z}
2779: \right) f^\phi_{i+}(\vec k,x)
2780: = \frac{1}{2\pi}\int_0^\infty dk_0 \, \big[{\cal C}_{\phi\tt ii}(k,x)
2781: + {\cal C}^\dagger_{\phi\tt ii}(k,x)
2782: \big]
2783: \,,
2784: \label{scalar_pos_freq}
2785: \end{equation}
2786: %
2787: which is the kinetic equation for the scalar distribution function
2788: %
2789: \begin{equation}
2790: f^\phi_{i+}(\vec k,x)
2791: \equiv n^\phi_i\big(\omega_{\phi i}(\vec{k},x),\vec{k},x\big)
2792: \,.
2793: \label{scalar_distribution_function}
2794: \end{equation}
2795: %
2796: Since $i\Delta^>(k,x)$ and $i\Delta^<(k,x)$ satisfy identical equations,
2797: we can immediately write
2798: %
2799: \begin{equation}
2800: \Big(-k \cdot \partial
2801: + \frac 12 (\partial_z M_{\tt ii}^2(z))\partial_{k_z}\Big)
2802: i{\Delta_{\tt ii}^{cp}}^<(k,x)
2803: = \frac 12
2804: \big[{\cal C}^*_{\phi\tt ii}(-k, x)
2805: + {\cal C}^T_{\phi\tt ii}(-k, x)
2806: \big]
2807: \,,
2808: \label{scalars-ke-11-22c+Col2}
2809: \end{equation}
2810: %
2811: where we took account of~(\ref{scalar_wigner_CP}).
2812: We can omit the transposition in the collision term, because we are
2813: only interested in the diagonal elements.
2814: Upon integrating over positive frequencies this becomes
2815: %
2816: \begin{equation}
2817: - \left(\del_t
2818: +\frac{\vec{k}\cdot{\nabla}}{\omega_{\phi i}}
2819: -\frac{(\partial_{z}{M_{\tt ii}^2(z)})}
2820: {2\omega_{\phi i}}\partial_{k_z}
2821: \right) f^\phi_{i-}(\vec{k},x) =
2822: \frac{1}{2\pi}\int_0^\infty dk_0 \,
2823: \big[
2824: {\cal C}_{\phi\tt ii}(- k, x)
2825: + {\cal C}^\dagger_{\phi\tt ii}(- k, x)
2826: \big]
2827: \,.
2828: \label{scalar_KinEq_antiparticles}
2829: \end{equation}
2830: %
2831: This transport equation is the CP-conjugate of~(\ref{scalar_pos_freq}) for
2832: antiparticles, where we defined the antiparticle distribution function as
2833: %
2834: \beq
2835: f^\phi_{i-}(\vec k,x)
2836: \equiv n^{\phi cp}_i \big(\omega_{\phi i}, \vec k, x \big)
2837: = - \big[ 1 + n^\phi_i \big(-\omega_{\phi i}, -\vec{k}, x \big)
2838: \big]
2839: \,.
2840: \label{particle-antiparticle}
2841: \eeq
2842: %
2843: This then implies the following definition for the CP-violating particle
2844: density:
2845: %
2846: \begin{equation}
2847: \delta f^\phi_i = f^\phi_{i+} - f^\phi_{i-}
2848: \,.
2849: \label{CP-violating density}
2850: \end{equation}
2851: %
2852: The relevant kinetic equation for $\delta f^\phi_i$
2853: is obtained simply by summing
2854: (\ref{scalar_pos_freq}) and~(\ref{scalar_KinEq_antiparticles}),
2855: %
2856: \begin{eqnarray}
2857: \left( \del_t
2858: + \frac{\vec{k}\cdot{\nabla}}{\omega_{\phi i}}
2859: - \frac{(\partial_z{M_{\tt ii}^2}(z))}{2\omega_{\phi i}}
2860: \partial_{k_z}
2861: \right)
2862: \delta f^\phi_i(\vec{k},x)
2863: &=& \frac{1}{2\pi}\int_0^\infty dk_0 \,
2864: \big[
2865: {\cal C}_{\phi\tt ii}(k, x)
2866: + {\cal C}^\dagger_{\phi\tt ii}(k, x)
2867: \nonumber\\
2868: &&\qquad\qquad\;\,\,
2869: +\,{\cal C}_{\phi\tt ii}(- k, x)
2870: + {\cal C}^\dagger_{\phi\tt ii}(- k, x)
2871: \big]
2872: \,.
2873: \label{scalar_pos_freq_deltaCP}
2874: \end{eqnarray}
2875: %
2876: It should be stressed that the absence of a CP-violating force in the
2877: flow term at linear order in $\hbar$ is made explicit in this equation.
2878: Further, this equation
2879: %Eq.~(\ref{scalar_pos_freq_deltaCP})
2880: makes it transparent how
2881: to extract CP-violating contributions from the collision term.
2882:
2883:
2884:
2885:
2886:
2887:
2888:
2889:
2890:
2891:
2892:
2893:
2894: \subsection{Applications to the stop sector of the MSSM}
2895:
2896:
2897: This analysis is relevant for example for a calculation of the CP-violating
2898: force in the {\it stop} sector $\tilde q = (\tilde t_L,\tilde t_R)^T$
2899: of the MSSM~\cite{HuetNelson:1995+1996,Riotto:1995,Riotto:1998,
2900: ClineJoyceKainulainen:1998,CarenaQuirosRiottoViljaWagner:1997,
2901: ClineJoyceKainulainen:2000+2001},
2902: in which the mass matrix reads
2903: %
2904: \begin{equation}
2905: M_{\tilde q}^2 = \left(
2906: \begin{array}{cc} m_Q^2 & y( A^* H_2 + \mu H_1) \\
2907: y( A H_2 + \mu^* H_1) & m_U^2 \end{array} \right)
2908: ,
2909: \label{mass-squarks}
2910: \end{equation}
2911: %
2912: where $m_{Q}^2$ and $m_U^2$ denote the sum of the soft SUSY-breaking
2913: masses, including D-terms and $m_t^2 = y^2 H_2^2$. Our analysis
2914: immediately implies that, for squarks in the quasiparticle picture,
2915: there is no CP-violating correction to the dispersion relation at first order
2916: in gradients, and hence there is no CP-violating semiclassical force in
2917: the flow term of the kinetic equation at order $\hbar$. This is in contrast
2918: to what was found
2919: in Refs.~\cite{HuetNelson:1995+1996,Riotto:1995,Riotto:1998,
2920: CarenaQuirosRiottoViljaWagner:1997}.
2921:
2922: To investigate whether there is CP-violation in the off-diagonal sector
2923: of the theory, we note that the mass matrix
2924: %~(\ref{mass-squarks})
2925: is diagonalized by the unitary matrix
2926: %
2927: \begin{equation}
2928: U = \left(
2929: \begin{array}{cc} \cos\theta & -\sin\theta\, {\rm e}^{-i\sigma}\\
2930: \sin\theta\, {\rm e}^{i\sigma} & \cos\theta \end{array} \right)
2931: \,,
2932: \label{U}
2933: \end{equation}
2934: %
2935: where
2936: %
2937: \beq
2938: \tan 2\theta = \frac{2y|A^*H_2 + \mu H_1|}{m_Q^2-m_U^2}
2939: \,,\quad
2940: %\nonumber\\
2941: \tan \sigma = \frac{|A| \tan\beta \sin\alpha + |\mu| \sin\zeta}
2942: {|A| \tan\beta \cos\alpha + |\mu| \cos\zeta}
2943: \,,
2944: \label{tan2theta}
2945: \eeq
2946: %
2947: with $A^*H_2 + \mu H_1 = |A^*H_2 + \mu H_1| {\rm e}^{i\sigma}$,
2948: $A^* =|A|{\rm e}^{i\alpha}$, $\mu = |\mu| {\rm e}^{i\zeta}$
2949: and $\tan\beta = H_2/H_1$. From this we easily obtain
2950: %
2951: \begin{equation}
2952: \Xi_\mu \equiv iU\partial_\mu U^\dagger = \left(
2953: \begin{array}{cc} 0 & i{\rm e}^{-i\sigma}\\
2954: -i{\rm e}^{i\sigma} & 0 \end{array} \right) \partial_\mu\theta
2955: + \frac 12 \left(
2956: \begin{array}{cc} -(1 - \cos 2\theta) & \sin 2\theta\, {\rm e}^{-i\sigma}
2957: \\
2958: \sin 2\theta\, {\rm e}^{i\sigma} & 1 - \cos 2\theta \end{array} \right)
2959: \partial_\mu\sigma.
2960: \label{Xi}
2961: \end{equation}
2962: %
2963: %where
2964: %%
2965: %\begin{eqnarray}
2966: % \theta' &=& \frac{m_Q^2-m_U^2}
2967: % {(m_Q^2-m_U^2)^2 + 4y^2|A^*H_2 + \mu H_1|^2}
2968: % \, y|A^*H_2 + \mu H_1|'
2969: %\nonumber\\
2970: % \sigma' &=& \frac{ |A| \cos\theta\sin(\sigma+\alpha)}
2971: % {|A| \tan\beta \cos\alpha + |\mu| \cos\zeta}
2972: % (\tan \beta)'
2973: %.
2974: %\label{sigma-theta-prime}
2975: %\end{eqnarray}
2976: %%
2977: %
2978: Since the off-diagonal elements of $\Xi_\mu$
2979: % in~(\ref{Xi})
2980: are precisely the
2981: ones involved in the mixing of the diagonal and off-diagonal
2982: equations, $\Xi_\mu$ is CP-violating whenever $\sigma \neq 0$
2983: and $\tan\beta$ varies at the phase
2984: interface, which is in general realized at the phase transition
2985: in the MSSM by complex $A$ and $\mu$ parameters.
2986:
2987:
2988:
2989:
2990:
2991:
2992:
2993:
2994:
2995:
2996:
2997:
2998:
2999:
3000:
3001: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3002: % %
3003: % OOOOO OOOOO OOOO O O O OOO O O O OOOO %
3004: % O O O O OO OO O O O OO O O O %
3005: % OOO OOOO OOOO O O O O O O O O O O O %
3006: % O O O O O O O O O O OO O O %
3007: % O OOOOO O O O O O OOO O O O OOOO %
3008: % %
3009: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3010:
3011: \cleardoublepage
3012: \section{Kinetics of fermions: tree-level analysis}
3013: \label{Kinetics of fermions: tree-level analysis}
3014:
3015: In this section we consider the dynamics of fermions in the presence of
3016: scalar and pseudoscalar space-time dependent mass terms. Our analysis is
3017: of relevance, for example, for electroweak baryogenesis calculations,
3018: and establishes the nature of the propagating quasiparticle states
3019: in the electroweak plasma at the first order electroweak phase transition.
3020: %
3021: In supersymmetric models baryogenesis is typically mediated
3022: by charginos and neutralinos~\cite{HuetNelson:1995+1996}.
3023: Since they both mix at the tree level,
3024: it is very important to consider the more general case of $N$ mixing
3025: fermions~\cite{KainulainenProkopecSchmidtWeinstock:2001}.
3026: %
3027: We derive the dispersion relation and
3028: kinetic equation for mixing fermions accurate to order $\hbar$.
3029: We also construct the equilibrium Wigner function to order $\hbar$,
3030: which can then be used for baryogenesis source calculations.
3031: This section is not just an overview of our original
3032: work~\cite{KainulainenProkopecSchmidtWeinstock:2001,
3033: KainulainenProkopecSchmidtWeinstock:2002}, but it also contains
3034: new insights and results.
3035:
3036:
3037:
3038:
3039: \subsection{Spin conservation and the fermionic Wigner function}
3040:
3041:
3042: In section \ref{Wigner representation and gradient expansion}
3043: we have already derived the equation of motion~(\ref{Wigner-space:fermionic_eom})
3044: for the fermionic Wigner function.
3045: For the moment we neglect the interactions induced by ${\cal L}_{\rm int}$
3046: in the Lagrangean~(\ref{lagrangean}).
3047: %
3048: %%
3049: %\begin{eqnarray}
3050: % iS^<_{\alpha\beta,ij}(u,v)
3051: % &=& -\left<\Omega| \bar{\psi}_{\beta,j}(v)\psi_{\alpha,i}(u) |\Omega\right>,
3052: %\label{S_less}
3053: %\\
3054: %iS^>_{\alpha\beta,ij}(u,v)
3055: % &=& \quad\left<\Omega|\psi_{\alpha,i}(u)\bar{\psi}_{\beta,j}(v)|\Omega\right>,
3056: %\label{S_greater}
3057: %\end{eqnarray}
3058: %%
3059: %The Wightman functions~(\ref{S_less}-\ref{S_greater}) are
3060: %two-point functions which are in general matrices in both spinor
3061: %($\alpha,\beta$) and fermionic flavor space ($i,j$).
3062: %
3063: %
3064: The equations for $S^<$ and $S^>$ are identical, so we omit the label in the
3065: following, indicating that the equations hold for both of them.
3066: The equation of motion is
3067: %
3068: \begin{equation}
3069: {\cal D}S
3070: \equiv \left( \kdag
3071: + \frac i2 \deldag
3072: - (m_h+i\gamma^5 m_a)
3073: \mbox{e}^{-\frac{i}{2}\stackrel{\leftarrow}{\del}\!\cdot\,\del_{k}}
3074: \right) S = 0 ,
3075: \label{S_less_eom}
3076: \end{equation}
3077: %
3078: where we introduced the symbol ${\cal D}$ for the kinetic
3079: derivative operator.
3080: %
3081: For planar walls propagating in $z$-direction the mass
3082: terms in~(\ref{S_less_eom}) simplify to
3083: $m_{h,a}(z)
3084: \mbox{e}^{\frac{i}{2}\stackrel{\leftarrow}{\del_z}\!\,\del_{k_z}}$,
3085: when written in the wall frame.
3086:
3087:
3088: Equation~(\ref{S_less_eom}) is the tree-level fermionic master equation,
3089: which contains all the necessary information to study the nature of the
3090: propagating fermionic states in the presence of a space-time dependent mass
3091: term. We shall now show that, for planar walls, Eq.~(\ref{S_less_eom})
3092: contains a conserved quantity,
3093: corresponding to the spin pointing in the $z$-direction in
3094: the rest frame of the particle. This will then allow us to recast $S^{<,>}$
3095: in a form which is block diagonal in spin, such that different spin blocks
3096: completely decouple, provided, of course, they decouple at the boundaries.
3097: A very nontrivial consequence of this observation is that
3098: a {\it quasiparticle picture}, when formulated in terms of the spin states,
3099: survives to order $\hbar$ in gradient expansion.
3100:
3101:
3102:
3103:
3104:
3105:
3106: \subsubsection{The 1+1 dimensional ($\vec k_\| = 0$) frame}
3107:
3108:
3109: In order to establish what precisely is the conserved quantity, we consider
3110: a particle with arbitrary momentum $\vec{k}$ and perform a boost
3111: into the frame where the particle momentum parallel to the wall vanishes,
3112: while the momentum component perpendicular to the wall remains unaffected:
3113: %
3114: \beq
3115: k
3116: = \left(k_0, k_x, k_y, k_z \right)
3117: \rightarrow \tilde{k}
3118: = \left(\tilde{k}_0, 0, 0, k_z \right)
3119: \quad,\quad
3120: \tilde{k}_0 = \mbox{sign}(k_0) \sqrt{k_0^2-\vec{k}_\|^2}
3121: \,.
3122: \eeq
3123: %
3124: In this frame the particle moves only perpendicular to the wall, which
3125: is why we call it ``1+1 frame''. The necessary boost is characterized by
3126: %
3127: \begin{equation}
3128: \vec v_\| = \frac{\vec k_\|}{k_0},
3129: \qquad
3130: \gamma_\| = \frac{k_0}{\tilde k_0}
3131: \,.
3132: \label{Lorentz-transform}
3133: \end{equation}
3134: %
3135: In the spinor space it is represented by the operator
3136: %
3137: \begin{equation}
3138: L(k) = \frac{k_0 + \tilde{k}_0
3139: - \gamma^0\vec{\gamma}\cdot\vec{k}_{\|}}
3140: {[{2\tilde{k}_0(k_0+\tilde{k}_0)}]^{1/2}}
3141: \label{L-Lambda}
3142: \end{equation}
3143: %
3144: and its inverse $L^{-1}(k_0,\vec{k}) = L(k_0,-\vec{k})$, which act like
3145: %
3146: \begin{equation}
3147: L(k)\kdag L^{-1}(k) \,=\,
3148: \gamma^0\tilde k_0 - \gamma^3 k_z \,\equiv \, \tilde{\kdag}
3149: \,.
3150: \label{transfercond}
3151: \end{equation}
3152: %
3153: The mass terms in the equation of motion are transformed as follows:
3154: %
3155: \begin{equation}
3156: L(k) m_{h,a}(x)
3157: \mbox{e}^{-\frac i2 \stackrel{\leftarrow}{\del}
3158: \!\cdot\,
3159: \del_{k}}
3160: L^{-1}(k)
3161: = m_{h,a}(\tilde{x})
3162: \mbox{exp}\Big(-\frac i2 \stackrel{\leftarrow}{\tilde{\del}}
3163: \!\cdot\,
3164: \left( \del_{\tilde{k}}
3165: +\left[ L(k)\del_kL^{-1}(k), \,\cdot\,\right]
3166: \right)
3167: \Big)
3168: \,.
3169: \label{mass-terms:transform}
3170: \end{equation}
3171: %
3172: Since $L(k)$ is a function of $k_0$ and $\vec k_\|$, the commutator
3173: term
3174: %in~(\ref{mass-terms:transform})
3175: does in general not even vanish
3176: for planar walls in the frames where $m=m(t,z)$.
3177: The exception is the wall frame of a planar wall, where
3178: $m=m(z)$, such that
3179: $\left(\del m\right)\cdot\left(L(k)\partial_k L^{-1}(k)\right) = 0$,
3180: and the commutator in~(\ref{mass-terms:transform}) vanishes.
3181: For this reason we work from now on in the {\it wall frame} of a planar wall.
3182: In section~\ref{Plasma frame} below we remark on how to transform our results
3183: to other frames, in particular to the plasma frame.
3184: Upon transforming~(\ref{S_less_eom}), we then get
3185: %
3186: \begin{equation}
3187: \tilde {\cal D} \tilde S
3188: \equiv \left( \tilde\kdag
3189: + \frac{i}{2}\tilde\deldag
3190: - (m_h + i\gamma^5 m_a)
3191: \mbox{e}^{\frac i2 \stackrel{\leftarrow}{\del_z}\!\,\del_{k_z}}
3192: \right) \tilde S = 0
3193: \quad ({\tt 1+1\; dim.\; frame}),
3194: \label{S_less_eom-tilde}
3195: \end{equation}
3196: %
3197: where the boosted Wigner function is
3198: %
3199: \begin{equation}
3200: \tilde S(\tilde{k}) = L(k) S(k) L^{-1}(k)
3201: \,.
3202: \label{S_less-boost}
3203: \end{equation}
3204: %
3205: Since the bubble wall, which is responsible for any dependence of the
3206: Wigner function on the average coordinate $x^\mu$, is symmetric in the
3207: $x$-$y$ plane, we can assume that the Wigner function has the same symmetry:
3208: %
3209: \begin{equation}
3210: \tilde S = \tilde S(\tilde k,\tilde t,z)
3211: \,.
3212: \label{tildeS<:spin-conserving}
3213: \end{equation}
3214: %
3215: This form holds true for homogeneous boundary conditions, which is the case
3216: for most of the time during a first order phase transition,
3217: during which the bubbles are large, almost planar and far away from each other,
3218: when measured in the units of a typical diffusion scale.
3219: The derivative in~(\ref{S_less_eom-tilde}) then reduces like
3220: $\tilde\deldag \rightarrow \gamma^0\partial_{\,\tilde t}+\gamma^3\partial_{z}$,
3221: so that the spin operator in $z$-direction
3222: %
3223: \begin{equation}
3224: \tilde S_z = \gamma^0\gamma^3\gamma^5
3225: \label{tildeSz}
3226: \end{equation}
3227: %
3228: commutes with the kinetic operator
3229: % in~(\ref{S_less_eom-tilde})
3230: %
3231: \begin{equation}
3232: [\tilde {\cal D}, \tilde S_z] = 0,
3233: \label{tildeS_z-commutes}
3234: \end{equation}
3235: %
3236: and hence in this frame the spin in $z$-direction is a good quantum
3237: number. Assuming that the plasma is in thermal equilibrium before the
3238: phase transition takes place, the fermions are described by the
3239: equilibrium Wigner function~(\ref{Green_fermionic_eq_<}), which is
3240: diagonal in spin in the 1+1 frame. The interaction with the bubble
3241: wall introduces no spin mixing, so that we
3242: can write $\tilde S$ in the spin-block diagonal form
3243: %
3244: \beq
3245: \tilde S = \sum_{s=\pm 1} \tilde S_s,
3246: \qquad
3247: \tilde S_s \equiv \tilde P_s \tilde S \tilde P_s,
3248: \label{tildeS-block}
3249: \eeq
3250: %
3251: where $\tilde P_s$ denotes the spin
3252: projector:
3253: %
3254: \begin{equation}
3255: \tilde P_s = \frac 12 (\mathbbm{1}+s\tilde S_z),
3256: \qquad
3257: \tilde P_s \tilde P_{s'} = \delta _{ss'} \tilde P_s,
3258: \qquad s = \pm 1
3259: .
3260: \label{tildePs}
3261: \end{equation}
3262: %
3263: The spin diagonal Wigner function lives in the subalgebra of the
3264: the Clifford algebra of the Dirac matrices which is spanned
3265: by the matrices commuting with $\tilde P_s$. We can
3266: write down a decomposition of $\tilde S_s$ by using a suitable
3267: hermitean basis of the subalgebra,
3268: %
3269: \beq
3270: \tilde S_s(\tilde{k}) = i\tilde P_s(k)
3271: \left[ s\gamma^3\gamma^5 \tilde g^{s}_0(\tilde{k})
3272: - s\gamma^3 \tilde g^{s}_3(\tilde{k})
3273: + {\mathbbm 1} \tilde g^{s}_1(\tilde{k})
3274: - i\gamma^5 \tilde g^{s}_2(\tilde{k})
3275: \right]
3276: \,,
3277: \label{S<tilde_decomposition}
3278: \eeq
3279: %
3280: with $\tilde{g}^s_a$, $a=0,1,2,3$ being scalar functions.
3281: If we choose the following chiral representation of the Dirac matrices
3282: %
3283: \beqa
3284: \gamma^\mu = \left(\begin{array}{cc}
3285: \mathtt{0} & \sigma^\mu \\
3286: \bar\sigma^\mu & \mathtt{0}
3287: \end{array}\right)
3288: \,,
3289: \label{chiral_basis_gamma}
3290: \eeqa
3291: %
3292: where $\sigma^\mu = (\mathbbm{1},\sigma^i)$,
3293: $\bar\sigma^\mu = (\mathbbm{1},-\sigma^i)$,
3294: and $\sigma^i$, $i=1,2,3$ are the usual $2\times 2$ Pauli matrices,
3295: then the spin operator is
3296: %
3297: \beq
3298: \tilde{S}_z = \gamma^0\gamma^3\gamma^5 = \mathbbm{1} \otimes \sigma^3
3299: \eeq
3300: %
3301: and the spin block diagonality of the Wigner function becomes explicit:
3302: %
3303: \begin{equation}
3304: - i \gamma^0 \tilde S_s
3305: = \frac 12 \rho^a \otimes (\mathbbm{1}+s\sigma^3) \tilde g_a^s
3306: \,.
3307: \label{tildeS<s}
3308: \end{equation}
3309: %
3310:
3311:
3312:
3313:
3314:
3315:
3316:
3317:
3318: \subsubsection{The 3+1 dimensional (moving) frame}
3319: \label{The 3+1 dimensional (moving) frame}
3320:
3321: Knowing that in the 1+1 frame the spin in $z$-direction is
3322: conserved by the interaction with the bubble wall, we can construct
3323: the corresponding conserved quantity in the original frame by boosting
3324: the spin operator $\tilde{S}_z$ back to the original frame:
3325: %
3326: \beq
3327: S_z(k)
3328: \equiv L^{-1}(k) \,\tilde{\!S}_z L(k)
3329: = \frac{1}{\tilde k_0}
3330: \left(k_0 \gamma^0 - \vec k_\| \cdot \vec\gamma\right)\gamma^3\gamma^5
3331: \,.
3332: \label{Sz}
3333: \eeq
3334: %
3335: In order to clarify the physical meaning of this operator, we recall that
3336: the covariant form of the spin operator is given by the
3337: Pauli-Lubanski tensor
3338: %
3339: \begin{equation}
3340: S_{PL}(k, n) \equiv - \frac{1}{e_0}\kdag\ndag\gamma^5
3341: \,,
3342: \qquad e_0 \equiv (k^2)^{1/2}
3343: \,,
3344: \label{Pauli-Lubanski}
3345: \end{equation}
3346: %
3347: which measures the spin of a particle with momentum $k$ in the direction
3348: $\vec n$ ($n^2 = -1$, $n\cdot k = 0$).
3349: For simplicity we choose the normalization for $S_{PL}$ such that the
3350: eigenvalues are $\pm 1$.
3351: In literature one often finds the normalization which
3352: corresponds to the eigenvalues $\pm\hbar/2$.
3353: %
3354: In the on-shell limit, $e_0$ reduces to the particle's mass.
3355: In the rest frame the Pauli-Lubanski tensor becomes
3356: $S_{PL}\stackrel{\vec k \rightarrow 0}{\longrightarrow}\gamma^0\npdag\gamma^5$,
3357: where $n'$ is the spin vector in the rest frame, which is related to
3358: $n$ by the corresponding Lorentz boost.
3359: To measure the spin in $z$-direction in the 1+1-dimensional frame, we use
3360: %
3361: \begin{equation}
3362: \tilde k^\mu = (\tilde k_0,0,0,k_z),
3363: \qquad
3364: \tilde n^\mu = \frac{1}{e_0}(k_z,0,0,\tilde k_0),
3365: \label{tildek-tildes}
3366: \end{equation}
3367: %
3368: and we find, of course,
3369: $S_{PL}(\tilde k,\tilde n) = \gamma^0\gamma^3\gamma^5 = \tilde{S}_z$.
3370: The same operator measures also the spin in $z$-direction in the rest frame,
3371: because the spin operator~(\ref{Pauli-Lubanski}) is invariant under boosts
3372: as long as the direction of the spin vector is parallel to the momentum,
3373: $\vec n\, || \vec k$.
3374: Now, by setting $S_{PL}(k,n)$ equal to $S_z(k)$ in~(\ref{Sz}), we find that
3375: $S_z$ measures spin in the direction corresponding to
3376: %
3377: \begin{equation}
3378: n^\mu(k) = \frac{1}{\tilde k_0 e_0}\left(
3379: \begin{array}{c} k_0 k_z \\
3380: k_x k_z \\
3381: k_y k_z \\
3382: \tilde k_0^2
3383: \end{array}
3384: \right)
3385: \,.
3386: \label{spin-direction}
3387: \end{equation}
3388: %
3389: The same result is of course obtained by boosting $\tilde n^\mu$
3390: in~(\ref{tildek-tildes}) to the original (moving) frame $k^\mu$.
3391: %
3392: In the highly relativistic limit we have $\tilde k_0^2 \rightarrow k_z^2$,
3393: $k_0^2 \rightarrow {\vec k}^2$,
3394: and our special spin vector~(\ref{spin-direction}) becomes proportional
3395: to $\vec k$, such that the spin operator $S_z$ approaches the helicity operator,
3396: %
3397: \begin{eqnarray}
3398: \hat H (\vec k) &=& -\frac{1}{e_0}\kdag\hdag\gamma^5
3399: \nonumber\\
3400: &=& \hat{\vec k}\cdot \gamma^0 \vec\gamma\gamma^5,
3401: \qquad
3402: h^\mu = \frac{1}{e_0}\left(
3403: \begin{array}{c} |\vec k| \\
3404: k_0 \hat{\vec k}
3405: \end{array}
3406: \right)
3407: \,,
3408: \label{helicity-operator}
3409: \end{eqnarray}
3410: %
3411: as one would expect. As usually, the helicity operator measures spin in the
3412: direction of a particle's motion, $\hat{\vec k} = \vec k/|\vec k|$.
3413: As a consequence,
3414: for light particles with momenta of order the temperature,
3415: $k\sim T \gg m$, the spin states we consider here can be approximated
3416: by the helicity states, which are often used in literature for baryogenesis
3417: calculations.
3418:
3419:
3420:
3421: The commutation relation~(\ref{tildeS_z-commutes}) in the 1+1 frame
3422: now implies that by construction the spin operator $S_z(k)$ commutes
3423: with the kinetic differential operator ${\cal D}$ in~(\ref{S_less_eom}),
3424: provided the bubble wall is stationary and $x$-$y$-symmetric in the wall frame.
3425: We can slightly relax this condition: $S_z$ and ${\cal D}$ commute even
3426: for non-stationary and $x$-$y$-dependent bubble walls, as long as the
3427: dependence is of the form
3428: %
3429: \beq
3430: S_s = S_s(k,t-\vec v_\|\cdot\vec x_\|, z)
3431: \,,
3432: \label{S<-tildeS<}
3433: \eeq
3434: %
3435: such that $\nabla_\| = - \vec{v}_\|\partial_t$ in ${\cal D}$.
3436: So we can treat time dependent problems, which can be used to study how the
3437: system relaxes from some initial conditions to the stationary state,
3438: admittedly only for quite special forms of the initial conditions.
3439:
3440:
3441: As a consequence of the above discussion,
3442: even in the original moving frame, the problem splits
3443: into two non-mixing sectors labeled by the spin $s$:
3444: %
3445: \beq
3446: S = \sum_{s=\pm 1} S_s
3447: \,,
3448: \qquad
3449: S_s \equiv P_s S P_s
3450: \,,
3451: \label{S-block}
3452: \eeq
3453: %
3454: where $P_s(k) = (\mathbbm{1}+sS_z(k))/2$, $s=\pm 1$ is the spin projector.
3455: We can again write down a decomposition of $S_s$,
3456: %
3457: \begin{equation}
3458: S_s
3459: = iP_s \left[ s\gamma^3\gamma^5 g^{s}_0
3460: - s\gamma^3 g^{s}_3
3461: + {\mathbbm 1} g^{s}_1
3462: - i\gamma^5 g^{s}_2 \right]
3463: \,,
3464: \label{S<_decomposition2}
3465: \end{equation}
3466: %
3467: using a set of matrices that commute with $P_s(k)$.
3468: This is nothing else than the boosted 1+1 dimensional spin diagonal
3469: Wigner function:
3470: %
3471: \beq
3472: \tilde{S}_s(\tilde{k}, \tilde{x}) = L(k) S_s(k, x) L^{-1}(k)
3473: \,.
3474: \eeq
3475: %
3476: As a consequence of the hermiticity property~(\ref{herm_ferm_Wigner}),
3477: the scalar functions
3478: $g^s_a(k, x) = \tilde{g}^s_a(\tilde{k}, \tilde{x})$ are all real.
3479: Since the direction in which spin is measured now depends on the momentum,
3480: it is not any more possible to find a representation of the Dirac algebra
3481: in which the block diagonality of $S_s$ is explicitly displayed.
3482:
3483:
3484:
3485:
3486:
3487:
3488: \vskip4truemm
3489:
3490:
3491:
3492:
3493:
3494: \subsection{Constraint and kinetic equations}
3495: \label{Constraint and kinetic equations}
3496:
3497: We are now ready to study the tree-level fermion dynamics governed by
3498: Eq.~(\ref{S_less_eom}).
3499: In order to get the component equations, we take the block diagonal
3500: form~(\ref{S<_decomposition2}) for the Wigner function $S_s$,
3501: multiply~(\ref{S_less_eom}) by
3502: $P_s(k)\{1,s\gamma^3\gamma^5,-is\gamma^3,-\gamma^5\}$ and take the trace.
3503: The resulting equations are
3504: %
3505: \begin{eqnarray}
3506: 2i \hat{k}_0 g^s_0
3507: -2is \hat{k}_z g^s_3
3508: -2i \hat{m}_h g^s_1
3509: -2i \hat{m}_a g^s_2
3510: &=& \Tr P_s(k) {\mathbbm{1}} {\cal C_\psi}
3511: \label{ga0_eom}\\
3512: %
3513: 2i \hat{k}_0 g^s_1
3514: -2s \hat{k}_z g^s_2 \;
3515: -2i \hat{m}_h g^s_0
3516: \,+2 \hat{m}_a g^s_3 \;
3517: &=& \Tr P_s(k)s\gamma^3\gamma^5 {\cal C_\psi}
3518: \label{ga1_eom}\\
3519: %
3520: 2i \hat{k}_0 g^s_2
3521: \,+2s \hat{k}_z g^s_1 \;
3522: -2 \hat{m}_h g^s_3 \;
3523: -2i \hat{m}_a g^s_0
3524: &=& \Tr P_s(k)\left(-is\gamma^3\right) {\cal C_\psi}
3525: \label{ga2_eom}\\
3526: %
3527: 2i \hat{k}_0 g^s_3
3528: -2is \hat{k}_z g^s_0
3529: \,+2 \hat{m}_h g^s_2 \;
3530: \,-2\hat{m}_a g^s_1 \;
3531: &=& \Tr P_s(k) \left(-\gamma^5\right) {\cal C_\psi}
3532: \,,
3533: \label{ga3_eom}
3534: \end{eqnarray}
3535: %
3536: where we used the shorthand notations
3537: %
3538: \begin{equation}
3539: \hat{k}_0 = \tk + \frac i2
3540: \frac{k_0\del_t+\vec k_\|\cdot\nabla_{\|}}{\tk},
3541: \quad\quad
3542: \hat{k}_z = k_z - \frac i2 \del_z
3543: \,,
3544: \label{shorthand_k0k3}
3545: \end{equation}
3546: %
3547: and
3548: %
3549: \begin{equation}
3550: \hat{m}_{h,a} =
3551: m_{h,a}(z) \mbox{e}^{\frac i2 \stackrel{\leftarrow}{\del_z}\!\,\del_{k_z}}
3552: \,.
3553: \label{hat-m-ha}
3554: \end{equation}
3555: %
3556: For completeness, we have added the contributions from the fermionic
3557: collision term. One should keep in mind that the dependence of the
3558: Wigner function on the time and parallel coordinates is restricted
3559: by~(\ref{S<-tildeS<}).
3560:
3561:
3562: In the general case with fermionic mixing, each of the
3563: functions $g_a^s$ is a hermitean matrix in flavor space.
3564: The mixing is mediated through the off-diagonal elements of the
3565: mass terms~(\ref{hat-m-ha}) and thus appears already at the leading order
3566: in gradient expansion; hence it would be inappropriate
3567: to work in the interaction basis~(\ref{ga0_eom}-\ref{ga3_eom}) without
3568: incorporating the flavor off-diagonal elements. Since
3569: when integrated over the momenta
3570: $\int d^4 k $ and $\int d^4 k (k^\mu/k_0)$,
3571: equations~(\ref{ga0_eom}-\ref{ga3_eom}) yield
3572: fluid equations, the same conclusions hold for fluid equations.
3573: This is in discord with the strategy advocated in a recent
3574: work~\cite{CarenaQuirosSecoWagner:2002} on chargino-mediated
3575: baryogenesis in the Minimal Supersymmetric Standard Model,
3576: where it was argued that, when considering the dynamics of CP-violating
3577: sources, one should work in the weak interaction basis
3578: for charginos (without taking account of the off-diagonals).
3579:
3580: Rather than considering the full dynamics of mixing fermions, we shall
3581: now argue that, in order to properly capture the dynamics of CP-violating
3582: densities to order $\hbar$, it suffices to work in the spin and flavor
3583: diagonal bases, provided one transforms into the mass eigenbasis,
3584: in which $m$ is diagonal. Here we generalize the analysis
3585: of Ref.~\cite{KainulainenProkopecSchmidtWeinstock:2001} to
3586: 3+1 dimensions and include the space-time transients that accord
3587: with~(\ref{S<-tildeS<}). More importantly, we fill in a gap in our
3588: original derivation in~\cite{KainulainenProkopecSchmidtWeinstock:2001}.
3589:
3590:
3591:
3592: \subsubsection{Flavor diagonalization}
3593: \label{Flavor diagonalization}
3594:
3595: In order to get to the mass eigenbasis, we diagonalize the fermionic
3596: mass matrix $m$. Since $m$ is in general nonhermitean,
3597: the diagonalization is exacted by the biunitary transformation
3598: %
3599: \begin{equation}
3600: m_d = U m V^\dagger
3601: \,,
3602: \label{mass_diagonalization}
3603: \end{equation}
3604: %
3605: where $U$ and $V$ are the unitary matrices that diagonalize
3606: $mm^\dagger$ and $m^\dagger m$, respectively.
3607: To make the analysis more transparent, we make
3608: an explicit separation of the spinor and flavor spaces by
3609: using a direct product notation $\otimes$.
3610: The master equation~(\ref{S_less_eom}) then becomes
3611: %
3612: \begin{equation}
3613: \left( \kdag \otimes {\mathbbm{1}}
3614: + \frac i2 \deldag \otimes {\mathbbm{1}} - \hat \bm
3615: \right) S = {\cal C}_{\psi},
3616: \label{S_eom_mixing}
3617: \end{equation}
3618: %
3619: where we have defined the mass term as
3620: %
3621: \begin{eqnarray}
3622: \hat{\bm} &=& P_R \otimes \hat m + P_L \otimes \hat m^\dagger
3623: \nonumber\\
3624: &=& {\mathbbm 1}\otimes \hat m_h + i\gamma^5 \otimes \hat m_a
3625: \,.
3626: \label{bm}
3627: \end{eqnarray}
3628: %
3629: It is now a simple matter to see that the mass matrix $\bm$ is diagonalized
3630: by the unitary matrices
3631: %
3632: \begin{eqnarray}
3633: \bX &=& P_L \otimes V + P_R \otimes U
3634: = {\mathbbm 1}\otimes \frac 12 (V+U) - \gamma^5 \otimes \frac 12 (V-U)
3635: \nonumber\\
3636: \bY &=& P_L \otimes U + P_R \otimes V
3637: = {\mathbbm 1}\otimes \frac 12 (V+U) + \gamma^5 \otimes \frac 12 (V-U)
3638: \label{rotation_matrices}
3639: \end{eqnarray}
3640: %
3641: as follows:
3642: %
3643: \begin{equation}
3644: \bm_d = \bX \bm \bY^\dagger
3645: \,.
3646: \label{diagonalization-full-m}
3647: \end{equation}
3648: %
3649: The Wigner function then transforms as
3650: %
3651: \begin{equation}
3652: S_d = \bY S \bX^\dagger
3653: \,,
3654: \label{S_diagonalization}
3655: \end{equation}
3656: %
3657: which is in general not diagonal.
3658: %
3659: From here on we work in the frame where the bubble wall is at rest,
3660: so that we can make a spin-diagonal ansatz for the rotated Wigner
3661: function $S_{sd}$ which is analogous to~(\ref{S<_decomposition2}).
3662: Note that the spin projector $P_s$ commutes with the rotation matrices.
3663: Since the mass term $\hat\bm$ mixes spinor and flavor, the rotation
3664: matrices $\bY$ and $\bX$ do so as well. As a consequence, the component
3665: functions $g^s_{ad}$ of the rotated Wigner function are not just the flavor
3666: rotated components $g^s_a$ of $S_s$. Indeed,
3667: inserting~(\ref{S<_decomposition2})
3668: into~(\ref{S_diagonalization}) leads to the following relations
3669: %
3670: \begin{eqnarray}
3671: g^s_{0d} &=& \frac 12 \left[ V (g^s_{0}+g^s_{3}) V^\dagger
3672: + U (g^s_{0}-g^s_{3}) U^\dagger \right]
3673: \label{g0d_g0}\\
3674: g^s_{3d} &=& \frac 12 \left[ V (g^s_{3}+g^s_{0}) V^\dagger
3675: - U (g^s_{0}-g^s_{3}) U^\dagger \right]
3676: \label{g3d_g3}\\
3677: g^s_{1d} &=& \frac 12 \left[ V (g^s_{1}-ig^s_{2}) U^\dagger
3678: +U (g^s_{1}+ig^s_{2}) V^\dagger
3679: \right]
3680: \label{g1d_g1}\\
3681: g^s_{2d} &=& \frac 12 \left[ V (g^s_{2}+ig^s_{1}) U^\dagger
3682: + U (g^s_{2}-ig^s_{1}) V^\dagger
3683: \right]
3684: \,,
3685: \label{g2d_g2}
3686: \end{eqnarray}
3687: %
3688: that is $g_{0}^s$ mixes with $g_{3}^s$ and $g_{1}^s$ mixes with $g_{2}^s$.
3689: We now transform~(\ref{S_eom_mixing})
3690: by multiplying it from the left by $\bX$ and from the right by $\bX^\dagger$
3691: to obtain
3692: %
3693: \begin{eqnarray}
3694: &&
3695: \left( \kdag \otimes {\mathbbm 1} + \frac i2 \mDdag
3696: - \bm_d\,
3697: {\rm e}^{\frac i2 \stackrel{\!\!\leftarrow\,}{{\mathbf D_z}}\partial_{k_z}}
3698: \right) \,\, S_d
3699: = {\cal C}_{\psi d}
3700: \,,
3701: \label{S_eom_mixing_diagonalization}
3702: \end{eqnarray}
3703: %
3704: where ${\cal C}_{\psi d} = \bX {\cal C}_{\psi} \bX^\dagger$,
3705: and we defined a `covariant' derivative
3706: %
3707: \begin{equation}
3708: {\mathbf D}_\mu
3709: = {\mathbbm 1}\otimes{\mathbbm 1}\partial_\mu
3710: - i [{\mathbbm 1}\otimes\Sigma_\mu,\cdot]
3711: - i\{\gamma^5\otimes\Delta_\mu,\cdot\}
3712: \,,
3713: \label{mD}
3714: \end{equation}
3715: %
3716: where $\Sigma_\mu$ and $\Delta_\mu$ are given in terms of the rotation
3717: matrices $U$ and $V$ as follows:
3718: %
3719: \begin{eqnarray}
3720: \Delta_\mu &=& \frac i2 \left( V \del_\mu V^\dagger
3721: - U \del_\mu U^\dagger \right)
3722: \label{mixing_Delta}\\
3723: \Sigma_\mu &=& \frac i2 \left( V \del_\mu V^\dagger
3724: + U \del_\mu U^\dagger \right)
3725: \,.
3726: \label{mixing_Sigma}
3727: \end{eqnarray}
3728: %
3729: Note that the covariant derivative~(\ref{mD}) can be obtained from
3730: %
3731: \begin{equation}
3732: \partial_\mu\bm = \partial_\mu(\bX^\dagger\bm_d\bY)
3733: = \bX^\dagger({\mathbf D}_\mu\bm_d)\bY
3734: \label{mD2}
3735: \end{equation}
3736: %
3737: %and the relations~(\ref{rotation_matrices}).
3738: The higher order derivatives are obtained simply by iteration.
3739: Already from~(\ref{S_eom_mixing_diagonalization}-\ref{mD}) it is clear
3740: that the kinetic and constraint equations, when written in the mass eigenbasis,
3741: will contain extra commutator and anticommutator terms.
3742: The choice of the rotation matrices $U$ and $V$ is not unique. After an
3743: $x$-dependent phase redefinition $U \rightarrow wU$ and $V \rightarrow wV$,
3744: where $w$ is a diagonal matrix with eigenvalues of absolute value 1, $U$ and
3745: $V$ still diagonalize $m$. This freedom to redefine the rotation matrices
3746: was the source of some problems in finding the correct physical source in
3747: the WKB approach. We will find, however, that only the diagonal elements of
3748: the matrix $\Delta_\mu$ are of relevance, and these are invariant under this
3749: reparametrization.
3750:
3751:
3752: We now project out the spinor structure in~(\ref{S_eom_mixing_diagonalization})
3753: by multiplying by $P_s(k)\{1,s\gamma^3\gamma^5,-is\gamma^3,-\gamma^5\}$ and
3754: taking the spinorial traces. The procedure is identical to the derivation
3755: of (\ref{ga0_eom}-\ref{ga3_eom}) at the beginning of
3756: section~\ref{Constraint and kinetic equations}, except for the subtlety
3757: related to the covariant derivative, because it contains spinor structure.
3758: The resulting equations are
3759: %
3760: \begin{eqnarray}
3761: &&\!\!\!\!\!\!\!\!
3762: \Big(2i\tilde k_0 - {\cal D}_t^-\Big) g^s_{0d}
3763: -s \Big(2ik_z + {\cal D}_z\Big) g^s_{3d}
3764: -2i {m}_{hd} \,
3765: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{1d}
3766: -2i {m}_{ad}
3767: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{2d}
3768: \label{g0d_eom}
3769: \\
3770: &&\phantom{sssssssssssssssssssssssssssssssssssssssssssssssssssssssssssss}
3771: = \Tr {\mathbbm 1}P_s{\cal C}_{\psi d}
3772: \nonumber\\
3773: %
3774: &&\!\!\!\!\!\!\!\!
3775: \Big(2i\tilde k_0 - {\cal D}_t^+\Big) g^s_{1d}
3776: -s \Big(2k_z - i {\cal D}_z\Big) g^s_{2d}
3777: -2i {m}_{hd}
3778: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{0d}
3779: +2 {m}_{ad}
3780: {\rm e}^{-\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{3d}
3781: \label{g1d_eom}
3782: \\
3783: &&\phantom{sssssssssssssssssssssssssssssssssssssssssssssssssssssssssssss}
3784: = \Tr (s\gamma^3\gamma^5) P_s{\cal C}_{\psi d}
3785: \nonumber
3786: \\
3787: %
3788: &&\!\!\!\!\!\!\!\!
3789: \Big(2i\tilde k_0 - {\cal D}_t^+\Big) g^s_{2d}
3790: +s \Big(2k_z - i {\cal D}_z\Big) g^s_{1d}
3791: -2 {m}_{hd}
3792: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{3d}
3793: -2i {m}_{ad}
3794: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{0d}
3795: \label{g2d_eom}
3796: \\
3797: &&\phantom{sssssssssssssssssssssssssssssssssssssssssssssssssssssssssssss}
3798: = \Tr \left(-is\gamma^3\right)P_s{\cal C}_{\psi d}
3799: \quad
3800: \nonumber\\
3801: %
3802: &&\!\!\!\!\!\!\!\!
3803: \Big(2i\tilde k_0 - {\cal D}_t^-\Big) g^s_{3d}
3804: -s \Big(2ik_z + {\cal D}_z\Big) g^s_{0d}
3805: +2 {m}_{hd}
3806: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{2d}
3807: -2 {m}_{ad}
3808: {\rm e}^{\frac i2 {\stackrel{\!\!\leftarrow}{D_z}\partial_{k_z}}} g^s_{1d}
3809: \label{g3d_eom}
3810: \\
3811: &&\phantom{sssssssssssssssssssssssssssssssssssssssssssssssssssssssssssss}
3812: = \Tr \left(-\gamma^5\right)P_s{\cal C}_{\psi d}
3813: \,,
3814: \nonumber
3815: \end{eqnarray}
3816: %
3817: where we have defined the derivatives
3818: %
3819: \begin{eqnarray}
3820: D_z m_{hd} &\equiv & \partial_z m_{hd}
3821: - i[\Sigma_z,m_{hd}] - \{\Delta_z, m_{ad}\}
3822: \nonumber\\
3823: D_z m_{ad} &\equiv & \partial_z m_{ad}
3824: - i[\Sigma_z,m_{ad}] + \{\Delta_z, m_{hd}\}
3825: \label{Dm_ha}
3826: \end{eqnarray}
3827: %
3828: and
3829: %
3830: \begin{eqnarray}
3831: {\cal D}_t^- &=& \gamma_\| \partial_t
3832: + \gamma_\|\vec v_\|\cdot\nabla_\|
3833: - is[\Delta_z,\cdot]
3834: \label{D_t-}
3835: \\
3836: {\cal D}_t^+ &=& \gamma_\| \partial_t
3837: + \gamma_\|\vec v_\|\cdot\nabla_\|
3838: - is\{\Delta_z,\cdot\}
3839: \label{D_t+}
3840: \\
3841: {\cal D}_z &=& \partial_z - i[\Sigma_z,\cdot]
3842: \label{D_z}
3843: \end{eqnarray}
3844: %
3845: Note that due to the $\gamma^5$ the anticommutator in~(\ref{mD}) has become
3846: the commutator in~(\ref{D_t-}). Moreover, the derivative
3847: ${\cal D}_t^+$ is not hermitean. Indeed, the
3848: anticommutator term in~(\ref{D_t+}) is antihermitean.
3849:
3850:
3851:
3852:
3853: \subsubsection{Constraint equations}
3854: \label{Constraint equations}
3855:
3856: The constraint equations correspond to the antihermitean
3857: parts of the equations (\ref{g0d_eom})-(\ref{g3d_eom}).
3858: %
3859: Since we are interested in the dispersion relation accurate to first order
3860: in $\hbar$, it suffices to consider these equations to first order in
3861: gradients:
3862: %
3863: \begin{eqnarray}
3864: \! 2\tilde k_0 g^s_{0d}
3865: - 2s k_z g^s_{3d}
3866: - \big\{m_{hd}, g^s_{1d}\big\}
3867: - \frac i2 \big[D_z m_{hd},\partial_{k_z} g^s_{1d}\big]
3868: - \big\{m_{ad}, g^s_{2d}\big\}
3869: - \frac i2 \big[D_z m_{ad}, \partial_{k_z} g^s_{2d}\big]
3870: \!\! &=&\! {\cal C}^s_{0d}
3871: \quad
3872: \nonumber\\
3873: \label{ce0d1}
3874: \\
3875: %
3876: 2\tk g^s_{1d} + s\{\Delta_z, g^s_{1d}\}
3877: + s(\partial_z - i[\Sigma_z,\cdot])g^s_{2d}
3878: - \big\{m_{hd}, g^s_{0d}\big\}
3879: - \frac i2 \big[D_z m_{hd},\partial_{k_z} g^s_{0d}\big]
3880: &&
3881: \label{ce1d1}
3882: \\
3883: + \frac 12 \big\{D_z m_{ad}, \partial_{k_z} g^s_{3d}\big\}
3884: -i \big[m_{ad}, g^s_{3d}\big]
3885: &=& {\cal C}^s_{1d}
3886: \nonumber\\
3887: %
3888: 2\tilde k_0 g^s_{2d}
3889: + s\{\Delta_z, g^s_{2d}\}
3890: - s(\partial_z - i[\Sigma_z,\cdot])g^s_{1d}
3891: - \frac 12 \big\{D_z m_{hd}, \partial_{k_z} g^s_{3d}\big\}
3892: + i\big[m_{hd}, g^s_{3d}\big]
3893: &&
3894: \label{ce2d1}
3895: \\
3896: - \big\{m_{ad}, g^s_{0d}\big\}
3897: - \frac i2 \big[D_z m_{ad}, \partial_{k_z} g^s_{0d}\big]
3898: &=& {\cal C}^s_{2d}
3899: \nonumber\\
3900: %
3901: \! 2 \tilde k_0 g^s_{3d}
3902: \!-\! 2s k_z g^s_{0d}
3903: \!+\! \frac 12 \big\{D_z m_{hd}, \partial_{k_z} g^s_{2d}\big\}
3904: \!-\! i\big[m_{hd}, g^s_{2d}\big]
3905: \!-\! \frac 12 \big\{D_z m_{ad},\partial_{k_z} g^s_{1d}\big\}
3906: + i \big[m_{ad}, g^s_{1d}\big]
3907: \!\! &=& \!{\cal C}^s_{3d}
3908: \,,
3909: \quad
3910: \nonumber\\
3911: \label{ce3d1}
3912: \end{eqnarray}
3913: %
3914: where
3915: %
3916: \begin{eqnarray}
3917: {\cal C}^s_{0d} &=& \frac{1}{2i} \, {\rm Tr}{\mathbbm 1}
3918: \Big( P_s(k){\cal C}_{\psi d}
3919: -P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
3920: \nonumber\\
3921: {\cal C}^s_{1d} &=& \frac{1}{2i} \, {\rm Tr}(s\gamma^3\gamma^5)
3922: \Big( P_s(k){\cal C}_{\psi d}
3923: -P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
3924: \nonumber\\
3925: {\cal C}^s_{2d} &=& \frac{1}{2i} \, {\rm Tr}(-is\gamma^3)
3926: \Big( P_s(k){\cal C}_{\psi d}
3927: -P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
3928: \nonumber\\
3929: {\cal C}^s_{3d} &=& \frac{1}{2i} \, {\rm Tr}(-\gamma^5)
3930: \Big( P_s(k){\cal C}_{\psi d}
3931: -P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
3932: \,.
3933: \label{ced:collision-terms:0-3}
3934: \end{eqnarray}
3935: %
3936: The trace is only to be taken in spinor space, not in the fermionic
3937: flavor space.
3938: For simplicity we now perform the analysis for two mixing
3939: fermions. Our findings are however valid for an arbitrary number of mixing
3940: fermions. Taking account of the fact that the off-diagonals are of the order
3941: $\hbar$, equations~(\ref{ce0d1}-\ref{ce3d1}) now imply
3942: the following diagonal equations accurate to order $\hbar$,
3943: %
3944: \begin{eqnarray}
3945: 2\tilde k_0\,{g_{0d}^s}_{{11}}
3946: - 2{m_{hd}}_{{1}}\, {g_{1d}^s}_{{11}}
3947: - 2{m_{ad}}_{{1}}\, {g_{2d}^s}_{{11}}
3948: - 2sk_z \, {g_{3d}^s}_{{11}}
3949: \phantom{s}
3950: &=& {{\cal C}^s_{0d}}_{{11}}
3951: \label{ced:diag0}
3952: \\
3953: - 2{m_{hd}}_{{1}}\, {g_{0d}^s}_{{11}}
3954: + 2\tilde k_0\,{g_{1d}^s}_{{11}}
3955: + 2s{\Delta_z}_{{11}} \, {g_{1d}^s}_{{11}}
3956: + s\partial_z \, {g_{2d}^s}_{{11}}
3957: + {(D_zm_{ad})}_{{11}}\partial_{k_z} {g_{3d}^s}_{{11}}
3958: \phantom{}
3959: &=& {{\cal C}^s_{1d}}_{{11}}
3960: \label{ced:diag1}
3961: \\
3962: - 2{m_{ad}}_{{1}}\, {g_{0d}^s}_{{11}}
3963: - s\partial_z \, {g_{1d}^s}_{{11}}
3964: + 2\tilde k_0\,{g_{2d}^s}_{{11}}
3965: + 2s{\Delta_z}_{{11}} \, {g_{2d}^s}_{{11}}
3966: - {(D_zm_{hd})}_{{11}}\partial_{k_z} {g_{3d}^s}_{{11}}
3967: \phantom{}
3968: &=& {{\cal C}^s_{1d}}_{{11}}
3969: \label{ced:diag2}
3970: \\
3971: - 2sk_z \, {g_{0d}^s}_{{11}}
3972: - {(D_zm_{ad})}_{{11}}\partial_{k_z} {g_{1d}^s}_{{11}}
3973: + {(D_zm_{hd})}_{{11}}\partial_{k_z} {g_{2d}^s}_{{11}}
3974: + 2\tilde k_0\,{g_{3d}^s}_{{11}}
3975: \phantom{s}
3976: &=& {{\cal C}^s_{3d}}_{{11}}
3977: \,,
3978: \label{ced:diag3}
3979: \end{eqnarray}
3980: %
3981: and similarly for the $(22)$-components.
3982: Solving the latter three equations in gradient expansion in terms of
3983: ${g_{0d}^s}_{{11}}$ we find
3984: %
3985: \begin{eqnarray}
3986: {g_{1d}^s}_{{11}} &=& \frac{1}{\tilde k_0}
3987: \Big[
3988: {m_{hd}}_{{1}}\, {g_{0d}^s}_{{11}}
3989: - \frac{1}{\tilde k_0} s{\Delta_z}_{{11}} \,
3990: ({m_{hd}}_{{1}} \, {g_{0d}^s}_{{11}})
3991: - \frac{1}{2\tilde k_0} s\partial_z \,
3992: ({m_{ad}}_{{1}} \, {g_{0d}^s}_{{11}})
3993: \nonumber\\
3994: && \phantom{ssss}
3995: - \frac{s}{2\tilde k_0}{(D_zm_{ad})}_{11}(1 + k_z\partial_{k_z})
3996: {g_{0d}^s}_{{11}}
3997: + \frac 12{{\cal C}^s_{1d}}_{{11}}
3998: \Big]
3999: \label{ced:diag1b}
4000: \\
4001: {g_{2d}^s}_{{11}} &=& \frac{1}{\tilde k_0}
4002: \Big[
4003: {m_{ad}}_{{1}}\, {g_{0d}^s}_{{11}}
4004: + \frac{1}{2\tilde k_0}s\partial_z \,
4005: ({m_{hd}}_{{1}} {g_{0d}^s}_{{11}})
4006: - \frac{1}{\tilde k_0} s{\Delta_z}_{{11}}
4007: ({m_{ad}}_{{1}} {g_{0d}^s}_{{11}})
4008: \nonumber\\
4009: && \phantom{ssss}
4010: + \frac{s}{2\tilde k_0}{(D_zm_{hd})}_{{11}}
4011: (1+k_z \partial_{k_z}){g_{0d}^s}_{11}
4012: + \frac 12 {{\cal C}^s_{2d}}_{{11}}
4013: \Big]
4014: \label{ced:diag2b}
4015: \\
4016: {g_{3d}^s}_{{11}} &=& \frac{1}{\tilde k_0}
4017: \Big[
4018: sk_z \, {g_{0d}^s}_{{11}}
4019: + \frac{1}{2\tilde k_0}
4020: {m_{hd}}_{{1}}{(D_zm_{ad})}_{{11}}\partial_{k_z}
4021: {g_{0d}^s}_{{11}}
4022: - \frac{1}{2\tilde k_0}
4023: {m_{ad}}_{{1}}{(D_zm_{hd})}_{{11}}\partial_{k_z}
4024: {g_{0d}^s}_{{11}}
4025: + \frac 12{{\cal C}^s_{3d}}_{{11}}
4026: \Big]
4027: \,.\quad
4028: \label{ced:diag3b}
4029: \end{eqnarray}
4030: %
4031: Here ${m_{hd}}_{{1}}$ denotes the first diagonal element of $m_{hd}$,
4032: ${m_{ad}}_{{1}}$ is defined correspondingly.
4033: Inserting these relations into~(\ref{ced:diag0}) we get the
4034: constraint for the diagonal densities
4035: %
4036: \beqa
4037: &&
4038: \frac{2}{\tilde{k}_0}
4039: \Big( k^2
4040: -|m_d|^2_{i}
4041: + \frac{s}{\tilde k_0}
4042: \big[ |m_d|^2_{i}(\partial_z {\theta_d}_{i}
4043: + 2{\Delta_z}_{\tt ii})
4044: \big]
4045: \Big)\,{g_{0d}^s}_{\tt ii}
4046: \nonumber\\
4047: &&\hphantom{XXXXX}
4048: = {{\cal C}^s_{0d}}_{\tt ii}
4049: + \frac{1}{\tilde k_0}
4050: \big( {m_{hd}}_{i}{{\cal C}^s_{1d}}_{\tt ii}
4051: + {m_{ad}}_{i}{{\cal C}^s_{2d}}_{\tt ii}
4052: + sk_z {{\cal C}^s_{3d}}_{\tt ii}
4053: \big)
4054: \,,
4055: \label{ced:diag0b}
4056: \eeqa
4057: %
4058: where we defined
4059: %
4060: \begin{eqnarray}
4061: |m_d|^2_{{i}}
4062: &=& {m_{hd}}_{i}^2 + {m_{ad}}_{i}^2
4063: \nonumber\\
4064: |m_d|^2_{{i}}\partial_z {\theta_d}_{i}
4065: &=& {m_{hd}}_{{i}}\partial_z{m_{ad}}_{{i}}
4066: - {m_{ad}}_{{i}}\partial_z{m_{hd}}_{{i}}
4067: \,,
4068: \label{ced:diag0c}
4069: \end{eqnarray}
4070: %
4071: where $i=1,2$.
4072: Note that the energy shift in~(\ref{ced:diag0b}) can be written
4073: in terms of the rotation matrices $U$ as
4074: %
4075: \begin{equation}
4076: |m_d|^2_{i}(\partial_z {\theta_d}_{i} + 2{\Delta_z}_{\tt ii})
4077: = - \Im[U(m\partial_z m^\dagger)U^\dagger]_{\tt ii}
4078: \,.
4079: \label{shift:U}
4080: \end{equation}
4081: %
4082:
4083:
4084:
4085:
4086: \subsubsection{Quasiparticle picture and dispersion relation}
4087: \label{Quasiparticle picture and dispersion relation}
4088:
4089: Equation~(\ref{ced:diag0b})
4090: %~(\ref{ced:diag0b}-\ref{ced:diag0c})
4091: specifies the spectral
4092: properties of the plasma excitations. Since it
4093: %~(\ref{ced:diag0b})
4094: is an algebraic constraint, the {\it quasiparticle picture}
4095: remains a valid description of the plasma to order $\hbar$,
4096: as it was already pointed out
4097: in~\cite{KainulainenProkopecSchmidtWeinstock:2001,
4098: KainulainenProkopecSchmidtWeinstock:2002}.
4099: %Note that in deriving~(\ref{ced:diag0b}) we have made no assumption
4100: %about the space-time dependences of $g^s_{ad}$.
4101: Since the constraint equation
4102: contains no time dependence, it measures genuine
4103: spectral properties independent of the actual dynamical
4104: populations of the states.
4105:
4106: The homogeneous solution of the constraint equation~(\ref{ced:diag0b})
4107: has the following spectral form
4108: %
4109: \beqa
4110: {g_{0d}^{<s}}_{\tt ii}(k,x)
4111: &=& 2\pi \delta\Big( k^2
4112: - |m_d|^2_{i}
4113: + \frac{s}{\tilde k_0}
4114: \big[ |m_d|^2_{i}(\partial_z {\theta_d}_{i}
4115: + 2{\Delta_z}_{\tt ii})
4116: \big]
4117: \Big)
4118: |\tilde{k}_0|\,n_{si}(k,x)
4119: %
4120: \nonumber\\
4121: %
4122: &=& 2\pi \sum_{\pm} \frac{\delta(k_0\mp\omega_{\pm si})}
4123: {2\omega_{\pm si}Z_{\pm si}} \,
4124: |\tilde{k}_0|\,n_{si}(k,x)
4125: \,,
4126: \label{spectral-solution:fermions}
4127: \eeqa
4128: %
4129: with the following {\it dispersion relations} for fermions
4130: %
4131: \begin{equation}
4132: \omega_{si} = \omega_{0i}
4133: - \frac{s}{2\tilde \omega_{0i}\omega_{0i}}
4134: \big[ |m_d|^2_{i}(\partial_z {\theta_d}_{i}
4135: + 2{\Delta_z}_{\tt ii})
4136: \big]
4137: ,
4138: \label{dispersion-relation:fermions}
4139: \end{equation}
4140: %
4141: where
4142: %
4143: \begin{eqnarray}
4144: \omega_{0i} &=& ({\vec k^2 + |m_d|^2_{i}})^\frac 12
4145: \nonumber\\
4146: \tilde\omega_{0i} &=& ({k_z^2 + |m_d|^2_{i}})^\frac 12
4147: \,.
4148: \label{dispersion-relation:fermions2}
4149: \end{eqnarray}
4150: %
4151: The normalization factors are
4152: %
4153: \beq
4154: Z_{si} = 1 - \frac{s|m_d|^2_{i}(\partial_z {\theta_d}_{i}
4155: + 2{\Delta_z}_{\tt ii})}{2\tilde \omega_{0i}^3}
4156: \,.
4157: \label{Zsipm}
4158: \eeq
4159: %
4160: We have normalized the solution~(\ref{spectral-solution:fermions})
4161: such that $n_{si}(k,x)$ represents the particle density in phase space
4162: $\{k,x\}$.
4163: We emphasize that the spectral solution~(\ref{spectral-solution:fermions})
4164: and the dispersion relation~(\ref{dispersion-relation:fermions})
4165: are valid in general for all plasma excitations:
4166: equilibrium and stationary excitations,
4167: as well as for non-equilibrium, time dependent transients that conserve spin,
4168: that is for ${g^s_{0d}} = {g^s_{0d}}(k,t-\vec k_\|\cdot\vec x_\|,z)$,
4169: assuming of course planar symmetry in the wall frame.
4170:
4171:
4172: Since in the on-shell limit the Wigner functions $S^<$ and $S^>$ satisfy
4173: identical Kadanoff-Baym equations~(\ref{Wigner-space:fermionic_eom}),
4174: and in addition they are related by the fermionic
4175: sum rule~(\ref{sum_rule_fermions}), based on the above analysis
4176: of $S^<$ we can easily reconstruct the spectral solution for the
4177: Wigner function $S^>$, which we write as
4178: %
4179: \begin{equation}
4180: S^>_d = \sum_{s=\pm 1} iP_s(k)
4181: \left[ s\gamma^3\gamma^5 g^{>s}_{0d}
4182: - s\gamma^3 g^{>s}_{3d}
4183: + {\mathbbm 1} g^{>s}_{1d}
4184: - i\gamma^5 g^{>s}_{2d} \right]
4185: ,
4186: \label{S>_decomposition}
4187: \end{equation}
4188: %
4189: where the spectral solution for $g^{>s}_{0d}$ reads
4190: %
4191: \begin{equation}
4192: g_{0d\tt ii}^{>s}(k,x)
4193: = - 2\pi \sum_{\pm} \frac{\delta(k_0\mp\omega_{\pm si})}
4194: {2\omega_{\pm si}Z_{\pm si}}
4195: \, |\tilde{k}_0| \, [1-n_{si}(k,x)]
4196: \,.
4197: \label{spectral-solution:fermions>}
4198: \end{equation}
4199: %
4200: Note that $g_{0d\tt ii}^{>s}$
4201: obeys the same equation~(\ref{ced:diag0b}) as $g_{0d\tt ii}^{<s}$.
4202: Similarly, $g_{1d\tt ii}^{>s}$, $g_{2d\tt ii}^{>s}$
4203: and $g_{3d\tt ii}^{>s}$ are related to $g_{0d\tt ii}^{>s}$
4204: the same way as indicated in
4205: Eqs.~(\ref{ced:diag1b}-\ref{ced:diag3b}).
4206:
4207:
4208:
4209:
4210: \subsubsection{Plasma frame}
4211: \label{Plasma frame}
4212:
4213: An interesting question is of course how to generalize the solution for
4214: the Wigner function~(\ref{S<_decomposition2})
4215: to other Lorentz frames, an important example being the plasma rest frame.
4216: The natural solution is to apply a Lorentz boost
4217: on the Wigner function. The simplest boost
4218: operator corresponding to a boost $v$ in $z$-direction
4219: can be easily constructed in analogy with~(\ref{L-Lambda}), and it reads
4220: %
4221: \begin{equation}
4222: L(\Lambda_z) = \frac{\gamma + 1 - \gamma v\cdot\gamma^0\gamma^3}
4223: {[{2(\gamma+1)}]^{1/2}},
4224: \label{L-Lambdaz}
4225: \end{equation}
4226: %
4227: where $\gamma = (1-v^2)^{-1/2}$.
4228: The Wigner function then transforms as
4229: %
4230: \begin{eqnarray}
4231: \breve{S}_s &=& L^{-1}(\Lambda_z)S_sL(\Lambda_z)
4232: \nonumber\\
4233: &=& - \breve{P}_s(k)
4234: \left[
4235: \gamma(1+v\gamma^0\gamma^3)
4236: \left( s\gamma^3\gamma^5 \tilde g^{s}_0
4237: - s\gamma^3 \tilde g^{s}_3
4238: \right)
4239: + {\mathbbm 1} \tilde g^{s}_1
4240: - i\gamma^5 \tilde g^{s}_2
4241: \right]
4242: \,,
4243: \label{boost:v}
4244: \end{eqnarray}
4245: %
4246: where
4247: %
4248: \begin{eqnarray}
4249: \breve{P}_s(\breve{k})
4250: = \frac{\gamma(\breve k_0 - v\breve{k}_z)}{\breve{\tilde k}_0}
4251: \gamma^0\gamma^3\gamma^5
4252: - \gamma(1+v\gamma^0\gamma^3)
4253: \frac{\breve{\vec k}_\|}{\breve{\tilde k}_0}
4254: \cdot\vec \gamma_\|\gamma^3\gamma^5
4255: \label{brevePs}
4256: \end{eqnarray}
4257: %
4258: is the spin operator
4259: with $\breve{\tilde k}_0^2
4260: = \gamma^2(\breve k_0 - v\breve{k}_z)^2-\breve{\vec k}_\|^2$, and
4261: $g_a^s=g_a^s(\breve k,\breve x)$. We leave as an exercise
4262: to the reader to find the direction of the spin vector $\breve s^\mu$
4263: corresponding to the spin operator~(\ref{brevePs}).
4264: To get the final form for the Wigner function in the new frame,
4265: one also needs to transform the coordinates and derivatives
4266: % from $k,x$ to $\breve k,\breve x$
4267: in~(\ref{ced:diag1b}-\ref{ced:diag3b}) and~(\ref{spectral-solution:fermions}),
4268: which makes the final expression for $i\breve S^<_s$ rather cumbersome.
4269: To get the Wigner function
4270: $i\breve S_s$ in the {\it plasma frame}, one needs to choose the boost
4271: that corresponds to $v = v_w$ and $\gamma = \gamma_w = (1-v_w^2)^{-1/2}$.
4272: This exercise underlines clearly the advantages of working in the
4273: wall frame, in which the Wigner function has a particularly simple form.
4274: Some authors~\cite{CarenaMorenoQuirosSecoWagner:2000,
4275: CarenaQuirosSecoWagner:2002} choose nevertheless to work in the plasma frame,
4276: although in a simplistic disguise.
4277:
4278: \subsubsection{Kinetic equations}
4279: \label{Kinetic equations}
4280:
4281: We now take (minus) the hermitean part of
4282: Eqs.~(\ref{g0d_eom}-\ref{g3d_eom}) and get the following
4283: {\it kinetic equations}
4284: %For a planar wall moving in $z$ direction we have
4285: %
4286: \begin{eqnarray}
4287: &&
4288: {\cal D}_t^- g^s_{0d}
4289: % \{\gamma_\|\partial_t - is[\Delta_z,\cdot]\} g^s_{0d}
4290: + s {\cal D}_z g^s_{3d}
4291: % + \{s(\partial_z -i[\Sigma_z,\cdot]) - i\gamma_\|[\Delta_t,\cdot]\} g^s_{3d}
4292: - \{m_{hd}
4293: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4294: g^s_{1d}\}
4295: + i[m_{hd}
4296: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4297: g^s_{1d}]
4298: \nonumber\\
4299: &&
4300: \phantom{sssssssssssssss}
4301: - \{m_{ad}
4302: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4303: g^s_{2d}\}
4304: + i[m_{ad}
4305: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4306: g^s_{2d}]
4307: = {\cal K}^s_{0d}
4308: \quad
4309: \label{ked0}
4310: \\
4311: %
4312: &&
4313: ({\cal D}_t^+)_h g^s_{1d}
4314: +2sk_z g^s_{2d}
4315: - \{m_{hd}
4316: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4317: g^s_{0d}\}
4318: + i[m_{hd}
4319: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4320: g^s_{0d}]
4321: \phantom{sssssssss}
4322: \nonumber\\
4323: &&
4324: \phantom{sssssssssssssss}
4325: - \{m_{ad}
4326: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4327: g^s_{3d}\}
4328: - i[m_{ad}
4329: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4330: g^s_{3d}]
4331: = {\cal K}^s_{1d}
4332: \quad
4333: \label{ked1}
4334: \\
4335: %
4336: &&
4337: ({\cal D}_t^+)_h g^s_{2d}
4338: -2sk_z g^s_{1d}
4339: + \{m_{hd}
4340: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4341: g^s_{3d}\}
4342: + i[m_{hd}
4343: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4344: g^s_{3d}]
4345: \phantom{sssssssss}
4346: \nonumber\\
4347: &&
4348: \phantom{sssssssssssssss}
4349: - \{m_{ad}
4350: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4351: g^s_{0d}\}
4352: + i[m_{ad}
4353: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4354: g^s_{0d}]
4355: = {\cal K}^s_{2d}
4356: \quad
4357: \label{ked2}
4358: \\
4359: %
4360: &&
4361: {\cal D}_t^- g^s_{3d}
4362: +s {\cal D}_z g^s_{0d}
4363: - \{m_{hd}
4364: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4365: g^s_{2d}\}
4366: - i[m_{hd}
4367: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4368: g^s_{2d}]
4369: \nonumber\\
4370: &&
4371: \phantom{sssssssssssssss}
4372: + \{m_{ad}
4373: \cos\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4374: g^s_{1d}\}
4375: + i[m_{ad}
4376: \sin\big(\frac 12 {\stackrel{\leftarrow}{D_z}\!\partial_{k_z}}\big),
4377: g^s_{1d}]
4378: = {\cal K}^s_{3d}
4379: \label{ked3}
4380: \,,
4381: \quad
4382: \end{eqnarray}
4383: %
4384: where $({\cal D}_t^+)_h = \gamma_\|(\partial_t -i[\Sigma_t,\cdot])
4385: + \gamma_\|\vec v_\|\cdot \nabla_\| $
4386: denotes the hermitean part of the derivative, and
4387: %
4388: \begin{eqnarray}
4389: {\cal K}^s_{0d} &=& -\frac{1}{2} \, {\rm Tr}{\mathbbm 1}
4390: \Big( P_s(k){\cal C}_{\psi d}
4391: +P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
4392: \nonumber\\
4393: {\cal K}^s_{1d} &=& -\frac{1}{2} \, {\rm Tr}(s\gamma^3\gamma^5)
4394: \Big( P_s(k){\cal C}_{\psi d}
4395: +P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
4396: \nonumber\\
4397: {\cal K}^s_{2d} &=& -\frac{1}{2} \, {\rm Tr}(-is\gamma^3)
4398: \Big( P_s(k){\cal C}_{\psi d}
4399: +P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
4400: \nonumber\\
4401: {\cal K}^s_{3d} &=& -\frac{1}{2} \, {\rm Tr}(-\gamma^5)
4402: \Big( P_s(k){\cal C}_{\psi d}
4403: +P^\dagger_s(k){\cal C}_{\psi d}^\dagger \Big)
4404: \,.
4405: \label{ked:collision-terms:0-3}
4406: \end{eqnarray}
4407: %
4408: Again the trace must only be taken in spinor space.
4409: Since we are interested in the order $\hbar$ effects, we can truncate
4410: the kinetic equations~(\ref{ked0}-\ref{ked3}) to second order in gradients.
4411: For example, the kinetic equation for particle density of spin
4412: $s$~(\ref{ked0}) reads
4413: %
4414: \begin{eqnarray}
4415: &&
4416: %{\cal D}_t^- g^s_{0d}
4417: (\gamma_\|\partial_t + \gamma_\|\vec v_\|\cdot\nabla_\|
4418: - is[\Delta_z,\cdot]) g^s_{0d}
4419: % + s {\cal D}_z^- g^s_{3d}
4420: + s(\partial_z -i[\Sigma_z,\cdot]) g^s_{3d}
4421: \nonumber\\
4422: &&
4423: \phantom{sssssss}
4424: - \frac 12 \{D_z m_{hd}, \partial_{k_z} g^s_{1d}\}
4425: + i[m_{hd}, g^s_{1d}]
4426: - \frac 12 \{D_z m_{ad}, \partial_{k_z} g^s_{2d}\}
4427: + i[m_{ad}, g^s_{2d}]
4428: = {\cal K}^s_{0d}
4429: \,,
4430: \quad
4431: \label{ked0-2}
4432: \end{eqnarray}
4433: %
4434: where we dropped the second order terms in the commutator, which is legitimate
4435: since they contribute only to the off-diagonal equations,
4436: and hence to order $\hbar^2$.
4437:
4438: From~(\ref{ked0-2}) we can immediately write the kinetic equation
4439: for the diagonal densities
4440: %
4441: \begin{eqnarray}
4442: && \gamma_\|(\partial_t + \vec v_\|\cdot\nabla_\|) {g^s_{0d}}_{11}
4443: + s\partial_z {g^s_{3d}}_{11}
4444: - (D_z m_{hd})_{11} \partial_{k_z} {g^s_{1d}}_{11}
4445: - (D_z m_{ad})_{11} \partial_{k_z} {g^s_{2d}}_{11}
4446: \nonumber\\
4447: &\mp& is({\Sigma_{z}}_{12}{g_{3d}^s}_{21}-{\Sigma_{z}}_{21}{g_{3d}^s}_{12}
4448: + {\Delta_{z}}_{12}{g_{0d}^s}_{21}-{\Delta_{z}}_{21}{g_{0d}^s}_{12})
4449: \label{ked0-diagB}
4450: \\
4451: &-& \frac 12\Big[
4452: (D_z m_{hd})_{12} \partial_{k_z} {g^s_{1d}}_{21}
4453: + (D_z m_{hd})_{21} \partial_{k_z} {g^s_{1d}}_{12}
4454: + (D_z m_{ad})_{12} \partial_{k_z} {g^s_{2d}}_{21}
4455: + (D_z m_{ad})_{21} \partial_{k_z} {g^s_{2d}}_{12}
4456: \Big]
4457: = {{\cal K}^s_{0d}}_{11}
4458: .
4459: \nonumber
4460: \end{eqnarray}
4461: %
4462: In order to decouple the diagonal and off-diagonal equations we make use
4463: of the off-diagonal equations, which to leading order in gradients are
4464: %
4465: \begin{eqnarray}
4466: && \!\!\!\!\!\!\!\!
4467: i\delta(m_{hd}){g_{1d}^s}_{12}
4468: + i\delta(m_{ad}){g_{2d}^s}_{12}
4469: \label{ke-off0}
4470: \\
4471: &&
4472: %\phantom{s}
4473: = - is{\Delta_{z}}_{12} \delta({g_{0d}^s})
4474: + \frac 12 (D_z m_{hd})_{12}\partial_{k_z} \Tr({g_{1d}^s})
4475: + \frac 12 (D_z m_{ad})_{12}\partial_{k_z} \Tr({g_{2d}^s})
4476: - is{\Sigma_{z}}_{12} \delta({g_{3d}^s})
4477: + {{\cal K}^s_{0d}}_{12}
4478: \nonumber
4479: \\
4480: && \!\!\!\!\!\!\!\!
4481: i\delta(m_{hd}){g_{0d}^s}_{12}
4482: + 2sk_z {g_{2d}^s}_{12}
4483: - \Tr(m_{ad}){g_{3d}^s}_{12}
4484: \label{ke-off1}
4485: \\
4486: &&
4487: %\phantom{s}
4488: = \frac 12 (D_z m_{hd})_{12}\partial_{k_z} \Tr({g_{0d}^s})
4489: - \frac i2 (D_z m_{ad})_{12}\partial_{k_z} \delta({g_{3d}^s})
4490: + {{\cal K}^s_{1d}}_{12}
4491: \nonumber
4492: \\
4493: && \!\!\!\!\!\!\!\!
4494: i\delta(m_{ad}){g_{0d}^s}_{12}
4495: - 2sk_z {g_{1d}^s}_{12}
4496: + \Tr(m_{hd}){g_{3d}^s}_{12}
4497: \label{ke-off2}
4498: \\
4499: &&
4500: %\phantom{s}
4501: = \frac 12 (D_z m_{ad})_{12}\partial_{k_z} \Tr({g_{0d}^s})
4502: + \frac i2 (D_z m_{hd})_{12}\partial_{k_z} \delta({g_{3d}^s})
4503: + {{\cal K}^s_{2d}}_{12}
4504: \nonumber
4505: \\
4506: && \!\!\!\!\!\!\!\!
4507: \Tr(m_{ad}){g_{1d}^s}_{12}
4508: - \Tr(m_{hd}){g_{2d}^s}_{12}
4509: \label{ke-off3}\\
4510: && = - is{\Sigma_{z}}_{12} \delta({g_{0d}^s})
4511: + \frac i2 (D_z m_{ad})_{12}\partial_{k_z} \delta({g_{1d}^s})
4512: - \frac i2 (D_z m_{hd})_{12}\partial_{k_z} \delta({g_{2d}^s})
4513: - is{\Delta_{z}}_{12} \delta({g_{3d}^s})
4514: + {{\cal K}^s_{3d}}_{12}
4515: \,.
4516: \nonumber
4517: \end{eqnarray}
4518: %
4519: Combining~(\ref{ke-off0}) and~(\ref{ke-off3}) we obtain
4520: %
4521: \begin{eqnarray}
4522: {g_{1d}^s}_{12} &=& \frac{1}{\delta(|m_d|)^2}
4523: \Big\{
4524: - s(D_z m_{ad})_{12}\delta(g_{0d}^s)
4525: - is[{\Delta_z}_{12}\delta(m_{ad})
4526: - i{\Sigma_z}_{12}\Tr(m_{hd})]\delta(g_{3d}^s)
4527: \nonumber\\
4528: &&
4529: \phantom{sssssssss}
4530: - \frac i2 [ \Tr(m_{hd})(D_z m_{hd})_{12} \partial_{k_z}\Tr(g_{1d}^s)
4531: - \Tr(m_{ad})(D_z m_{ad})_{12} \partial_{k_z}\delta(g_{1d}^s)]
4532: \nonumber\\
4533: &&
4534: \phantom{sssssssss}
4535: - \frac i2 [ \Tr(m_{hd})(D_z m_{ad})_{12} \partial_{k_z}\Tr(g_{2d}^s)
4536: + \Tr(m_{ad})(D_z m_{hd})_{12} \partial_{k_z}\delta(g_{2d}^s)]
4537: \nonumber\\
4538: &&
4539: \phantom{sssssssss}
4540: - i\Tr(m_{hd}) {{\cal K}^s_{0d}}_{12}
4541: + \delta(m_{ad}) {{\cal K}^s_{3d}}_{12}
4542: \Big\}
4543: \label{ke-off1b}
4544: \\
4545: {g_{2d}^s}_{12} &=& \frac{1}{\delta(|m_d|)^2}
4546: \Big\{\phantom{ss}
4547: s(D_z m_{hd})_{12}\delta(g_{0d}^s)
4548: + is[{\Delta_z}_{12}\delta(m_{hd})
4549: + i{\Sigma_z}_{12}\Tr(m_{ad})]\delta(g_{3d}^s)
4550: \nonumber\\
4551: &&
4552: \phantom{sssssssss}
4553: - \frac i2 [ \Tr(m_{ad})(D_z m_{hd})_{12} \partial_{k_z}\Tr(g_{1d}^s)
4554: + \Tr(m_{hd})(D_z m_{ad})_{12} \partial_{k_z}\delta(g_{1d}^s)]
4555: \nonumber\\
4556: &&
4557: \phantom{sssssssss}
4558: - \frac i2 [ \Tr(m_{ad})(D_z m_{ad})_{12} \partial_{k_z}\Tr(g_{2d}^s)
4559: - \Tr(m_{hd})(D_z m_{hd})_{12} \partial_{k_z}\delta(g_{2d}^s)]
4560: \nonumber\\
4561: &&
4562: \phantom{sssssssss}
4563: - i\Tr(m_{ad}) {{\cal K}^s_{0d}}_{12}
4564: - \delta(m_{hd}) {{\cal K}^s_{3d}}_{12}
4565: \Big\}
4566: \,,
4567: \label{ke-off2b}
4568: \end{eqnarray}
4569: %
4570: where
4571: %
4572: \begin{equation}
4573: \delta(|m_d|)^2 = \Tr(m_{hd})\delta(m_{hd})
4574: + \Tr(m_{ad})\delta(m_{ad})
4575: \,.
4576: \label{deltamd2}
4577: \end{equation}
4578: %
4579: From~(\ref{ke-off0}-\ref{ke-off3}) and~(\ref{ke-off1b}-\ref{ke-off2b})
4580: we then get
4581: %
4582: \begin{eqnarray}
4583: {g_{0d}^s}_{12} &=& \frac{1}{\delta(|m_d|)^2}
4584: \Big\{
4585: -2k_z[ {\Sigma_z}_{12}\delta(g_{0d}^s)
4586: +{\Delta_z}_{12}\delta(g_{3d}^s) ]
4587: + \frac 12 {\Sigma_z}_{12}\delta(|m_d|^2)\partial_{k_z}\Tr(g_{0d}^s)
4588: \nonumber\\
4589: &&
4590: \phantom{sssssssss}\,
4591: + sk_z[ (D_z m_{ad})_{12}\partial_{k_z}\delta(g_{1d}^s)
4592: - (D_z m_{hd})_{12}\partial_{k_z}\delta(g_{2d}^s)]
4593: \nonumber\\
4594: &&
4595: \phantom{sssssssss}\,
4596: - \frac 12 [ \Tr(m_{hd})(D_z m_{ad})_{12}
4597: - \Tr(m_{ad})(D_z m_{hd})_{12} ]
4598: \partial_{k_z}\delta(g_{3d}^s)
4599: \nonumber\\
4600: &&
4601: \phantom{sssssssss}\,
4602: - i\Tr(m_{hd}) {{\cal K}^{s}_{1d}}_{12}
4603: - i\Tr(m_{ad}) {{\cal K}^{s}_{2d}}_{12}
4604: - 2isk_z {{\cal K}^{s}_{3d}}_{12}
4605: \Big\}
4606: \label{ke-off0b}
4607: \\
4608: {g_{3d}^s}_{12} &=& \frac{1}{\delta(|m_d|)^2}
4609: \Big\{
4610: -2k_z[ {\Delta_z}_{12}\delta(g_{0d}^s)
4611: +{\Sigma_z}_{12}\delta(g_{3d}^s) ]
4612: + \frac 12 {\Delta_z}_{12}\delta(|m_d|^2)\partial_{k_z}\Tr(g_{0d}^s)
4613: \nonumber\\
4614: &&
4615: \phantom{sssssssss}\,
4616: - isk_z[ (D_z m_{hd})_{12}\partial_{k_z}\Tr(g_{1d}^s)
4617: + (D_z m_{ad})_{12}\partial_{k_z}\Tr(g_{2d}^s)]
4618: \nonumber\\
4619: &&
4620: \phantom{sssssssss}\,
4621: + \frac i2 [ \delta(m_{ad})(D_z m_{ad})_{12}
4622: + \delta(m_{hd})(D_z m_{hd})_{12} ]
4623: \partial_{k_z}\delta(g_{3d}^s)
4624: \nonumber\\
4625: &&
4626: \phantom{sssssssss}\,
4627: - 2isk_z {{\cal K}^s_{0d}}_{12}
4628: - \delta(m_{ad}) {{\cal K}^s_{1d}}_{12}
4629: - \delta(m_{hd}) {{\cal K}^s_{2d}}_{12}
4630: \Big\}
4631: \,.
4632: \label{ke-off3b}
4633: \end{eqnarray}
4634: %
4635: Equations~(\ref{ke-off1b}-\ref{ke-off3b}) are the off-diagonal densities
4636: correct to leading (first) order in gradients.
4637: We are now ready to compute the contribution from the off-diagonal densities
4638: in the kinetic equation~(\ref{ked0-diagB}). Making use of the hermiticity
4639: properties
4640: ${\Delta_z}_{12}^* = {\Delta_z}_{21}$,
4641: ${\Sigma_z}_{12}^* = {\Sigma_z}_{21}$,
4642: ${g_{ad}^s}_{12}^* = {g_{ad}^s}_{21}$, and~(\ref{ke-off1b}-\ref{ke-off3b}),
4643: after some algebra one finds that the contribution from the off-diagonal
4644: densities vanishes:
4645: %
4646: \begin{eqnarray}
4647: {\Sigma_{z}}_{12}{g_{3d}^s}_{21}-{\Sigma_{z}}_{21}{g_{3d}^s}_{12}
4648: + {\Delta_{z}}_{12}{g_{0d}^s}_{21}-{\Delta_{z}}_{21}{g_{0d}^s}_{12}
4649: &=& 0
4650: \nonumber\\
4651: (D_z m_{hd})_{12} \partial_{k_z} {g^s_{1d}}_{21}
4652: + (D_z m_{hd})_{21} \partial_{k_z} {g^s_{1d}}_{12}
4653: + (D_z m_{ad})_{12} \partial_{k_z} {g^s_{2d}}_{21}
4654: + (D_z m_{ad})_{21} \partial_{k_z} {g^s_{2d}}_{12}
4655: &=& 0
4656: \,, \qquad
4657: \label{ked0-diagC}
4658: \end{eqnarray}
4659: %
4660: where we neglected the collision terms that are suppressed by derivatives.
4661: With the help of this remarkable result and the constraint
4662: equations~(\ref{ced:diag1b}-\ref{ced:diag3b})
4663: the kinetic equation~(\ref{ked0-diagB}) finally reduces to the familiar
4664: form~\cite{KainulainenProkopecSchmidtWeinstock:2001,
4665: KainulainenProkopecSchmidtWeinstock:2002}
4666: %
4667: \begin{equation}
4668: \frac{1}{\tilde{k}_0}\Big[ k\cdot\del
4669: - \frac 12 \Big(\partial_z|m_d|^2_{i}
4670: - \frac{s}{\tilde k_0}
4671: \partial_z\left[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})
4672: \right]\Big)\partial_{k_z} \Big]{g^s_{0d}}_{\tt ii}
4673: = {{\cal K}^s_{0d}}_{\tt ii}
4674: \,.
4675: \label{kinetic_3+1}
4676: \end{equation}
4677: %
4678: Note that this equation holds for both $g^{s<}_{0d}$ and $g^{s>}_{0d}$.
4679: Recall furthermore that for spin conservation
4680: the space-time transients are constrained to satisfy
4681: ${g_{0d}^s}_{\tt ii} = {g_{0d}^s}_{\tt ii}(k,t-\vec v_\|\cdot \vec x_\|,z)$.
4682: The collisional contributions from the constraint equations are
4683: higher order in gradients and can be consistently neglected.
4684:
4685:
4686:
4687:
4688: The functions $g_{ad}^s$ ($a=0,1,2,3$) satisfy four apparently
4689: different equations~(\ref{ked0}--\ref{ked3}).
4690: These functions are related
4691: by the constraint equations~(\ref{ce0d1}--\ref{ce3d1}), which reduce the
4692: number of independent functions to a single one, $g_{0d}^s$, projected on the
4693: quasiparticle shell~(\ref{ced:diag0b}).
4694: %
4695: We will show in an appendix in Paper~II
4696: % Appendix~\ref{Consistency of the fermionic kinetic equations},
4697: that all four kinetic equations are actually mutually
4698: dependent in a self-consistent manner, and thus equivalent to the kinetic
4699: equation for $g_{0d}^s$, {\it plus} the on-shell
4700: equation~(\ref{ced:diag0b}). Our proof includes not only the flow term,
4701: but also the collisional sources.
4702: The equivalence of these equations including the collision term is
4703: a very nontrivial consistency check of our approach to the kinetics
4704: of fermions.
4705:
4706:
4707:
4708:
4709:
4710:
4711:
4712:
4713:
4714:
4715: \subsection{Boltzmann transport equation for CP-violating fermionic densities}
4716: \label{Boltzmann transport equation for CP-violating fermionic densities}
4717:
4718:
4719: Since our primary interest is transport of CP-violating densities,
4720: we show here how -- starting with the Kadanoff-Baym equations for
4721: fermions~(\ref{Wigner-space:fermionic_eom})
4722: ({\it cf.} also~(\ref{kinetic_3+1}))
4723: -- one obtains the on-shell Boltzmann transport equations. The crucial
4724: difference with respect to the scalar case discussed in
4725: section~\ref{Boltzmann transport equation for CP-violating scalar densities}
4726: is the CP-violating {\it semiclassical force} in the flow term of the kinetic
4727: equation for fermions. The force appears at second order in
4728: gradients, or equivalently at first order in an expansion in $\hbar$.
4729:
4730:
4731: \vskip 0.1in
4732:
4733: We begin our analysis with recalling the C and CP transformations for
4734: the fermionic fields,
4735: % in the representation~(\ref{chiral_basis_gamma}),
4736: %
4737: \begin{eqnarray}
4738: \psi^c(u)
4739: &\equiv& {\cal C} \psi(u) {\cal C}^\dagger
4740: = i\gamma^0\gamma^2 {\bar\psi}^T(u) ,
4741: \nonumber\\
4742: {\bar\psi}^c(u)
4743: &\equiv& {\cal C}\bar \psi(u) {\cal C}^\dagger
4744: = \psi^T(u) i\gamma^0\gamma^2
4745: \,,
4746: \label{C-transformations-psi}
4747: \end{eqnarray}
4748: %
4749: and for Dirac's $\gamma$-matrices,
4750: %
4751: \begin{eqnarray}
4752: {\cal C}{\gamma^\mu}^T {\cal C}^\dagger
4753: &=& i\gamma^0\gamma^2 {\gamma^\mu}^T (i\gamma^0\gamma^2)^\dagger
4754: = - \gamma^\mu
4755: \nonumber\\
4756: {\cal C} \gamma^5 {\cal C}^\dagger
4757: &=& i\gamma^0\gamma^2 \gamma^5 (i\gamma^0\gamma^2)^\dagger
4758: = \gamma^5
4759: \,.
4760: \label{C-transformations-gamma}
4761: \end{eqnarray}
4762: %
4763: In the relativistic limit
4764: the Standard Model fermions are chiral and couple differently to
4765: the weak interactions, so that both charge
4766: {\cal C} and parity {\cal P} symmetries are strongly violated.
4767: But the combined symmetry {\cal CP} is violated only very weakly,
4768: hence it is natural to consider antiparticles to be
4769: related to particles by a CP transformation.
4770: Therefore we also quote the parity transformations,
4771: %
4772: \begin{eqnarray}
4773: \psi^p(u)
4774: &\equiv& {\cal P}\psi(u) {\cal P}^\dagger
4775: = \gamma^0 \psi(\bar u),
4776: \qquad\;\,
4777: {\cal P} \gamma^\mu {\cal P}^\dagger
4778: = \gamma^0 \gamma^\mu \gamma^0
4779: = {\gamma^\mu}^\dagger = \gamma_\mu
4780: \nonumber\\
4781: {\bar\psi}^p(u)
4782: &\equiv& {\cal P}\bar\psi(u) {\cal P}^\dagger
4783: = {\bar \psi}(\bar u) \gamma^0
4784: \,,\qquad
4785: {\cal P} \gamma^5 {\cal P}^\dagger
4786: = \gamma^0 \gamma^5 \gamma^0 = - \gamma^5
4787: \,.
4788: \label{P-transformations-psi+gamma}
4789: \end{eqnarray}
4790: %
4791: Combining these relations with the definitions of the fermionic
4792: Wightman functions~(\ref{Green_fermionic_index}),
4793: we easily find the relations for their C- and CP-transformations:
4794: %
4795: \begin{eqnarray}
4796: S^<(u,v)
4797: &\stackrel{{\cal C}}{\longrightarrow}&
4798: \gamma^0\gamma^2 {S^>}^T(v,u) \gamma^0\gamma^2
4799: \label{Sc<}
4800: \\
4801: S^<(u,v)
4802: &\stackrel{{\cal CP}}{\longrightarrow}&
4803: - \gamma^2 {S^>}^T(\bar v,\bar u) \gamma^2
4804: \label{Scp<}
4805: \,,
4806: \end{eqnarray}
4807: %
4808: which in the Wigner representation become
4809: %
4810: \begin{eqnarray}
4811: S^<(k,x)
4812: &\stackrel{{\cal C}}{\longrightarrow}&
4813: \gamma^0\gamma^2 {S^>}^T(-k,x) \gamma^0\gamma^2
4814: \equiv
4815: {S^c}^<(k,x)
4816: \label{C_transforms_fermions}
4817: \\
4818: S^<(k, x)
4819: &\stackrel{{\cal CP}}{\longrightarrow}&
4820: \, - \gamma^2 {S^>}^T(-\bar{k}, \bar{x}) \gamma^2
4821: \hphantom{X}
4822: \equiv
4823: {S^{cp}}^<(\bar{k}, \bar{x})
4824: \,;
4825: \label{CP_transforms_fermions}
4826: \end{eqnarray}
4827: %
4828: analogous relations hold for ${S^{c}}^>$ and ${S^{cp}}^>$.
4829: As we did in the scalar case
4830: in~(\ref{scalar_wigner_C}--\ref{scalar_wigner_CP}),
4831: we employ an additional inversion of
4832: the spatial parts of position $\bar x^\mu = (x^0,-x^i)$
4833: and momentum $\bar k^\mu = (k^0,-k^i)$ in the definition of $S^{cp}$.
4834: In the case of several mixing fermions, the Wigner function is a matrix
4835: in flavor space which is affected by the transposition, too.
4836:
4837:
4838:
4839:
4840:
4841: \vskip 0.1in
4842:
4843:
4844: We have argued in section~\ref{Reduction to the on-shell limit}
4845: that a weak coupling reduction to the on-shell limit of the equation of
4846: motion~(\ref{Wigner-space:fermionic_eom})
4847: for the fermionic Wigner function results in the equation
4848: %
4849: \begin{equation}
4850: \Big( \kdag
4851: + \frac i2 \deldag_{\!x}
4852: - \left(m_h(x)+i\gamma^5m_a(x)\right)
4853: \mbox{e}^{-\frac i2 \overleftarrow{\del}_{\!x} \cdot \del_k}
4854: \Big) S^{>,<}(k,x)
4855: = {\cal C}_\psi(k,x) \,.
4856: \label{ferm_eom_wigner}
4857: \end{equation}
4858: %
4859: Taking account of the hermiticity property~(\ref{Wigner-space:fermionic_eom}),
4860: hermitean conjugation and transposition lead to
4861: %
4862: \begin{equation}
4863: \Big( \kdag^{\,T}
4864: - \frac i2 \deldag_{\!x}^{\,T}
4865: - \left(m_h^*(x)+i\gamma^5m_a^*(x)\right)
4866: \mbox{e}^{\frac i2 \overleftarrow{\del}_x \cdot \del_k}
4867: \Big)S^{>,<}(k,x)^T
4868: = -\gamma^0{\cal C}^*_{\psi}(k,x)\gamma^0
4869: \,.
4870: \label{ferm_eom_wigner_T}
4871: \end{equation}
4872: %
4873: By commuting $\gamma^0\gamma^2$ through and reversing the sign of the
4874: 4-momentum, we get the following equation for the C-conjugate of the
4875: Wigner function~(\ref{C_transforms_fermions}),
4876: %
4877: \begin{equation}
4878: \Big( \kdag
4879: + \frac i2 \deldag_{\!x}
4880: - \left(m^*_h(x)+i\gamma^5m^*_a(x)\right)
4881: \mbox{e}^{-\frac i2 \overleftarrow{\del}_{\!x} \cdot \del_k}
4882: \Big) S^{c<,>}(k,x)
4883: = \gamma^2{\cal C}^*_\psi(-k,x)\gamma^2
4884: \,.
4885: \label{ferm_eom_wigner_C}
4886: \end{equation}
4887: %
4888: Similarly, by commuting $\gamma^2$ through~(\ref{ferm_eom_wigner_T}),
4889: we arrive at the equation for the CP-conjugate
4890: Wigner function~(\ref{CP_transforms_fermions}):
4891: %
4892: \begin{equation}
4893: \Big( \bar\kdag
4894: + \frac i2 \deldag_{\!\bar x}
4895: - \left(m^*_h(x)-i\gamma^5m^*_a(x)\right)
4896: {\rm e}^{-\frac i2 \overleftarrow{\del}_{\!\bar x}\cdot\partial_{\bar k}}
4897: \Big) S^{cp<,>}(k, x)
4898: = - \gamma^0\gamma^2{\cal C}^*_\psi(-k,x)\gamma^0\gamma^2
4899: \,.
4900: \label{ferm_eom_wigner_CPb}
4901: \end{equation}
4902: %
4903:
4904:
4905: Several remarks are now in order. Note first that
4906: the presence of imaginary elements in the mass matrices,
4907: $m_h \equiv (1/2)(m+m^\dagger)$ and $m_a \equiv (1/2i)(m-m^\dagger)$,
4908: may violate both C and CP symmetry. This is true, of course, provided
4909: the imaginary parts cannot be removed by field redefinitions.
4910: Second, while C is not in general violated
4911: by a (real) antihermitean mass term $m_a$,
4912: CP is violated, provided $m_a=m_a(x)$ is space-time dependent, such that
4913: it cannot be removed by field redefinitions
4914: \footnote{An example where
4915: these types of CP-violation can be of relevance for baryogenesis
4916: has recently been considered in~\cite{GarbrechtProkopecSchmidt:2003},
4917: where a model of {\it coherent baryogenesis} has been constructed, which
4918: is typically operational on grand-unified scales.}. Finally,
4919: Eqs.~(\ref{ferm_eom_wigner_C}-\ref{ferm_eom_wigner_CPb}) illustrate where
4920: in the collision term are potential sources of C and CP-violation.
4921: As expected, both C and CP may be violated in the presence of complex
4922: Yukawa couplings, appearing in collision or mass terms.
4923: This can be seen from the complex conjugation of the collision term,
4924: ${\cal C}_\psi^*$.
4925: Further, the weak interaction vertices contain $P_{L,R}$, which is also
4926: a property of our model Lagrangean~(\ref{Yukawa}).
4927: %~(\ref{lagrangean}-\ref{Yukawa}).
4928: %This can be seen, for example, from the one-loop expressions for
4929: %the self-energies given in Paper~II.
4930: %~(\ref{Pi:ab2}-\ref{Sigma:ab2}).
4931: Now since $\gamma^2 P_{L,R}\gamma^2 = P_{R,L}$ flips chirality, while
4932: $(\gamma^0\gamma^2) P_{L,R}(\gamma^0\gamma^2) = P_{L,R}$ leaves it invariant,
4933: we conclude that the weak interaction vertices violate C,
4934: but they are invariant under CP, as they should.
4935: %
4936: Using the explicit expressions for the collision terms in Paper~II
4937: %~(\ref{coll_ferm:psi0}-\ref{coll_ferm:psi1})
4938: and~(\ref{The 3+1 dimensional (moving) frame}),
4939: one can show that the right hand side of
4940: equation~(\ref{ferm_eom_wigner_CPb}) is equal to ${\cal C}_\psi$, with
4941: all Wigner functions, both scalar and fermionic, replaced by their CP-conjugate
4942: counterparts, and the Yukawa coupling matrices $y$ replaced by $y^*$.
4943: Further details of the analysis of the collision term
4944: %~(\ref{ferm_eom_wigner_CPb})
4945: are left for Paper~II.
4946:
4947:
4948: \vskip 0.1in
4949:
4950:
4951: A careful look at equation~(\ref{ferm_eom_wigner_CPb}) shows that the
4952: kinetic operator for $S^{cp}(k)$ commutes with $P_s(\bar k)$
4953: rather than with $P_s(k)$.
4954: So the spin-diagonal ansatz for the CP-conjugate Wigner function, already
4955: written in the mass diagonal basis, is
4956: %
4957: \begin{equation}
4958: S^{cp}_{d\tt ii}(k,x)
4959: = \sum_s S^{cp}_{sd\tt ii}(k,x)
4960: = \sum_s iP_{s}(\bar{k})
4961: \big[ s\gamma^3\gamma^5 g_{0d\tt ii}^{cps}
4962: -s\gamma^3 g_{3d\tt ii}^{cps}
4963: +\mathbbm{1} g_{1d\tt ii}^{cps}
4964: -i\gamma^5 g_{2d\tt ii}^{cps} \big](k,x)
4965: \label{Sc<b}
4966: \,.
4967: \end{equation}
4968: %
4969: Now there are several ways to obtain the constraint and kinetic equation
4970: for the component function $g_{0d\tt ii}^{cps}$. First, one could repeat
4971: all the steps from the previous section. Since the mass term in the
4972: CP-conjugate equation~(\ref{ferm_eom_wigner_CPb}) is the complex conjugate
4973: of the mass term appearing in the equation for the original
4974: Wigner function~(\ref{ferm_eom_wigner}),
4975: we would begin by carrying out the flavor transformation, this time using the
4976: rotation matrices $\bX^*$ and $\bY^*$, then taking the appropriate spinorial
4977: traces. The antihermitean and hermitean parts would lead to constraint and
4978: kinetic equations for the CP-conjugate component functions, respectively, and
4979: finally to the desired equations for $g_{0d\tt ii}^{cps}$.
4980:
4981: Another possibility is to note that equation~(\ref{ferm_eom_wigner_CPb}) for
4982: $S^{cp}$ is obtained from equation~(\ref{ferm_eom_wigner}) for $S$ by
4983: replacing the mass by its complex conjugate and replacing all
4984: explicit occurrences of the momentum $k$ by $\bar{k}$, as well as
4985: sending all spatial derivatives $\del_x$ to $\del_{\bar{x}}$, and finally
4986: replacing the collision term on the right hand side,
4987: as indicated in Eq.~(\ref{ferm_eom_wigner_CPb}). When we apply exactly
4988: the same changes to the constraint and the kinetic equation for $g_{0d}$,
4989: which includes $\theta\rightarrow-\theta$ and $\Delta_z\rightarrow-\Delta_z$
4990: because of the complex conjugated mass, we find the respective equations
4991: for $g_{0d}^{cp}$.
4992:
4993: The simplest method, however, is based on the observation that, because
4994: of relation~(\ref{CP_transforms_fermions}), the Wigner function $S$ contains
4995: complete information about its CP-conjugate $S^{cp}$, and the same then
4996: holds for the component functions $g_0$ and $g_0^{cp}$ as well.
4997: Relation~(\ref{CP_transforms_fermions}) is covariant under the flavor rotation,
4998: %
4999: \beq
5000: S_d^{cp<}(k, x)
5001: \equiv \bY^* S^{cp<}(k,x) \bX^T
5002: = - \gamma^2 \left( \bY S^>(-k, x) \bX^\dagger \right)^T \gamma^2
5003: = - \gamma^2 {S_d^>}^T(-k, x) \gamma^2
5004: \,,
5005: \eeq
5006: %
5007: so we have to match the decomposition~(\ref{Sc<b}) for $S^{cp<}$ with
5008: %
5009: \beq
5010: - \gamma^2 [S^>_{d\tt ii}(-k, x)]^{\,T}\gamma^2
5011: = \sum_s iP_{s}(-\bar{k})
5012: \big[-s\gamma^3\gamma^5 g_{0d\tt ii}^{>-s}
5013: +s\gamma^3 g_{3d\tt ii}^{>-s}
5014: +\mathbbm{1} g_{1d\tt ii}^{>-s}
5015: +i\gamma^5 g_{2d\tt ii}^{>-s} \big](-k, x)
5016: \label{eom:g2S>Tg2}
5017: \,.
5018: \eeq
5019: %
5020: This implies the identification
5021: %
5022: \beq
5023: g_{0d\tt ii}^{cps<}(k,x) = - g_{0d\tt ii}^{>-s}(-k, x)
5024: \,,
5025: \label{g0cp-g0>}
5026: \eeq
5027: %
5028: where we used $P_s(-k) = P_s(k)$.
5029: Looking at the constraint equation~(\ref{ced:diag0b}) or the spectral
5030: ansatz~(\ref{spectral-solution:fermions>}) for $g_{0d\tt ii}^{>s}$,
5031: we immediately see that the component functions of the CP-conjugate
5032: Wigner function live on the same energy shells $\omega_{si}$ as the ones
5033: of the original Wigner function.
5034: %
5035: In other words, the dispersion relation for antifermions is identical
5036: to that for fermions, a conclusion that can also be reached by considering
5037: the CP-conjugated constraint equations.
5038: %
5039: Therefore we can make the spectral ansatz
5040: %
5041: \begin{equation}
5042: g_{0d\tt ii}^{<cps}(k, x)
5043: = 2\pi\sum_{\pm} \frac{\delta(k_0\mp\omega_{\pm si})}
5044: {2\omega_{\pm si}Z_{\pm si}}
5045: \, |\tilde{k}_0| \, n^{cp}_{si}(k,x)
5046: \,.
5047: \label{spectral-solution:fermions>cp}
5048: \end{equation}
5049: %
5050: Inserting this and the spectral ansatz~(\ref{spectral-solution:fermions>})
5051: for $g^>_0$ into~(\ref{g0cp-g0>}) shows that the particle density at
5052: negative energies is related to the density of antiparticles. For $k_0$
5053: on the positive or negative energy shell we have
5054: %
5055: \begin{equation}
5056: n^{cp}_{si}(k_0, \vec k,x)
5057: = 1 - n_{-si}\big(-k_0,-\vec k,x\big)
5058: \label{ncp_n}
5059: \,.
5060: \end{equation}
5061: %
5062: In order to find the kinetic equation for the CP-conjugate component function,
5063: let us first recall the kinetic equation~(\ref{kinetic_3+1})
5064: for $g_{0d\tt ii}^{s}$ from section~\ref{Kinetic equations}:
5065: %
5066: \begin{equation}
5067: \frac{1}{\tilde{k}_0}\Big[ k\cdot\del
5068: - \frac 12 \Big(\partial_z|m_d|^2_{i}
5069: - \frac{s}{\tilde k_0}
5070: \partial_z\left[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})
5071: \right]\Big)\partial_{k_z} \Big] {g^{s}_{0d}}_{\tt ii}(k,x)
5072: = {{\cal K}^s_{0d}}_{\tt ii} (k,x)
5073: \,.
5074: \label{kinetic_3+1_repeat}
5075: \end{equation}
5076: %
5077: Again we make use of relation~(\ref{g0cp-g0>}) and find
5078: %
5079: \beq
5080: \frac{1}{\tilde{k}_0}\Big[ k\cdot\del
5081: - \frac 12 \Big(\partial_z|m_d|^2_{i}
5082: - \frac{s}{\tilde k_0}
5083: \partial_z\left[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})
5084: \right]\Big)\partial_{k_z} \Big]
5085: g_{0d\tt ii}^{cps}(k, x)
5086: = - {\cal K}^{-s}_{0d\tt ii}(-k, x)
5087: \,.
5088: \label{kinetic_3+1_cp}
5089: \eeq
5090: %
5091: Needless to say, all ways of obtaining these equations yield the same
5092: results, although this might not be obvious for the collision term. Starting
5093: with the right hand side of equation~(\ref{ferm_eom_wigner_CPb}),
5094: one would expect to find
5095: %
5096: \beq
5097: {\cal K}^{cps}_{0d\tt ii}(k, x)
5098: \equiv \frac 12 \mbox{Tr}
5099: \left( \mathbbm{1} P_s(\bar{k})
5100: \gamma^0\gamma^2 {{\cal C}^*_{\psi d}}_{\tt ii}(-k, x)
5101: \gamma^0\gamma^2
5102: + \text{h.c.}
5103: \right)
5104: \label{coll_cp_a}
5105: \eeq
5106: %
5107: as the collisional contribution to the kinetic equation for $g^{cp}_0$.
5108: But since this is a trace in spinor space and we are looking at the
5109: diagonal elements in flavor space, we can apply a transposition to find
5110: %
5111: \beqa
5112: {\cal K}^{cps}_{0d\tt ii}(k, x)
5113: &=& \frac 12 \text{Tr}
5114: \left(
5115: \left(\gamma^0\gamma^2 P_s(\bar{k}) \gamma^0\gamma^2\right)^T
5116: {{\cal C}^\dagger_{\psi d}}_{\tt ii}(-k, x)
5117: + \text{h.c.}
5118: \right)
5119: \nonumber\\
5120: &=& \frac 12 \text{Tr}
5121: \left(
5122: P_{-s}(k)
5123: {{\cal C}_{\psi d}}_{\tt ii}(-k, x)
5124: + P^\dagger_{-s}(k)
5125: {{\cal C}_{\psi d}}_{\tt ii}^\dagger(-k, x)
5126: \right)
5127: = - {\cal K}^{-s}_{0d\tt ii}(-k, x)
5128: \,.
5129: \label{coll_cp_b}
5130: \eeqa
5131: %
5132:
5133:
5134:
5135: We define the distribution function for particles as the projection of the
5136: phase space densities on the positive mass shell,
5137: %
5138: \beq
5139: f_{si+}(\vec{k}, x) \equiv n_{si}(\omega_{si}, \vec{k}, x)
5140: \,,
5141: \label{onshell-density}
5142: \eeq
5143: %
5144: and then the kinetic equation for $g_0^{cp}$ suggests to define the
5145: distribution function for antiparticles as
5146: %
5147: \beq
5148: f_{si-}(\vec{k}, x) \equiv n^{cp}_{si}(\omega_{si}, \vec{k}, x)
5149: = 1 - n_{-si}(-\omega_{si}, -\vec{k}, x)
5150: \,.
5151: \label{onshell-density-cp}
5152: \eeq
5153: %
5154: We will verify in Paper~II
5155: %section~\ref{sec_fluid-currents}
5156: that $f_{si\pm}(\vec{k})$
5157: indeed measure the density of particles and antiparticles with momentum
5158: $\vec{k}$ and spin $s$.
5159: To obtain the Boltzmann equations for these densities
5160: we have to integrate~(\ref{kinetic_3+1_repeat}) and~(\ref{kinetic_3+1_cp})
5161: over positive frequencies:
5162: %
5163: \beqa
5164: \frac{1}{Z_{si}}
5165: \big( \del_t
5166: +\vec v_{si}\cdot \nabla_{\vec x}
5167: + F_{si}\partial_{k_z} \big) f_{si+}
5168: &=& \int _0^\infty \frac{dk_0}{\pi} {{\cal K}^s_{0d}}_{\tt ii}(k,x)
5169: \label{ferm_kin_pos}
5170: \\
5171: \frac{1}{Z_{si}}
5172: \big( \del_t
5173: +\vec v_{si}\cdot \nabla_{\vec x}
5174: + F_{si}\partial_{k_z} \big) f_{si-}
5175: &=& \int _0^\infty \frac{dk_0}{\pi} {{\cal K}^{cps}_{0d}}_{\tt ii}(k, x)
5176: \,,
5177: \label{ferm_kin_pos_cp}
5178: \eeqa
5179: %
5180: with the quasiparticle velocity and semiclassical force
5181: %
5182: \begin{eqnarray}
5183: \vec v_{si} &=& \frac{\vec{k}}{\omega_{si}}
5184: \label{velocity_si+}
5185: \\
5186: F_{si} &=& - \frac{1}{2\omega_{si}}
5187: \Big(\partial_z|m_d|^2_{i}
5188: - \frac{s}{\tilde \omega_{si}}
5189: \partial_z\left[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})
5190: \right]
5191: \Big)
5192: \,.
5193: \label{force_si+}
5194: \end{eqnarray}
5195: %
5196: Note that, in addition to the classical terms,
5197: both $\vec v_{si}$ and $F_{si}$ contain spin dependent
5198: contributions arising from derivatives of the pseudoscalar mass term,
5199: and thus a potential for inducing CP-violating effects.
5200:
5201: The Boltzmann equation~(\ref{ferm_kin_pos_cp}) for the antiparticle densities
5202: can also be obtained by integrating the kinetic equation for $g_0$
5203: over negative energies, and then using~(\ref{onshell-density-cp}).
5204: This procedure has often been advocated as a way of identifying CP-violating
5205: densities.
5206: It works, because the complete information about $S^{<cp}$ is contained in
5207: $S^>$, which in the on-shell limit is completely determined by $S^<$, as
5208: can be shown with the help of the fermionic spectral sum
5209: rule~(\ref{sum_rule_fermions}).
5210:
5211: \vskip 0.1in
5212:
5213: For practical baryogenesis calculations one often deals with
5214: problems close to thermal equilibrium, in which case it is convenient to treat
5215: the Boltzmann equations~(\ref{ferm_kin_pos}) and~(\ref{ferm_kin_pos_cp})
5216: in linear response approximation with respect to deviations from
5217: thermal equilibrium. We note that the thermal equilibrium distribution
5218: function in the wall frame has the form
5219: %
5220: \begin{equation}
5221: f^{\rm eq}_{si} = \frac{1}{e^{\beta \gamma_w(\omega_{si} + v_w k_z)}+1},
5222: \label{distribution-fn-eq}
5223: \end{equation}
5224: %
5225: with $\beta =1/T$, $\vec v_w = v_w \hat z$ is the wall velocity
5226: and $\gamma_w = (1-v_w^2)^{-1/2}$ is the corresponding boost factor.
5227: This is a generalization of the thermal equilibrium distribution
5228: function~(\ref{FermiDirac}) which includes both the effects of CP-violation
5229: from a pseudoscalar mass and of a moving plasma on thermal equilibrium.
5230: We show in Paper~II
5231: %section~\ref{Collision term}
5232: that this form for the equilibrium
5233: distribution function leads to the correct energy conservation
5234: in the collision term. Furthermore, the distribution
5235: function~(\ref{distribution-fn-eq}) satisfies
5236: %
5237: \begin{equation}
5238: Z_{si}^{-1}\big(\partial_t + \vec v_{si}\cdot\nabla_{\vec x}
5239: + F_{si}\partial_{k_z}\big) f^{\rm eq}_{si}
5240: = - v_w f^{\rm eq}_{si}(1-f^{\rm eq}_{si})Z_{si}^{-1}\beta F_{si}
5241: \label{distribution-fn-eq_eom}
5242: \end{equation}
5243: %
5244: such that, for a wall at rest, Eq.~(\ref{distribution-fn-eq}) satisfies
5245: the flow equation exactly, as one would expect from the correct equilibrium
5246: distribution function. Note that the degeneracy between the states
5247: with opposite spin is broken by this equilibrium,
5248: because particles with opposite spin satisfy different
5249: energy-momentum dispersion relations.
5250:
5251:
5252:
5253:
5254: Our main interest is the transport equation for a CP-violating
5255: distribution function that can eventually lead to the creation of
5256: a Baryon asymmetry. We first define
5257: %
5258: \beq
5259: f_{si\pm} = f^{\rm eq}_{si}
5260: + \frac 12 \delta f_i^{\rm even}
5261: + \delta f_{si\pm}
5262: \,,
5263: \label{distribution-fn_fsi_decompose}
5264: \eeq
5265: %
5266: where $\delta f_{si\pm}$ are spin dependent correction functions.
5267: Since spin dependence occurs only at order $\hbar$ in our Boltzmann equations
5268: and all sources are wall velocity suppressed, as for example can be seen
5269: in~(\ref{distribution-fn-eq_eom}), we find $\delta f_{si\pm} = O(v_w\hbar)$.
5270: %
5271: The distributions $f_{si\pm}$ can also contain a CP-even deviation from thermal
5272: equilibrium, $\delta f_{i}^{\rm even}$, which is relevant for the dynamics
5273: of phase
5274: interfaces~\cite{MooreProkopec:1995,HuberJohnSchmidt:2001,JohnSchmidt:2000+2001}.
5275: %
5276: Note that, since $\delta f_i^{\rm even}$ is not suppressed by powers of
5277: $\hbar$, it does not in general decouple from the equation for the
5278: densities $\delta f_{si\pm}$. In order to properly treat the dynamics
5279: of CP-violating densities, one has to solve also for the dynamics of
5280: CP-conserving densities!
5281: %
5282: Together with $\delta f_{i}^{\rm even} = O(v_w\hbar^0)$,
5283: Eq.~(\ref{distribution-fn-eq_eom}) implies that it suffices to
5284: solve the equation for $\delta f^{\rm even}_i$ to leading (classical)
5285: order in gradients, which is obtained by
5286: adding~(\ref{ferm_kin_pos}) and~(\ref{ferm_kin_pos_cp}),
5287: %
5288: \begin{equation}
5289: \big(\del_t + \vec v_{0i}\cdot\nabla_{\vec x}
5290: + F_{0i} \partial_{k_z}
5291: \big) \delta f_{i}^{\rm even}
5292: + \beta f_0(1-f_0) v_w \frac{\del_z|m_d|^2_i}{\omega_{0i}}
5293: = \int _0^\infty \frac{dk_0}{\pi}
5294: \left( {\cal K}^s_{0d\tt ii}(k) + {\cal K}^{cps}_{0d\tt ii}(k) \right)
5295: \,,
5296: \label{eom:delta_fi}
5297: \end{equation}
5298: %
5299: where
5300: %
5301: \begin{equation}
5302: \vec v_{0i} = \frac{\vec k}{\omega_{0i}}
5303: \,,\qquad
5304: F_{0i} = - \frac{\partial_z |m_d|^2_{i}}{2\omega_{0i}}
5305: \label{v0i:F0i}
5306: \end{equation}
5307: %
5308: are the classical particle velocity and the classical force, respectively,
5309: and $f_{0i} \equiv 1/(e^{\beta\omega_{0i}}+1)$.
5310:
5311:
5312:
5313: Based on the first order correction $\delta f_{si\pm}$, we can
5314: form two CP-violating distribution functions:
5315: %
5316: \beqa
5317: \delta f^v_{si} &\equiv& \delta f_{si+} - \delta f_{si-}
5318: \label{distribution-fv}
5319: \\
5320: \delta f^a_{si} &\equiv& \delta f_{si+} - \delta f_{-si-}
5321: \,.
5322: \label{distribution-fa}
5323: \eeqa
5324: %
5325: As we will explain in Paper~II,
5326: % section~\ref{sec_fluid-currents},
5327: $\delta f^v_{si}$ and $\delta f^a_{si}$ are related to the vector and axial vector
5328: density in phase space, respectively.
5329: Working to second order in gradients and to linear
5330: order in $v_w$, we obtain the equation for $\delta f^v_{si}$
5331: by subtracting~(\ref{ferm_kin_pos_cp}) from~(\ref{ferm_kin_pos}):
5332: %
5333: \beq
5334: \big( \del_t
5335: + \vec v_{0i} \cdot \nabla_{\vec x}
5336: + F_{0i} \del_{k_z}
5337: \big) \delta f^v_{si}
5338: = \int _0^\infty \frac{dk_0}{\pi}
5339: \left( {\cal K}^s_{0d\tt ii}(k) - {\cal K}^{cps}_{0d\tt ii}(k) \right)
5340: \,.
5341: \label{boltzmann:deltaf-v}
5342: \eeq
5343: %
5344: Note that the flow term contains {\it no source} whatsoever.
5345: This means that a CP-violating semiclassical force~(\ref{force_si+})
5346: cannot create vector charge densities
5347: (this remains no longer true, however, when one includes the possibility of
5348: coherent particle production, which can be exacted by a dynamical tracing
5349: of flavor mixing~\cite{GarbrechtProkopecSchmidt:2003}).
5350:
5351: \vskip 0.05in
5352:
5353: The transport equation for $\delta f^a_{si}$ is obtained
5354: by changing the sign of $s$ in~(\ref{ferm_kin_pos_cp}) and then subtracting
5355: it from~(\ref{ferm_kin_pos}). With the help
5356: of~(\ref{distribution-fn-eq_eom}) we find
5357: %
5358: \beq
5359: \big( \del_t
5360: + \vec v_{0i} \cdot \nabla_{\vec x}
5361: + F_{0i} \partial_{k_z}
5362: \big)
5363: \delta f^a_{si}
5364: + s{\cal S}_i^{\rm flow}
5365: = \int _0^\infty \frac{dk_0}{\pi}
5366: \left( {\cal K}^s_{0d\tt ii}(k) - {\cal K}^{cp-s}_{0d\tt ii}(k) \right)
5367: \,.
5368: \label{boltzmann:deltaf-a}
5369: \eeq
5370: %
5371: This equation now has a source:
5372: %
5373: \beqa
5374: {\cal S}_i^{\rm flow}
5375: &=& - \frac 12 v_w f_{0i}(1-f_{0i})\beta
5376: \Big\{ \delta F_i
5377: + F_{0i}\frac{\delta\omega_i}{\omega_{0i}}
5378: + F_{0i}\delta\omega_i\beta(1-2f_{0i})
5379: + F_{0i}\delta Z_i
5380: \Big\}
5381: \label{boltzmann:deltaf-a_source}
5382: \\
5383: &&
5384: + \frac 12 \Big\{
5385: \delta Z_i
5386: \big( \del_t
5387: + \vec v_{0i}\cdot\nabla_{\vec x}
5388: + F_{0i} \partial_{k_z}
5389: \big)
5390: + \frac{\delta\omega_i}{\omega_{0i}}
5391: \vec v_{0i} \cdot \nabla_{\vec x}
5392: + \Big[F_{0i}\frac{\delta\omega_i}{\omega_{0i}}
5393: + \delta F_i \Big] \partial_{k_z}
5394: \Big\}
5395: \delta f_{i}^{\rm even}
5396: \,,
5397: \label{boltzmann:deltaf-a_even}
5398: \quad\quad
5399: \end{eqnarray}
5400: %
5401: where the definitions
5402: %
5403: \beqa
5404: \delta F_i &\equiv&
5405: \frac{\partial_z[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})]}
5406: {\omega_{0i}\tilde\omega_{0i}}
5407: \nonumber\\
5408: \delta\omega_i
5409: &\equiv& \frac{[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})]}
5410: {\omega_{0i}\tilde\omega_{0i}}
5411: \quad,\quad
5412: \delta Z_i
5413: \equiv \frac{\omega_{0i}}{\tilde\omega_{0i}^2} \, \delta\omega_i
5414: %\nonumber\\
5415: % \delta Z_i
5416: % &\equiv& \frac{[|m_d|^2(\partial_z\theta_d+2\Delta_z)]_{ii}}
5417: % {\tilde\omega^3_{0i}}
5418: \label{boltzmann:definitions}
5419: \eeqa
5420: %
5421: have been used.
5422: We now pause to comment on the physical meaning of the various terms appearing
5423: in the Boltzmann transport
5424: equation~(\ref{boltzmann:deltaf-a}).
5425: First, the CP-violating density, $\delta f_{si}^a$, evolves on phase space
5426: according to the standard Boltzmann flow derivative,
5427: $d/dt \equiv \del_t + \vec v_{0i}\cdot\nabla_{\vec x} + F_{0i} \partial_{k_z}$.
5428: Second, various CP-violating sources are displayed
5429: in~(\ref{boltzmann:deltaf-a_source}).
5430: The first term, $\delta F_i$, is the source
5431: arising from the semiclassical force that has been usually taken
5432: into account in the
5433: WKB-literature~\cite{JoyceProkopecTurok:1996,ClineJoyceKainulainen:2000+2001}.
5434: Note that the second term has a similar origin.
5435: Indeed, it arises from the CP-violating deviation in the energy-momentum
5436: relation appearing in the force term, $-(\partial_z |m_d|_i)/2\omega_{si}$.
5437: The third term originates from the CP-violating split in the dispersion
5438: relation in the equilibrium solution~(\ref{distribution-fn-eq}),
5439: and it thus vaguely resembles a local version of spontaneous baryogenesis.
5440: And finally, the fourth term arises from the gradient ``renormalization''
5441: of the Wigner function. All four source terms are of the same order
5442: in gradients, and hence, {\it a priori}, they are all equally important.
5443: %
5444: In Paper~II
5445: % section~\ref{Fluid equations indeed}
5446: we take moments of the
5447: Boltzmann equation~(\ref{boltzmann:deltaf-a}) in order to obtain
5448: fluid equations. Since the source term~(\ref{boltzmann:deltaf-a_source})
5449: is symmetric in momentum, it will contribute to the zeroth moment equation.
5450: The parametric dependence of this source becomes explicit by
5451: rewriting it like
5452: %
5453: \beqa
5454: \int \frac{d^3k}{(2\pi)^3} {\cal S}_i^{\rm flow}
5455: &=& v_w \Big( - \partial_z[|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})]
5456: \frac{{\cal I}_a}{(2\pi)^2}
5457: \nonumber\\
5458: &&\hphantom{XX}
5459: + \beta^2 \left(\del_z|m_d|^2_i\right)
5460: [|m_d|^2_i(\partial_z\theta_{di}+2{\Delta_z}_{\tt ii})]
5461: \frac{{\cal I}_b}{(2\pi)^2}
5462: \Big)
5463: \,.
5464: \label{parametric-flow-source}
5465: \eeqa
5466: %
5467: In figure~\ref{figure:flow-source} we plot the dimensionless integrals
5468: ${\cal I}_a$ and ${\cal I}_b$ as a function of the mass.
5469: When $|m_d|_i \ll T$ ($|m_d|_i \gg T$), the source ${\cal I}_a$ (${\cal I}_b$)
5470: dominates. The sources are comparable in strength when $|m_d|_i \sim T$.
5471: %
5472: %
5473: \begin{figure}[t]
5474: \unitlength=1in
5475: \begin{center}
5476: \psfrag{Ia}[r]{$\beta^2|m|^2 {\cal I}_a$}
5477: \psfrag{Ib}[r]{$\beta^4|m|^4 {\cal I}_b$}
5478: \psfrag{m}{$\hphantom{XXXXXXXXXXXX}\beta|m|\rightarrow$}
5479: \includegraphics[width=3.5in,angle=-90]{A_pics/IaIb.eps}
5480: \end{center}
5481: \lbfig{figure:flow-source}
5482: \caption{
5483: The integrals ${\cal I}_a$ and ${\cal I}_b$ in Eq.~(\ref{parametric-flow-source})
5484: as a function of the mass.
5485: We scaled ${\cal I}_a$ with $|m|^2$ because this is the way it appears
5486: in the source
5487: term, analogously ${\cal I}_b$ is scaled with $|m|^4$. Note that these
5488: contributions enter the flow source with different signs.
5489: }
5490: \end{figure}
5491: %
5492: %
5493: Third, the terms~(\ref{boltzmann:deltaf-a_even})
5494: that couple the CP-even deviation form equilibrium have never been considered
5495: in literature. Nevertheless, they are formally of the same order as
5496: the source terms~(\ref{boltzmann:deltaf-a_source}), and thus cannot be
5497: neglected. While the first term in~(\ref{boltzmann:deltaf-a_even})
5498: can be reexpressed in terms of the sum of the collision terms, as
5499: indicated by Eq.~(\ref{eom:delta_fi}), the latter terms cannot,
5500: and represent genuine dynamical source terms, which have so far not
5501: been included in baryogenesis calculations. Fourth, the
5502: right-hand-side of Eq.~(\ref{boltzmann:deltaf-a}) contains
5503: the collision terms, to be considered in detail in Paper~II.
5504: %section~\ref{Collision term}.
5505: Finally, we remind the reader that all of the
5506: source terms in Eq.~(\ref{boltzmann:deltaf-a})
5507: are second order in derivatives (first order in
5508: $\hbar$), and are suppressed by the wall velocity.
5509:
5510:
5511:
5512:
5513:
5514:
5515:
5516:
5517:
5518:
5519:
5520:
5521:
5522:
5523:
5524:
5525: \subsection{Applications}
5526:
5527: In this section we have developed a formalism for the treatment of
5528: CP-violating effects of space-time dependent scalar and pseudoscalar
5529: mass terms in kinetic theory. We have shown how the effect of a CP-violating
5530: shift in the dispersion relation induces an order $\hbar$ semiclassical
5531: force in the flow term of the Boltzmann transport equation. The force
5532: acts on particles and antiparticles of opposite spin in opposite directions.
5533: Since we have included the possibility of fermionic mixing,
5534: this formalism is suitable for studies of baryogenesis in supersymmetric
5535: theories at a strongly first order electroweak phase transition,
5536: in which baryon production is typically biased by CP-violating
5537: chargino or neutralino currents. Further, our formalism is also suitable for
5538: studying baryogenesis problems from coupling of fermions to the bubble wall
5539: in two Higgs doublet models. For completeness we shall now discuss
5540: in some detail the case of charginos in the Minimal Supersymmetric
5541: Standard Model (MSSM) and its nonminimal extension, and finally remark
5542: on baryogenesis sources in two Higgs doublet models.
5543:
5544: \subsubsection{Minimal Supersymmetric Standard Model (MSSM)}
5545: \label{Minimal Supersymmetric Standard Model (MSSM)}
5546:
5547: Now that we have a general expression for the semiclassical source in the case of
5548: mixing fermions, we want to study two explicit examples, which are of relevance
5549: for baryogenesis. First we compute the source in the transport equations for the
5550: chargino sector of the MSSM. The chargino mass term reads
5551: %
5552: \begin{equation}
5553: {\overline\psi_R}\, m \, \psi_L +{\rm h.c.}
5554: \,,
5555: \label{chargino-mass}
5556: \end{equation}
5557: %
5558: where $\psi_R = ({\tilde W}_R^+,{\tilde h}^+_{1,R})^T$
5559: and $\psi_L = ({\tilde W}_L^+,{\tilde h}^+_{2,L})^T$
5560: are the chiral fields in the basis of winos.
5561: The mass matrix is
5562: %
5563: \begin{equation}
5564: m = \left( \begin{array}{cc} m_2 & gH_2^* \\
5565: gH_1^* & \mu \end{array} \right)
5566: \,,
5567: \label{CharginoMatrix_MSSM}
5568: \end{equation}
5569: %
5570: where $H_1$ and $H_2$ are the Higgs field vacuum expectation values and
5571: $\mu$ and $m_2$ are the soft supersymmetry breaking parameters, which introduce
5572: CP-violation~\footnote{We keep the possibility of complex Higgs vacuum
5573: expectation values, because
5574: we will reuse the formulas in the NMSSM case.}.
5575: Since for a reasonable choice of parameters there is no transitional
5576: CP-violation in the MSSM, we can take the Higgs expectation values
5577: $H_1$ and $H_2$ to be real
5578: \cite{HuberJohnSchmidt:2001,HuberJohnLaineSchmidt:2000}.
5579: The matrix that diagonalizes $mm^\dagger$ can be parametrized
5580: as~\cite{ClineJoyceKainulainen:2000+2001}
5581: %
5582: \begin{equation}
5583: U = \frac{\sqrt{2}}{\sqrt{\Lambda(\Lambda+\Delta)}}
5584: \left( \begin{array}{cc} \frac 12 (\Lambda+\Delta) & a \\
5585: -a^* & \frac 12 (\Lambda+\Delta)
5586: \end{array} \right) \,,
5587: \label{Rotation_MSSM}
5588: \end{equation}
5589: %
5590: where
5591: %
5592: \begin{eqnarray}
5593: a &=& g(m_2H_1 + \mu^*H_2^*) \nonumber\\
5594: \Delta &=& |m_2|^2 - |\mu|^2 + g^2(h_2^2 - h_1^2) \nonumber\\
5595: \Lambda &=& \sqrt{\Delta^2 + 4|a|^2} \,,
5596: \label{Rotation_MSSM_parameters}
5597: \end{eqnarray}
5598: %
5599: and $h_i = |H_i|$.
5600: The mass eigenvalues of the charginos are given by
5601: %
5602: \begin{equation}
5603: m_\pm^2 = \frac 12 \left( |m_2|^2 + |\mu|^2 + g^2(h_1^2+h_2^2) \right)
5604: \pm \frac{\Lambda}{2} \,.
5605: \label{chargino_eigenstates}
5606: \end{equation}
5607: %
5608: Upon inserting (\ref{CharginoMatrix_MSSM}-\ref{chargino_eigenstates})
5609: into~(\ref{shift:U}), it is straightforward to show
5610: that the chargino source term can be recast as
5611: %
5612: \begin{equation}
5613: \left[|m_d|^2(\partial_z\theta+2\Delta_z)\right]_{\pm}
5614: = \mp \frac{g^2}{\Lambda} \Im(\mu m_2)
5615: \partial_z\left(h_1h_2\right)
5616: \,.
5617: \label{chargino_source}
5618: \end{equation}
5619: %
5620: The sources figuring in the transport equation written for
5621: charginos in~(\ref{boltzmann:deltaf-a_source}) can be easily reconstructed
5622: from equations~(\ref{Rotation_MSSM_parameters}) and~(\ref{chargino_source}).
5623: The result~(\ref{chargino_source}) agrees with the one found
5624: by WKB methods in~\cite{ClineJoyceKainulainen:2000+2001}.
5625: In~\cite{CarenaMorenoQuirosSecoWagner:2000,CarenaQuirosSecoWagner:2002}
5626: however a different dependence on the Higgs fields was obtained.
5627:
5628:
5629: \subsubsection{Nonminimal Supersymmetric Standard Model (NMSSM)}
5630: \label{Nonminimal Supersymmetric Standard Model (NMSSM)}
5631:
5632: We now consider an extension of the MSSM which contains a
5633: singlet field $S$ in the Higgs sector, which can induce additional
5634: CP-violation in the Higgs sector. In particular,
5635: the Higgs vacuum expectation values may become complex.
5636: The singlet couples to higgsinos, and therefore we obtain the mass matrix
5637: by generalizing the higgsino-higgsino component of the chargino mass
5638: matrix~(\ref{CharginoMatrix_MSSM})
5639: %
5640: \begin{equation}
5641: \mu \rightarrow \tilde{\mu} = \mu + \lambda S \,,
5642: \label{MSSM_to_NMSSM}
5643: \end{equation}
5644: %
5645: where $\lambda$ is the coupling for the higgsino-higgsino-singlet interaction.
5646: The field content we consider is the same as in the MSSM, so the mass matrix is
5647: %
5648: \begin{equation}
5649: m = \left( \begin{array}{cc} m_2 & gH^*_2 \\
5650: gH^*_1 & \tilde{\mu} \end{array} \right) \,.
5651: \label{CharginoMatrix_NMSSM}
5652: \end{equation}
5653: %
5654: This matrix is still diagonalized by $U$ in~(\ref{Rotation_MSSM}).
5655: We write the Higgs expectation values as
5656: %
5657: \begin{equation}
5658: H_i = h_i{\rm e}^{i\theta_i} \quad,\quad i=1,2 \,,
5659: \end{equation}
5660: %
5661: where only one phase is physical. With the gauge
5662: constraint~\cite{HaberKane:1985}
5663: %
5664: \begin{equation}
5665: h^2_1\theta'_1 = h^2_2\theta'_2
5666: \end{equation}
5667: %
5668: we can write
5669: %
5670: \begin{equation}
5671: \theta_1' = \frac{h_2^2}{h^2}\theta'
5672: \quad,\quad
5673: \theta_2' = \frac{h_1^2}{h^2}\theta' \,,
5674: \end{equation}
5675: %
5676: where $\theta=\theta_1+\theta_2$ is the physical CP-violating phase,
5677: and $h^2=h_1^2+h_2^2$.
5678: Now everything is prepared to write the NMSSM-source term. It can
5679: be divided into three contributions, which have to be added. The first
5680: one is a generalization of the chargino source~(\ref{chargino_source})
5681: %
5682: \begin{equation}
5683: \left(|m_d|^2(\partial_z\theta_d+2\Delta_z)\right)_{h_1h_2\pm}
5684: = \mp \frac{g^2}{\Lambda} \Im (\tilde{\mu}m_2{\rm e}^{i\theta}) (h_1h_2)'
5685: \end{equation}
5686: %
5687: for the case involving a new scalar field $S$ and complex Higgs expectation values.
5688: In addition to this there are two new types of sources. One of them is
5689: proportional to a derivative of the CP-violating phase $\theta$ in the Higgs sector:
5690: %
5691: \begin{equation}
5692: \left(|m_d|^2(\partial_z\theta_d+2\Delta_z)\right)_{\theta\pm}
5693: = -\frac{g^2\theta'}{\Lambda}
5694: \bigg(
5695: \Big( \Lambda \pm (|m_2|^2+|\tilde{\mu}|^2)\Big)
5696: \frac{h_1^2h_2^2}{h^2}
5697: \mp \Re (\tilde{\mu}m_2{\rm e}^{i\theta})h_1h_2
5698: \bigg) \,.
5699: \end{equation}
5700: %
5701: Finally, there is a source that can be written as a derivative of the
5702: singlet condensate:
5703: %
5704: \begin{eqnarray}
5705: \left(|m_d|^2(\partial_z\theta_d+2\Delta_z)\right)_{S\pm}
5706: &=& \pm \frac{\lambda g^2}{\Lambda} \Im(m_2H_1H_2S') \\
5707: && + \frac{\lambda g^2}{2\Lambda}
5708: \Big( \Lambda \pm (|\tilde{\mu}|^2 + g^2h^2 - |m_2|^2) \Big)
5709: \Im(\tilde{\mu}^*S') \,.
5710: \nonumber
5711: \end{eqnarray}
5712: %
5713: The mass eigenvalues $m_\pm$, that is the diagonal elements of $|m_d|^2$,
5714: can be obtained from the corresponding expression~(\ref{chargino_eigenstates})
5715: in the MSSM part with the replacement $\mu\rightarrow\tilde{\mu}$.
5716:
5717:
5718:
5719:
5720:
5721:
5722:
5723:
5724:
5725:
5726:
5727:
5728:
5729:
5730:
5731:
5732:
5733:
5734:
5735:
5736:
5737:
5738:
5739:
5740:
5741:
5742: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5743: % %
5744: % %
5745: % OOOO OOO O O OOOO O O O OOOO O OOO O O %
5746: % O O O OO O O O O O O O O O OO O %
5747: % O O O O O O O O O O OOO O O O O O O %
5748: % O O O O OO O O O O O O O O O OO %
5749: % OOOO OOO O O OOOO OOOOO OOO OOOO O OOO O O %
5750: % %
5751: % %
5752: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5753:
5754: \cleardoublepage
5755: \section{Conclusion}
5756:
5757:
5758: In this work we perform a controlled first principle derivation of
5759: transport equations for a model Lagrangean of chiral fermions Yukawa-coupled
5760: to a complex scalar. Our treatment is accurate
5761: to first order in an $\hbar$ expansion, or, more concretely, to
5762: first order in a gradient expansion with respect to a slowly varying scalar
5763: background field, which is formally valid when
5764: $\hbar \partial \ll \hbar k$. Being valid to order $\hbar$,
5765: our treatment allows us to trace the propagation of CP-violating fluxes,
5766: which is of relevance, for example, for electroweak baryogenesis
5767: at a first order phase transition. We consistently include the collision term,
5768: although in this first paper of our series only in a formal way.
5769: The actual evaluation of these terms can be found in Paper~II.
5770:
5771: \vskip 0.05in
5772:
5773: We address an open question of electroweak scale
5774: baryogenesis mediated by mixing fermions, which couple in a CP-violating
5775: manner to a propagating bubble wall of a first order phase transition:
5776: which basis should be used to model the kinetics of mixing fermions?
5777: %
5778: We show that, if one is limited to a diagonal approximation,
5779: the mass eigenbasis is singled out as the only basis in which
5780: the diagonal and off-diagonal elements of the distribution function
5781: {\it decouple} at order $\hbar$ in a derivative expansion.
5782: Our derivation is valid when there are no nearly degenerate mass
5783: eigenvalues, that is when $\hbar k\cdot\partial\ll \delta(m_d^2)$,
5784: where $\delta(m_d^2)$ denotes the (minimum) split in the mass eigenvalues.
5785: No such claims can be made for the flavor (weak interaction) basis,
5786: in which flavor mixing is present already at the classical level $O(\hbar^0)$.
5787: This indicates that the use of the flavor basis in transport equations
5788: is at best problematic, unless flavor mixing is consistently included.
5789: %
5790: Of course, the final resolution of this problem can only come from
5791: a basis independent treatment, which would include the dynamics of
5792: both flavor off-diagonal and diagonal CP-violating densities.
5793: At the moment no such treatment is available, however.
5794:
5795:
5796:
5797:
5798:
5799:
5800: \vskip 0.05in
5801:
5802: We also address various other issues, which comprise a proof that
5803: the Kadanoff-Baym equation for a single scalar field can be reduced, in a
5804: weakly coupled regime, to an on-shell Boltzmann equation, which includes both
5805: the self-energy and the collision term, approximated at the same order in a
5806: coupling constant expansion.
5807: %
5808: An inclusion of the hermitean self-energies would be desirable, but has not
5809: been done yet.
5810: %
5811: These self-energies mix spin and moreover provide an additional source
5812: of CP-violation, which we expect to be of a similar strength as the source
5813: in the collision term (see Paper~II).
5814: %
5815: Further, we demonstrate that no CP-violating source is present at order
5816: $\hbar$ in the flow term of the scalar kinetic equation.
5817: %
5818: We then derive a Boltzmann
5819: transport equation for the relevant fermionic quasiparticles with
5820: a definite spin.
5821: We include all of the CP-violating sources arising from the flow term
5822: present at the order $\hbar$, which, of course, include the semiclassical
5823: force, but also a few, as-of-yet unencountered, sources. One of the new
5824: sources is induced by a CP-even deviation from the equilibrium distribution
5825: function, which itself is formally of order $\hbar^0$.
5826:
5827: \vskip 0.05in
5828:
5829: Our method represents a formalized and controlled implementation
5830: of the original heuristic semiclassical (WKB) treatment
5831: of the problem~\cite{JoyceProkopecTurok:1995,JoyceProkopecTurok:1996},
5832: in which one calculates the relevant (fermionic) dipersion relation
5833: accurate to order $\hbar$ by the means of a WKB analysis, and then inserts
5834: the result in the appropriate kinetic equation.
5835: As we have shown here, when applied with due care,
5836: the semiclassical approach leads to a correct semiclassical force
5837: in the flow term, but it does not permit the treatment of collisional
5838: sources, nor does it allow a critical assessment of the quasiparticle picture
5839: of the plasma or the implementation of any effects beyond the quasiparticle
5840: picture. None of these limitations apply for the approach presented here.
5841:
5842:
5843: \vskip 0.05in
5844:
5845: Our work is completed in Paper~II. There we
5846: discuss the collision term, which in the present paper was
5847: formally maintained in all equations, but never evaluated explicitly,
5848: and we derive and study a set of fluid equations, which form
5849: an approximation to the Boltzmann equations found here.
5850:
5851:
5852:
5853:
5854:
5855:
5856:
5857:
5858:
5859: \vskip 0.1in
5860:
5861: \section*{Acknowledgements}
5862: We would like to acknowledge collaboration and engaged discussions
5863: with Kimmo Kainulainen in earlier stages of this project.
5864: In particular, the results in section~\ref{Reduction to the on-shell limit}
5865: have been largely developed based on work with Kimmo Kainulainen.
5866: We would also like to thank Thomas Konstandin for constructive discussions.
5867: The work of SW was supported by the U.S. Department of Energy
5868: under Grant No. DE-AC02-98CH10886.
5869: SW also thanks the Alexander von Humboldt Foundation
5870: for support by a Feodor Lynen Fellowship.
5871:
5872:
5873:
5874:
5875:
5876:
5877:
5878:
5879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5880: % %
5881: % %
5882: % OOO OOOO OOOO OOOOO O O OOOO O O O %
5883: % O O O O O O O OO O O O O O O %
5884: % OOOOO OOOO OOOO OOO O O O O O O O %
5885: % O O O O O O OO O O O O O %
5886: % O O O O OOOOO O O OOOO O O O %
5887: % %
5888: % %
5889: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5890:
5891:
5892:
5893:
5894: \begin{appendix}
5895:
5896:
5897:
5898: \cleardoublepage
5899:
5900: \section{Gradient expansion in the off-diagonal scalar kinetic equation}
5901: \label{Gradient expansion in the off-diagonal scalar kinetic equation}
5902:
5903: We shall now prove that, provided
5904: $k\cdot\partial, (\partial \bar M_d^2)\cdot \partial_k \ll \delta(M_d^2)$,
5905: gradient expansion applies for the off-diagonal equation.
5906: Equation~(\ref{scalars-ke-12}) is of the form
5907: %
5908: \begin{equation}
5909: \Big(k\cdot\partial + \frac 12(\partial \bar M_d^2)\cdot \partial_k
5910: + \frac i2 \delta(M_d^2) - ik\cdot \delta(\Xi)\Big)
5911: i\Delta_{12}^< = {\cal S}_{12},
5912: \label{scalars-ke-12-2}
5913: \end{equation}
5914: %
5915: where ${\cal S}_{12} = {\cal S}_{12}[\Delta_d]$ represents the source
5916: composed of the diagonal elements. This equation can be solved by using
5917: the method of Green functions as follows. Consider for simplicity
5918: the following problem
5919: %
5920: \begin{equation}
5921: \big(k\cdot\partial + \frac i2 \delta(M_d^2)+\epsilon
5922: \big)G_\epsilon^r(k;x,x')
5923: = \delta^4(x-x'),
5924: \label{Green-function:eom}
5925: \end{equation}
5926: %
5927: which is solved by the retarded Green function
5928: %
5929: \begin{equation}
5930: G_\epsilon^r(k;x,x') = \frac{1}{k_0}\theta(t-t')
5931: \delta^3\Big(\vec x-\vec x' - \frac{\vec k}{k_0}(t-t')\Big)
5932: \exp\Big({-\Big[\frac i2 \frac{\delta(M_d^2)}{k_0}+\epsilon\Big](t-t')}\Big)
5933: ,
5934: \label{Green-function:eom2}
5935: \end{equation}
5936: %
5937: where $\epsilon$ represents an (infinitesimal) positive dissipation term.
5938: Formally, this solution contains rapid oscillations, since the exponent
5939: varies at the scale $1/\hbar$, which characterizes the off-diagonal elements.
5940: One may wonder what happened to these intrinsically quantum oscillations in
5941: our solution~(\ref{Delta12:solution}) and~(\ref{scalars-ke-12-solution}).
5942: In order to answer this question we consider the solution
5943: to~(\ref{scalars-ke-12-2})
5944: %
5945: \begin{eqnarray}
5946: i\Delta_{12}(k,x) &=& \int d^4 x' G_\epsilon^r(k;x,x') {\cal S}_{12}(k,x')
5947: \nonumber\\
5948: &=& \int ^t dt' {\cal S}_{12}\Big(k,\vec x-\frac{\vec k}{k_0}(t-t'),t'\Big)
5949: \frac{1}{k_0}
5950: \exp\Big({-\Big[\frac i2 \frac{\delta(M_d^2)}{k_0}+\epsilon\Big](t-t')}\Big)
5951: .
5952: \label{scalars-ke-12-solutionB}
5953: \end{eqnarray}
5954: %
5955: The shift in the position, $\delta \vec x = \vec v(t-t')$, corresponds to
5956: the retardation for a particle moving with the velocity $\vec v = \vec k/k_0$.
5957: A similar retardation shift results when the $\partial_k$-derivative
5958: is included in~(\ref{scalars-ke-12-2}-\ref{Green-function:eom}).
5959: Since ${\cal S}_{12}$ is varying very slowly when compared with the
5960: oscillatory term, we can expand around $t'=t$,
5961: %
5962: \begin{equation}
5963: {\cal S}_{12}\Big(k,\vec x-\frac{\vec k}{k_0}(t-t'),t'\Big)
5964: = {\cal S}_{12}(k,x)
5965: + (t'-t)\Big(\partial_t+\frac{\vec k}{k_0}\cdot\nabla\Big){\cal S}_{12}(k,x)
5966: + O(\partial^2)
5967: \end{equation}
5968: %
5969: and integrate~(\ref{scalars-ke-12-solutionB}) to obtain
5970: %
5971: \begin{eqnarray}
5972: i\Delta_{12}(k,x) &=& - \frac{2i}{\delta(M_d^2)}{\cal S}_{12}(k,x)
5973: - \frac{4}{\delta(M_d^2)^2}\, k\cdot\partial\, {\cal S}_{12}(k,x)
5974: + O(\partial^2).
5975: \label{scalars-ke-12-solution2}
5976: \end{eqnarray}
5977: %
5978: This converges provided
5979: %
5980: \begin{equation}
5981: k\cdot\partial \ll \delta(M_d^2),
5982: \label{convergence:off-diagonal}
5983: \end{equation}
5984: %
5985: which is precisely the criterion for the validity of the gradient expansion
5986: in the off-diagonal equations.
5987:
5988: \vskip 0.1in
5989:
5990: One may argue that the derivatives acting on the mass must be included as well.
5991: To study the role of these derivatives, we assume that
5992: both the source and the masses depend on the time variable only. In this case
5993: the off-diagonal equation of motion reads
5994: %
5995: \begin{equation}
5996: \big[k_0\partial_t + \frac 12 (\partial_t \bar M_d^2)\partial_{k_0}
5997: + \frac i2 \delta(M_d^2) - ik_0\delta(\Xi_t)
5998: \big]i\Delta_{12}(k_0,t) = {\cal S}_{12}(k_0,t)
5999: \label{Delta_12:eom1+1}
6000: \end{equation}
6001: %
6002: for which the retarded Green function equation is
6003: %
6004: \begin{equation}
6005: \big[k_0\partial_t + \frac 12 (\partial_t \bar M_d^2)\partial_{k_0}
6006: + \frac i2 \delta(M_d^2) - ik_0\delta(\Xi_t) + \epsilon
6007: \big]G_\epsilon^r(k_0,k_0';t,t')
6008: = \delta(k_0-k_0')\delta(t-t')
6009: \,.
6010: \label{Green-function:eom1+1}
6011: \end{equation}
6012: %
6013: When written in the Fourier space
6014: %
6015: \begin{equation}
6016: G_\epsilon^r(k_0,k_0';t,t') = \int \frac{d\kappa}{2\pi} e^{i\kappa(t-t')}
6017: g_\epsilon^r(k_0,k_0';\kappa)
6018: \label{Fourier:Green-function:eom1+1}
6019: \end{equation}
6020: %
6021: equation~(\ref{Green-function:eom1+1}) becomes
6022: %
6023: \begin{equation}
6024: \big[ik_0\kappa + \frac 12 (\partial_t \bar M_d^2)\partial_{k_0}
6025: + \frac i2 \delta(M_d^2) - ik_0\delta(\Xi_t) + \epsilon
6026: \big]g_\epsilon^r(k_0,k_0';\kappa)
6027: = \delta(k_0-k_0')
6028: \label{Green-function:eom1+1:Fourier}
6029: \end{equation}
6030: %
6031: where we have ignored the higher order gradients in time (that is
6032: we took $\partial_t \bar M_d^2$ and $\delta(M_d^2)$ to be time independent).
6033: The retarded solution of~(\ref{Green-function:eom1+1:Fourier}) is given by (the solution
6034: proportional to $-\theta(k_0'-k_0)$ corresponds to the advanced Green function),
6035: %
6036: \begin{equation}
6037: g_\epsilon^r(k_0,k_0';\kappa) = \frac{2}{\partial_t \bar M^2_d}\theta(k_0-k_0')
6038: \exp\Big(\frac{i}{\partial_t \bar M^2_d} \Big[
6039: (\kappa-\delta(\Xi_t))
6040: ({k_0'}^2-k_0^2) + (\delta(M_d^2)-2i\epsilon)(k_0'-k_0)\Big]\Big)
6041: .
6042: \label{Green-function:solution:Fourier:1+1}
6043: \end{equation}
6044: %
6045: This solution is not unique. Indeed, replacing ${k_0'}^2$ by
6046: ${k_0'}[a {k_0'} + (1-a) k_0 ]$ is also a solution, which however differs
6047: from~(\ref{Green-function:solution:Fourier:1+1}) at higher order in gradients,
6048: and hence this ambiguity is irrelevant.
6049: From~(\ref{Green-function:solution:Fourier:1+1}) we then have
6050: %
6051: \begin{equation}
6052: G_\epsilon^r(k_0,k_0';t,t')
6053: = \frac{2}{\partial_t \bar M^2_d}\theta(k_0-k_0')
6054: \delta\Big(t-t'-\frac{k_0^2-{k_0'}^2}{\partial_t \bar M^2_d}\Big)
6055: \exp\Big(\frac{i(k_0'-k_0)}{\partial_t \bar M^2_d}
6056: [\delta(M_d^2) - (k_0'+k_0) \delta(\Xi_t) - 2i\epsilon]\Big)
6057: ,
6058: \label{Green-function:solution:1+1}
6059: \end{equation}
6060: %
6061: such that Eq.~(\ref{Delta_12:eom1+1}) is solved by
6062: %
6063: \begin{eqnarray}
6064: i\Delta_{12}(k,x) &=& \int dk_0'dt'
6065: G_\epsilon^r(k_0,k_0';t,t') {\cal S}_{12}(k_0',t')
6066: \label{scalars-ke-12-solutionC}
6067: \\
6068: &=& \int_{-\infty} ^{k_0} dk_0'
6069: {\cal S}_{12}\Big(k_0',t-\frac{k_0^2-{k_0'}^2}{\partial_t \bar M^2_d}\Big)
6070: \frac{2}{\partial_t \bar M^2_d}
6071: \exp\Big(\frac{i(k_0'-k_0)}{\partial_t \bar M^2_d}
6072: [\delta(M_d^2) - (k_0'+k_0) \delta(\Xi_t) - 2i\epsilon]\Big)
6073: .
6074: \nonumber
6075: \end{eqnarray}
6076: %
6077: The shift in time
6078: $\delta t = -(k_0^2-{k_0'}^2)/(\partial_t \bar M^2_d)$ corresponds to
6079: the time retardation caused by the force term.
6080: Since in the limit of a very slowly varying field
6081: ($\partial_t \rightarrow 1/T \rightarrow 0$),
6082: ${\cal S}_{12}$ is varying very slowly in comparison with the rapidly oscillating
6083: term, we can expand around $t'=t$ and $k_0'=k_0$,
6084: %
6085: \begin{equation}
6086: {\cal S}_{12}\Big(k_0',t-\frac{k_0^2-{k_0'}^2}{\partial_t \bar M^2_d}\Big)
6087: = {\cal S}_{12}(k_0,t)
6088: + (k_0'-k_0)\Big(\partial_{k_0}
6089: + \frac{k_0'+k_0}{\partial_t \bar M^2_d}\partial_t\Big){\cal S}_{12}(k_0,t)
6090: + {\cal O}(\partial_t^2,\partial_{k_0}^2)
6091: \end{equation}
6092: %
6093: and integrate~(\ref{scalars-ke-12-solutionC}) to obtain
6094: %
6095: \begin{eqnarray}
6096: i\Delta_{12}(k_0,t) &=& - \frac{2i}{\delta(M_d^2) }
6097: \Big[1
6098: + \frac{2i}{\delta(M_d^2)}\Big(k_0\partial_t
6099: + \frac 12 (\partial_t\bar{M}_d^2)\partial_{k_0}
6100: - ik_0\delta(\Xi_t) \Big)
6101: \Big]{\cal S}_{12}(k_0,t)
6102: + {\cal O}(\partial_t^2).\;\;
6103: \label{scalars-ke-12-solution4}
6104: \end{eqnarray}
6105: %
6106: One of the higher order terms we have dropped is, for example,
6107: $4i[(\partial_t \bar M_d^2)/(\delta M_d^2)^3]\partial_t {\cal S}(k_0,t)$.
6108: This converges provided
6109: %
6110: \begin{equation}
6111: k_0\partial_t, (\partial_t\bar{M}_d^2)\partial_{k_0} \ll \delta(M_d^2).
6112: \end{equation}
6113: %
6114: From this and Eq.~(\ref{scalars-ke-12-solution2}-\ref{convergence:off-diagonal})
6115: we conclude that the criterion for validity of the gradient expansion
6116: in the off-diagonal equations reads
6117: %
6118: \begin{equation}
6119: k\cdot\partial, (\partial\bar{M}_d^2)\cdot\partial_{k},
6120: k\cdot\delta(\Xi) \ll \delta(M_d^2).
6121: \end{equation}
6122: %
6123:
6124:
6125:
6126:
6127:
6128:
6129:
6130:
6131:
6132: \end{appendix}
6133:
6134:
6135:
6136:
6137:
6138:
6139:
6140: % BIBLIOGRAPHY
6141:
6142: \cleardoublepage
6143:
6144: \begin{thebibliography}{999}
6145:
6146:
6147:
6148: \bibitem{PSW_2}
6149: T.~Prokopec, M.~G.~Schmidt, S.~Weinstock,
6150: ``Transport equations for chiral fermions to order $\hbar$ and electroweak baryogenesis: Part~II,''
6151: %Annals Phys.\ {\bf XXX} (2004) XXX
6152: arXiv:hep-ph/0406140.
6153: %%CITATION = HEP-PH 0312110;%%
6154:
6155:
6156:
6157: \bibitem{Schwinger:1961}
6158: Julian S.~Schwinger,
6159: ``Brownian Motion Of A Quantum Oscillator,''
6160: J.\ Math.\ Phys.\ {\bf 2} (1961) 407.
6161: %%CITATION = JMAPA,2,407;%%
6162:
6163: \bibitem{KadanoffBaym:1962} L.P.\ Kadanoff and Gordon\ Baym,
6164: {\it Quantum Statistical Mechanics}, Benjamin Press, New York (1962).
6165:
6166: \bibitem{Keldysh:1964}
6167: L.~V.~Keldysh,
6168: ``Diagram Technique For Nonequilibrium Processes,''
6169: Zh.\ Eksp.\ Teor.\ Fiz.\ {\bf 47} (1964) 1515
6170: [Sov.\ Phys.\ JETP {\bf 20} (1964) 1018].
6171: %%CITATION = ZETFA,47,1515;%%
6172:
6173: %\cite{ChouSuHaoYu:1985}
6174: \bibitem{ChouSuHaoYu:1985}
6175: K.~c.~Chou, Z.~b.~Su, B.~l.~Hao and L.~Yu,
6176: ``Equilibrium And Nonequilibrium Formalisms Made Unified,''
6177: Phys.\ Rept.\ {\bf 118} (1985) 1.
6178: %%CITATION = PRPLC,118,1;%%
6179:
6180: \bibitem{CalzettaHu:1986}
6181: E.~Calzetta and B.~L.~Hu,
6182: ``Nonequilibrium Quantum Fields: Closed Time Path Effective Action,
6183: Wigner Function And Boltzmann Equation,''
6184: Phys.\ Rev.\ D {\bf 37} (1988) 2878.
6185:
6186: %\cite{Kolb:qa}
6187: \bibitem{KolbWolfram:1980}
6188: E.~W.~Kolb and S.~Wolfram,
6189: ``Baryon Number Generation In The Early Universe,''
6190: Nucl.\ Phys.\ B {\bf 172} (1980) 224
6191: [Erratum-ibid.\ B {\bf 195} (1982) 542].
6192: %%CITATION = NUPHA,B172,224;%%
6193:
6194: \bibitem{KuzminRubakovShaposhnikov:1985}
6195: V.~A.~Kuzmin, V.~A.~Rubakov and M.~E.~Shaposhnikov,
6196: ``On The Anomalous Electroweak Baryon Number Nonconservation
6197: In The Early Universe,''
6198: Phys.\ Lett.\ B {\bf 155} (1985) 36.
6199: %%CITATION = PHLTA,B155,36;%%
6200:
6201: \bibitem{CohenKaplanNelson:1990+1991}
6202: %\cite{Cohen:py}
6203: %\bibitem{Cohen:py}
6204: A.~G.~Cohen, D.~B.~Kaplan and A.~E.~Nelson,
6205: ``Weak Scale Baryogenesis,''
6206: Phys.\ Lett.\ B {\bf 245} (1990) 561;
6207: %%CITATION = PHLTA,B245,561;%%
6208: %\bibitem{Cohen:1990it}
6209: A.~G.~Cohen, D.~B.~Kaplan and A.~E.~Nelson,
6210: ``Baryogenesis At The Weak Phase Transition,''
6211: Nucl.\ Phys.\ B {\bf 349} (1991) 727.
6212: %%CITATION = NUPHA,B349,727;%%
6213:
6214:
6215: %\cite{Shaposhnikov:jp}
6216: \bibitem{Shaposhnikov:1986}
6217: M.~E.~Shaposhnikov,
6218: ``Possible Appearance Of The Baryon Asymmetry Of The Universe
6219: In An Electroweak Theory,''
6220: JETP Lett.\ {\bf 44} (1986) 465
6221: [Pisma Zh.\ Eksp.\ Teor.\ Fiz.\ {\bf 44} (1986) 364].
6222: %%CITATION = JTPLA,44,465;%%
6223:
6224: %\cite{Shaposhnikov:tw}
6225: \bibitem{Shaposhnikov:1987}
6226: M.~E.~Shaposhnikov,
6227: ``Baryon Asymmetry Of The Universe In Standard Electroweak Theory,''
6228: Nucl.\ Phys.\ B {\bf 287} (1987) 757.
6229: %%CITATION = NUPHA,B287,757;%%
6230:
6231: \bibitem{KajantieLaineRummukainenShaposhnikov}
6232: %\bibitem{Kajantie:1997qd}
6233: K.~Kajantie, M.~Laine, K.~Rummukainen and M.~Shaposhnikov,
6234: ``A non-perturbative analysis of the finite T phase transition
6235: in SU(2) x U(1) electroweak theory,''
6236: Nucl.\ Phys.\ B {\bf 493} (1997) 413
6237: [hep-lat/9612006];
6238: %%CITATION = HEP-LAT 9612006;%%
6239: %\bibitem{Kajantie:1996mn}
6240: %K.~Kajantie, M.~Laine, K.~Rummukainen and M.~Shaposhnikov,
6241: ``Is there a hot electroweak phase transition at $m(H) >$ approx. $m(W)$?,''
6242: Phys.\ Rev.\ Lett.\ {\bf 77} (1996) 2887
6243: [hep-ph/9605288];
6244: %%CITATION = HEP-PH 9605288;%%
6245: %\bibitem{Kajantie:1996kf}
6246: %K.~Kajantie, M.~Laine, K.~Rummukainen and M.~Shaposhnikov,
6247: ``The Electroweak Phase Transition: A Non-Perturbative Analysis,''
6248: Nucl.\ Phys.\ B {\bf 466} (1996) 189
6249: [hep-lat/9510020].
6250: %%CITATION = HEP-LAT 9510020;%%
6251:
6252: \bibitem{RummukainenTsypinKajantieLaineShaposhnikov:1998}
6253: K.~Rummukainen, M.~Tsypin, K.~Kajantie, M.~Laine and M.~E.~Shaposhnikov,
6254: ``The universality class of the electroweak theory,''
6255: Nucl.\ Phys.\ B {\bf 532} (1998) 283
6256: [arXiv:hep-lat/9805013].
6257:
6258: \bibitem{CsikorFodorHeitger:1998}
6259: F.~Csikor, Z.~Fodor and J.~Heitger,
6260: ``Endpoint of the hot electroweak phase transition,''
6261: Phys.\ Rev.\ Lett.\ {\bf 82} (1999) 21
6262: [arXiv:hep-ph/9809291].
6263: %%CITATION = HEP-PH 9809291;%%
6264:
6265: %\cite{Aoki:1999fi}
6266: \bibitem{AokiCsikorFodorUkawa:1999}
6267: Y.~Aoki, F.~Csikor, Z.~Fodor and A.~Ukawa,
6268: ``The endpoint of the first-order phase transition of the
6269: SU(2) gauge-Higgs model on a 4-dimensional isotropic lattice,''
6270: Phys.\ Rev.\ D {\bf 60} (1999) 013001
6271: [arXiv:hep-lat/9901021].
6272: %%CITATION = HEP-LAT 9901021;%%
6273:
6274: %\cite{Buchmuller:1994qy}
6275: \bibitem{BuchmullerPhilipsen:1994}
6276: W.~Buchmuller and O.~Philipsen,
6277: ``Phase structure and phase transition of the SU(2) Higgs model
6278: in three-dimensions,''
6279: Nucl.\ Phys.\ B {\bf 443} (1995) 47
6280: [arXiv:hep-ph/9411334].
6281: %%CITATION = HEP-PH 9411334;%%
6282:
6283: %\cite{Bergerhoff:1994sj}
6284: \bibitem{BergerhoffWetterich:1994}
6285: B.~Bergerhoff and C.~Wetterich,
6286: ``The Strongly interacting electroweak phase transition,''
6287: Nucl.\ Phys.\ B {\bf 440} (1995) 171
6288: [arXiv:hep-ph/9409295].
6289: %%CITATION = HEP-PH 9409295;%%
6290:
6291: \bibitem{LaineRummukainen:2001}
6292: M.~Laine and K.~Rummukainen,
6293: ``Two Higgs doublet dynamics at the electroweak phase transition:
6294: A non-perturbative study,''
6295: Nucl.\ Phys.\ B {\bf 597} (2001) 23
6296: [arXiv:hep-lat/0009025].
6297:
6298: \bibitem{CarenaQuirosWagner:1996}
6299: M.~Carena, M.~Quiros and C.~E.~Wagner,
6300: ``Opening the Window for Electroweak Baryogenesis,''
6301: Phys.\ Lett.\ B {\bf 380} (1996) 81
6302: [arXiv:hep-ph/9603420].
6303:
6304: \bibitem{CarenaQuirosWagner:1998}
6305: M.~Carena, M.~Quiros and C.~E.~Wagner,
6306: ``Electroweak baryogenesis and Higgs and stop searches at LEP and the Tevatron,''
6307: Nucl.\ Phys.\ B {\bf 524} (1998) 3
6308: [arXiv:hep-ph/9710401].
6309:
6310:
6311: \bibitem{ClineKainulainen:1996}
6312: J.~M.~Cline and K.~Kainulainen,
6313: ``Supersymmetric Electroweak Phase Transition: Beyond Perturbation Theory,''
6314: Nucl.\ Phys.\ B {\bf 482} (1996) 73
6315: [arXiv:hep-ph/9605235].
6316:
6317: \bibitem{Laine:1996}
6318: M.~Laine,
6319: ``Effective theories of MSSM at high temperature,''
6320: Nucl.\ Phys.\ B {\bf 481} (1996) 43
6321: [Erratum-ibid.\ B {\bf 548} (1996) 637]
6322: [arXiv:hep-ph/9605283].
6323:
6324: \bibitem{Losada97}
6325: M.~Losada,
6326: ``High temperature dimensional reduction of the MSSM
6327: and other multi-scalar models,''
6328: Phys.\ Rev.\ D {\bf 56} (1997) 2893
6329: [arXiv:hep-ph/9605266].
6330:
6331: \bibitem{BodeckerJohnLaineSchmidt:1997}
6332: D.~Bodeker, P.~John, M.~Laine and M.~G.~Schmidt,
6333: ``The 2-loop MSSM finite temperature effective potential with stop condensation,''
6334: Nucl.\ Phys.\ B {\bf 497} (1997) 387
6335: [arXiv:hep-ph/9612364].
6336:
6337: \bibitem{LaineRummukainen:1998}
6338: M.~Laine and K.~Rummukainen,
6339: ``The MSSM electroweak phase transition on the lattice,''
6340: Nucl.\ Phys.\ B {\bf 535} (1998) 423
6341: [arXiv:hep-lat/9804019].
6342:
6343: \bibitem{PilaftsisWagner:1999}
6344: %\bibitem{Pilaftsis:1999qt}
6345: A.~Pilaftsis and C.~E.~Wagner,
6346: ``Higgs bosons in the minimal supersymmetric standard model
6347: with explicit CP violation,''
6348: Nucl.\ Phys.\ B {\bf 553} (1999) 3
6349: [hep-ph/9902371].
6350: %%CITATION = HEP-PH 9902371;%%
6351:
6352: %\cite{Csikor:2000sq}
6353: \bibitem{CsikorFodorHegedusJakovacKatzPiroth:2000}
6354: F.~Csikor, Z.~Fodor, P.~Hegedus, A.~Jakovac, S.~D.~Katz and A.~Piroth,
6355: ``Electroweak phase transition in the MSSM: 4-dimensional lattice
6356: simulations,''
6357: Phys.\ Rev.\ Lett.\ {\bf 85} (2000) 932
6358: [arXiv:hep-ph/0001087].
6359: %%CITATION = HEP-PH 0001087;%%
6360:
6361: \bibitem{Espinosa:1996}
6362: J.~R.~Espinosa,
6363: ``Dominant Two-Loop Corrections to the MSSM Finite Temperature Effective
6364: Potential,''
6365: Nucl.\ Phys.\ B {\bf 475} (1996) 273
6366: [arXiv:hep-ph/9604320].
6367: %%CITATION = HEP-PH 9604320;%%
6368:
6369: \bibitem{deCarlosEspinosa:1997}
6370: B.~de Carlos and J.~R.~Espinosa,
6371: ``The baryogenesis window in the MSSM,''
6372: Nucl.\ Phys.\ B {\bf 503} (1997) 24
6373: [arXiv:hep-ph/9703212].
6374: %%CITATION = HEP-PH 9703212;%%
6375:
6376: \bibitem{Pietroni:1993}
6377: M.~Pietroni,
6378: ``The Electroweak phase transition in a nonminimal supersymmetric model,''
6379: Nucl.\ Phys.\ B {\bf 402} (1993) 27
6380: [arXiv:hep-ph/9207227].
6381:
6382: \bibitem{DaviesFroggattMoorhouse:1996}
6383: A.~T.~Davies, C.~D.~Froggatt and R.~G.~Moorhouse,
6384: ``Electroweak Baryogenesis in the Next to Minimal Supersymmetric Model,''
6385: Phys.\ Lett.\ B {\bf 372} (1996) 88
6386: [arXiv:hep-ph/9603388];
6387: %%CITATION = HEP-PH 9603388;%%
6388:
6389: \bibitem{HuberSchmidt:1999}
6390: S.~J.~Huber and M.~G.~Schmidt,
6391: ``SUSY variants of the electroweak phase transition,''
6392: Eur.\ Phys.\ J.\ C {\bf 10} (1999) 473
6393: [arXiv:hep-ph/9809506].
6394:
6395: \bibitem{HuberJohnLaineSchmidt}
6396: %\bibitem{Huber:2000sa}
6397: S.~J.~Huber, P.~John, M.~Laine and M.~G.~Schmidt,
6398: ``CP violating bubble wall profiles,''
6399: Phys.\ Lett.\ B {\bf 475} (2000) 104
6400: [hep-ph/9912278].
6401: %%CITATION = HEP-PH 9912278;%%
6402:
6403: \bibitem{JoyceProkopecTurok:1995}
6404: M.~Joyce, T.~Prokopec and N.~Turok,
6405: ``Electroweak baryogenesis from a classical force,''
6406: Phys.\ Rev.\ Lett.\ {\bf 75} (1995) 1695
6407: [Erratum-ibid.\ {\bf 75} (1995) 3375]
6408: [arXiv:hep-ph/9408339].
6409: %%CITATION = HEP-PH 9408339;%%
6410:
6411: \bibitem{JoyceProkopecTurok:1996}
6412: M.~Joyce, T.~Prokopec and N.~Turok,
6413: ``Nonlocal electroweak baryogenesis. Part 2: The Classical regime,''
6414: Phys.\ Rev.\ D {\bf 53} (1996) 2958
6415: [arXiv:hep-ph/9410282].
6416: %%CITATION = HEP-PH 9410282;%%
6417:
6418: \bibitem{ClineJoyceKainulainen:1998}
6419: J.~M.~Cline, M.~Joyce and K.~Kainulainen,
6420: ``Supersymmetric electroweak baryogenesis in the WKB approximation,''
6421: Phys.\ Lett.\ B {\bf 417} (1998) 79
6422: [Erratum-ibid.\ B {\bf 448} (1998) 321]
6423: [arXiv:hep-ph/9708393].
6424: %%CITATION = HEP-PH 9708393;%%
6425:
6426: %\cite{Kainulainen:1999kc}
6427: \bibitem{Kainulainen:1999}
6428: K.~Kainulainen,
6429: Proceedings of COSMO 99, Trieste, Italy, Oct 1999,
6430: p.\ 383, World Scientific 2000
6431: ``Sources for electroweak baryogenesis,''
6432: [arXiv:hep-ph/0002273].
6433: %%CITATION = HEP-PH 0002273;%%
6434:
6435: \bibitem{ClineKainulainen:2000}
6436: J.~M.~Cline and K.~Kainulainen,
6437: ``A new source for electroweak baryogenesis in the MSSM,''
6438: Phys.\ Rev.\ Lett.\ {\bf 85} (2000) 5519
6439: [arXiv:hep-ph/0002272].
6440:
6441: \bibitem{ClineJoyceKainulainen:2000+2001}
6442: %\bibitem{Cline:2000kb}
6443: J.~M.~Cline and K.~Kainulainen,
6444: ``A new source for electroweak baryogenesis in the MSSM,''
6445: Phys.\ Rev.\ Lett.\ {\bf 85} (2000) 5519
6446: [arXiv:hep-ph/0002272];
6447: %%CITATION = HEP-PH 0002272;%%
6448: %\bibitem{Cline:2000nw}
6449: J.~M.~Cline, M.~Joyce and K.~Kainulainen,
6450: ``Supersymmetric electroweak baryogenesis,''
6451: JHEP {\bf 0007} (2000) 018
6452: [hep-ph/0006119]
6453: %%CITATION = HEP-PH 0006119;%%
6454: [Erratum, arXiv:hep-ph/0110031];
6455: %%CITATION = HEP-PH 0110031;%%
6456:
6457: \bibitem{HuberJohnSchmidt:2001}
6458: %\cite{Huber:2001xf}
6459: %\bibitem{Huber:2001xf}
6460: S.J.~Huber, P.~John and M.G.~Schmidt
6461: ``Bubble walls, CP violation and electroweak baryogenesis in the MSSM,''
6462: Eur.\ Phys.\ J.\ C {\bf 20} (2001) 695
6463: [arXiv:hep-ph/0101249].
6464: %%CITATION = HEP-PH 0101249;%%
6465:
6466: \bibitem{HuberSchmidt:2000}
6467: S.J.~Huber and M.G.~Schmidt
6468: ``Electroweak baryogenesis in a SUSY model with a gauge singlet,''
6469: Nucl.\ Phys.\ B {\bf 606} (2001) 183
6470: [arXiv:hep-ph/0003122];
6471: %%CITATION = HEP-PH 0003122;%%
6472:
6473: \bibitem{HuberSchmidt:2001}
6474: S.~J.~Huber and M.~G.~Schmidt,
6475: ``Baryogenesis at the electroweak phase transition for
6476: a SUSY model with a gauge singlet,''
6477: arXiv:hep-ph/0011059.
6478: %%CITATION = HEP-PH 0011059;%%
6479:
6480:
6481: \bibitem{CohenKaplanNelson:1991}
6482: A.~G.~Cohen, D.~B.~Kaplan and A.~E.~Nelson,
6483: ``Spontaneous baryogenesis at the weak phase transition,''
6484: Phys.\ Lett.\ B {\bf 263} (1991) 86.
6485: %%CITATION = PHLTA,B263,86;%%
6486:
6487: %\cite{Dine:1994vf}
6488: \bibitem{DineThomas:1994}
6489: M.~Dine and S.~Thomas,
6490: ``Electroweak baryogenesis in the adiabatic limit,''
6491: Phys.\ Lett.\ B {\bf 328} (1994) 73
6492: [arXiv:hep-ph/9401265].
6493: %%CITATION = HEP-PH 9401265;%%
6494:
6495: \bibitem{ComelliPietroniRiotto:1995}
6496: D.~Comelli, M.~Pietroni and A.~Riotto,
6497: ``Particle currents in a space-time dependent
6498: and CP violating Higgs background: A Field theory approach,''
6499: Phys.\ Rev.\ D {\bf 53} (1996) 4668
6500: [arXiv:hep-ph/9506278];
6501: %%CITATION = HEP-PH 9506278;%%
6502: D.~Comelli, M.~Pietroni and A.~Riotto,
6503: ``Particle currents on a CP violating Higgs background
6504: and the spontaneous baryogenesis mechanism,''
6505: Phys.\ Lett.\ B {\bf 354} (1995) 91
6506: [arXiv:hep-ph/9504265].
6507: %%CITATION = HEP-PH 9504265;%%
6508:
6509: \bibitem{HuetNelson:1995+1996}
6510: P.~Huet and A.~E.~Nelson,
6511: ``CP violation and electroweak baryogenesis in extensions
6512: of the standard model,''
6513: Phys.\ Lett.\ B {\bf 355} (1995) 229
6514: [arXiv:hep-ph/9504427];
6515: %%CITATION = HEP-PH 9504427;%%
6516:
6517: P.~Huet and A.~E.~Nelson,
6518: ``Electroweak baryogenesis in supersymmetric models,''
6519: Phys.\ Rev.\ D {\bf 53} (1996) 4578
6520: [arXiv:hep-ph/9504427].
6521: %%CITATION = HEP-PH 9504427;%%
6522:
6523: \bibitem{Riotto:1995}
6524: A.~Riotto,
6525: ``Towards a Nonequilibrium Quantum Field Theory Approach
6526: to Electroweak Baryogenesis,''
6527: Phys.\ Rev.\ D {\bf 53} (1996) 5834
6528: [arXiv:hep-ph/9510271].
6529: %%CITATION = HEP-PH 9510271;%%
6530:
6531: \bibitem{Riotto:1998}
6532: A.~Riotto,
6533: ``The more relaxed supersymmetric electroweak baryogenesis,''
6534: Phys.\ Rev.\ D {\bf 58} (1998) 095009
6535: [arXiv:hep-ph/9803357].
6536: %%CITATION = HEP-PH 9803357;%%
6537:
6538: \bibitem{CarenaQuirosRiottoViljaWagner:1997}
6539: M.~Carena, M.~Quiros, A.~Riotto, I.~Vilja and C.~E.~Wagner,
6540: ``Electroweak baryogenesis and low energy supersymmetry,''
6541: Nucl.\ Phys.\ B {\bf 503} (1997) 387
6542: [arXiv:hep-ph/9702409].
6543: %%CITATION = HEP-PH 9702409;%%
6544:
6545: \bibitem{CarenaMorenoQuirosSecoWagner:2000}
6546: M.~Carena, J.~M.~Moreno, M.~Quiros, M.~Seco and C.~E.~Wagner,
6547: ``Supersymmetric CP-violating currents and electroweak baryogenesis,''
6548: Nucl.\ Phys.\ B {\bf 599} (2001) 158
6549: [arXiv:hep-ph/0011055].
6550: %%CITATION = HEP-PH 0011055;%%
6551:
6552: \bibitem{CarenaQuirosSecoWagner:2002}
6553: M.~Carena, M.~Quiros, M.~Seco and C.~E.~M.~Wagner,
6554: ``Improved results in supersymmetric electroweak baryogenesis,''
6555: Nucl.\ Phys.\ B {\bf 650} (2003) 24
6556: [arXiv:hep-ph/0208043].
6557: %%CITATION = HEP-PH 0208043;%%
6558:
6559: \bibitem{KainulainenProkopecSchmidtWeinstock:2001}
6560: K.~Kainulainen, T.~Prokopec, M.~G.~Schmidt and S.~Weinstock,
6561: ``First principle derivation of
6562: semiclassical force for electroweak baryogenesis,''
6563: JHEP {\bf 0106} (2001) 031 [arXiv:hep-ph/0105295].
6564: %%CITATION = HEP-PH 0105295;%%
6565:
6566: \bibitem{KainulainenProkopecSchmidtWeinstock:2002}
6567: K.~Kainulainen, T.~Prokopec, M.~G.~Schmidt and S.~Weinstock,
6568: ``Semiclassical force for electroweak baryogenesis: Three-dimensional derivation,''
6569: arXiv:hep-ph/0202177.
6570: %%CITATION = HEP-PH 0202177;%%
6571:
6572: \bibitem{KainulainenProkopecSchmidtWeinstock:2002b}
6573: K.~Kainulainen, T.~Prokopec, M.~G.~Schmidt and S.~Weinstock,
6574: ``Some aspects of collisional sources for electroweak baryogenesis,''
6575: in proceedings of COSMO-01 [arXiv:hep-ph/0201245].
6576: %%CITATION = HEP-PH 0201245;%%
6577:
6578: \bibitem{KainulainenProkopecSchmidtWeinstock:2002c}
6579: K.~Kainulainen, T.~Prokopec, M.~G.~Schmidt and S.~Weinstock,
6580: ``Quantum Boltzmann equations for electroweak baryogenesis
6581: including gauge fields,'' in proceedings of COSMO-01
6582: [arXiv:hep-ph/0201293].
6583: %%CITATION = HEP-PH 0201293;%%
6584:
6585: \bibitem{KonstandinProkopecSchmidtWeinstock:2004}
6586: Thomas Konstandin, Tomislav Prokopec, Michael G.~Schmidt and Steffen
6587: Weinstock, in preparation.
6588:
6589:
6590: % 2pi
6591:
6592: \bibitem{LuttingerWard:1960} J.~M.~Luttinger and J.~C.~Ward,
6593: Phys.\ Rev.\ {\bf 118} (1960) 1417.
6594:
6595: \bibitem{CornwallJackiwTomboulis:1974}
6596: J.~M.~Cornwall, R.~Jackiw and E.~Tomboulis,
6597: ``Effective Action For Composite Operators,''
6598: Phys.\ Rev.\ D {\bf 10} (1974) 2428.
6599:
6600: \bibitem{BergesSerreau:2002}
6601: J.~Berges and J.~Serreau,
6602: ``Parametric resonance in quantum field theory,''
6603: arXiv:hep-ph/0208070.
6604: %%CITATION = HEP-PH 0208070;%%
6605:
6606: \bibitem{AartsAhrensmeierBaierBergesSerreau:2002}
6607: G.~Aarts, D.~Ahrensmeier, R.~Baier, J.~Berges and J.~Serreau,
6608: ``Far-from-equilibrium dynamics with broken symmetries
6609: from the 2PI-1/N expansion,''
6610: Phys.\ Rev.\ D {\bf 66} (2002) 045008
6611: [arXiv:hep-ph/0201308].
6612:
6613: %%CITATION = HEP-PH 0201308;%%
6614: \bibitem{CalzettaHu:2002}
6615: E.~A.~Calzetta and B.~L.~Hu,
6616: ``Thermalization of an interacting quantum field
6617: in the CTP-2PI next-to-leading-order large N scheme,''
6618: arXiv:hep-ph/0205271.
6619: %%CITATION = HEP-PH 0205271;%%
6620:
6621: \bibitem{KhlebnikovTkachev:1996}
6622: S.~Y.~Khlebnikov and I.~I.~Tkachev,
6623: %``Classical decay of inflaton,''
6624: Phys.\ Rev.\ Lett.\ {\bf 77} (1996) 219
6625: [arXiv:hep-ph/9603378].
6626: %%CITATION = HEP-PH 9603378;%%
6627:
6628: \bibitem{ProkopecRoos:1996}
6629: T.~Prokopec and T.~G.~Roos,
6630: ``Lattice study of classical inflaton decay,''
6631: Phys.\ Rev.\ D {\bf 55} (1997) 3768
6632: [arXiv:hep-ph/9610400].
6633: %%CITATION = HEP-PH 9610400;%%
6634:
6635: \bibitem{Berges:2002}
6636: J.~Berges
6637: ``Nonequilibrium quantum fields and the classical field theory limit,''
6638: Nucl.\ Phys.\ A {\bf 702} (2002) 351
6639: [arXiv:hep-ph/0201204].
6640: %%CITATION = HEP-PH 0201204;%%
6641: \\
6642: %\bibitem{Berges:2001fi}
6643: J.~Berges,
6644: ``Controlled nonperturbative dynamics of quantum fields out of equilibrium,''
6645: Nucl.\ Phys.\ A {\bf 699} (2002) 847
6646: [arXiv:hep-ph/0105311].
6647: %%CITATION = HEP-PH 0105311;%%
6648:
6649: \bibitem{BergesBorsanyiSerreau:2002}
6650: J.~Berges, S.~Borsanyi and J.~Serreau,
6651: ``Thermalization of fermionic quantum fields,''
6652: Nucl.\ Phys.\ B {\bf 660} (2003) 51
6653: [arXiv:hep-ph/0212404].
6654: %%CITATION = HEP-PH 0212404;%%
6655:
6656: \bibitem{AartsBerges:2001}
6657: G.~Aarts and J.~Berges,
6658: ``Classical aspects of quantum fields far from equilibrium,''
6659: Phys.\ Rev.\ Lett.\ {\bf 88} (2002) 041603
6660: [arXiv:hep-ph/0107129].
6661: %%CITATION = HEP-PH 0107129;%%
6662:
6663: \bibitem{BergesCox:2000}
6664: J.~Berges and J.~Cox,
6665: ``Thermalization of quantum fields from time-reversal invariant
6666: evolution equations,''
6667: Phys.\ Lett.\ B {\bf 517} (2001) 369
6668: [arXiv:hep-ph/0006160].
6669: %%CITATION = HEP-PH 0006160;%%
6670:
6671:
6672:
6673:
6674:
6675: \bibitem{ThickWalls}
6676: J.~M.~Moreno, M.~Quiros and M.~Seco,
6677: ``Bubbles in the supersymmetric standard model,''
6678: Nucl.\ Phys.\ B {\bf 526} (1998) 489
6679: [hep-ph/9801272];
6680: %%CITATION = HEP-PH 9801272;%%
6681: %\bibitem{Cline:1999wi}
6682: J.~M.~Cline, G.~D.~Moore and G.~Servant,
6683: ``Was the electroweak phase transition preceded by a color broken phase?,''
6684: Phys.\ Rev.\ D {\bf 60} (1999) 105035
6685: [hep-ph/9902220];
6686: %%CITATION = HEP-PH 9902220;%%
6687: G.~D.~Moore,
6688: ``Electroweak bubble wall friction: Analytic results,''
6689: JHEP {\bf 0003} (2000) 006
6690: [hep-ph/0001274];
6691: %%CITATION = HEP-PH 0001274;%%
6692: %\bibitem{Cline:1998hy}
6693: J.~M.~Cline and G.~D.~Moore,
6694: ``Supersymmetric electroweak phase transition:
6695: Baryogenesis versus experimental constraints,''
6696: Phys.\ Rev.\ Lett.\ {\bf 81} (1998) 3315
6697: [hep-ph/9806354].
6698: %%CITATION = HEP-PH 9806354;%%
6699:
6700: \bibitem{LeBellac:1996}
6701: M.~Le~Bellac, ``Thermal Field Theory,''
6702: Cambridge University Press (1996).
6703:
6704:
6705:
6706:
6707:
6708: \bibitem{Henning:1995} P.\ Henning,
6709: P.~A.~Henning,
6710: ``Thermo Field Dynamics For Quantum Fields With Continuous Mass Spectrum,''
6711: Phys.\ Rept.\ {\bf 253} (1995) 235.
6712: %%CITATION = PRPLC,253,235;%%
6713:
6714: \bibitem{GreinerLeupold:1998}
6715: C.~Greiner and S.~Leupold,
6716: ``Stochastic interpretation of Kadanoff-Baym equations
6717: and their relation to Langevin processes,''
6718: Annals Phys.\ {\bf 270} (1998) 328
6719: [arXiv:hep-ph/9802312].
6720: %%CITATION = HEP-PH 9802312;%%
6721:
6722:
6723: \bibitem{Leupold:2000}
6724: S.~Leupold,
6725: ``Life time of resonances in transport simulations,''
6726: Nucl.\ Phys.\ A {\bf 695} (2001) 377
6727: [arXiv:nucl-th/0008036].
6728: %%CITATION = NUCL-TH 0008036;%%
6729:
6730: \bibitem{ItzyksonZuber:1980} C.~Itzykson, J.~B.~Zuber, Quantum Field Theory,
6731: McGraw-Hill Inc., (1980); ISBN 0-07-032071-3, Chapter 3.
6732:
6733: \bibitem{MooreProkopec:1995}
6734: %\bibitem{Moore:1995ua}
6735: G.~D.~Moore and T.~Prokopec,
6736: ``Bubble wall velocity in a first order electroweak phase transition,''
6737: Phys.\ Rev.\ Lett.\ {\bf 75} (1995) 777
6738: [arXiv:hep-ph/9503296];\\
6739: %%CITATION = HEP-PH 9503296;%%
6740: G.~D.~Moore and T.~Prokopec,
6741: ``How fast can the wall move?
6742: A Study of the electroweak phase transition dynamics,''
6743: Phys.\ Rev.\ D {\bf 52} (1995) 7182
6744: [hep-ph/9506475].
6745: %%CITATION = HEP-PH 9506475;%%
6746: %\bibitem{Moore:1995ua}
6747:
6748:
6749: %\cite{John:2000zq}
6750: \bibitem{JohnSchmidt:2000+2001}
6751: P.~John and M.~G.~Schmidt,
6752: %``Do stops slow down electroweak bubble walls?,''
6753: Nucl.\ Phys.\ B {\bf 598} (2001) 291
6754: [Erratum-ibid.\ B {\bf 648} (2003) 449]
6755: [arXiv:hep-ph/0002050].
6756: %%CITATION = HEP-PH 0002050;%%
6757:
6758:
6759:
6760:
6761: %\cite{GarbrechtProkopecSchmidt:2003}
6762: \bibitem{GarbrechtProkopecSchmidt:2003}
6763: B.~Garbrecht, T.~Prokopec and M.~G.~Schmidt,
6764: %``Coherent baryogenesis,''
6765: arXiv:hep-ph/0304088.
6766: %%CITATION = HEP-PH 0304088;%%
6767:
6768: %\cite{Huber:1999sa}
6769: \bibitem{HuberJohnLaineSchmidt:2000}
6770: S.~J.~Huber, P.~John, M.~Laine and M.~G.~Schmidt,
6771: %``CP violating bubble wall profiles,''
6772: Phys.\ Lett.\ B {\bf 475} (2000) 104
6773: [arXiv:hep-ph/9912278].
6774: %%CITATION = HEP-PH 9912278;%%
6775:
6776:
6777: \bibitem{HaberKane:1985}
6778: H.~E.~Haber and G.~L.~Kane,
6779: ``The Search For Supersymmetry: Probing Physics Beyond The Standard Model,''
6780: Phys.\ Rept.\ {\bf 117} (1985) 75.
6781:
6782:
6783:
6784:
6785:
6786: \end{thebibliography}
6787:
6788: \cleardoublepage
6789:
6790: \end{document}
6791:
6792:
6793: